EPA
            United States
            Environmental Protection
            Agency
               Office of Health and
               Environmental Assessment
               Washington DC 20460
EPA-600/8-82-004B
December 1983
External Review Draft
            Research and Development
Health Assessment    Review
Document for           Draft
Dichloromethane      
-------
                                        EPA-600/8-82-004B
                                              December 1983
                                                 Review Draft
         Health Assessment Document
                             for
                  Dichloromethane
                (Methylene Chloride)
                             NOTICE

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

-------
                                  DISCLAIMER

    This report 1s an external  draft  for review purposes only and does not
constitute Agency policy.   Mention of trade names or commercial products does
not constitute endorsement or  recommendation for use.
          U.S. Envtr6iimsntal PrtWctlOT

-------
                                    PREFACE

    The Office of Health and Environmental  Assessment,  1n 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 A1r 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.
                                    111

-------
                         TABLE OF CONTENTS (continued)
5.   HEALTH EFFECTS OF DICHLOROMETHANE 	    5-1
    5.1  HUMAN HEALTH EFFECTS 	    5-1
         5.1.1  Acute Exposures 	    5-1
         5.1.2  Chronic Effects 	    5-6
    5.2  EFFECTS ON LABORATORY ANIMALS 	    5-7
         5.2.1  Acute Effects 	    5-7
         5.2.2  Chronic Effects 	    5-17
    5.3  TERATOGENICITY, MUTAGENICITY, AND CARCINOGENICITY 	    5-22
         5.3.1  Teratogenicity, Embryotoxicity,  and Reproductive
                Effects 	    5-22
         5.3.2  Mutageni city 	    5-27
         5.3.3  Evaluation of the Carcinogenicity of Methylene
                Chloride 	     5-53
    5.4  REFERENCES 	     5-101

         APPENDIX 	     A-l
                                       vi
005DC1/E                                                                12/22/83

-------
                                LIST OF TABLES

Table                                                                      Page

3-1   Synonyms and identifiers for dichloromethane 	    3-2
3-2   Selected properties of dichloromethane 	    3-3
3-3   Producers of dichloromethane 	    3-5
3-4   Consumption of dichloromethane 	    3-6
3-5   Reaction rate data for OH + CH2C12 	    3-8
3-6   Ambient air levels of dichloromethane 	    3-13
3-7   Effects of dichloromethane on freshwater species in acute
        tests 	    3-25
4-1   Pulmonary absorption of DCM by human subjects (sedentary
        conditions) 	    4-5
4-2   Effect of exercise on physiological parameters for volunteers
        exposed to DCM	    4-7
4-3   Body burdens of rats after inhalation exposure to 14C-DCM for
        6 hours 	    4-9
4-4   DCM concentrations in rat whole blood and plasma at apparent
        steady-state conditions of a 6-hour inhalation exposure 	    4-11
4-5   Tissue concentrations of DCM in rats exposed to 200 ppm DCM for
        4 days for 6 hours dai ly 	    4-12
4-6   Tissue distribution of 14C-activity 48 hours after 6-hour
        inhalation exposure or oral dosage of rats to 14C-DCM 	    4-14
4-7   Comparison of postexposure pulmonary elimination half-times
        of DCM for humans and rats 	    4-16
4-8   Blood carboxyhemoglobin concentrations of rats exposed
        to CO and DCM by inhalation 	    4-22
4-9   Fate and disposition of 14C-DCM in rats (412-430 mg/kg),
        injected intraperitoneally 	    4-27
4-10  Fate of 14C-DCM in rats after a single 6-hour inhalation
        exposure 	    4-30
4-11  Body burdens and metabolized 14C-DCM in rats after inhalation
        exposure to 14C-DCM 	    4-30
4-12  Fate of DCM in rats 48 hours after single oral doses 	    4-32
4-13  lr\ vitro covalent binding of 14C-DCM to microsomal protein
        and lipid 	    4-42
4-14  Comparative covalent binding of DCM, carbon tetrachloride, and
        trichloroethylene to lipid and protein in rat hepatocytes 	    4-43
4-15  Blood COHb and Hb concentrations in rats exposed to DCM 	    4-52
5-1   COHb concentrations in nonsmokers exposed to DCM at
        250 ppm (869 mg/m3) for 5 days 	    5-6
5-2   Acute lethal toxicity of DCM 	    5-8
5-3   Summary of cardiotoxic action of 5% dichloromethane 	    5-11
5-4   Mutagenicity testing of DCM in bacteria 	    5-28
5-5   Gene mutations and mitotic recombination of yeast 	    5-38
5-6   Gene mutations in multicel lular eukaryotes jj} vivo 	    5-40
5-7   Gene mutations in mammalian cells in culture 	    5-44
5-8   Tests for chromosomal aberrations 	    5-46
5-9   Tests for sister-chromatid exchange 	    5-49
5-10  Analytical  analysis of methylene chloride 	    5-54
5-11  Cumulative percent mortality of rats 2-year methylene chloride
        inhalation study 	    5-56
5-12  Summary of total  tumor data for rats administered methylene
        chloride for 2 years by inhalation 	    5-60
                                      VI I
005DC1/E                                                               12/22/83

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

Table                                                                      Page

5-13  Time-to-tumor, palpable mass, and histopathology data for
        salivary gland region sarcomas in individual male rats exposed
        to methylene chloride by inhalation for 2 years 	     5-61
5-14  Summary of salivary gland region sarcoma incidence in male
        rats in a 2-year inhalation study with methylene chloride 	     5-63
5-15  Cumulative percent mortality of hamsters, 2-year methylene
        chloride i nhalati on study 	     5-64
5-16  Summary of total tumor data for hamsters administered methylene
        chloride by inhalation for 2 years 	     5-68
5-17  Monthly mortality data for male rats on a 2-year inhalation
        toxicity and oncogenicity study 	     5-73
5-18  Monthly ,nortality data for female rats on a 2-year inhalation
        toxicity and oncogenicity study 	     5-74
5-19  Non-neoplastic liver lesions in male rats 	     5-75
5-20  Non-neoplastic liver lesions in female rats 	     5-76
5-21  Summary of mammary gland tumors in female rats 	     5-77
5-22  Group assignment of Fischer 344 rats administered monthly
        chloride in deionized drinking water for 24 months 	     5-78
5-23  Mean daily consumption of methylene chloride in 24-month
        chronic toxicity and oncogenicity study in Fischer 344 rats ...     5-79
5-24  Incidencp of hepatocellular tumors in male and female Fischer
        344 rats administered methylene chloride in deionized drinking
        water for 104 weeks 	     5-81
5-25  Historical control data of liver neoplasia in female
        Fischer 344 rats at Hazelton Laboratories America, Inc	     5-82
5-26  Pulmonary tumor bioassay in Strain A mice 	     5-83
5-27  Methylene chloride analyses 	     5-86
5-28  Observed and expected deaths, 1964-1980, 1964 hourly male
        methylene chloride cohorts from Kodak 	     5-88
5-29  Malignant neoplasms, observed and expected deaths, 1964-1980,
        1964 hourly male methylene chloride cohorts from Kodak 	     5-89
5-30  Selected methylene chloride chronic animal studies 	     5-104
5-31  Incidence rates of salivary gland region sarcomas in male
        Sprague-Dawley rats in the Dow Chemical Company (1980)
        i nhalati on study 	     5-105
5-32  Relative carcinogenic potencies among 54 chemicals evaluated
        by the Carcinogen Assessment Group as suspect human
        carci nogens 	     5-110
 005DC1/E                              Vl"                               12/22/83

-------
                                LIST OF FIGURES

                                                                           Page

3-1   The effect of oxygen doping of the carrier gas on the ECD
        response to several halogenated methanes at a detector
        temperature of 300°C 	     3-21
4-1   Inspired and expired air concentrations during a 2-hour, 100-ppm
        inhalation exposure to DCM for a 70-kg man, and the kinetics
        of the subsequent pulmonary excretion 	     4-4
4-2   Plasma levels of DCM in rats during and after DCM exposure
        for 6 hours 	     4-8
4-3   DCM venous blood levels in rats immediately after a single
        6-hour inhalation exposure to various concentrations of DCM ...     4-10
4-4   Pulmonary elimination of 14C-DCM following oral administration
        to rats of a single dose of 1 or 50 mg/kg (squares) 	     4-18
4-5   Carboxyhemoglobin concentrations in male nonsmokers exposed to
        increasing concentrations of DCM for 1, 3, or 5 h day
        for 5 days	     4-23
4-6   Carboxyhemoglobin concentrations in rats after inhalation exposure
        to increasing concentrations of DCM for single exposures
        of 3 hours 	     4-24
4-7   Blood CO content of rats after 3-hour inhalation exposure with
        1000 ppm dichloromethane, dibromomethane, and diiodomethane,
        respectively 	     4-25
4-8   Rates of production of CO from DCM given to rats 	     4-28
4-9   Enzyme pathways of the hepatic biotransformation of
        dihalomethanes 	     4-35
4-10  Proposed reaction mechanisms for the metabolism of
        dihalomethanes to CO, formaldehyde, formic acid, and inorganic
        halide 	     4-36
4-11  Blood COHb level in men during an 8-hour exposure for 5 conse-
        cutive days to 500 ppm and 100 ppm DCM 	     4-46
4-12  Blood COHb concentrations in rats during and after a 6-hour
        inhalation exposure to DCM 	     4-48
5-1   Histogram representing the frequency distribution of the
        potency indices of 54 suspect carcinogens evaluated by the
        Carcinogen Assessment Group 	     5-109
005DC1/E                                                               12/22/83

-------
                            AUTHORS AND REVIEWERS
 Stephen P. Bayard, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.

 David  L. Bayliss, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.

 I.W.F. Davidson, Department of Physiology and Pharmacology, Bowman Gray
     School of Medicine, Winston-Salem, N.C.

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

 Mark Greenberg, Environmental Criteria and  Assessment Office, U.S.
     Environmental Protection Agency, Research Triangle Park, N.C.

 Bernard H. Haberman, Carcinogen Assessment  Group,  U.S. Environmental
     Protection Agency, Washington, D.C.

 Jean C. Parker, Office of  Solid Waste, U.S.  Environmental  Protection  Agency,
     Washington, D.C.

 Dharm  Singh, Carcinogen Assessment Group, U.S. Environmental Protection
     Agency, Washington, D.C.
005DC1/E                                                               12/22/83

-------
The following individuals were asked to review an early draft of this document
and submit comments:

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

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

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

Dr. I. W. F. Davidson
Dept.  of Physiology/Pharmacology
The Bowman Gray School of Medicine
Winston-Salem, NC  27103

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

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

All Members of the
Interagency Regulatory Liaison Group
Subcommittee on Organic Solvents
                                       XI
005DC1/E                                                                12/22/83

-------
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, Ph.D.
Chao W. Chen, Ph.D.
Margaret M.L. Chu, Ph.D.
        Herman J. Gibb,  B.S., M.P.H.
        Bernard H. Haberman, D.V.M., M.S.
        Charalingayya B. Hiremath, Ph.D.
        Robert E. McGaughy, Ph.D.
        Dharm V. Singh,  D.V.M., Ph.D.
        Todd W. Thorslund, Sc.D.
Participating Members of the Reproductive Effects Assessment Group

John R. Fowle III, Ph.D.
Ernest R. Jackson, M.S.
Casey Jason, M.D.
K.S. Lavappa, Ph.D.
Sheila L. Rosenthal, Ph.D.
Carol N. Sakai, Ph.D.
Daniel S. Straus, Ph.D., Consultant
Vicki Vaughan-Dellarco, Ph.D.
Peter E. Voytek, Ph.D.
Gary M. Williams, M.D., Consultant

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

Members of the Agency Work Group on Solvents
Elizabeth L. Anderson
Charles H.' Ris
Jean C. Parker
Mark M. Greenberg
Cynthia Sonich
Steve Lutkerihoff
James A. Stewart
Paul Price
William Lappenbush
High Spitzer
David R. Patrick
Lois Jacob
Arnold Edelman
Josephine Brecher
Mick Ruggiero
Jan Jablonski
Charles Delos
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 Drinking Water
Consumer Product Safety Commission
Office of Air Quality Planning and Standards
Office of General Enforcement
Office of Toxics Integration
Office of Water Regulations and Standards
Office of Water Regulations and Standards
Office of Solid Waste
Office of Water Regulations and Standards
Office of Pesticide Programs
Office of Emergency and Remedial Response
005DC1/E
                                      XI I
                                                                       12/22/83

-------
The following individuals attended a review workshop to discuss draft EPA
documents on organic compounds which included an early draft of this
document:

Dr. Mildred Christian
Argus Laboratories
Perkasie, PA  18944

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

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

Dr. Herbert Cornish
School of Public Health
University of Michigan
Ann Arbor, MI  48197

Dr. I. W. F. Davidson
Dept.  of Physiology/Pharmacology
The Bowman Gray School of Medicine
Winston-Sal em, NC  27103

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

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

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

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

Dr. James Withey
Food Directorate
Bureau of Food Chemistry
Ottawa, Canada
                                      XIII
005DC1/E                                                                12/22/83

-------
                          1.   SUMMARY AND CONCLUSIONS

     DichVoromethane (methylene  chloride,  DCM)  is a solvent  that  is  widely
used for a variety of purposes.   Annual  production of DCM in the United States
is about 269,000 metric tons.   Approximately 85 percent of the DCM consumed in
the United States  is lost directly to the environment through dispersive use,
largely by evaporation to the atmosphere.  Natural sources have been proposed
for DCM but none is believed to contribute significantly to ambient  concentra-
tions.   Although ambient air  and water measurements are  rather scarce, they
indicate that DCM  is  found  in a variety of urban and non-urban areas of the
United States and  in  other  regions of the world.   The background atmospheric
concentration is about 35 parts-per-trillion (ppt).  Concentrations of DCM  in
urban areas could  be  one  or two orders of magnitude higher than background.
DCM is not expected to accumulate in the atmosphere; estimates of half-life
vary from 20 days to 1 year.   Hydroxyl radical  attack is probably sufficiently
rapid to prevent most,  if not all,  DCM from reaching the stratosphere.
     DCM has been  detected in both natural (fresh and seawater) and municipal
waters in various  geographical areas  of the United States.  Concentrations  of
dichloromethane  in  the low parts-per-billion  range have  been  measured in
surface water and  drinking  water.   DCM does not  appear to be formed to any
large extent during the chlorination process.   DCM's short evaporation half-life
from moving water  probably  allows  most of the compound dissolved in water to
be transported  into the atmosphere.   DCM also is  readily  degraded by bacteria
in concentrations  up to 400 parts-per-million (ppm).  The  extent to which DCM
enters groundwater  from surface  waters  is unknown.   Some DCM is deposited in
landfills.   Where  leaching  is possible, the compound may enter groundwater
systems because  it does  not easily adsorb to  clay,  limestone,  and/or peat
moss, and retention in the soil  is unlikely.  There  is  no  evidence for  signi-
ficant bioaccumulation of DCM in the food chain.   Data concerning the ecological
consequences of DCM in the  environment indicate that it is biodegradable under
both aerobic and anaerobic  conditions.   DCM appears to have  low toxicity to
aquatic organisms.
 Ambient is a term used in this and other documents of this nature to
 refer to the  environment  and should not be construed to  include  indoor and
 occupational settings.

005DC1/A                             1-1                              12/22/83

-------
     As with other  solvents  of  this class,  inhalation of DCM in air followed
by lung absorption  is  the  most  rapid route of  entrance  into the body.   DCM
also is well absorbed  into the body after oral ingestion.  Absorption through
the intact skin occurs to some extent,  but is  relatively  a much slower process.
DCM is appreciably more water-soluble and less 1ipid-soluble than its congeners,
carbon tetrachloride  and chloroform.  Because of  DCM's  high solubility in
water and  lipids,  it  is probably distributed throughout all body fluids and
tissues.   DCM's half-time  of  elimination from adipose tissue (6 to 6% hours)
is consistent with reports  that  DCM can be found in such  tissue 24 hours after
both single and chronic exposures.   DCM readily crosses the blood-brain barrier,
as evidenced  by its  narcotic effect at higher exposure  levels.   DCM also
probably crosses the  placenta and  distributes into the developing fetus, but
studies in experimental animals  indicate that effects on  the fetus are unlikely
at levels commonly experienced by humans.
     Following ingestion,  the absorption of DCM is virtually complete.   When
ambient air exposure  occurs,  the amount of DCM absorbed  increases  in direct
proportion to its concentration  in  inspired air,  the duration of exposure,  and
physical  activity.   Absorbed DCM is eliminated primarily  by pulmonary excretion
of the  unaltered parent compound  (about 85 percent).  About  2 percent is
excreted unchanged  in  the  urine, while a small amount is eliminated by other
routes.  Ten percent  or less  is  metabolized before  elimination  from the  body.
At very  low concentrations, essentially  all of the  compound  absorbed into  the
body may  be metabolized.  DCM is  known  to be metabolized by  the liver to
carbon monoxide (CO).   Carboxyhemoglobin (COHb) is formed from the  interaction
of CO  and  hemoglobin;  CO dissociates at  the  lung  and is eliminated.  Blood
normally contains about 0.5 COHb at all times; therefore, CO formed  from a few
ppt or ppb of  DCM vapor is of no  practical consequence.   However,  at higher
exposure levels (up to 500 ppm and above), the COHb level would be  expected to
reach  a  maximum  between 12 and   15  percent  in  man.   This  level  is below that
considered hazardous  for normally  healthy individuals, but could place addi-
tional stress on people with compromised cardiovascular systems.
     The effects  of COHb from DCM  metabolism  are additive to COHb formed from
exogenous  CO.   The  toxicities of CO and  DCM  can  differ markedly.   However,
persons exposed  to  levels  of DCM  that  do not exceed the  U.S.  Occupational
Safety and Health  Administration (OSHA) standard of 500 ppm (1737  mg/m3)  may
have blood COHb  levels that exceed  those  allowable from direct CO exposure.

005DC1/A                             1-2                              12/22/83

-------
     The results of animal  experimentation  by several  investigators indicate
that carbon dioxide (C02),  formaldehyde, and formic acid are additional meta-
bolites of DCM.   At least two pathways exist in rat liver for the metabolism of
DCM.  Together,  the microsome oxidative dehalogenation  and cytosol glutathione
transferase dehalogenation  systems could account for the CO and C02 generated
from the metabolism of  DCM.   Neither the microsomal system  nor  the cytosol
system  is  inducible by microsomal inducers,  such  as  phenobarbital or DCM.
     The weight  of evidence  from the available literature  indicates that
adverse toxicologic effects  (other  than  carcinogenicity and mutagenicity) in
humans are unlikely to occur at ambient air and water levels found or expected
in  the  general  environment or even at higher  levels  sometimes observed  in
urban areas.  In  fact,  available experimental data do  not indicate that  any
adverse toxicologic effects  are  induced  in humans  at a threshold limit value
(TLV®) of 100 ppm (347 mg/m3) DCM.  The potential  direct adverse  health effect
associated with exposure  levels  that greatly exceed 100  ppm  (347 mg/m3) is
primarily neurological.  The  lowest concentration  reported to  affect  eye-hand
coordination was 200 ppm (694 mg/m3).
     Liver and  kidney damage  is  unlikely to occur  as a  result  of  DCM  exposure
at  environmental  levels.   Hepatotoxicity  has  not been  reported  in  any human
case report, even  following  fatal overexposure.  Only minimal  hepatic changes
were observed in  animal  studies, even at doses ranging from the  average  LD50
(about 2 g/kg)  to near-lethal levels.  Animal studies  also indicate  that DCM
has low nephrotoxicity.
     Direct cardiac effects  of DCM in  humans  also  are unlikely because of the
low levels  of DCM found in the  environment.  Animal studies  have shown  that
acute exposure  levels exceeding  20,000 ppm  (69,480  mg/m3) are  required before
a decrease in myocardial contractility and other effects on cardiac performance
are observed.
     Available evidence hints that the teratogenic potential of DCM in experi-
mental animals is minimal.  However, a definitive assessment of such  potential
in humans can be made only after further testing in appropriate rodent species,
in accordance with current teratologic and reproductive testing methodologies.
     The weight of evidence with respect to mutagenic potential shows that DCM
is  capable  of causing  gene mutations and has the  potential  to  cause such
effects  in  exposed human  cells.   Positive responses to  DCM exposure were
observed in four  different microorganisms.   However, additional  studies are
needed to determine the strength of evidence for mammalian systems.

005DC1/A                             1-3                              12/22/83

-------
     The evidence  for  the  carcinogenic potential of DCM  is  based upon six
chronic (lifetime) studies in which  DCM was administered to rodent species.
Four of the studies involved  rats,  one involved mice,  and  one involved hamsters.
Chronic inhalation studies with  rats  and hamsters were conducted by the Dow
Chemical Company.  One  rat study (Dow Chemical Company,  1980)  showed  a small
increase in the number of benign  mammary tumors compared  to  controls in female
rats at all doses and in male rats  at the highest dose,  as well  as a statisti-
cally significant  increased  incidence  of ventral cervical sarcomas, probably
of salivary gland  origin  in  male rats.  In hamsters, there was an  increased
incidence of lymphosarcoma in females only;  this increase  was not statistically
significant after  correction for survival.  In the second inhalation study  in
rats by Dow Chemical  Company (1982)  there were no compound-related increased
incidences of any  tumor type, but the  highest  dose was appreciably  lower than
previously employed.   A borderline hepatocellular neoplastic nodule response
in female  Fischer  344  rats was  observed only  in the  chronic drinking water
study, which was conducted under  the  auspices  of the National Coffee Association.
However, while  the response was  significant with  respect  to matched controls,
the incidence was  within  the range of  historical control values  at the per-
forming laboratory.
     Two epidemiological  studies that  focused on DCM have  been reported.
Although neither study showed excessive risk to subjects,  there were sufficient
deficiencies to  prevent the  studies  from being  judged as showing no effect.
     The weight of evidence for  carcinogenicity in animals is limited, according
to the  criteria of the International Agency for  Research on Cancer (IARC).
This conclusion is based upon the statistically positive salivary gland sarcoma
response in male  rats  (Dow Chemical, 1980) and the borderline hepatocellular
neoplastic  nodule  response  in  the rat  (National  Coffee  Association,  1982).
The occurrence  of limited carcinogenic  activity  is  reason  to suggest  that
additional testing is warranted  to adequately clarify the carcinogenic potential.
When the absence  of  epidemiological  evidence  is  considered  with  the  limited
animal evidence, the  overall  evaluation of DCM, according to IARC criteria, is
a Group 3  chemical in  that it cannot  be classified as  to its carcinogenic
potential for humans.
     However, note must be made  of the aggregate  findings for both mutagenicity
and carcinogenicity  as  a  singular adverse response.  A  positive finding of
gene mutations  is  usually taken  as support for  the  likelihood of a chemical
having a carcinogenic potential.   In this case, the direct carcinogenic evidence
005DC1/A                             1-4                              12/22/83

-------
for animals is limited using IARC terminology; however, the positive mutageni-
city findings should be viewed as increasing the concern about the adequacy of
carcinogenicity testing  to date.  The mutagenicity  findings,  on the other
hand, when considered in their own right, are a qualitative finding for potential
human adverse  effect although further studies  are  required to clarify the
potential  to damage genetic material  in man.
RECOMMENDATIONS FOR FURTHER STUDIES
     It is apparent  that  further research  is  needed in several areas.  This
list does  not indicate research priorities.

     (1)  Teratogenicity and  Reproductive Effects.  To conclusively determine
          the teratogenic  potential  of  DCM with respect  to  humans,  it is
          desirable to conduct more  testing on appropriate rodent species at
          various exposure levels.
     (2)  Mutagenicity.  Additional tests for chromosomal  aberrations should
          be conducted.
     (3)  Pharmacokinetics.  To  more  fully  define the pharmacokinetics so as
to explain the carcinogenic responses or lack thereof and to clarify questions
and hypothesis regarding mutagenicity.
005DC1/A                             1-5                              12/22/83

-------
                               2.  INTRODUCTION
     Dichloromethane  (DCM)  is  a high-volume industrial chemical widely  used
to remove paint, clean metal, and propel aerosol sprays.  This document provides
an evaluation of the health effects of DCM and a review of the relevant available
scientific  literature.   To  provide  a  perspective  in  evaluating health  effects
associated with DCM, this document contains background chapters relating to ana-
lytical methodologies, production, sources and emissions, and ambient  air con-
centrations.
     DCM  is  released  into the ambient air  as  a result  of evaporation  during
its production, storage, and manufacturing or during general consumer use.   It
is believed to be derived from natural sources, but such formation is not believed
to contribute significantly to global  concentrations.
     Information on the effects of DCM has been derived primarily from studies
involving individuals  exposed  occupationally or accidentally  to DCM.   In such
exposures, the  concentrations  of  DCM  associated with adverse  effects to  human
health were  either  unknown  or  greatly exceeded the concentrations measured  in
ambient air.  Controlled exposure studies have established that vapor inhalation
is the principal route by which DCM enters the body.   DCM is eliminated from the
body primarily as the parent compound via the lungs.   DCM is metabolized in ani-
mals and in humans to carbon monoxide (CO), which elevates the carboxyhemoglobin
content of  blood.   The  distribution,  storage,  and metabolism  of DCM to CO and
carbon dioxide  (CO;,)  help explain its effects upon  humans.   Epidemiological
studies provide some information about the impact of DCM on human health, but it
is necessary to rely on animal studies to assess any indications of potentially
harmful effects for chronic low-level  exposures.
     The primary concerns regarding the potential  impact of DCM on human health
are narcosis effects associated with acute high-level exposures and any mutagenic
or   carcinogenic   effects   potentially   associated   with   chronic
low-level  exposures.
 Ambient is used  in  this and other documents of this nature to  refer  to  the
 environment and  should  not  be  construed to include indoor and  occupational
 settings.
005DC1/B                                  2-1                         11-10-83

-------
     The U.S.  Occupational  Safety and Health Administration (OSHA) health stan-

dard requires that a worker's exposure to DCM at no time exceed a TWAa of 500 ppm
                                                                  (R)
in the workplace air in any work shift of a 40-hour week.   The TLV  for an 8-hour

TWA concentration is 100 ppm with a 15-minute allowable excursion for an average

concentration of 500 ppm.
 TWA (time-weighted average)  is  defined as the time-weighted average concen-
 tration for a  normal  8-hour workday and a 40-hour workweek to which  nearly
 all workers may be exposed repeatedly, day after day, without adverse effect.
005DC1/B                                  2-2                         11-10-83

-------
                 3.  DICHLOROMETHANE:  BACKGROUND INFORMATION

3.1  PHYSICAL AND CHEMICAL PROPERTIES
     Dichloromethane (methylene  chloride,  DCM)  is one member  of  a  family  of
saturated aliphatic halogenated compounds.  Other common names or synonyms are
shown in Table 3-1.  The Chemical Abstracts Service Registry Number for DCM is
000075092.  Dichloromethane is a colorless, nonflammable, volatile liquid that is
completely miscible with a variety of other solvents (Anthony, 1979).   The impor-
tant physical  properties  of  DCM are shown in Table 3-2.   For example, DCM is
highly  volatile  (vapor  pressure  of  350 torr at  room temperature).   Hence,  the
most common mode of entry into the body is by inhalation.  The ambient air con-
centration for compounds of this nature is often expressed in parts-per-million
(ppm),  parts-per-billion (ppb), and parts-per-trillion (ppt).  At standard tem-
perature and pressure, 1 ppm is equivalent to 3.474 mg/m3.
     In the  absence  of  moisture at ordinary  temperatures, DCM is relatively
stable  when  compared with  its  congeners,  chloroform and  carbon tetrachloride.
In dry  air, DCM decomposes at temperatures exceeding 120°C (Anthony, 1979).  At
elevated temperatures (300°  to 450°C),  DCM tends to carbonize when its vapor
contacts steel and metal chlorides.   Moisture initiates hydrolysis of DCM, pre-
dominantly to hydrogen chloride (HC1), with trace amounts of phosgene (Anthony,
1979).   However, even trace amounts of phosgene detract from the paint-stripping
qualities of DCM (De Forest,  1979).   To retard production of phosgene via hydro-
lysis,   inhibitors are generally added to commercial  preparations of DCM (De Forest,
1979).   Protection against hydrolysis also is attained by the addition of phenolic
compounds (Anthony, 1979).
     Although anhydrous  DCM is noncorrosive to common metals, the HC1  produced
from hydrolysis  of DCM  initiates  corrosive action with aluminum and iron
(Anthony, 1979).  The addition of epoxides to consume the HC1 affords protection
against this corrosion (Anthony, 1979).   To minimize the decomposition of DCM,
storage containers should be  galvanized or lined with a phenolic coating (Anthony,
1979).   Commercial grades of DCM contain  a variety of  stabilizers to minimize
decomposition (McKetta and Cunningham, 1979).   Cyclohexane, thymol,  hydroquinone,
p-cresol, and low-boiling amines have been used as stabilizers (De Forest, 1979).
Aerosol  preparations containing DCM often use propylene oxide as a stabilizer.
For preparations designed to  be used in degreasing applications, special inhibitor
mixtures are used.   One  such  mixture includes propylene oxide, butylene oxide,
cyclohexane, and N-methyl  morpholine (De Forest, 1979).
005DC1/C                                  3-1                          12/7/83

-------
           TABLE 3-1.   SYNONYMS AND IDENTIFIERS FOR DICHLOROMETHANE

Chemical Abstracts Service registry number:   000075092
Chemical formula:         CH2C12
Structural formula:                      Cl
                                   H -  C  - H
                                        Cl
Synonyms:
 Dichloromethane
 Methylene dichloride
 Methylene bichloride
 Methylene chloride
     The experimentally determined  average  evaporative half-life of DCM from
water is 18 to 25 minutes (Oil ling, 1977).  In three separate experiments, Dill ing
used solutions of an average depth of 6.5 cm, containing approximately 1 ppm DCM.
The solutions were stirred at 200 rpm in a 250-ml beaker.  These experimentally
determined half-lives agreed with the value (20.7 minutes) obtained by the fol-
lowing formula:

                              t,  _ 0.06391d                           (3-1)
                               *-    ki

where d is the solution depth and kx is the liquid exchange constant (cm/min).
The formula  is an adaptation of the common  equation for the half-life of a
substance undergoing a first-order reaction.  The calculated evaporative half-
life may not be accurate for DCM in natural aquatic  systems.
005DC1/C                                   3-2                           12/7/83

-------
              TABLE 3-2.  SELECTED PROPERTIES OF DICHLOROMETHANE
Molecular formula
Formula weight
Boiling point (760 mm Hg)
Melting point
Vapor density
Density of saturated vapor
Density
Solubility
Explosive limits in oxygen
Flash point
Autoignition temperature
Relative evaporation rate

Vapor pressure
Conversion factors
 (25°C; 760 mm Hg)
Concentration in saturated air
Log octane/water partition
        CH2C12
        84.94
        40°C (760 mm Hg)
        -95 to -97°C
        2.93 (air = 1)
        2.06 (air = 1)
        1.326 g/ml (20°C)
        2.0 g/100 ml water at 20°C;
        soluble in ethanol,
        ethyl ether, acetone,
        and carbon disulfide
15.5 to 67% by vol
None
624° to 662°C
14 (water = 1)
71 (ether = 100)
Temp °F Temp °C
50 10
68 20
77 25
86 30
95 35
ume



mm Hg
230
349
436
511
600
     1 mg/1 = 1 g/m3 = 288 ppm
     1 ppm = 3.474 mg/m3 = 3.474 ng/1
     550,000 ppm (25°C)
     1.25
Hardie, 1969.
Anthony, 1979.
Weast, 1969.
 Dilling (1977) reported values from literature ranging from 19,400 ppm at 25°C
 to 22,700 ppm at 1.5°C.
American National Standards Institute Inc.,  1970.
Christensen and Luginbyhl, 1974.
bDe Forest, 1979.
005DC1/C
3-3
11-10-83

-------
3.2  ENVIRONMENTAL FATE AND TRANSPORT
     Dichloromethane is principally  used  as an aerosol propellant,  degreasing
solvent, paint  stripper,  and  thinner in paints and lacquers.   Because of its
volatility and  dispersive  use pattern,  much of the DCM produced worldwide is
emitted into the atmosphere.  Almost all of the emissions are from anthropogenic
sources.  DCM also is formed from natural  sources, but natural sources are not
believed to contribute significantly to global concentration (National Academy
of Sciences, 1978).
3.2.1  Production
     Dichloromethane is produced commercially in the United States, predominantly
via the following reaction (De Forest, 1979; Anthony, 1979):
                            FeCl3 or ZnCl
               HC1 + CH.OH  13Qu , 18Qoc *  CH3C1 + H20
                                                                      \3~c.)
               CHaCl + Clz  	»  CHi;Clz + HC1.

     In this vapor  phase  reaction sequence, yields of 95 percent  are usual.
Dimethyl ether is the secondary byproduct of the hydrochlorination.
     A less common method used to obtain DCM is direct reaction of methane with
chlorine at 485^ to 510«*C (Anthony, 1979).  Methyl chloride, chloroform, carbon
tetrachloride,  and HC1  are coproducts.   However, both reactions are used in the
chemical manufacturing industry so that the HC1 can be recycled (Anthony, 1979).
     In the liquid  phase,  DCM can be produced by  refluxing and distilling a
mixture containing methanol, HC1, and zinc chloride at 100° to 150°C.  However,
this method is  not widely used (Anthony, 1979).
     According to one source,  production in the United States is carried out by
five major companies at seven  sites  (Table 3-3).   However,  the U.S.  EPA  Toxic
Substances Control Act Public  Inventory showed that in 1977 there were six manu-
facturers and 13 importers (U.S. EPA, 1980b).
     According to statistics gathered by the U.S. International Trade Commission,
the United States annual  production of DCM was estimated to be 256,000 metric tons
in 1980 (U.S.  International Trade Commission,  1980) and 269,000 metric tons in
1981 (U.S.  International  Trade Commission, 1982).
     Edwards et al. (1982) estimated 1981 world production at 825,000 metric tons.
005DC1/C                                  3-4                          12/7/83

-------
                   TABLE 3-3.   PRODUCERS OF DICHLOROMETHANE
                                                     Annual  capacity as  of
                                                        January 1,  1979
Company
Allied Chemical
Diamond Shamrock
Dow Chemical
Dow Chemical
Stauffer Chemical
Vulcan Materials
Vulcan Materials
Location
Moundsville, WV
Belle, WV
Freeport, TX
Plaquemine, LA
Louisville, KY
Le Moyne, AL
Geismar, LA
Wichita, KA
(metric tons x 103)
23
50
92
88
28
37
60
Source:  Chemical Marketing Reporter,  August 6,  1979,  p.  9.

3.2.2  Use
     Dichloromethane is used  as a paint remover,  a urethane foam-blowing agent,
a vapor degreasing  and dip solvent for metal cleaning,  a solvent for aerosol
products,  a solvent in the pharmaceutical  industry, a  solvent in the manufacture
of polycarbonates by polymerization, and an extractant for caffeine, spices, and
hops.  DCM is used in the manufacture  of plastics, textiles, photographic film,
and photoresistant coatings, as a solvent carrier  in the manufacture of herbi-
cides and  insecticides,  and as a component  of rapid-drying paints  and adhe-
sives, carbon removers,  and brush cleaners.  Other minor applications include
use as a  low-pressure  refrigerant,  as a low-temperature heat transfer medium,
and as an air-conditioning coolant (Ahlstrom and Steele, 1979).   Distribution of
DCM by major use is shown in Table 3-4.   One of  the fastest growing segments of
the DCM market is the aerosol  sector because DCM is being substituted for some
chlorofluorocarbons as a solvent, vapor pressure depressant, and flame retardant.
Consumption by the aerosol industry is expected  to grow (Chemical and Engineering
News, 1982).
005DC1/C                                  3-5                          12/7/83

-------
                  TABLE 3-4.   CONSUMPTION OF DICHLOROMETHANE
Use
Paint remover
Metal degreasing agent
Aerosol propel! ant
Blowing agent for foams
(urethane)
Exports
Other
1980 metric tons, 103
1980
73
49
46
20
44
12
Total
1977
30
20
19
8
18
5
metric tons, %
1978
29
18
21
9
15
8
Source:  SRI International,  Chemical  Economics Handbook,  1980.

     Dichloromethane is  expected to retain popularity  as  a paint remover.
Although DCM competes with trichloroethylene and perchloroethylene as a solvent,
it is preferred as a paint remover because of better performance (Lowenheim and
Moran, 1975).
3.2.3  Emissions
     Emissions from dispersive uses are the major source of DCM in the environ-
ment.  Of the total amount of DCM produced in the United States, approximately
85 percent is estimated to be lost into the environment through sewage treatment
plants and surface waters, deposited on land,  or lost to the atmosphere.   The actual
losses during production, transport, and storage are not well documented, but it
appears that such losses represent only a very small percent of the DCM entering
the environment from product manufacture and use.  Most of the losses in production,
transportation, and  storage  are  fugitive (that  is, transient)  releases  from
leaky pump  seals,  valves,  and joints.   The dispersive uses of DCM are varied
and widespread and are distributed geographically approximately with the industri-
alized population in the United States.  Although most of the losses are to the
atmosphere, DCM  is  relatively soluble  in water.  The most effective means of
removal of DCM  from water is air stripping, which transfers the chemical from
water to the atmosphere (National Academy of Sciences, 1978).
005DC1/C                                  3-6                          12/7/83

-------
3.2.4  Persistence of PCM
3.2.4.1  Atmospheric Degradation—Reaction with hydroxyl  radicals  (OH*)  is  the
principal process by which many organic chemicals,  including DCM,  are scavenged
from the  troposphere (Crutzen and Fishman,  1977;  Singh,  1977; Altshuller,
1980).   These radicals are produced by irradiating  ozone  (03),  and the resultant
singlet oxygen atoms  then  react  with water vapor.   The tropospheric lifetime
of  a  compound can  be  related  to the OH- concentration by  the expression:

                         1 lifetime  =     l
                                         k [OH]                       (3.3)
where k is the rate constant of reaction.

     Available evidence indicates that the lifetime of DCM in the troposphere
is less than  1 year.   The modelling approaches of Crutzen and Fishman (1977)
and Singh  (1977)  indicate  that the range of the average OH concentration is
between 2  x 10? and 6 x 10? molecules/cm3.  Using an average concentration of
3 x 10? molecules cm3,  Altshuller (1980) calculated a tropospheric lifetime
for DCM of 1.4 years.   The rate constant expression of Davis et al.  (1976) and
a tropospheric temperature of 265K were used.
     Singh et  al.  (1979)  computed a 1-year lifetime  for  DCM  using the rate
data  reported  by the National  Aeroneutics  and Space Administration  (NASA,
1977) and  the  National  Bureau of Standards (1978) and a temperature of 265K.
An average OH concentration  of  4 x 10? molecules/cm3 was  employed in the
computation.    More  recently,  Singh et al. (1983)  calculated  an atmospheric
residence  time of 0.9  ± 0.3 year, based  upon a comparison of atmospheric
budget data with  available emissions  data.   Since the mean concentrations in
the northern  hemisphere are 1.8  times the southern hemisphere values,  a short
lifetime for DCM is indicated.
     However, Cox et al.  (1976) calculated a 0.3-year lifetime in photokinetic
studies in which  DCM  competed with nitrous acid as the target  of OH  attack.
The lifetime was  derived  from a rate constant of 10.4 x 10 14 cnrVmolecule •
sec at 298K.   An average OH concentration of 1 x 10e molecules/cm3 was assumed.
Use of 4 x 10s molecules/cm3 for the OH concentration would have resulted in a
calculated lifetime  of 0.76  year;  this  value is closer to those  of Singh
(1977) and Altshuller  (1980).   Davis  et al.  (1976) calculated  a lifetime of


005DC1/C                                  3-7                          12/7/83

-------
0.39 year from a rate constant of 8.7 x 1014 at 265K and an average OH concen-
tration of 9 x 105 molecules/cm3.
     Determination of the  rate  constant for the reaction of OH with DCM has
been the  focus  of various  investigations (Davis et  al.,  1976;  Cox et al.,
1976; Howard and Evenson, 1976;  Perry et al.,  1976).   The values are in general
agreement (Table 3-5) with the exception of Butler et al.  (1978).
3.2.4.2  Aquatic Biodegradation—Recent evidence  indicates  that DCM is bio-
degradable under both aerobic and anaerobic conditions.   Brunner and Leisinger
(1978) first reported the isolation of facultative methylotroph with the ability
to utilize DCM as a sole carbon source for growth.   The  organism was tentatively
identified as a Pseudomonas species.   Brunner et al.  (1980) observed that utili-
zation of DCM  is  caused  by an inducible enzyme, the  activity of which  is only
partially inhibited under anaerobic conditions.
     Rittman and  McCarty (1980)  investigated  sewage microorganisms for their
ability to  biodegrade DCM.   Bacterial  cultures were enriched from  a seed of
primary sewage effluent over a 12-month period.  DCM supported bacterial growth
when the mineral salts culture medium was supplemented with sodium bicarbonate.
     Klecka (1982) investigated the fate and effects of  DCM in a system simula-
ting a municipal  waste  water treatment plant.   After acclimation, the sludge
was  used  in closed-bottle  respirometer studies.  Respirometer studies demon-
strated the  disappearance  of 10 mg DCM/1 (14C-label led) within 4 hours, with
49 percent of the parent compound recovered as 14C~labelled C02 after 50 hours.
At 1 mg DCM/1, disappearance resulted within 3 hours, with a 66-percent conver-
sion of the parent compound to C02 after 50 hours.   Rate constants, calculated
for  the biodegradation  of  1,  10,  and 100  mg/1  by activated  sludge,  were 1.42,
1.61, and 0.35/hour, respectively.
     In  inhibition  studies on the effect  of  DCM  on oxygen consumption and
organic carbon degradation by activated sludge, Klecka (1982) found no signifi-
cant effect over a 24-hour period.  Modelling of a continuously mixed activated
sludge reactor  indicated that the rate of  biodegradation was  about 12 times
greater than the rate of volatilization.
     Wood et  al.  (1981) demonstrated the  degradation of DCM  under anaerobic
conditions,  using  sediment-water  samples  spiked with 200 ug DCM.   Degradation
was  observed to proceed  via methyl chloride, although accumulation  in the sam-
ples was not observed.
005DC1/C                                  3-8                          12/7/83

-------
                                       TABLE 3-5.   REACTION RATE DATA FOR OH +
(k x 10 14cmVmolecule •  sec)      Temperature,  K
                                                              Arrhenius expression
                                                                                               Reference
CO
14.5 ±2.0

15.5 ± 3.4

11.6 ± 0.5


 8.7

10.4

 2.7 ± 1
                                       298.5

                                       296

                                   245 to 375


                                       265

                                       298

                                       302
                                                         4.27 ± 0.63 x 10"12
                                                          exp [-(1094 ± 81/T)]
                                                         5.2 x 10~1*exp(-1094/T)
Perry et al.,  1976

Howard and Evenson, 1976

Davis et al.,  1976




Cox et al., 1976

Butler et  al., 1978

NASA,3 1977; National
 Bureau of Standards,
 1978
   NASA preferred value; reliability of log k judged to be ± 0.2 at 230K.

-------
3.2.5  Products of PCM
3.2.5.1  Atmospheric Simulation Studies—Pilling et al.  (1976) observed that DCM
was not  very  reactive  in a chamber atmosphere containing nitric oxide  (NO) or
nitrogen dioxide (N02).  Ozone-air mixtures containing 10 ppm (34.7 mg/m ) DCM
and 5  ppm  NO  or 16.8 ppm NO^ were exposed  to ultraviolet (UV) radiation at an
intensity  about 2.6 times that of natural  sunlight at noon on a summer day in
Freeport,  Texas.  After 21 hours of exposure in the presence of NO,  less  than
5 percent  of  the  DCM  had  disappeared.   Similarly, less  than  5  percent dis-
appeared in an N02 atmosphere after a 7.5-hour exposure.   The effect of varying
UV intensity or concentration on the rate of photodecomposition  was not investi-
gated.    Relative  humidity  in the photolysis reactor was  35  to  40 percent.
Further  investigation  of this  simulated trospheric reaction  showed anomalous
behavior, possibly occurring on the chamber walls  (Dilling and Goersch, 1980).
Dichloromethane was judged to contribute to oxidant formation less than other
halogenated compounds investigated, e.g., tetrachloroethylene, trichloroethylene,
and vinyl chloride.
     Butler et al. (1978)  proposed that the OH attack on DCM  in  the presence of
02 may result in formation  of phosgene (COC12) via the  reaction sequence
                         CH2C12 + OH -> -CHCl^ + H20
                         •CHC12 + 02 •* -CHC1202                       (3-4)
                         •CHC1202 •* COC12 + OH.

This pathway was suggested to account for a low rate constant of the reaction of
OH with DCM in the presence and absence of CO in the test atmosphere.  Production
of C02 was followed.
     Chlorine-sensitized photooxidation of DCM in the presence of C12 in dry air
resulted in CO and CO^ as the major carbon-containing products (Spence et al . ,
1976).  After  5 minutes  of  irradiation of  20 ppm DCM and 5 ppm C12  in air,  19
ppm CHi-Cl^, was  consumed.   The product distribution was  CO  (5 ppm); HC1 (38
ppm); phosgene (2 ppm); formyl chloride (0 ppm); and C02 (12 ppm).  The product
distribution is illustrative of a chain reaction:

                              •CHC1202 •* CIO + HCOC1
                               HCOC1 + Cl -» HC1 + COC1

005DC1/C                                  3-10                         12/7/83

-------
                               COC1 + 02 -» C02 + CIO                  (3-5)
                               COC1 •* CO + Cl

Chlorine-sensitized photooxidation  of  DCM is not expected  to be  significant
under real atmospheric conditions because Cl will react with species other than
halocarbons (Spence et al., 1976).
3.2.5.2   Hydrolysis—The  hydrolysis of  DCM  in natural waters is influenced by
acidic and  basic  conditions.   When  DCM  is hydrolyzed  in water at  temperatures
ranging from 80° to 150°C, the hydrolysis in an acidic solution was reported to
proceed at a rate corresponding to  a measured half-life of 13.75 days (Fells and
Moelwyn-Hughes, 1958).
     Radding et al.  (1977) reported a maximum hydrolytic half-life of 704 years
(100 to  150°C  at  pH 7);  this  is  in sharp contrast  to the aqueous  reactivity
results found by Oil ling et al. (1975) in which the half-life was about 18 months
(25°C).
3.2.5.3   Sorpti on—Dill ing et al.  (1975) found that DCM could be adsorbed to dry
bentonite clay and peat moss when these absorbents were added to a sealed solution
containing DCM.  However, the DCM that leaches from landfills adsorbs very little
to clay,  limestone,  and/or peat moss, so retention in the  soil is  unlikely.
These authors  consider  evaporation  of DCM from water a more  important process
than adsorption.
3.3  LEVELS OF EXPOSURE
     Dichloromethane  has  been detected  in  ambient air and  in  surface  and
drinking waters at numerous locations throughout the United States.
     Dispersion models  have been  used  to  estimate population  exposure to
ambient  DCM.   The predicted maximum annual  average to  which people may be
                               2
exposed is  14.3 ppb (0.050 mg/m ) from  living near  DCM production  facilities.
People living  near  other DCM sources such  as  organic solvent cleaning and
paint  and varnish  removal operations  are  expected  to  be exposed  to
concentrations that  do not exceed  7.1  to 14.3 ppb (0.025  to  0.050 mg/m )
averaged over a year's time (Systems Applications,  Inc.,  1983).
     Average background  mixing ratios are  approximately  30  to  50 ppt.   In
their review of ambient air data, Brodzinsky and Singh (1983)  concluded that
background levels  are about 50 ppt,  with many urban levels one or two orders of
magnitude higher.   Ambient air levels of DCM at various locations are shown in
Table 3-6.  Singh et  al.  (1983) reported that the average northern  hemisphere

005DC1/C                                  3-11                         12/7/83

-------
                                                      PRELIMINARY DRAFT

                                      TABLE  3-6.  AMBIENT AIR  LEVELS  OF DICHLOROMETHANE
       Location        Type  of  site
                                        Date of measurement/
                                         analytical  method
                                Mixing ratio
                                     (ppb)
                                                     Reference
   Arizona
Phoenix
    California

    Mill  Valley
Riverside
                      Urban
April 23 - May 6, 1979/
GC-EC
ro
    Point Arena
    San  Jose
    Point  Arena
                  Background subject   Jan.  11 -27,  1977/
                   to urban trans-     GC-EC
                   port
                      Urban
    Badger  Pass        High  altitude
                  Marine coastal
                  Urban
                  Marine coastal
    Los  Angeles        Urban
April 25 - May 4, 1977/
GC-EC
                                       May 5-13,  1977/
                                       GC-EC
May 23-30, 1977/
GC-EC
Aug.  21-27,  1978/
GC-EC
Aug.  30
GC-EC
                                       April  9
                                       GC-EC
        - 21,  1979/
                     Max 5.1552
                     Min 0.0859
                     Avg 0.8936 ± 0.9886
                     Max 0.087
                     Min 0.038
                     Avg 0.055 ± 0.014

                     Max 0.473
                     Min 0.033
                     Avg 0.111 ± 0.094

                     Max 0.126
                     Min 0.009
                     Avg 0.044 ± 0.025

                     Max 0.102
                     Min 0.013
                     Avg 0.045 ± 0.022

                     Max 1.920
                     Min 0.060
                     Avg 0.401 ± 0.352
Singh et al., 1981
- Sept.  5, 1978/     Max 0.080
                     Min 0.016
                     Avg 0.039 ± 0.017

                     Max 12.0288
                     Min  0.6014
                     Avg  3.7511 ± 2.6203
                                                       Singh et al., 1979
Singh et al., 1979
                                                       Singh et al., 1979
Singh et al., 1979
Singh et al., 1979
Singh et al.,  1979
Singh et al.,  1981

-------
   PRELIMINARY DRAFT
TABLE 3-6.  (continued)
Location
Oakland
Upland*
Kansas
Jetmar
to
,L Louisiana
CJ
Baton Rouge
Geismar
Missouri3
St. Louis
Nevada
Reese River
Date of measurement/
Type of site analytical method
Urban June 28 - July 10, 1979/
GC-EC
Urban Aug. 17 - Sept. 23, 1977/
GC-MS
Remote June 1-7, 1978/
GC-EC
Urban March 3 - May 20, 1977/
GC-MS
Urban March 1, 1977/
GC-MS
Urban May 30 - June 8, 1980/
GC-EC
High altitude May 14-20, 1977/
GC-EC
Mixing ratio
(ppb) Reference
Max 2.4058 Sinqh et al.. 1981
Min 0.0859
Avg 0.4155 ± 0.3146
Max 13 Pellizzari and Bunch ,
Min 0.54 1979
Avg 8.5 ± 4.5
Max 0.105 Singh et al., 1979
Min 0.033
Avg 0.054 ± 0.015
Max 0.55 Pellizzari et al.. 1979
Min 0.046
Ayj 0.25 ± 0.17
Max 0.33 Pellizzari, 1978a
Min 0.33
Avg 0.33 ± 0.062
Max 0.62 Singh et al., 1980
Min 0.16
Avg 0.39 ± 0.14
Max 0.099 Singh et al., 1979
Min 0.015
                Avg 0.052 ± 0.022

-------
                                                        PRELIMINARY DRAFT

                                                     TABLE 3-6.  (continued)
CO
I
Location
New Jersey3
Bridgeport
Edison
North Pacific
37°N
0°N - 33°N
Panama
Canal Zone
Southern Hemi
0° to 42°S
77° to 90°S
32° to 55°S
77° to 90°S
Type of site
Urban
Urban
Ocean
Marine
Marine
sphere
Marine
Remote
Marine
Remote
Date of measurement/
analytical method
Sept 22, 1977/
GC-MS
March 24 - July 1, 1977/
GC-MS
April 1977/GC-MS
Oct. 1976/GC-MS
July 1977/GC-MS
Oct. 1976/GC-MS
Jan. 1977/GC-MS
Oct. 1977/GC-MS
Nov. 1977/GC-MS
Mixing ratio
(ppb) Reference
Max 0.26 Pellizzari and Bunch, 1979
Min 0
Avg 0.13 ± 0.19
Max 69 Pellizzari, 1978a; 1977
Min 0
Avg 26 ± 30
Avg 0.030 ± 0.008 Cronn et al. , 1977
Avg 0.033 ± 0.046 Robinson, 1978
Max - Cronn and Robinson,
Min - 1979
Avg 0.034
Avg 0.035 ± 0.003 Robinson, 1978
Avg 0.034 ± 0.004
Avg 0.040 ± 0.002
Avg 0.033 ± 0.001

-------
                                                    PRELIMINARY  DRAFT
                                                  TABLE 3-6.   (continued)
Location Type of site
Texas
Aldine* Suburban
Houston* Urban
Virginia*
Front Royal Urban
Washington
Pullman Rural
Pullman Rural
Date of measurement/
analytical method

June 22 - Oct. 20, 1977/
GC-MS
June 28, 1977, -
May 24, 1970/GC-MS, GC-EC

Sept. 29 - Nov 16, 1977/
GC-MS

Dec. 1974 - Feb. 1975/
GC-MS
Nov. 1975/GC-MS
Mixing ratio
(ppb) Reference

Max 1.30 Pellizzari et al . , 1979
Min 0.29
Ayj 0.84 ± 0.50
Max 1.30 Pellizzari et al., 1979
Min 0 Singh et al. , 1980
Ayg 0.57 ± 0.32

Max 21 Pellizzari, 1978b
Min 0.49
Avg 8.5 ± 4.5

Max - Grimsrud and Rasmussen,
Min - 1975
Avg <0.005
Avg 0.035 Rasmussen et al., 1979
 GC-EC = gas chromatography-electron capture.
 GC-MS = gas chromatography-mass spectrometry.

*Data obtained from summary report of Brodzinsky and Singh, 1983.

-------
background mixing ratio is approximately 38 ppt and the global  average is 29 ppt.
The Washington State University group reported free troposphere values between
30 and 40  ppt (Cronn et al., 1977; Robinson, 1978; Cronn and Robinson 1979;
Grimsrud and Rasmussen, 1975; Rasmussen et al., 1979).
     Pellizzari and  Bunch  (1979)  have compiled a  list  of the  sites in the
United States  at which  Pellizzari and coworkers have identified DCM.  DCM was
                                 ®
sampled by  adsorption  onto Tenax  coupled with analysis by  high  resolution
gas chromatography-mass spectrometry  (GC-MS).   The highest  DCM concentration
reported in New Jersey was at a waste disposal site in  Edison.   A  level  of 360
ppb was measured during an  11-minute  sampling period in March 1976.  During a
75-minute sampling period  (November 1977), 55 ± 0.1 ppb DCM was detected at a
site in Staten Island,  New York.  Longer sampling times (up to 8  hours) were
required at  sites in Virginia, West Virginia, and  Pennsylvania to detect DCM.
During a 7.25-hour  period,  about 70 ppb DCM were detected at a site in Front
Royal, Virginia (October 1977).   In the southwest, a level of about 1 ppb was
reported at  sites in Houston, Texas,  during a 3-hour sampling period (October
1977).  During  considerably longer  sampling  periods (up to 24 hours),  lower
concentrations were  reported  for sites  in Louisiana, e.g., Baton  Rouge and
Geismar.   Sampling during a 48-hour period in Upland, California (August 1977)
indicated concentrations of about 12 ± 9 ppb DCM.
3.3.1  Analytical  Methodology
     There are  four  practical  methods to measure  air  concentrations  of the
halogenated hydrocarbons.

     (1)  Gas chromatography with an electron-capture detector.
     (2)  Gas chromatography-mass spectrometry.
     (3)  Long-path  infrared  absorption spectroscopy, usually with  preconcen-
tration  of  whole  air  and  then  separation  of the compounds by  gas
chromatography.
     (4)  Infrared solar  spectroscopy,  using  the  solar spectrum  at large
zenith angles to obtain greatest path lengths through the atmosphere.

     Each method has advantages  and disadvantages and applications for which
it is best suited.   A major drawback with these techniques is that they do not
allow real-time continuous  measurements  of the halocarbons at ambient levels
in the environment.

005DC1/C                                  3-16                         12/7/83

-------
     The two most widely used systems for identifying and measuring trace amounts
of DCM that occur in ambient air are gas chromatography-mass spectrometry (GC-MS)
and gas chromatography-electron capture detection (GC-EC).  Both systems have a
limit of detection below 30 ppt.  The GC-EC method has been reviewed by Pellizzari
(1974) and by Lovelock (1974).  The electron-capture detector is specific in that
halogenated hydrocarbons are quantitated while non-halogenated hydrocarbons do
not respond.  Thus,  high background  levels  of  non-halogenated hydrocarbons  in
ambient air or water samples do not interfere with measurements of halogenated
hydrocarbons.  In a complex mixture in which several compounds may have similar
retention times, alteration of the operating parameters of the GC-EC system will
usually provide separation of the components.  Linearity over a wide concentra-
tion range is achieved when the electron capture detector is used in the constant
current mode.  In this mode, the change in pulse frequency is linearly related to
sample concentration.  Nitrogen, or a 95-percent argon/5 percent methane mixture,
is commonly used as the carrier gas.
3.3.2  Sampling of Ambient Air and Water
     Contamination, absorption, and adsorption are common problems of the methods
used to analyze air and water for DCM content.   Four general approaches are used
to collect samples of air for analysis of trace gas concentration:  (1) cryogenic
sampling in which liquid helium or liquid nitrogen is used to cool a container
to extremely low temperatures; (2) pump-pressured samples, in which a mechanical
pump is used without cryogenic assistance to fill a sampler to a positive pres-
sure relative to the surrounding atmosphere; (3) ambient or subambient pressure
sampling in which an evacuated  container  is  simply opened and allowed  to fill
until  it has reached ambient pressure at the sampling location;  and (4) adsorption
of selected gases on adsorbants such as molecular sieves or activated charcoal.
Contamination and other problems are more serious in low-pressure sampling than
in high-pressure sampling systems (National  Academy of Sciences, 1978).
     Water samples are  subject  to the same  possibility of contamination and
other problems that exist in air  sampling.   DCM  aqueous samples must be care-
fully sampled, transported, and stored because of its volatility  and the com-
plexity of the samples, especially those containing chlorine or other oxidants.
A technique that is often used involves filling and sealing a serum bottle with-
out air space and storing it just above freezing (Kopfler et al., 1976).   Water
samples generally require additional  preparation before analysis.   Normally, such
samples may be concentrated  by various water analysis techniques, but direct
aqueous injection is used occasionally in GC analysis.
005DC1/C                                  3-17                         12/7/83

-------
3.3.2.1  Sampling and Detection in Ambient Air—Several  common approaches are
used to sample ambient  air  for trace gas analysis,  including  the  following
approaches (National  Academy of Sciences,  1978).
     (1)  Pump-pressure  samples:   A mechanical  pump is  used  to  fill  a stainless
steel or glass container  to a positive pressure relative to the surrounding
atmosphere.
     (2)  Ambient pressure samples:   An evacuated chamber is opened and allowed
to fill until  it has reached ambient pressure at the sampling location.  If
filling is conducted  at  high altitude,  the sample may become contaminated upon
return to ground level.
     (3)  Adsorption  on  molecular sieves,  activated charcoal,  or other sorbents.
     (4)  Cryogenic samples:  Air is pumped into a container and liquefied,  and
the partial  vacuum that  is created allows  more air to enter.   This  method allows
collection of several thousand liters of air.
     Singh et al. (1979) have satisfactorily measured ambient levels of DCM  by
analysis with GC-EC.   Samples were pressurized in stainless  steel vessels, then
                                                                      ®
preconcentrated by freezeout on 100/120 mesh glass beads (use of Tenax  monomer
was discontinued because oxygen oxidized the monomer and interfered with electron
capture detection).  Separation was performed on a column containing 0.2 percent
Chromosorb W 1500 80/100 mesh on Carbopack C.   A post-column Ascarite water  trap
was used to remove water before electron-capture detection.   Dual detectors  were
used to provide a coulometric response.
     Harsch et al. (1979) used GC-EC to identify and measure the level of DCM in
samples of ambient air.   A 500-ml sample was preconcentrated using the freezeout
concentration method  (Rasmussen et al., 1979).   Halocarbons  were desorbed onto a
stainless steel column packed with 10 percent SF-96 on 100/120 mesh Chromosorb
W.  The reported detection limit was 26 ppt.  Good separation from chlorofluoro-
carbon 113 (trichlorotrifluoroethane) was reported.
     Cronn and Harsch (1979) reported a GC-MS detection limit of 6 ppt for a 500-
ml aliquot of  DCM.   Cronn et  al.  (1977) also reported a  detection  limit  of 20
ppt for a 100-ml aliquot.  Pressurized air samples were separated on a column of
Durapack n-octane on 100/120 mesh Porasil  C.  Cronn et al. (1976) have compared
GC-MS with GC-EC in  terms of precision and sensitivity.   In general, GC-MS
offered great  specificity  but could not  equal  GC-EC  in reproducibility for
the 11 halocarbons studied.  For mass spectrometry, the detection limit for DCM
was  9  ppt  with a percent standard deviation of 13.  The detection limit with
temperature-programmed GC-EC was 4 ppt and the percent standard deviation was 7.3.
005DC1/C                                  3-18                         12/7/83

-------
     Pellizzari  and  Bunch  (1979)  reported  an  estimated  detection  limit  of  200
ppt using a high-resolution GC-MS system in which DCM was first adsorbed onto a
     ®
Tenax  GC.  The  accuracy of analysis was reported as ±30 percent.   The inherent
analytical  errors  are a  function of several  factors,  including:   (1) the
ability  to  accurately determine  the  breakthrough volume; (2) the  accurate
measurement of the ambient air volume sampled; (3) the percent recovery of DCM
from the  sampling  cartridge  after a period of storage;  and  (4) the  reproduci-
                                            ®
bility of thermal  desorption from the Tenax   cartridge and  its introduction
                                                   (R)
into the analytical system.  Oxidation of the Tenax  monomer was not reported.
Singh et al. (1982) have cautioned that the adsorption of halogenated compounds
        ®
on Tenax  may not reliably reflect ambient air levels because measurements by some
investigators have been  reported  as less  than background (100  to 200 ppt).
     Difficulties  in  using a  coulometric approach to  the GC-EC quantification
of DCM  in  air samples have been reported by  Lillian and Singh (1974).  These
investigators were unable  to measure  DCM accurately with detectors in series
because of a greater-than-coulometric response.  Dichloromethane was reported to
have a very low  ionization efficiency.  The observed response might be attributed
to the products  of ionization having greater electron affinities than the reac-
tants.
     Cox et al.  (1976)  reported that polyglycol  stationary-phase chemically
bonded to porous glass (Durasil Low Kl) was the only material found to separate
DCM from other halocarbons during a GC-EC analysis.   Satisfactory separation and
analysis of DCM was reported by Grimsrud and Rasmussen (1975) with a 50-foot SCOT
OV-101 column by GC-MS.
     When the freezeout  concentration method of  Rasmussen et al. (1979) was
applied to  a  500-ml  aliquot  of air,  the detection  limit for DCM with GC-EC
analysis was 4 ppt and the percent standard deviation was 26.2.   The GC column
contained 10 percent SF-96 on 100/120 mesh Chromosorb W.
     The National  Institute for Occupational  Safety and Health (NIOSH)  method
P & CAM 127 (NIOSH, 1974) is  recommended for measurement of DCM in samples where
the concentration is  greater  than 0.05 mg/sample.   This method uses adsorption
on charcoal  followed  by desorption with carbon disulfide.  Analysis is made by
gas chromatography with flame ionization detection.   The mean relative standard
deviation of the method is 8  percent.
     Grimsrud and Miller (1979) have reported an improved GC-EC method by which
the detector response of DCM  is enhanced by the addition of 02 in the carrier

005DC1/C                                  3-19                         12/7/83

-------
gas.   At the  highest  02  doping (5 ppth) the response of the detector to DCM
(960 ppb) was enhanced 57-fold.   The enhancement is depicted in  Figure 3-1.   A
constant-current electron-capture detector was  used.
3.3.2.2  Sampling and Detection in Water—A gas  purging  and trapping method
suitable for  DCM  has  been described by Bellar et al. (1979).  Water samples
are purged by bubbling with helium or nitrogen at 23°C.   The halocarbons  are
adsorbed onto  a porous polymer trap as the gas is vented.   Quantification is
                     ®
made by GC-MS.  Tenax  GC (60/80 mesh) was considered an effective adsorbent
for compounds  that boil above approximately 30°C.  A recommended general pur-
pose column is an 8-foot  by 0.1-inch inside diameter (i.d.) stainless steel  or
glass tube packed with 0.2 percent Carbowax® 1500 on Carbopack®  - C (80/100 mesh).
For a  sample  volume  of  5 ml, the  range  of the limit of detection is 0.1  to
1.0 ug/1.
3.3.2.2.1  Sample preservation (water).  Bellar et al.  (1979) recommended  that
water samples be stored in narrow mouth glass  vials.   Vials are  filled to  zero
head space and covered with a Teflon -faced silicone rubber septum.   Screw
caps are suitable seals.   The presence of chlorine in water samples results in
an increase  in the  concentration of certain halomethanes (not including DCM)
upon storage.
3.3.2.2.2  Soil and sediment (water).   Dichloromethane has been  found in drinking
and surface water at a variety of locations in the United States.  In a recent
National Academy of Sciences report (1977), DCM formation resulting from chlo-
rination treatment of water was reported.
     In a survey for volatile organics in five drinking water supplies, Coleman
et al.  (1976) found  that DCM was  common  to  all  the cities evaluated (i.e.,
Cincinnati, Miami, Philadelphia,  and Ottumwa,  Iowa).  Concentrations were  not
reported.  Analysis was performed using GC-MS.
     In a 1975 survey by  the U.S. Environmental Protection Agency (EPA) (1975a),
DCM was detected  in 9 of  10 water supplies.  Lawrence, Massachusetts,  had  the
highest concentration (1.6 ug/1).  A mean concentration of less  than 1 ug/1 in
finished water was reported in  a  survey  of Region V water  supplies  (U.S.  EPA,
1975b).  This  survey indicated 8 percent of the finished-water supplies contained
detectable DCM.
     Dichloromethane was  not among the major halogenated hydrocarbons detected
by Dowty et al. (1975a) in New Orleans drinking water.  In another report, DCM
was detected  in finished waters in the New Orleans area by GC-MS (Dowty et al.,
1975b), and  it was  also  found  in  Mississippi  River  clarifier effluent.   The
005DC1/C                                  3-20                         12/7/83

-------
                             /             300°
                         12345
                           O2 CONCENTRATION, ppth
Figure 3-1.   The effect of oxygen doping of the carrier gas on the ECD  response
             to several halogenated methanes at a detector temperature  of  300°C.

Source:  Grimsrud and Miller, 1979.
005DC1/C
3-21
11-10-83

-------
chlorocarbon was sorbed onto poly p-2,6-diphenyl  phenylene oxide  (35/60 mesh).
Vapors were desorbed  onto  a capillary chromatographic column and quantitated
by mass  spectrometry.  Raw water influent was purified after clarifier treat-
ment (sedimentation and some chemical  treatment)  to an extent that DCM concen-
tration  dropped  32 percent.  However, the  concentration  in finished water
increased after chlorination.  Dichloromethane also was detected  in commerci-
ally bottled artesian well  water.
     A slight  increase in DCM concentration in chlorinated finished water was
observed by Bellar et al.  (1974).   Finished water having 2.0 ug/1 DCM resulted
after a complete treatment process of raw river water containing  no detectable
DCM.   Water samples also were collected from various locations in a sewage treat-
ment plant.  Before  treatment,  the water contained 2.36 ppb DCM  (8.2 ug/1).
Before chlorination and after preliminary treatment, the water contained 0.8 ppb
DCM (2.9 ug/1).  After chlorination, the effluent contained 1 ppb DCM (3.4 ug/1).
These tests show that DCM and other chlorocarbons may have formed as a result of
the chlorination treatment.   The most notable chlorocarbon was chloroform, for
which an  increase  of 7.1 ug/1 to  12.1 ug/1 was  observed after chlorination.
GC-MS analysis was performed using a  headspace  preconcentration technique.
     Dichloromethane was detected  at 32  of  204 surface water sites  from which
samples were collected (Ewing et al., 1977).  Sites were  located  near  heavily
industrialized  river basins across the  United States.   Concentrations were
reported as being  greater  than  1  ppb  (1 ug/1).   Samples were collected from
July 1975  to December 1976.   Of the 204 sites, 91 were near major rivers and
57 were  in  tidal areas and  estuaries.  Samples (125 ml) were held at  60°C and
                                            (R)
stripped.  Volatiles  were  sorbed onto Tenax  GC  and  desorbed  onto Carbowax
1500 columns and analyzed by mass  spectrometry.
     Dichloromethane was not among the contaminants  detected by Sheldon  and
Hites (1978) in raw Delaware River water collected from August 1976 to March
1977.
     Pellizzari and Bunch (1979) reported detecting DCM in untreated Mississippi
River water  (Jefferson  Parish,  Louisiana) at a  mean  concentration of 2.581
ug/1.  The  highest value reported was 15.8 ug/1.   Determinations were made
from February  7 to August 5, 1977.   A  mean concentration of 0.13  ug/1 was
reported  by  Pellizzari  and Bunch  (1979) in tap  water from Jefferson  Parish.
The highest level was 1.1 ug/1.
005DC1/C                                  3-22                         12/7/83

-------
      Data  for DCM concentrations  in general ambient waters  in EPA's Storet files,
 covering the period January 1978  to April 15, 1981, indicate that  levels ranged
 from  0 to  120 (jg/1 (U.S. Environmental Protection Agency, 1981).   Sediment sam-
 pling detected DCM in 60 of 118 cases.  Concentrations were from 427 to 433 ppb.
 Ambient  soil  concentrations  of DCM are unknown  (U.S.  Environmental Protection
 Agency, 1981).
      Singh et al. (1983) have  detected DCM  in seawater samples  from the eastern
 Pacific Ocean.  Mean surface concentrations of 2 ng/1 were measured.
 3.4   ECOLOGICAL EFFECTS
 3.4.1 Effects on Aquatic Organisms
      Dichloromethane has been  tested for acute toxicity  in a limited number of
 aquatic species.  The information  in this chapter focuses upon  observed DCM con-
 centrations that were  reported to result in adverse effects under laboratory
 conditions.  Such parameters of toxicity are not easily  extrapolated to environ-
 mental situations.  Test  populations  may not be representative of the entire
 species because  susceptibility to the test substance at different lifestages
 may vary considerably.   Guidelines for the use of these  data in the development
 of criteria levels for DCM in water are discussed in the Ambient Water Quality
 Criteria  for  Halomethanes (U.S.  Environmental  Protection  Agency, 1980b).
      The toxicity of DCM  to  fish and other aquatic organisms has  been gauged
 principally by  flow-through  and  static  testing methods.  The  flow-through
 method exposes the organism(s) continuously to a constant concentration of DCM
 while oxygen  is  continuously replenished and waste products are  removed.  A
 static test exposes the organism(s) to the added initial concentration only.
 Results from  both types of tests are commonly used as initial   indications of
 the potential  of substances to cause adverse effects.
 3.4.1.1   Effects on Freshwater Species—Results  of  flow-through   and  acute
 static tests  with  DCM and freshwater  species  (fish and invertebrates)  are
 shown in Table 3-7.
     Alexander et al. (1978) used both flow-through and  static  methods to  in-
 vestigate the acute toxicity  of several  chlorinated solvents,   including DCM,
 to adult  fathead minnows  (Pimephales promelas).  Studies were  conducted in
accordance with  EPA  procedures described  by the Committee  on  Methods for
Toxicity Tests with Aquatic Organisms  (1975).  Alexander et al. reported that
the fish were observed  for the following effects:  loss  of  equilibrium, mela-
nization,  narcosis,  and swollen, hemorrhaging gills.   The DCM concentration that

005DC1/C                                  3-23                         12/7/83

-------
                                   PRELIMINARY DRAFT
                     TABLE 3-7.   EFFECTS OF DICHLOROMETHANE ON FRESHWATER SPECIES IN ACUTE TESTS
      Species                         Type of test         LC5oa            EC5o                 Reference

Bluegill (Lepomis macrochirus)        static             224,000 pg/1         -              U.S. EPA, 1980a


Fathead minnow (Pimephales            static             310,000 pg/1         -              Alexander et al., 1978
  promelas)


Fathead minnow (Pimephales            flow-through       193,000 jjg/1         -              Alexander et al., 1978
  promelas)


Daphnia magna                         static                -              224,000 M9/1      U.S. EPA, 1980a


a96-hour test.

b48-hour test.

-------
produced one or more of these observable effects in 50 percent of the fish (EC50)
was  99.0  mg/1  for DCM.   Fish affected  during  the exposure were transferred to
static freshwater aquaria at the end of the 96-hour exposure period.  Only the
fish that were severely affected by high concentrations of the chemical did not
recover.  However, short exposures to these compounds at the sublethal level  seem
to produce only reversible effects.
     Chronic test  data  concerning life cycle or embryo-larval tests  are  not
available  nor  were data  found on the  chronic toxicity to  invertebrates.
3.4.1.2  Effects on Saltwater Species—Static tests with mysid shrimp resulted
in an LC50 value of 256,000 ug/1.  No data exist on the chronic effects of DCM.
     In a  96-hour static test with Sheepshead minnow (Cyprinodon variegatus)
the  LC50 value was 331,000 ug/1 (U.S.  Environmental Protection Agency, 1980a).
3.4.2  Effects on Plants
     The 96-hour  EC50 values  for  DCM,  based on  chlorophyll a and  cell  numbers
of the freshwater alga,  Selenastrum capricornutum,  were greater than the highest
test concentration (662,000 ug/1).  The 96-hour EC50 value based on chlorophyll
a and cell numbers of the saltwater alga, Skeletonema costatum, was greater than
the  highest  test  concentration (662,000 ug/1) (U.S. Environmental Protection
Agency, 1980a).
     Few studies on the effects on vascular plants  are available (U.S. Environ-
mental Protection Agency, 1981).  Lehman and Paech  (1972) tested the effect of
DCM vapors on the photosynthetic fixation of 14C02  by alfalfa seedlings.   At a
very high concentration (21 percent),  DCM reduced photosynthesis by 82 percent.
3.4.3  Bioconcentration Potential
     Bioconcentration refers  to  increased concentration  of a substance in the
tissue of an organism (e.g., fish) relative to the  ambient water concentration
under steady-state conditions.  A measure of the potential  for organic chemicals
to bioconcentrate  in the  fatty tissues of aquatic  organisms is given  by  the
octanol/water partition coefficient (Neely  et al., 1974).    The log octanol/
water partition coefficient  for DCM has been measured at 1.25  (Hansch  et  al. ,
1975), but no  steady-state  bioconcentration  factor (BCF) has  been measured.
The BCF represents the ratio of the chemical concentration in  the organism to
that in the  water.   An  approximate BCF for DCM of 5.2 has  been  calculated
using the relationship  Log BCF = 0.76  Log P -  0.23  (Veith et al.,  1979).   This
estimated BCF places DCM in the low range of bioconcentration values,  suggesting
that the potential for bioconcentration in lipids is low.  However, no infor-

005DC1/C                                  3-25                         12/7/83

-------
(nation is available  concerning  DCM  levels in living organisms  (Pearson  and
McConnell, 1975) or the rate of transport of DCM through food chains.
     However, the BCF  alone may not be the most useful measure of the overall
fate of a substance  in water or its potential  for  producing toxic  effects.
Chemical  and biological degradation  of  the substance,  volatilization,  desorption,
and the depuration rate are among the key determinants  of toxicity.
3.5  CRITERIA, REGULATIONS, AND STANDARDS
     Permissible levels of DCM in the working environment have been  established
in various countries.  The OSHA health  standard requires that a worker's  exposure
to DCM at no time exceed 500 ppm (1,737 mg/m3) TWA in  any 8-hour work day of a
40-hr week,  with an acceptable ceiling  concentration of 1000 ppm (3,474 mg/m3),
that should  not  exceed 2000 ppm (6,948 mg/m3) for more than 5 minutes in any
2-hour period.  The American Conference of Government  Industrial Hygiene  (ACGIH)
TL\r for inhalation exposure [200 ppm (695 mg/m3)],  proposed for prevention of
narcotic effects or liver injury and for protection against excessive carboxy-
hemoglobin formation,  has  recently  been lowered to 100 ppm (347 mg/m3).   The
8-hour TWA value  in  the  Federal Republic  of Germany is 100  ppm;  in the German
Democratic Republic and Czechoslovakia  it is 144 ppm;  and in Sweden  it is 100 ppm.
The acceptable ceiling concentration in the USSR is 14 ppm (49 mg/m3).  NIOSH has
recommended  that  occupational  exposure to DCM not exceed 75 ppm (261 mg/m3),
determined as a  TWA  for up to  a  10-hour work day of a 40-hour week, in the
absence of exposure to carbon monoxide  above a TWA of 9 ppm for up to a 10-hour
work day.
     The Ambient Water Quality Criteria for Halomethanes (U.S. EPA,  1980a) indi-
cates that for  the maximum protection  of  human  health  from  the  potential  car-
cinogenic effects caused by exposure to any of several  halomethanes, including
DCM,  chloromethane,   bromomethane,   bromodichloromethane,  tribromomethane,
dichlorodifluoromethane, and/or trichlorofluoromethane,  through ingestion of
contaminated  water  and  contaminated aquatic organisms,  the  ambient water
concentration should be zero based on the assumption of no threshold for these
chemicals.   However, zero  level may not be attainable; therefore, the levels that
may result  in incremental  increase  of  cancer  risk over the  lifetime  are  esti-
mated at  10  5,  10 6,  and 10 7.  The  corresponding recommended criteria,  based
on data for  chloroform, are 1.9 ug/1, 0.19 ug/1, and 0.019 ug/1, respectively.
If the above estimates are made for consumption  of aquatic  organisms only,
excluding consumption  of water, the levels are 157 ug/1, 15.7 ug/1, and 1.57
ug/1,  respectively.   Estimates for  water consumption  only were not made.
005DC1/C                                  3-26                         12/7/83

-------
     For halomethanes where one criterion is derived for an entire class of com-
pounds, the Agency does not state that each chemical in the class is a carcinogen.
The intended interpretation of the criterion is that the risk is less than 10 5
whenever the total concentration of all halomethanes in water is less than the
criterion.   In a hypothetical case where all of the halomethanes in a sample are
non-carcinogenic, the criterion would be too strict.  In most cases where halo-
methanes are detected, a mixture of compounds occurs and in calculation of the
criterion,  the assumption is made that all components have the same carcinogenic
potency as  chloroform.
     The derived  water quality  criterion, based on  non-carcinogenic  risks  and
assuming a daily water intake of 2 liters and consumption of 6.5 grams of fish
and shellfish  per day (bioconcentration  factor 0.91),  would be 12.4 mg/1.
005DC1/C                                  3-27                         12/7/83

-------
3.6  REFERENCES


Ahlstrom,  R.  C.  and J.  M.  Steele.   Methylene chloride.  In:    Kirk-Othmer
     Encyclopedia of Chemical Technology.   A.  Standen,  ed.,  Interscience  Pub-
     lishers, New York, NY, 1979.

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

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

Anonymous.   Methylene  Chloride.   In:   Chemical  Marketing  Reporter,  p.  9,
     August 6, 1979.               ~

Anonymous.  Methylene Chloride.   In:  Chemical and Engineering News.  December 20,
     1982.  pg.  29.               ~~

Anthony, T.  Chlorocarbons and chlorohydrocarbons.  Jji:  Kirk Othmer's Encyclo-
     pedia of Chemical  Technology.   Wiley-Interscience, 1979.

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

Bellar,  T. A., W.  L.  Budde, and J. W.  Eichelberger.   The  identification  and
     measurement of volatile organic compounds in aqueous environmental samples.
     Chapter 4.    In:   Monitoring Toxic Substances,  ACS  Symposium  Series  94.
     D. Schuetzle, ed., American Chemical  Society, 1979.

Brodzinsky, R.,  and H.  B. Singh.   Volatile Organic Chemicals in the Atmosphere:
     An Assessment of Available Data.   EPA-600/S 3-83-027.  U.S. Environmental
     Protection Agency, June 1983.

Brunner,  W.,  and T.   Leisinger.   Bacterial degradation of  dichloromethane.
     Experentia 34:1671,  1978.

Brunner, W., D.  Staub,  and T.  Leisinger.  Bacterial  degradation of dichlorome-
     thane. Appl. Environ. Microbiol.  40:950-958, 1980.

Butler, R., I. J. Solomon, and A. Snelson.  Rate constants for the reaction of
     OH with  halocarbons  in the  presence of 02 + N2.  J. Air Pollut. Control
     Assoc. 28:1131-1133, 1978.

Coleman, W.  E.,  R. D.  Lingg, R.  G.  Melton,  and F. C.  Kopfler.   The occurrence
     of  volatile  organics  in five drinking water supplies  using gas  chromato-
     graphy/mass spectrometry.    Chapter 21.   In:  Identification and Analysis
     of Organic  Pollutants  in  Water.   L.   H.  Keith,  ed.,  Ann Arbor Science,
     Ann Arbor,  MI,  1976.

Committee  on  Methods for Toxicity Tests with Aquatic Organisms.  Methods for
     acute  toxicity  tests  with  fish, macroinvertebrates,  and amphibians.
     Ecol. Res.  Ser.   EPA-660/3-75-009.  U.S. Environmental Protection Agency,
     1975.

005DC1/C                                  3-28                         12/7/83

-------
Cox, R. A.,  R.  C.  Denwent, A.  E.  J.  Eggleton, and J.  E.  Lovelock.   Photo-
     chemical oxidation of  halocarbons  in the troposphere.   Atmos.  Environ.
     10:305-308, 1976.

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

Cronn, D.  R. ,  and  E.  Robinson.  Determination of  trace gases in  Learjet  and
     U-2 whole-air samples collected during the intertropical convergence zone
     study.   Report submitted to NASA,  August 1978.  In:  1977 Intertropical
     Convergence Zone Experiment.  I. G. Poppoff, W. A.  Page, and A.  P. Margozzi,
     NASA TMX78577, 1979.

Cronn, D.  R., R. A. Rasmussen, and E. Robinson.  Report for Phase I.  Measure-
     ment of Tropospheric Halocarbons by Gas Chromatography-Mass Spectrometry.
     Report prepared for the U.S. Environmental Protection Agency, August 1976.

Cronn, D.  R., R. A. Rasmussen, and E. Robinson.  Report for Phase II.  Measure-
     ment of Tropospheric Halocarbons by Gas Chromatography/Mass Spectrometry.
     Report  prepared  for the  U.S.  Environmental  Protection Agency,  October
     1977.

Crutzen, P.  J., and J.  Fishman.  Average concentrations of OH- in the troposphere
     and the budgets of CH4, CO, H2, and CH3CC13.  Geophys. Res. Lett. 4:321-324,
     1977.

Davis, D.  D.,  G.  Machado, B.  Conaway,  Y.  Oh,  and  R.  Watson.   A  temperature-
     dependent  kinetics study of the reaction of OH with CH3C1, CH2C12, CHCla,
     adn CH3.   Br.  J.  Chem. Phys. 65(4):1268-1274, 1976.

De Forest, E.  M.   Chloromethanes.   In:   Encyclopedia of  Chemical  Processing
     and Design.   J. J.  McKetta ancTW. A.  Cunningham,  eds., Marcel  Dekker,
     Inc., 1979.

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

Dilllng, W.   L.  and H.  K.  Goersch.  Dynamics, Exposure, and Hazard Assessment
     of Toxic Chemical.   R. Hague, ed.  Ann Arbor Science  Publishers,  Inc. Ann
     Arbor,  Michigan, 1980.  pg. 111.

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

Dilling, W.  L., C.  J.  Bredeweg, and N. B.  Terfertiller.  Simulated atmospheric
     photodecomposition rates of methylene chloride, 1,1,1-trichloroethane,
     trichloroethylene, tetrachloroethylene,  and other  compounds.   Environ.
     Sci.  Technol.  10(4):351-356, 1976.
005DC1/C                                  3-29                          12/7/83

-------
Dowty, B. J. ,  D.  R.  Carlisle, and J. L. Laseter.  New Orleans drinking water
     sources tested by  gas  chromatography/mass spectrometry.  Environ.  Sci.
     Technol. 9(8):762-765,  1975a.

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

Edwards, P.  R., I. Campbell, and G. S. Milne.  The impact of chloromethanes on
     the environment.   Part 2.  Chemistry  and Industry, 4 September 1982,  pp.
     619-622.

Ewing, B. B.,  E.  S. K.  Chian, J.  C.  Cook,  C.  A.  Evans,  P. K.  Hopke, and E.  G.
     Perkins.  Monitoring  to  Detect  Previously Unrecognized  Pollutants  in
     Surface Waters.   EPA-560/6-77-015,  U.S.  Environmental  Protection  Agency,
     July 1977.

Fells, I., and E.  A. Moelwyn-Hughes.  The  kinetics of the hydrolysis of methy-
     lene chloride.   J. Chem.  Soc. (London) 1326-1333, 1958.

Grimsrud, E. P.,  and  D. A.  Miller.   A  new approach  to  the  trace analysis of
     mono-  and di-halogenated organics,  an analysis  of  methyl  chloride in the
     atmosphere.   In:   National  Bureau  of Standards Special Publication 519,
     Trace Organic Analysis:  A New Frontier  in Analytical Chemistry,  Proceed-
     ings of the  9th Materials  Research Symposium, April  10-13,  1978.   Issued
     April, 1979, pp.  143-151.

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

Hansch, C. , A.  Vittoria, C.  Silipo, and P.V.C. Jow.  Partition coefficient  and
     the  structure-activity relationships  of the anesthetic gases.  J.  Med.
     Chem.  18(6): 546-548, 1975..

Harsch,  D.  E. , D. R.  Cronn,  and  W.  R.  Slater.   Expanded  list of halogenated
     hydrocarbons measurable  in ambient air.  J. Air  Pollut.  Control  Assoc.
     29(9):975-976, 1979.

Howard, C.  J., and K. M. Evenson.   Rate constants for the reactions of OH  with
     ethane  and  some  halogen substituted  ethanes at 296K.   J.  Chem.  Phys.
     64:4303,  1976.

Klecka,  G.  M.   Fate and effects  of methylene chloride in  activated sludge.
     Submitted to Appl.  Environ.  Microbiol,  1982.

Kopfler,  F.  C. ,  R.  G.   Melton,  R.  D.  Lingg,  and  W. E. Colman.   Identification
     and  analysis of  organic pollutants in  water.   L.  Keith, ed. , Ann Arbor
     Science Publishers, Inc.,  Ann Arbor,  MI,  1976.  pp.  87-104.

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

Lovelock,  J.   E.  The  electron  capture detector:  Theory and practice.  J.
     Chromatogr.  99:3-12, 1974.


005DC1/C                                  3-30                         12/7/83

-------
Lowenheim, F. A.,  and  M.  K.  Moran.  Methylene Chloride.  Jji:   Faith,  Keyes,
     and Clark's Industrial  Chemicals,  Fourth Edition, John Wiley, New York,
     1975.  pp.  530-538.

McKetta, J.  J.  and W.  A.  Cunningham, eds. ,  Encyclopedia of Chemical Processing
     and Design,   pp.  267, 1979.

National Aeronautics and  Space  Administration.   Chlorofluoromethanes and the
     Stratosphere.   Robert  D.  Hudson,  ed.   National  Aeronautics  and  Space
     Administration Reference Publication 1010, 1977.

National Academy of Sciences.   Drinking Water and Health.  National Research
     Council, 1977.  pp.  743-745.

National Academy of Sciences.  Non-fluorinated halomethanes in the environment.
     Panel on  low-molecular-weight halogenated  hydrocarbons.   Coordinating
     Committee  for Scientific and Technical Assessments  of Environmental
     Pollutants, 1978.

National  Bureau  of Standards,  Special  Publication 513 "Reaction rate and
     photochemical  data for  atmospheric chemistry," 1977, R. F. Hampson, Jr.
     and  D.  Garvin (eds.).  National  Bureau of Standards, Washington,  DC,
     1978.

National  Institute for Occupational Safety and Health  (NIOSH).  NIOSH Manual
     of Analytical  Methods, HEW Publication No. (NIOSH) 75-121, 1974.

Neely, W.B., D.B.  Branson,  and G.E. Blan.   Partition  coefficient to measure
     biconcentration  potential  of organic chemicals   in fish.  Environ. Sci.
     Technol. 8(13):1113-1115, 1974.

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

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

Pellizzari, E.  D.  Analysis of Organic Air Pollutants  by Gas Chromatography  and
     Mass Spectroscopy.  EPA-600/2-77-100.   U.S. Environmental  Protection Agency,
     1977.

Pellizzari,  E.  D.  Measurement of Carcinogenic  Vapors  in  Ambient Atmospheres.
     EPA-600/7-78-062.   U.S. Environmental Protection  Agency, 1978a.

Pellizzari,  E.  D.  Quantification of Chlorinated Hydrocarbons  in Previously
     Collected Air Samples.  EPA-450/3-78-112.   U.S.  Environmental Protection
     Agency, 1978b.

Pellizzari, E.  D. and J.  E. Bunch.  Ambient Air Carcinogenic Vapors:   improved
     Sampling and  Analytical Techniques and  Field Studies.   EPA-600/2-79-081,
     May, 1979.

Pellizzari,  E.  D. , M.  D.  Erickson, and R.  A. Zweidinger.   Formulation of a
     Preliminary Assessment of Halogenated Organic  Compounds in Man  and Environ-
     mental Media.   EPA-560/13-179-006.  U.S. Environmental  Protection Agency,
     1979.
005DC1/C                                  3-31                         12/7/83

-------
Perry, R. A. ,  R.  Atkinson,  and J.  N. Pitts.  Rate constants for the reaction
     of OH  radicals with CHFC12 and  CH3C1  over  the temperature  range 298-493K
     and with CH2C12 at 298K.   J.  Chem.  Phys. 64:1618, 1976.

Radding,  S.  B., D. H.  Liv,  H.  L.  Johnson, and T. Mill.  Review of the environ-
     mental  fate  of selected chemicals.   EPA-560/5-77-003,  U.S.  Environmental
     Protection Agency, 1977.

Rasmussen,  R. A. , D. E. Harsch, P. H. Sweany, J.  P.  Krasnec,  and D.  R.  Cronn.
     Determination of  atmospheric halocarbons by  a temperature  programmed gas
     chromatographic freezeout concentration method.   J.  Air Pollut. Control
     Assoc.  27:579,  1979.

Rittman,  B.  E.  and P.  L.  McCarty.   Utilization of dichloromethane by suspended
     and fixed-film bacteria.  Appl.  Environ.  Microbiol.  39:1225-1226, 1980.

Robinson, E.  Analysis of Halocarbons in Antarctica.   Report 78/13-42 prepared
     for the National  Science Foundation, December 1978.

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

Singh, H. B.   Atmospheric  halocarbons.   Evidence in favor of reduced average
     hydroxyl radical  concentrations  in  the troposphere.  Geophys.  Res.  Lett.
     4(3):101-104, 1977.

Singh, H. B. ,  L.  J.  Salas,  and R. E. Stiles.   Selected man-made halogenated
     chemicals in the  air  and  oceanic environment.   J. Geophy.  Res. 88:3675-
     3683, 1983.

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

Singh, H. B.,  L.  J. Salas, R.  Stiles, and H.  Shigeishi.   Atmospheric Measure-
     ments  of  Selected Hazardous  Organic Chemicals.   Second Year Interim  Re-
     port.  Report prepared  for  the  U.S.  Environmental Protection  Agency by
     SRI International, Menlo Park,  California, 1980.

Singh, H. B. ,  L.  J.  Salas,  H.  Shigeishi, A. J. Smith,  E.  Scribner, and L.  A.
     Cavanagh.    Atmospheric  distributions,  sources  and  sinks  of  selected
     halocarbons, hydrocarbons, SF6,  and N20.   EPA-600/3-79-107, Final  report
     submitted to the U.S.  Environmental Protection Agency  by SRI International,
     November 1979.

Singh, H. B., L.  J.  Salas,  and R.  E.  Stiles.  Distribution  of selected  gaseous
     organic mutagens  and suspect carcinogens  in  ambient  air.    Environ.  Sci.
     Technol. 16(12):872-880, 1982.

Systems Applications,  Inc.   Human  Exposure  to  Atmospheric  Concentrations of
     Selected Chemicals,  Vol II, as amended by  the U.S. Environmental Protection
     Agency to reflect public comments.   EPA contract  no. 68-02-3066, National
     Technical Information Service,  Springfield, VA,  PB83-265249, 1983.
005DC1/C                                  3-32                          12/7/83

-------
 Spence,  J.  W.,  P.  L.  Hanst, and B. W.  Gay,  Jr.   Atmospheric oxidation  of
      methyl  chloride,  methylene  chloride,  and chloroform.   J.  Air Pollut.
      Control  Assoc.  26(10):994-996,  1976.

 SRI  International.   Dichloromethane.  In:  Chemical  Economics Handbook.   SRI
      International,  Menlo  Park, CA,  198TJT

 U.S.  Environmental   Protection  Agency.   Preliminary assessment of  suspected
      carcinogens  in  drinking water.  Report  to Congress.   EPA-560/14-75-005.
      PB  260961,  1975a.

 U.S.  Environmental  Protection Agency.   Region V Joint Federal/State Survey of
      Organics  and Inorganics  in Selected  Drinking Water  Supplies, 1975b

 U.S.  Environmental   Protection  Agency.   Ambient Water Quality Criteria  for
      Halomethanes.   U.S. EPA.  Available  from:  National Technical  Information
      Service,  Springfield,  VA.  (NTIS PB81-117624), 1980a.

 U.S.  Environmental  Protection Agency,  Office  of Toxic  Substances.   Computer
      Printout.   Production Statistics  for Chemicals  in  the Non-Confidential
      Initial  TSCA  Inventory.   Retrieved November 14, 1980b.  Washington,  DC,
      U.S. EPA.

 U.S.  Environmental Protection Agency.  Environmental  Risk  Assessment of Dichloro-
      methane.   Draft report.   Office of Toxic Substances,  14 September 1981.

 U.S.  International   Trade  Commission 1980.   Synthetic  Organic  Chemicals.
      United States Production & Sales, 1979.  Washington,  DC.

 U.S.  International   Trade  Commission,  1982.    Synthetic  Organic  Chemicals.
      United States Production and Sales, 1981.  Washington, DC.

 Veith, G.D., K.J. Maule, and C.J. Petrocelli.  An evaluation  of using partition
      coefficients and  water solubility  to estimate bi concentration  factors
      for organic  chemicals in fish.  In: Aquatic Toxicology.  J.  G. Eaton,
      P. R.  Parrish,  A.  C.  Hendricks, eUs.  American  Society  for Testing  and
      Materials, 1979.  pp  116-129.

Wood,  P. R. ,  F.  Z.   Parsons, J. DeMarco,  H.  J. Harween, R.  F.  Lange,  I.  L.
      Payan, L.  M. Meyer, M.  C.  Ruiz, and  E.  D.  Ravelo.   Introductory study of
      the biodegradation of the chlorinated methane, ethane, and ethene compounds.
      Progress  Report,  submitted  to  the  U.S.  Environmental  Protection Agency,
      June 1980.

Wood, P.  R., F. Z. Parsons, J. DeMarco,  H. J. Harween, R.  F.  Lang,  I. L.  Payan
      and M.  C. Ruiz.   Introductory study of the biodegradation of the chlorinated
      methane,  ethane,  and  ethene  compounds.   Presented  at  the American Water
     Works  Association Meeting, June 1981.
005DC1/C                                  3-33                         12/7/83

-------
              4.   METABOLISM AND  PHARMACOKINETICS OF DICHLOROMETHANE

      Dichloromethane  (DCM) is a  colorless  liquid with a pleasant smell.  The odor
 threshold  is  about  100 ppm (347  mg/m3)  (May, 1966; Leonardos, 1969).  Because of
 its  relatively  high vapor pressure at room temperatures (350 to 500 torr), DCM is
 readily  absorbed  into the body following inhalation, and the great majority of
 severe poisoning  from this solvent occurs  from inhalation exposures.  Reports of
 poisoning  in  man  by oral ingestion are  rare (Llewellyn, 1966; Stewart and Hake,
 1976; Friedlander et al., 1978), and total recovery has followed the swallowing
 of quite large  doses  (Roberts et al., 1976).
      Dichloromethane has been used as a general industrial solvent for at least
 6 decades  (Lehmann  and Schmidt-Kehl, 1936), and  its  narcotic and  anesthetic
 properties  have been  known to clinical  medicine  for over  50 years  (Bourne and
 Stekle,  1923).  However, most of the available information on DCM's metabolism
 and  pharmacokinetics is derived  from studies that were conducted within the last
 10 years because  of the resurgence of interest in DCM following the demonstration
 of its metabolism to carbon monoxide (CO) (Stewart et al., 1972a,b).  These stu-
 dies  in animals and humans have been carried out at relatively high oral or in-
 halation concentrations  (100  to 1000 ppm) of  DCM.   Studies  of the pharma-
 cokinetics  of DCM for the low chronic exposure circumstances expected of the
 general ambient environment are  lacking.
 4.1  ABSORPTION,  DISTRIBUTION,  AND PULMONARY ELIMINATION
 4.1.1  Oral Absorption
     Absorption of DCM through the intestinal  mucosa after oral  ingestion appears
 to be rapid and complete.   McKenna and Zempel  (1981) obtained virtually complete
 recovery (92  to 96 percent) of 14C-DCM radioactivity in urine,  feces,  and exhaled
 air  of rats after single  oral  (gavage) closes of 1 or 50 mg/kg DCM  in water.
 Pritchard and Angelo  (1982) observed a  peak blood concentration occurrence at
 less than 10  minutes in B6C3F,  mice following  an oral  gavage  dose of 50 mg/kg DCM
 in either water or corn oil.
4.1.2  Dermal Absorption
     Absorption of  DCM through  the skin from  direct liquid contact or by the
 immersion of  hands  or arms is a  slow process.   Early animal studies by Schutz
 (1958) showed that DCM penetrates the skin and can be  absorbed  into the body by
this route.   Schutz exposed the  shaved skin of rats to direct  liquid  contact
005DC5/B                           4-1                           11-14-83

-------
for periods up to 20 minutes and found that kidney damage occurred after only
2 minutes of contact exposure.   In  contrast,  Stewart  and Dodd  (1964)  attempted to
quantify the rate of absorption through human skin by immersing the thumbs of
volunteers into liquid DCM and  then determining  the appearance and concentration
of DCM  in the breath.  An estimate of the amount of DCM entering the body was
made by  comparison  of  these  measurements  with breath concentrations  obtained
following controlled inhalation exposures.   From these breath  analyses,  Stewart
and Dodd concluded  that DCM is very slowly absorbed.  They also suggested that
although the amount of DCM absorbed depends  on the area of the exposed  skin,  the
slow rate of absorption would prevent toxic  quantities of DCM  from being taken
into the body from direct contact with the skin  of the hands and forearms.   Immer-
sion in DCM was found to be accompanied by excruciating pain within a few minutes,
which would serve as an effective deterrent.
4.1.3  Pulmonary Uptake
     Inhaled DCM rapidly equilibrates across the alveolar endothelium because
of its water and lipid solubility and the  very large  lung alveolar surface area.
DCM is  appreciably  more  water  soluble (2 grams per  100 ml),  but less  lipid
soluble  than  its  congeners,  chloroform and carbon tetrachloride;  its  lipid
partition coefficient  (olive oil/water)  is  180  at 37°C with a blood/air co-
efficient of 7.9 (Lehmann and Schmidt-Kehl,  1936;  Morgan et al., 1972;  Lindqvist,
1978).
4.1.3.1  Studies in Human—The  magnitude of  DCM  uptake into the body (dose, burden)
primarily depends on several  parameters:   inspired air concentration, pulmonary
ventilation, duration of exposure,  the rates of  diffusion into blood and tissues
and solubility  in blood and the various tissues.  The concentration  of  DCM in
alveolar air, in equilibrium with pulmonary  venous blood content, approaches  a
minimum difference with the concentration in the inspiratory air until  a steady-
state condition is reached.   After  tissue and total body equilibrium is reached
during  exposure, uptake  is  balanced by elimination through the  lungs and  by
other routes,  including  metabolism.   The  difference  between alveolar and in-
spiratory air  concentrations,  together with  the  ventilation  rate (about  6
1/min at rest), provides a means of calculating  uptake during  exposure:

                          Q = (C.    - C , ) V • T                    (4-1)
                          x   v insp    alvy
005DC5/B                           4-2                           11-14-83

-------
 where Q is  the  quantity  absorbed,  C  is  the  concentration  (inspired and alveolar)
 in mg/1,  V  is the  alveolar  ventilation  rate in  1/min, and T  is the duration of
 exposure  in minutes.   The percent  retention is  defined as (C.    - C  , )/C.    x
                                                             insp    alv'   insp
 100,  and  percent retention  x  quantity inspired  (V  • T -C.    ) is equal to  uptake.
      Figure 4-1 (from Riley et al.,  1966) illustrates the overall  time-course
 of absorption and  elimination during a  2-hour inhalation exposure of  100 ppm  (347
 mg/m^)  DCM  for  a 70-kg man.   During exposure, the  alveolar air concentration  of
 DCM can be  described  by an exponentially rising curve with  three components.
 At the beginning  of exposure, an initial  rapid  rate of  uptake occurs  [0  to 50
 ppm (0 to 173 mg/m^) alveolar air],  followed by a  second  slower  uptake  [50 to
 65 ppm  (173 to  208  mg/m3) alveolar air], and finally a very  slow rate occurs  as
 equilibrium is  approached at  70  ppm (243 mg/rn^) alveolar air concentration.   The
 total  quantity  of  DCM absorbed  and  retained  in the body during exposure  is
 represented by  the  area between  the alveolar and inspired environmental air con-
 centration  curves.   Complete  equilibrium or steady-state conditions  are  not
 attained  by the end of the  2-hour  exposure  to DCM, as shown  by the slowly  rising
 alveolar  concentration curve.   The three uptake compartments of the exponential
 alveolar  curve  correspond to  equilibrium  attained  by first-order passive  dif-
 fusion  of DCM from blood.   First,  DCM diffuses through a vessel-rich group (VRG)
 of  tissues  with high blood  flow  (VRG:  brain, heart, kidneys, liver, and endo-
 crine and digestive systems),  then it diffuses more slowly through the lean body
 mass  (muscle group, MG:  muscle  and skin) and last through adipose tissues (fat
 group,  FG).   With the termination  of exposure, blood and alveolar air DCM  concen-
 trations decline in parallel  in  an exponential manner with three components of
 pulmonary elimination and desaturation from VRG, MG, and FG body compartments.
 The area under the alveolar elimination curve in Figure  4-1  is proportional to
 the quantity of DCM absorbed during the 2-hour exposure.
     DiVincenzo and Kaplan  (1981a) obtained serial  breath excretion curves from
 four to six volunteers experimentally exposed to 50, 100,  150,  and 200 ppm (173,
 347,  520, and 694  mg/m^) DCM for 7.5 hours under sedentary conditions.   Pul-
monary  uptake was  rapid  during the first hour  (exposure  was interrupted at
4  hours,  for 30 minutes).   During  post-exposure, DCM levels  decreased rapidly
 in  an exponential  manner.   Less  than 0.1 ppm (0.347 mg/m3)  DCM was detected
 in  the  end  tidal air of the individuals exposed to 50,  100,  or  150 ppm (173,
005DC5/B                           4-3                           11-14-83

-------
                DURING

              EXPOSURE
          AFTER EXPOSURE
         100
      ui
      a

      E
      o


      o
      ui
      z
      ui
                             ROOM CONCENTRATION
             RESPIRATORY

              ABSORPTION!
                                   ABSORPTION AND EXCRETION

                                    OF METHYLENE CHLORIDE
EXHALED AIR CONCENTRATIONS
                                     .	

                          RESPIRATORY EXC~REtT6N
                                  TIME, hours.
Figure 4-1.   Inspired and expired air concentrations during a 2-hour, 100-ppm

             inhalation exposure to DCM for a 70-kg man, and the kinetics of

             the subsequent pulmonary excretion.


             Source:   Riley et al. , 1966.
005DC5/B
      4-4
11-14-83

-------
 347,  or 520  mg/m3)  DCM 7  hours  after  exposure was  terminated.   For  the  200  ppm
 (694  mg/m3)  exposures,  the  mean post-exposure end  tidal  air  concentration of  DCM
 decreased  to 1  ppm  (3.47  mg/m3) at  16 hours.  Post-exposure  elimination of  DCM
 was  less than 5 percent  of the amount absorbed.   A  related  exposure  study  in
 which volunteers were  exposed to 100,  150,  and  200 ppm  (347,  520, and 694 mg/m3)
 DCM for 7.5  hours daily,  for 5  consecutive  days, also indicated that  uptake and
 elimination  was directly  proportional  to  the magnitude  of  exposure, thus confirm-
 ing previous findings  by DiVincenzo  et al. (1972) and  Stewart et al. (1973).
      The retention  of  DCM  as a percentage  of  inspired air  concentration is
 independent  of  that concentration at  equilibrium.   Retention values  for DCM,
 reported by  different  investigators,  are  shown  in  Table 4-1.  These values  have
 a  large range and vary with the duration  of exposure.   Variation in the values
           TABLE 4-1.  PULMONARY ABSORPTION OF DCM BY HUMAN SUBJECTS
                             (SEDENTARY CONDITIONS)
Inhalation
concentration,
Investigator ppm
DiVincenzo and Kaplan, 1981a



Lehmann and
Schmidt-Kehl, 1936



Riley et al. , 1966
DiVincenzo et al. , 1972


Astrand et al . , 1975

50
100
150
200

662
806
1,152
1,181
44-680
100
100
200
250
500
Exposure,
hr
7.5
7.5
7.5
7.5

0.3
0.5
0.5
0.5
2
2
4
2
0.5
0.5
Retention,
%
70
60
63
60

74
75
72
72
31
53
41
51
55
55
Engstrom and
  Bjurstrom, 1977
750
     34
005DC5/B
4-5
11-14-83

-------
also is caused by differences in body weights  of  the  subjects  and  differences  in
body composition (proportion of  adipose  tissue to lean mass).   For  exposures
greater than  1  hour,  the  mean  retention  approximates 42 percent of  uptake  of
DCM or approximately 125 mg/hr  for an exposure of 100 ppm (347 mg/m3),  assuming
a resting ventilation rate of 6 1/min.
     For short exposures,  the quantity (dose)  of  DCM  absorbed  into the  body is
theoretically directly proportional to the concentration of DCM in the  expired
air.  This relationship has been confirmed experimentally (Lehmann and  Schmidt-
Kehl, 1936; Astrand et al., 1975).  The body burden of DCM also increases with
exposure  duration  and with  physical  activity (increased  ventilation  and
cardiac output)  at  a  given inhaled air concentration  (Engstrom  and Bjurstrom,
1977; Astrand et al., 1975; DiVincenzo and Kaplan, 1981b).   Astrand et  al.  (1975)
found that physical  activity during  exposure to  250  and 500 ppm DCM (869 and
1737 mg/m3) for 0.5 hour decreased retention from 55  percent in a  resting stage
to  40  percent during  activity,  but doubled  the amount of DCM  absorbed because
of  a three-fold increase of ventilation rate (6.9 to  11 I/minute).  DiVincenzo
and Kaplan (1981b) also found that physical exercise  during exposure increased
pulmonary  uptake.   Exposure  of  three males  to  100  ppm DCM  (347  mg/m3) for  7.5
hours, during which they exercised on a treadmill for 5 minutes of each 15-min
period, resulted also in an estimated doubling of average cardiac  output and as
much as an eight-fold increase in the average alveolar ventilation rate (Table 4-2).
Effects upon  blood carboxyhemoglobin (COHb) and conversion of DCM to CO are dis-
cussed elsewhere in this chapter.
     The quantity of  DCM absorbed is dependent also on body weight and fat con-
tent of the body.  Engstrom and Bjurstrom (1977) showed that for  an exposure to
750 ppm  DCM  (2606  mg/m3)  for 1  hour,  the  amount  of DCM absorbed into the body
was directly  proportional  to body weight and  to body fat content expressed
as  a percentage  of body weight.   Obese subjects  (average body fat is 25 percent
of  body weight)  absorbed 30 percent more DCM than  lean subjects (average body  fat
is  8 percent  of  body  weight).  Biopsy analysis of  subcutaneous  adipose tissue  of
obese  subjects  revealed concentrations of 10.2 and 8.4 mg  DCM/kg  tissue weight
at  1 and 4 hours post-exposure, respectively.  These concentrations,  although
lower  than those found  in  adipose tissue of lean  subjects, represented a great
total  storage amount in obese  subjects  because  their total  fat  stores  were
greater.   Significant amounts  of DCM were found  in adipose tissue (1.6 mg/kg)
of  obese subjects 22 hours after exposure,  indicating that elimination of DCM
 005DC5/B                           4-6                            11-14-83

-------
          TABLE 4-2.  EFFECT OF EXERCISE ON PHYSIOLOGICAL PARAMETERS
                         FOR VOLUNTEERS EXPOSED TO DCM


Work
intensity
(ml 02 min
Volunteer (kg-1)
1

2


3


Source:

4
14
4
15
19
4
16
28
DiVincenzo and


Aerobic
capacity, %

25

25
45

45
70
Kaplan (1981b).
, .

Heart
rate,
beats/mi n
56
84
68
96
119
68
123
145


Estimated
average
cardiac
output,
1/min
5
11
5
12
14
5
15
21


Estimated
average
alveolar
venti lation,
1/min
8
18
6
34
45
6
28
46

* •
 to equivalent Og consumption measured during pre-exposure aerobic capacity
 testing.  Oxygen consumption at rest was taken as (4 ml 02/min • kg) body
 weight, based on preexposure testing.

from the  FG  compartment proceeds at  a slow  rate.  Furthermore, the  residual
DCM in  adipose  tissue is additive to the next day exposure, particularly  in
obese people.  However, with multiple daily exposure, a total body equilibrium
and a constant  adipose tissue concentration are  eventually  achieved with a
given concentration of DCM in the inspired air.
4.1.3.2  Studies in Animals--McKenna et al.  (1982) used 250-gram male Sprague-
Dawley  rats  to  estimate the total body burden of DCM resulting from a 6-hour
inhalation exposure of 50,  500,  and  1500 ppm (173,  1737, and 5211 mg/m^)  of
14C-DCM.  After 6  hours,  the animals were in apparent total  body equilibrium
with the  inhaled  concentrations  of radioactive DCM as suggested by "plateau"
blood concentrations  (Figure  4-2).   The  body burden was calculated  from the
total  radioactivity recovered  from exhaled  air,  urine and feces, and carcass
analysis, during the  first 48  hours after exposure.  Table 4-3  shows the body
burden associated with each exposure.   The increase in body burden of 14C-acti-
vity was less than proportionate to the increments of DCM inhaled concentration.
005DC5/B
4-7
11-14-83

-------
          10.0
      o
      O
       M

      O
      <
                                                          A 50 ppm
                                                          • 500 ppm
                                                          • 1500 ppm
                                        I   I   I   I   I   I   I
          0.01 —
         0.001
Figure 4-2.   Plasma levels of DCM in rats during and after  DCM  exposure
             for  6  hours.   Data points  represent  mean ± standard  deviation
             for two to four rats.

             Source:  McKenna et al., 1982.
005DC5/B
4-8
11-14-83

-------
                TABLE 4-3.  BODY BURDENS OF RATS AFTER INHALATION
                        EXPOSURE TO 14C-DCM FOR 6 HOURS
Exposure
concentration,
ppm
50
500
1500

Number of
rats
3
3
3
Total body burden
mg Eq 14ODCM/kg
± S.D.
5.53 ± 0.18
48.41 ± 4.33
109.14 ± 3.15
Source:  McKenna et a!., 1982.
S.D. = standard deviation.
4.1.3.3  Blood/Air Relationship—The  blood  concentration  of  DCM during  inhal-
ation  and  in the  elimination phase  after  exposure  parallels  alveolar DCM
concentration.  This  predictable  relationship  is  defined  by  the solubility  of
DCM.  Astrand et al.  (1975) showed that for men exposed to 250 and 750 ppm DCM
(869 and 2606 mg/m^)  for 1.5 hours,  the  arterial  blood (mg/1) to alveolar
air  (mg/1)  concentration ratios were  constant  and averaged 10.3 and 11.1,
respectively,  over three-fold changes in  alveolar  concentrations.   These
jjn  vivo Ostwald  coefficients  agree with the value found  for blood/air  (7.9)
and for water/air (7.2) (Lindqvist, 1978; Morgan et al., 1972) at 37°C in vitro.
The high water/air coefficient suggests that DCM  is  dissolved  in plasma water
as well as in lipid components of blood.
     In the exposure study of DiVincenzo and Kaplan (1981a) described previously
(Section 4.1.3.1), exposure and post-exposure blood concentrations of DCM were
directly proportional  to  the  magnitude of  exposure.   Blood/air coefficients
calculated from the data suggest that the 200 ppm (694 mg/m^) exposure concen-
tration may be approaching that level at which saturation of metabolism occurs,
as evidenced by an increased level of DCM in venous blood.
     MacEwen et al. (1972) determined the blood DCM concentration in dogs that
were continuously exposed for 16 days to 1000 and 5000 ppm DCM (3470 and 17,370
mg/ma).  The  blood DCM concentrations were 36 and 183 mg/1, respectively, in
direct proportion to exposure concentration.  Total equilibrium can be assumed to
have occurred in these animals.   Ostwald coefficients of 10.4 and 10.5 for the
005DC5/B                           4-9                           11-14-83

-------
two exposure concentrations agree with the above values noted in man.  Similar
values can be calculated from the data of Latham and Potvin (1976); Figure 4-3
shows the proportional  relationship  they found in rats between DCM blood and
inspired  air  concentrations over a  range of 1,000 to 8,000  ppm (3,474 to
27,792 mg/m3) during a 6-hour exposure.
     In contrast to  these  findings  in man and other animals of a direct pro-
portional  relationship  between  inspired air concentration  of  DCM and blood
level, McKenna et al. (1982) reported a greater than proportionate increase of
blood concentration with inhalation exposure concentration  in rats.  Table 4-4
gives the apparent steady-state concentration of DCM in plasma and whole blood
of Sprague-Dawley rats exposed for 6 hours each to 50, 500, and 1500 ppm DCM (173,
        o>
        fc 0.60
        o
        CM
        U
        O  0.40
        ui
        _i
        O
        O
        O
        _J
        CO
0.20
                     INHALATION
                        1000       2000       4000       8000
                          EXPOSURE CONCENTRATION, ppm
 Figure  4-3.   DCM  venous  blood  levels  in  rats  immediately after a single
              6-hour  inhalation exposure  to  various  concentrations of DCM.
              Source:   Latham and  Potvin,  1976.
 005DC5/B
                         4-10
11-14-83

-------
         TABLE 4-4.  DCM CONCENTRATIONS IN RAT WHOLE BLOOD AND PLASMA
      AT APPARENT STEADY-STATE CONDITIONS OF A 6-HOUR INHALATION EXPOSURE

  Exposure                      DCM concentration         Plasma/blood
concentration,  Number of   	pg/ml ± S.D.	        distribution
ppm
50
500
1500
rats
3
3
3
Whole blood
0.22 ± 0.04
-
39.53 ± 3.71
Plasma
0.05 ± 0.01
2.38 ± 0.42
8.94 ± 0.39
coefficient
0.23
-
0.23
Source:  McKenna et al., 1982.
S.D. = standard deviation.

1737, and 5211 mg/m3).  The data indicate that the whole blood and plasma levels
of DCM increase disproportionately with an increase in inspired air concentra-
tion.  Furthermore, calculation  of  the  blood/air  ratio for these  data provide
increasing values  of  5.75,  5.97, and 7.59 for  50,  500,  and 1,500 ppm (173,
1737, and 5211 mg/m3), respectively.  McKenna et al. suggest that the resultant
increase in blood DCM concentration is greater than that predicted by increments
in the inspired air because of a rate-limited metabolism of DCM in the rat.   Thus,
at low inspired air concentrations (below saturation of metabolism), the blood/air
ratio is less (because of rapid metabolism) than that at high inspired air con-
centrations (above saturation, of metabolism); and indeed, the blood/air coeffi-
cient (7.59)  observed at  1,500 ppm  DCM  (5,211 mg/m3) agrees with  coefficients
observed by others.
4.1.4  Tissue Distribution
     Because of DCM's water solubility,  it probably distributes throughout the
body water, and  its  lipid solubility allows its  distribution  into all  body
tissues and cellular lipids and particularly into adipose tissue.  Engstrom and
Bjurstrom (1977)  determined  that the Ostwald  coefficient for subcutaneous
adipose tissue from human buttocks  is 51 at 60°C;  this  value  indicates that
the  tissue/blood  partition coefficient  may be  about  7  at body temperature
(37°C).
     DCM readily crosses  the  blood-brain barrier  even at relatively low vapor
exposure concentrations,  as evidenced by its impairment of manual and mental
performance at 500 ppm (1737  mg/m3) (Winneke and  Fodor,  1976; Winneke, 1981).
005DC5/B                           4-11                          11-14-83

-------
DCM also crosses the placenta and may affect fetal  development (Schwetz et al.,
1975; Anders and Sunram, 1982).
4.1.4.1  Animal Studies—Tissue concentrations of  DCM  increase with exposure
concentrations and  duration  and,  for any given tissue, are dependent on the
largely  unknown  tissue partition coefficients.   Savolainen et  al.  (1977)
exposed  rats  chronically  to  DCM [200 ppm (695 mg/m3),  6  hours daily,  for 5
days] and determined  DCM  concentrations  in peri renal fat and other tissues.
The resulting data, shown in Table 4-5,  indicate  that significant amounts of DCM
remained in peri renal  fat 18 hours after the previous  exposure of day  4, and
markedly increased  further with a 6-hour exposure on day 5. At termination  of
the  last 6-hour  exposure  (fifth daily exposure), the ratio  of tissue to  blood
concentrations was  about  one for brain  and liver,  but  6.6 for perirenal fat.
     Previous work by DiVincenzo et al.  (1972) in man and Carlsson and  Hultengren
(1975) in rats indicated that little uptake by adipose  tissue occurs.   However,
single, short exposure periods of 2 hours were used in  their studies,  and of the
smaller amount absorbed, 95 percent was  probably accomodated in VRG and MG com-
partments because the FG compartment receives only 5 percent of the cardiac out-
put.

            TABLE 4-5.  TISSUE CONCENTRATIONS OF DCM IN RATS EXPOSED TO
                      200 ppm DCM FOR 4  DAYS FOR 6 HOURS DAILY3
Exposure
DCM concentrations
nmoles/g tissue wet weight ± S.D.
on the fifth
day, hr
0
2
3
4
6
Tissue to
Cerebrum Cerebellum
-
73 ±20 57 ± 20
119 ± 33 86
57 ± 8 95 ± 8
83 90
blood coefficient after the fi

Blood Liver
-
90 ± 10 85 ± 2
79 ± 3 82 ± 1
120 ± 10 101 ± 13
100 ±1 83 ± 10
fth daily exposure
Perirenal
fat
113 ± 29
526 ± 94
537 ± 33
608 ± 58
659 ± 77

                0.83:1        0.90:1       1.0:1       0.83:1         6.59:1
aSavolainen et al., 1977.
S.D. = standard deviation
 005DC5/B                           4-12                          11-14-83

-------
     McKenna and coworkers (1981; 1982) have studied DCM distribution in tissues
of rats by observing 14C-activity 48 hours after a single 6-hour inhalation of
50, 500, or 1500 ppm DCM (173, 1737, and 5211 mg/rn^) and after single oral doses
of 1 or 50 mg/kg.  The results are shown in Table 4-6.   Following the inhalation
exposures and the oral dosage, the highest concentrations of 14C-activity were
found in the liver, kidney, and lung.  The observed 14C-activity in epididymal
fat was  consistently  lower than that observed  in  either whole blood or the
remaining carcass.  The tissue 14C-activity increased with dose, but it did not
increase in direct proportion with increasing dose.  Intact DCM was not detected
in any of the tissues assayed; therefore, the observed radioactivity is presumed
to represent nonvolatile metabolites of DCM.  These results contrast with those
of Savolainen et al. (1977), who found that in perirenal adipose tissue, signifi-
cant amounts of intake DCM remained 18 hours after the last of four daily 6-hour
inhalation exposures to 200 ppm DCM (895 mg/m3) (Table 4-5).
     The results  of  the  experiments of Savolainen  et al. and  McKenna  et  al.
indicate that total body equilibrium of DCM to inspired air concentration is not
achieved in the  rat within  a  single  6-hour  inhalation period even  though  this
response is suggested by the achievement of a "plateau"  blood concentration as
shown in Figure  4-2.   Of interest also is the evident difference between the
amount of tissue  14C-metabolites  associated with inhalation and oral dosage.
A close  correspondence exists for tissue metabolites of DCM for an oral dose
of 50  mg/kg and  a 6-hour inhalation exposure  of  50  ppm,  although 50  ppm
(173 mg/m^) provides  a  terminal  body burden of  only  5.5 mg/kg (Table 4-3).
4.1.5  Pulmonary Elimination
     Pulmonary excretion  is the mechanism  of  elimination of  virtually all
unchanged DCM from  the  body.   Less than 2 percent of estimated body doses of
DCM have been detected  as unchanged compound  in the urine of  human  subjects
exposed to  100 ppm and 200 ppm (347  and 695 mg/m^)  for 2 hours (DiVincenzo  et
al.,  1972)  and  in  the  urine of  dogs  exposed to  5000  ppm  (17,370 mg/m^)
(MacEwen et al.,  1972).
4.1.5.1  Studies  in Humans—Figure 4-1 shows schematically the time-course of pul-
monary elimination of DCM after inhalation exposure.  The parameters of elimin-
ation equilibration of the body are  the same as those of assimilation  equili-
bration.   After termination of exposure, DCM immediately begins to be eliminated
005DC5/B                           4-13                          11-14-83

-------
from the body via the lungs.  Alveolar air equilibrates with pulmonary venous
blood whose concentration  becomes  a  function of the first-order diffusion of
DCM from tissues, the arterial blood flow/tissue mass, and the relative solu-
bilities of DCM in tissues.  Figure 4-1 shows that alveolar DCM concentration
follows an  exponential  decay curve with  three  major components reflecting
desaturation of the VRG,  MG,  and FG compartments,  respectively.   The  half-
times of  elimination  of  DCM  from  these compartments have not  been firmly
established.  Riley et  al.  (1966)  measured expired  air  concentration  after
termination of  exposure and  found  half-times of 5  to  10  minutes  for the
VRG compartment, 50 to  60 minutes  for the MG compartment,  and 400 minutes
for the FG compartment.   DiVincenzo et al.  (1972),  who exposed subjects to 100
ppm and 200 ppm (347 and 695 mg/m3) for 2- and 4-hour periods, felt that "very

   TABLE 4-6.   DISTRIBUTION OF 14C-ACTIVITY IN TISSUE 48 HOURS AFTER 6-HOUR
               INHALATION EXPOSURE  OR ORAL DOSAGE OF RATS TO  14C-DCM
Exposure method
      mgEq 14C-Activity by exposure
       concentrations, tissue ± S.D.
Inhalation
Liver
Kidney
Lung
Brain
Epididymal fat
Skeletal muscle
Testes
Whole blood
Remaining carcass
Oral
Liver
Kidney
Lung
Brain
Epididymal fat
Skeletal muscle
Testes
Whole blood
Remaining carcass
50 ppm
8.4 ± 1.5
3.3 ± 0.1
1.9 ± 0.2
0.8 ± 0.3
. 0.5 ± 0.2
1.1 ± 0.1
1.1 ± 0.2
1.1 ± 0.2
1.3 ± 0.2
1 mg/kg
0.40 ± 0.04
0.15 ± 0.01
0.99 ± 0.01
0.04 ± 0.06
0.02 ± 0.002
0.08 ± 0.02
0.04 ± 0.002
0.06 ± 0.003
0.05 ± 0.002
500 ppm
35.6 ± 7.5
16.2 ± 2.4
11.0 ±1.3
4.2 ± 1.3
6.5 ± 0.5
4.4 ± 1.9
5.5 ± 1.3
8.1 ± 1.9
5.6 ± 0.9
50 mg/kg
6.67 ± 0.69
2.92 ± 0.21
1.67 ± 0.16
0.63 ± 0.10
0.33 ± 0.0006
0.86 ± 0.04
0.92 ± 0.10
1.41 ± 0.11
0.99 ± 0.60
1500 ppm
44.2 ±3.5
30.5 ± 0.2
16.5 ±1.6
6.7 ± 0.2
4.1 ± 0.9
7.7 ± 0.7
8.1 ± 0.5
8.9 ± 1.7
8.6 ± 1.4










Source:   McKenna et al., 1982; McKenna and Zempel,  1981.

S.D.  = standard deviation.
                           t
Number of animals in each group = 3
005DC5/B
4-14
11-14-83

-------
 little vapor" reached the fat stores and muscle tissues under these conditions.
 DiVincenzo  et  al.  found DCM to have a half-time value in blood of 40 minutes
 following 2 hours of exposure and prolonging exposure to 4 hours had no signi-
 ficant effect on the half-time.  DiVincenzo and Kaplan (1981a,b) have recently
 extended previous  studies by following the  course  of  pulmonary  elimination  in
 individuals exposed  for 7.5  hours  to  50,  100,  150,  and 200 ppm  DCM  (173,  347,
 520, and 694 mg/m3) and those exposed for 7.5 hours daily for 5 consecutive days
 to the three highest concentrations.  The authors concluded that post-exposure
 elimination of DCM in breath is a minor route of elimination over the concentra-
 tion range used; post-exposure elimination was less than 5 percent of the amount
 absorbed.  Pulmonary elimination of DCM was rapid at all  exposure concentrations.
 Less than 0.1 ppm (0.347 mg/m3) was detected in post-exposure breath samples at
 7 hours.  Die-away of DCM in blood paralleled that observed with pulmonary eli-
 minations.   Breath elimination times were prolonged by physical exercise.  Morgan
 et al.  (1972),  using  isotopically  labeled DCM, estimated the  half-time  of DCM
 in the  VRG  compartment  as 23 minutes.  Engstrom and Bjurstrom (1977)  reported
 the half-time of the MG compartment for lean subjects was about 60 minutes, and
 they obtained a longer time value for obese subjects.   For both lean and obese
 subjects, biopsies indicated that residual concentrations of DCM existed in adi-
 pose tissue nearly  24  hours  after exposure, suggesting that the half-time of
 elimination from the  FG compartment is fairly long.  From the  post-exposure
 alveolar concentration  curves prepared by Stewart  et  al. (1976a) for  subjects
 exposed to inhalation concentrations of 50 to 500 ppm DCM (173 to 1,736 mg/m3)
 for 1 to 7.5 hr/day for 5 successive days, the half-time of elimination for MG
 and FG  can  be  estimated as  60 to 80 minutes,  and  240 minutes,  respectively.
 Thus, the best  guesses  from  these studies  for the  half-times of elimination
 from the VRG,  MG,  and FG are 8 to 23 minutes, 40 to 80 minutes, and 6 to 6.5
 hours, respectively.   The long half-time  from elimination of the adipose tissue
 compartment and reports that DCM remains  in this  compartment 24 hours after sin-
 gle and chronic  exposures  indicate that  the concentration of DCM in  adipose
 tissue may only slowly,  over a period of  days, achieve equilibrium with inspired
 air concentration,  particularly of daily  exposures of short duration.
4.1.5.2  Studies in Animals--The estimates of half-times  of pulmonary excretion
of DCM in animals are derived from postexposure alveolar concentration curves.
Blood DCM decay concentration curves are  a less reliable  parameter of pulmonary
005DC5/B                           4-15                          11-14-83

-------
excretion because these  curves  are  influenced by metabolism,  the other major
route of elimination  of  DCM.   McKenna et al.  (1982) determined the pharmaco-
kinetic half-times for elimination of DCM from plasma of rats  following a single
inhalation exposure to 500 or 1500 ppm (1737  or 5211 mg/m3) for 6 hours.   Figure
4-2 shows that two first-order processes governed the disappearance of DCM from
plasma with apparent half-lives of 2 and 15 minutes for the rapid and slow com-
ponents of elimination,  respectively.  The kinetic  parameters calculated from
these curves are summarized in Table 4-7.
     The kinetic parameters for rats differ greatly from those of man.  Presumably,
the half-times for the fast and slow components correspond to  the half-times for
VRG and MG compartments of man.  The half-time for the FG or adipose tissue com-
partment was not established in the rat because plasma decay was only followed
for 1.5 hours following inhalation exposure (Figure 4-2).  The rates of elimi-
nation of  DCM  from  plasma  in  the  rat  are first order  and independent  of  dose,
although there is considerable evidence of extensive metabolism in these animals
(Section 4.2).
         TABLE 4-7.  COMPARISON OF POST-EXPOSURE PULMONARY ELIMINATION
                     HALF-TIMES OF DCM FOR HUMANS AND RATS
Subject
  Exposure        DCM
   method     concentration
                                             ^ first-order components (minutes)
                                             Alpha       Beta       Gamma
Human
Rat
Inhalation    50-500 ppm
                                             8-23
Inhalation
500 ppm (6 hr)
                                              40-80
                                          360-390


Rat Gavage


1500 ppm (6 hr) 2.4
1.06
1 mg/kg
50 mg/kg

14.8
15.5

12.6
12.6
-
~"

46.5
46.6
      McKenna  and Zempel (1981) also have determined the kinetics of pulmonary
 excretion  of  DCM in rats  after  a  single  oral  dose  (via gavage)  of  1  or 50  mg
 14C-DCM per body weight.   Figure 4-4  shows  the time course  (for  5  hours) of DCM
 005DC5/B
                        4-16
                                                                  11-14-83

-------
 in  exhaled  air  of  these  rats.   Pulmonary elimination of DCM of rats receiving 1
 mg/kg  dose  is characterized by a  two-component  decay  curve  representing two
 first-order processes  with half-times of elimination of  12.6  minutes  and 45.6
 minutes,  respectively.   The pulmonary excretion of DCM  of rats  given the 50
 mg/kg  dose  approximates  zero-order kinetics for the first hour after admini-
 stration  (possibly rate-limited by absorption)  and then  describes the  same
 first-order kinetics of the 1  mg/kg  dose.   These  results suggest that  the
 half-time of pulmonary elimination from the FG compartment for rats approxi-
 mates  45  minutes,  versus 6 to  6.5  hours  for man.   However,  as noted  for  rats,
 single  inhalation  or oral  exposures in man  probably do not result in total  body
 equilibrium,  particularly  with the fat compartment, and  elimination  from this
 compartment may be longer  than  indicated from these experiments.
     Withey and  Collins  (1980) determined  the kinetics of elimination of DCM
 from the blood of Wistar rats after intravenous administration of 3, 6,  9,  12,
 or  15  mg/kg of  DCM given intrajugularly  in  1 ml  of water.   For the  four  lower
 doses,  the  blood decay curves best fitted a  first-order two-compartment  model,
 but the highest dose fitted a first-order three-compartment model.  Withey  and
 Collins suggested  that the apparent change  from a two- to three-compartment
 system  occurred  as a consequence of a "difference in  biological  response to
 the dose  either as a consequence of  the  magnitude of  the dose or,  in some
 cases,  due  to the  varying response of different animals  to the same dose."
 This interpretation  is probably true  because the k (rate  constant for elimi-
 nation  from blood out of body,  principally via pulmonary excretion and/or meta-
 bolism) was relatively high for the lower doses and consistent with a half-time
 of body elimination of only 4.5 minutes.   Any significant equilibration  of  blood
 DCM with the  fat compartment was very unlikely during these experimental  circum-
 stances.  Further support for limited distribution  of DCM is the reported kinetics
 for the highest dose (15 mg/kg) that  fit a  three-compartment  system.   In this
 case,  k  was 0.09/min,  that is, a half-time of body elimination of 7.7 minutes.
 This result reflects  that more  time was required for equilibration into  fat to
 occur.   The half-time of elimination from the fat compartment was given  as  31.5
minutes, although the rate constant k5l [movement of DCM from the fat compartment
to the central compartment (blood)] was given as 0.006/min, indicating a half-
time of 115 minutes.   The volume of distribution (Vd)  of DCM was  calculated as
48.8 ml or about 12 percent of  body weight,  which was  surprisingly low for both
a water- and lipid-soluble compound that  is  known to  diffuse into all  the major
organ  systems.

005DC5/B                           4-17                          11-14-83

-------
                100.0
                 10.0
              i
             s>
             3   1.0
             (0
             o
             o
                  0.1 —
                    0
234

  TIME, hours
Figure 4-4.  Pulmonary elimination of  14C-DCM  following  oral  administration
             to rats of a single dose  of 1 or  50 mg/kg 14C-DCM (squares).
             Ordinate is percent dose  administered  during  0.5-hour collection
             periods with means ± SEM  for groups of three  rats.   SEM is the
             standard error of the mean.
005DC5/B
  4-18
11-14-83

-------
4.2  METABOLISM OF DCM
     DCM is known to be metabolized in man to CO and in experimental animals to
CO, carbon dioxide (C02), formaldehyde, and formic acid. The CO production results
in an  elevation  of blood  COHb  content,  from which CO dissociates  in the  lung,
followed by elimination of CO.   Experiments in man and animals have shown that
the metabolism of DCM is dose-dependent and limited by hepatic enzyme saturation.
4.2.1  Evidence for Metabolism to Carbon Monoxide
     Before 1972, most  of the  absorbed dose  of  DCM in  man was thought to be
excreted unaltered in exhaled air, while a small  amount was found in the urine
(MacEwen et al., 1972; DiVincenzo et al., 1972; Heppel  et al., 1944).   Metabolism
of DCM to CO was not known to occur.  However, in 1972, Stewart et al.  discovered
that the COHb  concentration  in blood increased  in persons exposed  to 200 to
1000 ppm DCM (695 to 3474 mg/m3) from 1 to 2 hours.   The COHb levels continued
to rise  beyond cessation  of exposure and decreased more  slowly  than  when
similar levels of COHb were induced by breathing CO.   Stewart and his associates
(1972a,b) proposed that  CO was the end product  of  DCM metabolism.   In  the
following year,  Fodor et  al.  (1973) showed that  COHb  blood  levels from DCM
exposure were  further elevated by concomitant exposure to diiodomethane and
dibromomethane, thus indicating that these dihalomethanes also are metabolized
to CO.   At first, this unique halocarbon metabolism was not generally accepted,
because the increased COHb levels might reflect  a change in the rate of  endo-
genous CO production  or  excretion that is associated  with  heme  degradation
by the microsomal hepatic heme  oxygenase system in man (Coburn, 1973;  Tenhunen
et al., 1969).   However,  Stewart et al. (1972a,b, 1973) observed no evidence
of an  enhanced  metabolism of hemoglobin in their subjects  exposed to DCM and
the subjects did not excrete increased  amounts of urobilinogen in their  urine
either during or after exposure.
     A second  hypothesis  put forth  to explain the origin of  the  excess  COHb
postulated a DCM-induced conformational  change in hemoglobin,  which increased CO
affinity.   Hence, more COHb would be formed,  with a longer  biological  half-life,
from endogenous and exogenous  CO sources at the  same  ambient concentrations.
Settle (1971)  had shown that xenon  and, to a much larger degree,  cyclopropane
bind to myoglobin and  increase  CO affinity.   Following this observation,  Nunes
and Schoenborn (1973)  used X-ray diffraction to demonstrate that  DCM also binds
to sperm whale myoglobin  and suggested that this binding  might  increase CO
affinity.   Settle (1975) then  investigated the CO binding to  human  hemoglobin
005DC5/B                           4-19                          11-14-83

-------
(Hb).   He observed that CO binding to human Hb in vitro at 37°C is increased in
the presence of  DCM  (1000 ppm,  3474 mg/m3).  The amount of COHb formed at a
given CO concentration  is doubled.  Also, measurement of the P50 values of Hb
(partial pressure of 02 or CO at which 50 percent saturation occurs at 37°C) in
the presence and absence of DCM  (1000 ppm) indicated a six-fold increase in CO
affinity.   From these observations, Settle (1975) suggested that the increased
COHb seen in vivo may be caused  by an increase in CO affinity and not to meta-
bolism.   More recently, Collison et al.  (1977) determined the Haldane affinity
constant for CO for both human and rat blood equilibrated at 37°C with air con-
taining only CO and CO plus DCM  (10,000 ppm).   No difference in the Haldane con-
stants  for  human  blood (mean value, 227) or  rat blood (mean value, 179) was
found.  Negative results also were obtained when the absolute affinity of CO was
measured in a nitrogen-DCM atmosphere.   Dill et al.  (1978) redetermined the P50
value for human and rat blood in the presence and absence of CO (2500 ppm) and
DCM (800 ppm).   DCM was found to  have no effect on the P50 values; therefore,
Dill concluded that DCM does not significantly increase the affinity of CO for
hemoglobin.
     The biotransformation of DCM to CO and CO^ has  now been confirmed by many
metabolism  studies.   The  hepatic  metabolism  of  DCM has  been  unequivocally
shown to be  the  origin of the CO  responsible for  increased COHb blood con-
centrations.  Independently, several groups have shown by administering 13C-DCM
or 14C-DCM to rats that labeled  CO subsequently appears in COHb with essentially
the same  specific activity (Carlsson and  Hultengren,  1975;  Miller et  al.,
1973;  Kubic  et al. ,  1974;  Zorn, 1975).  Furthermore, Fodor  and coworkers
(1973,  1976) and  Kurppa et al.  (1981) have demonstrated  that rats exposed to
CO  and  DCM, singly  and in combination,  effected an additive increase in
blood COHb  levels  (Table  4-8).   In addition, many investigators have shown a
dose-response relationship between injected or inhaled DCM and increased blood
COHb levels in both experimental animals and men (Figures 4-5 and 4-6) (DiVincenzo
et al., 1972;  Astrand et al. , 1975; Fodor et al.,  1973;  Ratney et al. , 1974;
Stewart et al., 1973, 1976a,b; Forster et al., 1974; Roth et al., 1975; Ciuchta
et al., 1979; Hake et al., 1974).
     The extensive metabolism of  halogenated  hydrocarbons  to CO  is apparently
unique  to the dihalomethanes.   This metabolism is not observed to any significant
extent  with  chloroform, carbon  tetrachloride, methyl  chloride, methyl  iodide,
trichlorofluoromethane, dichlorodifluoromethane, carbon disulfide, formaldehyde,
005DC5/B                           4-20                          11-14-83

-------
formic  acid,  or  methanol  (Miller et al., 1973; Kubic et al., 1974; Fodor and
Roscovanu,  1976;  Rodkey and  Collison,  1977).  Fodor  and  Roscovanu (1976)
state, without giving data, that chloroform, bromoform, and iodoform are meta-
bolized to CO in the rat, thus increasing blood COHb levels; this has not been
the observation of other investigators.   According to Fodor and Roscovanu (1976),
of the dihalomethanes, the bromo-iodo-halides are more extensively metabolized
to CO than is DCM (Figure 4-7), thus increasing COHb to a significantly greater
extent than DCM (Miller et al., 1973; Kubic et al., 1974; Fodor and Roscovanu,
1976; Rodkey and Collison, 1977).
4.2.2  Evidence for Dose-Dependent Metabolism of DCM:   Michaelis-Menten Kinetics
4.2.2.1  Studies in Humans—McKenna et al.  (1980) have reported in abstract form
kinetic studies  carried out  in six  healthy  male volunteers  exposed  to  200  and
350 ppm DCM  (694 and 1214 mg/m3) during each of two 6-hour exposure periods.
These investigators report that the metabolism of DCM was dose-dependent, based
on comparison of the  kinetics of the two doses during  and following  exposure.
Comparison was made  of  the DCM blood levels and of the concentrations of DCM
in expired air.  Blood  COHb levels and  exhaled  CO concentrations were less
than those expected  for the  350  ppm (1214 mg/m3) exposure.   Calculations made
of the rate of DCM metabolism during exposures were consistent with Michaelis-
Menten kinetics for DCM metabolism.
     More recently,  DiVincenzo and  Kaplan (1981a)  evaluated the conversion  of
DCM to COHb and CO in sedentary,  nonsmoking individuals (11 males and 3 females)
exposed to DCM levels of 50, 100, 150,  or 200 ppm (173, 347, 520,  or 694 mg/m3),
for 7.5h, or  for 7.5  h daily at  100, 150, or  200 ppm (347,  520, or  694 mg/m3)
for 5 consecutive  days.   Between 25 and 34  percent  of the absorbed DCM was
excreted in expired air as CO, during and after exposure.  The authors estimated
that as much as 70 percent of the inhaled vapor was absorbed; less than 5% was
expired as unchanged  DCM.  Materials balance  indicates that as much  as 70%  of
the amount  absorbed  may be converted to C02.   Blood  COHb  levels  increased
directly with the magnitude of exposure  and did not reach a plateau  after  7.5
hours of exposure.  At 200 ppm (684 mg/m3), the peak blood  COHb level  was  6.8
percent.  At  the recommended  TLV  for DCM  for 7.5 h,  COHb  levels were about
3 percent less than the increase in blood COHb levels  produced by  an exposure
to CO at  its  recommended  TLV  of 35 ppm.  DiVincenzo and Kaplan (1981b) also
observed that physical exercise resulted in higher blood levels than those found
with sedentary individuals.   Physical exercise (Table 4-2) increased absorption
of DCM, the  biotransformation  of DCM to CO, blood COHb levels, and pulmonary

005DC5/B                           4-21                          11-14-83

-------
         TABLE  4-8.   BLOOD CARBOXYHEMOGLOBIN  CONCENTRATIONS  OF  RATS
                     EXPOSED  TO  CO  AND  DCM BY INHALATION
Exposure concentration j>pm
DCM CO
100 (0.5 - 2)a
1000 (0.5 - 2)a
0 100
100 100
1000 100
None None
100
1000
1000 100
COHb
saturation
%
6.2
12.5
10.9
16.4
19.0
0.7 ± 0.2b
8.8 ± 1.9b
6.2 ± 0.9b
14.6 ± 1.3b
aAmbient air CO concentration.
 Mean ± standard deviation, N= 5/group.
Source:  Fodor et al., 1973; Kruppa et al.,  1981.
excretion of CO.  Individuals were exposed to 100 ppm DCM (347 mg/m3)  for 7.5
hours.   The authors concluded that workers performing physical exercise while
exposed to DCM at the recommended TLV  of 100 ppm (347 mg/m3) are unlikely  to
                              /a
exceed the COHb biological TLV  suggested by the National Institute of Occupational
Safety and Health (NIOSH).
4.2.2.2  Studies in Animals--DiVincenzo and Hamilton (1975) provided the first
information on the extent to which DCM is metabolized in rats.  These  investi-
gators injected rats intraperitoneally with 14C-DCM in corn oil  and determined
fate and disposition of radioactivity in exhaled air, urine, feces, and carcass
1, 8, and 24 hours after single doses ranging from 412 to 930 mg/kg.   Volatile
005DC5/B                           4-22                          11-14-83

-------
                               100     200     300     400

                                     INHALED AIR, ppm.
                         500
Figure 4-5.    Carboxyhemoglobin concentrations in male nonsmokers exposed  to
              increasing concentrations of DCM for 1, 3, or 5 h/day for  5  days.
              Pre-exposure values averaged 0.8%, but with 3- and 7.5-hour  expo-
              sures the pre-exposure values were above this baseline value on
              the mornings following exposure.

              Source:  Stewart and associates (1972, 1973, 1974).
005DC5/B
4-23
11-14-83

-------
                          200
  400     600

INHALED AIR, ppm
1000
Figure 4-6.    Carboxyhemoglobin concentrations in rats after inhalation exposure
              to increasing concentrations of DCM for single exposures of 3 hours.
              The values are corrected for pre-exposure COHb concentration and
              calculated from the data of Fodor et al., 1973.
005DC5/B
   4-24
          11-14-83

-------
65
60
55
50
-I
en
n
. 40
tu
£ 35
0 30
0
0 25
O
O
5! 20
15
10
5

. >
-
-
.

"

.

-
.


•
-
•
-
I
















-.

< '

























• •

,.












~
-
-
-



_

-
_


~
-
-
-
FA i A i A i
vH«wiA CH^Br* CH«l*k
Figure 4-7.  Blood CO content of rats after 3-hour inhalation exposure with
             1000 ppm dichloromethane, dibromomethane, and diiodomethane,
             respectively.

             Source:   Fodor and Roscovanu, 1976.
005DC5/B
4-25
11-14-83

-------
compounds in exhaled air were collected,  identified,  and quantified by gas  chro-
matography (GC) and radiotracer assay.   Recovery of radioactivity was  essentially
100 percent 24  hours  after administration.   About 98 percent of total radio-
activity was eliminated in exhaled air, and less than 2 percent was eliminated
in urine or feces (Table 4-9).  Some 90 percent of injected DCM was eliminated  un-
metabolized in exhaled air. Most of this  elimination (85 percent) occurred  within
2 hours.   Only 2 percent of the dose was  metabolized to CO,  3 percent  to C02,  and
1.5 percent to  an  unidentified volatile  compound (Table 4-9).   These  results
indicate that  less than 7 percent  of  the  dose is metabolized  in  the rat.
     Rodkey and  Collison (1977) considered that the small proportion  of meta-
bolized DCM found  by  DiVincenzo and Hamilton  (1975)  could  be  caused  by the
high dose of DCM used; i.e.,  metabolic transformation of DCM may be limited by
DCM excretion being more  rapid than metabolism.   Rodkey and Collison  carried
out balance  studies with  small  doses  of 14C-DCM administered  to  rats (17
mg/kg) by either  inhalation  or intraperitoneal injection.  The  animals were
placed in a closed rebreathing system that  trapped C02 and CO after conversion
                                                      ®
to C02 by passing  through  a catalyst bed of Hopcalite .   About 76 percent  of
14C-radioactivity was  recovered as 14CO (46.9 percent) and 14C02 (28.9 percent).
The remaining 24 percent could have been  exhaled as unchanged 14C-DCM, because
no radioactivity was  recovered  in carcass tissues.  Their results  also  showed
that DCM is directly metabolized to CO without isotopic dilution. The  extent of
the conversion was surprisingly great at  this low dose, and it is independent  of
the mode of administration.  For each mole  of DCM metabolized,  about 0.5 mole  of
CO and 0.3 mole of C02 were produced.
     In a second experiment,  Rodkey and Collison (1977) investigated the rela-
tionship between dose and  extent of  metabolism.  DCM  was  administered to rats
by intraperitoneal  injection  or by vaporization of the dose in  their closed
rebreathing system with a CO trap.   DCM doses from 6.8 to 69 mg/kg were given.
CO production and  DCM disappearance were calculated  from  the  change  in gas
phase  composition  as  determined by GC.  A control period was used to  measure
the endogenous rate of CO production.  When DCM was added to the system, there
was an immediate increase in the initial  rate of CO production (about 35 times
endogenous for  all  doses)  and  a rapid  disappearance of DCM  from the gas phase
(90 percent in 30 minutes).  CO continued to be produced for more than 2 hours
after  nearly complete disappearance of gaseous DCM.  Figure 4-8  shows the rates
of CO  production for various doses of DCM.   The initial rates (about  25 umole/
005DC5/B                           4-26                          11-14-83

-------
             TABLE 4-9.  FATE AND DISPOSITION OF 14ODCM IN RATS
                 (412-430 mg/kg) INJECTED INTRAPERITONEALLY

                                            Percent of dose (averages)
                                      2 hr             8 hr           24 hr
 Breath
  Unchanged  14C-DCM                    84.5             94.0           91.5
  14CO                                  0.14             1.43           2.15
  14C02                                 0.55             1.53           3.04
  Unidentified  14C                      0.40             0.80           1.49
  Total                                85.59            97.76          98.18

 Urine
  Unidentified  14C                     <0.01            <0.01           1.06

 Feces
  Unidentified  14C                     <0.01            <0.01           0.07
 Carcass
  Unidentified  14C                      3.09             2.06           1.53
  (mainly in liver)

 Source:  Divincenzo and Hamilton (1975).

 hr/kg body weight) are similar for all doses, but for lower doses they progres-
 sively decrease  after  1  to 2 hours to the endogenous rate as the DCM dose is
 metabolized.  For a very high dose of DCM (69 mg/kg), CO was produced at a nearly
 constant rate over a 6-hour period (COHb, 44 percent).   These observations sug-
 gest a saturation of the metabolizing enzymes even at the lowest dose (6.8 mg/kg),
 giving initially zero-order kinetics followed by first-order kinetics as the DCM
 concentration in the inhaled air is decreased below enzyme saturation.   The total
 amount of CO produced was related to the  dose of DCM.  For lower doses,  the moles
 of CO produced per mole of DCM were similar and averaged 0.48;  at a high DCM dose
 (69 mg/kg),  the ratio was 0.62 after 10.5 hours of exposure, suggesting  that sub-
 strate-induced enzyme formation may occur with long exposure to high doses of DCM.
Similar results were obtained with germ-free rats obviating intestinal  bacteria
005DC5/B                           4-27                          11-14-83

-------
              200
              150
           8
           §
           1
           o
           o
           I
100
              50
                                                         793
                               I
                                      I
I
                               246

                           TIME AFTER CH2d2 ADDITION, hours
Figure 4-8.  Rates of production of CO  from  DCM  given  to  rats.   Each curve
             represents changes above endogenous CO  rate  after the dose
             (in umoles/kg body weight) was  given by inhalation.

             Source:  Rodkey and Collison, 1977.
005DC5/B
                      4-28
        11-14-83

-------
as a source of CO.  In normal rats, inhibited methane production by intestinal
bacteria.  The  same results  also were  obtained with  dibromomethane, dichloro-
methane, bromochloromethane, and diiodomethane in respective order of magnitude
and rate of CO production.
     Rodkey and Collison's (1977) important finding that the metabolism of DCM
to CO in rats is rate-limited by enzyme saturation to about 25 pmole/hr/kg body
weight (b.w.) explains the seemingly low conversion observed by DiVincenzo and
Hamilton (1975) in this same species.   Recalculation of their data (Table 4-5)
gives a  DCM metabolism to CO of about  19 pmole/hr/kg b.w.   Comparable  results
for mice were also obtained  by Yesair  et al.  (1977).  These investigators ad-
ministered 1.0 and 100 mg/kg 14C-DCM in corn oil  by intraperitoneal injection.
Exhaled 14CO, 14C02,  and  unmetabolized 14C-DCM were trapped (14CO after oxida-
                   ®
tion with  Hopcalite   to  14C02  in aqueous potassium hydroxide, and 14C-DCM on
coconut charcoal) and quantified by GC and radiotracer assay.  The 1 mg/kg dose
(11.76 pmole/kg) was quantitatively metabolized to CO (0.45 mole/mole DCM) and
C02 (>0.50 mole/mole  DCM).   In the exhaled air  collected  for 12 hours, the
larger  dose  (11.76  pmole/kg) yielded 470 (jmole/kg  of  unmetabolized 14C-DCM
(40 percent dose), 0.20 mole 14CO,  and 0.25 mole  14C02.  Hence,  in mice  under
these experimental conditions, a  12 umole/kg dose of DCM  (1  mg/kg) does  not
saturate the metabolizing enzymes,  whereas 1200 pmole (100 mg) DCM/kg saturates
the enzymes and is metabolized at a constant rate of about 20 pmole/hr/kg b.w.
The remainder of the dose is excreted unchanged in the exhaled air.
     More recently, McKenna et al.  (1982) exposed rats to 50, 500, and 1500 ppm
14C-DCM (174,  1737, and 5211 mg/m3) for 6 hours.   They also found that the net
uptake and metabolism of DCM to CO and C02 did not increase in direct proportion
to the incremental increase of DCM exposure concentration (Table 4-10).  Further-
more,  increasing amounts  of  unchanged DCM in exhaled air were found with in-
creasing exposure concentration (Table 4-10).   At the end of the 6-hour exposure,
the body burden of DCM was calculated from total  radioactivity recovered during
the first 48 hours following exposure, and the body burdens had not increased to
the incremental  increase of DCM exposure concentration (Table 4-11).
     McKenna et al. estimated  the  amount of  DCM metabolized  during each  in-
halation exposure by  subtracting the  unchanged DCM  recovered  in  expired  air
(Table 4-10) from the calculated body burdens.  The  results are  summarized  in
Table  4-11.
005DC5/B                           4-29                          11-14-83

-------
          TABLE 4-10.  FATE OF 14ODCM IN RATS AFTER A SINGLE 6-HOUR
                             INHALATION EXPOSURE
                                       % body burden (x ± S.D., n=3)
i a i ainc uc i
measured
Expired CH2C12
Expired C02
Expired CO
Urine
Feces
Carcass
Skin
Cage wash
50 ppm
5.42 ± 0.73
26.20 ± 1.21
26.67 ± 3.00
8.90 ± 0.39
1.94 ± 0.19
23.26 ± 1.62
6.85 ± 1.62
0.75 ± 0.33
500 ppm
30.40 ± 7.10
22.53 ± 4.57
18.09 ± 0.81
8.41 ± 0.90
1.85 ± 0.68
11.65 ± 1.87
6.72 ± 0.13
0.24 ± 0.23
1500 ppm
55.00 ± 1.92
13.61 ± 1.20
10.23 ± 1.68
7.20 ± 0.74
2.33 ± 0.05
7.24 ± 0.65
3.97 ± 0.15
0.43 ± 0.15
Source:   McKenna et al., 1982.
 Mean ± standard deviation, number of animals in each group = 3.
             TABLE 4-11.   BODY BURDENS AND METABOLIZED 14C-DCM
                IN RATS AFTER INHALATION EXPOSURE TO 14C-DCM
Exposure
concentration
50 ppm
500 ppm
1500 ppm
Total body burden,3
mgEq 14CH2CI2/kg
5.53 ± 0.18
48.41 ± 4.33
109.14 ± 3.15
Metabolized 14CH2CI^,a
mgEq 14CH2CI2/kg
5.23 ± 0.32
33.49 ± 0.33
49.08 ± 1.37
Metabolized
14C-DCM, %
94.6
69.2
45.0
 Values are mean standard deviation
Source:  McKenna et al., 1982.
Number of animals in each group = 3

     Of particular interest is  the finding that the percent of body burden meta-
bolized decreased with the increase of body burden or increase of the level of
DCM exposure.   McKenna et al.  plotted the data in Table 4-11 in accordance with
the following linear form of the Michaelis-Menten equation:
005DC5/C                           4-30                          11-14-83

-------
                           dC = _K   . dC/ , + v                       (4-2)
                           dt     m   dt/ 3   vmax

 where  dC/dt  is  the (jg/kg DCM metabolized during the fixed time of the experiment
 and  S  is  the  exposure concentration in parts per million.  Estimates of V    and
                                                                         ITlaX
 K  ,  as  determined from the y-intercept and slope, were used with a computer pro-
 gram of the Michaelis-Menten equation to derive the values V    = 65.55 ± 2.54
                              ^                             max
 mg/kg  DCM metabolized,  and K =  493 ±  57 ppm  DCM.   Deviation  from  first-order
                                              3
 kinetics  occurred  at about 250 ppm (867  mg/m )  or one-half  k .   Therefore,
 zero-order  kinetics  with  saturation of DCM metabolizing  capacity  can be  ex-
 pected  at two to three times Km> or between 1000 and 1500 ppm of inhaled DCM
 (3,470  and  5,211 mg/m3).   These values  correspond  to  a  body  burden  of  about
 100  mg/kg for the rat (Table 4-11); therefore, they are in general accord with
 the  observations  of  Rodkey and Coll i son (1977) for the rat and Yesair et al.
 (1977)  for the  mouse.
     Methods  of determining  the  kinetic behavior of DCM metabolism during or
 after  inhalation exposure are  subject to unknown  error  from the indirect
 methods of  estimating both body burden  and  amount or rates  of metabolism.
 McKenna and  Zempel  (1981)  investigated the kinetics of DCM metabolism in the
 rat  after doses of  1 or 50  mg/kg  14C-DCM.  Table  4-12 gives  the disposition
 of DCM  after single oral doses in terms of 14C-equivalent.
     The  results  of  this study clearly  indicate that the  metabolism  of  DCM  in
 rats after oral dosing is dose-dependent.  Rats given a 1 mg/kg dose metabolized
 a greater percentage of the oral  dose  (88 percent) than those given a 50 mg/kg
 dose (28 percent).  Moreover, the rates of pulmonary elimination of unmetabolized
 DCM  (Figure 4-4)  and of the metabolites C02 and CO were first-order and were
 essentially unaffected by the dose despite large differences in the amounts of
 DCM and of metabolites excreted.   Therefore, the dose-dependent fate of DCM is
 caused by the saturation of metabolic  pathways.   Comparison of the near-saturating
 inhalation body  burden  (100  mg/kg)  (Table 4-11) shows a fair correspondence.
     In summary, the metabolism  of DCM is dose-dependent and follows Michaelis-
Menten kinetics in the rat and mouse.   In these species,  saturation of metabolic
capacity with zero-order kinetics  (about 25 mg/kg/hr) occurs at about  50 to
100 mg/kg orally or 1000 ppm DCM  (3,470 mg/m3) in  inspired air for an inhalation
period of 6  hours.   At doses of  1 mg/kg orally or  50 ppm (173 mg/m3) inhalation,
90 percent is metabolized, and at enzyme saturation doses 30  to 45 percent  is

005DC5/C                           4-31                          11-14-83

-------
      TABLE 4-12.  FATE OF DCM IN RATS 48 HOURS AFTER SINGLE ORAL DOSES

                                           Percent 14C-DCM dose by dose
Parameter                                         concentration3
Measured                                 1 mg/kg50 mg/kg
Expired CH2C12                        12.33 ± 1.43             72.09 ± 0.07
Expired C02                           35.01 ± 0.85              6.33 ± 0.39
Expired CO                            30.92 ± 1.67             11.87 ± 0.07
Urine metabolites                      4.52 ± 0.05              1.96 ± 0.05
Feces                                  0.93 ± 0.02              0.25 ± 0.02
Carcass                                5.84 ± 0.24              2.40 ± 0.24
Skin                                   1.56 ± 0.05              1.15 ± 0.06
Cage wash                              0.53 ± 0.04              0.08 ± 0.01

Source:  McKenna and Zempel,  1981.
aValues are mean ± standard deviation.   Number of animals in each group = 3.
metabolized.  Dose-dependent metabolism  of  DCM occurs also in man; at 100 to
200 ppm  (347  to 694 mg/m3) inhalation concentration, 50 to 60 percent of the
body burden  is  metabolized, as judged from  retention values of DCM determined
in man (Table 4-1).  Experimental studies in man suggest that as much as 95 per-
cent of the absorbed dose may be metabolized at these exposure levels.
4.2.2.3  Effect of Dosing Vehic1e--Pritchard and Angelo  (1982) have described
a  physiologically-based  pharmacokinetic  model  (Bishoff  model)  for mice and
used it to simulate the distribution, metabolism, and elimination of DCM after
both acute and chronic dosing.   Preliminary results indicate that the kinetics
are dependent  on the route and  vehicle  used for administration.  Adminis-
tration of  DCM  in water by oral gavage  or  by  intravenous injection  yields
similar blood and tissue profiles,  but administration in 50 percent polyethylene
glycol/water shows a rapid blood elimination and a slow liver elimination, while
oral gavage with corn oil as a vehicle shows a slower rate of tissue clearance
than that found with a water vehicle.
     Withey et  al.  (1982)  have investigated the absorption of DCM in fasting
,rats following oral gavage of equivalent doses (125 mg/kg) in 4 ml of water or


005DC5/C                           4-32                          11-14-83

-------
corn oil.   The  post-absorptive peak blood concentration averaged three times
higher for a water vehicle than for corn oil (121 ug/ml vs. 44 (jg/ml),  while the
time to peak blood concentration averaged three times longer for corn oil than
for the water vehicle (16.3 minutes vs. 5.2 minutes).  Aside from these differ-
ences in vehicle-mediated absorption, absorption was apparently rapid with both
vehicles and occurred to comparable extent because the ratio of the areas under
the blood concentration curves averaged 1.25 for water:oil.  The post-absorption
kinetics of blood elimination of DCM associated with the corn oil vehicle also
was extended (tj for oil was 49 minutes vs. 32 minutes for water).   These vehicle-
               's
related differences in absorption, kinetics of elimination, and tissue clearance
are probably related to the differences of absorption and diffusion of water and
corn oil across the gastrointestinal (GI) tract mucosa.  In contrast to aqueous
absorption into the portal system and then to the liver, corn oil and other lipids
are extensively transported via the mucosal lymphatic system, which slowly drains
by way of the left lymphatic thoracic duct at the junction of the internal jugu-
lar and subclavian veins and hence into the systemic circulation via the superior
vena cava.   The absorption, first-pass metabolism, tissue distribution, and eli-
mination kinetics probably are greatly affected by the volume of lipid vehicle
as well  as  its  nature when  introduced  into  the  stomach  of  a  rat.  While  these
considerations are unlikely to affect the pharmacokinetics of DCM in man in any
practical way, they are of importance in relation to the modes of dosing employed
in  long-term  carcinogenicity tests  of  DCM and other lipophilic compounds.
4.2.3  Enzyme Pathways of DCM Metabolism
     Figures 4-9 and  4-10 summarize current knowledge of the enzyme pathways
involved in the biotransformation of DCM and other dihalomethanes.   The scheme
is based on studies in  vivo  and  in  vitro with hepatocytes  and with  microsomal
and cytosol preparations.   The preponderance of evidence from in vivo experiments
indicates that these enzyme pathways are unique to the dihalomethanes and give
rise to both CO and C02 in nearly equimolar amounts.   However, CO is an end pro-
duct of microsomal oxidation, while  C02  is  an end product  of cytosolic enzyme
systems via  the metabolism  of  formaldehyde and formic  acid (Figure 4-9).
4.2.3.1  Microsomal Oxidation—The primary reaction of the dihalomethanes appears
to be an oxidative dehalogenation first  described by  Kubic and Anders  (1975).
These workers found that  DCM and other dihalomethanes were metabolized by rat
liver microsomal fractions  to  CO with inorganic halide  release.  The  system
required both NADPH and molecular 02 for maximal activity.  These experiments
005DC5/C                           4-33                          11-14-83

-------
               MICROSOMAL
               MIXED
              FUNCTION
             OXIDASE
            NADPH
                             CYTOSOL
                              GLUTATHIONE
                                TRANSFERASE
                          BINDING TO CELLULAR
                           MACROMOLECULES
                                   HCHO
   COHb
CO
                                PULMONARY
                                ELIMINATION
Figure 4-9.   Enzyme  pathways of the hepatic biotransformation of dihalomethanes.
005DC5/C
                  4-34
11-14-83

-------
Microsomal Pathway
               MFO
NADPH
-H
-X
>
Covalent
OH
Nonenzymatic
rearrangement
t
.sH Spontaneous
f ^ ^
binding < - 	 ^»
lipid ^0 Decomposition
protein ~H, -X
Cytosolic Pathway
                    formyl haTide
                                                       CO
    X—
GS - C
             /*CU              «^  r*C  ^ /M i V
             bon   	"*  uo ^*LUpA
             transferase        S-halomethy1gluthathi one

                                  HOH    Nonenzymatic
                                      ..  hydrolysis

                                  GS	CH2OH
                                 S-hydroxymethyl gluthathione
                            NAD
 formaldehyde
dehydrogenase
                                            \
                                             H2C=0 + GSH
                                            formaldehyde
S-formyl gluthathione

  HOH
S-formyl gluthathione
  hydrolase
            + GSH
   formic acid
Figure 4-10.
005DC5/C
      Proposed reaction mechanisms for the metabolism of dihalomethanes
      to CO, formaldehyde, formic acid, and inorganic halide.
      Source:  Ahmed et al., 1980.

                            4-35
                                11-14-83

-------
were carried out in a closed vessel, substrates were added without carrier sol-
vent, and  CO was determined by the  gas chromatographic headspace method. With
dibromomethane as the substrate,  3.6 moles of bromide were produced per mole of
CO.  In  the absence of NADPH, microsomal  fractions dehalogenated the methanes
without  CO formation.  Anaerobic conditions  substantially reduced  the  rate  of
conversion, although some CO formation (20 percent maximal)  occurred.   Equimolar
substrate concentrations of dichloromethane, bromochloromethane, dibromomethane,
and diiodomethane added  to microsomes produced the least amount of CO  for di-
chloromethane, while diiodomethane yielded the greatest amount (seven times DCM).
Liver microsomes were 5 times more active than lung microsomes and 30 times more
active than kidney microsomes.  Hogan et  al.  (1976)  also found DCM to  be con-
verted to  CO by  rat  liver microsomes requiring aerobic conditions  and  a NADPH
generating system.   These workers noted a high correlation between in vitro CO
production and microsomal cytochrome P450 content.
     Further evidence of the participation  of the P45d mixed-function  oxidase
system in  the  metabolism of dihalomethanes  is the observation that dibromo-
methane  and dichloromethane  added  to microsomal  cytochrome P450 preparations
produce  type 1 binding  spectra (Kubic and  Anders,  1975;  Cox et al., 1976).
However, Cox et  al.  (1976)  found that the  affinity  for P456  is less for DCM
(K,, 10  mM) than for chloroform (1C, 3 mM)  or carbon tetrachloride  (K<-, 1.5
mM) although carbon tetrachloride and chloroform  do  not give  rise  to signifi-
cant amounts  of CO  in vivo  (Miller et al. ,  1973; Kubic  and Anders,  1975;
Rodkey and Collison,  1977).   Both Kubic  and  Anders  (1975)  and Hogan et al.
(1976)  found   that  phenobarbital  in vivo  pretreatment  induced additional
in  vitro CO production,  while  cobaltous  chloride, which depletes  microsomal
cytochrome P456»  reduced CO microsomal production.   Furthermore,  SKF  525A,
ethylmorphine  and  hexobarbital  (type 1 substrates) inhibited in vitro micro-
somal conversion of dibromomethane to CO  (Kubic and Anders, 1975).
     Kubic and Anders and coworkers (Kubic and Anders, 1978; Ahmed and Anders,
1978; Stevens  and Anders, 1978, 1979) have studied the mechanism of the reaction
of  DCM with deuterium and with 1802.  The reaction shows a prominent deuterium
isotope  effect with a comparative rate of about 12 percent of that of the hydrogen
isotope, indicating that carbon-hydrogen  bond breakage is the rate-limiting step.
Studies  with 1802 showed that the oxygen  appearing in CO is derived from molecular
oxygen  rather  than  from water.   On the  basis of  these  studies,   Kubic and
Anders and coworkers  proposed the  reaction  mechanism  shown in Figure 4-10.
 005DC5/C                              4-36                         11-14-83

-------
 P45o~roediated  hydroxylation of  dihalomethanes  yields  the intermediate, hy-
 droxydihalomethane  (XgCHOH), which rearranges spontaneously to a formyl halide
 (XCHO) with  the  loss of one halogen atom.  The  resulting  formyl halide  is known
 to  readily decompose to yield CO.
 4.2.3.2   Cytosolic  Pathways—In  addition to being metabolized to  CO,  DCM  is
 also  converted to formaldehyde,  formic acid, inorganic halide, and CO^.  Kuzelova
 and Vlasak (1966) detected  formic acid in the urine of DCM-exposed workers and
 suggested that DCM  was metabolized via formaldehyde to formic acid.  Originally,
 Heppel and  Porterfield (1948)  reported  the  conversion  of dibromomethane and
 bromochloromethane  to  stoichiometric amounts  of formaldehyde and inorganic
 halide by a  9000-gram  supernatant fraction of rat liver and by liver slices and
 homogenates.  The system did not require 0%  Dut was glutathione-dependent.  Kubic
 and Anders  (1975)  more recently confirmed these  findings and localized this
 metabolic pathway to the cytosol.  Ahmed and Anders (1976, 1978) extended these
 findings and showed that the cytosol system  is a glutathione transferase that is
 found only  in  the  liver and it  requires no  cof actor other than glutathi one
 (cysteine is not a  substitute) and is not  inducible by  phenobarbital  or by
 repeated  exposure  to DCM or dibromomethane.   Furthermore,  the reaction was
 inhibited by reagents that react with sulfhydryl groups,  such as diethyl maleate
 and parachloromercuribenzoate,  as well as known substrates for glutathione trans-
 ferases.   The  substrate order of activity is diiodo  > dibromo = bromochloro >
 dichloromethane; this  order is  the same as  that found  for oxidative  dehalo-
 genation by Kubic and Anders (1975).   This pathway probably does not contribute
 to  CO production via a metabolism of formaldehyde to CO  because formaldehyde
 administration does not produce an increase of COHb in animals or humans (Kubic
 et  al.,  1974; Rodkey and Collison,  1977;  Kasselbart and Angerer,  1974).
     The cytosolic pathway has  been investigated in detail by Ahmed and Anders
 (1976, 1978), who have proposed the reaction sequence  shown  in  Figure 4-9.
These workers observed  that both dibromomethane and bromochloromethane have
 identical kinetic constants, suggesting that the initial and rate-limiting step
 is  a  displacement of halide,  with glutathione (GSH) interaction,  to give the
conjugate S-halomethylglutathione.   This  conjugate is postulated to undergo rapid
nonenzymatic hydrolysis to  produce S-hydroxymethylglutathione, and thus yield
formaldehyde and regenerate glutathione.   Because the addition of nicotinamide
005DC5/C                           4-37                          11-14-83

-------
adenine dinucleotide (NAD ) to incubation mixtures decreased yields of formalde-
hyde (Ahmed and Anders, 1976), the postulated intermediate,  S-halomethylgluta-
thione, a known substrate of hepatic cytosolic enzyme formaldehyde dehydrogenase,
also undergoes conversion of S-formylglutathione,  and with enzymatic hydrolysis
(S-formyl glutathione hydrolase) yields formic acid and regenerated GSH.   Ahmed
and Anders (1978) found that removal of formaldehyde dehydrogenase from cytosolic
fraction by precipitation with ammonium hydroxide  fractionation resulted in pro-
duction of only formaldehyde.   Thus, the proposed  reaction sequence for the meta-
bolism of dihalomethanes accounts for the experimentally observed stoichiometric
ratio of  two  inorganic  halides  formed  to one  formaldehyde and the  requirement
(but nonconsumption) of GSH.  Furthermore,  the sequence serves as a detoxifica-
tion mechanism for dihalomethanes that is dependent on cellular availability of
glutathione, and  the cellular availability of NAD determines the  ratio of the
products (formaldehyde/formic acid) formed.
4.2.3.3  Carbon Dioxide Formation—The cytosolic glutathione transferase system
for dehalogenation of dihalomethanes with the production of formaldehyde and for-
mic acid appears to be the major source of C02 as  an end product of the metabolism
of DCM in the intact rodent (Rodkey and Collison,  1977; DiVincenzo and Hamilton,
1975; Yesair et al., 1977; McKenna et al.,  1982).   Neely (1964) has demonstrated
that formaldehyde  is  almost quantitatively metabolized to  C02.   When Neely
injected  14C-formaldehyde  (14CHO)  intraperitoneally into  rats at dose levels
of 0.25  and  2.5  gmole/kg b.w.,  he recovered 82 percent of the dose as C02 in
exhaled air  collected over  a  24-hour period.  Peak  concentrations  of  14C02  in
exhaled air  occurred  1  hour after administration.  Neely suggested that the
formation of C02 occurred from formaldehyde entering the 1-C metabolic pool with
transfer to glycine by the folic acid cycle to give serine.   Transamination of
serine to pyruvate provides entry to the tricarboxylic acid cycle and completes
the oxidation to  C02.   In support of  this metabolic  route,  small  amounts of
14Olabeled serine and methionine were found in the urine.   However, when 14C-DCM
was administered to rats, no 14CHO was detected in the breath, serum,  or tissues
2 hours later, although substantial changes in tissue formaldehyde content were
noted (Rodkey and Collison, 1977; DiVincenzo and Hamilton, 1975).
     An additional  source  of  C02 production  from the  metabolism of dihalo-
methanes is the in vivo oxidation of CO to C02.   Carbon monoxide is metabolized
to C02  by various animal tissues (Fenn, 1970).   Luomanmaki and Coburn (1969)
also have demonstrated that 14C02 exists in the expired air of humans breathing
005DC5/C                           4-38                          11-14-83

-------
14CO during a 4-hour period in a closed rebreathing system.   Carbon monoxide is
metabolized by combining with the reduced form of tissue cytochrome oxidase in
the presence of a low 02 tension, being released as C02 (Coburn, 1970).   However,
the amount of C02 generated by oxidation of CO is very small, because the rate of
conversion is less than 5 percent endogenous CO from heme catabolism (Tzagoloff
and Wharton, 1965).   On the other hand, the metabolic rate of conversion of CO
to C02  appears  to increase as a function of  body  stores of  CO  and thus  as  a
function of blood COHb  concentration.   In dogs, at 10  to 15  percent COHb, the
metabolic conversion rate equals the formation of endogenous CO (Luomanmaki  and
Coburn, 1969).   Increased  CO  production and elevated  COHb that results  from
catabolism of DCM may stimulate  the metabolic production of  C02 from CO, thus
contributing to the total C02 produced by DCM metabolism.
4.2.3.4  Pathway Utilization Ratio—The microsome oxidative dehalogenation and
cytosol glutathione transferase dehalogenation systems (Figure 4-9) account for
the CO and C02 generated from the metabolism of the dihalomethanes.   Because the
microsomal system is apparently saturated and rate-limiting at low doses (Section
4.2.2), the relative molar amounts  of CO and C02 produced should provide an index
of the activity of the two pathways.   However, Yesair et al.  (1977) found nearly
equal molar amounts of CO and C02 with both low and high saturating doses of DCM
in mice.  For low doses in rats, Rodkey and Collison (1977) found 1.6 times as
much CO produced  as  C02, suggesting greater metabolism (at low doses) by the
microsomal oxidative pathway.   McKenna and Zempel  (1981) observed a CO:C02 ratio
of 0.9 for a low oral dose of DCM (1 mg/kg) to rats and a ratio of 1.9 for high
saturating dose (50 mg/kg), indicating a greater utilization of the microsomal
oxidative pathway at "saturating doses" (Table 4-12).   However, McKenna et al.
(1982) found for a low inhalation dose of DCM (50 ppm or 174 mg/m3) and a meta-
bolic saturating inhalation dose (1500 ppm or 5211 mg/m3),  no significant pre-
ference for either pathway (Table  4-10).   Clearly, the important factors of
hepatic content of glutathione  and of P45o plaY major  roles  in the relative
utilization of the two pathways of  metabolism.
     Though it  has been observed that equimolar doses  of dibromomethane and
diiodomethane produce COHb  concentrations  greater  than those produced by DCM
in rats (Fodor and Roscovanu,  1976; Roth et al., 1975; Rodkey and Collison,  1977),
in microsomal  preparations (Kubic and Anders,  1975) and in cytosol preparations
(Ahmed and Anders, 1976) no information is available on the ratio of CO to C02
produced by these compounds.  The use of isolated hepatocytes may prove useful
and avoid many  of the difficulties inherent with  whole animal  experiments.

005DC5/C                           4-39                          11-14-83

-------
4.3  DCM-INDUCED CHANGES IN HEPATIC ENZYMES
     The metabolism of  the dihalomethanes by the microsomal oxidative dehalo-
genation pathway  (but apparently not the cytosol pathway) can be modulated by
inducers of microsomal  mixed-function oxidase system.   Pretreatment of animals
with phenobarbital was  found by some workers to increase blood COHb levels and
microsomal  production of CO (Kubic et al.,  1974; Kubic and Anders,  1975;  Hogan
et al.,  1976;  Stevens  et al., 1980), but other investigators  found no effect
or observed a  decrease  (Miller et al., 1973; Roth et al., 1975).  Roth et al.
(1975) suggested  that enhanced metabolism  may initially be induced by pheno-
barbital, but  the  resulting  increase in local  microsomal  levels of CO may be
sufficient to inhibit cytochromic P450 oxidative dechlorination.   Of particular
interest to  human exposure  in the industrial  setting  is the  finding that
chronic  daily  exposures  of rats to DCM  substantially  increased metabolism and
COHb blood concentrations, suggesting  that  this dihalomethane can induce its
own metabolism  (Kubic et al., 1974;  Rodkey and Collison, 1977).  Heppel and
Porterfield (1948) also  reported that  repeated administration of bromochloro-
methane  to rats led to  an increased  rate of dehalogenation.  However, Haun et
al. (1972) found  that  continuous  exposure  of mice to 100 ppm  DCM (347 mg/m3)
for 4 to 12 weeks  decreased the hepatic content of cytochrome  P45o-
     Daily exposures for  shorter  periods  do not appear to influence  hepatic
P450 content,  but  they  do modify the activity of other enzyme  systems.  Kurppa
and Vainio (1981)  exposed  rats at 500 and  1000 ppm DCM (1735  and 3470 mg/m3)
6  hours  per day for  5  and 10  days.  They reported no change in hepatic P450
content but they observed a marked decrease of NADPH-cytochrome C (35 percent)
and a two-fold increase in UDP-glucuronsyl  transferase.   The  hepatic GSH content
remained unchanged. Toftgard et al. (1982)  exposed rats to 500, 1500, and 3000
ppm DCM (1735, 5211,  and 10,422 mg/m3) 6 hours/day for 3 days  and found no change
in hepatic P4s0 content, but a dose-related increase of microsomal  metabolism of
biphenyl and of benzopyrene  was noted.  These changes were postulated to be
caused by a change in  the proportions  of different cytochrome P4s0  isozymes
resulting from an  DCM-inducing effect on some specific forms  of cytochrome P4s0-
In contrast to these observations,  Pritchard et al.  (1982) recently administered
DCM to male mice for 3  days by gavage (5,  50, 100, 250, 500,  1000 mg/kg doses in
corn oil) and found no  significant changes  in hepatic weight,  microsomal protein,
P450> cytochrome  bs  contents,  or activities of aminopyrene N-demethylase and
biphenyl 4-hydroxylase.   Furthermore,   28 days  of  administration of  DCM  in
005DC5/C                           4-40                          11-14-83

-------
drinking water,  to  provide  daily  doses  ranging  from  5  to  1000  mg/kg,  also  did
not affect these parameters or hepatic glutathione content.
4.4  COVALENT BINDING TO CELLULAR MACROMOLECULES
     The likelihood  of  significant covalent binding of reactive  metabolites
from the metabolism of  OCM to cellular  macromolecules is predicated on the
postulated reactive  intermediates:   formyl  chloride  from  microsomal  oxidative
metabolism and  S-chloromethyl glutathione  and  formaldehyde from  cytosolic
metabolism (Figure 4-10).  These compounds may be capable of acylating cellular
nucleophiles.   S-chloromethyl glutathione  is  structurally similar  to the
reactive bis-halomethyl  ethers.
     Anders  et  al.  (1977)  have studied  the  extent and pattern of binding  to
microsomal  lipid and protein  after aerobic incubation of rat hepatic microsomes
with 14C-DCM.   Table 4-13  shows  that  metabolites  of DCM become  covalently
bound to both microsomal protein  and lipid  under conditions optimal  for meta-
bolism of DCM to CO.  Furthermore, microsomes from rats pretreated with pheno-
barbital showed increased binding.  Thus, the formyl  chloride  intermediate may
either acylate macromolecules or decompose to CO.
     Cunningham  et  al.  (1981) have  investigated covalent  binding of inter-
mediates from the metabolism  of 14C-DCM by rat hepatocytes.  Rat hepatocytes in
suspension have  been  shown  to metabolize DCM and other dihalomethanes  to  CO
(Stevens et  al.,  1980).   The results of  irreversible  binding  of 14C-DCM to
cellular macromolecules  in  this  system, in comparison to 14C-carbon trichlo-
roethylene, are given in Table 4-14.
     The covalent binding of carbon  tetrachloride to  lipids and  protein was
greatly enhanced in the  absence of 0- and was almost eliminated in the presence

                  TABLE  4-13.  I_N VITRO COVALENT BINDING OF 14C-DCM
                       TO MICROSOMAL PROTEIN AND LIPID
Conditions
     nmoles 14C bound/mg/min ± SD
     Protein                Lipid
Normal rat microsomes
Microsomes from pheno-
  barbital-treated rats
    0.24 ± 0.02
    0.57 ± 0.03
0.27 ± 0.02
1.97 ± 0.19
Source:   Anders et al., 1977.
S.D.  = standard deviation.
005DC5/C
4-41
11-14-83

-------
presence of 02 (Table 4-14).  In contrast, binding of DCM and trichloroethylene
was enhanced in the presence of oxygen; this is consistent with the microsomal
oxidative metabolism of these compounds to an epoxide and formyl chloride, re-
spectively. Glutathione-depleted hepatocytes showed markedly decreased covalent
binding of 14ODCM to both lipids and protein, suggesting inhibition of cytosolic
DCM GSH conjugation with a decreased production of reactive S-chloromethyl glu-
tathione and of formaldehyde.  However, at least part of the decrease in binding
may be caused by diethylmaleate inhibition of microsomal oxidation (Stevens et
al., 1980), as well as by diethyl maleate depletion of cellular GSH.   Phenobar-
bital pretreatment also markedly decreased binding from 14C-DCM hepatocyte meta-
bolism.   The explanation of this effect is not readily apparent because pheno-
barbital is not  known  to deplete cellular glutathione  but does increase  DCM
microsomal  oxidative metabolism  to  CO in hepatocytes (Stevens et al., 1980).
Perhaps the most  important observation made by  Cunningham  et al.  was that
14C-DCM metabolism by isolated hepatocytes did not result in the alkylation of
the nucleic acids  RNA or DNA, whereas metabolites of carbon tetrachloride  and
trichloroethylene labeled nucleic acids under the conditions.
     Reynolds  and Yee (1967) studied labeling patterns of 14C-DCM and 14C-formal-
dehyde in  rat  liver  after  jm vivo injection.   They found similar patterns of
labeling of DCM and its metabolite,  formaldehyde.  Binding occurred most often
at the amino  acid  locus corresponding to serine and on the acid-soluble cell
constituents,  with smaller amounts  in lipid and  nucleic acids.  However,  for-
maldehyde  or   formic  acid  can  directly  combine with  tetrahydrofolic
acid and consequently be  incorporated into de novo  nucleic  acid  synthesis;
therefore,  the association of 14C-activity with nucleic acids does not necessarily
indicate covalent binding.
4.5  KINETICS  OF CARBOXYHEMOGLOBIN FORMATION
     Blood COHb accumulates when the amount of endogenous or exogenously derived
CO in the  body exceeds  that of pulmonary elimination.   Pulmonary elimination
of CO is a first-order process that involves the exchange of hemoglobin-bound CO
with oxygen and the diffusion of CO across capillary and alveolar  endothelium
into alveolar  lung space.   In humans,  the half-time of elimination of CO is 4 to
5 hours (NIOSH, 1972;  Lambertsen, 1974),  and in the rat the time is 1.8 to 2.5
hours (McKenna et al.,  1982; McKenna and Zempel,  1981).   The half-time of pul-
monary elimination is independent of the blood COHb concentration but is depen-
dent on factors such as lung  function and dysfunction,  pulmonary rate and  am-
plitude, regional  blood flow, and cardiac function.

005DC5/C                           4-42                          11-14-83

-------
   TABLE 4-14.   COMPARATIVE COVALENT BINDING OF DCM,  CARBON TETRACHLORIDE,
         AND TRICHLOROETHYLENE TO LIPID AND PROTEIN IN RAT HEPATOCYTES
Atmosphere3
(0^/N^)
Substrate
ecu.
TCE
DCM
Protein
0.09
9.63
8.53
Lipid
0.06
5.08
11.96
GSH Depletion*3, %
Protein
100
74
44
Lipid
100
74
38
Phenobarbital
induction, %
Protein
145
288
44
Lipid
120
238
40
aRatio of covalent binding observed under oxygen and nitrogen atmosphere.
 Values are the mean of three to six determinations for each condition.
 Thirty minutes prior to hepatocyte isolation, rats were treated intraperi-
 toneally with 0.6 ml/kg of diethylmaleate.   Values are expressed as the mean
 percent of binding observed in hepatocytes  from untreated rats, with N=5.
GRats were pretreated with three consecutive daily doses of phenobarbital
 (80 mg/kg, intraperitoneally) beginning 4 days before isolation of hepatocytes.
 Values are expressed as the mean percent of binding observed in hepatocytes from
 untreated rats, with N=5.
Source:  Cunningham et al., 1981.
TCE = trichloroethylene
     Ordinarily,  the  sole  endogenous source of CO,  and hence COHb, is the
physiologic catabolism of heme by the hepatic microsomal heme oxygenase pathway
(Coburn, 1973;  Tenhunen  et al., 1969).   The endogenous rate of CO production
from this  source  in the  normal  human is about  20 umoles/hr, producing a blood
COHb concentration of  approximately 0.4 percent.   A close linear correlation
exists between the molar rate of CO production and the percent COHb saturation;
only 10 to 15  percent  of the total  body CO  is  not  associated with  hemoglobin.
Most of this 10 to 15 percent is bound to hemoproteins such as myoglobin and heme
cytochromes;  about 1 percent is dissolved in body water.  The rate of endogenous
CO  production  from heme catabolism is  markedly inhibited  by exogenous CO
sources producing COHb levels of approximately 12 percent saturation, suggesting
that blood levels  of  12 percent saturation  are sufficient  to  inhibit,  by  CO
binding, hepatic microsomal oxidase systems  involved in hemoglobin degradation
to  CO  (Coburn,  1970).  The  half-life of COHb  (4 to  5  hours) arising  from heme
catabolism or exogenous environmental CO (e.g., ambient-air CO, tobacco smoking)
005DC5/C                           4-43                          11-14-83

-------
is decreased by  increased  alveolar  ventilation or increased inspired partial
pressure of oxygen (Lambertsen,  1974).
     Since Stewart and his  associates (1972a,b) reported the remarkable  increase
of blood COHb (up to 15 percent  from 0.6 percent pre-exposure)  in persons  acutely
exposed by  inhalation  to DCM vapor,  numerous investigations of the phenomenon
have been made in experimental animals  and humans.   Studies have  been undertaken
to determine the  dose-response  relationship  of blood COHb level  with DCM  air
concentration,  with duration of exposure, with time-course of COHb blood con-
centration rise and decline, and  with  the magnitude of its occurrence in  the
industrial setting.  Because CO generated from DCM is additive to exogenous
environmental CO,  DCM exposures  at high levels  could pose an additional  health
burden.  Of  particular  concern  are  smokers who maintain significant constant
levels of COHb, i.e.,  4.6 to 5.2 percent (Stewart et al., 1974a,b;  Kahn  et al.,
1974), and  others  who  may  have  increased sensitivity to CO toxicity, such as
pregnant women and persons  with  cardiovascular  disease or pulmonary dysfunction.
Indeed, Stewart et al.  (1972a,b) have noted that exposure to concentrations of
                                         (R)
DCM that do not exceed the  industrial TLV^ (100 ppm, 347 mg/m3)  may yield  COHb
levels exceeding those allowable from exposure  to CO (35 ppm,  38.5 mg/m3;  about
5 percent COHb blood concentration).
     As previously  discussed, the recent findings  of  DiVincenzo  and Kaplan
(1981a,b) with sedentary individuals and those engaged  in physical  exercise
while being  exposed to DCM indicate that blood  levels of  COHb are increased
but are within the  biological' TLV®  recommended  by  NIOSH (1976).   NIOSH has
recommended that  the TWA exposure to DCM should not produce COHb levels that
exceed 5 percent.   DiVincenzo and  Kaplan (1981a) found that an 8-hour exposure
to 150 ppm DCM (520 mg/m3)  in sedentary individuals is equivalent to an  8-hour
exposure to 35 ppm CO;  either exposure  has been shown to result in blood levels
of COHb of about 5 percent.  In  their companion study in which individuals either
exercised or smoked cigarettes during  exposure,  DiVincenzo and Kaplan (1981b)
observed an  additive increase in  blood COHb concentrations.  Individuals were
exposed for  7.5 hours  to 100 ppm  DCM (347 mg/m3).   The effects of  exercise  on
physiological parameters are presented in Table  4-2.   Moderate to  heavy work-
loads produced about a  two-fold increase  in blood  COHb  levels compared to  the
increase for sedentary individuals.
     From review of the literature,  an important observation that can be made of
DCM-induced COHb is that the blood concentration, although DCM-dose dependent and
005DC5/C                           4-44                          11-14-83

-------
exposure-time dependent,  never exceeds 10 to 12 percent in either humans or ani-
mals in  usual  circumstances  of free breathing and normal pulmonary function
(Figures 4-5, 4-6, 4-11).   This limiting blood concentration is clearly deter-
mined by the resultant of first-order pulmonary elimination of CO and zero-order
kinetics of hepatic DCM metabolism to CO (Section 4.2.2).   Because COHb is essen-
tially confined to the blood compartment, a one-compartment open kinetic model
with zero-order input can be used to describe the time course of blood COHb as
follows:
                                                                      (4-3)
where V  is  the  volume of the Hb compartment (90 percent CO is distributed in
this compartment)  (Luomanmaki and Coburn, 1969), kQ is the zero-order rate of
COHb formation,  k   is the first-order rate constant for pulmonary elimination
of CO from  COHb, and  C  is the the concentration at any time t of COHb formed.
Then,
                                                                      (4-4)
This equation  (Wagner, 1975) describes the time course of rising COHb concen-
tration with zero-order  formation  of COHb.   With long periods of exposure to
DCM (6 to 8 hours), the relation becomes
C
COHbt
ko
Vke
1-e "kgt
                          COHb4
                                 = k /Vk
                                    o   e
                                   (4-5)
That is, a  steady-state  plateau concentration of blood  COHb  is  reached  (see
Figure 4-12).  Therefore, maximal blood COHb is mandated when KQ) V, and kg are
constants.
4.5.1  Studies in Human
     In a series of studies, Stewart and his associates (Stewart et a!., 1972a,b,
1973, 1974a,b;  Forster et al.,  1974; Hake  et al., 1974)  have  shown  that  blood
COHb levels  achieved  in  response to DCM exposure are  related to the  inhaled
concentration and to the duration of exposure.   Male nonsmokers were exposed to
005DC5/C
4-45
11-14-83

-------
                                                       CH2CI2, ppm
                                                       O—0500
                               EXPOSURE, days.
Figure 4-11.
Blood COHb level in men during an 8-hour exposure  for  5  consecu-
tive days to 500 ppm and 100 ppm DCM.  COHb percent  saturation is
equal to |jg CO per ml blood divided  by 2.5.

Source:   Fodor and Roscovanu, 1976.
005DC5/C
                     4-46
11-14-83

-------
DCM for 1, 2, and 7.5 hours daily,  5 days/week.   Blood COHb levels were deter-
mined for daily  pre-exposure  and post-exposure  times.   Their data, which are
replotted in part in Figure 4-5, indicate that maximum COHb concentrations occur
with 400 to 500 ppm DCM (1390 to 1737 mg/m3) exposure and increase with duration
of exposure.  Similar results were reported in women nonsmokers exposed 1, 3, and
7.5 hours to 250 ppm DCM (869 mg/m3) for 5 consecutive days (Hake et al., 1974),
and in male volunteers exposed to 100 and 500 ppm DCM (347 and 1737 mg/m3) daily
for 5 days (Fodor and Roscovanu, 1976) (Figure 4-11).  In each of these studies,
the time course of decay of COHb levels to pre-exposure levels occurred within
24 hours so that a consistent increment in COHb  with daily exposure was not ob-
served (Figure 4-11);  however,  the  half-time of COHb disappearance was signi-
ficantly longer than expected.  The longer apparent half-time of pulmonary eli-
mination of  CO  produced  by hepatic metabolism of DCM results from storage of
DCM in adipose tissue (Section 4.1.2), with conversion to CO continuing subse-
quent to termination of exposure.  In these circumstances, the biological half-
life of COHb derived from DCM (10 to 15 hours) is proportional to the body burden
of DCM, in contrast to the constant half-life of 4 to 5 hours from CO inhalation
(Stewart et al., 1972a,b, 1976a,b;  Fodor et al., 1973; Peterson, 1978; Peterson
and Stewart, 1970b).  However, when workplace exposures were simulated, the bio-
logic half-life  for  COHb was  similar to that for CO  (DiVincenzo  and  Kaplan,
1981a).
     DiVincenzo and  Kaplan  (1981a)  found that blood  COHb  saturation  in seden-
tary male volunteers is not attained during a 7.5-hour exposure to DCM at con-
centrations ranging from 50 ppm (173 mg/m3) to 200 ppm (694 mg/m3).  Peak blood
COHb during  the  200-ppm  exposure was 6.8  percent.   At the TLV   for  DCM,  the
authors estimated that an  8-hour exposure would produce  a  blood COHb level
of about 3  percent.   A linear  relationship was found between blood  COHb  and
DCM exposure concentrations.  Repeated exposure to DCM levels of  100,  150,  or
200 ppm (347,  620,  or 694 mg/m3) for 7.5  hours daily for 5  consecutive  days
produced slightly higher  blood  COHb levels than those found for single expo-
sures.   Following  the weekend,  blood  COHb levels  returned to  pre-exposure
levels.   Blood COHb declined at a slower rate than DCM in expired breath during
post-exposure.   About  24  hours  was  required for blood COHb to return to pre-
exposure values.
     The findings in  a companion study by  DiVincenzo  and  Kaplan  (1981b),  with
smokers and  individuals  engaged in  physical activity indicated that exposure
005DC5/C                           4-47                          11-14-83

-------
100.0
     E '   I   I   I   '   1   '   I   I   I   '   I  I   I  '   I   I   I   'E
 10.0

  0.1
                   EXPOSURE'
       J  1  I   I   I   I   1
                                             A 50 ppm
                                             • 500 ppm
                                             • 1500 ppm
I   I   I   1   I   I   I   I   I   ,   I   ,   I
    0123456/01234

                              TIME, hours


     Figure 4-12. Blood HbCO concentrations in rats during and after a
     6-hour inhalation exposure to DCM. Each data point is the mean
     ± standard error for two to four rats.
     Source:  McKenna et al. (1982).
                                 4-48

-------
to 100  ppm  DCM (347 mg/m3) may result in slightly higher blood COHb than in
sedentary nonsmokers.   The investigators  concluded that workers performing phy-
sical exercise while exposed to DCM at 100 ppm (347 mg/m3) are unlikely to exceed
the COHb biological TLV® recommended by NIOSH (1976).
     Ratney  et al.  (1974)  studied  the blood COHb  levels  (calculated  from
alveolar concentrations) of  a  group of young male adults in their workplace
where large  quantities of  DCM were  used (a plastic-film plant).  Workroom air
concentrations of DCM averaged 200  ppm (695 mg/m3) with no measurable CO.  At
the beginning of the work day,  blood COHb levels averaged 4.5 percent.  After an
8-hour  exposure, COHb  levels  rose to about 9 percent, then declined exponen-
tially  to 4.5 percent by the next working day (16 hours later) with an  apparent
half-time of CO pulmonary elimination of  13 hours.  These prolonged half-times
for CO  in  alveolar  air and for COHb  in  blood are difficult to interpret in
light of the contrasting  results  of Stewart  et al. (1972a,b)  and DiVincenzo
and Kaplan  (1981a,b).  Ratney et  al.  (1974) reported  that pre-exposure  levels
of CO in expired air and COHb in blood were elevated above the values expected
for nonsmokers.  Three of  the  seven individuals that were monitored were not
previously exposed  to  DCM;  therefore, it is surprising that the CO levels in
their alveolar air and blood COHb  values  were  as high as  reported  in the
paper.   Ratney et al. (1974) reported a mean  pre-exposure CO concentration in
expired air of about 29 ppm.   Normal levels of CO in alveolar air are expected
to range between 1 and 5 ppm.
     The studies of Stewart and associates,  Fodor and Roscovanu, and Ratney et
al. are consistent  in  their findings and lead to  the  following conclusions.
     (1) COHb  derived  from CO  of DCM metabolism  is additive to COHb derived
from CO of exogenous sources.
     (2) Blood COHb concentration attained from DCM metabolism is a function of
both DCM inhalation concentration and of  exposure duration,  with a maximal attain-
able concentration of 10 to 12  percent COHb.
     (3) A steady-state blood COHb concentration in sedentary individuals is not
attained within 8 hours for single exposures of DCM.
     (4) The  half-time of  first-order pulmonary elimination of COHb  is  4 to 5
hours,  although a longer pseudo half-time of 10 to 15 hours is observed because
of continued metabolism of DCM post-exposure,  from solvent stored in fat compart-
ment during exposure.
005DC5/C                           4-49                          11-14-83

-------
     (5) Blood COHb concentrations  from multiple  daily  exposures  reach  a  steady-
state level  within 3 to 5 days,  and the level  is  only slightly  higher than  that
after a single 8-hour exposure.
     These observations on the kinetics of COHb formation  during  DCM inhalation
exposure are consistent with Michaelis-Menten  kinetics  for DCM  metabolism to CO
(Section 4.2.2).   The data plot of  Stewart and associates  of Figure 4-5 (7.5-hour
exposures) illustrates dose-dependent  formation of COHb and is of the  form of
the Michaelis-Menten equation:
                                       V    S
                                dC  =   max                           (4-6)
                                dt     K  + S
                                        m
where dC/dt  is the rate of COHb formation and S is the exposure concentration
in parts per million.   Estimates of V    and K  were  obt
   K     ^                           max      m
data in accordance with the linear form of the equation.
in parts per million.   Estimates  of V    and  K  were  obtained  from plotting the
                                     max      m
                         d£  - -K    dC     s + y                     (4.?)
                         dt      m   dt          max

This form provides values of V    s 15% COHb/7.5 hr and K  = 200 ppm (694 mg/m3)
                                                         Ml
DCM.  These  values  indicate  a' maximum obtainable COHb blood level  in humans
caused by DCM exposure in air of about 25 percent, and a saturation- of hepatic
metabolism of DCM  to  CO in humans at 400 ppm DCM (1388 mg/m3)  in inhaled air
(2 x K ).
     The COHb blood levels resulting from DCM inhalation are dependent on pul-
monary function status as well as on metabolism of DCM to CO.   Physical  activity
or exercise  during exposure to DCM markedly increases pulmonary absorption and
body retention  (Section  4.1.3),  but tends to diminish the maximum blood COHb
level achieved  by  the end of the exposure period, although higher blood COHb
levels are  attained  3 to 4 hours  after  exposure when compared with control
sedentary subjects  (Stewart  and Hake, 1976; Astrand et al., 1975; DiVincenzo
and  Kaplan,  1981b).   These  findings  may  be  explained by  the decrease  in  half-
life of  pulmonary  CO  elimination affected by  the increased  pulmonary  ventila-
tion rate and increased cardiac  output with physical activity (Lambertsen, 1974).
Conversely,  compromised  pulmonary function  (e.g., adult respiratory distress

005DC5/C                           4-50                           11-14-83

-------
syndrome, emphysema,  asthma) can be expected to increase COHb blood levels by
decreasing CO pulmonary elimination.   A comprehensive kinetic model  describing
in humans the parameters of absorption, distribution, and elimination  of DCM, as
well as metabolism to CO and COHb,  would be valuable in evaluating the effects of
physical activity, pulmonary and cardiovascular diseases, and concomitant envi-
ronment xenobiotic exposures including drugs such as alcohol and barbiturates,
of which very little is known.   Hake (1979) has developed a computer simulation
model, based on the Coburn-Forster-Kane equation,  describing the kinetics of COHb
formation using the experimental data of Stewart and his associates.   He was able
to show that physical activity during exposure to DCM led to a lower blood COHb
than the sedentary level, because while increased ventilation increased DCM uptake,
it also increased the pulmonary elimination rate constant (k ) for CO.
4.5.2  Studies in Animals
     The kinetics of COHb formation from DCM exposure elucidated in humans have
largely been confirmed in animals.    Figure 4-6 shows the relationship between
exposure concentration and COHb levels in rats reported by Fodor et al.  (1973).
For a 3-hour exposure, maximal  levels of COHb (12.5 percent) are achieved with
about 1000 ppm DCM (3,474 mg/m3).   Fodor and Roscovanu (1976) noted that a 3-hour
exposure with 200 ppm DCM  (695  mg/m3)  in humans produced COHb levels  of  about
4.3 percent, but  found  nearly  twice this level in the rat (Figure 4-6); this
difference suggests that metabolic capacity for CO formation is  greater in rats
than in humans.   Hogan et al.  (1976) found that rats exposed to  440 ppm DCM (1529
mg/m3) for 3 hours had maximal  COHb levels of about 7 percent, and exposure to
2300 ppm (7,990 mg/m3) produced no further increase.  Pretreatment of the animals
with phenobarbital  increased the rate  of rise  of COHb  levels  and the  time  the
maximum level was maintained,  but  did not increase the highest  level.  These
investigators suggest  that endogenous CO  inhibition of the P450 metaboli-
zing system  by  CO  binding to the cytochrome may occur at very high levels of
DCM exposure.  However,  their observations also are consistent with the Michaelis-
Menten kinetics of hepatic CO formation from DCM,  with saturation of metabolic
conversion from exposure  to 400 ppm (1,388 mg/m3).  Kurppa and  Vainio (1981)
exposed rats to  500  and 1000 ppm DCM  (1737  and 3474 mg/m3), 6  hours/day, 5
days/week for 1 and 2 weeks.  The test results are given in Table 4-15.   Since
maximal blood COHb levels of 8 to 9 percent occurred with 500 ppm (1737 mg/m3)
exposure and no further increase was obtained with 1000 ppm (3474 mg/m3) or with
005DC5/C                           4-51                          11-14-83

-------
longer duration of exposure,  these  results also indicate that saturation of
metabolism of DCM to  CO in rats occurs at a DCM exposure  concentration  of 500 ppm
(1737 mg/m3) or less.
     McKenna et al.  (1982) exposed rats to 50,  500,  and 1500  ppm  DCM (174,  1737,
and 5211 mg/m3) for 6 hours;  during testing steady-state  conditions  were reached
between production of CO, maintenance of a given circulation  blood COHb concen-
tration, and pulmonary CO excretion.  These results are shown in Figure 4-12.
The blood percent COHb at steady-state was not  linearly related  to exposure con-
centration but reached a maximum of about 12.5  percent.  Kinetic  parameters for
the production of CO  and formation of COHb can  be estimated on the clear assump-
tion of  dose-dependent Michaelis-Menten  kinetics (Section 3.2.2).   Estimates
calculated from the data in Figure 4-12 are V  =13.5  percent COHb/6-hour expo-
sure, and K  = 170 ppm DCM (590 mg/m3).  These values indicate a  saturating DCM
exposure concentration of  about  350 ppm (1,215  mg/m3) (2 x K ) for the  rat.
     The determination of  Michaelis-Menten kinetics of DCM metabolism  to CO
and an  upper  limit to  the attainable  blood COHb  concentration provides  an ex-
planation of  the  observations  of  Roth et al. (1975) and of Haun et  al.  (1971,
1972).    Roth  et al. found that concentrations  of COHb  in the blood  of  rabbits
after very short exposures (20 minutes) to DCM inhalation concentrations ranging
from 2000 to  12,000  ppm (6948 to 41,688 mg/m3) were a linear function of DCM
exposure concentration.   COHb  levels were approximately 5.5 percent at 2000
ppm  (6,948  mg/m3)  and  13 percent  at 12,000 ppm (41,688 mg/m3).  However, with
    TABLE 4-15.  BLOOD COHb AND Hb CONCENTRATIONS IN RATS EXPOSED TO DCMa
Exposure
(ppm)
None
500
1000
Carboxyhemogl obi n, %
Five days
0.
8.
9.
4
1
2
± 0.1
± 0.6
± 0.6
Ten
0.4
8.1
8.5
days
± 0.
± 0.
± 0.

2
7
6
Hemoglobin, g/1
Five
155.2
144.0
151.5
days
± 5.5
± 21.2
± 6.0
Ten
156.
155.
158.
days
6 ±
4 ±
0 ±
6.2
3.5
4.5
a
 The  results are given as mean ± standard deviation of five rats per group.
Source:   Kurppa and Vainio, 1981.
 005DC5/C                           4-52                           11-14-83

-------
4 hours  of  exposure at about 7000 ppm  (24,318 mg/m3), steady-state blood COHb
concentrations of 14 percent were attained.   Phenobarbital pretreatment of the
rabbits  decreased the blood level of COHb achieved with a given DCM exposure,
although in  parallel  experiments,  phenobarbital  stimulated rabbit microsomal
benzene  hydroxylase and benzphetamine N-demethylase activity.   Therefore,  Roth
et al. (1975) suggested that increased CO production from DCM at the microsome
in response  to a phenobarbital-induced  increase of the cytochrome P45o system
inhibits further DCM metabolism.  A more likely explanation of these results is
that the short exposures did not provide sufficient time for the establishment
of body  equilibrium to inhaled DCM; therefore, even at high inhalation concentra-
tions, the  DCM presented to the  hepatic-metabolizing system was on the linear
portion  of the Michaelis-Menten curve.
     In  a comparison of dogs and monkeys continuously exposed to 25 and 100 ppm
DCM (87  and 348 mg/m3) for 6 to 13 weeks, Haun et al.  (1971) found that steady-
state levels of blood COHb were maintained throughout the exposure period, and
these concentrations were directly proportional to the exposure concentration.
However, these concentrations result from low-level exposures; therefore,  they
are expected to be on the linear portion of the Michaelis-Menten function.  Haun
et al. noted that dogs had higher steady-state blood concentrations of DCM than
monkeys  at these inhalation concentrations,  but monkeys  had the higher COHb blood
levels,  suggesting that monkeys have a greater hepatic capacity for CO formation.
4.5.3  Comparison of Kinetics of Humans and Rats
     The kinetics of DCM metabolism to CO and of COHb formation are strikingly
similar  for  humans and rats.  Both  species demonstrate Michaelis-Menten dose-
dependent kinetics with similar estimates for V  (15 and 13.5 percent COHb/time
unit, respectively) and K  (200 and 170 ppm DCM, respectively).  Hence, the satur-
ating DCM inhalation concentration  for  the metabolism of  DCM  to CO and forma-
tion of  COHb are comparable for the two species (350 vs.  400 ppm;  1215 vs. 1388
mg/m3).
     The metabolism of DCM  involves two pathways:  the microsomal oxidative
pathway  leads to CO, and the cytosolic  pathway using GSH  converts DCM  to  for-
maldehyde,  formic acid,  and C02.  Comparison of the kinetic parameters estimated
above for the CO pathway in the rat with those obtained  by McKenna et al.  (1982)
for the total overal1  (both pathways) metabolism of DCM  in these rats (Section
4.2.2) reveals that the K  = 400 ppm (1388 mg/m3) for overall  metabolism is about
005DC5/C                           4-53                          11-14-83

-------
twofold greater than for the CO pathway alone.   Furthermore,  this  figure  is  an
average overall value, suggesting that the cytosolic GSH-dependent pathway has
a much higher «m value than that of the CO pathway.   Andersen (1981)  has  called
the CO metabolism a high-affinity, low-capacity pathway and the GSH-dependent
pathway low-affinity but high-capacity.
4.6  MEASURES OF EXPOSURE AND BODY BURDEN IN HUMANS
     In the  controlled  laboratory  setting,  estimating the DCM absorbed into
the body by  comparing inspired and alveolar air concentrations, or by measur-
ing blood  levels of DCM and then extrapolating these parameters to body dose,
is an imprecise task.   However, the goal  of this research  is  to develop a suffi-
cient data base and knowledge of the kinetics  of absorption,  disposition,  eli-
mination,  and metabolism of DCM to enable assessment of the body burden of DCM
from acute or  chronic exposure in the industrial or ambient setting where air
concentrations and exposure period vary widely.  Air monitoring, an important
control measure, cannot be used as a reliable  index  of body burden.   At present,
monitoring levels in  blood  or alveolar air and measuring  blood COHb levels
are the available approaches  to  estimating in human body  burdens  from recent
exposures.  However, in addition to the lack of complete pharmacokinetic know-
ledge necessary to  interpret  these determinations  into accurate and  reliable
measures of body burden, the results also are  subject to unknown interindividual
variation from factors such as anthropometric  differences, metabolism and work
load,  age  and sex, as  well  as modifications from  drugs  and environmental
xenobiotics.
     Stewart and associates (1976a) advocate the use of breath analysis to moni-
tor DCM exposure because  it is a  noninvasive method and avoids the problems
associated with the multiple blood sampling required for determinations of blood
DCM and COHb levels.   In addition, breath samples can be readily collected with
little inconvenience  in  the  immediate post-exposure period and also  at later
periods.   Analysis of DCM in those samples by infrared spectroscopy or gas liquid
chromatography provides both an identification and  a measurement of the magnitude
of exposure.
     Stewart and his  associates  have constructed a "family"  of post-exposure
breath decay curves  spanning  20 hours from controlled  and known  inhalation
exposures  of volunteers  in  the laboratory.  The concentration  of DCM  in  the
alveolar  air during and after exposure  is  related  directly  to the average
inhalation exposure  concentration.   When the duration  of  exposure  and work
005DC5/C                           4-54                          11-14-83

-------
intensity are  known,  the  total body burden can be estimated by reference to
"standard" breath  concentration  curves.   Because  of large variations between
individuals exposed under identical conditions, only very approximate DCM body
doses can be estimated.  The 2-hour period following exposure appears to be the
most reliable breath sampling time for estimating the TWA DCM exposure concen-
tration.  Stewart  et  al.  (1976a) provide little  information on  the statis-
tical reliability  or  the  interindividual  variation expected in the  predictive
values  for exposure burdens, although  they point  out the  desirability of con-
structing "individualized" breath decay curves for every person expected to be
exposed to DCM.  Peterson (1978), using the same experimental  exposure data as
Stewart et al., has developed empirical equations relating post-exposure breath
concentrations with exposure time, duration,  and blood COHb levels.
4.7  SUMMARY AND CONCLUSIONS
     The metabolism and pharmacokinetics of DCM have been studied extensively.
At ambient temperature, DCM is a volatile liquid with high lipid solubility and
modest solubility in water; therefore, the principal  routes of entry to the body
are by pulmonary and oral  absorption.   Comparatively fewer data are available on
the metabolism and pharmacokinetics of absorption and excretion of DCM in humans,
but these kinetics have been studied extensively in the rodent.   Absorption from
the GI tract is rapid and complete, occurring by first-order passive processes in
the rat; the kinetics of peroral absorption and post-absorption disposition and
elimination are influenced  by  the dosing vehicle.  Pulmonary absorption also
occurs by first-order  diffusion  processes in both humans  and rats.  There are
three distinct components  whose rate constants correspond to tissue loading of
at least three major body compartments.  At equilibrium with inspired air con-
centration,  a  blood/air partition coefficient of about 10.5 at 37°C has been
observed.
     Distribution of DCM in the tissues is consistent with its  lipophilic nature
and modest water solubility.  The chemical readily crosses the  blood-brain and
placenta!  barriers and.distributes into breast milk.   Concentrations occurring
in all major  tissue  organs are dose-related  to inspired air concentration or
to oral dosage.   During  inhalation exposure,  the quantity of DCM absorbed is
dependent also on body weight and fat content of the body; the  adipose tissue/
blood coefficient at 37°C  is about 7,  and about 0.8 to 1.0 for  brain and liver
tissues.  The  rate of  tissue loading with a  given inspired air concentration
is increased with physical  activity, and with exposure duration, with steady-
state body equilibrium requiring more than 6  hours.

005DC5/C                           4-55                          11-14-83

-------
     The kinetics of elimination of DCM from the  body are  complex and are domi-
nated by two major and parallel  occurring processes:   1) pulmonary elimination of
unchanged DCM,  and 2) hepatic metabolism of DCM.   Pulmonary elimination follows
first-order kinetics, is independent of body dose,  and exhibits  at least three
distinct body compartments with  half-times for humans of about 8 to 23 minutes,
40 to 80 minutes, and 360 to 390 minutes, and half-times for rats of 1 to 2 min-
utes, 12 to 16 minutes, and 47 minutes, respectively.   The longest half-time is
associated with the lipid and adipose tissue body compartment.   The differences
between humans and  rats  for  the half-times of pulmonary elimination  from body
compartments are presumed to be  caused by differences in pulmonary and cardio-
vascular functions.   Hepatic metabolism is body-dose  dependent and saturable, and
follows nonlinear Michaelis-Menten kinetics.   Body dose, as given by oral dosing,
also is subject to the kinetics  of first-pass hepatic metabolism and first-pass
pulmonary excretion.  The estimated K  for overall  metabolism of DCM in the rat,
as obtained from inhalation exposures, is about 400 ppm.   This inhalation concen-
tration for 6 hours produces an  estimated body dose of approximately 50 mg/kg.
For the body dose of 50 mg/kg, 70 percent of the  DCM is metabolized, and for a
5.5 mg/kg body dose resulting from 50 ppm for a 6-hour exposure, 95 percent is
metabolized.  The inhalation concentration saturating overall metabolism  of
DCM is about 800 ppm (2 x K  ).
     The hepatic metabolism of DCM occurs by two  enzymic pathways:  1) a cyto-
chrome P450-mediated microsomal  oxidative dehalogenation to CO,  and 2) a cyto-
solic glutathione transferase dehalogenation  system yielding  formaldehyde  and
formic acid, which are further metabolized to CO^.   The activities of these two
pathways are approximately equal to low body burdens of DCM, and both pathways
are saturable at high body burdens. Covalent binding to cellular macromolecules
(proteins and lipids) by putative reactive intermediates  of metabolism, formyl
chloride from microsomal oxidative metabolism and S-chloromethyl glutathione and
formaldehyde from cytosolic  metabolism,  has  been shown to  occur  in  vitro with
rat hepatic microsome preparations and with intact isolated rat  hepatocytes and
in vivo after 14C-DCM injections to rats.  No evidence was found in these studies
for the occurrence of covalent binding to DNA.
     The occurrence of increased blood COHb levels in humans and animals exposed
to DCM  is a consequence  of hepatic microsomal oxidative metabolism of DCM to CO.
The  COHb  produced is additive to  COHb formed  from  exogenous  CO.    A  functional
relationship exists between  the DCM inhalation concentration, duration of exposure,
005DC5/C                           4-56                          11-14-83

-------
and the time course and peak blood COHb  level.  The blood COHb level achieved
is determined by the nonlinear kinetics (Michaelis-Menten kinetics) of hepatic
metabolism of DCM to CO and the parallel-occurring linear kinetics (first-order
kinetics) of pulmonary elimination of CO from circulating COHb.   Because hepatic
metabolism of DCM to CO is saturable, the zero-order kinetics of CO production
constrains blood COHb accumulation to an upper, limited concentration fo 12 to
15 percent COHb, as observed experimentally.   However, human studies have demon-
strated that exposures  to DCM at levels up to about  150 ppm (620 mg/m3) are
unlikely to exceed the biological TLV© for blood COHb  (5 percent) recommended
by NIOSH (1976).  Kinetic parameters for the production of CO and formation of
COHb,  calculated from experimental DCM exposure data  of blood COHb concentra-
tions  in humans  and  rats, have similar values.  Estimations of V  for humans
and rats are 15 and 13.5 percent COHb per 7-hour exposure,  respectively, and
estimates of K   are  200 and 170 ppm, respectively.   These  findings  indicate
that the CO pathway is saturated with an inspired air  concentration  of  appro-
ximately 400 ppm DCM (2 x K ) in both humans and rats.
005DC5/C                           4-57                          11-14-83

-------
4.8  REFERENCES

Ahmed, A.  E.,  and M. W. Anders.  Metabolism  of  dihalomethanes  to  formaldehyde
     and inorganic chloride.   Drug Metab. Dispos. 4:357-361, 1976.

Ahmed, A.  E. ,  and M. W. Anders.  Metabolism  of  dihalomethanes  to  formaldehyde
     and  inorganic  halide.  II.  Studies  on  the mechanism of  the  reaction.
     Biochem. Pharmacol. 27:2021-2025, 1978.

Anders, M. W. and J. M. Sunram.  Transplacental  passage of dichloromethane and
     carbon monoxide.  Toxicol. Lett. 12:231-234, 1982.

Anders, M. W.,  V. L. Kubic and A.  E. Ahmed.  Metabolism of halogenated methanes
     and  macromolecular binding.   J.  Environ.  Pathol. Toxicol.  1:117-121,
     1977.

Astrand,  I., P.  Ovrum, and A. Carlsson.  Exposure to methylene chloride.  I.
     Its concentration  in  alveolar  air  and blood  during  rest and  exercise and
     its metabolism.  Scand.  J. Work Environ. Hlth. 1:78-94, 1975.

Bourne, W. ,  and  R.  L.  Stekle.  Methylene  chloride  in  anesthesia.   Can.  Med.
     Assoc. J.  13:432-433, 1923.

Carlsson, A., and M. Hultengren.   Exposure to methylene chloride.   III.   Meta-
     bolism of l4C-labeled methylene chloride in  rat.  Scand.  J. Work Environ.
     Hlth. 1:104-108, 1975.

Ciuchta,  H.  P.,  G.  M.  Savell, and  R. C.  Spikes.   The  effect of alcohols and
     toluene upon  methylene   chloride  induced carboxyhemoglobin  in  rat  and
     monkey.  Toxicol.  Appl.  Pharmacol.  49:1-20,  1979.

Coburn, R.  E.   The  carbon monoxide  body  stores.   In:   Biological  effects of
     carbon monoxide.   Proceedings of a Conference7~New York Academy of  Science,
     New  York, January 12-14, 1970.  Ann. N. Y.  Acad.  Sci.  174:11-22,  1970.

Coburn, R. F.  Endogenous carbon monoxide metabolism.  Ann.  Rev. Med. 24:241-250,
     1973.

Collison,  H. A.,  F.  L. Rodkey, and  J. D. O'Neal.  Effects of  dichloromethane
     on hemoglobin function.   Biochem. Pharmacol. 2i5: 557-558,  1977.

Cox,  P.  J. ,  L.  J.  King, and  D.  V.   Parke.   The  binding  of trichlorofluoro-
     methane and other  haloalkanes to cytochrome  P-4bu under aerobic and anaerobic
     conditions.  Xenobiotica 6:363-375,  1976.

Cunningham,  M.  L. ,  A.  J.  Gandolfi,  K.  Brendel, and I. G.  Sipes.   Covalent
     binding of  halogenated volatile solvents to  subcellular macromolecules  in
     hepatocytes.   Life Sci.   29:1207-1212, 1981.

Dill,  D.  C. , P.  G.  Watanabe, and J. N.  Norris.    Effect of methylene chloride
     on the  oxyhemoglobin dissociation curve of rat and human  blood.  Toxicol.
     Appl.  Pharmacol.  46:125-129, 1978.

DiVincenzo,  G.  D.,  and M.  L.  Hamilton.   Fate and disposition of 14C-methylene
     chloride  in the rat.  Toxicol.  Appl.  Pharmacol.  32:385-393,  1975.

005DC5/D                            4-58                            11-14-83

-------
DiVincenzo,  G.  D.  and C. J.  Kaplan.   Uptake, metabolism, and elimination of
     methylene  chloride  vapor by  humans.  Toxicol. Appl.  Pharmacol.  59:130-140,
     1981a.

DiVincenzo,  G.  D.  and C. J.  Kaplan.   Effect of exercise  or  smoking on the
     uptake,  metabolism,  and excretion of methylene  chloride vapor.   Toxicol.
     Appl. Pharmacol. 59:141-148,  1981b.

DiVincenzo,  G.  D. ,  F.  J.  Yanno,  and B.  D.  Astill.   Human and canine exposures
     to  methylene  chloride  vapor.  Am.  Ind.  Hyg. Assoc.  J. .33:125-135,  1972.

Engstrom,  J. ,  and  R. Bjurstrom.    Exposure  to methylene  chloride.  Content  in
     subcutaneous  adipose  tissue.  Scand.  J. Work Environ.  Hlth.  3:215-224,
     1977.

Fenn, W.  0.   The burning of CO  in  tissues.   Ann. N. Y.  Acad.  Sci. 174:64-71,
     1970.

Fodor, G.  G. ,  and  A. Roscovanu.   Increased blood CO content  in  humans and
     animals caused  by incorporated halogenated  hydrocarbons.   Zbl.  Bakt. Hyg.
     162:34-40, 1976.

Fodor, G. , D.  Prajsnar,  and H-W.  Schlipkoter.   Endogenous CO-Bildung  durch
     inkorporierte Halogankohlenwasser-Staffe der Methanreihe:  Staub-Reinhalt.
     Luft  33:258-259, 1973.

Forster,  H.  V.,  S.  Graff, C. L.  Hake,  R.  Soto, R.  D.  Stewart.  Pulmonary-
     hematologic studies on humans during exposure to  methylene  chloride.
     Report  No.  NIOSH-MCOW-ENVM-MC-74-4.   The Medical College  of  Wisconsin.
     Department of  Environmental  Medicine,  Milwaukee,  WI, April 1974.   17 pp.

Friedlander,  B.  P.,  T. Hearne, and S.  Hall.   Epidemiological  investigation  of
     employees  chronically  exposed to  methylene chloride.  J. Occup.  Med.
     20:657-666, 1978.

Hake, C.  L.   Simulation  studies  of blood carboxyhemoglobin levels associated
     with  inhalation exposure to  methylene chloride.   Toxicol.  Appl. Pharmacol.
     48:A56 (abstr), 1979.

Hake, C.   L. ,  R.  D.   Stewart, H.  V.  Forster, A. J. Lebrun,  J. E. Peterson, and
     A.  Wu.   Results of  the  controlled  exposure  of human females  to  the vapor
     of methylene  chloride.   Report No. NIOSH-MCOW-ENVM-MC-74-3.  Milwaukee,
     Wis.  The  Medical   College  of Wisconsin,  Department of  Environmental
     Medicine, Milwaukee, WI, March 1974.  22 pp.

Haun, C.   C. ,  E.  S.   Harris,  and  K.  I.  Darmer.   Continuous animal  exposure to
     methylene chloride.   AMRL-TR-71-120, #10, _In:   Proceedings 2nd  Conference
     Environmental   Toxicology,  Wright-Patterson  Air  Force  Base,  OH,  pp.
     125-135, 1971.

Haun, C.  C., E.  H.  Vernot, K. I.   Darmer, and  S. S. Diamond.  Continuous  animal
     exposure to low levels  of  dichloromethane.  AMRL-TR-130,  #2.   In:   Pro-
     ceedings of the 3rd Conference on Environmental  Toxicology, Wright-Patterson
     Air  Force  Base, OH, Aerospace Medical  Research Laboratory,  1972.   pp.
     199-208.

005DC5/D                            4-59                           11-14-83

-------
Heppel, L. A.,  and  V.  T.  Porterfield.   Enzymatic  dehalogenation of certain
     brominated and chlorinated compounds.  J. Biol.  Chem.  176:763-769,  1948.

Heppel, L. A.,  P.  A.  Neal,  T. L.  Perrin,  N.  L.  Orr, and V. T. Porterfield.
     Toxicology of dichloromethane I.  Studies on effects of daily inhalation.
     J. Ind.  Hyg.  Toxicol. 26:8-16, 1944.

Hogan,  G. K. ,  R.  G.  Smith, and  H.  H.  Cornish.   Studies on the microsomal
     conversion of  CH2C12 to  CO.   Toxicol. Appl. Pharmacol. 37(1):112,  1976.

Kahn, A.,  R.  B.  Rutledge, G.  L.   Davis,  J.  A.  Altes, G.  E.  Gantner, C.  A.
     Thornton,  and  N.  D.  Wallace.   Carboxyhemoglobin  sources  in the metro-
     politan  St.  Louis population.  Arch.  Environ.  Hlth.  29:127-135, 1974.

Kasselbart, V., and J.  Angerer.   Influence of dichloromethane on  the disappear-
     ance  rate  of  ethanol  in  the blood of rats.   Int. Arch.  Arbeitsmed.
     33:231-236, 1974.

Kubic, V.  L. , and M. W. Anders.   Metabolism of dihalomethanes to  carbon  monoxide.
     II.   I_n vitro  studies.   Drug Metab. Dispos.  3:104-112, 1975.

Kubic, V.  L.  and  M. W. Anders.    Metabolism in dihalomethanes  to carbon  mon-
     oxide. III.  Studies on the mechanism of the reaction.  Biochem. Pharmacol.
     27:2349-2355,  1978.

Kubic, V.  L. ,  M.  W. Anders, R.  R.  Engel,  C.  H.  Barrow,  and W.  S.  Caughey.
     Metabolism of  dihalomethanes  to carbon monoxide.  I.  Iji vitro studies.
     Drug Metab. Dispos.  2:553-557,  1974.

Kurppa, K. and  H. Vainio.   Effects of  intermittent dichloromethane inhalation
     on blood carboxyhemoglobin concentration and drug metabolizing  enzymes  in
     rat.   Res. Comm.  Chem.  Path.   Pharmacol. 32:535-544, 1981.

Kurppa, K. ,  H.  Kivisto,  and H.  Vainio.   Dichloromethane  and  carbon monoxide
     inhalation:  carboxyhemoglobin  addition, and drug metabolizing  enzymes  in
     rat.  Int. Arch.  Occup. Environ. HHh. 49:83-87, 1981.

Kuzelova, M., and R. Vlasak.  Der  Einfluss  von Methylenchlorid  auf  die Gesundheit
     der  Arbeiter bei  der  Herstellung  von  Filmfolien und Ermittlung  der
     Ameisensaure als  Stoffwechselproduct des  Methylenechlorids.   Praco. Lek.
     18:167-170, 1966.

Lambertsen,  C.  J.   Effects  of excessive pressures  of oxygen,  nitrogen,  carbon
     dioxide,  and  carbon  monoxide.  Implications  in  aerospace,  undersea,  and
     industrial environments.  In:   Medical Physiology, ed. Mountcastle, V.  B.
     Vol.  2,  13th ed., C. V. Mosby Co.,  Saint Louis,  MO,  1974.

Latham,  S. ,  and M.  Potvin.    Microdetermination  of dichloromethane  in blood
     with  a  syringeless  gas  chromatographic injection system.   Chemosphere
     6:403-411, 1976.

Lehmann,  K.  B., and L. Schmidt-Kehl.   (The thirteen  most  important chlorinated
     aliphatic  hydrocarbons from  the standpoint  of  industrial  hygiene.)   Arch.
     Hyg.  116:131-268, 1936.  (Ger)


005DC5/D                           4-60                           11-14-83

-------
 Leonardos,  G. ,  0.  Kendall, and  H.  Bernard.  Odor  threshold  determinations  of
     53  odorant chemicals.   J.  Air  Pollut.  Contr.  Assoc.  19:92-95,  1969.

 Lindqvist,  T.   Partition coefficients of  some  commonly used solvents.  Ind.
     Environ.  Xenobiot.  Proceedings International  Conference,  1977.   Prague,
     J.  R.  Fonts, and  I. Gut., eds.,  pp.  165-167.  Excerpta  Medica,  Amsterdam-
     Oxford,  1978.

 Llewellyn,  0.  P.  Halogenated hydrocarbons  used  as solvents.   Ann. Occup. Hyg.
     9:199-208, 1966.

 Luomanmaki,  K. ,  and R. E.  Coburn.  Effects  of metabolism  and  distribution  of
     carbon  monoxide  on blood and  body  stores.  Am. J.  Physiol.  271:354-363,
     1969.

 MacEwen,  J.  D. ,  E.  H.  Vernot, and  C. C.  Haun.   Continuous animal  exposure  to
     dichloromethane.   Report  AMRL-TR-72-28.  Aerospace Medical Research Lab.
     Wright-Patterson  Air  Force  Base, OH,  1972.  pp. 33.

 May, J.   Odor thresholds of solvents for assessment of solvent odors in the
     air.   Staub-Reinhalt.   Luft 26:34-38,  1966.

 McKenna, M. J., and J.  A. Zempel.   The dose-dependent metabolism  of  14C-methylene
     chloride  following oral  administration to rats.   Food Cosmetics Toxicol.
     19:73-78, 1981.

 McKenna, M. J., J. A.  Zempel, and W.  H.   Braun.   The pharmacokinetics of  inhaled
     methylene  chloride in rats.   Toxicol.  Appl.  Pharmacol.  65:1-10, 1982.

 McKenna, M. J., J. H.  Saunders,  W.  R. Boeckler,  R.  J. Karbowski,  K.  D. Nitschke,
     and  M.  B.  Chenoweth.  The pharmacokinetics  of inhaled methylene chloride
     in  human volunteers.  Paper #176, 19th  Annual Meeting,  Society  of Toxicology,
     Washington, D.C.,  March 9-13,  1980.

 Miller,  V.  L. ,  R.  R.   Engel, and M. W. Anders.  J_n vivo  metabolism of dihalo-
     methanes to carbon monoxide.   Pharmacologist  15^:190 (abstr),  1973.

 Morgan,  A. ,  A.  Black,  and  D.  R.  Belcher.   Studies on the  absorption of  halo-
     genated  hydrocarbons  and their  excretion in  breath  using  38C1 tracer
     techniques.  Ann. Occup.  Hyg.  H:273-282, 1972.

 National  Institute  for Occupational Safety and Health.   Criteria for a Recom-
     mended Standard for Occupational Exposure to  Carbon Monoxide.   HSM  73-11000,
     1972.

 National  Institute  for  Occupational  Safety  and Health.  Criteria   for  a
     Recommended  Standard  in  Occupational  Exposure to  Methylene Chloride,
     HEW No. 76-138, 1976.

Neely,  W. B.  Metabolic rate of  formaldehyde-  14C  intraperitoneally  administered
     to the rat.  Biochem.  Pharmacol. 13:1137-1142, 1964.

Nunes,  A.  C.  and  B.  P. Schoenborn.   Dichloromethane  and myoglobin function.
     Mol. Pharmacol. 9:835-839,  1973.


005DC5/D                            4-61                            11-14-83

-------
Peterson, J. E.   Post-exposure  relationship  of carbon monoxide  in blood and
     expired air.   Arch Environ.  HHh.  21:172-173, 1970a.

Peterson, J. Modeling  the  uptake metabolism and excretion of dichloromethane
     by man.  Am.  Ind.  Hyg.  Assoc.  39:41-47,  1978.

Peterson, J.  E. ,  and  R. D.  Stewart.   Absorption and elimination  of  CO by
     inactive young men.   Arch.  Environ.  HHh.  21:165-171, 1970b.

Pritchard,  A.,  and M.  J.  Angelo.   Comparative  pharmacokinetic  studies of
     methylene chloride  in oil versus drinking water.  Report commissioned  by
     National Coffee Association of U.S.A.  Inc.  Available  August,  1982.

Pritchard,  A.,  B.  Doonan  and  R. Rourke.  Methylene  chloride:   effects on
     microsomal  enzymes in mice.   Study commissioned by National  Coffee Assoc-
     iation of U.S.A.  Inc., 1982.

Ratney, R. S., D.  H.  Wegman,  and H.  B.  Elkins.   ^n vivo conversion of methylene
     chloride to  carbon  monoxide.  Arch.  Environ.  HHh.  28:223-226,  1974.

Reynolds, E. S., and A. G.  Yee.   Liver parenchyma! cell injury.   V. Relationships
     between patterns  of chloromethane-14C incorporation  into constituents  of
     liver ji_n vivo and  cellular injury.   Lab. Invest. 16:591-603, 1967.

Riley,  E. C. , D.  W.  Fassett, and W. L.  Sutton.   Methylene  chloride vapor  in
     expired air of human subjects.   Am.  Ind. Hyg. Assoc.  J. 27:341-348, 1966.

Roberts, C.  J.  C., and  F. P.  F.  Marshall.  Recovery after "lethal" quantity of
     paint remover.  Brit.  Med.  J.  6000:20-21,  January 3, 1976.

Rodkey,  F.  L. ,  and H.  R.  Collison.   Biological  oxidation  of 14C-methylene
     chloride to  carbon monoxide and carbon dioxide by the rat.  Toxicol.
     Appl. Pharmacol.  40:33-38,  1977.

Roth, R. P., R.  T. Drew, R. J.  Lo, and J. R.  Fouts.  Dichloromethane  inhalation,
     carboxyhemoglobin  concentration,  and drug metabolizing enzymes in  rabbits.
     Toxicol. Appl. Pharmacol.  33:427-437, 1975.

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

Schutz,  E.   Effect of  polyethylene glycol 400  on percutaneous absorption  of
     active  ingredients.   Arch.  Exp.  Path.  Pharmakol.  232:237-238,   1958.

Settle, W.  Ph.D thesis, University of California, 1971.

Settle, W.   Role of changes  in carbon monoxide-hemoglobin binding  in  methylene
     chloride toxicity.  Fed. Proc. Fed.  Am. Soc. Exptl.  Biol. 34:229 (Abstr.),
     1975.

Stevens,  J.  L., and M.  W.  Anders.   Studies  on  the mechanisms of  metabolism of
     haloforms  to carbon  monoxide.   Toxicol.  Appl.  Pharmacol.  45:297-298,
     1978.


005DC5/D                           4-62                            11-14-83

-------
 Stevens, J.  L.,  and M. W.  Anders.   Metabolism  of  haloforms  to carbon monoxide.
      III.   Studies  on the mechanism of the reaction.   Biochem. Pharmacol. 28:
      3189-3194,  1979.

 Stevens, J.  L. ,  J.  H. Ratnayake,  and M.  W. Anders.   Metabolism  of dihalo-
      methanes  to carbon  monoxide.   IV.  Studies  in isolated rat hepatocytes.
      Appl.  Pharmacol.  55:484-489,  1980.

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

 Stewart, R.  D. ,  and C. L.  Hake.  Paint-remover hazard.   JAMA J. Am.  Med.  Assoc.
      235:398-401, 1976.

 Stewart, R.  D. ,  C.  L. Hake,  and A. Wu.  Use  of  breath analysis  to monitor
      ethylene  chloride exposure.   Scand.  J.  Work Environ.  Hlth.  2:57-70,
      1976a.

 Stewart, R.  D. ,  R.  S. Stewart,  W.  Stamm, and  R.  P. Seelen.   Rapid estimation
      of  carboxyhemoglobin level  in  fire  fighters.  JAMA J. Am. Med.  Assoc.
      235:390-392, 1976b.

 Stewart, R.  D. ,  H.  V. Forster,  C.  L.  Hake,  A. J.  Lebrun, and J.  E.  Peterson.
      Human  responses  to  controlled exposure  of  methylene  chloride vapor.
      Report  No.  NIOSH-MCOW-ENVM-MC-73-7.   Milwaukee,  WI, The Medical  College
      of  Wisconsin,  Department of  Environmental Medicine, December,  1973.   82
      pp.

 Stewart, R.  D. ,  H.  V. Forster,  C.  L.  Hake, A. J. Lebrun and J. E.  Peterson.
      Methylene chloride:    Development  of  a  biologic standard for  the industrial
      worker  by breath analysis.  Report No.  NIOSH-MCOW-ENVM-MC-74-9.  Milwaukee,
      WI, The Medical College.of  Wisconsin,  Department  of Environmental  Medicine,
      December, 1974b.

 Stewart, R.  D. ,  T.  N.  Fisher, M. J.  Hosko,  J.  E.  Peterson,  E. D.  Baretta, and
      H.  C.  Dodd.  Carboxyhemoglobin elevation after  exposure to dichloro-
      methane.  Science 176:295-296,  1972a.

 Stewart, R.  D. ,  T.  N. Fisher, M.  J. Hosko,  J. E.  Peterson,  E.  D.  Baretta and
      H.  C.  Dodd.   Experimental  human  exposures to methylene chloride.   Arch.
      Environ. Hlth.  25:342-348,  1972b.

 Stewart, R.  D.,  E.  D.  Baretta, L.  R.  Platte, E.  B.  Stewart, J. H.  Kalbfleish,
      B.  Van  Yserloo,  and A.  A.  Rimm.   Carboxyhemoglobin levels in  American
      blood donors.   JAMA J. Am.  Med. Assoc.  229:1187-1195,  1974a.

Tenhunen, R. , H.  S.  Marver, and  R.   Schmid.   Microsomal  hemeoxygenase;  characteri-
      zation of the enzyme.  J. Biol. Chem.  244:6388-6394, 1969.

Toftgard, R., 0.  G.  N. Nilsen, and J.  S. Gustafsson.   Dose-dependent induction
      of rat  liver microsomal  cytochrome P~450  ar|d microsomal  enzymatic  activities
      after inhalation of toluene and dichloromethane.    Acta  Pharmacol.  Toxicol.
      51:108-114,  1982.


005DC5/D                           4-63                            11-14-83

-------
Tzagoloff, A. , and  0.  C.  Wharton.   Studies on the  electron  transfer  system.
     LXII.   The reaction of cytochrome oxidase with carbon monoxide.  J. Biol.
     Chem.  240:2628-2632, 1965.

Wagner, J. G.   Fundamentals  of  Clinical Pharmacokinetics.  Drug Intelligence
     Publications, Inc., Hamilton,  IL, 1975.

Winneke, G. The neurotoxicity of dichloromethane.  Neurobehav. Toxicol. Teratol.
     3: 391-395, 1981.

Winneke, G. and G. G.  Fodor.   Dichloromethane produces narcotic effects.   Int.
     J. Occup. Hlth. Safety 45(2):34-37, 1976.

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

Withey, J. R. ,  B.  T.  Collins, and  P.  G.  Collins.   Effect of vehicle on the
     pharmacokinetics  and  uptake of  four  halogenated hydrocarbons  from the
     gastrointestinal  tract  of the  rat.   Preprint  paper  submitted to  J.  Appl.
     Toxicol., December, 1982.

Yesair, D. W. ,  P.  Jaques,  P. Shepis, and R. H.  Liss.  Dose-related pharmaco-
     kinetics of 14C-methylene chloride in mice.   Fed. Proc. Am. Soc. Exp.  Biol.
     36:998 (abstr), 1977.

Zorn, H.   In:  Bericht  uber die 14 Jahrestagung  der Deutschen Gesellschaft fur
     Arbeitsmedizin e.   V., G. Lehnert, D. Szadkowski  und  H. J. Weber, eds. , p.
     343, Centner Verlage, Stuttgart, 1975.
 005DC5/D                            4-64                            11-14-83

-------
                      5.  HEALTH EFFECTS OF DICHLOROMETHANE

 5.1   HUMAN  HEALTH EFFECTS
 5.1.1 Acute  Exposures
 5.1.1.1   Nervous System and Behavior—Virtually all  experimental  exposure
 studies  of  DCM using humans have  studied  acute  exposures  and  neural  or  behav-
 ioral  dependent  variables.  Using between three and eight subjects, Stewart et
 al.  (1972b) reported that subjects had symptoms of "lightheadedness" and had
 difficulty  with  speech  articulation at DCM levels higher than 868 ppm (3,012
 mg/m3)   for more than I hour.  They also reported anecdotal data about increased
 amplitude in  certain visual evoked potentials at about the same exposure levels;
 however,  no statistical  analyses were  possible  because  of the small  number of
 subjects.   In a  later study, they reported that 50, 100, or 250 ppm (173, 347,
 or 867 mg/m3) DCM had no effects on the Romberg equilibrium test, vigilance,
 coordination,  or arithmetic skill (Stewart et al., 1973).
      Winneke  (1974) and Winneke and Fodor (1976) exposed female human subjects
 to 0,  300,  500,  and 800 ppm (0, 1041, 1730, and 2,776 mg/m3) DCM for up to 3
 hours.   They  measured  performance  on a series of hand-eye coordination tasks
 including tapping pursuit  rotor and hand steadiness.   They also measured the
 critical  flicker fusion  (CFF)  frequency and performance on an auditory  signal
 detection task.  Concentration-related decrements in signal detection, beginning
 at 300 ppm  (1041 mg/m3), were observed.  CFF frequency also decreased in a con-
 centration  related  fashion.   Many  motor decrements were  observed  at  800 ppm
 (2,776 mg/m3), the  only concentration at which these  tests  were conducted.
 Fodor  and Winneke (1971) reported a similar study in which 300 and 800 ppm (1,041
 and  2,776 mg/m3) both  decreased CFF frequency  and impaired signal  detection
 behavior; however,  the  two  DCM levels did not  produce  effects  significantly
 different from each other.
     Gamberale et al. (1975) investigated the effect of DCM on psychomotor and
 cognitive performance in  14 men,  ages 20 to 30 years, divided at random into
 two  groups.    The first group was initially exposed to progressively  increased
 concentrations of DCM—250,  500, 748 and 997 ppm (870, 1,740,  2,600, and 3,470
mg/m3)--through  a face  mask.   Seven days later they were observed under con-
 trol  conditions.   The second group was studied under identical conditions,  but
 in reverse order.  The subjects were exposed to each concentration for 30 minutes
with no break in exposure;  total exposure time for each individual was two  hours.

005DC4/A                             5-1                               11/14/83

-------
At each concentration, a peak plateau of alveolar DCM occurred in about 10 minutes.
Heart rate showed no change.   During the two hours of exposure to DCM,  no sta-
tistically significant  impairment  in  any measured performance was observed,
although reaction time appeared to be more irregular during exposure  than during
control conditions.   Numerical ability and short-term memory were unaffected.
The experiment did not consider any latency effects.
     Putz et al.  (1976) exposed six men and six women (ages 18 to 40  years) to
201 ppm (700 mg/m3) of 99.5 percent pure DCM or to 7 ppm (80 mg/m3)  CO  for 4 hours
in an  8-m3  chamber with unspecified dynamics.   Alveolar concentrations were
measured hourly,  and peripheral blood from a finger was  taken before  and after
exposure.   Peak  alveolar  CO  after 4 hours of  OCM exposure was about 17 ppm
(60 mg/m3), and COHb was 5.1 percent.   Alveolar CO after CO exposure  for 4 hours
was 4 ppm (50 mg/m3), and COHb was 4.85 percent.   Thus,  there was COHb  and alveo-
lar CO equivalence between the two compounds.   Eye and hand coordination at the
end of the exposure period was depressed in subjects exposed to either  compound,
compared with  controls, with  DCM  depression being  greater  than CO depression.
The authors  concluded that CO  may  have been primarily responsible for  the  de-
creases in performance after exposure to both compounds.
     These preceding  studies  can  only be considered a  preliminary survey  of
possible  CMS  and behavioral  consequences of DCM  exposure.  Only one  study
(Winneke, 1974)  reported  concentration  effects data on CFF and signal  detec-
tion, both possibly involving the alertness of  the subjects.  CFF is  a  primitive
measure of  CMS activation.   The lowest concentration of DCM reported to have
affected behavior was 200 ppm  (694 mg/m3) (Putz et al.,  1976).  They also repor-
ted vigilance  decrements  in  addition  to  compensatory tracking  decrements.   In
spite of these few preliminary findings, the results are consistent with the hypo-
thesis that  DCM  effects are  depressant  and  also agree well  with  the  available
laboratory animal data, reported elsewhere in this chapter.
     Data are  required on the effects of DCM  upon other behaviors.  No  data
were found  regarding  cognitive skills,  and  no  concentration-effects  data  were
available for  motor  skills or hand-eye  coordination.   No  quantitative tests
of  electrophysiological  effects of DCM  in  humans were  found.   Finally,  no
data  regarding the effects of DCM  in combination with other  pollutants  or
medicinals were  found.
     Putz et  al.  (1976)  concluded that  DCM  produced its effects  via  its  prin-
cipal metabolite, COHb.   Stewart et al.  (1972a) had suggested this possibility,

005DC4/A                              5-2                               11/14/83

-------
 Winneke  (1974)  presented  evidence  suggesting that DCM  is more  toxic  than CO.  He
 compared subjects  exposed to  50 or 100 ppm CO  for 5 hours to subjects exposed to
 300,  500,  and 800  ppm  (1041,  1730,  and 2776 mg/m3) DCM for 4 hours.  OCM-exposed
 subjects showed more effects  on the  same tasks.  Comparison to data  presented by
 Putz  et  al.  (1976)  reveals  that equivalent COHb  levels were produced by 70 ppm
 CO  and 200 ppm  (694 mg/m3) DCM.   Examination  of the  behavioral  results  of  the
 study by Putz  et  al.  reveals  that DCM always  produced  greater  decrements  than
 did CO.   Logically  DCM should be more toxic than CO, since it not only produces
 COHb,  but,  because of  its  lipophilic qualities, it probably has some primary
 neural effect.  Additional  data are  required to explore this issue.
 5.1.1.2   Other  Experimental DCM Effects—Stewart et  al.  (1972a,b)  reported no
 effects  of exposure to  500  or 1,000  ppm (1,730 or 3,460 mg/m3) DCM for 2 hours
 on various hematological,  hepatic, or renal measures.  Continuous electrocardio-
 graphic  monitoring  during  one to two exposures to 1000 ppm (5,460 mg/m3) was not
 reported to  cause arrhythmias.  Gamberale et al. (1975) reported that DCM expo-
 sure  up  to 1,000 ppm (3,460 mg/m3) produced no alterations in heart  rate.   Based
 on  laboratory animal  exposures,  it is  not  reasonable to  expect  other internal
 organ  system involvement  in healthy humans with  low-level  acute exposures.
      Stewart and  Dodd  (1964)  studied groups of  three  to five men and women
 (25 to 62 years old) who  immersed  their thumbs in various chlorinated aliphatic
 solvents.  Within  the  first 2 minutes of  immersion  in DCM, all  subjects  re-
 ported an  intense  burning sensation on the dorsal  surface  of their thumbs.
 Within 10  minutes,  a  feeling of coldness  or  numbness,  alternating with the
 burning  sensation, was  reported; the slightest movement of the thumb triggered
 waves  of searing  and  excruciating pain.   Following removal  from the solvent,
 their  thumbs were  described  as  "numb  and cold" or "asleep."   A very mild
 erythema  and a  hint of white  scaling were  the  only overt signs of  irritation.
Within 1 hour,  both the erythema and paresthesia subsided.
     Studies of human  blood jji  vitro from individuals homozygous  for sickle
cell anemia (Matthews et al.,  1977, 1978) have shown that DCM caused hemolysis,
preferentially of  sickle and other abnormal cell types.   These reports suggest
that acute exposure to relatively high concentrations of DCM may result in anemia,
particularly in individuals who  are predisposed to  hemolytic anemias, e.g., per-
sons with  sickle cell  anemia or glucose-6-phosphate dehydrogenase deficiency.
However,  more compelling evidence is required  to ascertain  that such an effect
exists.

005DC4/A                             5-3                                11/14/83

-------
5.1.1.3  Accidental Exposure—A number of  case  reports  have  implicated  expo-
sure to DCM as one factor in human fatalities.   Repeated paint  stripping acti-
vities were associated with  the  death of a 66-year-old  man  in a report by
Stewart and Hake  (1976).   Although  mixed exposure occurred,  DCM is strongly
implicated as a causative  agent  because  of the striking association between
exposure and  chest  pain,  the high correlation  between  cardiovascular  effects
of pure DCM and a  commercially-formulated paint stripper (Aviado et al., 1977),
and the well-established metabolism  of  DCM to  CO.   Neither air levels of DCM
nor COHb levels were discussed in the report.
     DCM was a probable factor in the death of  a workman who  became unconscious
and fell  into  a vat containing the  solvent.  Death  occurred  after repeated
myocardial infarctions (Kuzelova et  al., 1975).  Overexposure to DCM may have
been a  factor  in  the death of one of four men  rendered unconscious in a fac-
tory where  DCM was  used  to extract oleoresin  from  dried  plant materials.
Death was attributed  "presumably"  to CNS depression (Moskowitz and Shapiro,
1952).
     Excessive prolonged exposure to DCM was judged by  Bonventre et al.  (1977)
to be  the  principal  factor in the death  of  a  13-year-old boy who had  used
paint remover containing DCM.  Neither  COHb nor ambient concentrations were
reported.   Levels   of  DCM  found in the liver, blood, and brain were 14.4 mg/
100 g, 51 mg/100 ml, and 24.8 mg/100 g,  respectively.
     High COHb levels  in blood were  found  in a 40-year-old male who was found
unconscious after  exposure  to  an unknown concentration  of DCM leaking  from
a degreasing factory (Benzon et al., 1978).  When examined 1.5 hours after the
incident, this individual  had  an initial arterial blood COHb  of 19 percent,
decreasing to  11  percent after 20 hours, and to 4  percent  after 28 hours.   He
was discharged from the hospital  with no further complications.  Another exposed
male, 50 years old and with a history suggestive of ischemic  heart disease, did
not lose consciousness.  When he  was examined,  his COHb level  was 11 percent,  and
an early mid-diastolic heart murmur was  found at the lower left sternal  border.
The ECG showed left anterior fascicular  block and sinus bradycardia.   The follow-
ing day, the ECG showed bundle branch block and nonspecific ST-T wave changes;  at
this time the COHb level was 6 percent.   The myocardial fraction of creative phos-
phokinase was  not elevated,  suggesting no  myocardial injury.   Subsequent  ECGs
showed no further changes.   Thirty-four hours after his admission to a hospital,
blood levels of COHb had decreased to 3 percent.  He was subsequently discharged.

005DC4/A                             5-4                                11/14/83

-------
      Exposure  to  DCM  vapors  as  well  as  skin  contact with  DCM  resulted  in  typical
 symptoms  of  solvent intoxication  for a  worker who was exposed  for about 4  hours
 to DCM vapors  emanating  from an open drum  and to DCM liquid from dipping  his bare
 hands into the liquid to  clean  copper gaskets (Hughes,  1954).  There was  no pro-
 vision for air ventilation in the workrooms.  Symptoms  included excessive  fatigue,
 weakness,  sleepiness, 1ightheadedness,  chilly sensations, and nausea.  About
 2  hours after  the exposure,  cough and substantial pain  developed and pulmonary
 edema was diagnosed upon  his admission  to  a  hospital.   DCM was considered  to be
 the  pronounced respiratory irritant.  In contrast to the  observation of Stewart
 and  Dodd  (1964),  no adverse  dermal effects were noted.
      Asphyxia  was judged to  be  the  cause  of  death  of a  20-year-old  male  found
 slumped over a tank  containing DCM  (Winneke et al., 1981).  Autopsy findings
 included  chemical burns  on  the forehead,  moderate  edema of the brain,  and
 pulmonary congestion  with focal hemorrhage.   The level  of DCM in postmortem
 blood was 29.8 mg percent.
      Systematic conclusions  from accidental  exposure data are difficult.    Exact
 exposure  levels and durations are unknown.   Simultaneous  exposure to other sub-
 stances is almost always the  case.   Nonetheless, some observations can be  sum-
 marized.
      Human case studies involving myocardial infarction have been reported and
 have  included  fatalities resulting from, or closely associated with, exposure to
 DCM.   Nonfatal  exposures  have caused ECG  changes that  were similar  to those
 induced by CO.  (It is as yet unclear what the relative contributions of DCM and
 its metabolite, CO, are to those effects.)  The case histories of certain exposed
 individuals  suggest the  existence of underlying cardiovascular disease.   This
 effect may,  therefore, be significant to this human subpopulation.   Hepatoto-
 xicity has not been  reported in any  human case report, even  following fatal
 exposures.  The only evidence of human nephrotoxicity resulting from DCM exposure
was a finding of  congested kidneys following a fatal  exposure.  Ocular toxicity
 other  than eye irritation and congested conjunctivae has  not  been reported in
 humans exposed to DCM.
5.1.2  Chronic Effects
 5.1.2.1   Experimental  Exposure—Stewart  and  his colleagues  extended  their
 studies of the effects of DCM  on human subjects to include longer exposure.
 In one study  (Hake et al.,   1974), male volunteers were exposed 5 days/week
for 5 weeks to DCM concentrations  of 50  ppm (173 mg/m3)  in week 1,  250 ppm (867

005DC4/A                             5-5                               11/14/83

-------
mg/m3)  in week 2, 250 ppm (867 mg/m3) in week 3, 100 ppm (346 mg/m3) in week 4,
and 500 ppm (1730 mg/m3) in week 5.  Three subjects were exposed for 1 hour/day,
three subjects were exposed for 3 hour/day, and four subjects were exposed for
7.5 hour/day.   This complex experimental design was further confounded by attempts
to separate smoking  and nonsmoking populations.  The information that can be
gathered from this report is summarized in Table 5-1.   Alongside is a summary of
a companion report from Stewart's group (Forster et al., 1974), which exposed nine
female  subjects to 250 ppm (867 mg/m3) DCM for 1,  3, or 6.5 hour/day for 5 days.

            TABLE 5-1.  COHb CONCENTRATIONS IN NONSMOKERS EXPOSED
                    TO DCM AT 250 ppm (869 mg/m3)  FOR 5 DAYS
Exposure Time
h/day
Pre-exposure
1
3
7.5
Mean of Daily Maximum
Male (n)a
0.9 (5)
3.3 (1)
7.0 (1)
9.6 (3)
Observed Average COHb (%)
Female (n)
1.4 (8)
3.5 (3)
5-1 (1)
10.1 (4)
aHake et al.,  1974.
bForster et al.,  1974.

     Carboxyhemoglobin levels up to 10 percent were reported in 10 men (20 to
39 years old) and  9 women (20 to 41 years old) exposed to DCM concentrations
between 40 and 500  ppm (139 and 1,737 mg/m3) by inhalation (Peterson, 1978).
Exposure durations were  between  1  hour and  7.5  hours/day  for  not more than
5 successive days.
5.1.2.2  Accidental  Exposure--Degenerative nervous system disorder was reported
in a 39-year-old chemist  occupationally exposed  for about 5 years to  air con-
centrations of pure DCM estimated to range from 660 to 900 ppm (2,293 to 3,127
mg/m3)  (Weiss, 1967).  No liver or kidney damage or ECG alterations were re-
ported.  Dermal  contact  had  also occurred, and erythema and fissures  appeared
on the hands and forearms.   This individual's progressive visual and  auditory
illusions and hallucinations were correlated with exposure to DCM.   Neurologic
and  psychiatric  examinations  excluded an   underlying  psychosis.   Toxic
encephalosis due  to occupational exposure  to DCM was diagnosed.

005DC4/A                             5-6                               11/14/83

-------
      Collier  (1936)  reported four cases of occupational  exposure to paint re-
 mover containing  approximately  96  percent  DCM.  The men,  all of  whom were  pro-
 fessional  painters,  had been exposed  to  lead for 5  to 14 years.  During one
 autumn,  while  they were  removing paint, the workers complained of  loss  of  appe-
 tite,  dullness, faintness,  and  giddiness while using the  remover and during  the
 following  few  hours.   Two  of the  workers  were examined by the author and  the
 findings were  reported  in detail.  One  painter, aged 42,  complained of  leg and
 arm  pains,  precordial  pain, great fatigue, and  blurred  vision.   The author
 diagnosed  the  symptoms  as slight chronic lead intoxication and acute DCM-induced
 toxemia,  since the  acute symptoms  subsided upon  cessation of working  with the
 paint  remover.  The second  painter, aged 45,  experienced  tingling  of the  hands
 and  feet,  in addition to symptoms  of fatigue  and  drowsiness.  Upon cessation of
 exposure to DCM,  all of the symptoms subsided.
      Barrowcliff  (1978)  and Barrowcliff and Knell (1979)  reported  that  an in-
 dividual exposed  to  300 to  1,000  ppm  (1,042  to  3,474  mg/m3) DCM for  3  years
 developed  bilateral  temporal lobe degeneration.   This was  thought to  result
 from  chronic CO intoxication as a  result of exposure to DCM.
 5.2   EFFECTS ON LABORATORY  ANIMALS
     Much  of  the  experimental  research on  the effects  of  DCM has been  done on
 laboratory animals.   Such work can be used to elucidate the  general principles
 of DCM action  and to discover which target organs  or systems are involved.
 Generalization of laboratory animal findings to humans, however, is not straight-
 forward.  Human health effects must, in the final analysis,  be assessed in man.
 5.2.1  Acute Effects
 5.2.1.1  Lethality--Qn1y preliminary work  has been done  on  the  lethality of
 DCM.    No systematic data exist which would make it possible  to specify a time-
 by-dose-by-lethality function.
     Table 5-2 is a summary of lethality data.  The inhalation data are remark-
 ably consistent between rats and mice.   The one report on guinea pigs suggests
 that they  may  be  more  sensitive to DCM but procedural variation  cannot  be
 ruled  out.   Data  on shorter exposures  are  required since  accidental exposures
 and substance abuse frequently involve  high concentration, short-term profiles.
A curve of LCbu for a range of exposure times  would be valuable.
     Oral and intraperitoneal injection LDbu data appear to be  remarkably con-
sistent across investigations and species.   Exceptionally low values (Zakhari,
1965)  and  high values  (Ugazio  et al.,   1973)  are  possibly due  to procedural

005DC4/A                             5-7                               11/14/83

-------
 o
 o
                                           TABLE 5-2.  ACUTE LETHAL TOXICITY OF DCM
 i
 CO

Route of
Administration
Coral
Intr aperi tonea 1

Inhalation




Species
mice
rats
rats
mice
mice
mice
mice
mice
dogs
mice
mice
rats
rats
guinea
pigs
Duration of
Dose Exposure
1987 mg/kg
4368 mg/kg
2388 mg/kg
448 mg/kg
1330 mg/kg
1995 mg/kg
1990 mgAg
1990 mgAg
1260 mgAg
14,100 ppm 6 hours
16,100 ppm 7 hours
ppm
25,600-28,000 1.5 hours
ppm
16,100-18,150 6 hours
ppm
11,550 ppn 6 hours

Effect
LD50
LD50
LD50
LD50
100%
survival
20%
survival
LD50
LD50
LD50
LD50
LJD50
(a)
(a)
LD50

Reference
Zakhari (1977)
Ugazio et al. (1973)
Kimura et al. (1971)
Zakhari (1977)
Plaa and Larson (1965)
Plaa and Larson (1965)
Klaassen and Plaa (1966)
Gradiski et al. (1974)
Klaassen and Plaa (1967)
Gradiski et al. (1974)
Svirbely et al. (1947)
Berger and Fodor (1968)
Berger and Fodor (1968)
Balmer et al. (1976)


               (a) Cessation of brain electrical activity in all  animals.   1 ppm = 3.474 rng/m3.
CO
OJ

-------
differences.  The  mean  LD50  from  Table  5-2,  excluding  the  low  and  high  values
mentioned,  for injection and oral doses in rodents is near 2000 mg/kg.  No data
for other sites of injection were found.  Oral and intraperitoneal data cannot
be compared with inhalation data because of the "first pass effect" which occurs
following oral or  intraperitoneal administration.
     Morbidity resulting after short-term exposure includes effects on several
different organs.  Ocular damage, liver damage, kidney damage, changes in car-
diac parameters, and increased pancreatic bile duct flow all occur.
     The course  of events  during lethal exposures to  DCM  via  inhalation  was
described by  Berger  and Fodor (1968) who exposed  rats to  DCM  concentrations
ranging from  2,800 to  28,000 ppm (9.7  to 97.3 g/m3).  In  these  exposures an
initial period  of  excitation was followed by  a deep narcosis  accompanied by
a decrease  in muscle tone and a  reduction in  brain electrical  activity.   Rats
exposed to  concentrations to  25,000  to  28,000  ppm  (87  to 97.3  g/m3) ceased to
exhibit electroencephalographic  (EEC) activity after  only 1.5 hours; those
exposed to  concentrations between 16,000 and 18,000 ppm (55.6 and 62.5 g/m3)
ceased to exhibit  EEC activity after 6 hours.
5.2.1.2  Nervous System and Behavior—The nervous  system  is a likely target
organ because DCM  is  lipophilic  and thus is  concentrated  in myelin.   Brain
tissue has  been  shown  to have high  concentrations of  DCM  following exposure
(Savolainen et  al.,  1977;  Bergman,  1978; see also Section 4.1.3).   Pankow
et al. (1979)  showed  that  nervous tissue function is affected by DCM dosage.
They demonstrated a linear decrease in sciatic nerve conduction velocity, using
intraperitoneal  injections  of DCM at dose levels  of 1 to 6 mmoles/kg.
     EEC and  rapid eye  movement  (REM) have  been  used  to  characterize sleep
during DCM inhalation.   These data form a concentration-related continuum from
500 to 3000 ppm  (1,735  to  10,410 mg/m3)  in  which REM sleep was  reduced  in
duration beginning at about  500  ppm (1,735  mg/m3) (Fodor and Winneke, 1971).
At 3,000 to 9,000  ppm  (10,410 to 31,230 mg/m3), sleeping  time was increased
with a further  decrease in  REM sleep (Berger  and  Fodor, 1968).  Higher con-
centrations eventually produced a flat EEC after  coma (Berger and Fodor, 1968).
The above may be  viewed as  a continuum  of  effects beginning with REM sleep
reduction at  500 ppm  (1,735  mg/m3) and ending in  brain death after 1.5 hours
at 27,000 ppm (93.7 mg/m3)  or 6 hours at 17,000 ppm (59 mg/m3).
     Measures  of general activity level  are commonly used variables  to char-
acterize behavioral effects.   However,  general  activity is not  a  simple  or

005DC4/A                             5-9                               11/14/83

-------
unidimensional  measure.   Frequently,  some measures  of activity level  are affected
though others are not.   A concentration of 5,000 ppm (17,370 mg/m3)  DCM admini-
stered for 1 hour  to  male rats was  sufficient  to  decrease running  activity
(Heppel and Neal ,  1944).   Running  activity increased after exposure, but was
lower than activity in the same male  rats when they had not been exposed to DCM.
Similarly, Thomas et al.  (1971) reported that the spontaneous activity of mice
was decreased by a 3-hour exposure  to 1,000 ppm (3,474 mg/m3) of DCM.  Weinstein
et al. (1972) exposed female mice in groups to about 5,000 ppm  (17,370  mg/m3)
DCM  for  24 hours  and observed progressive decreased spontaneous activity.
     Little can  be said  about  the effects of acute DCM exposure upon the
nervous system and behavior  of laboratory animals.  The effect appears to be
consistently depressive  above  500 ppm  (1,737 mg/m3).  However,  to say  that  an
effect is "depressive" is to oversimplify matters.   Also,  only activity level or
sleep has been measured with DCM acute exposure in laboratory animals.   No work
has  been  found on  the effects  of acute  DCM exposure  upon motor  tasks,  sensory
discrimination, schedule-controlled behavior, or response acquisition in labora-
tory animals.  Similarly, no research has been published concerning  the effects
of acute DCM exposure on any electrophysiological variable other than nerve con-
duction velocity  and  in  a secondary  manner,  brain  electrical  activity.   There
have also been no  follow-up studies to show whether acute exposure effects were
(or were not) reversible.
5.2.1.3  Cardiovascular  Effects—Aviado et al. (1977) used pentobarbital-anes-
thetized,  artificially  ventilated, open-chested dogs to examine the cardio-
vascular  effects  of exposure to 0.5,  1.0,  2.5,  and 5 percent pure DCM  (5,000,
10,000, 25,000,  and 50,000 ppm) (1 ppm = 3.474  mg/m3)  and  to the same  concen-
trations of a paint stripper composed of 90.2 percent DCM, 4.2  percent  methanol,
3.2  percent  isopropanol, and 2.4 percent toluene.   Pure DCM or paint stripper
was  administered  via  an endotracheal catheter  for 5 minutes.   At the  higher
doses  (2.5 percent and 5 percent), all  hemodynamic effects were consistent  with
primary  depression of myocardial contractility (Table 5-3):   left ventricular
dp/dt  (the time-dependent  rate  of rise  of  ventricular pressure, which  is a  mea-
sure  of myocardial  contractility), fell,  as did  LV pressure.   Left ventricular
end-diastolic pressure  (left ventricular  filling pressure)  rose.  However,  car-
diac  output  fell.   Mean  arterial pressure  fell,  but  calculated  peripheral  vas-
cular resistance  (MAP-central  venous pressure/cardiac output)  rose.   Heart  rate
fell  nonsignificantly when 5 percent DCM (50,000 ppm)  (174 g/m3) was employed.

005DC4/A                              5-10                              11/14/83

-------
         TABLE 5-3.   SUMMARY OF CARDIOTOXIC ACTION OF 5% DICHLOROMETHANE
Function
HR
MAP
LVP
CVP
LVEDP
LVdP/dt
CO
SV
VR
Taylor et
1%
0%
0%
0%
9%
17%
14%
14%
25%
al.
t



t
4-
4-
4-
t
(1976)
(NS)



(NS)




Aviado et
11%
4%
4%
—
125%
22%
36%
33%
44%
al.
4,
4-
4-

t
4-
4-
4-
t
. (1977)b
(NS)
(NS)
(NS)


(NS)



 Comments:
          Rabbits

Pentobarbi tal-anesthetized
Spontaneously breathing
Close-chested
1-min.  exposure (except
   1.5 min.  for CO)
Static dose (5%)
           Dogs

Pentobarbital-anesthetized
Artificially ventilated
Open-chested
5-min. exposure
Increasing doses (0.5, 1.0,
   2.5, and 5.0%)
  Key:         HR - Heart Rate
              MAP - Mean Arterial  Pressure
              LVP - Left Ventricular Pressure
              CVP - Central  Venous Pressure
            LVEDP - Left Ventricular End-Diastolic Pressure
          LVdP/dt - Left Ventricular Rate of Time-Dependent Pressure Change
               CO - Cardiac  Output
               SV - Stroke Volume
               VR - Vascular Resistance (Peripheral  or Systemic)
               RR - Spontaneous Respiration Rate
               MV - Minute Volume
               NS - Not significantly different from control values at p <0.05

  Results presented are for  the highest dose (5%) of dichloromethane.


     The authors concluded that the effects were compound-related and that DCM

exerted a negative inotropic action.
     The effects of  DCM  on  cardiovascular function were studied in groups of
six male New  Zealand white  rabbits (Taylor et al. ,  1976).   The animals (2.4-

3.4 kg b.w.) were anesthetized with pentobarbital and bilaterally vagotomized;

results were recorded for 1-minute exposures to 5 percent (50,000 ppm, 174 g/m3)

DCM in  air  (except for cardiac output, which  was  measured over a period of

1.5 min.),  enriched to 40 percent  oxygen, and delivered via an endotracheal tube

to spontaneously breathing animals.  No changes were observed in the mean arterial
005DC4/A
              5-11
                   11/14/83

-------
pressure, left ventricular pressure, left ventricular end-diastolic pressure,
or heart rate.   However,  significant depression of left  ventricular dP/dt (rate
of rise in ventricular pressure,  an index of myocardial  contractility),  cardiac
output, and stroke  volume  occurred.   Exposure to DCM resulted  in  increased
peripheral vascular resistance,  apparently to compensate for the decreased car-
diac output and to maintain arterial pressure (Table 5-3).
     Aviado's  study in dogs correlates  well  with Taylor's findings  in rabbits.
At 5 percent,  DCM decreased left ventricular dP/dt,  cardiac output, and  stroke
volume.  An increase in left ventricular end-diastolic pressure  and peripheral
vascular resistance was  observed  in both animal models.   The mean arterial
pressure  and  left ventricular pressure were  not  affected  in either study.
     One difference between  the  studies of Aviado and Taylor was the lack of
change in the  heart rate of  rabbits compared with an apparent decrease in the
heart rate of  dogs.   This could  be due  to a difference in the length of  exposure,
since rabbits  were exposed for only 1 minute, and dogs were exposed for  5 minutes.
Also, vagotomy of the rabbits may have  prevented a parasympathetically mediated
fall in heart  rate.   The  only other difference between the  two studies was that
left ventricular end-diastolic pressure did not change in rabbits,  but signifi-
cantly increased in dogs.
     Trained male beagles (size  unspecified) were repeatedly exposed to  37,000
and 70,000 mg/m3  (10,650 and 20,150 ppm) pure DCM  (Reinhardt et al. , 1973).
After 5 minutes of  exposure  to  DCM, 0.008 mg/kg epinephrine was injected in-
travenously.   Sensitization to epinephrine did not develop.  Higher concentra-
tions of DCM could not be given  because the animals became  hyperactive,  similar
to Stage II general  anesthesia.
     In two reports,  Aviado  and  coworkers (Aviado,  1977; Zakhari,  1977) men-
tioned arrhythmias occurring in  dogs exposed to 5 percent (50,000 ppm; 174 g/m^)
DCM.  The simultaneous administration of alcohols prevented these arrhythmias.
In a subsequent review Aviado (1975) stated 0.5 percent of  (5,000 ppm; 17 g/m3)
DCM sensitized the heart  to epinephrine.  Levels of 2.5 and 5.0  percent  (25,000
and 50,000 ppm)  were  reported to induce  arrhythmias  in  anesthetized monkeys
(Aviado, 1975; Belej et al. , 1974) but produced only tachycardia in unanesthe-
tized monkeys  (Aviado, 1975).  However, Aviado and coworkers have not described
their experimental protocols or  the nature of the arrhythmias in their dog and
monkey experiments.
005DC4/A                             5-12                              11/14/83

-------
     Aviado and Belej (1974)  ^xposed Swiss mice (25 to 35 g) five times to DCM
concentrations of  20  and 40  percent v/v in oxygen.  A face mask of (unspeci-
fied)  description  was  used.   Cardiac arrhythmia was produced  by the higher
concentration.  Intravenous  epinephrine  0.006 mg/kg sensitized the heart  to
arrhythmias induced by the lower dose.
     Loyke (1973)  induced chronic renal hypertension in 13 Sprague-Dawley rats
(100 g, sex not specified).   Both poles of one kidney were ligated and the con-
tralateral one was removed.   After maintenance of high systolic pressure for
3  months,  11  experimental  rats and three  controls were  injected subcutane-
ously  with 2  mg/kg unspecified DCM, biweekly for  15 doses.  Two hypertensive
rats were used for positive controls and 11 normotensive rats as naive controls.
DCM  reduced the  systolic blood pressure of the hypertensive rats from 200 to
160 mm  Hg.  The positive controls  remained hypertensive.   DCM  was  ineffective
in reducing blood pressure in normotensive rats.    No changes ascribable to DCM
were seen in the liver.
     Douglas et al. (1976) and Wilkinson et al. (1977) used spontaneously hyper-
tensive rats  weighing 400 mg  (sex  not  specified).  Six  were  injected  sub-
cutaneously with  90 mg/kg  unspecified DCM.  Six others were  used as controls.
DCM  reduced the  blood pressure of  the  hypertensive  rats  by  about  10  percent
but  was ineffective in  reducing  blood pressure of  the  normotensive rats.
     Adams and Erickson  (1976) exposed eight trained mongrel  dogs  (size  and
sex  not specified)  to  1,700, 3,400, 6,800, and 17,400 mg/m3 (489, 979, 1,957
and  5,009  ppm) DCM for  2  hours via a permanent tracheotomy.   The animals
were exposed  repeatedly  but  randomly  to all concentrations.  Frequency  of ex-
posure  was not given.   COHb, which continued  to increase for  2 hours after
exposure,  was  both time- and dose-responsive.  The slopes of the  increase
showed  that the higher blood  levels increased  faster at higher concentrations
than at lower concentrations.   Cardiovascular effects resembled  those induced
by epinephrine, except that the heart rate did not change.  DCM increased blood
pressure,  coronary flow,  and  inotropic action, and induced arrhythmias.
     Pryor et al.  (1978)  reported an experiment with an open-chest mongrel dog
(size and sex unspecified).   The blood concentration of DCM  after  1.5 minutes
of exposure to vaporized DCM  of unspecified concentration with oxygen supple-
mentation  was 2,320 ug/ml and without oxygen  supplementation was 2,680
A second exposure to DCM was given 1 hour after the first.  After both
005DC4/A                             5-13                              11/14/83

-------
exposures, heart rate slowed,  blood pressure fell,  and aortic blood flow decreased.
During recovery after the first exposure,  tachycardia and hypertension developed.
After the second exposure,  the animal  died.
     Differences in the  above  studies  concerning the effects of DCM on blood
pressure, inotropic action, and induction  of arrhythmias may be due to the use
of different species of animals and different doses of DCM.
     DCM  used  in the  study of cardiopulmonary effects generally was at a very
high level  (>  30,000  ppm;  104 mg/m3)  in a short-term exposure (£ 5 minutes).
Knowledge of effects  of  such  levels would be important  for accidental or sub-
stance abuse exposures  but are otherwise  not particularly useful  for health
assessment.   For such short exposures, thresholds  for effects  are  quite high.
It  is  hard  to  interpret  such  short-term exposures  considering  that asymptotic
blood  level  is  not approached until 1 to 2 hours.  Data on longer duration
acute exposures are not available.
5.2.1.4   Hepatic,  Pancreatic,  and  Renal Effects—Idaassen  and  Plaa  (1966)
injected  Swiss-Webster mice  (25 to 35 g) intraperitoneally  with analytical
grade  DCM.   Doses   of  13,300  mg/kg had no  effect  on Bromsylphalein (BSP)®
retention or on serum glutamic pyruvic transaminase (SGPT) activity 24 hours
after  injection.  No  histopathological change was  seen upon examination of the
1iver.
     In  the dog, however, effects on the liver by  DCM were reported by Klaassen
and  Plaa (1967).   They derived an  ED5Q for SGPT elevation of  798  mg/kg  (the
LDj-n was  1,260  mg/kg).  Histopathology showed moderate neutrophilic infiltration
in  the sinusoids  and portal  areas.  Necrosis was  not seen.  At "near lethal"
doses, there was vacuolization of centrilobular  hepatocytes.
     Reynolds and Yee  (1967) gavaged fasted male Charles River rats (100 to  300 g)
with  doses  of  14C-DCM up  to  2,210  mg/kg  dissolved in mineral  oil.   The  rats
were sacrified  24 hours  later.  No  liver necrosis  was seen and no  change in  glu-
cose-6-phosphatase  activity  was observed.   Labelled  DCM (14C) was found in
hepatic  lipids  and  proteins, minimally in the lipids  and in  high concentrations
in  proteins.   Similar patterns of  incorporation  were  found in  liver microsomes.
     Weinstein  et  al. (1972) exposed female ICR mice (13 to 20 per group,  23
to  27  g) continuously to approximately 4,893 ppm (17,000 mg/m3) vaporized tech-
nical  grade DCM in  Thomas  Domes, at pressures slightly  below ambient.  Tempera-
ture,  humidity, C0?)  and air  flow  were monitored and  controlled.   During a  24-
hour period,  there was a  progressive decrease of  spontaneous  activity.  Body
weight decreased,  liver weight increased absolutely  and as a ratio  to body
005DC4/A                             5-14                               11/14/83

-------
weight, liver triglycerides  increased  (indicating  liver toxicity), glycogen
decreased, and protein synthesis was reduced as shown by reduction 3H-leucine
incorporation.
     Many histological changes appeared.   Fatty  infiltration after 24 hours
involved the entire lobule;  centrilobular hepatocyte nuclei  became smaller and
dense, and ballooning developed.   Staining showed glycogen  reduction.   Progres-
sive changes began  after  12 hours  of exposure.   Smooth (SER) and rough (RER)
endoplasmic reticulum showed changes:   polysomes  broke down, ribosomal  particles
detached  from RER  in the centrilobular cells, RER  membranes broke up into
vesicles, pernuclear cisternae dilated, and lipid droplets  increased.   Mitochon-
dria,  however, were  unaffected.  The authors concluded that  the  pattern  of
liver  damage  was similar to  that  following carbon tetrachloride exposure.
     Morris et al.  (1979) exposed male Hartley guinea pigs  weighing 500 to 750
g for  6  hours  to a 5,181 ppm  (18,000  mg/m3) vapor concentration of reagent
grade DCM.  Animals were sacrified  immediately after exposure and liver samples
were taken.  Liver  and serum were analyzed  for triglycerides, which increased
markedly in the former and decreased in the latter.   Liver  phospholipids showed
no change, neither did serum-free fatty acid.  Liver slices were incubated with
14C-palmitic acid.   The  uptake of  14C was similar to controls.   Uptake after
14C-leucine did not differ between  controls and exposed samples.
     Differences in protein  syntheis in the  above two studies in protein  syn-
thesis may be due  to different periods of  exposure and the  use of different
animal species.
     Harms et al.  (1976) cannulated the bile duct of male Sprague-Dawley  rats
(350 to 450 g) after pretreatment with intraperitoneal injections of 670 mg/kg
DCM dissolved in corn oil.   Tritiated 3H-insulin was instilled into the duct
24 hours later.   DCM seemed to induce an increase in pancreatic bile duct flow,
which was unrelated to any observed hepatic effect.   Hamada and Peterson (1977),
in a follow-up study, intraperitoneally injected  860 mg/kg  DCM (dissolved in corn
oil) in male Sprague-Dawley rats  (280 to 320 g).   The bile  duct was cannulated,
and bile duct pancreatic flow and its contents were measured and compared with
controls.  DCM induced increased pancreatic  bile duct flow,  decreased  protein
concentration, and increased chloride,  sodium, and potassium.  Bicarbonate was
unaffected.   There  was no  statistically significant difference in wet weight
of the pancreas or in total  bile  flow;  this may indicate a  reduced hepatic bile
flow.  An experiment with secretin  indicated that these changes were not related
to  that  substance  or, after an ancillary  experiment  with  atropine,  to any
cholinergic effect.
005DC4/A                             5-15                              11/14/83

-------
     Plaa and Larson (1965)  injected 10 male  Swiss  mice  (18  to  30  g)  intraperi-
toneally with 1330 mg/kg of  DCM (source and quality unidentified)  dissolved in
corn oil.  No glycosuria or  proteinuria was detected 24  hours  later.   Two  sur-
viving mice of the 10 that had been injected  with  1,995  mg/kg  DCM  had proteinuria
but not  glycosuria.   No  renal  histopathology was  seen after the lower dose.
Proteinuria following the higher dose suggested that there might have been some
tubular damage.
     Klaassen and  Plaa (1966) also  injected male Swiss-Webster mice  (25 to 35
g) intraperitoneal ly  with 13.3  g/kg analytical grade DCM.   These  mice showed
no glycosuria, proteinuria,  or changes in BSP excretion  24 hours after injection.
Two groups of 10 mice each were gavaged for  3 days  with  5 g/kg 60  percent  ethanol
and equicaloric  solutions of  dextrose.   All  animals were then  injected with
DCM, and urine was collected and analyzed 24  hours  later.  BSP excretion remained
within control limits, indicating the absence of a  renal lesion, but the authors
reported, "A few kidneys exhibited hydropic  degeneration with  minimal necrosis
of the convoluted tubules."
     Kluwe et al. (1982) have investigated the effects of DCM  on renal tubular
cells of  adult  male Fischer 344 rats by injecting DCM (1300 mg/kg),  i.p., in
corn oil.   Renal  proximal  tubular swelling  was observed in  the cortex and in
the outer medulla.  Tubular cell functions measured included organic ion transport,
ability  to  maintain a potassium  gradient, and  the  rate  of oxygen  utilization.
None of  these functions  were  altered by  DCM.   Rats  were  sacrificed 2,  12,  24,
48, or 96 h after  injection and kidneys removed and thin slices prepared.   DCM
was not observed to alter glomerular, distal  tubular, inner  medullary or papil-
lary cell  morphology.   Although DCM did cause an increase in the fraction of
water in renal tissue, the lack of effect on other parameters suggested to the
authors that little or no functional cell disturbance had occurred.
5.2.1.5  Other effects-- Ballantyne et al. (1976) studied the effects of liquid
DCM on various ocular parameters by  instilling 0.1 and 0.01 ml of DCM in rabbits'
eyes.  Lachrymation persisted for a week, inflammation of the lids and conjunc-
tivae  for  2 weeks, conjunctiva! edema for a week,  sloughing  for  3  days,  and
increased  corneal  thickness  for 9 days.   Iritis and keratitis  appeared within
6  hours  and  lasted  for 7 and 14  days,  respectively.  Intraocular pressures  in-
creased  also.   Increased corneal thickness developed in rabbits exposed to DCM
vapor  at  concentrations  of 1,750 and  17,500 mg/m3  (504 and 5,040 ppm).
     Sahu  and Lowther (1981) observed that  inhalation  of DCM by  2-month-old
Sprague-Dawley  rats led to pulmonary  injury,  presumably  rupture  of type  II
005DC4/A                             5-16                               11/14/83

-------
alveolar cell  membranes  and release of cell contents into the airways.  Rats
were exposed to about 4000 ppm (13,896 mg/m3) for 5 h/day, 5 days/wk, for 4 weeks.
Pulmonary secretions were obtained by lung lavage.   The soluble supernatant of
the  lung homogenate  and  the cell-free  lung  lavage were  used  for determination
of  lung  lipid peroxidation.  Lipid  peroxidation was  significantly elevated
(P < 0.05).
5.2.1.6  Summary of  Effects  of Acute Exposure—Of all  the organ systems stu-
died, the central nervous system appears to be affected by DCM at levels rang-
ing  from  500  to 1,000  ppm (1,730 to 3,470  mg/m3).   While only short-term
exposures were used, levels of over 20,000 ppm (69 mg/m3) were required before
cardiac function changes were produced.  Hepatic effects were reported at expo-
sure levels as low as 5,000 ppm (17,300 mg/m3).   While it is difficult to compare
results of injection studies to those of inhalation exposures, it appears that
half the LDbU  is required to produce hepatic or renal changes.  The above com-
parisons are  based  upon  the lowest  level  reported.  The  behavioral  data were,
however, in  two  cases,  for levels of  exposure  at  or below  1,000  ppm.  The
weight-of-evidence for short-term exposures indicates that the CNS is the pri-
mary target organ for DCM.
     The concentrations of DCM necessary to depress cardiac function (25,000 to
50,000 ppm)  in acute experiments  are so  high that chronic long-term  exposures
of humans to  levels  considerably in excess  of  250  ppm (865 mg/m3) would be
unlikely to  have any effect.  However, comparatively  little  research  has been
done on acute DCM effects on laboratory animals.
5.2.2  Chronic Effects
5.2.2.1  CNS Effects—In early studies of DCM chronic exposure (Heppel et al.,
1944; Heppel and Neal,  1944), dogs,  monkeys, rabbits, guinea pigs, and rats were
exposed 4 hours/day, 5 days/week for 8 weeks to 5,000 or 10,000 ppm (17,300 or
34,600 mg/m3) DCM.   Unfortunately, only qualitative observations of behavioral
"symptoms"  were reported.  No symptoms were noted in groups exposed to 5,000 ppm
(17,300 mg/m3).  At 10,000 ppm (34,600 mg/m3) all species were affected but in
different ways.  Dogs became excitable and hyperactive.  Monkeys became pro-
gressively more inactive and, by the end of each daily exposure, lay prostrate
with barely perceptible  respiration.   Rabbits  first were  excitable  and then
became inactive toward  the  end  of each  daily session.  Guinea pigs  and rats
simply became  more  inactive.  All  species appeared well  within 1  hour after
cessation of exposure.   Only monkeys (n = 2) appeared  to  develop  behavioral

005DC4/A                             5-17                              11/14/83

-------
tolerance.   No other  cumulative  effects  were reported over the course of the
8 weeks.
     Weinstein et al.  (1972) exposed female mice to 5,000 ppm DCM continuously
for 7 days.  Only qualitative behavioral observations were reported.  For the
first few hours of exposure, the animals  exhibited an increased activity level
and increased food and water intake.   This  was followed by decreased activity,
"hunched"  posture,  dehydration,  and the  appearance of  roughened  yellow
coat.   By  the  fourth  day of exposure, many of the mice had adapted  so that by
the end of the study they were virtually normal.
     Thomas et al.  (1972) studied activity level  in  mice  continuously exposed
to 25, 100, or 1,000 ppm (87,  347, or 3,470 mg/m3) for 14 weeks.   The lowest expo-
sure level  elevated activity level.   There  was no effect at 100 ppm (347 mg/m3),
and at 1,000 ppm (3470 mg/m3), there was  reduced activity level.   The increased
activity level in the 25 ppm (87 mg/m3)  exposed group was, if replicable, pro-
bably due to the increased sensory stimulation due to the odor of DCM.
     Savolainen et al. (1977)  exposed male  rats to 500 ppm (1,735 mg/m3) DCM, 6
hours/day for 4 days.   Exposed rats engaged in more grooming behavior than con-
trols during the first exposure hour but by the 17th exposure hour (the 3rd day)
exposed rats no longer differed from controls.  Other behaviors were observed,
but by their absence in the results summary,  it may be assumed that only grooming
was affected.   Biochemical  analysis of tissue from the right cerebral hemisphere
showed no  difference  in protein, RNA, or glutathione levels as compared with
levels in  control animals.  Relatively small  increases in acid proteinase and
nonspecific cholinesterase activities were  reported, but the determinations for
treated animals were  made  by  only two assays and, therefore, may be of ques-
tionable significance.
     Since only one study  of  activity level  is available, it is difficult to
conclude much about neurobehavioral effects of chronic exposures.  As in acute
exposure studies, no  other  behaviors, such as sensory, motor response acqui-
sition, or  schedule-controlled  behaviors,  were studied.    Where  effects  were
seen, no follow-up  studies  were  conducted  to  determine the irreversibility  of
effects.
     Suggestions  of species differences in DCM  sensitivity  occurred in the
studies by  Heppel  et al.  (1944)  and  Heppel  and  Neal  (1944).   Based on  these
qualitative observations, one could conjecture that  DCM  behaves in a manner
similar to  other  anesthetics:  there is an  early excitatory phase  followed

005DC4/A                              5-18                              11/14/83

-------
by progressive  debilitation.   If  some species had different sensitivities to
DCM, they  might remain  in the excitatory phase or progress more rapidly into
debi1itation.
     Several reports mentioned an increased tolerance to DCM over the duration
of chronic exposures (Weinstein et al. ,  1972; Savolainen et al. , 1977).   Heppel
et al.  (1944) also  reported monkeys  to have  adapted  somewhat.   Probably  adap-
tation, like sensitivity, is a function of species, exposure level, and duration.
     Because of the paucity of data and the qualitative approaches used, these
conclusions must be regarded as tentative.  Effects have been reported at expo-
sure levels  as  low  as 1,000 ppm (3,470 mg/m3)  but  not  lower.  The  one case  of
increased activity level at 25 ppm (87 mg/m3) must be discounted.
5.2.2.2  Hepatic and Renal Effects—Many of the "high-dose" chronic animal stu-
dies with DCM have revealed a certain degree of liver and kidney involvement as
target organs.   The magnitude of this involvement increases in such a way that
extremely high doses of DCM (depending upon dosage and duration of exposure) can
produce toxic effects upon these organs.   Heppel et al. (1944) observed moderate
centrilobular congestion and fatty degeneration of the liver in dogs and guinea
pigs exposed to 10,000 ppm (34,740 mg/m3), 4 hours/day, 5 days/wk for 8 weeks.
Weinstein et al.  (1972)  reported  identical findings after  exposing mice to
5,000 ppm (17,370 mg/m3) continuously for 7 days.   High mortality was observed
by MacEwen  et al.  (1972) where 14 weeks  of  continuous exposures at  1,000 or
5,000 ppm  (3,480 or  17,400 mg/m3) resulted in  severe toxic  effects  and a  high
degree of mortality in mice,'rats, dogs,  and monkeys.
     More importantly, chronic  mouse studies by Weinstein and Diamond (1972)
and Haun et al.  (1972) have revealed that continuous exposure to even such low
levels of DCM as 100 ppm (347 mg/m3) can effect changes in both liver function
and cell architecture.
     Weinstein and Diamond (1972)  exposed ICR mice (17 to 25 g) continuously for
3 days to 10 weeks to 100 ppm (347 mg/m3) of chemical grade DCM in a Thomas Dome
with specified  dynamic  characteristics  and at 96.66 Pa (725 mm Hg) pressure.
Twelve groups of  16  mice each were used.   Except for one instance at week 2,
body weights were comparable to controls.   Liver weights followed body weights,
and in all  cases, liver to body weight ratios were within control  limits.  Tri-
glycerides increased approximately threefold at week 2, nearly fourfold by week 3,
then decreased to about double at week 4.   Four mice withdrawn at 3 days showed
no abnormalities, but at 7 days, centrilobular fat accumulation was seen, accom-
panied by a  decrease in  liver glycogen.   These abnormalities persisted to the
005DC4/A                             5-19                              11/14/83

-------
termination of the experiment at 10 weeks.   During this  time,  the nuclei  enlarged.
No other histopathology was seen under the  light microscope.   Under the electron
microscope, autophagic vacuoles containing  debris appeared in  the hepatocytes.
The smooth and rough endoplasmic reticulum  showed no changes.
     Haun et al.  (1972) exposed mice,  rats,  dogs, and monkeys  to 25 and 100 ppm
(87 and  347  mg/m3)  reagent grade DCM for continuous  exposure  periods up to
100 days in  Thomas Domes at ambient pressure.  This  is the longest continuous
exposure study reported.   Strain,  sex,  and size of the  test  animals were not
specified.   Exposure to the lower concentrations had no  observable effect,  but
exposure to  the  higher concentrations resulted  in  positive  fat stains and
vacuolization.  The rats showed nonspecific tubular degeneration and regenera-
tion in the kidneys, but no changes in organ-body weight ratios.
     Norpoth et al. (1974), when studying the possibility of enzyme  induction
in inhalation of  hydrocarbon  solvents,  tested DCM.   Male SPF  Wistar rats (80
to 100 g)  were exposed 5  hours/day to vapors containing 0,  500, or 5,000 ppm
(0, 1,737, or 17,370  mg/m3 DCM) of the  hydrocarbon  solvents  for 10 days or
250 ppm (869 mg/m3 DCM) of the solvents for 28 days.   There were 50 animals in
the control  group  and six animals in each  of the exposed groups.  At the end
of these exposure  periods,  the animals were  killed  and  the concentration  of
liver  cytochrome  P4S(j and microsomal aminopyrine  demethylase  activity were
determined.
     Following the 10-day exposure, a significant increase in liver cytochrome
P45u in animals  exposed to 500 ppm (1,737 mg/m3) but not in animals exposed to
5,000 ppm  (17,370 mg/m3) of the compound was reported.  In contrast, aminopyrine
demethylase activity was not elevated in animals exposed to 500 ppm (1,737 mg/m3)
DCM but was substantially elevated in those exposed to 5,000 ppm (17,370 mg/m3)
DCM.  After 28 days of exposure to DCM at 250 ppm (869 mg/m3), no changes were
noted in liver enzymes, liver weight, or histologic appearance.
     These results  indicate  that differential enzymatic  induction  can be pro-
duced  by exposure  to  DCM.   The inverse  dose-response relationship noted for
cytochrome P45u may be due to the combined effect of DCM and CO.  Acclimatiza-
tion  may have occurred by the 28th day of exposure to ameliorate the effects
seen  at  10 days.   In  addition, the doses given  in the second exposure period
(28 days)  were much smaller than during the first period.
      Loyke (1973) induced chronic renal hypertension  in 13 Sprague-Dawley  rats
(100  g,  sex  not specified).  Both poles of one kidney were ligated and the con-
tralateral one was removed.  After maintenance of high systolic  pressure for 3
005DC4/A                             5-20                              11/14/83

-------
months,  11  experimental  rats  ^nd  3  controls  were  injected  subcutaneously with
2  mg/kg  unspecified  DCM biweekly for 15  doses.   Two  hypertensive rats were
used  for positive  controls and 11  normotensive rats  as  naive  controls.  DCM
reduced  the  systolic  blood pressure of the hypertensive rats from 200 to 160
mm  Hg.   The  positive controls remained hypertensive.   DCM was ineffective in
reducing  blood  pressure  in normotensive rats.  No  changes  ascribable  to DCM
were  seen in the liver.
      In  the  Dow Chemical  (1980) chronic  inhalation  study in rats,  exposure to
500,  1500, or 3500 ppm (1,735, 5,205, or 12,145 mg/m3),  for 6 hours/day, 5 days/
week,  and 2  years  resulted in exposure-related  non-neoplastic  hepatic  lesions
in  both  males  and  females.  Grossly, the  effect was most prominent in  females
exposed  to  3500 ppm  and consisted of increased numbers  of dark or pale foci.
The percentages of total rats with any degree of vacuolization were 17 percent,
38  percent,  45  percent,  and 54 percent  in  the males of the  0,  500,  1,500, and
3,500 ppm exposure groups, respectively, and 34 percent,  52 percent, 59 percent,
and 65 percent  in the females, respectively.   The degree of severity tended to
increase with the dose.  Further information relating to protocol and additional
observations are discussed in Section 5.3.3.1.1.
      In the Dow Chemical (1980) 2-year inhalation study  in hamsters at the same
levels described above for rats,  a variety of gross and  histopathologic observa-
tions were  recorded  for hamsters  sacrificed  at 6-, 12-, or 18-month interim
kills.  Histopathologically, exposure-related differences consisted of hamsters
with amyloidosis of the liver, kidney,  adrenals, thyroid, and spleen.   Further
observations are discussed in Section 5.3.3.1.3.
     In the Dow Chemical (1982) 2-year inhalation study  in Sprague-Dawley rats
(Spartan substrain) exposed to 50, 200,  and 500 ppm (174, 696, and 1,735 mg/m3),
no definite exposure-related histopathologic findings were noted.  An exception
was the interim sacrifice of females at 15 months.  Some  had a focus or foci  of
altered liver cells.   There were  significant increases in non-neoplastic liver
lesions (i.e.,  hepatocellular vacuolization and multinucleated hepatocytes)  in
female rats  at 500 ppm (1,935 mg/m3).   Further observations are discussed in Sec-
tion 5.3.3.1.4.
     Histopathologic  changes in liver cells of Fischer 344 rats exposed to vari-
ous levels of  DCM  in  drinking water over  a 2-year exposure period  have been
reported by  the National Coffee Association (1982).   Details are described in
Section 5. 3. 3.1.5.

005DC4/A                             5-21                              11/14/83

-------
5.2.2.3  Morbidity and Mortality--Heppel et al.  (1944) conducted two inhalation
experiments.  One used 1,700 mg/m3 (489 ppm) commercial DCM for 7 hours daily,
5 days/week for 6 months.   The other involved exposure for 4 hours/day, 5 days/
week, for  7 to 8 weeks at a  concentration of 34,000 mg/m3  (9,789 ppm).  Dogs,
rabbits, guinea pigs, and rats were used in each experiment.   Two monkeys were
added to the  experiment  at  the higher  concentration.   Animals  were exposed
together in a single chamber.  Temperature and humidity were uncontrolled.   At
the lower dose, 3 of 14 male guinea pigs died.   They had fatty degeneration of
the liver  and pneumonia.  No other animals' deaths attributed to DCM exposure
were reported.
     At the higher dose,  the dog experiment was  terminated after six exposures
because of  the continuing  Stage II excitement  reactions.  Three rabbits and
one rat died  during  the  course of the experiment.   Each animal  had extensive
pulmonary congestion.  Clinical observations  in the dogs  at  the  lower dose
showed no changes in blood pressure,  blood chemistry, or liver function tests,
and at autopsy, organ weights of liver, kidneys, heart, lungs, and spleen were
similar to  controls.   No  lesions  due to DCM were found.   All animals  showed
clinical effects after exposure to the higher concentration.   Aside from the
responses in  dogs  noted  above,  the other animals showed progressive signs of
depression, usually becoming prostrate at the end of each daily exposure period.
All animals,  including dogs, recovered  rapidly  upon removal from the chamber,
and fed well.  At  autopsy,  dogs and guinea pigs showed fatty degeneration of
the liver.   Pulmonary  congestion  was found in the rabbits.  Monkeys and rats
showed no lesions related to DCM exposure.
5.2.2.4  Summary of Effects of Chronic Exposures—Chronic exposure to high levels
of DCM has  been reported to produce alterations  in behavior and in hepatic and
renal function.  When  deaths occurred, they were frequently due to pulmonary
congestion. Apparently these effects begin  at about 500 to 1,000 ppm with  be-
havioral effects in  evidence early,  followed by changes  in internal organs.
No data are available  to determine if the effects seen were reversible after
cessation of  chronic  exposures.   This issue could best be addressed through
additional  research.
5.3  TERATOGENICITY, MUTAGENICITY, AND CARCINOGENICITY
5.3.1  Teratogenicity, Embryotoxicity, and Reproductive Effects
     It is not possible,  on the basis of limited available data, to define the
full potential of  DCM  to produce  adverse teratogenic  or reproductive effects.

005DC4/A                             5-22                              11/14/83

-------
Human epidemiology  studies  tMt  evaluate the effects of  DCM  on the exposed

population are difficult to conduct.  Each of the available mammalian studies
had methodological drawbacks that do not allow for conclusive evaluation of the
ability of DCM to produce  a teratogenic response over a wide range of doses,

which should  include  doses  high  enough to produce signs of maternal toxicity

and lower doses  that  do not produce this  effect.   Other studies in chicken

embryos have indicated that DCM disrupts embryogenesis in a dose-related manner

(Elovaara et  al.,  1979).   However,  since administration of DCM directly into

the air space  of chicken  embryos is not comparable to administration of dose

to animals with a placenta, interpretation of this result related to the poten-

tial of DCM to cause adverse human reproductive effects is not possible.  Another

preliminary study  in  rats  indicated that adverse behavioral effects may occur
after exposure to low levels  of DCM.   However, additional behavioral studies

have not been conducted to  more fully evaluate this effect.
5.3.1.1  Animal Studies—The following discussion subscribes to basic viewpoints

and definitions  of  the terms  "teratogenic" and "fetotoxic" as  summarized  by

the Office of Pesticides and Toxic Substances (U.S.  EPA, 1980):


     Generally, the term "teratogenic"  is  defined as the tendency to produce
physical and/or functional  defects in offspring J_n 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, reduc-
tions 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
categories.   Rather, the spectrum of adverse embryonic/fetal effects is contin-
uous, and all  deviations  from the normal  must  be  considered as examples of
developmental  toxicity.  Gross morphological  terata represent but one aspect
of this spectrum, and while the significance of such structural  changes is more
readily evaluated,  such effects are not necessarily more  serious  than certain
effects which are ordinarily  classified as fetotoxic—fetal  death  being the
most obvious example.

     In view of the spectrum of effects at issue, the Environmental Protection
Agency (EPA) suggests that  it might be useful  to consider developmental  toxicity
in terms  of three basic subcategories.   The first subcategory would be  embryo
or fetal  lethality.   This  is,  of course, an irreversible  effect  and may occur
with or without  the occurrence of gross  terata.  The  second subcategory would
be teratogenesis  and would  encompass those changes (structural and/or functional)
which are induced prenatally,  and which is 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 carcino-
genicity) which are attributable to i_n utero exposure.  The third category would


005DC4/A                             5-23                              11/14/83

-------
be embryo or fetal  toxicity as comprised of those effects  which are potentially
reversible.   This subcategory would  therefore  include such effects as weight
reductions,  reduction in  the  degree  of skeletal ossification, and delays in
organ maturation.
     Two major problems  with a definitional scheme of this nature must be pointed
out, however.   The first is that the  reversibility of any  phenomenon is extremely
difficult to prove.  An organ such as the  kidney, for example, may be delayed
in development and then  appear to "catch up."   Unless a series of specific kidney
function tests is performed on the neonate, however,  no conclusion may be drawn
concerning permanent organ  functional  changes.   This same uncertainty as to
possible long-lasting after effects  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.

5.3.1.2  Mice--Swiss-Webster  mice were exposed  via  inhalation to 1,250 ppm
(4,350 mg/m3)  of  DCM for  7 hours daily during days 6 through 15 of gestation
(Schwetz et al. ,  1975).   This level  was cited as twice the maximum excursion
limit for industrial exposure.   Two  control groups were similarly exposed to
filtered room  air.   Day 0  of  gestation was  determined when a  vaginal plug was
observed.  Caesarean sectioning of dams was performed on day  18  of gestation.
     Dams were evaluated for body weight gain and various  organ weights.   Mater-
nal COHb  level  determinations were performed on blood samples collected  via
orbital  sinus puncture immediately following the third and tenth (last) exposure.
Following Caesarean sectioning, fetuses were weighed, measured (crown-rump length),
sexed, and examined for external malformation.   One-half of the fetuses in each
litter were examined for  soft-tissue malformations (free-hand sectioning) and
one-half were  examined, following staining, for skeletal  malformations.   One
fetus in  each  litter was  randomly selected and  evaluated  using  histological
techniques following serial sectioning.
     In  this  study,  maternally toxic effects of  DCM exposure were observed,
consisting of a significant increase in body weight,  a significant increase in
absolute liver weight, and significant increases in COHb values with return to
control  levels after 24 hours.  On the basis of the maternal  liver weight obser-
vations,  a  minimal  toxicity may have occurred.   The cause of the increased
maternal weights is  unknown.  Since  only the absolute liver weight was reported
and not  the increase in liver weight per animal, the observed  increased maternal
weight  gain might not  have  been  a result  of gains  in  liver weight alone.
     Of  the 12 litters examined, a statistically significant  number of litters
contained fetuses that had a  single  extra  center of  ossification in the  sternum.

005DC4/A                            5-24                              11/14/83

-------
This  common  variation  in mice is thought  to  reflect the  degree of embryonic
development.   It  is  not known if this observation resulted from an accelera-
tion  in  development  or  was  a  chance  occurrence.   The litters  in exposed  group
were  heavier than control fetuses (5.74 g vs.  5.42 g);  however this may be caused
by  the  average slightly smaller  litter size compared to controls  (10  vs.  12).
The litters in this treatment group also had a lower incidence of delayed ossi-
fication of the sternebrae (17 vs. 23 percent), split sternebrae (8 vs. 18 per-
cent), and  ossification of  the skull  bones (25 vs. 36 percent).   Cleft palate
and "rotated kidney" were observed in two (17 percent)  of the DCM exposed fetuses,
and not  observed  in  any of  the control litters.   Because  of the low incidence
of  these  effects,  these effects  may  reflect  spontaneous  malformation rates.
5.3.1.3  Rats--A study  using the same design as that used for the mice (Section
5.3.1.2) was performed  in  Sprague-Dawley rats (Schwetz et al., 1975).   Rats
inhaled DCM  at 1,250 ppm (4,350  mg/m3) for 7  hours daily  on days  6 through  15
of  gestation,  with  day  0 being  the  day  spermatozoa were  observed  in  vaginal
smears.   Dams were Caesarean sectioned on day 21 of gestation.  All other pro-
cedures were identical   to those performed in the mouse  study with the exception
that  in the rat study,   food consumption was monitored.
      No effect  on maternal  body  weight or  food consumption was observed.  The
average absolute maternal liver weight was significantly increased in comparison
with  the control values but there was not effect on the relative weight of the
liver.  Carboxyhemoglobin values in  the  dams increased significantly during
exposure but returned to control  levels within 24 hours.
      In the 19 litters  evaluated, there was no effect on the average  number of
implantation sites per  litter,  litter size,  number of resorptions, or fetal
sex ratio and  body  weight.   The incidence of dilated renal  pelvis was signi-
ficantly increased,  but this observation might indicate a slight but  reversible
delay in development similar to delays in sternal ossification.   However, since
this  study evaluated only one dosage level, it is not possible to firmly estab-
lish the significance of this effect or its reversibility.
     Hardin and Manson  (1980)  used  Long-Evans rats to  evaluate the effect of
exposure to DCM via  inhalation at 4,500 ppm (15,660 mg/m3) for 6  hours daily,
7 days/week to  determine whether exposures before and during gestation  were
more  detrimental to  the developing  conceptus than exposures before gestation
only.   Minimal  maternal  toxicity, consisting of increased absolute and relative
liver weight and elevated COHb levels, was observed.  The litters of rats exposed

005DC4/A                             5-25                              11/14/83

-------
to DCM during  gestation  also  had lower fetal  body weights than controls.   No
other significant deleterious  effect was observed.
     Bornschein et al.  (1980)  reported on the behavioral  teratogenic effects
in the Long-Evans rats exposed to DCM from the Hardin and  Manson (1980) study.
Ten rats were evaluated for general  activity at 5, 10,  45, and 108 days of age,
avoidance learning at  approximately  4 months  of age, and activity following
avoidance learning  at approximately 5 months  of  age.  Fetuses delivered by
Caesarean section had  lower fetal  body weight than those  that were naturally
delivered.   Treatment-related  effects were reported for animals in the general
activity tests as early  as 10 days  of age (both sexes) and were still demon-
strated in male  rats  at  150 days of  age.  No adverse effects were observed in
growth rate, long-term food and water  consumption, wheel  running activity, or
avoidance learning.
     In the  Bornschein et  al.   (1980)  study, the number of rats per test group
was small, usually one male and one  female per  litter.  Therefore, this study
should be regarded as preliminary, and additional  studies are  needed  to fully
confirm these  effects.   Also,  the entire  field  of  behavioral teratology is in
its early stages of development (Buelke-Sam and Kimmel, 1979), and the signi-
ficance of alterations  in  behavioral effects  to human risk assessment is not
clearly defined.
005DC4/A                             5-26                              11/14/83

-------
 5.3.2   Mutagenlcity
     Dichloromethane has been tested for mutagenic activity in bacteria, yeast,
 insects,  nematodes, mammalian cells in vitro, and rodents.  These studies are
 discussed below and are summarized in Tables 5-4 to 5-9.
 5.3.2.1   Gene Mutations—
 5.3.2.1.1  Bacteria.  There are fourteen reports in the literature concerning
 the mutagenic potential of DCM in bacteria; the Salmonella histidine reversion
 assay was used in all of these studies (Simmon et al.,  1977; Simmon and Kauhanen,
 1978; Kanada and Uyeta, 1978; Jongen et al., 1978; McGregor, 1979; Snow et al.,
 1979; Green, 1980; Rapson et al., 1980; Barber et al.,  1980; Nestmann et al.,
 1980; Green, 1981; Gocke et al., 1981; Nestmann et al., 1981; and Jongen et al.,
 1982).  Kanada and Uyeta (1978) also tested DCM in the  B. subtil is rec assay.
 DCM tested positive in all  studies using Salmonella without or with metabolic
 activation in strains TA100, TA1535, or TA98 when assays were performed in
 sealed gas tight exposure chambers.  Negative responses were reported by Rapson
 et al.  (1980) and Nestman et al. (1980) in  standard assays, but these tests
 are judged to be inadequate because DCM was added directly to the agar medium,
 and no precautions were taken to prevent excessive evaporation and escape of
 the test material.  The tests were carried  out at 37°C, which is  very close to
 the boiling point of DCM (39°C).  It is very likely excessive evaporation
 occurred.  Data were presented in many of the reports,  and a clear dose-related
 response  is apparent for each, with 10-fold or greater  increases  in revertants
 observed at the highest doses compared to negative controls.  The doses employed
 and the responses observed are summarized in Table 5-4.
    The purity of the test material was not given in any report.   Because of
 this, most positive responses must be viewed with caution; they may be caused by
 substances other than DCM.   For instance, formaldehyde, a metabolite of DCM,
could conceivably also form nonenzymatically in the aqueous solutions used for
biological testing by hydrolysis of DCM (March, 1977).   However,  because of the
consistency of positive responses with several  different samples, it is more
likely the DCM itself is mutagenic.  For instance,  in their tests of DCM, Barber
et al.  (1980) used a redistilled sample estimated to be >99.9 percent pure,
                                 5-27

-------
                                                       TABLE 5-4.  MUTAGENICITY TESTING OF OCM  IN BACTERIA
 i
INi
CO
Reference Test System Strain
Simmon et al . Salmonella/S9 TA1535
1977 vapor exposure TA1537
TA1538
TA98
TA100







Simmon and Sa1mone11a/S9 TA100
Kauhanen 1978 vapor exposure






Activation
System
None











Aroclor-1254
Induced rat
liver
mlcrosome
S9 mix



Concentration/Results
(Extrapolated from
Fig. 17) 0, 50,
100, 200, 400, and
800 ul/9 liter
desiccator







0 and 1 ml/9 liter
desiccator for 6.5
and 8 hours





(Extrapolated

from Fig. 17)
TA100
Dose (ul) jtevertants/plate
0
50
100
200
400
800


Met.
»h Act.
6.5
+
8
+



170
210
300
400
650
1350
TA100
revertants

Treated Control
688 133
1344 130
830 174
912 158



Comments
1. Toxicity not reported.
2. Number of revertants
observed for TA100 not
specified numerically.
3. Data not presented for
strains other than TA100.
4. Purity and source of
compound not provided.
5. Positive response.



1. Toxicity not reported.
2. Purity and source of
compound not provided.
3. Used as a positive
control In the testing of
2-chl oroethy 1 -chl oro-
formate.
4. Positive response.

-------
                                                                 TABLE 5-4.  (continued)
 i
INi

Reference
Kanada and
Uyeta 1978





Jongen et al .
1978








Test System Strain
Salmonella/S9 TA98
and TA100
B. subtlUs
rec assay
testing


Salmonella/S9 TA98
vapor exposure TA100







Activation
System
PCB-lnduced rat
liver mlcrosome
S9 mix




Phenobarbltal-
Induced rat
liver mlcrosome
S9 mix






Concentration
Not reported






(ppm x 103)

0
5.7
11.4
14.1
22.8
57.0


Result
DCM reported negative
for both strains 1n
B. subtlUs and positive
for both in S. typhimurlum



TA100* TA98*
+S9 -S9 +S9 -S9
15Z+19 1Z9+1Z Zl+4 19+5
329~37 248+32 54+5 44+8
515+76 407+47 74+4 56+10
757~82 582+56 93+9 66+12
865+82 653+89 123+10 96+11
1201+191 740+94 149+42 110+42


Comments
1. Results summarized
1n abstract form.
2. Positive results
of "Ames" testing
supports reports by
other authors using
same system.
1. Testing conducted
1n gas tight perspex
boxes.
2. Only highest dose
exhibited less than
83% survival.
3. Purity of DCM not
reported.
4. Positive response.
                                                                                                                    (continued on the following page)

-------
                                                                   TABLE 5-4.   (continued)
tri
 i
CO
O
Reference Test System Strain
Snow et al. Salmonella/S9 TA98
1979 vapor exposure TA100

Activation
System
DCM-1nduced Syrian
Golden hamster liver
S9 mlcrosome mix

Dose
(ul /Chamber)
0
100
300
500
1000


+"59
"55
177
463
642
972
Result
TA100*
-S9
63
142
274
468
632

TA98
+S9
38
47
69
92
39


-S9
19
31
46
61
72
Comments
1. Purity of DCM not
reported.
2. No Information about
variability of results.
3. Positive response.

           *Mean calculated from three  plates/dose.
(continued on the following page)

-------
                                                                   TABLE 5-4.  (continued)
en
 i
CO

Reference Test System Strain
McGregor Salmonel1a/S9 TA1535
1979 vapor exposure







Nestmann et Salmonella/S9 TA1535
al. 1980 vapor exposure TA1537
TA1538
TA98
TA100




Activation
System
None . Atmospheric
Theoretical

0
0.5
1.0
2.0
4.0
10
Aroclor-lnduced
rat liver S9










Concentration/Results
Concentration
Actual

nd
0.14
0.33
0.67
1.60
nd









X Plate Concentration
ug

nd
245
600, 595, 530
1400
2425
nd









Revertants

15
20
25
50
75
80










Comments
1. Purity of DCM
not reported.
2. Positive response.






1. Data not presented.
2. Negative response
1n standard test.
3. Positive response
1n gas tight chamber.
Doubling 1n revertant
counts for TA1535;
sixfold Increase for
TA100.
                                                                                                                      (continued on the  following  page)

-------
                                                                  TABLE  5-4.   (continued)
       Reference
Test System
Strain
Activation
System
Dose
Result
Comments
       Green  1980    Salmonel1a/S9   TA1535
                     vapor exposure  TA100
                          Rat liver fractions
en
 i
CO
ro
                                          Dose
                                       (% in  air)
                                         0
                                         1.4
                                         2.8
                                         5.5
                                         8.3
                                            TA 100
                                         +S9     -S9
                                        "69+3     :
                                        283+10  267+20
                                        506T27
                                        825~34
                                                                    1050+88
                   462+28
                   872727
                   997+88
                     1.  Preliminary results
                     presented in abstract form.
                     2.  Metabolic studies
                     conducted in rat tissue and
                     TA100.  Similar metabolism
                     1n both systems.  Radiolabel
                     reported to bind to bacterial
                     DNA but not to rat liver DNA.
                     3.  Purity of DCM not
                     reported.
                     4.  Positive response.
                     Author thinks this is due to
                     close proximity of cyto-
                     plasmic enzymes and inter-
                     mediates to DNA in bacteria,
                     and that negative responses
                     would be obtained in higher
                     organisms.  Positive
                     responses in other tests
                     argue against this.  See
                     discussion in text.
                                                                                                                    (continued  on  the  following  page)

-------
TABLE  5-4.  (continued)
Revertants/Plate
TA1535
Reference Test System ppm Vapor

en
i
oo
OJ






Barber et al . Salmonella/S9
1980 vapor exposure
3

7

9

10

0

.600

,200

.100

,900

umoles/plate*
0

38

76

96

115

-S9
23

40

59

78

64

+S9
28

36

51

78

50

TA98
-S9
23

259

441

459

741

+S9
39

288

297

322

479

TA100
-S9
254

752

1440

2640

3060

+59
264

1152

960

1096

3240

Comments
1. Tested redistilled sample of DCM
> 99. 91 pure.
2. Revertants/nmole at highest dose
for TA1535, TA98, and TA100 were
0.0006, 0.006, and 0.03, respectively.
3. Data shown for testing In gas
tight chamber.
4. Negative response 1n standard
test; positive response in gas tight
chamber.
                                                   (continued  on the  following  page)

-------
                                                                  TABLE 5-4.  (continued)
CO
-pi
Activation
Reference Test System Strain System
Concentration/Result
Nestmann et Salmonella/S9 TA1535 Aroclor 1254-
al. 1981 vapor exposure TA1537 Induced rat
TA1538 liver S9
TA98
TA100
Material

Type Weight (mg)
Added
Paint 0
remover 203
370
790
1435
Mix 0
(90:5:5 0.1
v/v/v DCM/ 0.2
methanol/ 0.4
ethanol) 0.8
4NQO 0.001
(Average
values from
triplicate
plates In 4
experiments)
Vaporized
„
144
241
469
903
(ml)
0.1
0.2
0.4
0.8








his* Revertants/Platea

TA1535
16
22
14
23
31
13
15
24
25
34
28






TA100
144
310
433
563
785
154
268
401
789
1084
878






TA98
25
31
42
76
60
32
43
73
138
164
162





Comments
1. Levels
of DCM in
chambers related din
mutatlonal dose-effei
three paint removers
2. Data shown for 01
remover only. Other
similar response.
3. Purity of DCM noi
4. Positive responsi
removers likely due 1
Exposure Level (mg/1 )

Time
DCM
Max 5h
Averaged0 Calculated0
...
12.7
21.9
40.1
80.2

12.2
26.9
50.6
94.1







15.5
25.5
49.9
95.4

12.8
25.3
50.8
101.0








Measured

13.0
23.0
45.0
86.0

11.5
27.5
50.0
94.5






Methanol



<0.5
<0.5
0.7
1.9

<0.5
<0.5
0.9
2.6






exposure
sctly to the
ct curves of
•
rte paint
two gave
t reported.
; for paint
to DCM.
Ethanol d



...
...
...
—

<0.9
0.7
1.1
2.7






             concentration  against  time.   (c).   Calculated
             in  91  chamber,   (d).   Maximum measured.
 ,.  Determined from an area under curve for
rom amount vaporized assuming only DCM vaporized,
                                                                                                                   (continued on the following page)

-------
                                                                   TABLE  5-4.   (continued)
CO
CJ-,
Reference Test System Strain
Green Salmonella/S9 TA100
1981 vapor exposure











Gocke et Salmonella/39 TA1535
al. 1981 vapor exposure TA1537
TA1538
TA98
TA100


Jongen et Salmonella/S9 TA100
al . 1982 vapor exposure




Activation
System
Aroclor-1254
Induced rat liver
S9, microsomes,
and cytosol









Aroclor-1254
rat liver S9





Aroclor-1254
rat liver S9,
microsomes,
and cytosol


Dose
Result
Revertants
% Vapor
0
2.8
5.0
8.4







+S9
TOlT
458
700
950







-S9
TTO
386
720
900







Revertants
ul /desiccator
B
125
250
500
750

Activation

S9
Cytosol
Microsomes
+S9
30 + 0
54 + 7
68 + 32
105 ~ 7
203 _+ 32

0
150
150
150
150
-S9
40~T"0
85 + 7
110 + 14
195 + 21
295 _+ 7
% DCM
0.35 0.7
210 350
240 410
220 420
215 380
Comments
1. Bacterial and mammalian meta-
bolism similar.
2. Cytosol and glutathione
catalyze DCM to formaldehyde and
C02- S-chloromethylglutathione
1s a putative Intermediate.
3. DCM converted to carbon monoxide
1n the presence of microsotnes.
Formyl chloride is a putative
Intermediate.
4. Purity of DCM not given.
5. Positive response. See entry
for Green 1980.
1. Spontaneous rervertants for
TA100 too low.
2. No information presented for
toxicity.
3. Purity of DCM not given.
4. Equivocal positive response.


1.4 1. Purity of DCM not given.
550 2. Positive response.
730
810
610

-------
containing only traces of 1,1- and 1,2-dichloroethane,  chloroform,  chloromethane,
and an unidentified C^^Q aliphatic material  (Dr.  E.  Barber,  Eastman Kodak,
personal  communication).   Nestmann et al.  (1981)  reported  that their sample  was
"gas chromatographically  pure," and Gocke  et  al.  (1981)  checked their sample
for the "correct melting  point (sic)  and elementary  analysis."  From this
information we conclude that the consistent positive  responses in  Salmonella
can be attributed to DCM.  Barber et al.  (1980) conducted  their tests in a
chemically inert, closed  incubation system and analyzed the concentrations of
DCM in the vapor-phase head space and in the  aqueous  phase of a test plate by
gas-liquid chromatography (Barber et al.,  1981).   Based on this information, the
mutagenic responses at the highest dose  (i.e., 115 ^moles/plate)  for TA1535,
TA98, and TA100 were 0.0006, 0.006, and 0.03  revertants per umole,  respectively,
indicating DCM is a weak  mutagen for Salmonella under the  conditions of the
test.
    The results discussed above clearly  show  that  DCM is mutagenic  in Salmonella.
However,  questions have been raised about  the applicability of these results
to predicting mutagenicity in other species,  especially mammals.   DCM is metabol-
ized, apparently via mutagenic intermediates, to  CO  and C02 in both rodents  and
humans (see Chapter 4).  CO is produced by an oxidative dechlorination of DCM by
the microsomal 9450 mixed function oxidase system.  Formyl chloride is believed
to be an  intermediate in  this pathway.  A  second  cytosolic glutathione transferase
system dehalogenates DCM  to produce formaldehyde,  which is further  oxidized  to
C02-  This pathway is thought to proceed via  an S-chloromethyl  glutathione
intermediate (Ahmed and Anders, 1978; Kubic and Anders,  1975).  Formyl  chloride
and S-chloromethyl glutathione are highly  reactive alkylating agents.  They  are
highly unstable compounds, but like formaldehyde,  they  are likely mutagens if
they reach DNA.  Salmonella also metabolizes  DCM  to  C02 and CO apparently by
reaction  pathways similar to those occurring  in mammals (Green, 1980, 1981).
Because of the reactivity of formaldehyde, formyl  chloride, and S-chloromethyl
glutathione, and the proximity of bacterial DNA to bacterial  cytoplasmic
enzymes,  it has been hypothesized that these  chemical  substances  are more
effective as mutagens when they formed by  bacterial  metabolism than when they
are formed outside the bacterial  cell by rat  liver fractions  (Green, 1980,
1981).  The basis for this hypothesis is that rat  liver fractions  used for
                                       5-36

-------
metabolic activation have little effect on  increasing  the mutagenicity  of
methylene chloride in the Ames test.   The  implication  is made  that  as organismic
complexity is increased there is less  likelihood  of  DCM causing mutations.   It
is argued that compartmentalization of DMA  into the  nucleus  protects the genetic
material from exposure to the mutagenic metabolites  of DCM  (i.e., they  would
react with other cellular constituents first)  and thus there is little  or no
mutagenic risk.  The positive results  using eukaryotes, discussed in the follow-
ing sections, argue against this hypothesis.
5.3.2.1.2  Yeast.  Callen et al. (1980) studied the  ability  of DCM  obtained
from Fisher Scientific Company (purity not  reported) and six other  halogenated
hydrocarbons to cause gene conversion, mitotic recombination,  and reverse
mutations in Saccharomyces cerevisiae  (Table 5-5).   Strain  D7  log phase cells
were incubated for 1 hour in culture medium containing 0,  104, 157, and 209  mM
DCM.  The precent survival for these doses  were 100, 77, 42, and  <0.1,
respectively.  Due to the toxicity of  the  compound,  the genetic endpoints
were not measured at the highest dose.  The response for the other  doses
(0, 104, and 157 mM) expressed per 10^ survivors  were: gene conversion at
the trp-5 locus (18, 28, and 107); mitotic  recombination for ade-2  (310, 190,
and 4,490); total genetic alterations  for  ade-2  (3,300, 3,900, and  14,000);
and reverse mutations for ilv-1  (2.7,  4.4,  and 5.8).  A greater than twofold
dose-related increase over negative controls was  observed  for each  endpoint
measured.  No exogenously applied metabolic activation was  used in  these
experiments, which indicates that yeast metabolize DCM intracellularly  to
a mutagenic intermediate(s) that reaches nuclear  DMA.  In  another genetic
study employing yeast, Simmon et al. (1977) reported that  DCM (source and
purity not given, but stated to  be the highest available purity)  did not
increase mitotic recombination in strain D3 of Saccharomyces cerevisiae
when cells (1 x 10^) in suspension culture  were exposed for  4 hours at  30°C
(Table 5-5).  The doses used and the actual experimental values obtained for
mitotic recombination were not reported.  The discrepancies  between the work
by Callen et al. (1980) and Simmon et  al.  (1977)  may be due  to a  number of
factors including the different  strains used (D3  vs. D7),  exposure  time differ-
ences (4 hours vs. 1 hour), or differences  in the incubation temperature (30°C
vs. 37°C).  Call en et al. (1980) reported  that an increase  in the treatment
time of D7 cells with DCM from 1 hour  to 4  hours  significantly reduced  the level

                                   5-37

-------
                                                TABLE 5-5.  GENE MUTATIONS AND MITOTIC RECOMBINATION IN YEAST
                                                                                         Response/10^ Survivors
LTl
I
CO
co


Reference
Callen
et al. 1980





Simmon
et al. 1977



Test System
Saccharomyces
cerevlslae





Saccharomyces
cerevlslae
suspension test

Dose
Strain (mM) % Survival
07 0 100
104 77
157 42
209



D3


ade-2
trp-5 Total Genetic 1lv-l
Conversion Recombination Alterations Revertants Comments
18 • 310 3300 2.7 1. Positive response.
28 190 3900 4.4 2. Active metabolites
107 4490 14000 5.8 produced by this system
are made intracellularly
rather than by an
exogenously employed
activation system.
1. Data not provided,
but reported negatlv
for mitotlc recomblna-
                                                                                                                              tion.
                                                                                                                              2.  Strain differences
                                                                                                                              and differences in
                                                                                                                              treatment conditions
                                                                                                                              (i.e., time and temp-
                                                                                                                              erature) may be the
                                                                                                                              cause of differences
                                                                                                                              between this study and
                                                                                                                              that of Callen et al.
                                                                                                                              1980.
                                                                                                                              3.  Cytochrome P450
                                                                                                                              concentration not known.
                                                                                                                              Callen et al. (1980)
                                                                                                                              report different yeast
                                                                                                                              strains have different
                                                                                                                              levels.

-------
of induced mutation.  Other variables,  such as a lower level  of P45Q enzymes in
strain D3, could conceivably account for the discrepancy in the results.   At this
time, DCM is considered to be a positive mutagen in yeast.
5.3.2.1.3  Drosophila.  Two reports are available concerning the ability  of DCM
to induce sex-linked recessive lethal  mutations in Drosophila melanogaster
(Table 5-6).  Abrahamson and Valencia (1980) reported negative results, while a
positive response was reported by Gocke et al. (1981).  Abrahamson and Valencia
(1980) conducted their sex-linked recessive lethal tests using two routes of
administration:  feeding and injection.  Due to the low solubility of DCM in
aqueous solutions, high concentrations of the test substance were not used in
these experiments, and the negative response observed may be due, in part, to
the fact that sufficiently high doses were not tested.  In  the feeding study,
male Canton S flies were placed in culture vials containing glass microfiber
paper soaked with a saturated solution of 1.9 percent DCM in a sugar solution
(224 mM DCM) for 3 days.  (The feeding solution was added twice daily to  compen-
sate for evaporation of the compound).   At this dose, there was no evidence of
toxicity.  After mating, chromosomes from the 14,682 offspring of treated
parents and chromosomes from the 12,450 offspring of concurrent control  parents
were assessed for recessive lethal mutations.   No evidence  of mutagenicity was
observed.  DCM gave a level of 0.204 percent lethal mutations compared to 0.215
percent for controls.  However, because of the volatility and insolubility of
DCM, the actual dosing to the animals may have been much less than expected.
     In the injection study, 0.3 ul of an isotonic solution containing 0.2
percent DCM was administered to male flies.  This exposure  level resulted in 30
percent post-injection mortality.  However, the post-injection mortality  observed
for the controls was not reported.  Because the mortality observed in studies
such as this is due not only to the test chemical administered, but also  to the
damage caused by injection, concurrent negative controls upon which to base
conclusions concerning toxicity of the test chemical are necessary.  After
mating, 8,262 chromosomes from the offspring of treated parents and 8,723
chromosomes from the offspring of control parents were assessed for recessive
lethal  mutations.  No evidence of mutagenicity was observed by this route of
                                     5-39

-------
                                                 TABLE 5-6.  GENE MUTATIONS  IN  MULTICELLULAR EUKARYOTES IN VIVO
CJ-i
 i
-p.
o
Reference Test System
Abrahamson Orospphila
and Valencia sex-linked
1980 recessive
lethal test
Gocke et Drosophila
al. 1981 sex-linked
recessive
lethal test
Strain Chemical Route
FM6 EMS Fed
females, Tr1s Fed
Canton Neg. Controls Fed or Inj
S males DCM Fed
Inj
iFChromosomes
Tested I Lethal s
773 44
2442 35
94491 230
14682 34
8262 18
Base Lethal s/Brood
females, DCM (mM) 1 ^ 3
Berlin 0 19/7130
K males (0.27)
125 16/3632
(0.44)
620 8/1213
(0.66)
Total 24/4845
treated (0.50)
P < 0.05
8/5525 13/3416
(0.14) (0.38)
2/2579 6/1310
(0.08) (0.46)
3/735 5/1005
(0.41) (0.50)
5/3114 11/2315
(0.16) (0.47)
Corrected
Lethal s (X) Comments
5.69 1. No precaution taken
1.43 to design exposure
0.233 chambers to prevent
0.204 evaporation of the
0.157 test compound for
feeding experiment.
2. No concurrent
negative controls
reported for the
Injection experiment.
3. Negative response at
dose tested (224 mM).
1. Positive response for
Brood 1 indicating DCM
is mutagenlc to sperm in
Drospphila.
T. Higher dose used than
in test by Abrahamson and
Valencia (1980).
                                                                                                                       (continued on the following page)

-------
                                                                     TABLE  5-6.   (continued)
tr,
i
Reference
Samolloff
et al.
1980
Test System
Panagrelus
redivlvus
sex-linked
recessive
lethal test
Concentration
(mol/L)
DCM
W*
io-6
io-4
Proflavlne
10-8
io-6
io-4
Mutation Frequency
(Lethal mutations/105 loci)
6.
10.
9.
12.
10.
28.
0
1
8
5
0
6
(Survival
L2
Juveniles
l.OZ
1.02
1.00
Toxlcity
rel. to controls)
L2-L3
Molt
0.99
1.00
1.00
L3-L4
Molt
0.97
0.86
0.88
L4
Adult
Molt
0.4b
0.15
0.17
Comments
1. Equivocal positive re-
sponse.
2. Not dose-related.
3. Some positive controls
gave negative (e.g., EMS)
or only marginally positive
response (e.g., 3-methyl
cholanthrene).
4. Test system not
validated.

-------
administration.   Flies injected  with  0.2  percent DCM had 0.157 percent lethals
compared to 0.206 percent for controls.
     Gocke et al. (1981)  also tested.DCM  (Merck, Darmstadt, FRG, purity  not
given) for its ability to induce sex-linked  recessive  lethal mutations in
Drosophila.  Two  solutions,  125  mM  and 625 mM  in 2  percent DMSO and  5 percent
saccharose, were  fed to wild-type Berlin  K male flies  for an unreported  period
of time.  The highest dose (625  mM)  is reportedly close to the 1050.  These
males were then mated to Base females. Three  broods were scored (i.e.,  off-
spring from virgin females mated to treated  males on days 1 through  3, 4 through
6, and 7 through  10 after exposure).   There  were significantly more  lethals  (24
out of 4,845 chromosomes scored) for brood 1 (0.50  percent lethal  mutations)
compared to the negative controls (19 lethals  out of 7,130 chromosomes,  0.27
percent lethal mutations), P £0.05.   Elevated, but not statistically signifi-
cant increases in lethal  mutations  were noted  in broods 2 (0.16 percent  compared
to 0.14 percent lethals)  and 3 (0.47  percent compared  to 0.39 percent lethals)
of the treated flies compared to the controls.  As  noted in Table  5-6 the
incidence of lethals is dose-related. The incidence of lethals for  the  high
dose in brood 2 is elevated nearly  three-fold  over  the corresponding negative
control value (0.41 vs. 0.14 lethals, respectively), but this may  not be a
significant increase because of the small sample size  (735 vials).  This test
indicates DCM is  mutagenic to sperm in Drosophila  (a multicellular eukaryote).
The discrepancies in results between Abrahamson and Valencia  (1980)  and  Gocke
et al. (1981) may be due to stock specific differences (Canton  S.  vs. Berlin K)
in the metabolic  activation of DCM  or more likely  to the larger doses of DCM
employed by Gocke et al. (1981).
5.3.2.1.4  Nematodes.  In another sex-linked recessive lethal test,  Samoiloff
et al. (1980) tested DCM for its ability  to  mutate  the nematode Panagrellus
redivivus  (Table  5-6).  Individual  females homozygous  for the X-linked mutation
b7 (coiled phenotype in liquid medium) were  grown  for  120 hours  in the  presence
of several concentrations of DCM ranging  from 10~3M to 10~8M.   They were then
washed and mated to S-15 males who  carry  an  X-chromosome crossover suppressor
extending  at  least 15 recombination units to either side of  b7.   One hundred
                                   5-42

-------
 female progeny were collected and mated to wild-type (C-15) males and their
 progeny scored for the presence of the b7 phenotype.  The absence of b7 male
 progeny indicates lethality of the X-chromosome marked with b7 derived from a
 female grown on DCM.  Three replicate experiments were performed.  A non-dose-
 related increase in the level of lethals was observed in the progeny of DCM-
 treated worms compared to the negative controls.  For worms treated with 10-8,
 1CT6, and 10~4M DCM, the corresponding lethal mutations/105 loci  were 6.0, 10.1,
 and 9.8, respectively, compared to an estimated spontaneous mutation frequency of
 2.2 x 10~6 mutations/locus.  Some of the positive controls tested concurrently,
 such as proflavine, yielded a positive response (12.5, 10.0, and  28.6 lethals/105
 loci at 10-8, ID'6, and 10~4M, respectively).  But others, such as aflatoxin B
 and ethyl methanesulfonate (EMS), did not cause an increase in lethal mutations.
 This study suggests that DCM is mutagenic in nematodes, but firm  conclusions
 cannot be made because the assay is not validated and more importantly because
 of the negative responses obtained with some of the positive controls.
 5.3.2.1.5  Mammalian cells in culture.  Jongen et al. (1981) tested DCM for its
 mutagenic potential in several mammalian short-term tests.  Testing for the
 induction of forward mutations at the HGPRT locus will be described here (Table
 5-7).  Testing for the ability of DCM to cause sister chromatid exchanges (SCE),
 unscheduled DMA synthesis (UDS), and inhibition of DNA synthesis  (IDS) will be
 discussed later in the section on other indicators of DNA damage.
     In their testing of the ability of DCM to cause forward mutations, Jongen
 et al. (1981) incubated log phase CHO and V79 cells with 1, 2, 3, 4, and 5
 percent DCM or 1,  2, 3, and 4 percent DCM, respectively, at 37°C  for 1 hour in a
 closed glass container without exogenous S9 mix.  DCM was obtained from Merck
 (analytical  grade).  The cells were exposed to gaseous DCM and then DCM in
 solution for 15-minute intervals each by alternately tilting the  plates then
 placing them horizontally.  After growth to allow for an 8-day (CHO cells) or
6-day (V79 cells)  expression period,  mutant cells were selected in thioguanine-
containing medium.   DCM failed to increase the mutation frequency of either
cell  line at any  dose compared to controls.   However, DCM was not very cytotoxic
to either cell  line.   At the highest dose, survival  decreased only 20 percent.
                                    5-43

-------
TABLE 5-7.   GENE  MUTATIONS IN MAMMALIAN CELLS IN CULTURE
V79
Concentration Mutants/105
Reference
Jongen et
al. 1981







Test System
6-Thioguanine
resistance In
V79 and CHO
cells





(X)
DCM 0
1
2
3
4
5
EMS 0
2
4
Survivors
2
1.8
2
1.7
1.6

2
13
33
r.Hfl
Survival Mutants/105
(%)
100
98
95
85
80




Survivors
1.9
1.8
1.2
0.9
2.1
2.5



Survival
(%)
100
90
85
80
73
76




Comments
1. Equivocal negative response.
2. Highest dose only resulted 1n 20%
decrease 1n survival. Higher doses up
to about 80% toxicity should be tested.






-------
It would be appropriate to repeat the experiment using higher doses of DCM.
EMS yielded a positive, dose-dependent increase in mutation induction in V79
cells, but it was not tested in CHO cells.
     Based on the positive responses in bacteria, fungi,  nematodes, and insects,
DCM is judged to be capable of causing gene mutations.  Metabolic activation  to
highly reactive mutagem'c metabolites apparently accounts for this response,
and although these are thought to be unstable, they seem to be capable of
interacting with genetic material of higher eukaryotes.
5.3.2.2  Chromosomal Aberrations—Three studies on the ability of DCM to cause
chromosomal aberrations were evaluated.  Burek et al.  (1980)  subjected four
groups of 10 Sprague-Dawley albino rats (Spartan substrain, SPF-derived, 5 males
and 5 females) to 0, 500, 1,500, or 3,500 ppm DCM by  inhalation 6 hours/day,
5 days/week for 6 months.  The animals were then sacrificed,  bone marrow cells
were collected, chromosome preparations were made, and slides were coded and
subse quently analyzed.  Two hundred metaphases per animal  were scored and
aberrations were tabulated (See Table 5-8).  No increase in the total frequency
of abnormal cells or in the frequency of any specific  type of aberration was
noted in the treated animals compared to the controls.  There were 1.1 +_ 1.3,
0.6 +_ 0.7, 0.8 +_ 1.2, and 1.1 +_ 0.9 percent cells with chromosome aberrations
in animals treated with 0, 500, 1,500, and 3,500 ppm  DCM, respectively.
     Thilagar and Kumaroo (1983) treated CHO cells grown in either plastic or
glass culture flasks with 0, 2, 5, 10 and in one experiment 15 ul/ml  (i.e., 0,
31, 78, 156, and 234 mM) DCM for 2 hours with or 12 hours without S9 mix derived
from Aroclor-induced rat livers.  DCM was obtained from Fisher Scientific
(certified A.C.S., lot no. 713580).  After the exposure period, the cells were
washed, refed, and allowed to grow before being arrested at metaphase with
colcemid and harvested for chromosome preparation.  Slides were coded and read
"blind;" 100 cells were scored for each dose level (50 cells/duplicate flask).
DCM induced a dose-related increase in chromosome aberrations (see Table 5-8)
ranging from 0.02 aberrations/cell in the negative controls to 1.44 aberrations/
cell  at 15 ul/ml  (234 mM).  The response was greater  in cells treated in glass
culture flasks and was not dependent on the presence  of the exogenous metabolic
activation system.
                                    5-45

-------
                                                        TABLE  5-8.   TESTS  FOR  CHROMOSOMAL  ABERRATIONS
en
 i

Reference
Burek et
al. 1980








Reference
Thi lagan
and
Kumaroo
1983



Route of
Strain/Tissue Exposure
Male and female Inhalation
Sprague-Dawley
rat/bone marrow







Test System Dose RCG*
Cultured CHO DCM (ul/ml)
cells 0 100
2 98.4
5 75.3
10 66.7
TEM
TH? N.D.
Dose
ppm
0
500
1500
3500






Chromatld

2
4
8
12

22
Breaks
Chromatld
0.9 + 0.99
0.5 T 0.71
0.5 T 0.97
0.7 jf 0.48





Breaks
Isochromatld

0
0
14
34

30
Chromosome
0.2 + 0.42
0.2 7 0.42
0.1 ~ 0.32
0.2 jf 0.42






Exchange

0
2
8
10

42
D1 centrlcs
0
0
0
0 0.





Number of
Aberrations/Cell

0.02
0.06
0.34
0.56

0.96
Rings Exchanges Comments
0 0
0 0
0 0.1 + 0.
2 _* 0.42 0





Cells w/
Aberration (%)

2
6
26
38

66
1. 5 animals/sex/
dose.
32 2. 200 cells/
animals.
3. Dose to bone
marrow cells may
have been low.
4. Negative res-
ponse.

Comments
1. Positive response.
2. Four experiments
yielded similar
response.



            *RCG
Relative cell growth.
Reference
Gocke et
al. 1981
Strain/Tissue
Hale and female
NMRI mice /bone
marrow
Route of
Exposure
l.p.
Injection
Dose
ppm
No. injection x
0
2 x 425
2 x 850
2 x 1700
Breaks
Chromatld Chromosome Di Gentries Rings Exchanges
mg/kg Mlcronuleated Polychromatic Erythrocytes (%)
1.9
1.9
3.5
2.8
Comments
1. Suggestive pos-
itive response;
authors Indicate
negative response.
2. Dose to bone
marrow cells may
have been low.

-------
     Gocke et al. (1981) assessed the ability of DCM (Merck, Darmstadt;  purity
not given) to cause micronuclei in polychromatic erythrocytes (PCE).   Two male
and two female NMRI mice were used for each of three dose levels (425, 850, and
1,700 mg/kg/intraperitoneal injection).  The highest dose approximated the 1059
for mice.  Intraperitoneal injections of each dose were given at 0 and 24 hours,
the animals were sacrificed at 30 hours, bone marrow smears were made, and
1,000 PCEs per animal were scored for the presence of micronuclei.  An increase
in PCEs with micronuclei was observed at the two highest doses, but the response
was not dose-related and was not double the control  value.  Thus, the results
are considered suggestive of a positive response but are not conclusive.
(There were 1.9 percent micronuclei in the untreated controls compared to 3.5
percent micronuclei in the animals receiving two injections of 850 mg/kg, and 2.8
percent micronuclei at the highest dose).
     Based on the positive response reported by Thilagar and Kumaroo (1983), DCM
is tentatively judged to be capable of causing chromosomal aberrations.   The
negative responses reported by Burek et al. (1980) and Gocke et al. (1981) are
not inconsistent with these results.  Thilagar and Kumaroo (1983) exposed
mammalian cells in culture to DCM.  In the studies by Burek et al. (1980) and
Gocke et al. (1981), exposure occurred in vivo.  The dosage received in the bone
marrow cells may have been insufficient to yield a detectable positive response.
5.3.2.3  Other Indicators of DNA Damage—
5.3.2.3.1  Sister chromatid exchange (SCE).  Two papers have been published on
the ability of DCM to induce SCEs (Table 5-9).  Jongen et al. (1981)  tested the
ability of 0.5, 1.0, 2.0, 3.0, and 4 percent DCM (i.e., 58, 118, 235, 353, and
471 mM) to induce SCEs in Y79 cells.  Log phase cells were incubated at 37°C
for 1 hour in a closed glass container.  The cells were exposed to DCM in the
gaseous phase and in the medium by tilting the plates for 15 minutes, then
placing them horizontally.  The experiment was conducted seven times  and each
yielded a dose-related increase in SCEs/cell, which  approached but did not
exceed a twofold increase above the control  level.  Taken together, these
results are statistically significant (P < 0.001).  Increasing the exposure
time to 2 hours or 4 hours or using S9 from rat liver did not alter the  shape
                                  5-47

-------
of the dose-response curve,  which  plateaued  at  1  percent DCM.  The authors
suggest that this phenomenon is  due  to  a  saturation  of  the metabolic activation
system of V79 cells.
     Thilagar and Kumaroo (1983) exposed  CHO cells to 0, 2, 5, 10, and  in one
experiment 15 ul  DCM/ml  of medium  (0, 31,  78, 156, and  234 mM DCM) for  2 hours
with and 24 hours without metabolic  activation.   The cells were  grown for 24
hours in BrdUrd followed by  a mitotic shake  off,  fixing, and staining by a
fluorescence-plus-Giemsa technique,  and then the  coded  slides were scored
"blind." Slight dose-related elevations in SCE  values were noted (see Table 5-9),
but they never exceeded  a 50 percent increase at  the highest dose.  The authors
judged their test to be  negative,  but the results are consistent with those
reported by Jongen et al. (1981) discussed above.  Thus, DCM is  judged  to be
capable of causing DMA damage resulting in SCE.
5.3.2.3.2  DNA repair assays.  In  their study of  the genotoxic potential of
DCM, Jongen et al. (1981) also measured UDS  and IDS  in  V79 cells and primary
human fibroblasts (AH cells). These experiments  were conducted  by exposing 105
cells attached to glass  coverslips (UDS assay)  or to glass petri  plates (IDS
assay) to 0.5, 1.0, 2.0, 3.0, and  5.0 percent DCM (58,  118, 235,  353, and 471 mM,
respectively) without metabolic  activation.   UDS  experiments were done  in
duplicate, and at least  25 nuclei  of non-S phase  cells  were scored for  the
number of silver grains/nucleus  at each dose level.  DCM had no  detectable
effect on UDS in either  cell line.  In  the IDS  assays,  the relative rate of DNA
synthesis was determined radioisotopically immediately  after DCM exposure and
0.5, 1.5, and 3.5 hours  later.  The  average  of  duplicate samples revealed that
DCM inhibited DNA synthesis  in V79 and  AH cells at all  dose levels compared to
controls but that synthesis  recovered with time after exposure in all cases.
This is unlike the persistent inhibition  of  DNA synthesis by the positive
control 4-nitroquinoline-l-oxide.   The  authors  conclude that DCM was not induc-
ing genetic damage in cells  but  was  inhibiting  DNA synthesis by  an effect on
cell metabolism.
     Perocco and Prodi (1981) also performed a  UDS assay using DCM.  They
collected blood samples  from healthy individuals  for their studies, separated
                                  5-48

-------
TABLE 5-9.  TESTS FOR SISTER-CHROMATIO EXCHANGE
Reference Test System Dose
Results
Comments
SCE/Cell
Jongen et SCE/V79 cells X
al. 1981 In culture DCM

0
Experiment #


0
1

.26 +


0


.02


0


.30
7

+ 0


.03
1.
2.
SCEs
3.
Exposure time 1 hour.
Positive response. Significant Increases 1n
(P < 0.001)
Same type of
experiments 2-6
0.5
1.0
2.0
3.0
4.0
0
0
0

0
.40^
.46 _f
.45^
...
.51 _+
0
0
0

0
.02
.02
.03

.03

0
0
0
0

.47
.51
.58
.61
—
± °
± °
± °
±°

.02
.03
.03
.03
4.




DMSO did not




•
dose-response observed 1n
(data not shown).
Increase Incidence of SCE.




                                                              (continued on the following page)

-------
                             TABLE 5-9.   (continued)
Reference Test System Dose
SCE/Cel 1
(X _+ SO) Range of SCE's MI MI + M2 Comments
Thllagar Cultured CHO DCM (ul/ml)
en and cells 0
1 Kumaroo
o 1983 2

5

10

Trlethylene
10.28 i 3.17

11.36 +. 3.09

12.56 +_ 2.95

12.36 +_ 3.35

melamlne
5-17 0 3 97 1. Marginal, but not significant Increases
1n SCEs.
3-19 0 16 84 2. Three other experiments yielded similar
responses.
7-18 6 54 40 3. Results not Inconsistent with test by
Jongen et al. (1981) where highest dose
7-21 4 56 40 was three times greater (I.e., 471 mM vs.
156 mM).

0.025 ug/ral  47.74 i 4.76
39-61
49   47

-------
the lymphocytes, and cultured 5 x 105 cells in 0.2-ml  medium for 4  hours  at
37°C in the presence or absence of DCM (Carlo Erban, Milan,  Italy or Merck-
Schuchardt, Darmstadt, FRG, 97 to 99 percent pure).  The tests were conducted
both in the presence and in the absence of PCB-induced rat liver S9 mix.   A
comparison was made between treated and untreated cells for scheduled DMA
synthesis (i.e., DMA replication) and UDS.  No difference was noted between the
groups with respect to scheduled DMA synthesis measured as dpm of [3H] deoxythy-
midylic acid (TdR) after 4 hours of culture (2,661 _+ 57 dpm in untreated  cells
compared to 2,356 +_ 111 dpm in cells treated with 5 ul/ml [78 mM] DCM).  Subse-
quently, 2.5, 5, and 10 ul/ml (39, 78, and 156 mM) EDC was added to cells
cultured in 10 mM hydroxyurea to suppress scheduled DMA synthesis.   The amount
of unscheduled DMA synthesis was estimated by measuring dpm from incorporated
[3H]TdR 4 hours later.  At 10 ul/ml DCM, 532 +_ 31 and  537 +_ 39 dpm  were counted
without and with exogenous metabolic activation,  respectively.  Both values
were lower than corresponding negative controls of 715 +_ 24 and 612 +_ 26  dpm,
respectively.  No positive controls were run to ensure that the system was
working properly, although testing of chloromethyl methyl ether (CMME) with
activation resulted in a doubling of dpms over the corresponding negative
control values (1,320 +_ 57 at 5 ul/ml CMME vs. 612 _+ 26 untreated).  The  authors
calculated an effective DNA repair value (r) for each  chemical based on the
control and experimental values with and without metabolic activation.  DCM was
evaluated by the authors as negative in the test, but  they did not  state their
criteria for classifying a chemical as positive.   None of the experimental
values from cells treated with DCM had higher dpm values than the controls.
     Based on these experiments there is no evidence  that DCM specifically in-
hibits DNA synthesis or causes UDS.  However, UDS assays are designed to detect
the occurrence of a specific type of DNA repair (i.e., "long-patch" excision
repair).  Xenobiotics that cause damage that is repaired by another process may
not be readily detected.
5.3.2.4  Summary and Conclusions.  Dichloromethane has been tested  for its
ability to cause gene mutations (in Salmonella, yeast, Drosophila,  Panagrellus,
                                  5-51

-------
and cultured mammalian cells), chromosomal  aberrations  (in  rats,  mice,  and
cultured mammalian cells), and other indicators  of DNA  damage  (sister chromatid
exchange formation [SCE] in cultured cells, unscheduled DNA synthesis,  and
inhibition of DNA synthesis).
     Commercially available samples of DCM  have  been  shown  to  be  mutagem'c  in  a
wide range of organisms, including bacteria (Salmonella), fungi  (Saccharomyces),
nematodes (Panagrellus), and insects (Drosophila).  The responses were  weak under
treatment conditions used and were obtained without the addition  of metabolic
activation systems (e.g., S9 mix).  The data suggest  that DCM  is  metabolized to  a
mutagem'c metabolite(s).  Several  metabolism studies  (including one with  bacteria)
indicate such activation occurs "in vivo."   Negative  results have been  reported
for gene mutation tests in fungi  (Saccharomyces)  and  mammalian cells in culture,
but these may represent false negative results because  of the  treatment condi-
tions used.   DCM has also been reported to  induce chromosomal  aberrations in
cultured mammalian cells but not in bone marrow  cells from  animals exposed  j£
vivo, perhaps because a sufficient dose of  DCM did  not  reach the  bone marrow
cells to cause observable effects  there. DCM causes  a  weak increase in SCEs
but has not been shown to cause UDS or inhibit DNA  synthesis.
     Based on the weight of available evidence,  showing positive  responses  in
four different organisms, DCM is judged to  be capable of causing  gene mutations
with the potential to cause such effects in exposed human cells.   A positive
response in cultured mammalian cells indicates it causes chromosomal aberra-
tions, but additional  testing (e.g., another in  vivo  or in  vitro  chromosomal
aberration assay) is needed to resolve this point.  If  such tests are conducted,
care should be taken to insure that the test cells  are  sufficiently exposed to
DCM.  If such precautions are not  taken, false negative responses may be  ob-
tained.
                                  5-52

-------
5.3.3  Evaluation of the carcinogenicity of PCM
     The purpose of this section is to provide an evaluation of the likelihood
that DCM is a human carcinogen and, on the assumption that it is a human carcin-
ogen, to provide a basis for estimating its public health impact, including a
potency evaluation in relation to other carcinogens.   The evaluation of carcino-
genicity depends heavily on animal  bioassays and epidemiologic evidence.  How-
ever, other factors, including mutagenicity, metabolism (particularly in rela-
tion to interaction with DNA), and pharmacokinetic behavior, are important to
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 epidemi-
ologic evidence, the quantitative aspects of assessment,  and finally, a summary
and conclusions dealing with all of the relevant aspects  of the carcinogenicity
of DCM.  Further, the National Toxicology Program (NTP) rat and mouse gavage
bioassay draft technical report (1982) on DCM was cancelled because of data
discrepancies at their contract laboratory (memo from John A. Moore dated July
25, 1983).
5.3.3.1  Animal Studies--
5.3.3.1.1  Dow Chemical Company (1980) inhalation study in rats.  A total of
1,032 male and female Sprague-Dawley rats (129/sex/exposure concentration) were
exposed by inhalation to DCM at 0,  500, 1,500, or 3,500 ppm for 6 hours/day, 5
days/week (excluding holidays), in a 2-year toxicity  and  oncogenicity study.
Approximately 95 rats/sex/exposure concentration were part of the chronic
toxicity and oncogenicity portion of the study.  This number also included those
animals that died spontaneously, were killed moribund during the study, or were
killed at the end of the 2-year exposure.  The remaining  animals were sacrificed
as part of the cytogenetic studies or for one of the  interim kills at either 6,
12, 15, or 18 months of exposure.  The rats were received at 6 to 7 weeks of
age (males weighed 220 to 250 g; females weighed 170  to 200 g) from Spartan
Research Animals, Inc., Haslett, Michigan, and were individually marked for
identification with metal ear tags.  All rats were maintained on a 12-hour
light/dark cycle.  They were observed daily, including weekends and holidays,
for general  health status and signs of possible toxicity.
     Dichloromethane representative of technical  grade material  was obtained
from Dow Chemical Company, Plaquemine, Louisiana, and was used throughout the
                                    5-53

-------
exposure.  Fourteen different samples  of DCM were  analyzed  during  the  2 years
of animal exposure; each sample showed 99 percent  pure  DCM,  with a few trace
chemical  contaminants that varied slightly from sample  to sample,  as shown  in
Table 5-10.  The concentration of DCM  vapor in  chambers was  considered well
within the range of expected variability.  Hematologic  determinations, serum
clinical  chemistry, urinalysis, bone marrow collection, and  blood  carboxyhemo-
globin (COHb) determination were done  in animals sacrificed  at 6,  12,  15,  and
18 months (interim kills).  Plasma estradiol  determination  was done at the  12-
and 18-month interim kills.  This included samples from six  controls/sex  and
four high exposure animals (3,500 ppm)/sex from the 12-month kill, which  were
pooled together (two animals/sample) to give three control  samples and two
high exposure (3,500 ppm) samples/sex.  Ten individual  samples/sex (not pooled)
from the high exposure and control groups were  also sent from the  18-month  kill.
     All  animals that either died spontaneously, were killed in moribund  con-
dition, or were killed at the interim  or terminal  kills were subjected to
complete gross and microscopic pathological examinations by a veterinary  path-
ologist.   Liver samples for possible electron microscopic evaluation were
collected.
     In females exposed to 3,500 ppm,  there was a statistically significant
increase of mortality from the. 18th through the 24th months that may be exposure-
related.   The remaining treated groups in males or females  did not differ
significantly from the controls (Table 5-11).  There was no exposure-related
difference in body weights of either male or female rats exposed to 500,  1,500,
and 3,500 ppm DCM.
     Although some hematologic values  were increased and others were  decreased,
the mean values were within the normal range of biological  variability.   Serum
glutamic pyruvic transaminase  (SGPT),  blood urea nitrogen (BUN), and  serum
alkaline phosphatase (AP) values were  in the normal range.   It is  noted  that  the
females had  significantly increased (P < 0.025) plasma  estradiol levels  at 18
months, which may be related to the higher incidence of mammary tumors in the
exposed  (3,500 ppm) group.  Urinalysis findings were in the normal range, with
the exception of a few statistically significant values in  specific gravity in
males exposed to 1,500 ppm at  6 months and males and females exposed  to  3,500 ppm
at 12 months.  Rats exposed to 500, 1,500, or 3,500 ppm had elevated  COHb values
but with no  evidence of either dose-response or increased values with  prolonged
exposure.                           5_54

-------
                                 TABLE 5-10.   ANALYTICAL ANALYSIS OF DICHLOROMETHANE
                                             (Dow Chemical Company 1980)
1
Specific gravity 1.320
HC1, ppm 14.2
H20, ppm 207
Nonvolatile
material, ppm ND
Methyl chloride, ppm <1
en
^ Chloroform, ppm 60
en
Vinylidene
chloride, ppm 60
Trans 1,2-dichloro-
ethylene, ppm 560
Cyclohexane, ppm 365
Ethyl chloride, ppm 6
Vinyl chloride, ppm <1
Methyl bromide, ppm 23
Carbon tetrachloride, <2
ppm
2
ND
1.8
560

<7
<4

<27


52

561
385
9
--
--
__

3
ND
2.3
37

<7
<4

399


53

266
247
5
--
--
__

4
ND
2.9
55

<7
4.5

48


90

706
467
11
--
--
__

Sample Number
5 6 7
ND
1.8
52

<7
4.5

48


75

550
365
8
--
--
	

ND
1.1
27

<7
4.5

48


65

550
374
11
--
--
	

ND
2.5
112

<7
4.5

32


60

487
335
5
--
--
__

8
ND
--
--

ND
1

52


66

--
399
2
1
1
1

9
ND
--
--

ND
1

576


72

321
266
2
1
1
15

10
ND
1
324*

ND
1

64


62

653
426
2
1
1
. 1

11
ND
--
--

ND
1

562


78

323
262
2
1
1
13

12
ND
1
264*

ND
1

515


70

318
268
6
1
1
16

13
ND
1
340*

ND
1

547


77

337
288
1.5
1
1
20

14
ND
1
3

NL)
1

460


72

298
242
6
1
1
12


*0riginal analysis was lost; sample was subsequently reanalyzed.
ND = not determined.
-- = not detected.

-------
                    TABLE 5-11.  CUMULATIVE PERCENT MORTALITY OF RATS
                        2-YEAR DICHLOROMETHANE INHALATION STUDY
                              (Dow Chemical  Company 1980)
                        Males
Month of
 study    0 ppm  500 ppm  1,500 ppm  3,500 ppm
              Females
0 ppm  500 ppm  1,500 ppm  3,500 ppm
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
0.8
0.8
0.8
0.8
0.8
0.8
0.9
0.9
1.7
1.7
1.7
2.6
2.7
3.6
10.9
15.2
23.8
32.4
44.2
48.4
56.8
63.2
80.0
85.3
0
0
0
0.8
0.8
1.6
2.6
3.4
3.4
4.3
6.0
6.8
10.9
10.9
11.8
16.2
20.0
36.2
45.3
52.6
56.8
65.3
73.7
85.3
0
0
0
0
0
0.8
0.9
0.9
0.9
1.7
3.4
4.3
5.5
10.0
13.6
20.0
31.4
37.1
49.5
63.2
74.7*
83.2*
89.5
93.7
0
0.8
2.3
3.9
4.7
4.7
6.0
6.0
6.0
6.0
6.8
8.5
9.8
15.2*
21.4
29.0*
33.6
43.0
54.6
58.8
72.2*
80.4*
87.6
92.8
0
0
0.8
0.8
0.8
0.8
0.9
2.6
3.4
5.1
5.1
6.0
7.2
9.0
13.5
14.2
23.6
26.4
36.5
42.7
49.0
62.5
70.8
78.1
0
0
0
0
0.8
0.8
1.7
2.6
3.4
5.1
6.0
6.8
8.2
8.2
10.9
15.2
20.0
29.5
42.1
50.5
58.9
64.2
65.3
74.7
0
0
0.8
0.8
0.8
0.8
1.7
1.7
1.7
2.6
4.3
5.1
8.2
8.2
10.0
13.3
20.0
29.5
46.3
55.8
65.3*
73.7
81.1
86.3
0
0
0
0
0.8
0.8
0.9
0.9
2.6
2.6
4.3
6.0
12.5
12.5
19.6
20.6
29.9
42.1*
57.7*
68.0*
81.4*
86.6*
90.7*
95.9*

  *Significantly  different from  control  by  Fisher's  Exact Test,  P  < 0.05.

    5.3.3.1.1.1   Gross  and histopathologic  observations  of  rats  from  the 6-,  12-,
    15-,  and 18-month  interim  kills.   Numerous  gross and histopathologic observations
    were  recorded for  control  and  DCM-exposed rats at each  time  period, and most
    were  typical  of  spontaneous  or naturally-occurring lesions normally seen  in
    rats  of this  strain.   There  were many palpable masses in males and females.
    Some  palpable masses  appeared  to be  abscesses of the preputial or clitoral
    glands, while others  were  cyst-like  lesions  of the skin.  The  total number of
    masses  in the 3,500 ppm group  of males  was  significantly increased over the
                                       5-56

-------
controls at 15, 18, and 21 months, but not at 23  months.   Female  rats  exposed
to 500, 1,500, and 3,500 ppm showed an exposure-related increase  in total
number of masses.  There was also a trend of increased benign mammary  tumors  in
females exposed to 1,500 and 3,500 ppm.  The total  numbers of animals  with
benign mammary gland tumors were 9/28 (0 ppm), 10/29 (500 ppm),  11/29  (1,500
ppm), and 14/27 (3,500 ppm), whereas the total numbers of benign  mammary gland
tumors were 17/28 (0 ppm), 17/29 (500 ppm), 28/29 (1,500 ppm, P  = 9.23 x 1Q-4),
and 37/27 (3,500 ppm, P = 2.33 x lO"10).  These observations were apparent  only
when the cumulative results of the 6-, 12-, 15-,  and 18-month kills were evaluated.
     There were a few other observations that appeared to reflect exposure-related
lesions.  The liver was the only organ that exhibited definite exposure-related
non-neoplastic effects in both males and females  at all exposure  concentrations.
Grossly, the effect was most prominent in females exposed to 3,500 ppm and  con-
sisted of increased numbers of dark or pale foci.  The control group had an
incidence of 0/28, while the 3,500 ppm DCM-exposed female rats had a significantly
greater number of foci (11/27).  Some rats from the 3,500 ppm exposure group
had mottled livers or had an accentuated lobular  pattern to the  liver  (0/29
control males compared to 6/27 males exposed to 3,500 ppm).  Because of the
limited number of rats at each interim kill, this latter change  may be related
to exposure but may also be due to biological  variability.
     Histologically observed, exposure-related lesions were present in the
livers of both males and females exposed to 500,  1,500, or 3,500  ppm.   In males,
the total number of animals with any degree of vacuolization consistent with
fatty changes were 5/29 (0 ppm), 19/29 (500 ppm,  P = 2.05 x 10'4), 21/29 (1,500
ppm, P = 2.47 x ID"5), and 23/27 (3,500 ppm, P =  2.81 x 10'7).  The livers  of
female rats also had alterations considered to be related to DCM exposure.  The
total numbers of females with any degree of vacuolization consistent with fatty
changes were 13/28 (0 ppm), 20/29 (500 ppm, P  = 7.16 x 10'2), 20/29 (1,500  ppm,
P = 7.16 x 10-2), and 22/27 (3,500 ppm, P = 7.16  x 1Q-3).  Because of  these ef-
fects, it may be considered that this experiment  was performed at the  maximum
tolerated dose (MTD).
5.3.3.1.1.2  Gross and histopathologic observations of rats killed moribund or
dying spontaneously during the study and those from terminal  sacrifice (24  months).
     Non-neoplastic observations--The liver was affected in both  males and  females
exposed to 500, 1,500, or 3,500 ppm.  The percentage of total rats with any de-
gree of vacuolization was 17 percent, 38 percent, 45 percent, and 54 percent  in
                                   5-57

-------
the males of the 0,  500,  1,500  and  3,500  ppm  exposure groups,  respectively, and
34 percent, 52 percent,  59 percent,  and 65  percent  in the  females,  respectively.
Also, the degree of  severity tended to  increase with the dose.  The males  in all
exposure groups had  fewer cases of  grossly  observed mottled and/or  enlarged
adrenals.  These gross alterations  appeared to correspond  to  the  histologically
observed decrease in the number of  cases  of adrenal cortical  necrosis, but the
nodular hyperplasia  incidence (unilateral or  bilateral) was increased:   18/95 (0
ppm), 30/95 (500 ppm,  P  = 3.07  x 10-2), 31/95 (1,500 ppm,  P =  2.2 x 10-2), and
24/97 (3,500 ppm).
     Tumor or tumor-like lesions—The Sprague-Dawley strain of rats used in
this study historically  has had a high  spontaneous  incidence  of benign mammary
tumors.  The incidence varies slightly  from study to study, but normally exceeds
80 percent in females and about 10  percent  in males by  the end of a 2-year study.
The mammary gland tumors have been  classified, based on their predominant  morpho-
logical cellular pattern, as fibromas,  fibroadenomas, or adenomas.
     The benign mammary  tumor response  was  present  in males and,  to a lesser
extent, in females.   There was  a non-statistically  significant increase  in the
number of male rats  with a benign mammary tumor exposed to 3,500  ppm  (14/95
as compared to 7/95, 3/95, and  7/95 in  the  0, 500,  or 1,500 exposure  groups).
There was a slight increase in  the  total  number of  benign  mammary tumors in
males exposed to 0 ppm (8/95),  500  ppm  (6/95), 1,500 ppm  (11/95), or  3,500 ppm
(17/97, P = 4.6 x 10-2).
     The total number of female rats with a benign  mammary tumor  did  not
increase in any exposure group  (0,  500,  1,500, or 3,500 ppm groups  had totals
of 79/96, 81/95, 80/95,  and 83/97 benign  mammary tumors,  respectively).  How-
ever, the total number of benign mammary  tumors increased  in  an exposure-related
manner, with 165/96  in the controls, and  218/95, 245/95,  and  287/97 in the females
exposed to 500, 1,500, and 3,500 ppm, respectively. Expressed another way,  the
average number of benign mammary tumors per tumor-bearing  female  rat  increased
from 1.7 in the control  rats, to 2.3 in rats  exposed to 500 ppm,  to 2.6  in those
exposed to 1,500 ppm, and to 3.0 in those exposed to 3,500 ppm.  This effect is
exposure-related, and a dose-response relationship  was  apparent.  There  was  no
indication of an increased number or incidence of malignant mammary tumors in
either males or females.
                                    5-58

-------
     The number of malignant tumors (Table 5-12) increased in male rats exposed
 to 3,500 ppm.  This increase did not appear to correlate clearly with an in-
 creased number of any one tumor type or location.  However, this observation
 led Dow Chemical Company to re-evaluate the gross and histopathologic data on
 all tumors arising in or around the salivary glands.  Table 5-13 lists the
 specific individual animal data for these salivary gland area tumors, showing
 the palpable mass data, specific histopathologic diagnoses, and the number of
 sarcomas with metastases.  Table 5-14 summarizes the incidence of salivary
 gland region sarcomas in male rats.
     Grossly, these tumors were large (several centimeters in diameter), cystic,
 necrotic, or hemorrhagic.  They appeared to invade all adjacent tissues in the
 neck region, and often completely replaced the normal salivary gland tissue.
 Histologically, all were sarcomas.  They were composed of cells that varied from
 round to spindle-shaped, but that appeared to be of mesenchymal cell origin.
 Mitotic figures were frequently observed, as were necrosis and local invasion
 into adjacent tissues.  Most tumors had remnants of normal salivary acini or
 ducts that were caught up in the cellular proliferation.  Two were relatively
 small masses, and appeared to be arising in the interstitial  and capsular
 tissue of the salivary glands.
     One tumor of this type was found in the controls (1/93)  compared to 0/94
 in the 500 ppm exposure group, 5/91 in males from the 1,500 ppm exposure group,
 and 11/88 in males from the 3,500 ppm exposure group (P = 0.002).  Historically,
 a spontaneous incidence (0 to 2 percent) of this tumor type has been observed in
 Dow's laboratory.  Therefore, the 12.5 percent incidence (11  of the 88 rats,
 Table 5-14) found in the males from the 3,500 ppm group was higher than the
 corresponding controls of this study,  and was higher than expected based on
 historical  control  data for male rats  of this strain.  Also,  the males exposed
 to 1,500 ppm had five of these tumors, which was also slightly higher than
 expected,  but was not statistically signficant.   Therefore, this effort appeared
 to be exposure-related in the males exposed to 3,500 ppm.
     The total  number of male rats with malignant tumors was  similar in the
control, 500 ppm, and 1,500 ppm exposure groups.   Males exposed to 3,500 ppm
had an increase in  this category,  since 69 of the 124 rats had malignant tumors
compared to 57, 59,  and 57 in the 0, 500, and 1,500 ppm exposure groups, respec-
tively.
                                    5-59

-------
                             TABLE 5-12.  SUMMARY OF TOTAL  TUMOR DATA  FOR  RATS ADMINISTERED  DICHLOROMETHANE  FOR 2  YEARS BY  INHALATION
                                                                    (Dow Chemical  Company 1980)
 I
O^
o
Males
Spontaneous



Total number of rats
examined


Total number of rats
with a tumor


Total number of rats
with a benign tumor


Total number of rats
with a malignant tumor



Concentration
In ppm
0
500
1500
3500
0
500
1500
3500
0
500
1500
3500
0
500
1500
3500

Interim
kill
29
29
29
27
9
7
9
11
6
3
8
9
6
4
5
7
deaths and
killed
moribund
81
81
89
90
53.
52
67
68
33
29
42
37
39
45
48
55

Terminal
kill
14
14
6
7
13
13
6
7
11
11
6
5
12
10
4
7

Cumulative
total s
124
124
124
124
75
72
82
86
50
43
56
51
57
59
57
69

Interim
kill
28
29
29
27
14
15
14
14
14
15
14
14
2
3
1
2
Females
Spontaneous
deaths and
killed
moribund
75
71
82
93
67
68
79
91
65
60
73
84
23
28
31
32

Terminal
kill
21
24
13
4
21
24
13
4
21
24
12
4
13
14
7
2

Cumulative
total s
124
124
124
124
102
107
106
109
100
99
99
102
38
45
39
36

-------
  TABLE 5-13.   TIME-TO-TUMOR, PALPABLE MASS, AND HISTOPATHOLOGY  DATA  FOR  SALIVARY  GLAND  REGION  SARCOMAS  IN INDIVIDUAL  MALE  RATS
                                      EXPOSED TO DICHLOROMETHANE  BY INHALATION  FOR 2  YEARS
                                                     (Dow Chemical Company 1980)
cr>
Animal
number
76-3301

76-3733

76-3809


76-3722

76-3743

76-3749

76-3583

76-3698

76-3580
Month of study Size of tumor
Exposure tumor first when first
group (ppm) observed observed (cm)
0

1500

1500


1500

1500

1500

3500

3500

3500
15 5

16 3

Not observed


24 1.5

14 2

18 3

Not observed

14 3

21 2
Month of study Size of tumor
for necropsy at necropsy (cm)
15

17

19


24

14

20

24

15

24
7x6x4

5x4x2

Not detected
(slightly enlarged
salivary gland)
4x3x2

No size given
(6 grams)
5x4x4

Not detected

7x6x3

9 x 6 x 4.5
Evidence of
Histological distant
diagnosis metastases
Subcutaneous - undif-
ferentiated sarcoma
Salivary gland -
malignant schwannoma
Salivary gland -
malignant schwannoma

Salivary gland - undif-
ferentiated schwannoma
Subcutaneous - round
cell sarcoma
Subcutaneous - round
cell sarcoma
Salivary gland -
fibrosarcoma
Salivary gland -
carcinosarcoma
Subcutaneous -
None

None

None


None

None

None

None

Yes

Yes
                                                                                                neurofibrosarcoma

-------
TABLE 5-13.  (continued)
Animal
number
76-3663
76-3682
c^ 76-3578
r-o
76-3608
76-3621
76-3666
76-3671
76-3597
Month of study Size of tumor
Exposure tumor first when first
group (ppm) observed observed (cm)
3500 20 2
3500 21 6
3500 16 8
3500 18 2.5
3500 17 4
3500 15 3
3500 Not Observed
3500 19 6
Month of study
for necropsy
21
21
16
18
18
16
14
19
Evidence of
Size of tumor Histological distant
at necropsy (cm) diagnosis metastases
No size given Subcutaneous -
fibrosarcoma
7x4x4 Subcutaneous -
' undifferentiated
round cell sarcoma
7 x 7 x 3.5 Subcutaneous - undif-
ferentiated sarcoma
6x4x3 Subcutaneous -
pleomorphic sarcoma
5x5x3 Subcutaneous -
pleomorphic sarcoma
8x6x4 Subcutaneous -
pleomorphic sarcoma
3.5 Subcutaneous -
neurofibrosarcoma
7 x 7 x 3.5 Subcutaneous -
Yes
None
None
None
None
Yes
None
None
                                         fibrosarcoma

-------
    TABLE 5-14.  SUMMARY OF TALIVARY GLAND REGION SARCOMA INCIDENCE  IN MALE
             RATS IN A 2-YEAR INHALATION STUDY WITH  DICHLOROMETHANE
                          (Dow Chemical  Company 1980)
Dose
0 ppm
500 ppm
1500 ppm
3500 ppm
Incidence*
1/93 (IX)
0/94 (0%)
5/91 (5.5%)
11/88 (12.5%)
Fisher's exact test
—
—
(P = 0.10, N.S.)
(P = 0.002)

   *Cochran-Armitage test for linear trend, P<0.0001.
    N.S. = Not significant.

5.3.3.1.2  Dow Chemical  Company (1980)  inhalation study in hamsters.   A total
of 866 Golden Syrian hamsters [Ela: Eng (syr) strain;  Engle Laboratory Animals,
Inc., Farmersburg, Indiana] (107 to 109/sex/exposure concentration)  were ex-
posed by inhalation to 0, 500, 1,500, and 3,500 ppm of DCM.  The materials and
methods for experimental design are the same as mentioned previously  in the rat
portion of the Dow study.  The body weights of the hamsters were 61  to 70 grams
when they were received.  The hamsters  were marked by  unique toe clips for
group identification and by ear punch for individual identification  within the
cages.
     The mortality data for males  and females are presented in Table  5-15.
Female hamsters exposed to DCM at  3,500 ppm had a statistically significant
decreased mortality from the 13th  through the 24th month of the study.  Females
exposed to 1,500 ppm also had statistically significant decreased mortality
from the 20th through the 24th month.  This decreased  mortality in females
exposed to 3,500 and 1,500 ppm was considered to be exposure-related.  The
remaining exposure groups of male  (500, 1,500, and 3,500 ppm) and female (500
ppm) hamsters had no exposure-related differences in mortality.  Some hamsters
in each group had alopecia at 5.5  months into the study, but this alopecia was
secondary to a mange mite (Demodex species) infection.  This parasite did not
result in increased mortality or morbidity.  No treatment-related differences
were observed in the body weights  of either males or females exposed  to 500,
1,500, or 3,500 ppm of DCM.

                                    5-63

-------
               TABLE 5-15.   CUMULATIVE  PERCENT MORTALITY  OF  HAMSTERS,
                        2-YEAR DICHLOROMETHANE INHALATION STUDY
                              (Dow Chemical  Company  1980)
                         Males
Month of
 study    0 ppm  500 ppm  1500 ppm  3500  ppm
              Females

0 ppm  500 ppm  1500 ppm  3500 ppm
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
0
0
0
0
0
0
1.0
1.9
1.9
2.9
3.8
5.1*
6.4
10.6
14.9
23.4
24.5
31.9
36.0
37.1
41.6
56.2
61.8
82.0
0
0
0
0
0.9
1.9
1.9
3.8
5.8
5.8
5.8
6.7
10.1
14.1
17.2
18.2
19.2
24.2
29.8
43.6
51.1
61.7
63.8
78.7
0
0
0
0
0
0
1.0
1.9
1.9
2.9
2.9
3.9
10.2
13.3
14.3
17.3
19.4
24.5
31.2
41.9
54.8
68.8
75.3
88.2
0
0
0
0.9
3.7
3.7
5.8
5.8
6.7
9.6
11.5
12.5
17. 2t
24.2
26.3
33.3
34.3
37.4
42.6
55. 3t
60.6
69.1
74.5
85.1
0.9
0.9
1.9
1.9
2.8
3.7
3.9
4.9
4.9
5.8
9.7
13.3*
22.6
26.9
32.3
36.6
41.9
52.7
63.6
71.6
80.7
88.6
94.3
100.0
0
0
0.9
1.9
1.9
2.8
4.9
4.9
5.8
9.7
10.7
11.7
16.3
24.5
26.5
30.6
32.7
38.8
59.1
63.4
70.0
82.8
91.4
95.7
0.9
0.9
1.9
1.9
2.8
3.7
4.8
7.6
8.6
8.6
9.5
13.3
15.0
20.0
24.0
29.0
32.0
40.0
50.5
55. 8t
61.lt
68. 4t
75. 8t
89. 5t
0
0
0.9
0.9
0.9
0.9
2.9
3.9
3.9
4.9
4.9
4.9
7. It
11. 2t
15. 3t
23. 5t
27. 6t
30. 6t
40. 9t
45. 2t
53. 8t
72. Ot
80. 6t
90. 3t

    *Five males and five females died due to food deprivation.   These animals were
     subsequently deleted from mortality calculations.

    tSignificantly different from controls by Fisher's  Exact Test,  P  < 0.05.
         Based on the information available to the Carcinogen Assessment Group (CAG),
    it is very difficult to conclude whether the MTD was used.  Dow Chemical  Company
    has not submitted a 90-day dose-finding study, but a 30-day inhalation study
    has been reported in a letter from Dr. J. Burek to Dr.  D. Singh, dated May 1,
    1981.
                                        5-64

-------
     "The study was conducted prior to the two-year study, but has not been
reported.  CD-I mice, Golden Syrian hamsters, Sprague-Dawley and CDF (F-344)
rats were exposed to 0, 2,500, 5,000, or 8,000 ppm DCM vapor 6 hours/day, 5
days/week, for a total of 20, 19, 20 or 6-2/3 exposures, respectively, in 21-29
days.  Body weight data was obtained throughout the study.  Clinical chemistry
parameters were measured.  All animals underwent gross pathological examination
at the termination of the experiment.  The weights of the liver and kidneys
were recorded from animals in the 0, 2,500, and 5,000 ppm groups, and organ/body
weight ratios were calculated.  Animals exposed to 8,000 ppm DCM exhibited
anesthetic effects, increased blood urea nitrogen levels in Sprague-Dawley
rats, and decreased body weights in rats.  The animals exposed to 5,000 ppm
showed slight anesthesia, decreased body weight in male rats, increased SGPT
values in female mice and Sprague-Dawley rats and increased liver weights in
female mice, hamsters and rats.  Animals exposed to 2,500 ppm DCM appeared to
scratch more than controls and therefore appeared to be affected, but showed no
other- effect attributable to exposure.  The target organ in this study was the
liver.  Because of the results obtained, 8,000 and 5,000 ppm DCM were considered
to be too high a dose level for the two-year study, and 2,500 ppm did not
appear to have produced a severe enough response over the 30-day period.
Therefore, concentrations of 3,500 ppm of DCM was [sic] chosen as the top dose
for the two-year study."
     Dow conducted a subchronic dose-finding study for a period of only 30 days
rather than for the 90-day period that is usual in most animal bioassays.
Further, body weight or mortality rate in the experimental group of hamsters
does not decrease as compared to the controls.  Hematologic determinations,
serum clinical chemistry, urinalysis, bone marrow collection, and blood COHb
determination were done in animals at the 6-, 12-, 18-, and 24-month interim
kills.  No treatment-related effects were observed in any of the parameters
evaluated in male or female hamsters after 6, 12, 18, or 24 months of exposure
to 500, 1,500, or  3,500 ppm of DCM, respectively.
     Carboxyhemoglobin determinations were performed on the blood of male and
female hamsters following 22 months on test.  Males and females exposed to DCM
at 500, 1,500, or 3,500 ppm all had significantly elevated COHb values.  There
was a slight trend in a dose-response relationship in females, since the mean
                                    5-65

-------
COHb value for those exposed to 500 ppm was  23.6  percent,  while  the  values
for females exposed to 1,500 or 3,500 ppm were  30.2  percent  and  34.6 percent,
respectively.  However, an apparent dose-response relationship was not  observed
in males.  Dow Chemical Company indicated that  these data, when  compared to
those in the rat study, suggested that hamsters had  a greater degree of meta-
bolism of DCM to carbon monoxide.  Furthermore, the  apparent dose-response  in
females was surprising.  As a result, additional  hamsters  were exposed  to a
single 6-hour exposure, and their COHb values were determined.   There was no
apparent sex difference, and the dose-response  relationship  observed in females
after 22 months could not be verified.  Since there  was no dose-response rela-
tionship in male and female rats or- male hamsters, and since the female hamsters
exposed to a single 6-hour exposure did not  show  a dose-response relationship,
the apparent trend for an exposure-related increase  in female hamsters  at 22
months may be a cumulative effect.
5.3.3.1.2.1  Gross and histopathologic observations  of hamsters  from the 6-, 12-,
and 18-month interim kills.  A variety of gross and  histopathologic  observations
were recorded for  hamsters that were sacrificed at the 6-, 12-,  or  18-month
interim kills.  Histopathologically, exposure-related differences were  present,
and consisted of decreased numbers of hamsters  with  amyloidosis  of  the  liver,
kidney, adrenals,  thyroid, and spleen.  A few animals in each group  may or  may
not represent the trend of amyloidosis in males.
5.3.3.1.2.2  Gross and histopathologic observations  of hamsters  killed  moribund
or dying spontaneously during the study and  those from terminal  sacrifice
(24-months).  Neoplastic and non-neoplastic  observations.  Gross and histopath-
ologic examinations were conducted on all hamsters that died or  were killed
moribund during the study and on all surviving  hamsters at the  end  of the study.
The histopathologic observations for males and  females are presented in the Dow
Chemical Company report (1980), tables 124 and  127.   The observations shown
include all the neoplastic and non-neoplastic lesions recorded  for  these hamsters.
Most observations were within the normal or  expected range for  Golden Syrian
hamsters, as indicated by Dow Chemical Company.  The female  hamsters had in-
creased incidences of lymphosarcoma in the experimental group.   The incidences
were 1/96, 6/95, 3/95, and 7/97 (P = 0.033)  in  the 0, 500, 1,500, and 3,500 ppm
groups, respectively (letter from Hugh Farber,  Dow Chemical  Company, to EPA,
                                    5-66

-------
 dated April 14, 1981).  A re-evaluation of lymphosarcoma data of female
 hamsters by the CAG resulted in the following incidences:  1/91, 6/92, 3/91,
 and 7/91 (P = 0.032) in the 0, 500, 1,500, and 3,500 ppm groups, respectively.
 The differences between the denominators above reflect the CAG's use of
 actual numbers of animals examined (which did not include animals that were
 severely cannibalized, autolyzed, or missing), whereas the Dow denominators
 included the total number of animals.  Also, only a small number of mammary
 gland tissues were examined, and no lesions were found.  Table 5-16 summarizes
 the total tumor data.  The total number of hamsters with benign tumors increased
 significantly in females at 3,500 ppm; the total number of hamsters with malignant
 tumors increased significantly in males at 1,500 ppm.
 5.3.3.1.3  Summary of the Dow Chemical Company (1980) rat and hamster inhalation
 studies.  Based on all the data evaluated, the following points are considered
 to be major findings in the rat and hamster studies:
     1.  Male rats exposed to 1,500 or 3,500 ppm appeared to have an increased
 number of sarcomas in the ventral midcervical area near the salivary glands.
 There were 1/93, 0/94, 5/91, and 11/88 (P = 0.002) sarcomas in male rats exposed
 to 0, 500, 1,500, or 3,500 ppm, respectively.  Based on routine sections, special
 stains, and ultrastructural  evaluations, these tumors appeared to be of mesen-
 chymal cell origin; however, a myoepithelial cell origin of these cells could
 not be ruled out.  These tumors had some areas that morphologically resembled
 one cell type (i.e., neurofibrosarcoma, fibrosarcoma), and still other tumors
 had cell types that were undifferentiated or pleomorphic.  In some, one cell
 type was predominant, while in others, areas of all  of the above cell  types
were present,  depending on the area of the tumor examined.  Furthermore, the
origin of each of these tumors remains questionable.  All appeared to be arising
 in the midcervical  region, and all  involved the salivary glands.  Only two
tumors were small  enough to be localized within the salivary gland.  The rest
were larger tumors that clearly involved the salivary glands as well as adjacent
tissues, and could have been growing either into or out of the salivary glands.
However, they  probably arose within the salivary glands based on the two local-
ized small  tumors described above.
     Therefore,  there was an apparent association between the increased incidence
of sarcomas in the salivary  gland region of male rats and prolonged exposure
via inhalation to 1,500 or 3,500 ppm DCM.   There were no salivary gland sarcomas

                                    5-67

-------
               TABLE 5-16.  SUMMARY OF TOTAL TUMOR  DATA  FOR HAMSTERS ADMINISTERED  DICHLOROMETHAHE BY  INHALATION FOR 2 YEARS
                                                         (Dow Chemical Company  1980)









en
i
01
CO











Total number of
hamsters during
this period*

Total number of
hamsters with
a tumor


Total number of
hamsters with a
benign tumor

Total number of
hamsters with a
malignant tumor



Concentration
(ppm)
0
500
1500
3500
0
500
1500
3500

0
500
1500
3500
0
500
1500
3500


Interim
kills
15
10
10
15
4
3
2
4

4
2
2
3
0
1
0
1

Males
Spontaneous
deaths and
killed
moribund
76
74
82
78
18
17
17
18

14
10
6
10
6
8
12
8


Terminal
kill
16
20
11
14
3
9
8
7

2
7
5
6
1
4
3
1


Cumulative
total s
107
104
103
107
25
29
27
29

20
19
13
19
7
13
15t
10


Interim
kills
15
10
10
14
2
0
2
2

2
0
1
1
0
0
1
1

Females
Spontaneous
deaths and
killed
moribund
91
88
81
82
17
20
13
27

11
8
9
22
8
13
4
9


Terminal
kill
0
4
10
9
_
1
4
3

_
1
3
3
_
0
2
0


Cumulative
totals
106
102
101
105
19
21
19
32t

13
9
13
26t
8
13
7
10

*Does not Include hamsters that escaped from  their cages, or hamsters that were severely autolyzed, or
 severely cannibalized.  Also, this total  does  not Include 500 ppm or 1500 ppm male and female hamsters
 from the 6-month Interim kill because no  hlstopathology was done on these animals except for a liver
 special stain (GomoM's Prussian Blue Iron Reaction).

tSlgnlflcantly different from controls when analyzed by Fisher's Exact Test, P<0.05.

-none examined or not applicable.

-------
 in  female rats or in hamsters of either sex.  Further, it will be of interest
 to  find out what kind of lesions are present or absent in the ongoing National
 Toxicology Program inhalation study.
     2.  Male and female rats exposed to DCM had increased numbers of benign
 mammary tumors as compared to control values.  Female rats exposed to 500, 1,500,
 or  3,500 ppm of DCM had increased numbers of benign mammary tumors per tumor-
 bearing rat in comparison with the controls.  The increase was evident in the
 palpable mass data and the gross necropsy findings, which were confirmed by the
 histopathologic examination.  The total  number of female rats with benign
 mammary tumors was not statistically increased in any exposure group (0, 500,
 1,500, or 3,500 ppm groups had a total  of 79, 81, 80, and 83 animals with benign
 mammary tumors, respectively).  Sprague-Dawley rats have very high incidences
 of  spontaneous mammary tumors.  However, the total  number of benign mammary
 tumors has increased in an exposure-related manner with 165/92 in the controls
 and 218/90, 245/92, and 287/95 in the females exposed to 500, 1,500, or 3,500
 ppm, respectively.  Expressed another way, the average number of benign mammary
 tumors per female rat increased from 1.7 in the controls, to 2.3 in those
 exposed to 500 ppm, to 2.6 in those exposed to 1,500^ ppm, and to 3.0 in those
 exposed to 3,500 ppm.  This increase was considered to be exposure-related and
 dose-dependent.
     A mammary tumor response was present in male rats also, but to a lesser
 extent than in females.  The number of rats with benign mammary tumors in males
 exposed to 3,500 ppm increased but this  increase was not statistically signifi-
 cant.  The total number of benign mamary tumors in  males exposed to 1,500 or
 3,500 ppm increased slightly.  As was the case in females, these effects in
males exposed to 1,500 or 3,500 ppm were considered to be exposure-related.
     There were no mammary gland tumors  in male or  female hamsters.  Also only
28/92, 44/93,  30/94,  and 27/93 mammary  gland tissues were examined in the 0,
500, 1,500,  and 3,500 ppm groups, respectively.  Not a single lesion was recog-
nized in the mammary  gland tissues examined.  The CAG feels that a greater
number of hamster mammary gland tissues  should have been examined to better
evaluate the true incidence of mammary  tumors.
     3.  There was an increased incidence of lymphosarcoma in female hamsters.
The incidence  was 1/96, 6/95, 3/95,  and  7/97 (P = 0.033)  in the 0, 500, 1,500,

                                    5-69

-------
and 3,500 ppm groups,  respectively  (letter from Hugh  Farber,  Dow  Chemical  Com-
pany, to EPA, dated April  14,  1981).   A re-evaluation of  lymphosarcoma  data  of
female hamsters by the CAG resulted in the following  incidences:   1/91,  6/92,
3/91, and 7/91 (P = 0.032) in  the 0,  500,  1,500,  and  3,500  ppm  groups,  respectively.
The differences between the denominators above  reflect the  CAG's  use  of  actual
numbers of animals examined (which  did not include  animals  that were  severely
cannibalized, autolyzed, or missing),  whereas,  the  Dow denominators included the
total number of animals.  Dow  Chemical  Company  believed that  the  females exposed
to 3,500 ppm had better survival  (statistically significant)  than the controls
and thereby had a greater chance  to develop these tumors.   After  correction  for
survival (1/39 vs. 7/63) by the CAG.   These data are  not  statistically  signifi-
cant (P = 0.12).
     4.  There appears to be a question as to whether or  not  the  doses  given
the rat and hamster were at or near the MTD.  The body weights  of male  rats
increased particularly toward  the latter part of the  experiment,  whereas the
body weights of female rats were  unaffected in  any  experimental group.   Exposure
to 3,500 ppm resulted  in an increased mortality rate  in female  rats during the
last six months, but the male  rats  were unaffected  at any concentration.  On
the other hand, decreased mortality was observed in female  hamsters exposed  to
1,500 and 3,500 ppm, while mortality  in male hamsters was unaffected  at 500, 1,500,
and 3,500 ppm based on only a  30-day  rat and hamster  inhalation (dose-finding)
study.  Based on this  information,  it is difficult  to judge whether the  animals
were given a dose equal to the MTD.
5.3.3.1.4  Dow Chemical Company (1982) inhalation toxicity  and  oncongenicity
study in rats.  A 2-year inhalation study  of DCM with Sprague Dawley  rats and
Golden Syrian hamsters by Dow Chemical  Company  (1980) has been  described earlier
in this document.  In  that study, animals  were  exposed 6  hours/day, 5 days/week
for 2 years to DCM at  0, 500,  1,500,  and 3,500  ppm.  The  carcinogenic response
was positive in rats but negative in  hamsters.   In  rats,  the  liver appeared  to
be the target organ affected by exposure.   Hepatocellular vacuolization, consis-
tent with fatty change, was observed in male and female rats  inhaling 500,
1,500, or 3,500 ppm DCM.  There was an increased incidence  of multinucleated
hepatocytes upon exposure to 500, 1,500, or 3,500 ppm, and  an increased number
of foci and areas of altered hepatocytes at 3,500 ppm in  female rats.  Benign
                                   5-70

-------
mammary tumors were increased in male rats inhaling 1,500 or 3,500 ppm,  and in
female rats inhaling 500, 1,500, or 3,500 ppm DCM.   Male rats exposed to 1,500
or 3,500 ppm DCM had an increased number of sarcomas in the region of the
salivary gland.  Female rats inhaling 3,500 ppm DCM had an increased mortality
rate.  Carboxyhemogiobin levels in the blood of rats exposed to DCM were higher
than control levels; however, no differences were observed in COHb levels of
animals inhaling DCM at 500 ppm versus animals inhaling it at 3,500 ppm.  The
objective of this second study was to further investigate the toxicity of DCM
at concentrations far below those that may cause saturation of the metabolic
processes in rats.
     A total of 360 male and 492 female Sprague-Dawley rats (Spartan substrain,
6 to 8 weeks old) were used in this study.  Groups of 90 rats/sex were exposed
by inhalation to 0 (control), 50, 200, and 500 ppm DCM (technical grade, lot
#TA 05038, with purity of at least 99.5 percent) 6 hours/day, 5 days/week, for
20 months (males) or 24 months (females).  Occasionally an exposure was  shorter
than 6 hours, due to vapor generation or mechanical problem.  In addition, 30
extra female rats, identified as 500/0, were exposed to 500 ppm DCM for the
first 12 months of the study and were housed as control  rats for the duration
of the study (last 12 months).  Another 30 female rats,  identified as 0/500,
were housed in the same manner as control rats for the first 12 months of the
study and were exposed to 500 ppm DCM for the remaining 12 months of the study.
To determine the rate of DMA synthesis in the liver, 18 female rats were included
in each group.  After 6, 12, 15, and 18 months of exposure, five rats/sex/expo-
sure level were sacrificed.  In addition, five female rats from each of the
500/0 and 0/500 groups were sacrificed at the 18-month interim necropsy.
     Clinical laboratory tests for chemistry, plasma hormone levels, and DNA
were made on interim sacrificed rats.  Gross and microscopic examinations were
made of animals at interim and terminal sacrifice,  as well as of animals dying
spontaneously and those that were killed moribund during the study.
     As reported by the authors, the nominal and analytical concentrations of
DCM in the chambers were in close agreement, indicating no detectable loss or
decomposition of test material during vaporization.  Approximately 2 months
after the initial exposure to DCM, symptoms consistent with sialodacryoadenitis
(SDA virus)  were observed in male and female rats in each experimental  group,

                                    5-71

-------
including control  groups.   The rats  from all  exposure  groups appeared  to be
equally affected,  and the  symptoms were  not  apparent 3 weeks after  the initial
observation.
     No significant difference in body weight gain  was noticed  in male rats,
but the mean  body  weights  of female  rats at  50,  200, or  500 ppm were signifi-
cantly higher throughout the study period in comparison  with controls.  Although
the authors of this study  consider this  to be a  reflection of biological varia-
bility, the CAG considered that the  highest  dose was not the MTD because the
same strain of rat tolerated a dose  of 3,500 ppm by inhalation  in the  previous
study at the  same  laboratory (Dow Chemical Company, 1980).
     No increase in mortality rate from  that of  the control group was  observed
in male or female  rats.  According to the authors,  "...due to the high mortality
rate in all groups of male rats, the terminal  necropsy for male rats occurred
during the 21st month of exposure to methylene chloride  to ensure adequate numbers
of surviving  animals for pathologic  evaluation." This is not consistent with the
findings in Tables 5-17 and 5-18.
     No significant effect on absolute or relative  organ weight was seen in
male or female rats.  Blood COHb levels  were significantly elevated (P < 0.05)
above controls in  all experimental groups of rats.   Incorporation of ^H-thymidine
into hepatic  DNA was unaffected in female rats at 6 and  12 months.   DMA synthesis
in rats was not determined at 15 months, as  originally scheduled  in the protocol,
due to the number of mammary tumors  observed in  female rats at  200  and 500 ppm.
     Results from the 6-,  12-, 15-,  and  18-month interim necropsies showed no
definite exposure-related gross or histopathologic  findings in  male and female
rats from any of the interim sacrifices.  An exception was  the  interim sacrifice
of female rats at 15 months, where  1, 3, 4,  and  5 females inhaling  DCM at  0,
50, 200, or 500 ppm, respectively,  had  a focus or foci of altered cells in the
liver.  This effect was not apparent in  female rats from 6-,  12-, or 18-month
                                                    i
interim sacrifices, nor was it apparent  from rats dying  spontaneously, killed
moribund during the study, or terminally sacrificed.
     There were no significant histological  lesions observed  in other  organs,
with the exception of the liver.  Data  on the liver lesions are given  in  Tables
5-19 and 5-20.  In males, the incidence  of hepatocellular vacuolization increased
slightly (Table 5-19).  The liver lesions in female rats were significantly in-
creased for foci of altered cell, hepatocellular vacuolization, and multinucleated
                                    5-72

-------
             TABLE 5-17.  MONTHLY MORTALITY DATA FOR MALE RATS IN
    A TWO-YEAR DICHLOROMETHANE INHALATION TOXICITY AND ONCOGENICITY STUDY
                         (Dow Chemical Company 1982)
Month
of
study
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
Terminal sacrifice
0 ppm
0*
0
0
0
0
0
0
0
1(1/85)
1(1/85)
4(3/85)
4(3/85)
5(4/80)
15(12/80)
20(16/80)
33(25/75)
37(28/75)
47(35/75)
59(41/70)
70(49/70)
74(52/70)
50 ppm
0
0
1(1/90)
1(1/90)
1(1/90)
1(1/90)
2(2/85)
2(2/85)
4(3/85)
4(3/85)
4(3/85)
5(4/85)
9(7/80)
13(10/80)
18(14/80)
20(15/75)t
37(28/75)
45(34/75)
64(45/70)
70(49/70)
73(51/70)
200 ppm
0
0
0
1(1/90)
1(1/90)
1(1/90)
2(2/85)
4(3/85)
4(3/85)
4(3/85)
5(4/85)
7(6/85)
11(9/80)
14(11/80)
23(18/80)
31(23/75)
37(28/75)
53(40/75)
69(48/70)
79(55/70)
81(57/70)
500 ppm
0
0
0
0
0
0
0
0
1(1/85)
1(1/85)
1(1/85)
1(1/85)
4(3/80)
8(6/80)
13(10/80)
23(17/75)
25(19/75)
41(31/75)
50(35/70)
66(46/70)
73(51/70)
*Percent mortality (number dead/original number of animals minus animals
 sacrificed for an interim necropsy).

tSignificantly different from control value by Fisher's Exact Test.
                                   5-73

-------
           TABLE  5-18.  MONTHLY MORTALITY DATA FOR FEMALE RATS IN
   A  TWO-YEAR  DICHLOROMETHANE INHALATION TOXICITY AND ONCOGENICITY STUDY
                       (Dow Chemical Company 1982)
Month
of
study
1
2
3
4
5
6
7
8
9
0
1
2
3
4
5
6
7
8
9
10
11
12
13
Terminal
sacrifice
(after 24
months)
0 ppm
0*
0
0
0
0
0
0
0
1(1/85)
1(1/85)
2(2/85)
4(3/85)
4(3/80)
5(4/80)
9(7/80)
12(9/75)
15(11/75)
19(14/75)
30(21/70)
36(25/70)
43(30/70)
53(37/70)
60(42/70)



64(45/70)
50 ppm
0
0
0
0
0
0
0
0
0
0
0
2(2/85)
5(4/80)
5(4/80)
9(7/80)
15(11/75)
15(11/75)
17(13/75)
27(19/70)
36(25/70)
46(32/70)
53(37/70)
63(44/70)



76(35/70)
200 ppm
1(1/90)
1(1/90)
1(1/90)
1(1/90)
1(1/90)
1(1/85)
1(1/85)
2(2/85)
2(2/85)
2(2/85)
2(2/85)
2(2/85)
5(4/80)
5(4/80)
10(8/80)
13(10/75)
15(11/75)
21(16/75)
30(21/70)
34(24/70)
49(34/70)
54(38/70)
64(45/70)



67(47/70)
500 ppm
0
0
0
0
0
1(1/90)
1(1/85)
1(1/85)
1(1/85)
1(1/85)
1(1/85)
4(3/85)
6(5/80)
9(7/80)
15(12/80)
17(13/75)
19(14/75)
28(21/75)
37(26/70)
39(27/70)
47(33/70)
50(35/70)
54(38/70)



61(43/70)
0/500
0
0
0
0
0
0
0
0
0
0
0
0
3(1/30)
3(1/30)
3(1/30)
7(2/30)
7(2/30)
17(5/30)
24(6/25)
32(8/25)
36(9/25)
48(12/25)
60(15/25)



72(18/25)
500/0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
Ot
Ot
3(l/30)t
8(l/25)t
8(2/25)t
16(4/25)t
36(9/25)t
48(12/25)t



52(13/25)

*Percent mortality (number dead/original  number of animals minus animals
 sacrificed for an interim necropsy).

tSignificantly different from control value by Fisher's Exact Test.
                                 5-74

-------
             TABLE 5-19.   NON-NEOPLASTIC LIVER LESIONS IN MALE RATS
                          (Dow Chemical  Company 1982)
             Exposure level
                  (ppm)
                    Foci of
                 altered cells
                       Hepatocellular
                       vacuolization
Terminal
 kill
  0
 50
200
500
 1/18
 3/19
 5/13*
 7/19*
3/18
9/19*
7/13*
8/19
Moribund and
 spontaneous
 death
Combined
  0
 50
200
500

  0
 50
200
500
 6/52
 6/51
 6/57
 4/51

 7/70
 9/70
11/70
11/70
19/52
 9/51
14/57
20/51

22/70
18/70
21/70
28/70
   *Fisher's Exact Test, P < 0.05.
                                   5-75

-------
            TABLE  5-20.   NON-NEOPLASTIC LIVER LESIONS IN FEMALE RATS
                          (Dow  Chemical Company 1982)
             Exposure Level
                 (ppm)
   Foci  of
altered cells
Hepatocellular
vacuolization
Multinucleated
  hepatocytes
Termi nal
kill





Moribund and
spontaneous
death





Combined






0
50
200
500
0/500
500/0


0
50
200
500
0/500
500/0
0
50
200
500
0/500
500/0

9/25
5/17
10/22
17/27*
5/7
8/12


12/45
12/53
17/48
14/43
7/18
5/13
21/70
17/70
27/70
.31/70
12/25
13/25

23/25
15/17
21/22
23/27
6/7
12/12


18/45
27/53
23/48
30/43*
9/18
4/13
41/70
42/70
44/70
53/70*
15/25
16/25

4/25
4/17
5/22
16/27*
1/7
6/12*


4/45
2/53
7/48
11/42*
2/18
3/13
8/70
6/70
12/70
27/70*
3/25
9/25*

*Fisher's Exact Test,  P < 0.05.
0/500 = rats exposed to 500 ppm  DCM for first 12 months.
500/0 = rats exposed to 500 ppm  DCM for last 12 months.
hepatocytes at 500 ppm (Table 5-20)  as compared to  controls.   The  significance
of these alterations in the liver is not known.   Further,  the  number of liver
lesions appeared to be increased in  the 500/0 group at terminal  sacrifice,
combined sacrifice, and death as compared to controls, but this  was  not the
case in the 0/500 ppm group.  There  were no significant differences  in any tumor
type for liver, kidney, spleen, brain, salivary gland, lung,  skin, pancreas,
                                     5-76

-------
and mammary gland in male and female rats, with the exception of mammary gland
tumors, which were significantly higher in females (Table 5-21).
     In summary, there were significant increases in non-neoplastic liver lesions
(i.e., hepatocellular vacuolization and multinucleated hepatocytes) in female
rats at 500 ppm.  There was an increase in benign mammary tumors (adenoma,
fibroma, and fibroadenoma) in female rats.  The number of benign mammary
tumors/tumor-bearing rats observed in female rats was 2.0,  2.3,  2.2,  and 2.7
in rats inhaling DCM at 0, 50, 200, and 500 ppm,  respectively.   Female rats of
group 500/0 showed effects that were similar to rats exposed to  500 ppm for 24
months, but the 0/500 group did not differ from the control  group.
          TABLE 5-21.   SUMMARY OF MAMMARY GLAND TUMORS IN  FEMALE  RATS
                          (Dow Chemical  Company 1982)
Exposure
level (ppm)
Rats with a
benign mammary
tumor
(adenoma,
fibroadenoma,
or fibroma)
Total number
of benign
mammary tumors
(adenomas,
fibroadenomas,
and fibromas)
0
50
200
500
0/500
500/0
0
50
200
500
0/500
500/0
Number
of rats
70
70
70
70
25
25
70
70
70
70
25
25
Moribund and Terminal
spontaneous kill
35
43
41
32
17
11
69
97
91
68
35
27
17
15
20
23
6
12*
36
36
44
79
15
33
Cumulative
52
58
61*
55
23
23
105t
133
135
147
50
60

     *Significantly different from control when analyzed by Fisher's Exact
      Test, P < 0.05.
     tData could not be analyzed by Fisher's Exact Test.
     0/500 = rats exposed to 500 ppm DCM for first 12 months.
     500/0 = rats exposed to 500 ppm DCM for last 12 months.
                                    5-77

-------
     In conclusion, the results of this study  offer very  limited  evidence
of the carcinogenicity of DCM.   However,  the highest dose in  this study  is  half
of the lowest doses in both the NTP gavage study  (1982  draft)  and the  ongoing
NTP inhalation study.
5.3.3.1.5  National Coffee Association (1982)  study in  rats.   On  August  11,
1982, Hazelton Laboratories America, Inc., reported on  a  chronic  study in
Fischer 344 rats administered DCM in deionized drinking water for 24 months.
This study was sponsored by the National  Coffee Association  (NCA),  and utilized
1,000 animals in seven different dose groups or regimens  (Table 5-22).  The
actual  mean daily consumption levels of DCM in the  drinking water were similar
to the expected target levels (Table 5-23).
     Interim sacrifices were performed at 26,  52, and 78  weeks of treatment with
5, 10,  or 20 rats/sex/dose group, respectively, with the  exception of  the  animals
in Group 7 (the recovery group, which were only sacrificed with the remaining
terminal animals at 104 weeks).  The effects of compound  administration  were
evaluated using the following criteria:  survival,  body weight gains,  total food
consumption, water consumption, clinical  observations,  ophthalmoscopic findings,
clinical pathology, organ and tissue weights,  and gross and microscopic  pathology.

         TABLE 5-22.  GROUP ASSIGNMENT OF FISCHER 344 RATS ADMINISTERED
           DICHLOROMETHANE IN DEIONIZED DRINKING WATER  FOR 24 MONTHS
                                  (NCA 1982a)
Group Number
1.
2.
3.
4.
5.
6.
7.
Control
Control
Low-dose
Mid- dose 1
Mid-dose 2
High-dose
Hi gh-dose/recovery
Number of Animals
Males
85
50
85
85
85
85
25
Females
85
50
85
85
85
85
25
Target dose
(mg/kg/day)
0
0
5
50
125
250
250
    (78 weeks/26 weeks)

-------
        TABLE 5-23.  MEAN DAILV CONSUMPTION OF DICHLOROMETHANE IN 24-MONTH
          CHRONIC TOXICITY AND ONCOGENICITY STUDY IN FISCHER 344 RATS
                                  (NCA 1982a)
                   Target level
   Group           (mg/kg/day)         Males             Females
3 5
4 50
5 125
6 250
7 250*
5.85
52.28
125.04
235.00
232.13
6.47
58.32
135.59
262.81
268.72

    *The designated recovery group (Group 7) mean Is for the first 78
     weeks only.
     No compound-related findings were reported for either survival,  clinical
observations, opthalmoscopic findings, gross necropsy findings, or organ
weight data.  Throughout the study, small  but significantly decreased body
weight gains and water consumption were reported for both male and female rats
in Groups 5, 6, and 7.  Food consumption also decreased,  but this criteria was
only monitored for the first 13 weeks.  These decreased effects were  attributed
to DCM administration.  Low-magnitude statistical  increases in mean hemotocrit,
hemoglobin, and red blood cell  counts were noted in both  male and female animals
of Groups 4, 5, and 6.  In most cases, these were  within  the range of historical
control  values of Fischer 344 rats.
     Histopathological alterations were described  in the  livers of rats of
both sexes in Groups 4, 5, 6, and 7.   These changes consisted of an increased
incidence of foci/areas of cellular alteration in  Groups  4, 5, 6, and 7 and of
fatty changes in Groups 5 and 6 after 78 and 104 weeks  of treatment.   The
incidence of neoplastic nodules and/or hepatocellular carcinomas in female
Fischer  344 rats (Table 5-24) was derived  from the data presented in  Volumes
I-IV of  the August 11, 1982,  NCA report (NCA 1982a)  and in the Addition to the
                                    5-79

-------
                            TABLE  5-24.   INCIDENCE OF HEPATOCELLULAR TUMORS  IN MALE AND FEMALE FISCHER 344 RATS
                                  ADMINISTERED DICHLOROMETHANE  IN DEIONIZED  DRINKING WATER FOR 104 WEEKS
                                                              (NCA 1982a, b)
Cn
00
o
Diagnosis


Neoplastic nodule (NN)
Hepatocellular carcinoma
(HC)
Combined NN and HC
Total livers examined

Neoplastic nodule (NN)
Hepatocellular carcinoma
(HC)
Combined NN and HC
Total livers examined
Treatment group (mg/kg/day)
0

4(4.7.)
2(2.4)

6(7.1)
85

0(0)
0(0)

0(0)
85
0

5(10)
2(4)

7(14)
50

0(0)
0(0)

0(0)
49
5

2(2.4)
0(0)

2(2.4)
85

1(1.2)
0(0)

1(1.2)
85
50
Male*
3(3.6)
0(0)

3(3.6)
84
Female*
2(2.4)
2(2.4)

4(4. 8)t
83
125

3(3.5)
0(0)

3(3.5)
85

1(1.2)
0(0)

1(1.2).
85
250

1(1.2)
1(1.2)

2(2.4)
85

4(4. 7)t
2(2.4)

6(7. l)t
85
250 (Recovery)

4(16)
0(0)

4(16)
25

2(8.0)
0(0)

2(8. 0)t
25

             *Numbers in parentheses represent percent incidence for the particular lesion.

             tStatistically significant (P < 0.05)  by the Fisher's Exact Test using a combined control  incidence of 0/134.

-------
Final NCA Report of November 5, 1982 (NCA 1982b).  Male rats did not show
an increased incidence of liver tumors in treated animals versus controls
(Table 5-23).  These statistically significant increases in the incidences
of liver tumors in female rats were within the range of historical  control
values at this laboratory (Table 5-25), as presented in a letter from
Dr. John Kirschman of General Foods to Dr. Dharm Singh of the CAG,  dated
February 17, 1983.  Therefore, based on a review of the NCA study,  DCM
administered in deionized water at doses up to 250 mg/kg/day was borderline
for carcinogenicity to Fischer 344 rats.
5.3.3.1.6  National Toxicology Program (1982 draft) gavage study in rats and
mice.  The National Toxicology Program (NTP) conducted a 2-year carcinogenesis
bioassay of food-grade DCM and reported on their results in a draft technical
report dated September 22, 1982 (NTP, 1982).  Dichloromethane was administered
in corn oil by gavage to male and female Fischer 344/N rats and B6C3F1 mice.
The DCM was more than 99.5 percent pure, with vapor phase chromatographic
analysis detecting the presence of vinylidene chloride and trans-dichloroethylene
up to 0.4 percent.  A 13-week dose finding study was conducted to evaluate the
toxicity of the compound.  Based on survival, body weight gain, and histopathol-
ogical examination, doses of 500 and 1,000 mg/kg by gavage were selected for male
and female rats and mice for the 2-year study.
     The NTP announced on July 25, 1983 that the draft NTP report would not be
issued as a final  report due to discrepancies in experimental data  that compro-
mise a clear interpretation.  NTP further allowed that pending the  results of
an in-depth audit, select and relevant information from these gavage studies
might be incorporated into the future draft technical  report for DCM inhalation
studies.
5.3.3.1.7  Theiss  et a!. (1977).  A pulmonary tumor bioassay in mice was reported
by Theiss et al.  (1977).  Groups of 20 male strain A mice were injected intra-
peritoneally three times a week with 0, 160, 400, or 800 mg/kg DCM  for a total
of 16 or 17 injections.   Mice were sacrificed 24 weeks after the first injection,
and the lungs were examined under a dissecting microscope for surface adenomas.
Some adenomas were confirmed by histology.
                                     5-81

-------
             TABLE 5-25.   HISTORICAL  CONTROL  DATA OF  LIVER  NEOPLASIA  IN  FEMALE  FISCHER 344 RATS
                                   AT HAZELTON  LABORATORIES AMERICA,  INC.
                                                 (Kirschman,  1983)


en
i
oo
IN3




LIVER (No.


Neoplastic
% Incidence
Study: A ' B C
Group: 12 11
examined) 57 70 37 40


nodules 54----
8.8 5.7 —
Hepatocellular carcinoma

% Incidence

_.

D E
T T
38 43


1
2.3
1 1
2.6 2.3

F
1
11


1
9.1
1
9.1

G
~~7 T
14 39


4
28.6
1
7.1

H I
r T
40 17


2
11.8
--
__

J
T
13


2
15.4
1
7.8

Study
Total
419


19
4
5
1


s




.5

.2

— = Not reported.

-------
     Tumors were found at all three dose levels; however, due to poor survival
 and  the  small  number of animals, the increase in tumors did not reach statistical
 significance at the two highest dose levels (Table 5-26).  At the lowest dose,
 a  highly  significant increase in the number of tumors was observed (P = 0.013).
 Therefore, this study was marginally positive for carcinogen!city.
             TABLE 5-26.  PULMONARY TUMOR BIOASSAY IN STRAIN A MICE
                        (adapted from Theiss et al. 1977)

Dose
(mg/kg)
0
160
400
800

Total dose
given
0
2,720
6,800
12,800

No. of mice
at beginning
20
20
20
20
No. of mice
exami ned
for tumors
15
18
5
12


Tumors/mouse
0.27
0.94
0.80
0.50


Significance*
—
<0.013
>0.1
>0.1

*The test of significance used is the exact test of ratio of two Poisson parameters.


5.3.3.1.8  Heppel et al. (1944).  Heppel et al.  (1944) exposed dogs, rabbits,
guinea pigs, and rats to DCM by inhalation at levels of 5,000 ppm for 7 hours/day
and 10,000 ppm for 4 hours/day, 5 days/week, for 6 months.  No tumors developed
in any animals.
5.3.3.1.9  McEwen et al. (1972).  McEwen et al.  (1972) exposed dogs to DCM by
inhalation at 500 ppm for 14 weeks;  no tumors were reported, but edema of the
meninges of the brain occurred.  Neither this study nor the Heppel  et al.
(1944) study could have detected a carcinogenic  response because of the shortness
of the observation times and the fact that these studies were not originally
designed to test for carcinogenicity.
                                    5-83

-------
5.3.3.1.10  Other animal  studies in  progress.   The  NTP  sponsored  a 2-year  inhala-
tion study in male and female Fischer 344/N  rats  and  B6C3F1 mice  at  exposure  con-
centrations of 0, 1,000,  2,000,  and  4,000 ppm  for rats  and 0,  2,000,  and 4,000
ppm for mice.  The animals have  been sacrificed (April  1983),  but the pathology
report is not yet available.
     A 2-year study was sponsored by the National Coffee Association  in which
male and female B6C3F1 mice were administered  daily doses of 0, 60,  125, 185,
or 250 mg/kg DCM in drinking water.   The animals  have been sacrificed, and the
results of this study should be  available by the  end  of 1983.
5.3.3.1.11  Cell  transformation  studies.  Price et  al.  (1978)  exposed Fischer
rat embryo cell cultures  (F1706, subculture  108)  to DCM liquid at concentrations
of 1.6 x 102 and 1.6 x 103 yM for 48 hours.  Dichloromethane was  diluted with
growth medium to yield the appropriate doses.   The  DCM  sample, obtained from  the
Fisher Scientific Company, was >_ 99.9 percent  pure.   The cells were  grown  in
Eagles minimum essential  medium  in Earle's salts  supplemented  with 10 percent
fetal  bovine serum, 2 mM  L-glutamine, 0.1 mM nonessential ami no acids, 100 y  g
penicillin, and 100 yg streptomycin/ml.   Quadruplicate  cultures were  treated  at
50 percent confluency with each  dose.  After treatment, cells  were cultured in
growth medium alone at 37°C.   Transformation of cells treated  with either  dose
level  of DCM was observed by 23  and  30 days  of incubation, and was characterized
by progressively growing  foci composed of cells lacking contact inhibition and
orientation.  There was no transformation of cells  grown in medium alone or in
the presence of a 1:1,000 acetone concentration,  even after a  subculture.
Twenty and 27 microscopic foci per three dishes with  the low and  high DCM  doses
respectively, were found  in dishes inoculated  with  50,000 cells from cultures
treated four subcultures  earlier and held for  4 weeks at 37°C  in  a humidified
C02 incubator before staining.
     Subcutaneous injection of cells treated with 1.6 x 102 yM DCM five subcul-
tures earlier produced local  fibrosarcomas in  5/5 newborn Fischer 344 rats within
60 days following treatment.   The ability of cells  grown in growth medium  alone
to induce local fibrosarcomas was not determined; however, negative  responses
were obtained with cells  grown in the presence of a 1:1,000 concentration  of
acetone.  Exposure of cells to 3.7 x 10'1 yM 3-methylcholanthrene produced 124
microscopic foci  per three dishes in the inoculation  test described  above  by  37
                                   5-84

-------
 days of incubation, and local fibrosarcomas in 12/12 rats by 27 days following
 subcutaneous injection of cells.  The exposure of 3-methylcholanthrene was
 attained by initial dilution in acetone to 1 mg/ml  followed by further dilution
 in growth medium to 0.1 ug/ml (personal communication from Dr. P. J. Price).
     Dr. Price wrote a letter to the CAG dated Nov. 14, 1980, saying that "the
 analysis of methylene chloride showed a purity of 99.9 percent.  The original
 study was done in quadruplicate and in each case the Fischer rat cells were
 transformed.  Since the publication, the same batch of methylene chloride was
 sent to Andy Sivak at Arthur D. Little to be run against the Kakunago clone A31
 of BALB/c 3T3.  It did not transform his cells.  We then repeated the study in
 Fischer rat cells and at the same time tested methylene chloride sent to us by
 the National Coffee Producers Association.  The test was run in triplicate.
 The Fisher methylene chloride again transformed the cells, while the National
 Coffee Producers'  (supplied by Diamond Shamrock) was negative.' The differing
 responses in the two experiments may have been due to differences in the levels
 of impurities present in the samples used.  The chemical compositions of DCM
 samples from different suppliers are given in Table 5-27.
     The Fischer rat embryo cell line contains the genome of the Rauscher
 leukemia virus, but there is no basis for minimizing the positive results.
 Since the mode of action of DCM is not known, this transformation may be due to
 activation of the virus.
 5.3.3.2  Epidemiologic Studies—
 5.3.3.2.1  Friedlander et al. (1978), Hearne and Friedlander (1981).  Friedlander
 et al. (1978) performed mortality analyses of male Eastman-Kodak employees
 exposed to low levels of DCM.  This study was updated by Hearne and Friedlander
 (1981).  Measurements from 1959 to 1975 ranged from 30 to 120 ppm.  Both the
 original  1978 analysis and the 1981 update found no increase in neoplasms,
 heart disease, or any other cause of death in comparison with the two control
 groups, which were composed of other Kodak employees and of New York State
males.  The population was relatively stable and the workers were rotated
 throughout the work area, and thus exposure was averaged among all the workers.
Dichloromethane had been used for 30 years as the primary solvent in this
Eastman-Kodak operation.
                                   5-85

-------
                    TABLE 5-27.  DICHLOROMETHANE ANALYSES*

Methyl chloride
Vinyl chloride
Ethyl chloride
Vinylidine chloride
Carbon tetrachloride
Fisher D-123t
Lot 761542 Diamond ShamrockS
(ppm) (ppm)
0.5 <1
0.8 <1
329 <1
3.6 <1
Dow#
(ppm)
3
33
Chloroform               <0.1                      26                       <5

Trichloroethylene        <0.1                      <1                       <1

1,2-Dichlorethylene     369                        20                       86

Methyl bromide          —                       —                       11

Cyclohexane             —                       300                      305
     *Presented by Drs. Sivak and Kirschman at the Science Advisory Board
      Meeting, Sept.  4-5,  1980.

     tFisher sample used by Dr.  P. Price in two series  of cell  transformation studies.

     §This material is under test in NCA's chronic rat  study  on DCM in
      drinking water.  It was also used in Dr. Price's  follow-up study.

     #This material,  also analyzed by NCI, with the same results as given by
      Drs. Sivak and Kirschman,  was used in the following tests:

                NCA's 90-day studies in rats and mice
                Dr. Sivak's neoplastic transformation assay
                Dr. P. Price's follow-up series of cell  transformation tests
                NCI's chronic bioassay studies in rats  and mice presently under way


                                      5-86

-------
     Two separate mortality analyses were done.   In  the  earlier paper,  one
approach used the proportionate mortality ratio  to examine 334  deaths  of
DCM-exposed workers during 1956 to 1976.   Seventy-one  neoplasms were  found;  73
were expected based on other Kodak employee mortality  ratios.   Furthermore,  no
single site was over-represented.
     A second approach, included in the first paper  and  used exclusively in  the
update, was a cohort mortality study of all 751  employees who were in  the DCM
work area in 1964.  The results of the update are shown  in Tables 5-28 and
5-29, taken from that paper.  There were 110 deaths  during the  16-year follow-
up (retrospective).  Two control groups were used:   other Kodak employees,  and
New York State males.  The expected numbers of deaths  in the exposed  groups
based upon the control group experiences were 105 and  168, respectively. The
differences between the observed and the expected deaths based  on the  controls
are either not statistically significant (other Kodak  employees) or the expected
deaths based on New York State males are significantly increased over  the
observed numbers.
     The results show that malignant neoplasms accounted for 24 of the 110  deaths
in the study cohort, which was less than the 28.6 or 38.5 expected malignancies
based on the control data.  Nine of these 24 deaths  were from respiratory cancer
(7.6 and 13.6 were expected based on the control  groups) and seven were from can-
cers of the digestive organs (less than expected).   Only the two deaths associated
with brain or nervous tissue represented a higher than expected total  (SMR  =
169 and SMR = 227 vs. two control  groups), but these SMRs (standard mortality
ratios) were not statistically significant.
     Further stratification of the cohort focused on the 252 males with 20  years
or more of exposure who were employed in 1964.  In  this  group 59 deaths occurred:
13 were due to malignant neoplasms (17.8 and 24.7 expected based on the two
control groups) and 32 were due to circulatory diseases  (37.9 and 59.7 expected).
     A further analysis of the 252 males shows that this cohort had a  median
age of approximately 54 years in 1964.  With this group  the more common cancers
would have to be markedly increased for there to be  a  reasonable probability of
detecting tKe increase.  For example, following the  cohort for  16 years, cancer
mortality at this age would require 10 deaths from respiratory  cancer to detect
a significant result at the P = 0.05 level.  This represents an increase of at

                                    5-87

-------
                                        TABLE 5-28.  OBSERVED AND EXPECTED DEATHS, 1964-1980,
                                         1964 HOURLY MALE DICHLOROMETHANE COHORTS FROM KODAK
                                                    {Hearne and Friedlander, 1981)
en
i
oo
oo
Cohort exposed minimum

Diagnostic group
Malignant neoplasms
Circulatory diseases
Ischemic heart disease
Other circulatory
Respiratory diseases
Accidents
All other causes
All causes
No.
observed
13
32
27
5
5
2
7
59
KP* no.
expected
17 .77
37.91
27.34
10.57
3.64
2.35
4.27
65.94
20 years
NYSt no.
expected
24.72t
59.735
45.505
14.23
7.11
3.90
10.60
106. 06§
Total 1964 CH2C12 cohort
No.
observed
24
62
48
14
8
6
10
110
KP* no.
expected
28.64
59.11
43.40
15.71
5.17
5.53
6.63
105.08
NYSt no.
expected
38.54§
90.905
69.545
21.36
10.32
9.80
18.44
168.005

            *Based  on  age-sex  specific mortality of hourly men, rest of Kodak Park, 1964-1976.

            tBased  on  age-sex  specific mortality New York State (excluding New York City) for the years 1965, 1970, 1975,

            SObserved-expected differences significant (P £0.05).

-------
                             TABLE 5-29.  MALIGNANT NEOPLASMS, OBSERVED AND EXPECTED DEATHS,  1964-1980,
                                             1964 HOURLY MALE DICHLOROMETHANE COHORTS FROM KODAK
                                                    (Hearne and Friedlander, 1981)
i
oo
                                            Cohort exposed minimum 20 years
Total  1964 CH2Cl2 cohort

Diagnostic Group
Digestive organs
Pancreas
Other
Respiratory
Genitourinary
Brain and other nervous
Lymphatic and hematopoietic
Other malignant neoplasms
Total malignant neoplasms
No.
observed
5
2
3
4
0
1
0
3
13
KP* no.
expected
7.08
1.34
5.74
4.60
2.19
0.62
1.44
1.84
17.77
NYSt no.
expected
7.73
1.48
6.25
8.675
3.48
0.44
2.00
2.40
24.72S
No.
observed
7
3
4
9
1
2
1
4
24
KP* no.
expected
10.93
2.36
8.57
7.56
<=, 3.23
1.18
2.68
3.06
28.64
NYSt no.
expected
11.68
2.24
9.44
13.62
4.94
0.88
3.45
3.97
38.54§

            *Based on age-sex specific mortality of hourly men, rest of Kodak Park, 1964-1976.

            tBased on age-sex specific mortality New York State (excluding New York City) for the years 1965,  1970,  1975.

            §0bserved-expected differences significant (P < 0.05).

-------
least 100 percent over that expected,  the  expected  probability of  lung cancer
death for this cohort being 0.018  over the 16 years,  based on other Kodak
employees'  rates.  The statistical  power to detect  100  percent increases (at a
= 0.05, one-sided)  is about 95  percent for all malignancies  and 45 percent  for
respiratory cancer deaths.   The remaining  499 males with  less than 20 years
exposure were significantly younger (median age  about 36  years), and a follow-up
of this cohort for 16 years mortality  might fail  to detect even a  moderate
effect, since expected cancer mortality in this  age group is so low.
5.3.3.2.2  Ott et a!. (1983a, b, c, d. e).  Ott  et  al.  (1983a, b,  c, d, e)
recently reported the results of a health  evaluation  of employees  of one fiber
production plant who  were exposed  to DCM as part of a solvent mixture consisting
of this substance plus methanol and also acetone in a separate container.   A
second fiber production plant utilizing only acetone, but similar  to the first
in other respects,  was selected as a referent population. The investigation
focused primarily on  health effects occurring to the  cardiovascular system  stem-
ming from the metabolism of DCM to COHb in the body.  In  addition  to a cohort
mortality study, the  authors examined  several other health endpoints in the
still-employed group.  These involved  clinical evaluations,  electrocardiographic
monitoring, metabolism tests, and  evaluation of  oxygen  half-saturation pressures.
     Environmental  exposure in  the plant varied  from  140  ppm in areas of low
DCM use to 475 ppm in areas of  high DCM use, based  on an  8-hour time-weighted
average (TWA).  Methanol was present in smaller  quantities by a factor of  ten,
while acetone ranged  from 100 to 1,000 ppm TWA in both  plants  (Ott et al.,
1983a).  Industrial hygiene surveys were conducted  from September  1977 to
February 1978.
     To qualify for inclusion in the cohort, production employees  (both men
and women) had to have worked a minimum of three months in the preparation  or
extrusion areas of either plant during January 1, 1954  to January  1, 1977.   In
the DCM plant, both cellulose diacetate (acetate) and cellulose triacetate
(CTA) fibers were made side by  side.  Although acetone  was present in both
                                    5-90

-------
plants, it was the solvent of choice for making acetate fibers in the referent
plant.  The exposed cohort consisted of 1,271 persons versus 948 persons in the
referent plant, as follows:
                                                         Referent
                           PCM plant                       plant
   White men                   487                          696
       women                   615                          248
Nonwhite men                    64                            1
       women                   105                            3
             Total           T72TT
Persons were followed through June of 1977.  The authors noted that vital
status was not available for 226 (18 percent) of the DCM-exposed cohort and 112
(12 percent) of the companion plant.  The authors commented that few deaths
would be added from this last group based on "previous experience with the
social security follow-up mechanism."  Such a statement is subject to question,
however, without knowing the age distribution of this group.
     The authors found no excessive mortality from any cause  in the "exposed"
or the "referent" population, either when the group with unknown vital status
was presumed lost to follow-up at the time lost, or was presumed alive until
June of 1977; i.e., person-years would cease accumulating at  the time the
person was lost to follow-up in the former instance, but in the latter case,
person-years would continue to accumulate until  the end of the follow-up period,
as if the individual were still alive.  Altogether in the exposed category for
white men, 37 observed deaths out of 487 were seen versus 34.9 expected based
on the latter definition, or 30.4 expected based on the former definition  re-
garding the group with unknown vital status.  For white women, only 11 deaths
were observed versus either 15.9 or 13.5, depending upon the  choice of the first
or second definition above.  For malignant neoplasms in white men, there were
five observed deaths versus either 6.3 or 5.6 expected.  In white women, two ob-
served deaths were seen, versus either 5.2 or 4.5 expected.  No cancer deaths  were
reported in nonwhite males and females, probably due to the small  size of  these
select subgroups.  Since the authors were mainly interested in cardiovascular
                                    5-91

-------
effects, they decided to examine further only ischemic heart disease in terms
of duration of exposure and length of follow-up.   Even when considering only
individuals with a minimum of 10 years of exposure who were followed for a
minimum of ten years, only two male deaths were observed versus 1.7 expected
based on this cause.  Ischemic heart disease was  not observed as a cause of
death in women.  No corresponding data are available for any form of cancer by
latency or duration of employment.
     Several deficiencies are to be noted with respect to the use of this study
as a sensitive indicator of mortality.  Foremost  among these is the relatively
low and unusual distribution of mortality among the members of the cohort.
Because of the excessive numbers of observed and  expected deaths from external
causes compared to the numbers of observed and expected deaths from malignant
neoplasms, and the fact that deaths from accidents are the leading cause of
death in young males nationwide, it is quite likely that the cohort is a rela-
tively youthful group.  This is further evidenced by the surprisingly few
deaths observed overall  compared to the size of the cohort--!ess than 10 percent
in all instances (women, nonwhite women, and white females).  Unfortunately,
no age breakdown is available to confirm this observation.
     Additionally, only 310 of the exposed males  have had a chance to be followed
at least 17.5 years.  Some 241 exposed males entered the cohort after 1960 and
could not possibly have been followed 17.5 years.  Because of the 15- to 20-year
latency period involved with most human cancers,  cancer effects attributable to
DCM exposure would probably not have been expected to manifest themselves prior
to the 17th year.   Thus, the power of this study  to detect a statistically
significant elevated risk of cancer {as well as ischemic heart disease) is low.
     Another possible problem deals with the extent of follow-up of the cohort.
Almost 18 percent (226)  of the exposed cohort was lost to follow-up as of June
1977.  Although the authors discount that as unimportant, it should be of
concern that if the ages of the lost-to-follow-up group are relatively advanced,
the likelihood is great that enhanced mortality will be in the higher age
groups.  Whenever extraordinary means are employed to determine the vital  status
of a subgroup of the cohort for which all other methods of follow-up have
failed, the residual deaths found as a result of  this endeavor are usually
overly represented by sudden deaths due to heart  failure or accidents.  The
                                    5-92

-------
opportunity for leaving a record of their deaths is minimized because of the
nature of the deaths; hence, it becomes less likely that the vital  status can
be determined.
     Still another problem with this study is the "healthy worker"  effect.  In
most studies of this kind, workers at the time of their employment  and shortly
thereafter are generally somewhat healthier than the population from whence
they came.  Ill or infirm persons do not usually choose jobs that may be detri-
mental to their health.  This tendency usually results in as much as a 20 percent
deficit of mortality compared to that expected.
     In summary, this study is inadequate to assess cancer mortality in the
described cohort for the reasons stated.  The study focuses mainly  on heart
disease as a consequence of DCM exposure.
5.3.3.3  Quantitative Estimation.  This quantitative section deals  with the unit
risk for DCM in air and water and the potency of DCM relative to other carcino-
gens that the CAG has evaluated.  The unit risk estimate for an air or water
pollutant is defined as the lifetime cancer risk occurring in a hypothetical
population in which all individuals are exposed continuously from birth through-
out their lifetimes to a concentration of 1 yg/m^ of the agent in the air they
breathe or to a concentration of 1 yg/1 in the water they drink. This calcula-
tion provides a quantitative estimation of the impact of the agent  as a carcino-
gen.  Unit risk estimates are used to compare the carcinogenic potency of
several agents with each other and to give a crude indication of the population
risk that might be associated with air or water exposure to these agents, if
the actual exposures are known.
5.3.3.3.1  Procedures for the determination of unit risk for animals.  The data
used for the quantitative estimate are taken from one or both of the following:
1) lifetime animal studies, and 2) human studies where excess cancer risk has
been associated with exposure to the agent.  In animal studies it is assumed,
unless evidence exists to the contrary, that if a carcinogenic response occurs
at the dose levels used in the study, then responses will also occur at all
lower doses, with an incidence determined by an extrapolation model.
     There is no solid scientific basis for any mathematical extrapolation
model that relates carcinogen exposure to cancer risks at the extremely low
concentrations that must be dealt with in evaluating environmental  hazards.
                                     5-93

-------
For practical  reasons,  such low levels  of risk  cannot  be measured  directly
either by animal  experiments or by epidemiologic  studies.   We  must,  therefore,
depend on our current understanding of  the mechanisms  of carcinogenesis  for
guidance as to which risk model to use.   At the present time,  the  dominant view
of the carcinogenic process involves the concept  that  most cancer-causing
agents also cause irreversible damage to DMA and  are mutagenic.  There is
reason to expect that the quanta!  type  of biological response, which is  charac-
teristic of mutagenesis, is associated  with a linear non-threshold 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 higher doses,  there  can  be an
upward curvature, probably reflecting the effects of multistage  processes on
the mutagenic response.  The linear non-threshold dose-response  relationship is
also consistent with the relatively few epidemiologic  studies  of cancer  responses
to specific agents that contain enough  information to  make the evaluation
possible (e.g., radiation-induced leukemia, breast and thyroid cancer, skin
cancer induced by arsenic in drinking water, liver cancer  induced  by aflatoxins
in the diet).   Also, some evidence from animal  experiments is  consistent with
the linear non-threshold model (e.g., liver tumors induced in  mice by 2-acetyla-
minofluorene in the large scale EDgi study at the National Center  for Toxicolog-
ical Research, and the initiation stage of the  two-stage carcinogenesis  model
in rat liver and mouse skin).
     Because its scientific basis, although limited,  is the best of any of  the
current mathematical extrapolation models, the  linear  non-threshold model  has
been adopted as the primary basis for risk extrapolation in the  low-dose region
of the dose-response relationship.  The risk estimates made with this model
should be regarded as conservative, representing  the most  plausible upper limit
for the risk;  i.e., the true risk is not likely to be  higher  than  the estimate,
but it could be lower.
     The mathematical formulation chosen to describe  the linear  non-threshold
dose-response relationship at low doses is the  linearized  multistage model.
This model employs enough arbitrary constants to  be able to fit  almost any
monotonically increasing dose-response  data, and  it incorporates a procedure
                                   5-94

-------
 for estimating the largest possible linear slope (in the 95 percent confidence
 limit sense) at low extrapolated doses that is consistent with the data at all
 dose levels of the experiment.
 5.3.3.3.1.1  Description of the low-dose animal  extrapolation model.  Let P(d)
 represent the lifetime risk (probability) of cancer at dose d.  The multistage
 model has the form

           P(d) = 1 - exp [-(q0 + qjd + q2d + q2d2 + ... + qkdk]
 where
                           qj _> 0, i = 0, 1, 2,  .... k.
 Equivalently,
                 Pt(d) = 1 - exp [-(q^ + q2d2 + ... + qkdk)]
 where
                            P (d) =  P(d) - P(0)
                             t        i - P(0)

is the extra risk over background rate at dose d or the effect of treatment.
     The point estimate of the coefficents qj, i = 0, 1, 2, ..., k, and
consequently the extra risk function P^(d) at any given dose d, is calculated
by maximizing the likelihood function of the data.
     The point estimate and the 95 percent upper confidence limit of the extra
risk, Pt(d), are calculated by using the computer program GLOBAL79, developed by
Crump and Watson (1979).  At low doses, upper 95 percent confidence limits on the
extra risk and lower 95 percent confidence limits on the dose producing a given
risk are determined from a 95 percent upper confidence limit, q*, on parameter qj.
                                                               1
Whenever qi > 0, at low doses the extra risk Pt(d) has the approximate form
P-t(d) = qj x d.  Therefore, q* x d is a 95 percent upper confidence limit on the
extra risk, and R/q* is a 95 percent lower confidence limit on the dose producing
an extra risk of R.  Let LQ be the maximum value of the log-likelihood function.
The upper limit, q*, is calculated by increasing qj to a value q* such that
when the log-likelihood is remaximized subject to this fixed value, q*, for the

                                    5-95

-------
linear coefficient, the resulting maximum value of the log-likelihood  LI
satisfies the equation

                               2 (LQ - LI) =  2.70554

where 2.70554 is the cumulative 90 percent point of the chi -square  distribution
with one degree of freedom, which corresponds to a 95  percent upper limit
(one-sided).  This approach of computing the  upper confidence limit for the
extra risk, Pt(d), is an improvement on the Crump et al .  (1977)  model.   The
upper confidence limit for the extra risk calculated at low doses  is always
linear.  This is conceptually consistent with the linear non-threshold concept
discussed earlier.  The slope, q*, is taken as an upper bound of the potency  of
the chemical in inducing cancer at low doses.  (In the section calculating the
risk estimates, P^fd) will be abbreviated as  P.)
     In fitting the dose-response model, the  number of terms in  the polynomial
is equal to (h-1), where h is the number of dose groups in  the experiment,
including the control group.
     Whenever the multistage model does not fit the data sufficiently,  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
                                  (1-Pj)
is calculated where N^  is the number of animals in the in dose group,  X^  is
the number of animals in the i™1 dose group with a tumor response,  P.,-  is the
probability of a response in the i'th dose group estimated by fitting the
multistage model to the data, and h is the number of remaining groups.   The
fit is unacceptable whenever X^ is larger than the cumulative 99 percent 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.
                                      5-96

-------
5.3.3.3.1.2  Selection of data.   For some chemicals,  several  studies  using
different animal  species, strains,  and sexes and run  at several  doses and
different routes  of exposure are available.   A choice must be made as to which
of the data sets  from several  studies to use in the model.   The  procedures used
in evaluating these data are consistent with the approach of making a maximum-
likely-risk estimate.  They are  listed as follows:
     1.  The tumor incidence data are separated according to organ sites or tumor
types.  The set of data (i.e., dose and tumor incidence) used in the  model is the
set where the incidence is significantly higher statistically than the control
for at least one  test dose level or where the tumor incidence rate shows a
statistically significant trend  with respect to dose  level.  The data set that
gives the highest estimate of the lifetime carcinogenic risk, q*, is  selected
in most cases.  However, efforts are made to exclude  data sets that produce
spuriously high risk estimates because of a small  number of animals.   That is,
if two sets of data show a similar dose-response relationship and one has a
very small sample size, the set  of data having the larger sample size is selected
for calculating the carcinogenic potency.
     2.  If there are two or more data sets of comparable size that are identi-
cal with respect  to species, strain, sex, and tumor sites, the geometric mean
of q*, estimated  from each of these data sets, is used for risk  assessment.
The geometric mean of numbers Aj_, A2, ..., Am is defined as

                       (Ax x A2  x ...

     3.  If two or more significant tumor sites are observed in  the same study,
and if the data are available, the number of animals  with at least one of the
specific tumor sites under consideration is used as incidence data in the model
5.3.3.3.1.3  Calculation of human equivalent dosages.  It is appropriate to
correct for metabolism differences between species and absorption factors via
different routes  of administration.  Following the suggestion of Mantel and
Schneiderman (1975), it is assumed that ing/surface area/day is an equivalent
dose between species.  Since the surface area is approximately proportional to
the two-thirds power of the weight, as would be the case for a perfect sphere,
                                       5-97

-------
the exposure in mg/day per two-thirds power of the  weight is  also  considered
to be equivalent exposure.  In an animal  experiment,  this equivalent  dose  is
computed in the following manner.  Let
    Le = duration of experiment
    le = duration of exposure
    m  = average dose per day in mg during administration of  the agent (i.e.,
         during le), and
    W  = average weight of the experimental  animal.

Then, the lifetime average exposure is

                                d-  leXm
                                     Le x W2/3

     Inhalation—When exposure is via inhalation,  the calculation  of dose can  be
considered for two cases where 1) the carcinogenic agent is  either a completely
water-soluble gas or an aerosol  and is absorbed proportionally  to  the amount of
air breathed in, and 2) where the carcinogen is a  poorly water-soluble gas
that reaches an equilibrium between the air breathed and the body  compartments.
After equilibrium is reached, the rate of absorption of these agents is expected
to be proportional to the metabolic rate, which in turn is proportional to the
rate of oxygen consumption, which in turn is a  function of surface area.
     Case 1—Agents that are in the form of particulate matter  of  virtually
completely absorbed gases, such as sulfur dioxide, can reasonably  be expected
to be absorbed proportionally to the breathing  rate.  In this case the exposure
in mg/day may be expressed as

                                   m = I x v x  r

where I = inhalation rate per day in m3, v = mg/m3 of the agent in air and
r = the absorption fraction.
                                       5-98

-------
     The inhalation rates,  I,  for  various  species  can be calculated from  the
observations of the Federation of  American Societies for Experimental  Biology
(FASEB 1974) that 25 g mice breathe 34.5 liters/day and 113  g  rats breathe
105 liters/day.  For mice and  rats of other weights W (in  kilograms),  the sur-
face area proportionality can  be used to find breathing rates  in m3/day,  as
follows:

                   For mice, I = 0.0345  (W/0.025)2/3 m3/day
                   For rats, I = 0.105 (W/0.113)2/3 m3/day.

For humans, the value of 20 m3/day* is adopted as  a standard breathing rate
(ICRP 1977).
     The equivalent exposure in mg/W2/3  for these  agents can be derived from
the air intake data in a way analogous to  the food intake  data. The  empirical
factors for the air intake per kilogram per day,  i =  I/W,  based upon  the pre-
viously stated relationships,  are  tabulated as follows:

                   Species            W              i  =  I/W
                     Man            70                 0.29
                     Rats           0.35                0.64
                     Mice           0.03                1.3

Therefore, for particulates or completely  absorbed gases,  the  equivalent ex-
posure in mg/W2/3 is

                             m  =  Ivr  =  iWvr = iwl/3vr.
     *From "Recommendation of International Commission on Radiological  Protection,"
page 9.  The average breathing rate is 10? cm3 per 8-hour workday and 2 x 10' cm3
in 24 hours.
                                      5-99

-------
In the absence of experimental  information or a  sound  theoretical  argument  to
the contrary, the fraction absorbed,  r,  is assumed  to  be  the  same  for  all species.
     Case 2 — The dose in mg/day of partially soluble vapors is  proportional  to
the 02 consumption, which in turn is  proportional to W2'3 and is also  proportional
to the solubility of the gas in body  fluids, which  can be expressed as an absorp-
tion coefficient, r, for the gas.  Therefore, expressing  the  Q£ consumption as
02 = k W2'3, where k is a constant independent of species,  it follows  that

                               m = k  W2/3 x v x r
or
                               d =  J"   = kvr.
                                      ~
As with Case 1, in the absence of experimental  information or a sound theoretical
argument to the contrary, the absorption fraction, r,  is assumed to be the same
for all species.  Therefore, for these substances a certain concentration in
ppm or yg/m3 in experimental animals is equivalent to  the same concentration
in humans.  This is supported by the observation that  the minimum alveolar con-
centration necessary to produce a given "stage" of anesthesia is similar in man
and animals (Dripps et al . , 1977).  When the animals are exposed via the oral route
and human exposure is via inhalation or vice versa, the assumption is made, unless
there is pharmacokinetic evidence to the contrary, that absorption is equal by
either exposure route.
5.3.3.3.1.4  Calculation of the unit risk from animal  studies.  The 95 percent
upper-limit risk associated with d mg/kg2/3/day is obtained from GLOBAL79, and
for most cases of interest to risk assessment, can be  adequately approximated
by P(d) = 1 - exp (-q*d).  A "unit risk" in units X is the risk corresponding
to an exposure of X = 1.  This value is estimated by finding the number of
mg/kg2/3/day that corresponds to one unit of X and substituting this value into
the above relationship.  Thus, for example, if X is in units of yg/m3 in the
air, we have for case 1, d = 0.29 x 7C-1/3 x 10~3 yg/kg2/3/day, and for case 2,
d = 1, when yg/m3 is the unit used to compute parameters in animal experiments.
                                     5-100

-------
     If exposures are given in terms of ppm in air,  the following calculation
may be used:

                1 ppm = 1.2 x molecular weight (gas) mg/m3.
                              molecular weight (air)

Note that an equivalent method of calculating unit risk would be to use
mg/kg/day for the animal exposures, and then to increase the jth polynomial
coefficient by an amount

                          (Wh/Wa)J/3   j = 1, 2, .... k

and to use mg/kg/day equivalents for the unit risk values.   In the section
calculating the unit risk for animals,  the final q* will always be the upper-
limit potency estimate for human risk based on animal data.
5.3.3.3.1.5  Interpretation of quantitative estimates.   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 differences 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 equivalent doses for humans compared to animals
on a mg/surface area basis is virtually without experimental verification re-
garding carcinogenic response.  Finally, human populations are variable with
respect to genetic constitution, 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 as 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, and by the same route of exposure,
preferably inhalation.
                                     5-101

-------
     The quantitative aspect of carcinogen risk  assessment  is  included  here
because it may be of use in the regulatory decision-making  process,  e.g.,
setting regulatory priorities,  evaluating the adequacy  of technology-based
controls, etc.  However, with present technology,  only  imprecise  estimations
are possible concerning cancer risks to humans at  low levels of exposure.  At
best, the linear extrapolation model  used here provides a rough but  plausible
estimate of the upper limit of risk,  and while the true risk is probably  not
much more than the estimated risk,  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
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.
5.3.3.3.1.6  Alternative methodological approaches.  Methods used by the  CAG
for quantitative assessment are consistently conservative,  i.e.,  they tend toward
high estimates of risk.  The most important part of the methodology  contributing
to this conservatism is the linear  non-threshold extrapolation model.  There
are a variety of other extrapolation models that could  be used, all  of which
would give lower risk estimates.  These alternative models, the one-hit,  Probit,
and Wei bull models, have not been used by the CAG  in the following analysis,  but
are included for comparison in the  appendix.  The  CAG's position  is  that, given
the limited data available from these animal bioassays, especially at the
high-dose levels required for testing, almost nothing is known about the  true
shape of the dose-response curve at low environmental levels and  that the risk
estimates obtained by use of the linear non-threshold model represent plausible
upper limits only.
     Extrapolation from animals to humans could  also be done on  the  basis of
relative weight rather than on the basis of relative surface  area.  Although
the latter approach, used here, has more justification  in  terms  of human  pharma-
cological responses, it is not yet clear which of the two  approaches is more
appropriate for the assessment of carcinogenicity.   In  the  absence of informa-
tion on this point, it seems appropriate to use  the more conservative basis  for
extrapolation, or relative surface area.
                                    5-102

-------
5.3.3.3.2  Unit risk potency estimates and relative potency.   Neither of the
two available epidemiologic studies provided positive results from which to
derive a quantitative risk estimate for DCM exposure.  With respect to animal
data, three chronic studies are relevant to this discussion:  two inhalation
studies and one drinking water study.  These are summarized in Table 5-30.   In
the two inhalation studies (Dow Chemical Company 1980, 1982)  only the salivary
gland regions in male Sprague-Dawley rats in the 1980 study showed statistically
significant increased cancers.  These results are discussed in Section 5.3.3.1.1.2
and are presented in Tables 5-11 through 5-14.  In the drinking water study (NCA
1982), the increase over untreated controls in combined neoplastic nodules  and
hepatocellular carcinomas in female rats was the only statistically significant
finding (Table 5-24).  However, when compared with historical controls, these
results lose their statistical significance.
5.3.3.3.2.1  Unit risk (yg/m3) for inhalation studies.  The data used for esti-
mates of the unit risk for inhalation are presented in Table  5-31, which shows
the positive salivary gland region sarcomas for the inhalation bioassay.
Exposure was 6 hours/day, 5 days/week, for 2 years.  Equivalent dosages were
determined for humans from the animal dosages utilizing the equivalent dosage
methodology presented previously.  As described in the section on pulmonary
uptake and distribution, DCM is readily absorbed into the body following inhala-
tion and it equilibrates rapidly across the alveolar epithelium.  Therefore,
the CAG considers it a virtually completely absorbed gas especially at low
doses and determines equivalent human exposure as explained under Case 1
of the inhalation section.  As presented in Table 5-31, the nominal exposures
are nearly 15 times the human lifetime equivalent exposures.   This difference,
24/6 x 7/5 = 5.6, is partly due to the use of a continuous equivalent dosage.
There is an additional factor of about 2.6, however, which is attributable
to the nature of the method used for determining human equivalent dosages
for inhalation studies.  Put another way, if DCM had been determined to be  a
partially soluble vapor, the unit risk slope would be less by a factor of about
2.6.
                                     5-103

-------
                             TABLE 5-30.  SELECTED DICHLOROMETHANE CHRONIC ANIMAL STUDIES
en
i
i—»
O
Author
Dow (1980)
Dow (1980)
Animal species
and strains
Sprague-Dawley rats
Syrian hamsters
Numbers
95 rats/sex/dose
107-109 hamsters
sex/dose
Routes
Inhalation
6/hours/day
5 days/week
for 2 years
Inhalation
6 hours/day
5 days/week
for 2 years
Doses
0, 500, 1500,
and 3500 ppm
0, 500, 1500,
and 3500 ppm
Resul ts
Salivary gland
sarcoma in males
"
   Dow  (1982)
   NCA  (1982)
Sprague-Dawley rats
90 rats/sex/
dose
Fischer 344 rats
85 rats/sex/
dose; 135 male
controls; 134
female controls
Inhalation
6 hours/day
5 days/week
for 20 (males)
or 24 (females)
months

Deionized
drinking water
for 24 months
0, 50, 200, and
500 ppm
0, 5, 50, 125,
250, 250
(recovery)
Borderline neo-
plastic nodule
response of the
liver in females

-------
       TABLE 5-31.  INCIDENCE ".ATES OF SALIVARY GLAND REGION SARCOMAS IN
                        MALE SPRAGUE-DAWLEY RATS IN THE
          DOW CHEMICAL COMPANY (1980) DICHLOROMETHANE INHALATION STUDY
  Continuous human equivalent
  (animal nominal) exposures
     ppm
 and
ig/m3
      Incidence rates

  Number of rats with tumors/
total  rats examined (%)
0
34
103
240
(0)
(500)
(1
(3
,500)
,500)
1
3
8
0 (0)
.2
.6
.4
x 105
x 105
x 105
(1
(5
(1
.8
.2
.2
x 106)
x 106)
x 107)
1/93
0/94
5/91
11/88
(1%)
(0%)
(5%)
(12%)

* 1 ppm x 1.2 x 103 yg/m3 x 84.9 = 3.5 x 103 yg/m3
                            "ZO"

An exposure of 500 ppm DCM in air expressed as yg/m3 is

          500 ppm x 3.5 x 103 yg/m3 = 1.8 x 106 yg/m3
                         ppm

Since animal exposure was for 6 hours/day, 5 days/week, the animal  continuous
lifetime exposure equivalent was

           1.8 x 106 ug/m3 x  6 x 5 = 3.2 x 105 yg/m3
                             ?4~   7

The human equivalent dose for DCM is calculated first by determining the amount
actually breathed by the rats.  As presented in the inhalation dose equivalence
section, 350 g rats breathe

           I = 0.105 (0.350/0.113)2/3 = 0.223 m3 (air)/day.

Thus continuous animal  dose was 3.2 x 105 yg/m3 x 0.223 m3/day = 7.15 x 104 yg/
day, or, for a 350 g rat, 7.15 x 104 yg/day/0.350 kg = 2.0 x 105 yg/kg/day.
The human equivalent dose is

             2.0 x 105 yg/kg/day * (70/0.350)!/3 = 3.5 x io4 yg/kg/day
  or
3.5 x 104 yg/kg/day x 70 kg = 1.2 x 10$ yg/m3.
                      20 m3
                                    5-105

-------
     When the above incidence data were fitted with  the  continuous  human
equivalent exposures,  the linearized multistage model yielded  the following
value for the 95 percent upper limit of risk:
                       q* = 1.8 x ID'7 (  yg/m3)-!
5.3.3.3.2.2  Unit risk (mg/kg/day)  and (yg/1)  for oral  studies.   This unit risk
estimate should be used only under  the assumption that  DCM is  a  potential  human
carcinogen.  As discussed in the qualitative section, there is only limited
animal evidence and no human evidence to support that assumption.
     Since publication of a final NTP report on gavage  studies for rats and
mice on DCM has been cancelled,  there is no suitable oral  study  from which to
estimate a unit risk.
5.3.3.3.2.3  Comparison of animal and human inhalation  studies.   The study of
Kodak employees, yielding negative  cancer results, is compared with the positive
tumor results of the rat inhalation study (Dow, 1980).   In the latter study,  the
salivary gland region tumors in  male Sprague-Dawley rats led to  a 95 percent  upper-
limit slope estimate of q* = 1.8 x  10"7 (yg/m3)"!.  If  this slope factor is applied
to the human inhalation study, in which time-weighted average  exposure was esti-
mated as ranging from approximately 30 to 120 ppm, the  expected  impact of expo-
sure can be estimated.  For this purpose, 1 yg/m3 of DCM is equivalent to 2.9 x
10~* ppm.  Thus, the upper-limit slope in ppm is

      q* = 1.8 x 10-7 (ug/n,3)-l  x 3.45 x 103 yg/m3 = 6.2 x 10-4  ppm-l.
       1                                ppm
                                     5-106

-------
Since this upper-limit slope is based on continuous lifetime equivalent exposure,
the exposure range of 30 to 120 ppm time-weighted average must be adjusted to
lifetime equivalence as follows:

        30 ppm x 20 years x 240 days x  8 hrs = 1.88 ppm
                 Tff         "353"        "2T
and 7.52 ppm lifetime equivalence for the 120 ppm exposure group,  assuming
20 years of exposure for the 252 long-term exposure workers.   Based on this
upper-limit slope factor and the range of exposure, the group of 252 workers
could expect an additional  lifetime risk of between

          R = 6.2 x 10~4 ppm'1 x 1.88 ppm = 1.2 x 10~3  and 4.7 x 10~3.

For these 252 workers, this would translate to an upper limit of between
252 x 1.2 x 10-3 =0.3 and 1.2 excess lifetime cancer deaths.  Based on the
total expected Kodak employee deaths of 65.9 (Table 5-28)  for this 20 years
minimally exposed cohort, we would expect 26 percent of the cohort to die in
the 17-year follow-up period.  Transposing this 26 percent to the  expected
excess lifetime cancer deaths, an upper limit of between 0.1  and 0.3 excess
cancer deaths can be expected during the 17-year follow-up period.  The power
to detect this increase from 17.77 to 18.1 cancer deaths is less than 10 percent.
If these excess cancer deaths were from cancers of one  specific site, the power
would be greater, but not great enough to declare this  a negative  study.  Even
for a rare cancer, such as a liver cancer, the expected number of  cases would
be much less than 1.  Since only deaths can be observed, the  power to observe
one death from liver cancer in this cohort is quite small.
     Based on the above analysis, the study of Kodak workers  exposed to DCM,
showing no increase in cancer, cannot be judged as having negative results
because of its low power, which is related to low exposure from a  weak animal
carcinogen.
5.3.3.3.2.4  Relative potency.  One of uses of the unit risk  concept is to com-
pare the relative potencies of carcinogens.  To estimate relative  potency on  a
per mole basis, the unit risk slope factor is multiplied by the molecular weight,
                                   5-107

-------
and the resulting number is expressed in terms  of  (mMol/kg/day)~!.   This  is
called the relative potency index.
     Figure 5-1 is a histogram representing the frequency  distribution  of the
potency indices of 54 chemicals evaluated by the CAG  as  suspect  carcinogens.
The actual data summarized by the histogram are presented  in  Table  5-32.   Where
human data are available for a compound, they have been  used  to  calculate the
index.  Where no human data are available, animal  oral  studies and  animal
inhalation studies have been used,  in that order.  Animal  oral studies  are
selected over animal inhalation studies because most  of the chemicals have
been subjected to animal oral studies; this allows potency comparisons  by route.
     The potency index for DCM, based on salivary  gland region tumors in  male
Sprague-Dawley rats in the Dow inhalation study (1980),  is 5.3 x 10~2 (mMol/kg/
day)-1.  This is derived as follows:   the slope estimate from the Dow study,  1.8  x
10"7 (ug/m3)-!, in converted units of 6.3 x 10~4 (mg/kg/day)"1,  is  multiplied  by
the molecular weight of 84.9 to give  a potency  index  of 5.3 x 10~2. Rounding
off of the nearest order of magnitude gives a value of 10~1,  which  is the scale
presented on the horizontal axis of Figure 5-1. The  index of 5.3 x 10~2  is  the
least potent of the 54 suspected carcinogens.  Ranking of  the relative  potency
indices is subject to the uncertainty of comparing estimates  of  potency of
different chemicals based on different routes of exposure  to  different  species
using studies of different quality.  Furthermore,  all the  indices are based  on
estimates of low-dose risk using linear extrapolation from the observational
range.  Thus these indices are not valid for the comparison of potencies  in  the
experimental or observational range if linearity does not exist  there.   The
potency index for DCM, furthermore, is valid only  under the assumption  that
DCM is a potential human carcinogen.   The evidence for that is limited.
5.3.3.3.2.5  Summary of quantitative  estimation.  No  positive epidemiologic
studies exist from which to estimate  a unit risk for  exposure to DCM.   Only  one
animal data set has shown increased cancers from which a unit risk  assessment
could be estimated.  In the Dow Chemical Company (1980) rat inhalation  study,  there
were increased sarcomas in the salivary gland region  of male  rats;  this yielded  an
upper 95 percent limit of the potency estimate  of q*  = 1.8 x  10~7
                                    5-108

-------
  4th
quartile
                                     -j
      3rd
    quartile
                             -h
                            4x10*2
                                   2nd
                                 quartile
                        +
                      2x10*3
                                                                  1st
                                                                quartile
                                    1x10*1
              I

                         CN

                             'CO


                                 I
             I
1
          I
          -2
Jill
0246
   Log of Potency Index
                     I
                    8
Figure 5-1.  Histogram representing the frequency distribution of the
potency indices of 54 suspect  carcinogens evaluated by the Carcinogen
Assessment Group.
                                       5-109

-------
TABLE 5-32.  RELATIVE CARCINOGENIC POTENCIES AMONG 54 CHEMICALS EVALUATED
  BY THE CARCINOGEN ASSESSMENT GROUP AS SUSPECTED HUMAN CARCINOGENSl>2,3
Compounds
Acrylonitrile
Aflatoxin Bj
Aldrin
Ally! Chloride
Arsenic
B[a]P
Benzene
Benzidine
Beryllium
Cadmium
Carbon Tetrachloride
Chlordane
Chlorinated Ethanes
1,2-dichl oroethane
hexachloroethane
1 , 1 ,2 ,2-tetrachl oroethane
1 , 1 , 1- tri chl oroethane
1 , 1 ,2-tri chl oroethane
Chloroform
Chromium
DDT
Di chl orobenzi dine
1 , 1-di chl oroethyl ene
Dieldrin
Slope
(mg/kg/day)'1
0.24(W)
2924
11.4
1.19 x ID'2
15(H)
11.5
5.2 x 10-2(W)
234(W)
4.86
6.65(W)
1.30 x 10-1
1.61
6.9 x 10-2
1.42 x 10-2
0.20
1.6 x 10-3
5.73 x lO-2
7 x ID'2
41(W)
8.42
1.69
1.47 x 10-1(1)
30.4
Molecular
weight
53.1
312.3
369.4
76.5
149.8
252.3
78
184.2
9
112.4
153.8
409.8
98.9
236.7
167.9
133.4
133.4
119.4
104
354.5
253.1
97
380.9
Order of
magnitude
Potency (logio
index index)
1 x 10+1
9 x 10+5
4 x 10+3
9 x 10-1
2 x 10+3
3 x 10+3
4 x 10°
4 x 10+4
4 x 10+1
7 x 10+2
2 x 10+1
7 x 10+2
7 x 10°
3 x IQO
3 x 10+1
2 x 10-1
8 x 10°
8 x 10°
4 x 10+3
3 x 10+3
4 x 10+2
1 x 10+1
1 x 10+4
+1
+6
+4
0
+3
+3
+1
+5
+2
+3
+1
+3
+1
0
+1
-1
+1
+1
+4
+3
+3
+1
+4
                               5-110

-------
                              TABLE 5-32.   (continued)
Compounds
Dinitrotoluene
Di phenyl hydrazine
Epichlorohydrin
Bis(2-chloroethyl )ether
Bis(chloromethyl ) ether
Ethyl ene Di bromide (EDB)
Ethyl ene Oxide
Formal dehyde
Heptachlor
Hexachlorobenzene
Hexachlorobutadiene
Hexachl orocycl ohexane
technical grade
alpha isomer
beta isomer
gamma isomer
Methyl ene Chloride
(dichloromethane)
Nickel
Nitrosamines
Dime thy! nitrosamine
Diethylnitrosamine
Di butyl ni trosami ne
N-ni trosopyrrol i di ne
N-ni troso-N-ethyl urea
N-ni troso-N-methyl urea
N-ni troso-di phenyl ami ne
Slope
(mg/kg/day)-l
0.31
0.77
9.9 x 10-3
1.14
9300(1)
8.51
0.63(1)
2.14 x 10-2(1)
3.37
1.67
7.75 x 10-2

4.75
11.12
1.84
1.33
6.3 x 10-4

1.15(W)

25.9{not by q*)
43.5(not by q*)
5.43
2.13
32.9
302.6
4.92 x ID'3
Molecular
weight
182
180
92.5
143
115
187.9
44.0
30
373.3
284.4
261

290.9
290.9
290.9
290.9
84.9

58.7

74.1
102.1
158.2
100.2
117.1
103.1
198
Order of
magnitude
Potency (logio
index index)
6 x 10+1
1 x 10+2
9 x 10-1
2 x 10+2
1 x 10+6
2 x 10+3
3 x 10+1
6 x 10-1
1 x 10+3
5 x 10+2
2 x 10+1
. o
1 x 10+3
3 x 10+3
i f\
5 x 10+2
4 x 10+2
5 x 10-2

7 x 10+1

2 x 10+3
4 x 10+3
9 x 10+2
, o
2 x 10+2
i O
4 x 10+3
i M
3 x 10+4
1 x 10°
+2
+2
0
+2
+6
+3
+1
0
+3
+3
+1

+3
+3
+3
+3
-1

+2

+3
+4
+3
+2
+4
+4
0
PCBs
4.34
324
1 x 10+3
                                    5-111

-------
                              TABLE 5-32.   (continued)
Slope Molecular Potency
Compounds (mg/kg/day)-1 weight index
Phenols
2,4,6-trichlorophenol 1.99 x 1Q-2 197.4
Tetrachlorodioxin 4.25 x 105 322
Tetrachloroethylene 6 x 10'2 165.8
Toxaphene 1.13 414
Trichloroethylene 1.26 x 10-2 131.4
Vinyl Chloride 1.75 x 10-2(I) 62.5
4 x 10°
1 x 10+8
1 x 10+1
5 x 10+2
2 x 10°
1 x 100
Order of
magnitude
Oog10
index)
+1
+8
+1
+3
0
0

Remarks:
1. Animal slopes are 95 percent upper-limit slopes

based on the 1

inearized
        multi-stage 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 non-threshold model.

    2.   The potency index is a rounded-off  slope in (mMol/Kg/day)-1  and is cal-
        culated by multiplying the  slopes in  (mg/kg/day)-1 by the molecular
        weight of the compound.

    3.   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.
equivalent to q* = 6.3 x 10~4 (mg/kg/day)-1  by  the  oral  route.   However,  no

positive data exist with which to provide a  direct  oral  route estimate.   In

total, there is only limited evidence that DCM  is a potential human  carcinogen.

The unit risk estimate is valid only if one  accepts that limited evidence.

Further, if one chooses to express the potency  of DCM relative  to that of 54

chemicals evaluated as suspect carcinogens,  DCM is  the weakest,  ranking last.
                                    5-112

-------
5.3.3.4  Summary—Six chronic studies of DCM administered to animals have been
reported:  four in rats, one in mice, and one in hamsters.  The Dow Chemical
Company (1980) reported the results of chronic inhalation studies in rats and
hamsters.  The rat study showed a small  increase in the number of benign mammary
tumors compared to controls in female rats at all doses and in male rats at the
highest dose, as well as a statistically significant increased incidence of
ventral cervical sarcomas, probably of salivary gland origin, in male rats.  The
response pattern of the salivary gland tumors is unusual, consisting of sarcomas
only and appearing in males but not in females.  In hamsters, there was an
increased incidence of lymphosarcoma in females only, which was not statistically
significant after correction for survival.  The second inhalation study in rats
by the Dow Chemical Company (1982) reported that there were no compound-related
increased incidences of any tumor type,  but that the highest dose was far below
that of the previous Dow inhalation study in rats.   The National Coffee Associa-
tion (NCA) drinking water study in Fischer 344 rats reported that the incidence
of neoplastic nodules and/or hepatocellular carcinomas in female rats was
increased significantly with respect to matched controls, but their incidence
was within the range of historical control values at that laboratory.  The NTP
(1982) draft gavage study on rats and mice will not be published as a final
report due to data discrepancies.  Selected information from the gavage studies
may be incorporated into the future NTP inhalation  bioassay, pending the results
of the in-depth audit.
     There are some other inadequate animal studies in the literature.  One
study (Theiss et al., 1977) reported a marginally positive pulmonary adenoma
response in strain A mice injected intraperitoneally with DCM.  Two negative
animal inhalation studies were inadequate because they were not carried out for
a full lifetime (Heppel  et al., 1944; McEwen et al., 1972).
     Two other carcinogenic!ty studies of DCM in animals are currently in
progress:  an NTP 2-year bioassay by inhalation in  Fischer 344/N rats and
B6C3F1 mice, and an NCA 2-year drinking water study in B6C3F1 mice.
     Positive results in a rat embryo cell transformation study were reported by
Price et al. (1978).  The significance of their findings with regard to carcino-
genicity is not well understood at the present time.
                                      5-113

-------
     The epidemiologic data consists  of two studies:   Friedlander  et  al.  (1978),
updated by Hearne and Fried!ander (1981);  and Ott et  al.  (1983a, b, c,  d,  e).
Although neither study showed excessive risk, both showed sufficient  deficiencies
to prevent them from being judged negative studies.   The  Friedlander  et al.  study
(1978) lacked great enough exposure (based on animal  cancer  potency estimates)
to provide sufficient statistical  power to detect a potential  carcinogenic
effect.  The Ott et al.  study (1983a, b, c, d, e), among  other deficiencies,
lacked a sufficient latency period for site-specific  cancer.
     Only one animal  data set has shown increased cancers from which  a  unit  risk
assessment could be estimated.  In the Dow Chemical Company  (1980) rat  inhala-
tion study, there were increased sarcomas in the salivary gland region  of  male
rats; this yielded an upper 95 percent limit of the potency  estimate  of q* =  1.8 x
10~7 (ug/m3)-1, equivalent to q* = 6.3 x 10~4 (mg/kg/day)'1  by the oral  route.
This unit risk estimate  should be used only under the assumption that DCM  is  a
potential human carcinogen.  As discussed previously, there  is only limited
animal evidence and no human evidence to support that assumption.
5.3.3.5  Conclusions.  Animal studies show a statistically positive salivary  gland
sarcoma response in male rats (Dow, 1980)  and a borderline hepatocellular  neo-
plastic nodule response  in the rat (NCA, 1982).  There is also evidence that  DCM
is weakly mutagenic.   According to the criteria of the International  Agency
for Research on Cancer (IARC), the weight of evidence for carcinogem'city  in
animals is limited.
     There was an absence of epidemiologic evidence for the  carcinogem'city  of
DCM in a well-conducted epidemiologic study having long-term exposure.   However,
on the basis of animal data, the level of exposure to the individuals in the
study was too low to produce an observable increase in cancer.  The overall
evaluation of DCM, based on IARC criteria, is group 3, meaning that the chemical
cannot be classified as  to its carcinogem'city for humans.
     The unit risk for DCM, based on a rat inhalation study, is estimated  to be
1.8 x 10~7 for a lifetime exposure to 1 yg/m3 in air, but the above  is  true  only
under the assumption that DCM is a potential human carcinogen.  Even  under that
assumption, the potency of DCM is the lowest of the 54 chemicals which  the CAG
has evaluated as suspect carcinogens.
                                        5-114

-------
5.4  REFERENCES

Abrahamson, S. , and R.  Valencia.  1980.  Evaluation of substances of interest
     for genetic damage using Drosophila melanogaster.  Prepared for FDA
     Contract No.  233-77-2119.

ACGIH (American Conference of Governmental Industrial Hygienists).  1980.
     Threshold Limit Values for Chemical Substances and Physical Agents in the
     Workroom Environment with Intended Changes for 1980.  Cincinnati, OH:
     ACGIH, p. 23.

Adams, J.D., and H.H.  Erickson.  1976.  The effects of repeated exposure to
     methylene chloride vapor.  Prep. Annu. Sci. Meet., Aerosp. Med. Assoc.
     61-62.

Ahmed, A.E., and M.W.  Anders.  1976.  Metabolism of dihalomethanes to formal-
     dehyde and inorganic halide.  I.  Jjn vitro studies. Drug Metab. Dispos.
     4(4):  357-361.

Ahmed, A.E., and M.W.  Anders.  1978.  Metabolism of dihalomethanes to formal-
     dehyde and inorganic halide.   II.  Studies on the mechanism of the re-
     action.  Biochem.  Pharmacol.  27:2021-2025.

Anders, M.W., V.L.  Kubic, and A.E. Ahmed.  1978.  Bioorganic mechanisms of the
     metabolism of dihalomethanes to carbon monoxide, formaldehyde, formic
     acid and inorganic halide.  Int. Congress Ser. Excerpta Med.: Ind. Environ.
     Xenobiotica 440:22-24.

Anderson, E.W., R.J. Andelman, and J.M. Strauch.  1973.  Effect of low-level
     carbon monoxide exposure on onset and duration of angina pectoris: a
     study in ten patients with ischemic heart disease.  Ann. Intern. Med.
     79:46-50.

Anderson, R.F., D.C. Allensworth, and W.J. deGroot.   1967.  Myocardial toxicity
     from carbon monoxide poisoning.  Ann. Intern. Med. 67:1172-1182.

Anonymous.   1947.   Methylene chloride.  Ind.  Hyg. Newsletter 7:15.

Arthur D. Little (unpublished results, 1978).

Arthur D. Little (unpublished results, 1980).

Aviado, D.M., and M.A.  Belej.  1974.  Toxicity of aerosol propellants on the
     respiratory and circulatory systems.  I. Cardiac arrhythmia in the mouse.
     Toxicol.  2:31-42.

Aviado, D.M., S.  Zakhari, and T. Watanabe.  1977.  Non-fluorinated propellants
     and solvents for aerosols.  CRC Press, Cleveland.

Aviado, D.M.  1975.  Toxicity of aerosol propellants  in the respiratory and
     circulatory systems.   X.  Proposed Classification. Toxicol.  3:321-332.

Ballantyne, B., M F. Gazzard, and D.W. Swanston.  1976.  The ophthalmic toxi-
     cology of dichloromethane.  Toxicol. 6:173-187.


005DC2/B                            5-115                          12/22/83

-------
Barber, E.D., W.H.  Donish, and K.R.  Mueller.   1980.   Quantitative measurement
     of the mutagenicity of volatile liquids in the Ames Salmonella/microsome
     assay.  Abstract.

Barber, E.D., W.H.  Donish, and K.R.  Mueller.   1981.   A procedure for the
     quantitative measurement of volatile liquids in the Ames Salmonella/
     microsome assay.   Mutat.  Res.  90:31-48.

Balmer, M.F., F.A.  Smith, L.J. Leach, and C.L.  Yuille.  1976.  Effects in the
     liver of methylene chloride inhaled alone and with ethyl alcohol.  Amer.
     Ind.  Hyg. Assoc.  J. 37:345-352.

Barber, E.D., W.H.  Donish, and K.R.  Mueller.   1980.   Quantitative measurement
     of the mutagenicity of volatile liquids in the Ames Salmonella/mi crosome
     assay.  Abstract.   P-39,  Environmental Mutagen Society llth Annual Meeting.

Barber, E.D., W.H.  Donish, and K.R.  Mueller.   1981.   A procedure for the
     quantitative measurement of volatile liquids in the Ames Salmonella/
     microsome assay.   Mutation Research, 90:31-48.

Barrowcliff, D.F.,  and A.J. Knell.   1979.  Cerebral  damage due to endogenous
     chronic carbon monoxide poisoning caused by exposure to methylene chloride.
     J. Soc. Occup.  Med. 29:12-14.

Barrowcliff, D.F.  1978.  Chronic carbon monixide poisoning caused by methylene
     chloride paintstripper.  Med.  Sci. Law.  18(4):238.

Belej, M.A., D.G. Smith, and D.M. Aviado.  1974.  Toxicity of aerosol propel-
     lants in the respiratory and circulatory systems.  IV.  Cardiotoxicity  in
     the monkey.  Toxicol. 2:381-395.

Benzon, H.T. , L. Claybon, and E.A.  Brunner.  1978.  Elevated carbon monoxide
     levels from exposure to methylene chloride.  J. Am. Med. Assoc. 239(22):
     2341.

Berger, M., and G.G. Fodor.  1968.   CMS disorders under the  influence of air
     mixtures containing dichloromethane.  Zentralbl. Bakteriol.  (Ref.) 215:
     517.

Bergman, K.  1978.   Application of whole body autoradiography to  distribution
     studies of organic solvents.   Int. Symp. Control Air Pollut. Work Environ.
     (Proc.) Part 2, 128-139.

Bonventre, J., 0. Brennan, D.  Jason, A. Henderson, and M. L. Basots.  1977.
     Two deaths following accidental inhalation of dichloromethane and 1,1,1-tri-
     chloroethane.  J.  Anal. Toxicol. 1(4):158-160.

Bornmann, G. and A. Loeser.  1967.   Zur Frage winer Chronisch-Toxichen Wirkung
     von Dichlormethan.  Z. Lebensmittel-Unters.  Forsch.  136:  14-18.  (Read
     in translation.)

Bornschein,  R. L. , L. Hastings, and  J.  Manson.   1980.  Behavioral  toxicology  in
     the offspring of rats  following maternal exposure to dichloromethane.
     Toxicol. Appl. Pharmacol. 52:29-37.


005DC2/B                             5-116                          12/22/83

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

Burek, J.D., K.D.  Nitschke, T.J.  Bell, D.L. Wackerle, R.C.  Childs, J.E. Beyer,
     D.A.  Dittenber, L.W.  Rampy,  and M.J. McKenna.   1980.  Methylene Chloride:
     A Two-year Inhalation Toxicity and Oncogenicity Study in Rats and Hamsters.
     Final  Report.  Toxicology Research Laboratory,  Health and Environmental
     Sciences,  Dow Chemical Co.,  Midland, MI 48640.  Co-sponsored by Diamond
     Shamrock Corp., Dow Chemical Co., Imperial Chemical Industry Ltd., Stauffer
     Chemical Co.,  and Vulcan Materials Co..

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

Carlsson,  A., and M. Hultengren.   1975.   Exposure to methylene chloride. III.
     Metabolism of 14C-labelled methylene chloride in rat.   Scand. J. Work
     Environ. Health 1:104-108.

Clark, D.G. and D.J. Tinston.  1973.  Correlation of the cardiac sensitizing
     potential  of halogenated hydrocarbons with their physiochemical properties.
     Br.  J. Pharmacol.  49(2):355-357.

Collier,  H.  1936.   Methylene dichloride intoxication in industry -- a report
     of two cases.   Lancet 1:594-595.

Collison,  H.A.,  F.L. Rodkey, andJ.D. O'Neal.  1977.  Effect of dichloromethane
     on hemoglobin function.  Biochem. Pharmacol.  26:557-558.

Cox, C.R.   1972.   Regression model and life tables.  J.  Roy. Stat. Soc. B
     34:187-220.

Crump, K.S.  1979.   Dose-response problems in carcinogenesis.  Biometrics
     35:157-167.

Crump, K.S., H.A.  Guess,  and L.L. Deal.   1977.  Confidence intervals and test
     of hypotheses concerning dose-response relations inferred from animal
     carcinogenicity data.  Biometrics 33:437-451.

Crump, K.S., and W.W.  Watson.  1979.  GLOBAL79.  A Fortran program to  extrapo-
     late dichotomous animal carcinogenicity data to low dose.  Nat!.  Inst.  of
     Environ. Health Science Contract No. l-ES-2123.

Delzell,  E., and R.R.  Monson.  1982.  Mortality among rubber workers.  VI. Men
     with exposure to acrylonitrile.  J. Occup. Med. 24(10):767-769.

Dill, D.C., P.G.  Watanabe, and J.R. Morris.  1978.   Effect of methylene chloride
     on the oxyhemoglobin dissociation curve of rat and  human blood.   Toxicol.
     Appl.  Pharmacol.  46:125-129.

DiVincenzo, G.D.,  and M.L. Hamilton.  1975.  Fate and disposition of 14C-methy-
     lene choloride in the rat.  Toxicol. Appl. Pharmacol.  32:385-393.
005DC2/B                            5-117                          12/22/83

-------
DiVincenzo, G.D., F.J.  Yanno, and B.D. Astill.  1972.  Human and canine exposures
     to methylene chloride vapor.  Am. Ind. Hyg.  Assoc. J. 33:125-135.

Doll, R.   1971.   Weibull distribution of cancer.   Implications for models of
     carcinogenesis.   J. Roy. Stat.  Soc.  A 13:133-166.

Douglas,  B.H., J.S.  Wilkinson, R.F.  Williard, and W.L. Williams.  1976.
     Effect of dichloromethane on the blood pressure of spontaneously hyperten-
     sive rats (abstract).  IRCS Med. Sci. Libr.  Compend.  4:507.

Dow Chemical Company Report, 1979.

Dow Chemical Company.   1980.  Methylene chloride: a two-year inhalation toxicity
     and oncogenicity study in rats and hamsters.  FYI-OTS-0281-0097.  Follow-up
     response A.  Washington, D.  C.:   U.  S. Environmental Protection Agency, 0
     Office of Toxic Substances.

Dow Chemical Company.   1980.  Burek, J.D., K.D.  Nitschke,  T.J. Bell, D.L.
     Wackerle, R.C.  Childs, J.E. Beyer,  D.A.  Dittenber, L.W. Rampy, and M.J.
     McKenna.  Methylene chloride:   A two-year inhalation toxicity and oncogen-
     icity study in rats and hamsters.  Toxicology Research Laboratory, Health
     and Environmental  Sciences, Dow Chemical Company, Midland, MI (December
     31).  Unpublished.

Dow Chemical Company.   1982.  Nitschke,  K.D., T.J. Bell, L.W.  Rampy, and M.J.
     McKenna.  Methylene chloride:   A two-year inhalation toxicity and onco-
     genicity study in rats.  Toxicology Research Laboratory,  Health and
     Environmental Sciences, Dow Chemical Company, Midland, MI  (October 11,
     1983).

Dripps, R.D., J.E. Eckenhoff, and L.D. Vandam.  1977.  Introduction to Anes-
     thesia; the Principles of Safe Practice.  5th ed.  Philadelphia, PA:
     W.B. Saunders Co.,  pp. 121-123.

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

Environmental Criteria and Assessment Office.  1979.    Air Quality Criteria for
     Carbon Monoxide.   EPA-600/8-79-022, U.S. Environmental Protection Agency,
     Washington,  DC, October.

FASEB.   1974.  Biological data books.  2nd. ed., Vol.  III.  Edited by  Philip
     L. Altman and Dorothy S. Dittmen.  Federation of  American  Societies for
     Experimental Biology, Bethesda, MD.   Library of Congress No. 72-87738.

Filippova, L.M.   1967.   Genetic  activity of germinal systems.    Genetika 8:134-
     137.

Flury, F. , and F. Zernik.  1931.  Harmful Gases, Vapors,  Fogs,  Smoke  and Dust.
     Berlin:  Julius Springer, pp.   311-312.

Fodor, G.G., and A. Roscovanu.   1971.  Increased blood CO content in  humans
     and animals experimentally  exposed to industrial  solvent vapors  in England.
     HM  (ed.):  Proceedings of the  2nd International Clean  Air  Congress.  New
     York:  Academic Press, pp.   238-243.
005DC2/B                             5-118                          12/22/83

-------
 Fodor,  G.G.,  and A.  Roscovanu.   1976.   Increased  blood CO  content in humans
      and  animals caused  by  incorporated halogenated  hydrocarbons.   Zbl.  Bakt.
      Hyg.  162:34-40.

 Fodor,  G.G.,  and G.  Winneke.   1971.   Nervous  system  disturbances  in men  and
      animals  experimentally exposed  to  industrial  solvent  vapors  in England.
      Proceedings of  the  2nd International  Clean Air  Congress.   New York:
      Academic  Press,  pp. 238-43.

 Fodor,  G.G.,  D. Prajsnar, and  H.W. Schlipkoter.   1973.   Endogenous CO formation
      by incorporated halogenated  hydrocarbons  of  the methane  series.   Staub-
      Reinhalt  Luft.  33:260-261.

 Forster,  H.U., S. Graff, C.L.  Hake,  R.  Soto,  and  R.D.  Stewart.  1974.
      Pulmonary - Hematologic studies  on humans during exposure  to methylene
      chloride.  Report No.  N10SH-MCOW-ENVM-MC-74-4.   The Medical  College of
      Wisconsin, Dept. of Environmental  Medicine,  Milwaukee, Wis.,  17 pp,
      Apri1.

 Freireich, E.J., E.A. Gehan, D.P. Rail,  L.H.  Schmmidt and  H.E.  Skipper.   1966.
      Quantitative comparison of toxicity of anticancer agents in  mouse,  rat,
      hamster, dog, monkey,  and man.   Cancer Chemother.  Rep.   50:  219.

 Friedlander, B.R., F.T. Hearne, and  S.  Hall.   1978.   Epidemiologic investiga-
      tion of employees chronically exposed to  methylene chloride.   J.  Occup.
      Med.  20:657-666.

 Gamberale, F., B.A.  Annwall, and M.  Hultengren.   1975.   Exposure  to methylene
      chloride -- II.  Psychological  functions.  Scand.  J.  Work. Environ.
      Health 1(2):95-103.

 Gocke,  E., M.T. King, K.  Eckhardt, and  D. Wild.   1981.   Mutagenicity of  cos-
      metics ingredients licensed by  the  European  communities.   Mutat.  Res.
      90:91-109.

 Gradiski,  D. , J.L.  Megadur, M.  Baillot,  M.C. Daniere,  and  M.B.  Schuh.  1974.
      Comparative toxicity of the principle chlorinated  aliphatic  solvent. J.
      Eur.  Toxicol.  7:247.

Green, T.   1980.   The metabolism and mutagenicity of  methylene  chloride.
     Abstracts of papers, Society of Toxicology,  Inc.  19th annual  meeting,
     Wash., D.C., March 9-13, 1980.

Green, T.   1981.   The metabolic activation of dichloromethane and  chlorofluoro-
     methane in a bacterial  mutation assay using  Salmonella typhimurium.
     Unpublished Manuscript.  Imperial Chemical Industries, PLC,  Central
     Toxicology Laboratory,  Alderly Park, Macclesfield,  Cheshire  SK104,  United
     Kingdom.

Hake, C.L., R.D.  Stewart, H.U.  Forster,  A.J. Lebrun,  J.E.  Peterson,  and  A. Wu.
     1974.  Results of the controlled exposure of human  females to  the vapor
     of methylene chloride.   Report.  No. N10SH-MCOW-ENVM-MC-74-3.   Milwaukee,
     Wis.   The Medical College of Wisconsin, Dept. of Environmental  Medicine,
     22 pp, March.


005DC2/B                            5-119                           12/22/83

-------
Hamada, N.  and R.E.  Peterson.   1977.   Effect of chlorinated aliphatic hydrocar-
     bons on excretion of protein and electrolytes by rat pancreas.  Toxicol.
     Appl.  Pharmacol.  39:185.

Hardin, B.D., and J.M. Manson.   1980.  Absence of dichloromethane teratogenicity
     with inhalation exposure in rats.  Toxicol.  Appl. Pharmacol.  52:22-28.

Harms, J.S., R.E. Peterson, J.M.  Fujimoto, and C.P. Erwin.  1976.  Increased
     "bile duct-pancreatic fluid" flow in chlorinated hydrocarbon-treated
     rats.  Toxicol.  Appl. Pharmacol.   35:41-49.

Haun, C.C., E.S.  Harris, and K.I. Darmer.  1971.   Continuous animal exposure
     to methylene chloride.  Am.  RL-TR-71-120, #10.   IN: Proceed. 2nd Conf.
     Environ. Toxicol.  Wright-Patterson AFB, Ohio,   pp 125-135.

Haun, C.C., E.H.  Vernot, K.I.  Darmer, Jr., and S.S. Diamond.  1972.  Continuous
     animal exposure to low levels of dichloromethane AMRL-TR-.130, paper No.
     12.   In:  Proceedings of the 3rd Annual Conference on Environmental
     Toxicology,  Wright-Patterson Air Force Base, Ohio, Aerospace Medical
     Research Laboratory, pp.  199-208.

Hearne, F.T., and B.R. Friedlander.   1981.  Follow-up of methylene chloride
     study.  J.  Occup. Med. 23:660.

Heppel, L.A., and P.A. Neal.   1944.   Toxicology of dichloromethane (methylene
     chloride) -- II.   Its effect on running activity in the male rat.  J.
     Ind. Hyg. Toxicol. 26:17-21.

Heppel, L.A., P.A. Neal, T.L.  Perrin, M.L. Orr, andV.T. Porterfield.   1944.
     Toxicology of dichloromethane (methylene chloride).  I. Studies on effects
     of daily inhalation.  J.  Ind. Hyg. Toxicol.  26:8-16.

Hogan, G.K., R.G. Smith, and H.H. Cornish.  1976.  Studies on the microsomal
     conversion of dichloromethane to carbon monoxide.  Toxicol. Appl.  Pharmacol
     37:112.

Hughes, J.P.  1954.   Hazardous exposure  to some  so-called safe  solvents.   J.
     Med. Assoc.  156:234-237.

International Commission on Radiological  Protection (ICRP).  1977.   Recom-
     mendation of the  International  Commission on  Radiological  Protection.
     Publ.  No. 26, adopted Jan.  17,  1977.  Pergamon Press, Oxford, England.

Johnson, E.M.  1981.   Screening  for  teratogenic  hazards,  nature of the  problems.
     Ann. Rev. Pharmacol. Toxicol.   21:417-429.

Jongen, W.M.F., G.M. Alink, and  J.H.  Koeman.   1978.   Mutagenic  effect of
     dichloromethane on  Salmonella typhimurium.  Mutat.  Res. 56:245-248.

Jongen, W.M.F.,  E.G.M. Harmson,  G.M.  Alink, and  J.H.  Koeman.  1982.   The
     effect of glutathione conjugation  and microsomal oxidation on the  muta-
     genicity of dichloromethane in  S^_  typhimurium.   Mutat.  Res.  95:183-189.
005DC2/B                             5-120                           12/22/83

-------
Jongen, W.M.F., P.H.M.  Lohman, M.J.  Kettenhagen, G.M.  Alink, F. Berends, and
     J.H.  Koeman.   1981.   Mutagenicity testing of dichloromethane in short-term
     mammalian test systems.   Mutat.  Res. 81(2): 203-213.

Kahn, A.,  R.B. Rutledge,  G.L.  Davis,  J.A. Altes, G.E.  Gantner, C.A.  Thornton,
     and N.D.  Wallace.   1974.   Carboxyhemoglobin sources in the metropolitan
     St. Louis population.   Arch.  Environ. Health 29:127-135.

Kanada, T.,  and M.  Uyeta.   Mutagenic screening of organic solvents in microbial
     systems.   Mutat.  Res.  54:215, 1978.  (Abstract)

Kim, N.K.   Air pollution evaluations using risk assessment methodology.  1981.
     J. Air Pollut. Control  Assoc.  31(2): 120-122.

Kimura, E.T.,  D.M.  Ebert,  and P.W. Dodge.  1971.  Acute toxicity and limits of
     solvent residue for sixteen organic solvents.  Toxicol. Appl. Pharmacol.
     19:699-704.

Klaassen,  C.D. and G.L.  Plaa.   1966.   The relative effects of  various chlori-
     nated hydrocarbons on liver and kidney function in mice.  Toxicol.  Appl.
     Pharmacol. 9:139-131.

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.

Kluwe, W.M., F.W.  Harrington,  and S.E. Cooper.  1982.   Toxic effects of organ-
     ohalide compounds on renal tubular cells i_n vivo and J_n vitro.  J. Pharm.
     Exper.  Therap. 230(3):  597-603.

Kubic, V.L., and M.W.  Anders.   1975.   Metabolism of dihalomethanes to carbon
     monoxide.  II. In vitro studies.  Drug Metab. Dispos.   3:104-112.

Kuzelova,  M.,  and R. Vlasak.   1966.   The effect of methylene dichloride on the
     health of workers in production of film foils and investigation of formic
     acid as a methylene dichloride metabolite.  Prac. Lek.  18:167-70.

Kuzelova,  M. ,  J.  Cerny, S.  Havova, M. Hub, V. Kunor, and A.  Popler.  1975.
     Lethal  methylene chloride poisoning with severe chilblains  (Czech.).
     Prac.  Lek. 27(9):317-319.

Little, 1980.

Loyke, H.F.   1973.   Methylene chloride and chronic renal hypertension.  Arch.
     Pathol. 95:130-131.

MacEwen, J.D., E.H. Vernot,  and C.C.  Haun.  1972.  Continuous  animal exposure
     to dichloromethane.   AMRL-TR-72-28, Systems Corporation Report No. W-71005.
     Wright-Patterson Air Force Base, Ohio, Aerospace Medical  Research.

MacEwen, J.D., and E.H.  Vernot.  1971.  Toxic hazards research unit annual
     technical report.   U.S.N.T.I.S.  AD Rep. Issue No. 73:543.
005DC2/B                            5-121                           12/22/83

-------
Mantel, N. ,  and M.A. Schneider-man.  1975.  Estimating "safe" levels:  a hazar-
     dous undertaking.   Cancer Res. 35:1379-1386.

March, J.  1977.  Advanced Organic Chemistry:  Reactions, Mechanisms, and
     Structure.  2nd ed.  N.Y.:  McGrawHill, pp. 341-343.

Matthews, R.M., D.H. Martin, I.D. Kuntz and T.L.  James.  1978.  Antisickling
     behavior of dichloromethane i_n vitro.

Matthews, R.M., D.H. Martin, I.D. Kuntz and T.L.  James.  1977.  Antisickling
     behavior of dichloromethane i_n vitro (abstract).  Blood 50(5):113.

McGregor, D.B.  1979.  Practical experience in testing unknowns i_n vitro.
     Chapter 4.  Pages 53-71 in G.E. Paget, ed.  Topics in Toxicology.  Muta-
     genesis in submammalian systems:   Status and significance.  Baltimore,
     MD:  University Park Press.

Morris, J.B., F.A.  Smith and R.H. Carman.  1979.   Studies on methylene chloride
     induced fatty liver.   Exp. Mol. Pathol.  30:386-393.

Moskowitz, S. , and H. Shapiro.   1952.   Fatal  exposure to methylene chloride
     vapor.   Arch.  Ind.  Hyg. Occup. Med.  6:116-123.

MAS (National Academy of Sciences).  1977.  Drinking Water and Health.  Safe
     Drinking Water Committee,  Advisory Center on Toxicology, Assembly of  Life
     Sciences, National  Research Council, National Academy of Sciences,
     Washington, DC, p.  939. Available from:   Printing and Publishing Office,
     National Academy of Sciences, 2101 Constitution Ave., Washington, DC
     20418.

National  Coffee Association.  1982a.  24-month chronic toxicity and  onco-
     genicity study of methylene chloride in rats.  Final report.  Prepared by
     Hazelton Laboratories America, Inc.  , Vienna, VA.  (August 11, 1982).
     Unpublished.

National  Coffee Association.  1982b.  24-months chronic toxicity and onco-
     genicity study of methylene chloride in rats.  Addition to the  final
     report.  Prepared by Hazelton Laboratories America,  Inc., Vienna, VA
     (Nov. 5, 1982).  Unpublished.

National  Institute for Occupational Safety and Health.  1972.  Criteria  for a
     recommended standard for occupational exposure to carbon monoxide.  HSM
     78-11000.  NIOSH.

NIOSH  (76-138): 1976.  Criteria for a recommended standard  .  . . Occupational
     exposure to methylene chloride, U.S. Government Printing Office, Washington,
     D. C. 166 pp.

National  Toxicology Program (NTP).  1982.  Draft technical  report  on the
     carcinogenesis bioassay of dichloromethane  (methylene  chloride),  gavage
     study.   Research Triangle Park, NC,  and Bethesda, MD.

Nestmann, E.R., E.G.-H.  Lee, T.I. Matula, G.R. Douglas, and  J.C. Mueller.
     1980.  Mutagenicity of constituents  identified  in pulp  and paper  mill
     effluents  using the Salmonella/mammalian microsome assay.  Mutat.  Res.
     79:203-212.
005DC2/B                            5-122                           12/22/83

-------
Nestmann, E.R., R. Otson, D.7. Williams, and D.J. Kowbel.  1981.  Mutagenicity
     of paint removers containing dichloromethane.  Cancer Lett. 11:295-302.

Norpoth, K., U. Witting, M.  Springorum and C. Witting.  1974.   Induction of
     microsomal enzymes in the rat liver by inhalation of hydrocarbon solvents.
     Int. Arch. Arbeitsmed.   33(4): 315-321.

Ott, M.G.,  L.K. Skory, B.B.  Holder, J.M. Bronson, and P.R. Williams.  1980.
     Health Surveillance of Employees Occupationally Exposed to Methylene
     Chloride.  I. Mortality.   Scand. J. Work Environ. Health (in press),
     1980a.   Cited in Burek et al..

Ott, M.G.,  L.K. Skory, P.R.  Williams, J.M. Bronson and B.B. Holder.  1980.
     Health Surveillance of Employees Occupationally Exposed to Methylene
     Chloride.  II. Morbidity.  Scan. J. Work Environ. Health (in press),
     1980b.   Cited in Burek et al..

Ott, M.G.,  L.K. Skory, B.B.  Holder, J.M. Bronson, and P.R. Williams.  1980.
     Health evaluation of employees Occupationally exposed to methylene chlo-
     ride.   General Study Design and Environmental Considerations.  Scand.  J.
     Work Environ. Health 9:1-7.

Ott, M. G. ,  L. K.  Skory, B.  B. Holder, J. M. Bronson, and P. R. Williams.
     1980.  Health evaluation of employees occupationally exposed to methylene
     chloride. Clinical laboratory evaluation.  Scand. J. Work  Environ. Health
     9(Suppl.  1):  17-25.

Ott, M. G.,  L. K.  Skory, B.  B. Holder, J. M. Bronson, and P. R. Williams.
     1980.  Health evaluation of employees occupationally exposed to methylene
     chloride. Mortality.   Scand. J.  Work Environ. Health 9(Suppl. 1):8-16.

Ott, M. G.,  L. K.  Skory, B.  B. Holder, J. M. Bronson, and P. R. Williams.
     1980d.  Health evaluation of employees occupationally exposed to methylene
     chloride. Twenty-four hour electrocardiographic monitoring.  Scand. J.
     Work Environ. Health 9(Suppl.  1):26-30.

Ibid.  1983.   Health evaluation of employees occupationally exposed to methy-
     lene chloride.  Mortality.  Scand.  J. Work Environ. Health 9:8-16.

Ibid.  1983.   Health evaluation of employees occupationally exposed to methy-
     lene chloride.  Clinical  Laboratory Evaluation.  Scand. J. Work Environ.
     Health 9:17-24.

Ibid.  1983.   Health evaluation of employees occupationally exposed to methy-
     lene chloride.  Twenty-four hour electrocardiographic monitoring.  Scand.
     J. Work Environ.  Health 9:26-30.

Ibid.  1980.   Health evaluation of employees occupationally exposed to methy-
     lene chloride.  Metabolism data and oxygen half-saturation pressures.
     Scand.  J. Work Environ. Health 9:31-38.

Pankow, D.,  R. Gutewort, W.  Glatzel,  and K. Tietze.  1979.  Effect of dichloro-
     methane on the sciatic motor conduction velocity of rats.  Experentia
     35:373-374.
005DC2/B                            5-123                          12/22/83

-------
Perocco, P., and G.  Prodi.   1981.   DNA damage by haloalkanes in human lympho-
     cytes cultured iji vitro.   Cancer Lett. 13:213-218.

Peterson, J.  1978.   Modeling the uptake, metabolism and excretion of dichloro-
     methane by man.   Am.  Ind.  Hyg.  Assoc. 39(1):41-47.

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.

Price, P. J.,  C. M.  Hassett, and J.  I. Mansfield.  1978.  Transforming activi-
     ties of trichloroethylene and proposed industrial alternatives.  In Vitro
     14:290-293.

Pryor, G. T.,  R. A.  Howd,  R. Malik,  R. A. Jensen and C. S. Robert.  1978.
     Biomedical studies on the effects of abused inhalant mixtures.  Quart.
     Prog.  Rep. #7.  Contract #271-77-3402.  Rockville, MD.  Nat.  Inst. Drug.
     Abuse.

Putz, V. R., B. L.  Johnson and J.  V.  Setzer.  1976.  A comparative study of
     the effect of carbon monoxide and methylene chloride on human performance.
     J.  Environ. Pathol.  Toxicol.  2:97-112.

Rail, D.P.   Difficulties in extrapolating the results of toxicity studies  in
     laboratory animals to man.  1969.  Environ. Res.  2: 360-367.

Rapson,  W.H.,  M.A.  Nazar,  and V.V. Butsky.  1980.  Mutagenicity produced by
     aqueous chlorination of organic compounds.  Bull. Environ. Contam. Toxicol.
     24:590-596.

Ratney,  R.  S. ,  D.  W.  Wesman, and H.  B. Elkins.  1974.  Iji vivo conversion  of
     methylene chloride to carbon monoxide.  Arch. Environ. Health 28:223-226.

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

Reynolds, E. S., and A. G.  Yee.  1967.  Liver parenchymal cell  injury.  V.
     Relationships between patterns of chloromethane C14  incorporation into
     constituents of liver i_n vivo and cellular  injury.   Lab.  Invest. 16:591-
     603.

Riley, E. C.,  D. W.  Fassett, and W.  L. Sutton.  1966.  Methylene  chloride
     vapor in expired air of human subjects.  Amer. Ind.  Hyg.  Assoc. J. 27:341-
     348.

Roth, R. P., Lo. K.  Jai-Ruey, and R. T. Drew.  1973.   Dichloromethane  inhalation
     and drug metabolizing enzymes - the effect of chemical treatment with
     mixed function oxidase inducing and  inhibiting agents on  carboxyhemoglobin
     formation.  AMRL-TR-73-125, Paper No. 22.  J_n:  Proceedings  of the 4th
     Annual Conference on Environmental Toxicology, Wright-Patterson Air Force
     Base, Ohio, Aerospace Medical Research  Laboratory, pp. 279-289.
005DC2/B                            5-124                           12/22/83

-------
 Sahu,  S.C.,  and O.K.  Lowthei.   1981.   Pulmonary  reactions  to  inhalation of
     methylene chloride:   Effects  on  lipid  peroxidation  in rats.   Tox.  Lett.
     8:253-256.

 Samoiloff, M.R., S. Schulz,  Y.  Jordan,  K. Denich,  and  E. Arnott.   1980.   A
     rapid simple  long-term  toxicity  assay  for aquatic contaminants  using the
     nematode Panagrellus  redivivus.   Can.  J. Fish Aquat.  Sci.  37:1167-1174.

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

 Scharf, S. M., M.  D.  Thames  and R.  K.  Sargent.   1974.  Transmural  myocardial
     infarction after exposure  to  carbon monoxide  in coronary artery disease:
     Report  of a case.  New  Engl.  J.  Med. 291:85-86.

 Schutz, E.   1958.  Effect  of polyethylene glucol 400 on  percutaneous absorption
     of active ingredients.  Arch.  Exp. Pathol.  Pharmakol.  323:237-238.

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

 Settle, W.   1975.  Role of changes  in  carbon  monoxide-hemoglobin binding in
     methylene chloride toxicity.   Fed. Proc. 34:229.

 Simmon, V.F.  1978.  Structural  correlations of carcinogenic and mutagenic
     alkyl halides.   Pages 163-171  in  I.M.  Asher and C.  Zervos, eds.  Structural
     correlates of carcinogenesis  and  mutagenesis.   A  guide to testing  priori-
     ties?   HEW Publication  No.  (FDA)78-1046.

 Simmon, V.F., and K.  Kauhan'en.   1978.   In vitro microbological  mutagenicity
     assays  of 2-chlorethyl  chloroformate.  Final  report,  Contract No.
     68-03-11-74.  Prepared  for  U.S.  Environmental  Protection  Agency,  National
     Environmental Research  Center, Water Supply Laboratory,  Cincinnati,  OH
     45268.

 Simmon, V. F., V. Kauhanen,  and  R.  G.  Tardiff.  1977.  Mutagenic activity of
     chemicals identified  in drinking water.  Dev.  Toxicol. Environ.  Sci.
     2:249-258.

 Skory,  L.K.   1980.   Health Surveillance of  Employees Occupationally  Exposed to
     Methylene Chloride.  V.   A  Review  of Effects  on Oxygen Transport.   Scand.
     J. Work. Environ. Health (in press), 1980.   Cited in  Burek et al.

 Skory,  L.K.,  M.G.  Ott, P.R. Williams, J.M. Bronson  and B.B. Holder.   1980.
     Health  Surveillance of  Employees Occupationally Exposed  to Methylene
     Chloride.   III.  Clinical Pathological  Evaluation.   Scand. J.  Work  Environ.
     Health  (in press),  1980a.   Cited  in Burek et  al.

Skory,  L.K.,  M.G.  Ott, P.R. Williams, J.M. Bronson  and B.B. Holder.   1980.
     Health  Surveillance of  Employees Occupationally Exposed  to Methylene
     Chloride.   IV. 24-Hour  EKG Monitoring.    Scand. J. Work Environ.  Health
     (in press),  1980b.  Cited in Burek  et al.

005DC2/B                            5-125                          12/22/83

-------
Snow, L.,  P.  McNair, and  B.C. Castro:  1979.  Mutagenesis testing of methylene
     chloride and 1,1,1-trichloroethane in Salmonella strains TA-100 and
     TA-98.   Personal Communication from Northrop Services, Inc., P.O. Box
     12313,  Research Triangle Park, NC, 27709, September 19

Stewart,  R.  D.,  and C.  L.  Hake.   1976.  Paint remover hazard.  J. Med. Assoc.
     235(4):398-401.

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.,  E.  D. Baretta, L. R. Platt, E. B. Stewart, J. H. Kalbfleish,
     B.  Van  Yserloo, and A.  A. Rimm.  19474.  Carboxyhemoglobin levels in
     American blood donors.   J.  Am. Med.  Assoc. 229:1187-1195.

Stewart,  R.  D.,  H.  V. Forster, C. L. Hake, A. J.  Lebrun, and J. E. Peterson.
     1973.   Human responses to controlled exposure of methylene chloride
     vapor.   Report No. NIOSH-MCOW-ENVM-MC-73-7.  Milwaukee, WI, Dept. Environ-
     mental  Medicine, December.   82 p.

Stewart,  R.  D.,  T.  N. Fisher, M.  J. Hosko, J. E.  Peterson, E. D. Baretta, and
     H.  C.  Dodd.   1972a.   Carboxyhemoglobin elevation after exposure to dichlo-
     romethane.  Science 176(4032):295-296.

Stewart,  R.  D.,  T.  N. Fisher, M.  J. Hosko, J. E.  Peterson, E. D. Baretta, and
     H.  C.  Dodd.   1972b.   Experimental human exposure to methylene chloride.
     Arch.  Environ. Health 25:342-348.

Stokinger,  H.E.  and R.L.  Woodward.  1958.  Toxicologic methods for establishing
     drinking water standards.  J. Am. Water Works Assoc., April.

Su, G.  and K.A.  Wurzel.  1981.  A regulatory framework for setting air emission
     limits  for noncriteria pollutants.  J. Air Pollut. Control Assoc.  31(2):
     160-162.

Svirbely,  J.  L.,  B. Highman, W.  C. Alford, and W. F. Von Oettingen.  1974.
     The toxicity and narcotic action of mono-chloro-mono-bromo-methane with
     special  reference to inorganic and volatile  bromide in blood, urine and
     brain.   J.  Ind. Hyg.  Toxicol. 29:383-389.

Taylor,  G.  J., R. T. Drew, E. M.  Lores and T. A.  Clemmer:  1976.  Cardiac
     depression by haloalkane propellents, solvents, and inhalation anesthetics
     in rabbits.  Toxicol.  Appl.  Pharmacol. 38:379-387.

Theiss,  J.  C., G. D. Stoner, M.  B. Shimkin, and E. K. Weisburger.  1977.  Test
     for carcinogenicity of organic contaminants  of United States drinking
     waters  by pulmonary tumor response in strain A mice.  Cancer Res. 37:
     2717-2720.

Thilagar,  A.K.  and V. Kimaroo.  1983.  Induction  of chromosome  damage by
     methylene chloride in CHO cells.  Mut. Res.  116: 361-367.
005DC2/B                            5-126                           12/22/83

-------
Thomas, A. A., M. K. Pinkert^n, and J. A. Warden.   1972.   Effects  of  low level
     dichloromethane exposure on the  spontaneous  activity  of  mice.  AMRL-TR72-130,
     paper No. 14.  In:   Proceedings  of  the  3rd Annual  Conference  on  Environmental
     Toxicology, Wright-Patterson Air Force  Base, Ohio,  Aerospace  Medical
     Research Laboratory, pp. 223-227.

Thomas, A. A., M. K. Pinkerton, and J. A. Warden.   1971.   Effects  of  methylene
     chloride exposure on the spontaneous activity  of mice.   AMRL-TR-H-120.
     Paper No. 13.  In:   Proceedings  of  the  2nd Annual  Conference  on  Environ-
     mental Toxicology, Wright-Patterson Air Force  Base, Ohio,  Aerospace
     Medical Research Laboratory,  p.  185-189.

U.S. Environmental  Protection Agency.  1976.   Interim procedures and  guidelines
     for health  risk and  economic impact assessments of  suspected  carcinogens.
     Federal Register 41:21402 (May 25).

U.S. Environmental  Protection Agency.  1978.   Proposed  Guidelines  for Regis-
     tering Pesticides in the United  States.   Fed.  Regist. 43(163):37382-37888,
     August 22.

U.S. Environmental  Protection Agency.  1979.   Halomethanes:   Ambient  Water
     quality criteria.   NTIS PB-296797.  National Technical Information  Service,
     Springfield, VA.

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

U.S. Environmental  Protection Agency.  1979.   Proposed Health Effects Test
     Standards for  Toxic Substances Control  Act Test Rules and  Proposed  Good
     Laboratory  Practice Standards for Health  Effects.   Fed.  Regist.  44(145):
     44089-44092, July 26. '                                           ~

U.S. EPA (U.S. Environmental Protection  Agency).  1980a.   Ambient  Water  Quality
     Criteria for Halomethanes.   U.S.  EPA.   Available from:   National  Technical
     Information Service, Springfield, VA (NTIS PB81-117624).

U.S. EPA (U.S. Environmental Protection  Agency).  1980b.   Guidelines  and
     Methodology for the Preparation  of  Health Effect Assessment Chapters of
     the Ambient Water Quality Criteria  Documents.  U.S. EPA, Environmental
     Criteria and Assessment Office;  Office  of Health and  Environmental  Assess-
     ment; Office of Research and Development, Cincinnati, OH,  November  28.

U.S. EPA (U.S. Environmental Protection  Agency).  1981.  Advisory  Opinion for
     Dichloromethane (Methylene Chloride).   Office of Drinking  Water,  U.S.
     EPA,  Washington,  DC.  Draft.

Ugazio, G., E. Budino,  0. Danni  and Milillo:  Hepatotoxicity  and lethality of
     haloalkanes.   1973.   Biochem.  Soc.  Trans. 1:968-972.

Valencia,  R.:   Contract report to FDA, 1978.
005DC2/B                            5-127                          12/22/83

-------
Von Oettingen, W. F., C. C. Powell, N. E. Sharpless, W. C. Alford, and L. J.
     Pecora.  1949.   Relation between the toxic action of chlorinated methanes
     and their chemical and physiochemical properties.  NIH Bulletin 191.

Weinstein, R. S. , and S. S. Diamond.  1972.  Hepatotoxicity of dichloromethane
     (methylene chloride) with continuous inhalation exposure at a low-dose
     level.  AMRL-72-130, paper No. 13.  In:  Proceedings of the 3rd Annual
     Conference on Environmental Toxicology, Wright-Patterson Air Force Base,
     Ohio, Aerospace Medical Research Laboratories, pp. 209-222.

Weinstein, R. S., D. D. Boyd., and K. C. Back.  1972.  Effects of continuous
     inhalation of dichloromethane in the mouse—morphologic and functional
     observations.  Toxicol. Appl.  Pharmacol. 23:660-79.

Weiss, G.  Toxic encephalosis as an occupational hazard with methylene chloride.
     1967.  Zentrabl.  Arbeitsmed. 17:282-85

Wilkinson, J. S. , B. H. Douglas, R. F. Williard and W. L. Williams:  Reduction
     of blood pressure in the spontaneously hypertensive rat (SHR) with methylene
     chloride (CH2CL2) (Abstract).   Clin. Res. 25:48A, 1977.

Winneke, C.L., W.D.  Collom, and F.  Esposito.  1981.  Accidental methylene
     chloride fatality.  Forensic Sci. Intern. 18:165-168.

Winneke, G.  1974.  Behavioral effects of methylene chloride and carbon monoxide
     as assessed by sensory and psychomotor performance.  In:  C. Xintaras, B.
     L. Johnson, and I. DeGroot, eds., Behavioral Toxicology—Early Detection
     of Occupational Hazards,  Publication HEW No. (NIOSH) 74-126.  U.S. Dept.
     of Health, Education and Welfare, Public Health Service, Center for
     Disease Control, National Institute for Occupational Safety and Health.
     p. 130-144.

Winneke, G. and G. G.  Fodor:  Dichloromethane produces narcotic effects.
     Occup. Health Safety.   Mar/Apr:34-37, 1976.

Zakhari, S.  1977.  Methylene chloride.  Chapter 2  In: Non-fluorinated propellants
     and solvents for aerosols.  L. Golberg, ed.  CRC Press, Cleveland.
005DC2/B                            5-128                           12/22/83

-------
                                    APPENDIX

             COMPARISON OF RESULTS BY VARIOUS EXTRAPOLATION MODELS

     The estimates of unit risk based on animal  studies presented in the body
of this document are all calculated by the use of the linearized multistage
model.  The reasons for its use have been detailed herein.   Essentially, this
model is part of a methodology that estimates a conservative linear slope at
low extrapolation doses and is consistent with the data at  all  dose levels of
the experiment.  It is a nonthreshold model  which holds that the upper-limit of
risk predicted by a linear extrapolation to low levels of the dose-response
relationship is the most plausible upper limit for the risk.
     Other models have also been used for extrapolation,  and include the
three nonthreshold models presented here:  the one-hit, the log-Probit,  and
the Weibull.  The one-hit model is characterized by a continuous downward
curvature, but is linear at low doses.  It can be considered the linear  form or
first stage of the multistage model because of its functional form.  Because of
this and its downward curvature, the one-hit model  will always  yield estimates
of low-level risk that are at least as large as those of the multistage  model.
Further, whenever the data can be fitted adequately by means of the one-hit
model, estimates from the two procedures will be comparable.
     The other two models, the log-Probit and the Weibull,  are  often used to
fit toxicological data in the observable range,  because of  their general "S"
curvature.  The low-dose upward curvatures of these two models  usually yield
lower low-dose risk estimates than those of the one-hit or  multistage model.
     The log-Probit model was originally proposed for use in problems of
biological assay, such as the assessment of potency of toxicants and drugs,
and has usually been used to estimate such values as percentile lethal dose
or percentile effective dose.  Its development was strictly empirical, i.e.,
it was observed that several  log dose-response relationships followed the
cumulative normal probability distribution function.  In  fitting the cancer
bioassay data, assuming an independent background,  this becomes:

          P(D;a,b,c) = c + (1-c)  *  (a+blogioD)  a,b > o £ c < 1

                                    A-l

-------
where P is the proportion responding at dose D,  c  is an  estimate  of the
background rate, a is an estimate of the standarized mean  of individual
tolerances, and b is an estimate of the log dose-Probit  response  slope.
     The one-hit model  arises from the theory that a single molecule of a
carcinogen has a probability of transforming a single noncarcinogenic cell
into a carcinogenic one.  It has the probability distribution function:

                    P(D;a,b) = l-exp-(a+bd)   a,b  > 0

where a and b are the parameter estimates.   The  estimate a represents the
background or zero dose rate, and the parameter  estimated  by b represents
the linear component or slope of the dose-response model.   In discussing the
added risk over background, incorporation of Abbott's correction  leads to

                    P(D;b) = l-exp-(bd)  b > 0

Finally, a model from the theory of carcinogenesis arises  from the multihit
model applied to multiple target cells.  This model has  been termed here the
Weibull model.  It is of the form
                       P(D;b,k) = l-exp-(bdk)   b,k > 0
For the power of dose only, the restriction k > 0 has been placed on this model.
When k > 1, this model yields low-dose estimates of risks usually significantly
lower than either the multistage or one-hit models, which are linear at low
doses.  All three of these models usually project risk estimates significantly
higher at the low exposure levels than those from the log-Probit.
     The estimates of added risk for low doses for the above models are given
in Table A-l for the DCM inhalation study.  Both maximum likelihood estimates
and 95% upper confidence limits are presented.  All estimates incorporate
Abbott's correction for independent background rate.
                                    A-2

-------
TABLE A-l.  ESTIMATES OF LOW-DOSE  RISK  TO HUMANS BASED ON SALIVARY GLAND REGION SARCOMAS IN MALE RATS
       IN THE DOW CHEMICAL  COMPANY (1980) INHALATION STUDY DERIVED FROM FOUR DIFFERENT MODELS
          (All estimates Incorporate Abbott's correction for Independent background rate.)*
Dose
ug/m3
Males
1 ug/m3
10 ug/m3
100 ug/m3
1000 ug/m3
10,000 ug/m3
* Response fit
Maximum likelihood estimates of
additional risks
Multistage One-hit Wei bull Log-Prob1t Multistage
model model model model model

2.2 x 10-8 1.3 x 10-7
2.2 x 10-7 1.3 x 10-6
2.2 x 10-6 1.3 x 10-5
2.2 x 10-5 1.3 x 10-*
2.2 x 10-* 1.3 x 10-3
against human equivalent dosages
Human equivalent dose ug/m3

0
1.2 x 105
3.6 x 105
8.4 x 105

2.6 x 10-11 o 1.8 x 10-7
1.1 x 10-9 o 1.8 x 10-6
5.0 x 10-8 1.1 x 10-16 1.8 x 10'5
2.2 x 10-6 5.0 x 10-H 1.8 x 10'*
9.5 x 10-5 1.7 x 10-6 1.3 x 10'3
as presented In Table 5-31.
Tumors/Total
1/93
0/94
5/91
11/88
95% upper confidence limit of
additional risks
One-hit Welbull Log-Prob1t
model model model

2.0 x 10-7 4.8 x 10-1° 3.5 x 10-31
2.0 x 10-6 1.7 x 10-8 1.6 x 10-22
2.0 x 10-5 6.1 x 10-6 2.5 x 10-15
2.0 x 10-* 2.0 x 10-5 1.3 x 10'9
2.0 x 10-3 6.1 x 10-* 2.4 x 10-1




-------