United States
                Environmental Protection
                Agency
                Office of Health and
                Environmental Assessment
                Washington DC 20460
EPA-600/8-84-004A
March 1984
External Review Draft
                Research and Development
?,EPA
Health Assessment
Document for
Chloroform

Part 1  of 2
Review
Draft
(Do Not
Cite or Quote)
                               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.

-------
                                     ERRATA
     The attached pages are to be substituted  for  the  corresponding  pages  in
the Health Assessment Document for Chloroform  (September 1985)  in  which  certain
key individuals'  names were inadvertently omitted.  Most notably,  that of  one
of the principal  authors, Or. Jean C.  Parker,  was  not  included.
     In addition, several typographical  errors have  been corrected in the  front
matter of the document.

-------
                                    PREFACE

     The Office of Health and Environmental Assessment has prepared this
health assessment to serve as a "source document" for EPA use.   This health
assessment document was developed for use by the Office of Air  Quality
Planning and Standards to support decision-making regarding possible
regulation of chloroform as a hazardous air pollutant.  However, the scope of
this document has since been 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 in order to quantitatively identify the toxicity of
chloroform and related characteristics.  Observed effect levels and other
measures of dose-response relationships are discussed, where appropriate, to
place the nature of the health responses in perspective with observed
environmental levels.
     Any information regarding sources, emissions, ambient air  concentrations,
and public exposure has been included only to give the reader a preliminary
indication of the potential presence of this substance in the ambient air.
While the available information is presented as accurately as possible, it is
acknowledged to be limited and dependent in many instances on assumption
rather than specific data.  This information is not intended, nor should it be
used, to support any conclusions regarding risks to public health.
     If a review of the health information indicates that the Agency should
consider regulatory action for this substance, a considerable effort will be
undertaken to obtain appropriate information regarding sources, emissions, and
ambient air concentrations.  Such data will provide additional  information for
drawing regulatory conclusions regarding the extent and significance of public
exposure to this substance.
                                      TM

-------
                              TABLE OF CONTENTS



LIST OF TABLES	  x

LIST OF FIGURES	xiv

AUTHORS, CONTRIBUTORS, AND REVIEWERS 	 .... 	 xv1

1.   SUMMARY AND CONCLUSIONS	1-1

     1.1.   INTRODUCTION	1-1
     1.2.   PHYSICAL AND CHEMICAL PROPERTIES, AND ANALYSIS	1-2
     1.3.   PHARMACOKINETICS	1-2
     1.4.   HEALTH EFFECTS OVERVIEW	 1-5

            1.4.1.   Toxidty	.  .  „ 1-5
            1.4.2.   Reproductive Effects . 	  .  	 1-7
            1.4.3.   Mutagenicity	1-7
            1.4.4.   Carcinogenicity. 	 1-9

2.   INTRODUCTION	     2-1

3.   BACKGROUND INFORMATION	3-1

     3.1.   INTRODUCTION	3-1
     3.2.   PHYSICAL AND CHEMICAL PROPERTIES  ..... 	 3-2
     3.3.   SAMPLING AND ANALYSIS	3-4

            3.3.1.   Chloroform in Air	3-4
            3.3.2.   Chloroform in Water  	 3-5
            3.3.3.   Chloroform in Blood  	 3-6
            3.3.4. -  Chloroform in Urine  	 3-6
            3.3.5.   Chloroform in Tissue	3-6

     3.4.   EMISSIONS FROM PRODUCTION AND USE	3-7

            3.4.1.   Emissions from Production	  .  .  . 3-7

                     3.4.1.1.   Direct Production	3-7
                     3.4.1.2.   Indirect Production .	3-12

            3.4.2.   Emissions from Use.  .	3-20

                     3.4.2.1.   Emissions from Pharmaceutical
                                Manufacturing	3-21
                     3.4.2.2.   Emissions from Fluorocarbon-22
                                Production	3-22
                     3.4.2.3.   Emissions from Hypalon®
                                Manufacture	3-22
                     3.4.2.4.   Chloroform Emissions from Grain
                                Fumigation	3-23
                                      IV

-------
                        TABLE OF CONTENTS (continued)
                     3.4.2.5.    Chloroform  Losses from  Loading and
                                Transportation	3-24
                     3.4.2.6.    Miscellaneous Use Emissions	3-25
                     3.4.2.7.    Summary  of  Chloroform Discharges
                                from  Use	3-25
            3.4.3.    Summary	3-25
     3.5.    AMBIENT AIR CONCENTRATIONS	3-26
     3.6.    ATMOSPHERIC REACTIVITY  	 .3-32
     3.7.    ECOLOGICAL EFFECTS/ENVIRONMENTAL PERSISTENCE	3-33
            3.7.1.    Ecological  Effects	3-33
                     3.7.1.1.    Terrestrial	3-33
                     3.7.1.2.    Aquatic	3-34
            3.7.2.    Environmental  Persistence	3-37
     3.8.    EXISTING CRITERIA,  STANDARDS, AND GUIDELINES	3-40
            3.8.1.    Air	3-40
            3.8.2.    Water	3-42
            3.8.3.    Food	3-43
            3.8.4.    Drugs  and  Cosmetics	3-43
     3.9.    RELATIVE SOURCE CONTRIBUTIONS	3-43
     3.10.  REFERENCES FOR  CHAPTER  3	3-44
4.   DISPOSITION AND RELEVANT PHARMACOKINETICS  	 4-1
     4.1.    INTRODUCTION	*	4-1
     4.2.    ABSORPTION	4-2
            4.2.1.    Dermal Absorption	4-2
            4.2.2.    Oral	4-3
            4.2.3.    Pulmonary  Absorption	4-6
     4.3.    TISSUE  DISTRIBUTION	4-12
     4.4.    EXCRETION	4-20
            4.4.1.    Pulmonary  Excretion	4-20
            4.4.2.    Other  Routes of  Excretion	4-30
            4.4.3.    Adipose Tissue Storage	4-31
     4.5.    BIOTRANSFORMATION OF CHLOROFORM	4-32

-------
                       TABLE OF CONTENTS  (continued)
           4.5.1.   Known Metabolites	4-32
           4.5.2.   Magnitude of Chloroform Metabolism.  .......  .4-36
           4.5.3.   Enzymatic Pathways of Biotransformation	4-39

     4.6.   COVALENT BINDING TO CELLULAR MACROMOLECULES   	4-45

           4.6.1.   Proteins and Lipids	4-45

                    4.6.1.1.   Genetic Strain Difference	4-51
                    4.6.1.2.   Sex Difference	4-53
                    4.6.1.3.   Inter-species Difference	4-53
                    4.6.1.4.   Age Difference	4-56

           4.6.2.   Nucleic Acids	4-56
           4.6.3.   Role of Phosgene	 .  .4-58
           4.6.4.   Role of Glutathione	4-59

     4.7.   SUMMARY	4-61
     4.8.   REFERENCES FOR  CHAPTER 4	4-65

5.   TOXICITY.	5-1

     5.1.   EFFECTS  OF ACUTE  EXPOSURE  TO  CHLOROFORM.	5-1

           5.1.1.   Humans	5-1

                    5.1.1.1.  Acute  Inhalation  Exposure  in  Humans.  .  .  .5-1
                    5.1.1.2.  Acute  Oral  Exposure  in Humans 	 5-5
                    5.1.1.3.  Acute  Dermal  and  Ocular  Exposure
                               in Humans	5-6

            5.1.2.   Experimental Animals	5-7

                    5.1.2.1.   Acute  Inhalation  Exposure  in  Animals  ... 5-7
                    5.1.2.2.   Acute  Oral Exposure  in Animals ...... 5-8
                    5.1.2.3.   Acute  Dermal  and  Ocular  Exposure
                               in Animals	5-11
                     5.1.2.4.   Intraperitoneal  and  Subcutaneous
                               Administration in Animals	5-12

     5.2.   EFFECTS OF CHRONIC EXPOSURE TO CHLOROFORM	5-13

            5.2.1.   Humans	5-13

                     5.2.1.1.  Chronic Inhalation Exposure in Humans  .  .5-13
                     5.2.1.2.  Chronic Oral  Exposure in Humans	5-15

            5.2.2.   Experimental Animals	5-16

                     5.2.2.1.  Chronic Inhalation Exposure in Animals  . 5-16
                     5.2.2.2.  Chronic Oral  Exposure in Animals	5-17


                                      vi

-------
                        TABLE OF CONTENTS (continued)
     5.3.    INVESTIGATION  OF  TARGET ORGAN  TOXICITY  IN  EXPERIMENTAL
            ANIMALS	5-31

            5.3.1.    Hepatotoxicity	5-31
            5.3.2.    Nephrotoxiclty	5-39

     5.4.    FACTORS MODIFYING THE TOXICITY OF  CHLOROFORM	5-49

            5.4.1.    Factors  that Increase the Toxicity	5-50
            5.4.2.    Factors  that Decrease the Toxicity	5-58

     5.5.    SUMMARY:  CORRELATION OF EXPOSURE AND  EFFECT	5-60

            5.5.1.    Effects  of  Acute  Inhalation  Exposure	5-60
            5.5.2.    Effects  of  Acute  Oral Exposure	5-61
            5.5.3.    Effects  of  Dermal  Exposure	5-62
            5.5.4.    Effects  of  Chronic Inhalation  Exposure	5-63
            5.5.5.    Effects  of  Chronic Oral Exposure	5-64
            5.5.6.    Target Organ Toxicity	5-66
            5.5.7.    Factors  that Modify the Toxicity  of Chloroform .  .  .5-76

     5.6.    REFERENCES FOR CHAPTER 5	5-77

6.   TERATOGENICITY AND REPRODUCTIVE EFFECTS   	 6-1

     6.1.    SUMMARY	6-13
     6.2.    REFERENCES FOR CHAPTER 6	6-14

7.   MUTAGENICITY 	7-1

     7.1.    INTRODUCTION	7-1
     7.2.    COVALENT  BINDING  TO  MACROMOLECULES 	 7-1
     7.3.    MUTAGENICITY STUDIES IN BACTERIAL  TEST  SYSTEMS  	 7-4
     7.4.    MUTAGENICITY STUDIES IN EUCARYOTIC TEST SYSTEMS   	 7-10
     7.5.    OTHER STUDIES INDICATIVE OF DNA DAMAGE	7-16
     7.6.    CYTOGENETIC STUDIES   	 7-21
     7.7.    SUGGESTED ADDITIONAL TESTING	7-25
     7.8.    SUMMARY AND CONCLUSIONS	7-26
     7.9.    REFERENCES FOR CHAPTER 7	7-27

8.   CARCINOGENICITY	8-1

     8.1.    ANIMAL STUDIES	8-1

            8.1.1.    Oral  Administration (Gavage):   Rat	 8-2

                     8.1.1.1. National Cancer Institute (1976)  	 8-2
                     8.1.1.2. Palmer et al.  (1979)	8-9

            8.1.2.    Oral  Administration (Gavage):   Mouse  	 8-11
                                     vn

-------
                   TABLE OF  CONTENTS  (continued)
               8.1.2.1.  National Cancer  Institute  (1976)	8-11
               8.1.2.2.  Roe et al.  (1979)	8-14
               8.1.2.3.  Eschenbrenner and Miller (1945) 	 8-18
               8.1.2.4.  Rudali (1967)	8-21

       8.1.3.   Oral Administration  (Drinking Water):  Rat
               and Mouse	8-22

               8.1.3.1.  Jorgenson  et al.  (1985)	8-22

       8.1.4.   Oral Administration  (Capsules):  Dog	8-26

               8.1.4.1.  Heywood  et al.  (1979)	8-26

       8.1.5.   Intraperitoneal Administration:  Mouse	8-29

               8.1.5.1.  Roe et al. (1968)	8-29
               8.1.5.2.  Theiss et  al.  (1977) 	8-30

       8.1.6.   Evaluation  of Chloroform  Carcinogenicity
               by Reuber  (1979)	8-31
       8.1.7.   Oral Administration  (Drinking Water):  Mouse:
               Promotion of Experimental  Tumors	8-32

               8.1.7.1.  Capel et al.  (1979)  	8-32

8.2.   CELL TRANSFORMATION  ASSAY	8-38

       8.2.1.   Styles (1979)	8-38

8.3.   EPIDEMIOLOGIC STUDIES	  .8-41

       8.3.1.   Young  et al. (1981)	8-42
       8.3.2.   Hogan  et al. (1979)	8-46
       8.3.3.    Cantor et al.  (1978)	8-47
       8.3.4.   Gottlieb et al.  (1981) 	8-52
       8.3.5.   Alavanja et al.  (1978)	8-54
       8.3.6.    Brenniman et al.  (1978)   	8-56
       8.3.7.    Struba et al.  (1979)	8-58
       8.3.8.    Discussion	8-60

8.4.   RISK ESTIMATES FROM ANIMAL DATA	8-63

       8.4.1.    Possible Mechanisms Leading to a
                Carcinogenic Response for Chloroform	8-64
       8.4.2.    Selection of Animal  Data Sets	8-67

                8.4.2.1.  NCI 1976  Bioassay (Mice):  Liver
                          Tumors	8-67
                8.4.2.2.  NCI 1976  Bioassay (Rats):  Kidney
                          Tumors	•> .8-68
                                VI 11

-------
                   TABLE OF CONTENTS (continued)
                8.4.2.3.   Roe  et  al.  1979 Bioassay  (Mice):
                          Kidney  Tumors	8-69
                8.4.2.4.   Jorgenson et al.  1985  Bioassay
                          (Rats): Kidney Tumors	8-69

       8.4.3.    Interspecies Dose Conversion	8-71

                8.4.3.1.   General Considerations . .,	8-71
                8.4.3.2.   Calculation of Human Equivalent Doses . . 8-79

       8.4.4.    Choice of Risk Model	8-87

                8.4.4.1.   General Considerations 	 8-87
                8.4.4.2.   Mathematical Description  of Low-Dose
                          Extrapolation Model 	 8-90
                8.4.4.3.   Adjustment  for Less than  Lifespan
                          Duration of Experiment	8-91
                8.4.4.4.   Additional  Low-Dose Extrapolation .... 8-92

       8.4.5.    Unit Risk Estimates	8-93

                8.4.5.1.   Definition  of Unit Risk	8-93
                8.4.5.2.   Calculation of the Slope  of the
                          Dose-Risk Relationship for  Chloroform  .  .8-93
                8.4.5.3.   Risk Associated with  1 iig/m3  of
                          Chloroform  in Air	8-96
                8.4.5.4.   Risk Associated with  1 jig/L of
                          Chloroform  in Drinking Water  	 8-96
                8.4.5.5.   Interpretation of Unit Risk Estimates  . . 8-97
                8.4.5.6.   Reconciliation of Unit Risk Estimates
                          with Epidemiological  Evidence	8-98
                8.4.5.7.   Discussion   	8-98

8.5.   RELATIVE CARCINOGENIC POTENCY  	   8-100

       8.5.1.    Derivation of  Concept 	   8-100
       8.5.2.    Potency Index  	   8-100

8.6.   SUMMARY	8-106

       8.6.1.    Qualitative 	   8-106
       8.6.2.    Quantitative	8-110

8.7.   CONCLUSIONS	8-112
8.8.   REFERENCES FOR CHAPTER  8	8-115

APPENDIX 8A   COMPARISON AMONG DIFFERENT EXTRAPOLATION  MODELS. . . 8A-1

-------
                                LIST  OF  TABLES

Table                                                                   Page
3-1       Physical Properties of Chloroform  	   3-3
3-2       Chloroform Producers, Production Sites, and Capacities ....   3-9
3-3       Estimated Chloroform Discharges from Direct Sources  	 3-13
3-4       Ethylene Dichloride Producers, Production Sites, and
          Capacities	3-15
3-5       Chloroform Discharges from Indirect Sources  	 3-21
3-6       Chlorodifluoromethane Producers and Production Sites ..... 3-23
3-7       Chloroform Discharges from Use	3-26
3-8       Relative Source Contribution for Chloroform	3-27
3-9       Ambient Levels of Chloroform	3-28
3-10      Acute and Chronic Effects of Chloroform on Aquatic
          Organisms	«  •>	3-35
3-11      Values  for kOH	3-38
3-12      Summary of EXAMS Models of the Fate of Chloroform	3-41
4-1       Physical Properties of Chloroform  and Other
          Chloromethanes	4-4
4-2       Partition Coefficients for Human Tissue at 37°C	4-4
4-3       Retention and Excretion of Chloroform  by  Man  During  and
          After  Inhalation Exposure  to  Anesthetic Concentrations ....  4-8
4-4       Chloroform Content  in United  Kingdom Foodstuffs and
          in  Human Autopsy Tissue	 4-13
4-5       Concentration of Chloroform  in Various Tissues  of Two
          Dogs After 2.5  Hours  Anesthesia	4-16
4-6       Concentrations  of  Radioactivity  (Chloroform  Plus
          Metabolites)  in Various Tissues  of the Mouse (NMRI)   	 4-17
4-7       Tissue Distribution of  14C-Chloroform  Radioactivity
           in  CF/LP  Mice After Oral  Administration  (60  mg/kg)	4-19
 4-8        Pulmonary Excretion of  13CHC13 Following  Oral Dose 	4-25

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

Table                                                                    Page


4-9       Species Difference in the Metabolism of 14C-Chloroform .  .  .  .  4-28

4-10      Kinetic Parameters for Chloroform After I.V. Administration
          to Rats ...........................  4-29

4-11      Levels of Chloroform in Breath of Fasted Normal Healthy
          Men  ............................. 4-33

4-12      Cgvalent Binding of Radioactivity From 14C-Chloroform and
          14C-Carbon Tetrachloride in Microsomal Incubation In Vitro.  .  4-49

4-13      Mouse Strain Difference in Covalent Binding of Radioactivity
          From 14C-Chloroform  .....................  4-52

4-14      In Vivo Covalent Binding of Radioactivity From 14CHC13
          in Liver and Kidney of Male and Female Mice (C57BL/6)  ....  4-54
4-15      In Vitro Covalent Binding of Radioactivity from
          to Microsomal Protein from Liver and Kidney of Male and
          Female Mice (C57BL/6)  ....................  4-54
          Covalent Binding of Radioactivity from *4C-Chloroform and
          14C-Carbon Tetrachloride in Rat Liver Nuclear and Microson
4-16
                                                            Microsomal
          Incubation In Vitro  	  4-58
4-17      Effect of Glutathione, Air, No or 0,0:0? Atmosphere
          on the In Vitro Covalent Binding of CCT4, CHC13 and CBrCl3
          to Rat Liver Microsomal Protein  	  4-60

4-18      Effects of 24-Hour Food Deprivation on Chloroform and
          Carbon Tetrachloride In Vitro Microsomal Metabolism,
          Protein, and P-450 Liver Contents of Rats	4-62

5-1       Relationship of Chloroform Concentration in Inspired
          Air and Blood to Anesthesia	5-2

5-2       Dose-Response Relationships	   5-6

5-3       Effects of Inhalation Exposure of Animals to Chloroform,
          Five Days/Week for Six Months	5-18

5-4       Effects of Subchronic or Chronic Oral Administration of
          Chloroform to Animals  	  5-20

5-5       Target Organ Toxicity of Chloroform	5-67

6-1       Summary of Results of the Schwetz et al. (1974) Study	6-3

6-2       Summary of Results of the Murray et al. (1979) Study	6-6

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

Table                                                                   Page

6-3       Summary of Effects of  the Thompson et al.  (1974)  Study. . . . 6-10

7-1       Genetic Effects of Chloroform on Strain D7 of
          S. Cerevlslae	7-12

7-2       Mitotic Index, Anaphase/Metaphase, and Presence of
          Complete C-M1tosis 1n  Grasshopper Embryos  after Exposure
          to CHCla Vapor	7-25

8-1       Effect of Chloroform on Kidney Epithelial  Tumor Incidence
          in Osborne-Mendel Rats	8-6

8-2       Effect of Chloroform on Thyroid Tumor Incidence in Female
          Osborne-Mendel Rats .......  	  8-7

8-3       Toothpaste Formulation for Chloroform Administration	8-9

8-4       Effects of Chloroform on Hepatocellular Carcinoma Incidence
          in B6C3F1 Mice   	8-13

8-5       Kidney Tumor Incidence in Male ICI Mice Treated with
          Chloroform   .	  .	8-17

8-6       Liver and Kidney Necrosis and Hepatomas in Strain A Mice
          Following Repeated Oral Administration of Chloroform
          in Olive Oil	8-20

8-7       Relative Tumor  Incidence in Male Osborne-Mendel Rats
          Treated with Chloroform in Drinking Water  	 8-24

8-8       Liver Tumor  Incidence Rates in Female B6C3F1 Mice Treated
          with Chloroform  in Drinking Water	8-25

8-9       SGPT Changes  in  Beagle Dogs Treated with Chloroform	8-28

8-10      Effect of Oral  Chloroform Ingestion on the Growth of Ehrlich
          Ascite Tumors	8-35

8-11      Effect of Oral  Chloroform Ingestion on Metastatic
          "Tumor Takes"  with 816 Melanoma	8-36

8-12      Effect of Oral  Chloroform Ingestion on the Growth and
          Spread of the  Lewis Lung Tumor	8-37

8-13      Correlation  Coefficients Between  Residual  Mortality
          Rates  in  White Males  and THM  Levels  in Drinking  Water
          by  Region and  by Percent of  the  County Population
          Served  in the United  States	8-50
                                      Xil

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

Table                                                                   Page

8-14      Correlation Coefficients  Between Bladder Cancer
          Mortality Rates by Sex and BTHM Levels in Drinking
          Water by Region of the United States	8-50

8-15      Risk of Mortality from Cancer of the Rectum Associated
          with Levels of Organics in Drinking Water  	 8-54

8-16      Cancer Risk Odds Ratios and 95% Confidence Intervals
          (Chlorinated Versus Unchlorinated)   	  8-61

8-17      Incidence of Tumors in Experimental Animal Studies	8-68

8-18      Species Difference in  the Metabolism of 14c-chloroform
          (Oral Dose of 60 mg/kg)	8-78

8-19      Pulmonary Excretion of Chloroform Following Oral Dose.  .  .  .  8-78

8-20      Continuous Human Equivalent Doses and Incidence of
          Hepatocellular Carcinomas in Male and Female B6C3F1 Mice. . . 8-84

8-21      Continuous Human Equivalent Doses and Incidence of
          Renal Tubular-Cell Adenocarcinomas in Male
          Osborne-Mendel Rats	8-84

8-22      Continuous Human Equivalent Doses and Incidence of
          Malignant Kidney Tumors in Male ICI Mice	8-85

8-23      Continuous Human Equivalent Doses and Incidence of
          Renal Tubular-Cell Adenomas and Adenocarcinomas in Male
          Osborne-Mendel Rats	8-85

8-24      Upper-Bound Estimates of Cancer Risk of 1 mg/kg/day,
          Calculated by Different Models on the Basis of Different
          Data Sets	8-95

8-25      Relative Carcinogenic Potencies Among 55 Chemicals Evaluated
          by the Carcinogen Assessment Group as Suspect Human
          Carcinogens   	8-102

8A-1      Maximum Likelihood Estimate of the Parameters for Each
          of the Four Extrapolation Models, Based on Different
          Data Sets	8A-2
                                     xiii

-------
                               LIST OF FIGURES
Figure                                                                  Page

4-1       Rate of Rise of Alveolar (Arterial)  Concentration Toward
          Inspired Concentration For Five Anesthetic Agents of
          Differing Ostwald Solubilities  	  4-9

4-2       Arteriovenous Blood Concentrations of a Patient During
          Anesthesia with Chloroform   	 4-10

4-3       Exponential Decay of Chloroform, Carbon Tetrachloride,
          Perchloroethylene and Trichloroethylene in Exhaled
          Breath of a 48 Year-old Male Accidentally Exposed to
          Vapors of These Solvents   	 4-21

4-4       Relationship Between Total 8-Hour Pulmonary Excretion of
          Chloroform Following 0.5-g Oral Dose in Man and the
          Deviation of Body Weight From Ideal	4-26

4-5       Blood and Adipose Tissue Concentrations of Chloroform During
          and After Anesthesia in a Dog    	4-32

4-6       Metabolic Pathways of Chloroform Biotransformation  	4-34

4-7       Metabolic Pathways of Carbon Tetrachloride
          Biotransformation  	 4-41

4-8       Rate of Carbon Monoxide Formation After Addition of Various
          Halomethanes to Sodium Dithionite-reduced Liver Microsomal
          Preparations From Phenobarbito!-treated Rats   	 4-46

4-9       Effect of  Increasing Dosage of i.p.-Injected 14C-Chloroform
          on  Extent of Covalent Binding of Radioactivity in vivo  to
          Liver and Kidney Proteins of Male Mice 6  Hours
          after Administration	4-50

4-10      Comparison of  Irreversible Binding of Radioactivity from
          I4C-CHC13 to Protein and  Lipid of Microsomes from
          Normal Rabbit, Rat, Mouse, and Human Liver  Incubated
          In  Vitro at  37°C in 02	4-55

5-1       Probable Pathways of Metabolism of Chloroform  in the
          Kidney.	5-44

8-1       Survival Curves  for Fisher 344  Rats  in a  Cardnogenicity
          Bioassay on  Chloroform	8-4

8-2       Negative Result  in Transformation  Assay of  Chloroform
          which was  also Negative  in the  Ames  Assay	8-40
                                      XT V

-------
                         LIST OF FIGURES (continued)
8-3       Frequency Distribution of CHC13 Levels in 76 U.S.
          Drinking Water Supplies	8-49

8-4       Effect of Increasing Dosage of i.p.-injected 14c-chloroform
          on Extent of Covalent Binding of Radioactivity jn vivo
          to Liver and Kidney Proteins of Male Mice 6 Hours After
          Administration 	8-74

8-5       Comparison of Irreversible Binding of Radioactivity
          from 14C-CHC13 to Protein and Lipid of Microsomes from
          Normal Rabbit, Rat, Mouse, and Human Liver Incubated
          In vitro at 37°C in 02	8-75

8-6       Allometric Relationship (Y=aWn) Between Species Body
          Weight (1n order:  mouse, rat, squirrel monkey, and man)
          and the Amount Metabolized of a Common Oral Dose of
          Chloroform as Calculated from the Data of Fry et al.
          (1972) and Brown et al. (1974)	8-80

8-7       The Relationship Between the Equivalent Human Exposure
          Dose and Bioassay Tumor Incidence  	  8-86

8-8       Histogram Representing the Frequency Distribution of the
          Potency Indices of 55 Suspect Carcinogens Evaluated by
          the Carcinogen Assessment Group	8-101
                                      xv

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




     The EPA Office of Health and Environmental  Assessment (OHEA)  is

responsible for the preparation of this health assessment document.  The OHEA

Environmental Criteria and Assessment Office (ECAO/RTP) Research Triangle

Park, NC  27711, had overall responsibility for coordination and direction of

the document production effort (Si Duk Lee, Ph.D., Project Manager, ECAO/RTP,

919-541-4159).
AUTHORS

Larry Anderson, Ph.D.
Carcinogen Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

David Bayliss, M.S.
Carcinogen Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Chao W. Chen, Ph.D.
Carcinogen Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Joan P. Coleman, Ph.D.
Syracuse Research  Corporation
Syracuse, NY

I.W.F. Davidson, Ph.D.
Bowman Gray School of Medicine
Winston-Sal em,  NC

D.  Anthony Gray, Ph.D.
Syracuse Research  Corporation
Syracuse, NY

Si  Duk  Lee,  Ph.D.
Environmental  Criteria  and Assessment Office
Office  of Health  and Environmental  Assessment
U.S.  Environmental Protection  Agency
Research Triangle  Park, NC
Chapter 8
Chapter 8
Chapter 8
Chapter  5



Chapter  4



Chapter  3



Chapter  2
                                      xvi

-------
Jean C. Parker, Ph.D.                                            Chapters  4,8
Carcinogen Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Sheila Rosenthal, Ph.D.                                              Chapter  7
Reproductive Effects Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Carol Sakai, Ph.D.                                                   Chapter  6
Reproductive Effects Assessment Group
Office of Health and Environmental Assessment
U.S. Environmental Protection Agency
Washington, DC

Sharon B. Wilbur, M.A.                                               Chapter  5
Syracuse Research Corporation
Syracuse, NY


U.S. Environmental Protection Agency Peer Reviewers

Karen Blanchard
Office of Air Quality Planning and Standards
Research Triangle Park, NC

Lester D. Grant, Ph.D.
Office of Health and Environmental Assessment
Environmental Criteria and Assessment Office
Research Triangle Park, NC

Joseph Padgett
Office of Air Quality Planning and Standards
Research Triangle Park, NC

Jerry F. Stara, D.V.M.
Office of Health and Environmental Assessment
Environmental Criteria and Assessment Office
Cincinnati, OH


Environmental Criteria and Assessment Office Support Staff

F. Vandiver Bradow
Allen Hoyt
                                     XVI 1

-------
     The OHEA Carcinogen Assessment Group (CAG)  was responsible for
preparation of the sections on carcinogenicity.   Participating members of the
CAG are listed below (principal authors of present carcinogenicity materials
are designated by an asterisk [*]).
   Participating Members of the Carcinogen Assessment Group
   Roy E. Albert, M.D. (Chairman)
   Elizabeth L. Anderson, Ph.D.
   Larry D. Anderson, Ph.D.*
   Steven Bayard, Ph.D.
   David L. Bayliss, M.S.*
   Robert P. Beliles, Ph.D.*
   Chao W. Chen, Ph.D.*
   Margaret M.L. Chu, Ph.D.
   James C. Cogliano, 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.
Jean C. Parker, Ph.D.*
Charles H. Ris, M.S., P.E.
Dharm V. Singh, D.V.M., Ph.D.
Todd W. Thorslund, Sc.D.
     The OHEA Reproductive Effects Assessment Group (REAG) was responsible
for preparation of the sections on mutagenicity, teratogenicity, and
reproductive effects.  Participating members of the REAG are listed below
(principal authors of present sections  are designated by an asterisk  [*]).

     Participating Members of the Reproductive Effects Assessment Group
      Eric  D.  Clegg,  Ph.D.
      John  R.  Fowle,  III,  Ph.D.
      David Jacobson-Kram,  Ph.D.
      K.S.  Lavappa,  Ph.D.
      Sheila L.  Rosenthal,  Ph.D.*
 Carol  N. Sakai,  Ph.D.*
 Lawrence R.  Valcovic, Ph.D.
 Vicki  Vaughan-Dellarco,  Ph.D.
 Peter  E. Voytek,  Ph.D.  (Director)
                                     xvi i i

-------
External Peer Reviewers
Dr. Karim Ahmed
Natural Resources Defense Fund
122 East 42nd Street
New York, NY  10168

Dr. Eula Bingham
Graduate Studies and Research
University of Cincinnati (ML-627)
Cincinnati, OH  45221

Dr. James Buss
Chemical Industry Institute of
  Toxicology
Research Triangle Park, NC  27709

Dr. I.W.F. Davidson
Wake Forest University
Bowman Gray School of Medicine
Winston-Sal em, NC

Dr. Larry Fishbein
National Center for Toxicological
  Research
Jefferson, AR  72079
(501) 542-4390

Dr. Derek Hodgson
Department of Chemistry
University of North Carolina
Chapel Hill, NC  27514

Dr. Marshall Johnson
Thomas Jefferson Medical College
Department of Anatomy
1020 Locust Street
Philadelphia, PA  19107

Dr. Trent Lewis
National Institute for Occupational
  Safety and Health
26 Columbia Parkway
Cincinnati, OH  45226
(513) 684-8394

Dr. Richard Reitz
Oow Chemical, USA
Toxicological Research Laboratory
1803 Building
Midland, MI  48640
Dr. Bernard Schwetz
National Institute of
  Environmental Health Sciences
Research Triangle Park, NC  27709

Or. James Selkirk
Oak Ridge National Laboratory
Oak Ridge, TN  37820
(615) 624-0831

Dr. Samuel Shibko
Food and Drug Administration
Division of Toxicology
200 C Street, SW
Washington, DC  20204

Dr. Robert Tardiff
1423 Trapline Court
Vienna, VA  22180
(703) 276-7700

Dr. Norman M. Trieff
University of Texas Medical Branch
Department of Pathology
Galveston, TX  77550
(409) 761-1895

Dr. Benjamin Van Duuren
Institute of Environmental Medicine
New York University Medical Center
New York, NY  10016
(212) 340-5629

Dr. James Withey
Health and Protection Branch
Department of National Health &
  Welfare
Tunney's Pasture
Ottawa, Ontario  KIA 01Z  Canada

Mr. Matthew Van Hook
Consultant
1133 North Harrison Street
Arlington, VA  22205
                                     xix

-------
                           SCIENCE ADVISORY BOARD
                       ENVIRONMENTAL HEALTH COMMITTEE
                      CHLORINATED ORGANICS SUBCOMMITTEE


     The content of this  health assessment document on  chloroform was
independently peer-reviewed in public session  by the Chlorinated  Organics
Subcommittee of the Environmental Health Committee of the Environmental
Protection Agency's Science Advisory  Board.


               ACTING CHAIRMAN. ENVIRONMENTAL HEALTH COMMITTEE

Dr. John Doull, Professor of Pharmacology and  Toxicology, University of
     Kansas Medical Center, Kansas City, Kansas  66207


                 EXECUTIVE SECRETARY. SCIENCE ADVISORY BOARD

Dr. Daniel Byrd III, Executive Secretary, Science Advisory Board, A-101 F,
     U.S. Environmental Protection Agency, Washington,  DC  20460


                                   MEMBERS

Dr. Seymour Abrahamson, Professor of Zoology and Genetics, Department of
     Zoology, University of Wisconsin, 500 Highland Avenue, Madison,
     Wisconsin  53706

Dr. Ahmed E. Ahmed, Associate  Professor of Pathology,  Pharmacology, and
     Toxicology, The University of Texas Medical Branch, Galveston, Texas
     77550.

Dr. George T. Bryan, Professor of Human Oncology,  K4/528 C.S.C Clinical
     Sciences,  University  of Wisconsin, 500 Highland Avenue, Madison,
     Wisconsin  53792.

Dr. Ronald 0. Hood, Professor  and Coordinator,  Cell  and  Developmental Biology
     Section, Department of Biology, The  University  of Alabama,  and Principal
     Associate, R.D. Hood  and  Associates,  Consulting Toxicologists, P.O.  Box
     1927, University, Alabama  35486.

Dr. K.  Roger Hornbrook,  Department of  Pharmacology,  P.O.  Box  26901,
     University of Oklahoma,  Oklahoma  City, Oklahoma  73190.

Dr. Thomas  Starr,  Chemical Industry  Institute  of Toxicology,  P.O.  Box  12137,
     Research Triangle Park,  North Carolina   27709.
                                       xx

-------
Review
Draft
(Do Not
Cite or Quote)
EPA-600/8-84-004A
March 1984
External Review Draft
             Health Assessment
        Document  for Chloroform
                    Part 1  of 2
            External  Review Draft
                          NOTICE
This document is a preliminary draft. It has not been formally released by the U.S. Environmental
Protection Agency 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 AGENCY
                Office of Research and Development
             Office of Health and Environmental Assessment
             Environmental Criteria and Assessment Office
                 Research Triangle Park, NC 27711
                       March 1984

-------
                                 PREFACE
     The Office of Health and Environmental Assessment has prepared this health
assessment to serve as a "source document" for EPA use.  This health assessment
document was developed for use by the Office of Air Quality Planning and
Standards to support decision-making regarding possible regulation of chloroform
as a hazardous air pollutant.
     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 chemical's toxicity and related characteristics
are qualitatively identified.  Observed effect levels and other measures of
dose-response relationships are discussed, where appropriate, so that the
nature of the adverse health response are placed in perspective with observed
environmental levels.
     This document will  be subjected to a thorough copy editing and proofing
following the revision based on the  EPA's Scientific Advisory Board review
comments.

-------
                               TABLE OF CONTENTS

                                                                           Page

LIST OF TABLES	     vi

LIST OF FIGURES	       *

1 .    SUMMARY AND CONCLUSIONS	     1 -1

2.    INTRODUCTION	     2-1

3.    BACKGROUND INFORMATION 	     3-1

     3.1   INTRODUCTION 	     3-1
     3.2   PHYSICAL AND CHEMICAL PROPERTIES  	     3-2
     3.3   SAMPLING AND ANALYSIS 	     3-4

           3.3.1    Chloroform in Air	     3-4
           3.3.2   Chloroform in Water	     3-5
           3.3.3   Chloroform in Blood	     3-6
           3.3.4   Chloroform in Urine	     3-6
           3.3.5   Chloroform in Tissue	     3-6

     3.4   EMISSIONS FROM PRODUCTION AND USE  	     3-6

           3.4.1    Emissions from Production	     3-6
           3.4.2   Emissions from Use	    3-20
           3.4.3   Summary of Chloroform Discharges  from  Use	    3-26

     3.5   AMBIENT AIR CONCENTRATIONS 	    3-26
     3.6   ATMOSPHERIC REACTIVITY 	    3-32
     3.7   ECOLOGICAL EFFECTS/ENVIRONMENTAL  PERSISTENCE  	    3-33

           3.7.1    Ecological Effects	    3-33
           3.7.2   Environmental Persistence	    3-36

     3.8   EXISTING CRITERIA, STANDARDS, AND  GUIDELINES  	    3-39

           3.8.1    Air	    3-39
           3.8.2   Water	    3-41
           3.8.3   Food	    3-41
           3.8 .4   Drugs and Cosmetics	    3-42

     3.9   RELATIVE SOURCE CONTRIBUTIONS 	    3-42
     3.10  REFERENCES 	    3-43

4.    DISPOSITION AND RELEVANT PHARMACOKINETICS  	     4-1

     4.1   INTRODUCTION 	     4-1
                                       111

-------
                           TABLE OF CONTENTS (cont.)

                                                                           Page

     4. 2    ABSORPTION 	    4-2

           4.2.1    Dermal Absorption	    4-2
           4.2.2    Oral Absorption	    4-3
           4.2.3    Pulmonary Absorption	    4-7

     4.3    TISSUE DISTRIBUTION 	   4-12

     4.4    EXCRETION 	   4-21

           4.4.1    Pulmonary Excretion	   4-21
           4.4.2    Other Routes of Excretion	   4-31
           4.4.3    Adipose Tissue Storage	   4-32

     4.5    BIOTRANSFORMATION OF CHLOROFORM 	   4-34

           4.5.1    Known Metabolites	   4-34
           4.5.2    Magnitude of Chloroform Metabolism	   4-37
           4.5.3    Enzymic Pathways of Biotransformation	   4-40

     4.6    COVALENT BINDING TO CELLULAR MACROMOLECULES 	   4-4 7

           4.6.1    Proteins and Lipids	   4-47
           4.6.2    Nucleic Acids	   4-58
           4.6.3    Role of Phosgene	   4-59
           4.6.4    Role of Glutathione	   4-61

     4.7    SUMMARY 	   4-63
     4.8    REFERENCES	   4-67

5.    TOXICITY 	    5-1

     5.1    EFFECTS OF ACUTE EXPOSURE TO CHLOROFORM  	    5-1

           5.1.1    Humans	    5-1
           5.1 .2    Experimental Animals	    5-6

     5.2    EFFECTS OF CHRONIC EXPOSURE TO CHLOROFORM 	   5-13

           5.2.1    Humans	   5-13
           5.2.2    Experimental Animals	   5-15

     5.3    INVESTIGATION OF TARGET ORGAN TOXICITY IN EXPERIMENTAL
           ANIMALS 	   5-28

           5.3.1    Hepatotoxicity	   5-28
           5.3.2    Nephrotoxicity	   5-34

-------
                           TABLE OF CONTENTS (cont.)
     5.4   FACTORS MODIFYING THE TOXICITY OF CHLOROFORM  	    5-38

           5.4.1   Factors that Increase Toxicity	    5-39
           5.4.2   Factors that Decrease Toxicity	    5-44

     5.5   SUMMARY; CORRELATION OF EXPOSURE AND EFFECT 	    5-46

           5.5.1   Effects of Acute Inhalation Exposure	    5-46
           5.5.2   Effects of Acute Oral Exposure	    5-4?
           5.5.3   Effects of Dermal Exposure	    5-48
           5.5.4   Effects of Chronic Inhalation Exposure	    5-48
           5.5.5   Effects of Chronic Oral Exposure	    5-50
           5.5.6   Target Organ Toxicity	    5-51
           5.5.7   Factors that Modify the Toxicity of Chloroform	    5-62

     5.6   REFERENCES	    5-63

6.   TERATOGENICITY AND REPRODUCTIVE EFFECTS	    6-1

     6.1    REFERENCES	    6-1 6

7.   MUTAGENICITY	    7-1

     7.1    INTRODUCTION	    7-1
     7.2   COVALENT BINDING TO MACROMOLECULES	    7-1
     7. 3   MUTAGENICITY STUDIES IN BACTERIAL TEST SYSTEMS	    7-3
     7.4   MUTAGENICITY STUDIES IN EUCARYOTIC TEST SYSTEMS	    7-9
     7. 5   OTHER STUDIES INDICATIVE OF DNA DAMAGE	    7-14
     7.6   CHROMOSOME STUDIES	    7-19
     7.7   SUGGESTED ADDITIONAL TESTING	    7-21
     7.8    REFERENCES	    7-23

8.   CARCINOGENICITY 	    8-1

     8.1    ANIMAL STUDIES 	    8-1

           8.1.1   Oral Administration (Gavage): Rat	    8-2
           8.1.2   Oral Administration (Gavage): Mouse	   8-11
           8.1.3   Oral Administration (Capsules):  Dog	   8-21
           8.1.4   Intraperitoneal Administration:  Mouse	   8-24
           8.1.5   Evaluation of Chloroform Carcinogenicity by
                     Reuber (1979)	   8-26
           8.1.6   Oral Administration (Drinking Water):  Mouse:
                     Promotion of Experimental Tumors	   8 -26

     8 .2    CELL TRANSFORMATION ASSAY 	   8-32

           8.2.1   Styles (1979)	   8-32

-------
                      TABLE OF CONTENTS  (cont.)

                                                                     Page

8.3   EPIDEMIOLOGIC STUDIES 	   8-34

      8.3.1    Young et al.  (1981)	   8-37
      8.3.2   Hogan et al.  (1979)	   8-40
      8.3.3   Cantor et al. (1978)	   8-42
      8.3.4   Gottlieb et al.  (1981)	   8-46
      8.3.5   Alavanja et al.  (1978)	   8-48
      8.3.6   Brenniman et al.  (1978)	   8-50
      8.3.7   Struba (1979)	   8-51
      8.3.8    Discussion	   8-53

8.4   QUANTITATIVE ESTIMATION  	   8-56

      8.4.1    Procedures for the Determination of Unit Risk	   8-59
      8.4.2   Unit Risk Estimates	   8-70
      8.4.3   Comparison of Potency with Other Compounds.	   8-80
      8.4.4   Summary of Quantitative Assessment	   8-80

8.5   SUMMARY 	   8-85

      8.5.1    Qualitative	   8-85
      8.5.2   Quantitative	   8-88

8.6   CONCLUSIONS 	   8-89
8.7   REFERENCES 	   8-91
8.8   APPENDIX A:   Comparison Among Various Extrapolation
                     Models	    A-l

-------
                                 LIST OF TABLES


Table                                                                       Page

3-1     Physical Properties of Chloroform	       3-3

3-2     Chloroform Producers, Production Sites, and Capacities	       3-8

3-3     Chloroform Discharges from Direct Sources	      3-1 3

3-4     Ethylene Dichloride Producers, Production Sites and
          Capacities	      3-1 5

3-5     Chloroform Discharges from Indirect Sources	      3-21

3-6     Chlorodifluoromethane Producers and Production Sites	      3-23

3-7     Chloroform Discharges from Use	      3-27

3-8     Relative Source Contribution for Chloroform	      3-28

3-9     Ambient Levels of Chloroform	      3-29

3-1 0    Acute and Chronic Effects of Chloroform on Aquatic
          Organisms	      3-34

3-11     Values for knu	        3-38
                    Un

3-12    Summary of EXAMS Models of the Fate of Chloroform	      3-40

4-1     Physical Properties of Chloroform and Other
          Chloromethanes	       4-4

4-2     Partition Coefficients for Human Tissue at 37°C	       4-5

4-3     Retention and Excretion of Chloroform in Man During and
          After Inhalation Exposure to Anesthetic Concentations	       4-9

4-4     Chloroform Content in United Kingdom Foodstuffs and
          in Human Autopsy Tissue	      4-1 4

4-5     Concentration of Chloroform in Various Tissues of Two
          Dogs After 2. 5 Hours Anesthesia	      4-16

4-6     Concentration of Radioactivity (Chloroform Plus
          Metabolites) in Various Tissues of the Mouse (N MRI)	      4-18

4-7     Tissue Distribution of   C-Chloroform Radioactivity
          in CF/LP Mice  After Oral Administration (60 mg/kg)	      4-20

-------
                            LIST OF TABLES  (cont.)

Table                                                                      Page

4-8     Pulmonary Excretion of   CHCL~ Following Oral
           Dose:   Percent of Dose	   4-26

4-9     Species Difference in the Metabolism of   C-Chloroform	     4-29

4-4b    Kinetic Parameters for Chloroform After I.V. Administration
          to Rats	   4-30

4-1 0    Levels of Chloroform in Breath of Fasted Normal
          Healthy Men	   4-35

4-11      Cojplent Binding of Radioactivity From   C-Chloroform and
            C-Carbon Tetrachloride in Microsomal Incubation
          In Vitro	   4-49

4-12    Mouse Strain Difference in Covalent Binding of Radioactivity
          From   C-Chloroform	     4-53

4-13    In Vivo Covalent Binding of Radioactivity From   CHC1o
          in Liver and Kidney of Male and Female Mice (C57BL/6)	   4-55

4-1 4    In Vitro Covalent Binding of Radioactivity from   CHC1
          to Microsomal Protein from Liver and Kidney of Male and
          Female Mice (C57BL/6)	   4-56

                                               1 4
4-15    Covalent Binding of Radioactivity from   C-Chloroform and
            C-Carbon Tetrachloride in Rat Liver Nuclear and
          Microsomal Incubation In Vitro	   4-60

4-16    Effect of Glutathione, Air, N? or CO:  0? Atmosphere
          on the In Vitro Covalent Binding of C Cl^ and C Br Cl
          to Rat Liver Microsomal Protein	   4-62

4-1 7    Effects of 24-Hour Food Deprivation on Chloroform and
          Carbon Tetrachloride In Vitro Microsomal Metabolism,
          Protein, and P-450 Liver Contents of Rats	   4-64

5-1     Relationship of Chloroform Concentration in Inspired
          Air and Blood to Anesthesia	     5-2

5-2     Dose-Response Relationships	     5-7

5-3     Effects of Inhalation Exposure of Animals to Chloroform,
          Five Days/Week for Six Months	   5-17

5-4     Effects of Subchronic or Chronic  Oral Administration of
          Chloroform to Animals	   5-19
                                      vm

-------
                             LIST  OF  TABLES  (cont.)

Table

5-5     Target Organ Toxicity of Chloroform	    5-52

7-1     Genetic Effects of Chloroform on Strain D7 of
          S. Cerevisiae	    7-10

8-1     Effect of Chloroform on Kidney Epithelial Tumor Incidence
          in Osborne-Mendel Rats	     8-5

8-2     Effect of Chloroform on Thyroid Tumor Incidence in Female
          Osborne-Mendel Rats	     8-7

        Toothpaste Formulation for Chloroform Administration	     8-9

        Effects of Chloroform on Hepatocellular Carcinoma Incidence
          in B6C3F1  Mice	    8-13

8-5     Kidney Tumor Incidence in Male ICI Mice Treated with
          Chloroform	    8-17

8-6     Liver and Kidney Necrosis and Hepatomas in Strain A Mice
          Following Repeated Oral Administration of Chloroform
          in Olive Oil	    8-19

8-7     SGPT Changes in Beagle Dogs Treated with Chloroform	    8-23

8-8     Effect of Oral Chloroform Ingestion on the Growth of Ehrlich
          Ascites Tumors	    8-29

8-9     Effect of Oral Chloroform Ingestion on Metastatic Tumor Takes
          with B1 6 Melanoma	    8 -30

8-10    The Effect of Oral Chloroform Ingestion on the Growth and
          Spread of the Lewis Lung Tumor	    8 -31

8-11    Correlation Coefficients Between Residual Mortality
          Rates in White Males and THM Levels in Drinking Water
          by Region and by Percent of the County Population
          Served in the United States	    8-44

8-12    Correlation Coefficients Between Bladder Cancer
          Mortality Rates by Sex and  BTHM Levels in Drinking
          Water by Region of the United States	    8-45

8-13    Risk of Cancer of the Rectum  Mortality Associated
          with Level of Organics in Drinking Water	    8 -48
                                      IX

-------
                            LIST OF TABLES  (cont.)

Table                                                                      Page

8-1*4    Cancer Risk Odds Ratios and 95% Confidence Intervals
          (Chlorinated Versus Unchlorinated)	    8-54

8-15    Incidence of Hepatocellular Carcinomas in Female and Male
          B6C3F1  Mice	    8-71

8-16    Incidence of Tubular-Cell Adenocarcinomas in Male
          Osborne-Mendel Rats	    8-72

8-17    Incidence of Malignant Kidney Tumors in Male ICI Mice	    8-72

8-18    Upper-Bound Estimates of Cancer Risk of 1 mg/kg/day,
          Calculated by Different Models on the Basis of Different
          Data Sets	    8-75

8-19    Relative Carcinogenic Potencies Among 53 Chemicals Evaluated
          by the Carcinogen Assessment Group as Suspect Human
          Carcinogens	    8-82

-------
                                LIST OF FIGURES

 Figure                                                                      Page

 4-1      Rate  of Rise of Alveolar  (Arterial)  Concentration Toward
           Inspired Concentration  For Five Anesthetic Agents  of
           Differing Ostwald Solubilities	    4-10

 4-2      Arteriovenous Blood Concentrations of a Patient During
           Anesthesia with Chloroform	    4-11

 4-3      Exponential Decay of Chloroform, Carbon Tetrachloride,
           Perchloroethylene and Trichloroethylene in Exhaled
           Breath of 48 Year-old Male Accidentally Exposed to
           Vapors of These Solvents	    4-22

 4-4      Relationship Between Total 8-Hour Pulmonary Excretion of
           Chloroform Following 0.5-g Oral Dose in Man and the
           Deviation of Body Weight From Ideal	    4-27

 4-5      Blood and Adipose Tissue  Concentrations of Chloroform During
           and After Anesthesia in a Dog	    4-33

 4-6      Metabolic Pathways of Chloroform Biotransformation.
           (Identified CH Cl_ metabolites are underlines)	      4-36

 4-7      Metabolic Pathways of Carbon Tetrachloride
           Biotransformation	    4-42

 4-8      Rate  of Carbon Monoxide Formation After Addition of Various
           Halomethanes to Sodium Dithionite-reduced Liver Microsomal
           Preparations From Phenobarbitol-treated Rats	    4-46

 4-9      Effect of Increasing Dosage of i.p.-Injected
            C-Chloroform on Extent of Covalent Binding of
           Radioactivity In Vivo to Liver and Kidney Proteins of Male
           Mice 6 Hours After Administration	    4-51

 4-1 0     Comparison of irreversible binding of radioactivity from
            C-CHC1_ to protein and lipid of microsomes from
           normal rabbit, rat,  mouse, and human liver incubated
           in vitro at 37°C in 02	      4-57

8-1      Survival curves for Fisher 344 Rats in a Carcinogenicity
           Bioassay on Chloroform	    8-5

8-2      Negative Result in Transformation Assay of Chloroform
          which was also Negative in the Ames Assay	   8-35

8-3      Frequency distribution of CHC13 levels in 68  U.S.
          drinking water supplies.  The abscissa is linear in the
          logarithm of the level	   8-43

-------
                            LIST OF FIGURES (cont.)
8-4     Point and Upper-Bound Estimates of Four Dose-Response Models
          Over Low-Dose Region on the Basis of Liver Tumor Data
          for Female Mice	     8-76

8-5     Point and Upper-Bound Estimates of Four Dose-Response Models
          Over Low-Dose Region on the Basis of Liver Tumor Data
          for Male Mice	     8-77

8-6     Histogram Representing the Frequency Distribution of the
          Potency Indices of 53 Suspect Carcinogens Evaluated by
          the Carcinogen Assessment Group	     8-81
                                     XII

-------
     The Office of Health and Environmental Assessment  (OHEA), U.S. EPA,
is responsible for the preparation  of  this  health assessment document.  The
Environmental  Criteria and Assessment  Office  (ECAO/RTP), OHEA, had the
overall  responsibility for coordination  and the  document production effort.

                             Project Manager

                            Si Duk  Lee,  Ph.D.
               Environmental  Criteria  and Assessment Office
              U.S. EPA, Research Triangle Park,  N.C.  27711
                              (919)  541-4159


                          Authors and  Reviewers

               The principal  authors of  this  document are:

                           Larry Anderson Ph.D.
                       Carcinogen Assessment  Group
                        U.S.  EPA, Washington,  D.C.

                            David Baylis, M.S.
                       Carcinogen Assessment  Group
                       U.S.,  EPA, Washington,  D.C.

                           Chao W.'Chen, Ph.D.
                       Carcinogen Assessment  Group
                       U.S.,  EPA, Washington,  D.C.

                            Carol Sakai, Ph.D.
                  Reproductive Effects Assessment Group
                       U.S.,  EPA, Washington,  D.C.

                         Sheila Rosenthal,  Ph.D.
                  Reproductive Effects Assessment Group
                       U.S.,  EPA, Washington,  D.C.

                          I.W.F.  Davidson,  Ph.D.
                      Bowman  Gray School of Medicine
                           Winston  Salem, N.C.

                          D.  Anthony Gray,  Ph.D.
                         Syracuse Research  Corp.
                              Syracuse,  N.Y.

                          Sharon  B.  Wilbur, M.A.
                         Syracuse Research  Corp.
                              Syracuse,  N.Y.

                          Joan P. Coleman,  Ph.D.
                         Syracuse Research  Corp.
                              Syracuse,  N.Y.

-------
     The following individuals provided peer-review of  this  draft  or  earlier
drafts of this document:

U.S. Environmental Protection Agency

Joseph Padgett
Office of Air Quality Planning and Standards
U.S. EPA

Karen Blanchard
Office of Air Quality Planning and Standards
U.S. EPA

Jerry F. Stara, D.V.M.
Office of Health and Environmental  Assessment
Environmental Criteria and Assessment Office
U.S. EPA

Lester D. Grant, Ph.D.
Office of Health and Environmental  Assessment
Environmental Criteria and Asssessment Office
U.S. EPA

Participating Members of the Carcinogen Assessment Group

Roy E. Albert, M.D. (Chairman)
Elizabeth L. Anderson, Ph.D.
Larry D. Anderson, Ph.D.
Steven Bayard, Ph.D.
David L. Bayliss, M.S.
Chao W. Chen, Ph.D.
Margaret M. L. Chu, Ph.D.
Bernard H. Haberman, D.V.M., M.S.
Charalingayya B. Hiremath, Ph.D.
Robert E. McGaughy, Ph.D.
Dharm W. Singh, D.V.M., Ph.D.
Todd W. Thorslund, Sc.D.

Participating Members of the Reproductive Effects Assessment Group

Peter E. Voytek, Ph.D. (Director)
John R. Fowle, III, Ph.D.
Carol Sakai, Ph.D.
Ernest Jackson, M.D.
K.S. Lavappa, Ph.D.
Sheila Rosenthal, Ph.D.
Vicki Vaughn-Dellarco, Ph.D.
                                           XIV

-------
 External Peer Reviewers
Dr. Karim Ahmed
Natural Resources Defense Fund
122 E. 42nd Street
New York, N.Y.  10168

Dr. Eula Bingham
Graduate Studies and Research
University of Cincinnati (ML-627)
Cincinnati, Ohio  45221
(513) 475-4532

Dr. James Buss
Chemical Institute of Industrial
  Toxicology
Research Triangle Park, N.C.  27709

Dr. I.W.F. Davidson
Wake Forest University
Bowman Gray Medical School
Winston Salem, N.C.

Dr. Larry Fishbein
National Center for Toxicological
  Research
Jefferson, Arkansas  72079
(501) 542-4390

Dr. Derek Hodgson
Department of Chemistry
University of North Carolina
Chapel Hill, N.C.  27514

Dr. Marshall Johnson
Thomas Jefferson Medical College
Department of Anatomy
1020 Locust Street
Philadelphia, Pennsylvania  19107

Dr. Trent Lewis
National Institute of Occupational
  Safety and Health
26 Columbia Parkway
Cincinnati, Ohio  45226
(513)  684-8394

Dr. Richard Reitz
Dow Chemical, USA
Toxicology Research Laboratory
1803 Building
Midland, Michigan  48640
Dr. Marvin A. Schneiderman
Clement Associates, Incorporated
Arlington, Virginia  22209
(703) 276-7700

Dr. Bernard Schwetz
National Institute of Environmental
  Health
Research Triangle Park, NC  27709
(919) 541-7992

Dr. James Selkirk
Oak Ridge National Laboratory
Oakridge, Tennessee  37820
(615) 624-0831

Dr. Samuel Shibko
Food and Drug Administration
Division of Toxicology
200 C Street, S.W.
Washington, D.C.  20204
Telephone:

Dr. Robert Tardiff
1423 Trapline Court
Vienna, Virginia  22180
(703) 276-7700

Dr. Norman M. Trieff
University of Texas Medical Branch
Department of Pathology, UTMB
Galveston, Texas  77550
(409) 761-1895

Dr. Benjamin Van Duuren
Institute of Environmental Medicine
New York University Medical Center
New York, New York  10016
(212) 340-5629

Dr. James Withey
Health and Protection Branch
Department of National  Health &
  Welfare
Tunney's Pasture
Ottawa, Ontario
CANADA, KIA 01Z
                                         xv

-------

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








      Chloroform  is a dense,  colorless,  volatile liquid used  primarily in the




 production of chlorodifluoromethane  (90%) and for export (5/£).  Non-consumptive




 uses  (5%)  include  use  as  a  solvent,  as a  cleaning agent, and  as a  fumigant




 ingredient.   Direct  United States  production  of chloroform  in 1981  was 184




 million  kg,  while  indirect  production  is  estimated at  13-2 million kg  (=193




 million kg overall).  The amount of chloroform emitted to air is estimated  to be




 7.2 million kg,  emissions to water are 2.6 million kg,  and  emissions to  land are




 0.6 million kg.  Total United States emissions are estimated to be  10.4 million




 kg.




      Chloroform  is ubiquitous in the environment, having been found  in urban and




 non-urban locations.  The northern hemisphere background average  has been deter-



                      —1 2
 mined to be 14 ppt (10~    v/v),  while the southern hemisphere has  been determined




 to be <3 ppt.  The global average is 8 ppt.  For  the most part,  urban ambient air




 concentrations remain <1000 ppt,  and rural  or  remote locations can  be  <10 ppt.




 There are some notable exceptions, however,  the  reasons  for this are not readily




 apparent.    The  highest  values  reported   were  in Rutherford,   New  Jersey,




 (31,000 ppt) and Niagara Falls, New  York (21,611 ppt).




     Hydroxyl radical oxidation is the  primary  atmospheric reaction of chloro-




 form.  Based on  the rate constant  for reaction  with chloroform,  a  half-life of



 21.5 weeks is expected.   The principal products  from this  reaction  are  HC1 and




 CO,,.  It has been estimated that roughly 1/& of the tropospheric chloroform will




diffuse into  the stratosphere,  based on a  lifetime of 0.2  to 0.3  years  and a




troposphere-to-stratosphere turnover time of 30 years.  An EXAMS model of chloro-




form in water confirms other data that the major removal process for chloroform



in water is evaporation.
                                      1-1

-------
     The best analytical method  for detection of chloroform  appears  to be gas




chromatography with  electron  capture  or  electrolytic conductivity  detector.




This gives a detection limit of <5 ppt.




     The pharmacokinetics and metabolism  of chloroform  have been studied in both




humans and experimental animals.   Chloroform is rapidly and extensively absorbed




through the  respiratory and  gastrointestinal  tracts.   Absorption  through the




skin would make a significant  contribution  to  body  burden  only in instances of




contact of the skin with liquid chloroform.




     The available data suggest that,  for resting human at least 2 hours of




exposure are required to reach an apparent  body equilibrium with the inhaled




chloroform concentration.  The percentage of the inhaled chloroform concentration



retained in the body at "equilibrium"  would be -65%, and  is  independent of the




inhaled concentration.  The magnitude of chloroform uptake into the body  (dose or




body burden) is directly proportional to the concentration of chloroform in the




inspired air, the duration  of exposure,  and the  respiratory  minute volume, and




can be estimated by multiplying the  percent  retention by the total volume of air




breathed during exposure and by the exposure concentration.




     The absorption of chloroform from the  gastrointestinal tract appears to be




virtually complete,  judging  from  recovery  of  unchanged chloroform and metabol-




ites  in  the  exhaled air of  humans  and  in  the exhaled air,  urine,  feces, and




carcass of  experimental animals.   Peak  blood  levels  occurred at =1 hour after




oral administration of chloroform in olive  oil to humans or animals.




     After  inhalation or  ingestion,  highest  concentrations  of chloroform are




found in tissues with higher lipid contents. Results from the administration of




  C-labeled chloroform  to animals indicate  that the distribution of radioactiv-




ity (reflecting both chloroform and  its metabolites) may be  affected  by the route




of  exposure.   Oral administration  appeared to result  in the accumulation of a
                                       1-2

-------
greater proportion of radioactivity  in  the  liver than did inhalation exposure,




but differences  in  experimental protocols  make  this interpretation tentative.




Sex differences in the distribution of chloroform and  its metabolites were found




only  in  mice and not in  rats or  squirrel  monkeys.   The  kidneys  of male mice




accumulated strikingly more radioactivity than did those of female mice.  Other




than  the renal accumulation of radioactivity in male mice, the tissue distribu-




tion  of radioactivity after oral administration  was  similar in mice, rats, and




squirrel monkeys.




      Chloroform has been detected  in  fetal  liver.  Chloroform would be expected




to appear  in  human milk,  because  it has been found  in  cow's  milk, cheese, and




butter.




      Chloroform  is metabolized via microsomal cytochrome  P-450  oxidation to  a




reactive intermediate.  The primary end  product of chloroform metabolism is CC^?




but small  amounts of the reactive  intermediate bind covalently to  tissue macro-




molecules  or  conjugate with cysteine  and  glutathione.  Covalent binding of the




reactive intermediate to macromolecules is  considered to be responsible for the




hepato- and nephrotoxicity of chloroform.




      The   initial  product  of  cytochrome  P-450  oxidation of  cnloroform  is




trichloromethanol, which spontaneously  dehydrochlorinates to produce phosgene.




Phosgene is  thought  to  be the toxic  reactive  intermediate produced during the




metabolism of chloroform.   Phosgene reacts with water to yield C0_, with protein




to form a covalently  bound  product, and  with cysteine and glutathione.  While the




liver is the  primary site for chloroform metabolism, other  tissues,  including the




kidney, can also metabolize chloroform  to C0p.




      Interspecies comparisons  of   the magnitude  of  chloroform  metabolism have




been made only for the oral route.   Mice,  rats, and squirrel monkeys metabolized




85, 66, and "\Q%, respectively, of  a 60 mg/kg body weight dose of   C-chloroform
                                      1-3

-------
   1 4
to   C0_(measured  in  the  expired air).  Most  of the remainder  was  exhaled as



unchanged chloroform;  small amounts of radioactivity (2 to 8% of the dose) were



excreted in the urine  and  feces.  Human subjects  (1 or 2/dose) who ingested 0.1,



0.5, and 1.0 g of  3C-chloroform (=1.4,  7,  and  14  mg/kg body weight) metabolized



all of the  low dose, 50% of the intermediate dose, and only -35% of the high dose,



judging from the excretion of unchanged chloroform and  ^CO  through the lungs.


This difference indi^j '.!;._:;;  ./'.at the fraction of the dose metabolized is dose-


dependent.



     Regardless of  the route of entry into  the body,  chloroform  is  excreted



unchanged  through  the lungs and eliminated via  metabolism,  with  the primary



stable metabolite,  CO   also being excreted through the lungs.  High concentra-



tions of unchanged chloroform have been  found  in the bile  of squirrel monkeys



after oral  administration,  but not in the urine or feces.   The inorganic chloride



generated from chloroform metabolism is excreted via the urine.



     Decay  curves for the pulmonary excretion of unchanged chloroform in humans



appear  to  consist  of  three  exponential components.   The  terminal  component,



thought to  correspond to elimination  from adipose tissue, had  a  half-time of



36 hours.   This  long half-time,  together  with  the  relatively  high  levels of


chloroform  found in adipose tissues of humans known to have been  exposed only to



ambient levels of the  chemical and of animals exposed experimentally, indicates a



potential for bioaccumulation.



     The adverse health effects of exposure to chloroform include neurological,



hepatic, renal, and  cardiac effects.   These  effects  have been documented in



humans  as  well  as  in  experimental animals.   In  addition, studies with animals



indicate that chloroform is carcinogenic and may be teratogenic.



     Evidence  of  chloroform's  effects  on humans has been  obtained  primarily



during  the  use  of  this chemical as an  inhalation anesthetic.   In  addition to



depression  of  the  central  nervous  system,  chloroform anesthesia was associated
                                     1-4

-------
with cardiac arhythmias (and some cases  of cardiac arrest), hepatic necrosis and




fatty  degeneration,  polyuria, albuminuria,  and  in cases  of severe poisoning,




renal  tubular  necrosis.   When used for  obstetrical  anesthesia,  chloroform was




likely to produce respiratory depression in the infant.  Experimental exposures




of  humans  to  chloroform  have focused  only on  subjective  responses.   Humans




exposed experimentally to  chloroform  for 20 to 30 minutes have reported dizzi-




ness, headache, giddiness, and tiredness at concentrations >1000 ppm, and light




intoxication at concentrations above 4000 ppm.




     Similar  symptoms occurred   in  workers  employed  in  the manufacture  of




lozenges  containing  chloroform;  exposure  concentrations   ranged  from 20  to




237 ppm, with  occasional  brief  exposure  to =1000 ppm.   Additional complaints




were of gastrointestinal distress,  and  frequent  and  scalding micturition.   The




only  other  report of  adverse effects  stemming  from  occupational  exposure  to




chloroform was of enlargement  of the  liver; this report was compromised by the




apparent lack of suitable controls.




     Acute inhalation experiments with animals revealed that  single  exposures to




100 ppm were sufficient to produce mild  hepatic  effects in mice.   The exposure




level that would produce mild renal  effects  is not known, but frank toxic effects




occurred in the kidneys of male mice exposed  to 5  rag/I  (1025  ppm).  In subchronic




inhalation experiments, histological evidence of  mild  hepato- and nephrotoxicity




occurred in rats with exposures  to  as  low as 25  ppm,  7 hours/day for 6 months.




The effects were reversible  if exposure was terminated, and did not occur when



exposure was limited to 4 hours/day.




     Information on the effects of acute and long-term oral  exposure to chloro-




form is available  primarily from  experiments with animals.   Human  data  are




mainly in the form of case  reports and involve the abuse of medications contain-




ing not  only  chloroform,  but  other  potentially  toxic  ingredients as  well;
                                      1-5

-------
however, a fatal dose of as little as 1/3  ounce was reported.  As with inhalation
exposure, the  primary  effects  of oral exposure were  hepatic  and renal damage.
Narcosis also occurred with high doses, but this effect was not usually a focus
of concern  in  these  experiments.  Subchronic and chronic  toxicity experiments
with rats, mice, and dogs did not clearly  establish a no-effect level of exposure
for  systemic  toxicity.   Although a dose  level  of  1  7 mg/kg/day  of chloroform
produced  no adverse effect  in  four strains of mice,  the  lowest  dosage tested,
15 mg/kg/day, elevated some clinical chemistry indices of hepatic damage in dogs
and  appeared to affect a component of the  reticuloendothelial  system (histio-
cytes)  in their livers.
     No  controlled  studies have been performed to  define  dose-response thres-
holds for neurological or cardiac effects of ingested  or inhaled chloroform.  It
is not known whether subtle impairment of neurological or cardiac function might
occur at  levels as low as or lower than  those which affect the liver.
     Several substances,  which  are  of  interest  because  of  accidental  or
intentional human  exposure,  have  been  shown  to modify  the  systemic  toxicity
of chloroform,  usually  by  modifying  the metabolism of  chloroform to  the
reactive  intermediate.   Examples  of  substances  that potentiate  chloroform-
induced toxicity  are  ethanol,  PBBs,   ketones  and   steroids.   Factors  that
appear  to protect against  toxicity include disulfiram  and  high  carbohydrate
diets.
     Chloroform appears  to  have  teratogenic  potential  in  laboratory animals
when inhaled.   Chloroform was  selectively  more toxic  to the fetuses than to
 the  dams when  pregnant rats and mice were exposed to the vapor.   Delayed  fetal
 development occurred  at  an  exposure level  (30 ppm) that produced minimal mater-
 nal  effects.   Embryotoxic effects and low  but  statistically  significant  inci-
 dences  of teratogenic  effects occurred  at  an exposure level  that  produced mild
 to moderate maternal  effects (i.e.,  100  ppm).  When chloroform was  administered

                                       1-6

-------
 orally  (via  gavage)  to  pregnant  rats  and  rabbits,  however,  toxic  effects  on




 fetal development  occurred  only  at  dosage levels that produced  severe  toxic




 effects  in the dams; no teratogenic  effects were observed at any  dosage  level




 in  these  animals.



      It has  been demonstrated that  chloroform can  be metabolized  in vivo  and  in




 vitro to  a  substance(s) (presumably  phosgene)  that  interacts with protein and




 lipid.    However,  the  sole  experiment measuring  interaction of metabolically




 activated chloroform with DNA yielded a negative result.   This result was  judged




 as  inconclusive, because the specific  activity  of  the   CHC1  may have been too



 low.




     The  majority  of the  assays for  mutagenicity and genotoxicity  have also




 yielded  negative  results;   however,  many  of  these results  are inconclusive




 because of various inadequacies  in  the experimental  protocols used.   The major




 problem  is  with those  bacterial,  sister chromatid  exchange,   and  chromosome




 aberration studies that used reconstituted  exogenous activation  systems  (i.e.,




 S-9 mix).  In none  of these studies was it  shown  that  chloroform  was activated or




 metabolized by the activation system  used.   Metabolism of 2-aminoanthracene or



 vinyl compounds  (used as positive controls)  is  probably an  inadequate  indication




 that  the  activation  system can metabolize chloroform  because these  substances




 are not  halogenated  alkanes and  are  therefore  not  metabolized like  them.   A




 better indication  that  an  activation system is  sufficient  for  metabolism  of




 chloroform may be to  show that it metabolizes    CHC1_ to intermediates that bind




 to macromolecules.   A second problem with experimental protocols utilizing exo-




genous activation systems relates to the possiblity that any reactive metabolic




 intermediates formed  may  react  with  microsoraal or  membrane  lipid or protein




before reaching the DNA of the test organism.   A third potential problem occurs




in those  in vitro protocols  in which precautions were  not taken to  prevent escape



of volatilized CHC1-.






                                      1-7

-------
     Studies in which endogenous activation systems were  used  include those in




yeast,  in Drosophila  (sex-linked recessive lethal),  and  in mice  (bone-rnarrow




micronucleus,  sperm head  abnormalities and host-mediated assay).   The results




from several of these studies suggest that chloroform may be a weak mutagen.




     In summary,  with the  present data,  no  definitive conclusions can be reached




concerning the mutagenicity  of  chloroform.  However,  there  is some indication




(from the binding  studies and from the mutagenicity  tests  that utilized endo-




genous or in vivo metabolism) that chloroform may have  the  potential  to  be a weak




mutagen.   In  order  to  substantiate this,  only certain  well-designed in vivo




mutagenicity studies or studies with organisms possessing endogenous activation




systems are recommended.




     Chloroform  in  corn  oil  administered  at  maximally and  one-half  maximally




tolerated doses  by gavage  for  78  weeks  produced  a  statistically  significant




increase  in the  incidence  of  hepatocellular  carcinomas in  male  and  female




B6C3F1  mice  and  renal  epithelial  tumors  (malignant  and  benign)  in  male




Osborne-Mendel rats;  a  carcinogenic response of female  Osborne-Mendel rats to




chloroform was not apparent in this study.




     A  statistically significant  increase in  the  incidence  of  renal  tumors




(benign and malignant) was found in another study in male ICI mice  treated with




chloroform in either toothpaste or arachis oil by gavage for 80 weeks; however,




treatment with a gavage dose of chloroform in  toothpaste for  80 weeks did not




produce  a  carcinogenic  response in  female ICI  mice and male  mice  of the CBA,




C57BL, and CF/1 strains.  A  carcinogenic  response  was not observed in male and




female  Sprague-Dawley  rats  given  chloroform   in   toothpaste  by   gavage  for




80 weeks, but early mortality was high in control and  treatment groups.  Gavage




doses of chloroform in toothpaste did not show a carcinogenic effect in male and




female beagle dogs treated for 7 years; however, the treatment period  was short
                                      1-8

-------
 in relation to  the lifespan  of the beagle  dog.   Results  of preliminary




 toxicity tests  and the  carcinogenicity  studies  indicate  that  doses  of




 chloroform in  toothpaste given to mice,  rats,  and dogs in the carcinogen-




 icity studies  approached  maximally  tolerated doses.   However,   doses  of




 chloroform in  toothpaste  given to  mice and  rats  were lower than those




 given in corn  oil.




      Hepatomas were found in  NLC  mice  given  chloroform twice weekly  for




 an unspecified period of time  and  in female strain A mice given chloroform




 once  every 4 days for a total  of  30 doses  at a level which produced liver




 necrosis; however,  small numbers of animals were examined in pathology, the




 duration of studies was  below the lifetime of  the animals,  and no control




 group of NLC mice was apparent.  Although a carcinogenic effect of chloro-




 form  was not  evident in newborn  (C57  x DBA2 -  Fl) mice given  single or




 multiple subcutaneous doses during the  initial  8  days  of life and observed




 for their lifetimes,  the  dose levels used appeared well below  a maximum




 tolerated dose  and the  period of treatment  after  birth  was quite  short




 compared to lifetime treatment.   Chloroform  was  ineffective  at  maximally




 tolerated and  lower doses  in  a pulmonary adenoma bioassay  in   Strain A




 mice.  Although  an ability of chloroform to  promote growth and  spread of




 Tjewis  lung  carcinoma,  Erlich  ascites,  and  B16 melanoma cells  in mice has




been  shown,  the  mechanism by  which  chloroform produced  this  effect is




 uncertain, and  the  relevance  of  this  study  to  the  evaluation of  the




 carcinogenic potential of chloroform is presently not clear.




     There are no epidemic logic studies  of  cancer and  chloroform per  se.




There appears  to be an  increased  risk of  cancer of the bladder,  rectum,




and large  intestine from  chlorinated  drinking water  and,  by  inference,




possibly  from chloroform, the predominate contaminant.
                                      1-9

-------
     In conclusion, evidence that chloroform has carcinoqenic activity is

on increased incidences of hepatocellular carcinomas in mal e and female B6C3F1

mice, renal epithelial tumors in male Osborne-Mendel rats, kidney tumors in

male TCT mice, arid hepatomas in NLC and female strain A mice.  Applying the

International Agency for Research on Cancer  (IARC) criteria for animal studies,

this level of evidence would be sufficient for concluding that chloroform is

carcinogenic in experimental animals.

     Although there is limited evidence in humans for the carcinogenicity of

chlorinated drinking water, and by inference, possibly for chloroform, the human

evidence for chloroform itself is insufficient.  Considering both human and

animal evidence, the overall IARC classification would be Group 2B, meaning

that chloroform is probably carcinogenic in humans.

     Four data sets that contain sufficient  information are used to estimate

the carcinogenic potency of chloroform.  They are: liver tumors in female mice,

Liver tumors in male mice, kidney tumors in male rats, and kidney tumors in

male mice.  The carcinogenic potencies, calculated by the  linearized multistage

model on the basis of these four data sets,  are comparable within an order of

magnitude.  The two data sets for liver tumors in female and male mice give

slightly higher potency values than the two  data sets for kidney tumors in male

rats and male mice.  The geometric mean, g*  = 7 x 10-2/(mg/kg/day), of the
                                          1
potencies calculated from  liver tumors in male and female mice  (the most sensi-

i;ve species), is  taken to represent the carcinogenic potency of chloroform.

Rised on  this potency, the upper-bound estimate of the cancer risk due to  1

^cj/rn3 of  chloroform in air is P = 1 x 10-5.  The upper-bound estimate of

 the  cancer  risk due to 1 jug/liter in water if P = 2  x 10-6.  These estimates

appear  consistent  with the limited epidemiclogic data available for humans.

The  estimated potency of chloroform is in the  fourth quartile of 53 suspect

carcinogens that have been evaluated by the  Carcinogen Assessment Group.
                                     1-10

-------
                                2.   INTRODUCTION

     The U.S.  Environmental Protection Agency is responsible under the
authority of various laws for the identification, comprehensive assessment,
and as appropriate regulation, of environmental  substances which may be of
concern to the public health.   For example, under section 112 of the Clean
Air Act the EPA Administrator is directed to establish standards for any
air pollutant (other than those for which national ambient air quality
standards are applicable) which, in his judgment, "causes, or contributes
to, air pollution which may reasonably be anticipated to result in an
increase in mortality or an increase in serious  irreversible, or incapacitating
reversible, illness."
     Within EPA, the Office of Health and Environmental Assessment is
responsible for providing scientific assessments of health effects for
potentially hazardous air pollutants such as chloroform.  These health
assessment documents form the scientific basis for subsequent agency actions,
including the Administrator's judgment as to whether regulations or standards
may be appropriate.
     This Health Assessment Document for Chloroform represents a comprehensive
data base that considers all sources of chloroform in the environment, the
likelihood of human exposures and the possible consequences to man and
lower organisms from its absorption.  This information is integrated into a
format that can serve as the basis for qualitative and quantitative risk
assessments, while at the same time identifying  gaps in our knowledge that
limit present evaluative capabilities.   Accordingly, it is expected that
this document may serve the information needs of many government agencies
and private groups that may be involved in decision making and regulatory
activities.  (As with all such EPA documents, this preliminary draft is
made available to the scientific community and the general public so that
comments of interested individuals and organizations may be considered, and
the latest scientific evidence incorporated, in  the final draft.   This
draft will  also be reviewed by the Environmental Health Advisory Committee
of EPA's Science Advisory Board at a subsequently announced public meeting).
                                      2-1

-------
                           3.   BACKGROUND INFORMATION




3.1. INTRODUCTION




     This section provides background information supportive of the  human health




effects  data  presented in  subsequent  sections.   It is  not intended  to  be a




comprehensive review of analytical methodology, sources, emissions, air concen-




trations, or environmental transport and fate information.   In order to fulfill




this  purpose  and since  the literature  concerning chloroform is  vast,  only a




portion of the available literature was included.  Those articles included were




chosen because  of their relevance  to  the topic at  hand  and because  they were




representative of the literature as a whole.




     To  provide  the most  complete  overview  possible,  some non-peer reviewed




information has been added, from the Chloroform Materials Balance Draft Report by




Rem et al. (1982).  This is an updated  version of  the original Level I Materials




Balance:  Chloroform (Wagner et al., 1980).  As  described  by Wagner et al.  (1980):




     "A  Level  I  Materials Balance  requires  the  lowest  level  of effort  and




involves  a  survey of readily  available  information for constructing  the




materials balance.  Ordinarily, many assumptions must be made in accounting for




gaps  in  information;   however,  all  are  substantiated  to  the  greatest  degree




possible.   Where possible, the  uncertainties  in  numerical values are given,




otherwise they are estimated.   Data gaps are identified and  recommendations are




made for filling them.   A Level I Materials Balance relies heavily on the EPA's




Chemical Information Division  (CID)  as  a source  of data and references involving




readily available  information.   Most  Level  I Materials Balance  are  completed




within a 3-6 week  period;  CID literature searches  generally require a  2  week




period to complete. Thus,  the total time required  for  completion  of  a  Level I




materials balance ranges from  5-7 weeks."
                                      3-1

-------
     Because a greater  level  of effort  went  into the 1980 and  1982  Materials
Balance  reports  on chloroform  than would  normally be  devoted  to  background
(non-health  effects)  information  in   an  EPA   health   assessment   document,
information in this chapter is drawn from these reports.   However,  because such
information has not been peer-reviewed and  includes  some major assumptions,  it
should not be used to support  any  regulations or standards regarding risks  to
public health.


3.2. PHYSICAL AND CHEMICAL PROPERTIES
     Chloroform (CHC1 }  (CAS registry number 67-66-3)  is  a member of a family of
halogenated saturated aliphatic compounds.  Synonyms for chloroform include the
following:
     Chloroforme (French)
     Chloroformio (Italian)
     Formyl trichloride
     Methane trichloride
     Methane, trichloror
     Methenyl chloride
Methenyl trichloride
Methyl trichloride
NCI-C02686
Trichloromethaan (Dutch)
Trichloroform
Trichloromethane
Table 3-1  lists important physical  properties.  Cloroform is a colorless, clear,
dense,  volatile  liquid  with  an ethereal  non-irritating odor  (DeShon,  1979).
Chloroform is nonflammable; however, when  hot  chloroform vapors are mixed with
alcohol  vapors,  the  mixture  burns  with  a greenish  flame.   At  25°C  and  1
atmosphere, a 1  ppm concentration of chloroform in air is equal to 4.88 mg/m .


     Chloroform decomposes with prolonged exposure to sunlight regardless  of the
presence of air (DeShon,  1979).   It also decomposes in the dark in the presence
                                      3-2

-------
                                  TABLE  3-1

                      Physical Properties of Chloroform0
Molecular weight
Melting point (°C)
Boiling Point
Water-chloroform azeotrope (°C)
Specific gravity (25/4°C)
Vapor density (101kPa, 0°C, kg/m3)
Vapor pressure
°C kPa
-30 1.33
-20 2.61
-10 4.63
0 8.13
10 13.40
20 21.28
30 32.80
40 48.85
Solubility in water
°C
0
10
20
30
Log octanol/water partition coefficient
Conversion factors at 25 °C and 1 atm
1 ppm CHC1- in air equals 4.88 mg/nr
1 rag/m CHC1 in air equals 0.205 ppm
119.
-63.
61.
56.
1.
4.

torr
10.0
19.6
34.7
61.0
100.5
159.6
246.0
366.4

g/kg H^O
10.62
8.95
8.22
7.76
1.97
38
2
3
1
48069
36















 Source:   DeShon,  1979
h,
 Hansch and Leo,  1979

'Calculated
                                3-3

-------
of  air.    The  principal  decomposition  products  include  phosgene,  hydrogen




chloride,  chlorine,  carbon dioxide,  and water.   Ozone  causes chloroform  to




decompose rapidly.




     Chloroform forms  a  hydrate in water  at  0°C (CHCl^  •  18HJD,  CAS  registry




number 67922-19-4); the  hexagonal  crystal  decomposes at  1.6°C  (DeShon,  1979).




Chloroform is chemically stable to water, having a  hydrolysis half-life of 3100




years  in neutral  water  at  25°C  (Mabey and  Mill,  1978);  the  half-life  of




chloroform in air from hydroxyl radical reactions is 78 days (Hampson,  1980).




     The hydrogen  atom on chloroform  can  be  removed in  the presence of  warm




alkali metal  hydoxide  to  form a trichloromethyl anion (DeShon,  1979).  This anion




can condense with  carbonyl compounds.   Both wet and dry  chloroform  will  react




with aluminum, zinc, and iron.




     Small amounts  of  ethanol  are used to  stabilize chloroform from oxidation




during storage (DeShon,  1979).




3-3. SAMPLING AND ANALYSIS



3.3.1.   Chloroform in Air.  Chloroform  in  air can be  analyzed by a number of




methods; however, the method of Singh et al. (1980) appears to be substantially




free of  artifact  problems  and  completely quantitative.   In this method, an air




sample in  a  stainless steel canister  at  32 psig is connected  to  a  preconcen-




tration  trap consisting  of a 4" x 1/16" ID stainless  steel tube  containing glass



beads, glass wool,  or  3% SE-30 on acid washed  100/120 mesh  chromosorb W.   The




sampling line and  trap, maintained  at 90°C,  are flushed  with  air  from the




canister;  then the  trap  is immersed in liquid  0- and air is passed through the




trap,  the initial and final pressure being noted (usually between 30 and 20 psig)




on a  high-precision pressure guage.   The ideal gas law can be used to estimate




the volume of air passed through the trap.   The contents of the trap are desorbed




onto a chromatography  column by  backflushing it with an inert gas while holding

-------
the trap at boiling water temperature.  An Ascarite trap may be inserted before




the chromatography column to remove water.  Suitable columns include 2Q% SP-2100




and 0.1? CW-1 500 on Supelcoport (100/120 mesh, 6' x 1/8" stainless  steel) and 20%




DC-200 on Supelcoport (80/100 mesh, 33' x 1/8" Ni).  Both columns can be operated




at 45°C with a carrier gas flow of 40 mS,/min on the former column and 25 m£/min on




the latter.   An electron capture  detector operating  at 325°C was  found  to be




optimum.  It should be noted that the above authors found Tenax to be unsuitable




for air analyses  because  of the  presence  of artifacts  in the  spectrum  from




oxidation of the  Tenax monomer.   In addition,  when Tenax is used as a sorbent,




safe sampling volumes (i.e., that volume of air which,  if sampled over a variety




of circumstances, will not  cause significant  breakthrough) should  be adhered to.




Brown and Purnell (1979) determined the safe volume  for chloroform  per gram Tenax




to be 9.3 I  (flow rate  5-600 mS,/min; CHC1 cone. <250 mg/rn^ temp, up to  20°C) with




a safe desorption  temperature of 90°C.




     The detection limits of this method were not specified  and are dependent on




the volume of air  sampled.    Analyses as low  as 16  ppt have  been reported using




this method (Singh et al.,  1930).




3-3.2.  Chloroform in Water.  Chloroform in water can  be analyzed by the purge-




and-trap mejhi^od  (Method  502.1 )  as recommended by  the Environmental Monitoring




and Support Laboratory of the U.S. EPA (1981 a).   In this method, an inert gas is




bubbled through 5 ml of water at  a  rate of 40 m^/minute  for 11  minutes, allowing




the purgable  organic compounds  to  partition  into  the  gas.   The  gas  is passed




through a column containing  Tenax GC at 22°C, which traps most of the organics




removed from the  water.   The Tenax column is then heated rapidly to 130°C and




backflushed  with  helium (20-60  mil/minute,  4  minutes)  to  desorb  the  trapped




organics.   The  effluent of  the  Tenax  column is passed into an  analytical gas




chromatography column packed with 1 £ SP-1000 on Carbopack-B (60/80 mesh, 8'x 0.1"
                                      3-5

-------
I.D.) maintained at 40°C.   The  column is then temperature programmed starting at




45°C for  3  minutes and increasing at 8°C/minute  until  220°C is reached; it is




then held there for 15 minutes or until all  compounds  have eluted.   A halogen-




specific detector  (or GC-MS) having a  sensitivity of 0.10 |ig/Ji with a relative




standard deviation of <10$ must be used.




3-3-3   Chloroform in Blood.   Chloroform in blood  can be  analyzed  by using a




modified purge-and-trap method (Pellizzari et al., 1979).  This method involves




diluting an aliquot of whole blood (with anticoagulant) to =50 ml with prepurged,




distilled water.   The mixture  is  placed in a 100 mi, 3~neck round bottom flask




along with  a teflon-lined magnetic stirring  bar.  The necks  of  the flask are




equipped with a  helium inlet, a Tenax trap, and  a thermometer.  The Tenax  trap is




a 10 cm x 1.5 cm i.d.  glass tube containing pre-extracted (Soxhlet, methanol, 24




hrs) and  conditioned  (2?0°C,  30  m£/min helium  flow,  20 min)  35/60  mesh Tenax




(=1.6 g, 6 cm).  The sample is then heated to 50°C and  purged with a helium flow




rate of 25 mJt/min for  90 min.  Analysis can be performed as indicated in  Section




3.3.2.




3.3.4.   Chloroform in Urine.  Chloroform in urine  can  be  analyzed by using an




apparatus identical to the one described in Section 3.3.3, using  25 m& of urine



diluted to 50 mil instead of blood,.



3-3.5.  Chloroform in Tissue.  Chloroform in tissue can be analyzed by using an




apparatus identical to the one described  in  Section  3.3.3, using 5 g of tissue




diluted to 50 mS, and macerated  in an ice bath instead of blood.  The purge  time is




reduced to 30 minutes.




3.4. EMISSIONS FROM PRODUCTION AND USE




3.4.1.  Emissions  from Production.




     3.4.1.1.   DIRECT  PRODUCTION  —  Chloroform is produced  commercially in the




United States by two methods, chloririation of methane and chlorination  of methyl
                                       3-6

-------
chloride  produced  from methanol  and hydrogen  chloride  (Wagner et  al.,  1980;

DeShon, 1979).  The chemistry is summarized by the following reactions:



     Methane Chlorination

     ^CHjj + 10C12 ------- »• CH3C1 + CH2C12 + CHC13 + CCl^ + 10HC1


     Methanol Hydrochlorination

                    Catalyst
     HC1 + CH OH ----------- »• CH-C1 + H?0
             J     280 - 350°C  J
     3CH-C1 + 6C12 --------- >• CH2C12 + CHC13 + CCl^ + 6HC1
The methanol process accounts for 7^% of capacity, while methane accounts for 26%

of capacity (SRI International, 1983).

     In the chlorination of methane, natural gas is directly chlorinated in the

gas  phase  with chlorine  at  485-510°C  (Anthony,  1979).   The  product  mixture

contains all chlorinated methanes, which are removed by scrubbing and separated

by fractional distillation.

     In the  second process,  gaseous  methanol  and HC1 are  combined  over  a hot

catalyst to  form  methyl  chloride  (Ahlstrom  and Steele,   1979).    The  methyl

chloride is  then  chlorinated  with chlorine to  produce  CH  Cl?,  CHC1-, and CC1.

(DeShon, 1979).  The chlorination conditions for both processes can  be adjusted

to optimize chloroform production.

     United States production  is carried out by five  manufacturers at  seven sites

summarized in Table 3-2.

     The annual production of chloroform in the United States has risen from 35

million kg  (77  million Ibs)  in   1960  (DeShon,  1979) to  >18M million  kg  (405

million Ibs) in 1981  (USITC,  1982) with few declines.
                                      3-7

-------
                                     TABLE  3-2




              Chloroform Producers,  Production Sites,  and Capacities*
Producer
Diamond
Shamrock
Dow Chemical

Linden Chemicals
and Plastics, Inc.
Stauffer
Chemical Co.
Vulcan
Materials Co.
TOTAL
Production
Site
Belle, WV
Freeport, TX
Plaquemine, LA
Moundsville, WV
Louisville, KY
Geismar, LA
Wichita, KA

Capacity
Millions of kg
(Millions of Ibs)
18(40)
45(100)
45(100)
14(30)
34(75)
27(60)
50(110)

233(515)
Process
Methanol
Methane
Methanol
Methanol
Methanol
Methanol
Methanol and
Methane

•Source:  SRI International,  1983.
                                          3-8

-------
     3.4.1.1.1.   Chloroform  Emissions from the Methane  Chlorination Process —




Rem et al .  (1982)  reported that air emissions could come from process vents, in-




process  and  product storage  tanks,  liquid waste streams,  secondary emissions




(handling and disposal of process wastes), and fugitive emissions from leaks in




process valves, pumps, compressors and pressure relief valves.




     The emission factors  they   calculated  were based  on  a typical  methane




chlorination facility  as  reported by U.S.  EPA (1980a) having a  total  chloro-




methane capacity of 2 x 10  metric tons (441 x 10  Ibs),  operating continuously




(8760 hr/yr), and having a product mix of 20% CH_C1,  45?  CH2C12,  25% CHClo, and
    CClj..  The emission factor  for the uncontrolled recycle methane  inert  gas




purge  vent  for the  above  plant (0.014 kg/metric  ton)  was calculated  from  an




hourly CHC1-, emission rate  of  0.071 kg/hr reported by Dow Chemical Company for a




46,000 metric ton/yr facility  assuming  continuous (8760 hr/yr) operation (Beale,




n.d.).   The  uncontrolled  emission factor  for  the  distillation  area  emergency




inert gas vent (0.032 kg/metric ton)  was calculated from an emission factor  for




volatile organic compounds  (VOC )  of  0.20 kg/metric ton  of  total chloromethane




production and composition  data  showing chloroform to be  4? of  VOC (U.S. EPA,




1980a).  In-process and product storage emissions  (0.91  to  0.80  kg/metric ton,




depending on controls) were calculated  from emission equations for breathing  and




working losses from AP-61  (U.S. EPA,  198lb) assuming tanks to be half-full, have




95% emission  controls when present,  and a 12°C diurnal  temperature  variation




(U.S. EPA,  1980b).   Rem et al .  (1982) calculated the total chloroform emissions




to air  to  be 70.2 metric tons  (155  x  10  Ibs) by multiplying the  appropriate




factors  by  plant   capacity  use  after  including  secondary  emissions  (0.21




kg/metric ton) and fugitive emissions (5.5  kg/hr).




     Releases of  chloroform to water  come from scrubbers, neutralizers ,  and




cooling water  (Rem  et  al . ,  1982).   Based  on a 300 ppm CHC1- content  in  total
                                      3-9

-------
wastewater discharges averaging 68 il/min and assuming that 90% would volatilize,


Rem et al. (1982) calculated a release  factor of 0.023 kg/metric ton.  They then


added to this the emission from indirect contact cooling water (100 ppm,  5800 1


cooling water/metric ton CHC1_, 905?  evaporation) to calculate a release rate of

                                  o
3.3 metric tons of CHC1~ (7-3 x 10J  Ibs) per year to water.


     No quantifiable data were available to Rem et al.  (1982) regarding release


of chloroform to land from methane chlorination.


     3-4.1.1.2.   Chlorination Emissions  from  the Methanol  Hydrochlorination-


Methyl Chloride Chlorination Process —   Chloroform  emissions  to  air  come from


process  vents,  in-process  and product  storage  tank  emissions,   and  fugitive


emissions from leaks in  valves, pumps,  compressors,  and  pressure  relief  valves


(Rem  et  al.,  1982).   Rem  et al. (1982)  used  an uncontrolled  emission  factor


reported by Vulcan Materials Company for process vents (Hobbs, 1978), and assumed


continuous  (8,760  hr/year)  operation,  and  controls  sufficient  to  reduce


emissions 8Q% to obtain the emission factor for controlled process vents  (0.003


kg/metric ton for controlled;  0.015 kg/metric ton for  uncontrolled).   Storage


emissions  (0.176   kg/metric  ton  for  controlled,  0.88  kg/metric   ton  for


uncontrolled) were calculated from Hobbs  (1978) and from emission equations for


breathing and working losses from AP-42 (U.S. EPA, 198lb), assuming tanks to be


half-full, have 80% emission reduction controls, and a 12°C  diurnal temperature


variation  (U.S.  EPA,   1980b).    Fugitive  emission factors  (3-32  kg/hr  for


uncontrolled;  1.08  kg/hr for  controlled) for  volatile  organic  compounds were


used  (U.S.  EPA,  1980c)  along with  a  control  factor  of 67-5/6  based on leak


detection and repair (U.S.  EPA, 1980b).  Emission rates  were then calculated to


be 196 metric tons  (432 x 103 Ibs) based on plant capacity,  capacity use (66%},


the  level  of emission  controls used  at  each plant, and continuous  operation


(8,760 hr/year).
                                      3-10

-------
     Rera et al. (1982) assumed that releases of chloroform to water  from methyl




chloride  chlorination resulted  from  contamination of  cooling water  and  from




spent  acid  and spent caustic  streams.   For  cooling  water contamination,  they




assumed minor spills and leaks resulted in contamination of 100 mg chloroform/Ji




cooling  water,  that  5,800 8,  of cooling  water  per  metric  ton  of chloroform




produced was consumed, and  that 90/6 of the chloroform evaporated.   For spent  acid




and spent caustic streams  they assumed  that  0.04 kg of chloroform are released




for every metric ton  of chloromethane  produced.   They further stated that  this




release factor was not considered to be very reliable;  however, in the absence of




better data they used this factor to  approximate  emissions.   In addition,  they




assumed that 1/3 of the  chloromethane production consists of chloroform and  that




90% of the released chloroform evaporates.  Using these assumptions, Rem et al.




(1982) calculated that 0.070  kg of chloroform is released to water per metric ton




of chloroform produced,  or 8.0 metric tons (17.6  x  10  Ibs)  of chloroform were




released to water based on 1980 production levels.




     Rem et al. (1982) reported that the bottoms from chloroform distillation in




the methyl chloride chlorination  process are  the  feed for carbon tetrachloride




and perchloroethylene production, and that during their production a residue is




formed  that  contains  chloroform.    This  residue  is  landfilled or deep-well




injected.  This represents the only  known release of  chloroform from carbon-




tetrachloride/perchloroethylene production.




     Rem et al. (1982) assumed that  1.02 kg of residue is produced/metric ton of




chloroform from methyl  chloride production  (see  Wagner et  al.,  1980).   They




further  assumed  (as  did  Wagner et  al.,   1980)  that  18.4$  of the  residue is




chloroform and that 25% is landfilled. This results  in a release factor of 0.047




kg chloroform per metric ton of chloroform produced, or 5.4 metric tons  (12 x 10^




Ibs) based on 1980 production.
                                     3-11

-------
     3.^.1.1.3-  Summary of Direct Production — Direct production emits some 283




metric tons into the environment on  an  annual basis.  Greater  than  95%  (266.2




metric tons) of this is  emitted into  the air.  Direct production  accounts for =3$




of all environmentally  released chloroform,  and  -3.6%  of all chloroform released




to air.  Table 3-3 summarizes chloroform discharges from direct production.




     3.^.1.2.  INDIRECT PRODUCTION




     3.4.1.2.1.  Chloroform  Formation During Ethylene Bichloride Production —




Ethylene dichloride  (EDC) is produced by two methods,  direct  chlorination and




oxychlorination,  and is  used  principally  for vinyl  chloride monomer   (VCM)




production.  A  combination  of the two methods  is  used by most VCM  production




facilities  in  a process known as  the balanced process since the  HC1  from the




dehydrochlorination of EDC is used to produce more EDC from ethylene, the major




products of the overall reaction being VCM and H_0.









     Direct Chlorination
     Balanced Process
          HC1 + CH  -- CHC1
                                      3-12

-------
                                   TABLE  3-3
                                                           *
                  Chloroform Discharges from Direct Sources
                               Environment Release  (metric tons/year)
Source                       Air         Water        Land        Total
Methyl Chloride
  Chlorination              196            8             5.4        209.4


Methane Chlorination         70.2         3.3           	         73.5
Total:                      266.2        11.3           5.4        282.9
 Source: Rem et al. (1982)
                                         3-13

-------
     Chloroform is  formed  as a  byproduct  during EDC manufacture.   Rem  et al.




(1982)  estimated  chloroform  emissions to  air  from EDC  production  based  on




emission  factors  developed  by  the EPA  during  field  studies of  domestic EDC




production facilities (see  U.S. EPA,  1980d).  Domestic production facilities are




listed in Table 3-4.  Emission sources  (e.g., process vents, fugitive emissions,




storage) for each  plant,  plant capacity, capacity utilization  (56/6), and control




technology were all combined with the emission factors to determine the overall




chloroform  emissions  at the  current  level  of  control.    According  to  their




calculations, chloroform emissions to  the  atmosphere are =760 metric tons/year




(1,675 x 103 Ibs/year).




     Chloroform releases to water may  occur  during the  discharge of wastewater




from the process;  however,  the amount of chloroform present is unknown.




     Chloroform releases to land from  EDC  production reportedly  occur when the




light ends from EDC  distillation are landfilled (Rem et al.,  1982).  An estimated




217 metric tons (478 x 103  Ibs)  were landfilled in  1980.




     3.4.1.2.2.  Chlorination of Drinking Water — Chloroform in drinking water




arises when  humic  substances  or  methyl ketones  (e.g.,  acetone)  in water react




with a hypochlorite anion (Stevens et al.,  1976; NAS, 1978).  Hypochlorite is the



principal reactant  in chlorinated water above  pH 5.  Chloroform is produced by




the haloform reaction outlined below:








     R-COCH  + 30C1~	>• RCOCC1  + 30H~




     R-COCC1  + OH~	>• RCOO~ + CHC1








     Rem et al. (1982) used the  data from  the National  Organics Reconnaissance




Survey (NORS) (Symons et al., 1975) and the National Organics Monitoring Survey




(NOMS) (U.S. EPA,  n.d.) to  estimate the concentration of chloroform in drinking
                                      3-14

-------
                                   TABLE  3-4
        Ethylene Bichloride Producers,  Production Sites  and Capacities'
Producer
     Production
        Site
   Capacity
millions of Kg
(millions of Ibs)
Alantic Richfield Co.
  ARCO Chem. Co., div.

Borden Inc.
  Borden Chem. Div.
    Petrochems. Div.

Dow Chemical. U.S.A.
E.I. du Pont de Nemours and Co.,
  Conoco Inc., subsid.
    Conoco Chems. Co. Div.

Ethyl Corp.
  Chems. Group
Formosa Plastics Corp.  U.S.A.
Georgia-Pacific Corp.
  Chern. Div.

The BF Goodrich Co.
  BF Goodrich Chem.  Group

  Convent Chem. Corp.,  subsid,
PPG Iridust.,  Inc.
  Chems. Group
    Chera. Div.
     Port Arthur,  TX       204       (450)
     Geismar, LA           231       (510)

     Freeport, XX          726      (1600)
     Oyster Creek,  TX      476      (1050)
     Plaquemine, LA        862      (1900)
Inc.
318
102
249
386
(700)
(225)
(550)
(850)
     Lake Charles,  LA      524      (1155)
     Baton Rouge,  LA
     Pasadena,  TX

     Baton Rouge,  LA
     Point Comfort,  TX
     Plaquemine,  LA        748      (1650)
     Deer Park,  TX         145       (320)
     La Porte, TX          719      (1585)
     Calvert City, KY       454      (1000)
     Convent, LA           363       (800)
     Lake Charles,  LA      1225      (2700)
                                   3-15

-------
                               TABLE  3-4  (cont.)
Producer
Production
   Site
   Capacity
millions of Kg
(millions of Ibs)
Shell Chem. Co.
Union Carbide Corp.
  Ethylene Oxide Derivatives Div.
Vulcan Materials Co.
  Vulcan Chems., div.
Deer Park, TX
Norco, LA
Taft, LA
Texas City, TX
Geisraar, LA
635
544
 68'
 68
159
(1400)
(1200)
 (150)
 (150)
 (350)
                                                  TOTAL  9,206     20,295
 Source: SRI International,  1983
Captive use only
                                          3-16

-------
water.   These  surveys  provided  information on chloroform concentrations in 137




U.S.  cities.   To  determine the  amount  of  chloroform  generated,  the  authors



multiplied the volume  of water treated by  each  city by the chloroform concen-



tration in the drinking water.  The amount of chloroform generated by each city



was then summed and divided by the volume of water generated to give a weighted



average  concentration  of 41 ^g/ii.  This  was then multiplied  by  the estimated


                                              1 3
volume of water chlorinated annually (4.6 x  10   i/year) to yield the amount in



the U.S. (1,900 metric  tons, 4.2x10  Ibs).  Rem et al . ( 1982) stressed that this



value probably represents a minimum since  NORS was conducted during the winter



(hence, chloroform levels were low) and NOMS samples were iced when taken (hence,



may be lower than if allowed full contact time).   Thus,  the actual value may be



higher than this estimate.



     3.4.1.2.3-  Chlorination of  Municipal  Sewage — Chlorination of municipal



sewage results in  increased chloroform concentrations (NAS, 1978).   Municipal



wastewater generally  contains  a  lower concentration of  chloroform precursors



(humic materials)  than do  ambient waters;  therefore, the  amount  of chloroform



generated from wastewater Chlorination is  smaller (wagner et al . ,  1980).  Rem et



al . (1982) calculated  chloroform  production  from  wastewater  treatment  based on



analyses of the secondary effluent from 28 municipal  plants published by the EPA



(U.S. EPA, 1979a).  These  analyses showed that the  average  chloroform  concen-



tration increased by 9 >>g/&, from 5 to 14 ^g/£.  Rem et  al . (1982) assumed that
all municipal  wastewater was  chlorinated  and multiplied  the average  concen-



tration increase by the municipal wastewater  flow  (9.7  x 10   £/day)  listed in



U.S. EPA (198lc).  By this method,  320 metric tons (0.7 x  10  ibs) was calculated



to be produced annually.



     3-4.1.2.4.  Chlorination of Cooling Waters — Cooling water used in electric



power generating plants  is treated with chlorine as a biocide to prevent fouling
                                     3-17

-------
intake screens and condensers in both once-through and closed cycle systems (U.S.


EPA,  1980e).   Bern  et al. (1982) calculated  chloroform production based on the


size of the average power plant (hence, the volume of water required), the type


of cooling system used (once-through or recirculating), and the fact that 65% of


all power plants  chlorinate cooling water.  They calculate that 72 metric tons of

                     o
chloroform  (160  x  10  Ibs)  are discharged directly  into  water by once-through


systems and !90 metric tons of chloroform  (420  x  10  Ibs)  are  emitted into the


air by recirculating systems.  A total of 262 metric tons of chloroform (580 x


10  Ibs) are  produced annually from cooling water chlorination.


     3.4.1.2.5.  Chlorination in the  Pulp and Paper  Industry — Pulp and paper


mills emit  more  chloroform to the  environment than  any  other  single  source.


Chloroform is  produced during the bleaching of wood pulp, a process that whitens


the final paper product.  Rem et al.  (1982)  based their  estimation on information


contained in  a number  of documents concerned  with the pulp and paper industry


(U.S. SPA,  1980f; NCASI, 1977;  Metcalf and Eddy Inc.,  1972; TAPPI, 1963).  From


the  opeBating conditions, analytical data,   and  production steps  detailed in


these documents,  Rem et al. (1982)  determined  the quantity of chloroform produced


for each of nine different types of mills for which monitoring data existed and


applied these valuers to all mills for which no values from sampling existed.  The


authors determined that chloroform is  emitted at three different stages: into the


air- during the bleaching process itself, into the air during the detention time


in wastewacer treatment plants, and into the water from  treatment  plant effluent.


The amount, of  chloroform produced annually was  calculated to be 128 metric tons


(282 x  10   Ibs)  released  to  air  during bleaching operations, 3,985 metric tons


(8.78 x 10  Ibs)  released to  air from  wastewater detention (4,113 metric tons to

                                    o
air), and 298  metric tons  (657 x  10  Ibs) discharged into water from treatment


plants (4,411 metric tons or 9.72 x 10  Ibs total).
                                      3-18

-------
     3.4.1.2.6.   Chloroform Production from  Combustion of Leaded  Gasoline —




Chloroforom has been reported to  be a component of automobile exhaust (Harsch et




al., 1977).  Its presence is reportedly the result of using ethylene dichloride



and  ethylene  dibromide as  lead  scavengers in  leaded gasoline  (Lowenbach and




Schlesinger Associates, 1979).  Rem et  al.  (1982) cite other sources which state




that chloroform is  not  formed during the combustion of  leaded  fuel containing




ethylene dichloride.  The Emissions Testing and Characterization Branch of EPA's




Environmental Sciences Research Laboratory measured chlorocarbon emissions from




automobiles using leaded  gasoline and  found no  chloroform.  Rem et al.  (1982)




then reason  that  even  if  chloroform were  present  in automobile  exhaust, the




decrease in the usage of leaded gasoline will decrease the amount of chloroform




produced.



     The authors cite  an  EPA report  (U.S.  EPA,  1982a) which states that leaded




gasoline consumption is expected to drop by >75%, from 34 x 10  gallons to 8.3 x




109  gallons/year.   Rem et al. (1982) used the estimate of Wagner et al. (1980)




that W> of the ethylene dichloride  in gasoline would be converted to chloroform,




and  the new lead phase-down regulations (U.S.  EPA,  1982b) to calculate an annual




emission rate of 180 metric  tons  (397 x  103 Ibs)  in 1983 and 44 metric tons  (97 x




103  Ibs) by  1990.



     3.4.1.2.7.    Chloroform  Formation During Atmospheric  Trichloroethylene




Decomposition   —   Trichloroethylene  is  a  major   industrial   solvent   used




principally  for  vapor degreasing  of   fabricated  metal parts  (66/O  (Chemical




Marketing Reporter, 1981), and the majority of each year's  production is used for




replacement  of evaporative  loss  to the environment  (Rem et al.,  1982).   The




postulated  formation  of  chloroform  during   the  atmospheric decomposition of




trichloroethylene is based on laboratory experiments in which trichloroethylene,




N02,  H20,  and  a hydrocarbon  mixture  were  irradiated with  light  having the
                                      3-19

-------
intensity and spectral distribution of the  lower  troposphere  (U.S.  EPA,  1976).




Dichloroacetylchloride, phosgene,  chloroform,  and  HC1 were detected as products.




Rem et al. (1982) did not describe the method they used to determine the  amount




of chloroform  produced  from  this  reaction;  however, they estimated 780  metric




tons of chloroform are produced annually.




     3.4.1.2.8.   Miscellaneous Indirect  Source  — Wagner  et al.  (1980)  have




listed a number of other indirect sources  of  chloroform  that are difficult if not




impossible to quantify.  These  sources are:   chlorination wastewaters from the




textile industry, the food  processing  industry, breweries, combustion of tobacco




products  treated  with  chlorination  pesticides,   thermal  decomposition  of




plastics,  biological  production  in  red marine  algae,  and  the   reaction  of




chlorinated pollutants with humic  substances in natural waters.




     3.4.1.2.9.  Summary of  Indirect  Production  — Chloroform  is  produced and




emitted into the environment from  a variety of indirect sources.  These sources




account for =£4£ of all chloroform air emissions, 98?  of water discharges, and




36$ of all land discharges  (84.6$  of all environmental releases).




     The  environmental discharges  of  chloroform  from  indirect   sources  are




summarized in Table 3-5.



3.4.2.  Emissions from Use.  Chloroform is used consumptively  for the  production




of chlorodifluoromethane or Fluorocarbon-22 (accounts  for 9056 of domestic 1982




production, 60% for refrigerants,  30% for fluoropolymer) and  for exports (3% in




1982) (Anonymous, 1983).  Non-consumptive uses include its use as an extraction




solvent;  as  a solvent for penicillin, alkaloids, vitamins, flavors, lacquers,




floor polishes, artificial  silk manufacture, resins, fats, greases,  gums, waxes,




adhesives, oils,  and rubber;  as  a dry  cleaning  agent; as  an intermediate in




pesticide and dye manufacture;  and as  a  fumigant  ingredient  (Rem et al.,  1982;




DeShon,  1979;  Merck  Index, 1976).  The  great majority of chloroform used non-
                                      3-20

-------
                                   TABLE 3-5
                                                             o
                  Chloroform Discharges  from Indirect Sources
Environmental Release (metric
Source
Pulp and Paper Mills
Drinking Water Chlorination
Ethylene Dichloride
Manufacture
Trichloroethylene
Photodegradation
Municipal Wastewater
Chlorination
Cooling Water Chlorination
Automobile Exhaust
TOTAL
Air
4113
0
760
780
0
190
180
6023
Water
298
1900
	 b
0
320
72
0
2590
Land
0
0
217
0
0
0
0
217
tons/ year)
Total
4411
1900
977
780
320
262
180
8830
aSource:  Rem et al., 1982
 minor releases possible
                                           3-21

-------
consumptively is emitted into the environment since (except for expansions) the




chloroform  purchased  for these  uses  is make-up  solvent used to  replace  that




amount not recovered from processes (Wagner et al., 1980).




     3.4.2.1. EMISSIONS FROM PHARMACEUTICAL MANUFACTURING — Chloroform is used




as an extraction solvent during the manufacture of  some antibiotics and ster|ijo\ds,




and  during  the manufacture  of  certain  other  biological  and natural  pharraa-




ceuticals (Rem et al.,  1982).  It is also  used as a chemical intermediate.  Based




on a Pharmaceutical Manufacturing Association (PMA) Survey,  Rem et al.  (1982)




reported that =1000 metric tons of chloroform are released into the environment




(no year was specified; the PMA report was  dated 1978).   The distribution was as




follows:  57.0? (570  metric  tons)  to  air,  4.6? (46 metric tons)  to water, and




38.4? (384 metric tons) to land.




     3.4.2.2. EMISSIONS  FROM FLUOROCARBON-22 PRODUCTION -- The  single largest




use  of  chloroform  is for  Fluorocarbon-22  production   (Chlorodifluoromethane,




CHC1F ).  Fluorocarbon-22 producers and production sites are listed in Table 3-6.




Chloroform release to the environment  can occur  from  process emissions, fugitive




emissions, and storage emissions (Rem et al., 1982).   Rem et al. (1982) reported




that  the first source  listed  does   not  represent  a   significant source  of




chloroform   emissions   based  on  the  design  of  Fluorocarbon-22  production




facilities and the  process description.




     Storage  emission estimates were  based  on U.S.  EPA (1980g)  for chloroform




feedstock storage  in fixed  roof  tanks.   Rem et  al. (1982)  reported  that the




Allied  Chemical facility  at  El Segundo,  California,  uses a control system that




results in  complete capture  of chloroform  vapors.  Du  Pont  (and  all others by




assumption)  use refrigerated condensers that reduce the uncontrolled emission




factor  of 2.5 kg/metric  ton  by 66?.  Fugitive  emissions from leaks in valves,




pumps,  compressors, and relief valves  was estimated to result  in an  uncontrolled






                                      3-22

-------
                                  TABLE 3-6
             Chlorodifluoromethane Producers and Production Sites
Producer
Production
   Site
Allied Corp.
  Allied Chem. Co.
E.I. du Pont de Nemours  and  Co., Inc.
   Petro^hems. Dept.
   Freon  Products  Div.
Pennwalt Corp.
  Chems. Group
     Fluorochemicals  Div.
Baton Rouge, LA
Danville, IL
El Segundo, CA
Deepwater, NJ
Louisville, KY
Calvert City, KY
 Source:   SRI  International,  1983
                                   3-23

-------
emission rate of 0.75 kg/metric ton.   Total emissions to air were calculated by


Rem et al. (1982) to be 139 metric tons/year by multiplying the emission factors


by 1980 Fluorocarbon-22 production (97,500 metric tons).


     No estimate was made for emissions to wastewater because of a lack of data.


Emissions  to  land were  based on the  reported  practice  of landfilling  spent


catalyst by Allied.  Rem et al.  (1982)  assumed a catalyst contamination level of


1056 and a total emission of 2.0 metric tons/year.


     3.4.2.3- EMISSIONS  FROM HYPALON*  MANUFACTURE — Hypalon*  is  a chemically


resistant  synthetic  rubber made by  substituting  chloride and  sulfonyl  groups


onto  polyethylene.  The process involves  dissolving  polyethylene in chloroform


followed by reaction with chorine and sulfur dioxide.   Based on a Du Pont report,


Rem et al. (1982) estimated that 54.9 metric tons of chloroform were emitted into

                      •
the  air from  Hypalon   manufacture  in  1980,  based  on an  emission  inventory


conducted  by the Texas Air Control Board, Austin,  Texas.  No  information was


available for water or land emissions.


     3.^.2.4.   CHLOROFORM EMISSIONS FROM  GRAIN  FUMIGATION --  Chloroform  is a


registered pesticide for use on certain insects that commonly infest stored raw


bulk grains and is present in only one  product (Rem et al., 1982). This product,

          «
Chlorofume   FC.30  Grain  Fumigant  (Reg. No.   5382-15),  marketed  by  Vulcan


Materials  Company,  contains  12.2%  chloroform,  20.4?  carbon disulfide,  and 7.4?


ethylene dibromide.  Originally registered in 1968,  the EPA issued a "Notice of


Presumption Against Continued Registration of a Pesticide Product — Chloroform


(Trichloromethane)"  in  1976   because  of  oncogenic  effects  in  rats and  mice.


Continued study resulted in returning it to the normal registration  process  (U.S.


EPA,  1982c).   Based  on  a  personal  communication  with  D.  Lindsay of  Vulcan


Materials, Rem et al. (1980)  estimated that between 10,000 and 12,000 gallons of


chloroform/year  were  used in  grain  fumigants in  the  United  States.   Vulcan



                                      3-24

-------
reported  1981  sales of  7,000 gallons of Chlorofume   in 1981 or  5054 gallons


(19,131 &).  Based on its density, 28,400 kg (28.4 metric tons) was released to


the environment (air) in this way.


     3-4.2.5.  CHLOROFORM LOSSES FROM LOADING AND  TRANSPORTATION  — Rera et al.


(1982) estimated  chloroform  losses  from loading ships,  barges, tank cars,  and


tank trucks.  The method was based on the  degree of chloroform saturation of the


air expelled  from tanks during filling,  temperature,  vapor  pressure,  control


efficiency, and filling methods as described  by U.S. EPA  (1979b) and Environment


Reporter (1982).  The U.S. mode of  transportation was  taken  from  Sax (1981) as


follows:    rail,  40.3?;  barge,  47.856;  and truck,  11.9?.  Loading  losses  were


calculated to be 40.9 metric tons (90,200 Ibs).


     Transit  losses  result from  temperature and barometric  pressure changes.


The losses were  assumed to be the same for  barges, tank  trucks, and rail cars and


were estimated from the following equation:


                                  LT= 0.1 PW

                                          o
     where  LT= transit loss, Ib/week -  10J gal  transported


             P= true vapor pressure  of transported liquid,  psia


             W= density of condensed vapors Ib/gal





No reference or justification for the use  of  the formula was presented.  By this


method, and assuming 1 week transit  time, Rem et al. (1982)  calculated that  49-2


metric  tons   (0.11  x  10  Ibs)  chloroform   were  lost  to  the air  using  1980


production values.


     3.4.2.6. MISCELLANEOUS USE EMISSIONS — Rem et al.  (1982) cited the previous


materials  balance  (Wagner  et  al.,  1980)  in  predicting  the  emissions  from


chloroform contamination of  methyl  chloride,  methylene  chloride,  and  carbon


tetrachloride.  Chloroform is present to  some  extent in these products since  they
                                     3-25

-------
are all made by the same  process.   Assuming a contamination level  of  7.5 ppn,


17.5  ppm and  150  ppm for  methyl  chloride,  methylene  chloride,  and  carbon


tetrachloride, respectively, Rem  et al.  (1982) estimate  that  releases to air,


land and water would be 9.8, 0.6,  and 0.15 metric tons, respectively.


     Chloroform is also used in  a  variety of  products (see Section 3-4.2) and as


a  general  solvent.   Rem  et al.   (1982)  estimate  that,  while these  uses  are


generally declining,  laboratory  uses  in  particular  may  account  for 8.5?  of


production or  14,200  metric tons  of chloroform.   Rem  et al.  (1982),  however,


estimated the range of uncertainty to be + 50%.


     3.4.2.7.    SUMMARY  OF  CHLOROFORM  DISCHARGES  FROM  USE  —  Chloroform


discharges  from  manufacturing  facilities  that   use  chloroform  as a  process


ingredient account  for 12%  of  all chloroform emissions  to air,  1.8/t  of water


discharges, and 63/6 of all land  discharges,  or  12.7% of chloroform discharges to


all media.  Table 3-7 summarizes  chloroform  discharges to all media.


3.4.3.   Summary.   Chloroform is  produced  by  direct  and  indirect processes.


Direct   production  accounts  for   184  million   kg annually,  while  indirect


production accounts for =8.8 million kg annually.


     Direct  production of chloroform  and  processes  associated with its  use

                                            «
(i.e.,  Fluorocarbon-22 production,  Hypalon  manufacture,  loading  and  transit


losses,  grain  fumigation,  pharmaceutical  use) emit some  1.6  million kg to the


environment.  Virtually all of the indirectly produced chloroform is  emitted into


the  environment;  the  total  amount  of  chloroform emitted  is  =10.4  million kg.


This  represents  -5.6%  of  direct production.  The relative source contributions


from  all quantifiable sources are listed in  Table 3-8.


3.5.  AMBIENT AIR CONCENTRATIONS


      Monitoring  data  for  a number of U.S. and world locations  are  presented in


Table 3-9.  For the most  part,  ambient concentrations  remain <1000  ppt (1 ppt  =



                                      3-26

-------
                                   TABLE 3-7
                        Chloroform Discharges from Use
Environmental Release (metric
Source
Pharmaceuticals
Chlorodifluoromethane
Manufacture
Loading and Transit Losses

-------
                                                                            TABLE 3-8
                                                          Relative Source Contribution  for Chloroform
LO
 I
Environmental Release (metric
Source
Pulp and Paper Mills
Drinking water Chlorination
Pharmaceuticals
Ethylene Bichloride Manufacuture
Trichloroethylene Photodegradation
Municipal wastewater Chlorination
Cooling water Chlorination
Methyl Chloride Chlorination
Automobile exhaust
Chlorodifluoromethane Manufacture
Loading and Transit Losses
Methane Chlorination
•
Hypalon Manufacture
Grain Fumigation
Secondary Product Contamination
Laboratory Usage
TOTAL
Air
4113
0
570
760
780
0
190
196
180
139
90.1
70.2
54.9
28.4
9.8


7181.4
Jof
Total0
39
0
5.5
7.3
7.5
0
1.8
1.9
1.7
1.3
0.9
0.7
0.5
0.3
0.1


68.8
Water
298
1900
46
	 b
0
320
72
8
0
	
0
3.3
	
0
0.6


2647.9
Jof
Total
2.9
18
0.4
	
0
3.1
0.7
0.1
0
	
0
0.03
	
0
0.006


25.4
Land
0
0
384
217
0
0
0
5.4
0
2
0
	
	
0
0.2
	
608.6
Jof
Total
0
0
3.7
2.1
0
0
0
0.1
0
0.02
0
	
	
0
0.002
	
5.8
Total
4411
1900
1000
977
780
320
262
209.4
180
141
90.1
73.5
54.9
28.4
10.6
	
10,438°
tons/ year )
*« c
Total0
42
18
10
9.4
7.5
3.1
2.5
2.0
1.7
1.4
0.9
0.7
0.5
0.3
0.1
	

                        Source:  Rem et al.  (1982)
                        dashed lines indicate minor releases  possible
                       c
                        values are rounded
                        not included because of uncertainty

-------
mj
 I
                                                                       TABLE 3-9


                                                              Ambient  Levels  of Chloroform
Location
Alabama
Tuscaloosa
Talladega Forest
Arizona
Phoenix
California
Stanford Hills
Point Reyes
Los Angeles
Palm Springs
Yosemite
Mill Valley
Riverside
Badger Pass
Point Arena
Point Arena
Los Angeles
Oakland
Delaware
Delaware City
Kansas
Jetmore
Maryland
Baltimore
Montana
Western Montana
Type of Site

urban
rural

urban

clean
clean marine
urban
urban-suburban
remote-high altitude
clean marine
urban- suburban
remote-high altitude
clean marine
clean marine
urban
urban

NS

remote-continental

urban

remote
Date

2/77
2/77

4-5/79

11/75
12/75
1-5/76
5/76
5/76
1/77
4-5/77
5/77
5/77
8-9/78
4/79
6-7/79

7/74

6/78

7/74

3/76
Analytical
Method

GC-ECD
GC-ECD

GC-coulometry

GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD

GC-ECD

GC-ECD

GC-ECD

GC-MS
Concentration (ppt, v/v)
Max.

3000
200

514.0

217
114
724
616
24
36
310
38
42
48
223.5
60.1

<10

34

<10

NR
Min.

100
NR

27.1

12
15
23
20
12
4
24
2
8
12
24.3
13.1

<10

9

<10

NR
Average

800
100

111.4

33
37
102
99
17
25
25
16
20
18
88.2
32.1

NR

16

NR

9


Holzer
Holzer

Singh

Singh
Singh
Singh
Singh
Singh
Singh
Singh
Singh
Singh
Singh
Singh
Singh

Reference


et al
et al

et

et
et
et
et
et
et
et
et
et
et
et
et

Lillian

Singh


et

Lillian

Cronn



, 1977
, 1977

al., 1981

al.
al.
al.
al.
al.
al.
al.
al.
al.
al.
al.
al.


1979
1979
1979
1979
1979
1979
1979
1979
1979
1979
1981
1981

et al., 1975a

al.


, 1979

et al., 1975a


and Harsch, 1979
       1979
Nebraska
Reese River
New Jersey
Rutherford

Newark


remote-high altitude

urban

urban


5/77

1978

1978


GC-ECD

GC-MS

GC-MS


19

31,000

7500


6

NR

NR


13

4600

3900


Singh et al., 1979

Bozzelli
and Kebbekus, 1979
Bozzelli and
Kebbekus, 1979

-------
                                                                   TABLE 3-9 (cent.)
 I
uo
 o
Location Type of Site
Piscataway

Somerset (county)

Bridgewater
Township
Bound Brook
Patterson
Clifton
Fords
Newark
Passaic
Hoboken
Seagrlt
Seagrit
Sandy Hook
Sandy Hook
Bayonne
New York
Staten Island
New York City
New York City
White Face Mountain
White Face Mountain
Niagara Falls
Ohio
Wilmington Air
Wilmington Air
Texas
Houston
urban

urban

rural

urban
urban
urban
urban
urban
urban
urban
urban
urban
urban
urban
urban

urban
urban
urban
remote
remote
urban

Force Base
Force Base

urban
Date
1978

1978

1978

3/76
3/76
3/76
3/76
3/76
3/76
3/76
6/74
6/75
7/71
7/75
7/75

3/76
6/71
6/75
9/74
9/75
NS

7/71
7/75

6-7/77
Analytical
Method
GC-MS

GC-MS

GC-MS

GC-MS
GC-MS
GC-MS
GC-MS
GC-MS
GC-MS
GC-MS
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-ECD

GC-MS
GC-ECD
GC-ECD
GC-ECD
GC-ECD
GC-MS

GC-ECD
GC-ECD

GC-MS
Concentration (ppt,
Max.
2900

11,000

NR

NR
NR
NR
NR
NR
NR
NR
60
50
63
55
15,000

NR
480
150
250
350
21,611

4800
5000

11,034
Min.
NR

NR

NR

NR
NR
NR
NR
NR
NR
NR
<10
5
<10
10
<10

NR
<10
10
<10
6
215

<10
20

Trace
v/v)
Reference
Average
2200

5000

NDb

854
768
1700
3422
7582
854
427
40
35
30
25
1030

4268
160
200
9
8
NR

340
480

NR
Bozzelli and
Kebbekua, 1979
Bozzelli and
Kebbekus, 1979
Bozzelli and
Kebbekus, 1979
Pellizzari, 1977
Pellizzari, 1977
Pellizzari, 1977
Pellizzari, 1977
Pellizzari, 1977
Pellizzari, 1977
Pellizzari, 1977
Lillian et al., 1975a
Lillian et al., 1975b
Lillian et al., 1975a
Lillian et al., 1975b
Lillian et al., 1975a

Pellizzari, 1977
Lillian et al., 1975a
Lillian et al., 1975b
Lillian et al., 1975a
Lillian et al., 1975b
Pellizzari et al., 1979

Lillian et al., 1975a
Lillian et al., 1975a

Pellizzari et al., 1979

-------
                                                            TABLE 3-9 (oont.)
Location Type of Site
Washington
Pullman rural

Pullman rural
England
Liverpool/Manchester suburban

Organochlorine urban
manufacturer
Moel Faman, urban
Flintshire
Rannoch Moor, urban
Argyllshire
Rural areas rural
Ireland
Cork urban
Japan
Kobe NS
Atlantic Ocean
Northeast Atlantic
(Cape Blanc to
Lands End)
31°19'N 13°32'W to
19°51'N 05°51'W
Date

12/71 to 2/75

11/75

NS

NS

NS

NS

NS

1971

NS

7-8/72




Analytical
Method

GC-MS

GC-ECD

GC-ECD

GC-ECD

GC-ECD

GC

GC

GC-ECD

GC-ECD

GC




Concentration (jpj>tj v/v) Reference
Max. Min. Average

NR NR 20 Grlmsrud and
Rasmussen, 1975
NR NR 13 Rasmussen et al.,1977
C G
8 3 NR Pearson and McConnell,
1975
MO <10.1 NR Pearson and McConnell,
1975
0.1 <0.1 NR Pearson and McConnell,
1975
0.5 0.1 NR Murray and Riley, 1973

1.2 0.82 0.82 Murray and Riley, 1973

NR NR 26.5 Cox et al., 1976

9100 300 Okuno et al., 1971

0.96 0.11 0.35 Murray and Riley, 1973




 Subject Urban Transport





 Detection Limits 10 ppt





 ppb by mass




NR = Not reported; NS = Not specified




ND = Not detected

-------
   — 12
 10   ,  v/v),  and  .some <10  ppt.  There  are notable exceptions, however, although



 the  reasons  for this are not  readily apparent.



     Singh  (1977) and Singh et al.  (1979) have determined northern and southern


 hemisphere   background  concentrations  as  well  as  a  global   average.    The



 hemispheric  values  are 14 ppt  for the northern hemisphere  and  <3  ppt for the



 southern  hemisphere.   This difference in hemispheric  values suggests that the



 oceans  are not a  significant  source of chloroform,  but rather, that chloroform,



 for the most part, is anthropogenic.  The global average concentration  determined



 by Singh  et  al. (1979) is 8 ppt.



     An interesting point not presented in Table 3-9 is that chloroform concen-



 trations  above  an inversion  layer are significantly  lower  than concentrations



 below it.   In Wilmington,  OH, above an inversion layer, the chloroform concen-



 tration was <10 ppt, whereas  below it the concentration was  120 ppt (Lillian et


 al., 1975a).



 3.6. ATMOSPHERIC REACTIVITY



     The  principal  atmospheric   reactant   responsible  for  the  removal  of



 chloroform  is probably  the hydroxyl radical (Atkinson et al.,  1979; Graedel ,



 1978; Altshuller,  1980; Singh,  1977;   Crutzen  and Fishman,  1977).   Hydroxyl



 radicals  are  formed in  the lower  atmosphere  in two ways, first,  by the photo-



 dissociation of ozone (X <310) into 0 (1D) atoms  (Atkinson et al . , 1979).  These



 go on to react with either water, hydrogen or methane to form hydroxyl radicals.



The second important  source of hydroxyl radicals is  the reaction of hydroperoxyl



 radicals with nitric oxide.



     Hydroxyl  radical   reactions   probably  follow  the  course  outlined  below



 (Graedel,  1978):



           CHC13 + HO ---------- *  »CC13 + H20
           cci3
                                      3-32

-------
                    NO
          •OCC1  --------------- v COC1  (phosgene) + Cl«




           COC12 + H20 --------- >• C02 + 2 HC1
Pearson and McConnell  (1975) found HC1 and C02 as the only  products of chloroform




irradiation with UV (X >290nm)  light.   The half-life  reported by these workers




(23 weeks)  was of  the  same order  of  magnitude  as  that calculated  from  the




hydroxyl radical rate constant (11.5 weeks)  (Singh et  al., 1981).




     Chloroform will not  react  photolytically in the trospos phere; the UV cutoff




for chloroform is 175 nm (i.e., it will not absorb light  >175 nrn).  Callahan et




al.  (1979)  calculated that  roughly  1% of  the  tropospheric  chloroform  would




diffuse eventually into the stratosphere,  based on a  lifetime of 0.2-0.3 years




and a troposphere-to-stratosphere turnover time of 30  years.




3.7. ECOLOGICAL EFFECTS/ENVIRONMENTAL PERSISTENCE




3.7.1.  Ecological Effects.



     3.7.1.1.   TERRESTRIAL --  Data on  the  terrestrial  ecological  effects of




chloroform are  not available.   Significant effects  are not  expected because




chloroform is quite volatile and does  not accumulate in terrestrial (or aquatic)




environments, and is  diluted rapidly  and degraded  to  low concentrations in the




troposphere (NAS, 1978).   Conceivably,  acute  effects on wildlife can occur in the



vicinity of major chloroform spills, but significant chronic effects from long-




term exposure to low ambient levels is unlikely.



     3.7.1.2.  AQUATIC —  The  toxicity of chloroform to  aquatic organisms has




been  reviewed by  the U.S.  EPA  (1980h).    As  summarized in Table  3-10,  two




freshwater fish  (rainbow trout,  bluegill)  and  one  invertebrate (Daphnia magna)




species have been acutely tested under standard conditions;  LC^ concentrations




ranged from 28,900 to 115,000 [ig/Jl  (Bentley et al., 1975; U.S. EPA,  1978a), and
                                      3-33

-------
                                                    TABLE 3-10

                           Acute and Chronic Effects  of Chloroform on Aquatic  Organisms'
Species
Cladoceran
Daphnia magna
Rainbow trout
Salmo gairdneri
Rainbow trout
Salmo gairdneri
OJ
i Bluegill
-(=• Lepomis macrochirus
Bluegill
Lepomis macrochirus
Pink shrimp
Penaeus duorarum
Orangespotted sunfish
Lepomis humilis
Goldfish
Carassius auratus
Duration
48 -hr
96-hr
96 -hr
96 -hr
96 -hr
96 -hr
1-hr
30 to 60 rain
Concentration
(jj.g/2,) Method
28,900 S,Ub
66,800 S,U
43,800 S,U
115,000 S,U
100,000 S,U
81,500 S,U
106,890 to NS
152,700
97,000 to NS
167,000
Effect
LC50
LC50
LC50
LC50
LC50
LC50
death
5056
anesthetized
Reference
U.S. EPA, 1978a
Bentley et al . ,
Bentley et al . ,
Bentley et al. ,
Bentley et al . ,
Bentley et al . ,
Clayberg, 1917
Gherkin and
Catchpool, 1964

1975
1975
1975
1975
1975


Threespine stickleback    90-min
  Gasterosteus aculeatus
207,648
NS
 anesthesia    Jones,  1947a
with recovery

-------
                                                TABLE 3-10  (cont.)
Species
Ninespine stickleback
Pungitius pungitius
Duration
NS
Concentration
(|ig/O
148,320 to
296,640
Method
NS
Effect
Avoidance
Reference
Jones, 1947b
Rainbow trout
  (embryo-larval)
  Salmo gairdneri

Rainbow trout
  (embryo-larval)
  Salmo gairdneri
                         27 days
                         27 days
2,030
1,240
F,M
F,M
                                                                              LC
                                                                             g/Jl H
                         50 mg/Jl  ardness
                                                                              LC
                                                                             g/i, H
Birge et al., 1979
Birge et al., 1979
                        200 mg/i,  ardness
UJ
uo
Rainbow trout 23 days
(embryo)
Salmo gairdneri
10,600 F,Me 4056
teratogenesis
Birge et al., 1979
 Source:  U.S. EPA, 1980

 Static test, unmeasured concentration

 Saltwater species
H
 Corrected from vol/vol to p.g/£

eplow-through test, measured concentration

 Exposures began within 20 minutes of fertilization and ended 8 days after hatching.

hr = hour; rain = minutes; NS = Not stated

-------
the trout was more sensitive  than the bluegill.  With stickleback, goldfish, and

oranges potted  sunfish,  anesthetization  or  death  occurred  after exposure  to

97,000 to 207,000 \j.g/H chloroform for 30  to  90  minutes (Clayberg, 1917;  Jones,

1947a; Gherkin and Catchpool,  1964).   Only  one  test has been  conducted with
chloroform and  saltwater  organisms;  the 96~hour LC,-n  for the  pink  shrimp was

81,500 \ig/H  (Bentley et al., 1975).

     Embryo-larval tests with rainbow trout at 2 levels of hardness provided 27-

day LC(-n values  of  2030 and 1240 jig/S,  (Birge et  al.,  1979).   There  was  a 40$

incidence of teratogenesis in the embryos at hatching (23 day exposure at 10,600
     Bluegills bioconcentrated radiolabeled chloroform by a factor of 6 after a

14-day exposure, and  the  tissue  half-life was <1 day  (U.S. EPA,  1978a).   This

degree  of   bioconcentration   and  short  biological  half-life  suggest  that

chloroform residues would not  be  an environmental hazard to consumers of aquatic

life (U.S. EPA, 1980h).

3.7.2.  Environmental Persistence.  A number  of researchers  have reported the

dominance  of hydroxyl  radical oxidations  in the  fate of  chloroform  in the

atmosphere  (see  Section 3-6).   Singh et  al.  ( 198 1 ) calculated an atmospheric

residence time for chloroform  based on the NASA  reviewed rate constant reported

by  Hampson  (1980).   They  reported a 116-day  (16.6-week)  residence  time for a

hydroxyl radical concentration of  10  molecules/ cm .  This compares well to the

observed 33-week lifetime of chloroform in a sunlit flask (Pearson and McConnell,

1975).  This  lifetime was  based  on experiments conducted in northwest England,

which receives less intense sunlight  than  most of the U.S., and may account for

its longevity.

     According to recent hydroxyl radical measurements, tropospheric ambient air
                               £       r-i
concentrations  range  from  =10  to 10   molecules/mi,  (Atkinson  et  al.,  1979);
                                      3-36

-------
models of the troposphere have suggested a concentration ranging  between  2 to 6 x




105 molecules/mS, (Crutzen and Fishman, 1977; Singh, 1977).




     Table 3-11  summarizes  the literature  values  for kQH>  the temperature of




measurement, and the calculated lifetime  based on the indicated  hydroxyl radical


                                                                C              "3

concentration.   If  a hydroxyl  radical  concentration  of 2  x 10  molecules/cm




(typical for summer months;  winter concentrations are lower) (Singh et al., 1981)




is  assumed,  most of  the lifetimes  calculated  from  the  rate  constants  range




between 0.2 to 0.5 years (69 to 181 days, 122 average).




     Mabey and Mill  (1978) critically reviewed hydrolysis data available in the




literature.  They determined that chloroform had a hydrolysis half-life of >3,000


                                                                             _ij

years at pH 7 and 298  K.  This is based on a base hydrolysis rate of 0.602 x 10



                             -7
and a OH  concentration of  10   in neutral water.




     Dilling  et  al.  (1975)  and Billing  (1977) determined  the volatilization




half-life of  chloroform from water.  For a  1 ppm chloroform  solution stirred at




200 rpm,  the  time  for 50%  removal  was 21.5 minutes  (average);  90? removal was




accomplished  in  71  minutes.   The  addition  of  dry  granular  bentonite  clay,




dolomitic limestone, or peatmoss had little effect  on  the evaporation rate.  The




rapid volatilization of chloroform was seen also by Jensen  and Rosenberg (1975),




who reported that 0.1-1.0  ppn  solutions of  chloroform in  partly  open sunlit




aquaria lost  50-60$ of the  chloroform in 8 days as opposed  to only 5% in closed




aquaria.  Pearson and McConnell (1975) suggested that  the  presence of chloroform




in  ambient waters may be from aerial  transport  and washout.




     An EXAMS model  of the  fate of chloroform in a  pond, a  river, and an oligo-




trophic lake  and eutrophic lake revealed the dominant process in all  cases  to be


                                                                     -9    -1
volatilization.    Input   parameters   included  hydrolysis  (2.5  x  10    hr  ),




octanol/water partition  coefficient  (91),  vapor pressure (150.5 torr at 20°C),




solubility (8200 ppn), Henry's Law Constant (2.88  x 10~3),  reaeration rate ratio
                                      3-37

-------
OJ
 I
                                                          TABLE 3-11


                                                        Values for k
                                                                    OH
kOH
era







x 1014
molecule" sec"
6.51
10.1
16.8
11.4
6.4
7.4
10
[OH] Lifetime
K x 10~5 (years)
265 4 1.2
296
298 10 0.19
298
265 9 0.56
273
298
Reference
Singh et al . , 1979
Howard and
Evenson, 1976
Cox et al., 1976
Davis et al., 1976a

Hampson, 1980

       a(4.69 + 0.71) x  10~12 exp
- (2254 + 214/RT)
         Evaluated by NASA

-------
(0.583), alkoxy radical rate constant (0.7 M~1  hr~1)  (RO =  1014 M), and a stream




loading of  1  g/hr.   No photochemical or  bacterial  degradation parameters were




entered since chloroform has no UV absorbence >290 nm, and virtually no bacterial




degradation occurs with chloroform (Pearson  and McConnell,  1975; Bouever et al.,




1981).  Table 3-12 summarizes the EXAMS model  generated fate of chloroform.




3.8. EXISTING CRITERIA, STANDARDS, AND GUIDELINES




3.8.1.  Air.  The Occupational Safety and Health Administration (OSHA) currently




limits occupational exposure to chloroform to  a ceiling level of 50 ppm (40 CFR




1910.1000).  This ceiling level  is  not  to be  exceeded in  the workplace  at any




time.   To  protect  against mild central nervous  system  depression,  irritant




effects,  and fetal  abnormalities  (which  were  considered to  occur at  lower




exposure  levels  than  those causing liver  injury),  the National  Institute for




Occupational Safety  and  Health  (NIOSH)  recommended in  1974 that exposure  to




chloroform be limited to 10 ppm as a Time-Weighted Average  (TWA) exposure for up




to a 10-hour workday,  40 hour workweek.   A ceiling level of 50 ppm was proposed




for any 10-minute period  (NIOSH,  1974).  NIOSH lowered the recommended criterion




to 2 ppm TWA in 1976  (NIOSH, 1977) in response to a  positive NCI carcinogenesis




bioassay (NCI,  1976).  NIOSH recommended that exposure to halogenated anesthetic




agents, including chloroform,  be limited  to 2 ppn  because this is  the  lowest




detectable level using the recommended sampling and analysis techniques, and not




because a safe level  of airborne exposure could be defined.




     On the basis of  recent reports of  carcinogenicity  and embryotoxicity, the




American  Conference  of Governmental  Industrial Hygienists (ACGIH)  currently




classifies  chloroform as   an  Industrial  Substance  Suspect  of  Carcinogenic




Potential for Man (ACGIH,  1981).  The ACGIH recommends  a Threshold Limit Value




(TLV) of  10 ppm and a  15-minute  Short-Term  Exposure Limit  (STEL)  TWA of  50 ppm




for chloroform.
                                     3-39

-------
                                                                                        TABLE 3-12


                                                                    Summary of EXAMS Models of the Fate of Chloroform3
UJ
 I

Maximum total concentrations in water
column (mg/i)
Maximum concentration in sediments
(dissolved in pore water, mg/i,)
Maximum concentration in Bios
Plankton (|ig/g)
Benthos (ng/g)
Maximum total concentration in sediment
(mg/kg, dry weight)
Total steady accumulation (kg)
% in water column
$ in sediments
Disposition
chemical transformation (?)
biotransformation (J)
volatilization (J)
Volatilization half-life
exported (J)
export half-life
Mass Flux from Volatilization (kg/hr)
Self-Purification Time
River
9.92 x 10~7
9.85 x 10~7
2.58 x 10~^
2.56 x 10~s
3.19 x 10""6
9.15 x 10"^
96.93
3-07
0.00
0.00
1.74
36 hours
98.26
0.65 hours
1.7i» x 10~5
37 hours
Pond
2.50 x 10~3
1.36 x 10~3
6.49 x 10~2
3-53 x 10"^
6.50 x 10~3
5.43 x 10~2
91.9
8.1
0.00
0.00
93.35
40 hours
6.65
566 hours
9.33 x lO"1*
31 days
Oligotrophic
Lake
1.33 x lO"4
5.82 x TO"6
3.45 x 10~3
1.51 x 10~*
2.83 x 10~5
0.33
99.95
0.05
0.00
0.00
94.98
10 days
5.02
192 days
2.28 x 10~3
65 days
Eutrophic
Lake
1.26 x 10~4
4.57 x 10"6
3.27 x 10~3
1.19 x 10"4
9.35 x 10"6
0.3
99.94
0.06
0.00
0.00
95.57
9 days
4.1(3
196 days
2.29 x 10~2
56 days
                    TJased on load of 1.00 g/hr

-------
     Foreign industrial air standards  for  chloroform  include (Utidjian, 1976):


Bulgaria, 10 ppm;  Czechoslovakia,  10 ppn (50 ppn for brief exposures); Finland,


50 ppm; Hungary, 4 ppm  (20 ppra  for  brief exposures);  Japan, 50 ppm;  Poland, 10


ppn; Rumania,  10 ppm; Yugoslavia, 50 ppm;  West Germany, 10 ppm (Utidijian, 1976).


3.8.2.  Water.  As discussed in the Ambient Water Quality Criteria Document for


chloroform (U.S. EPA, 1980h), the EPA has proposed  an amendment that would add to


the National Interim Primary Drinking Water Regulations  a section on the control


of  organic  halogenated chemical  contaminants.    The  proposed limit  for total


trihalomethanes  in  drinking  water, which includes  chloroform  as   the major


constituent, is 100 ng/&.  Although some estimates  of cancer  risk were  performed,


this  limit  was  set   primarily  on  the  basis  of  technological  and  economic


feasibility,  and  initially  will apply  only to water supplies  serving  >75,000


consumers.  The basis  and purpose  of this  regulation are discussed in  a report


that was  prepared  by the Office of Drinking Water (U.S. EPA, 1978b).


     The  U.S.  EPA  (1980h) recently  derived cancer-based ambient water  criteria


for  chloroform.   Since zero level  concentrations of chloroform  will never be


attainable  in  chlorine-treated water,   levels  that  may result  in incremental

                                                                            _5
increases of  cancer risk over the lifetime were estimated at risks of  1  x 10   ,

   r         t-t
10   ,  and 10   .  The corresponding recommended criteria, which were derived with


the  tumor incidence data from  the  NCI bioassay with female mice (NCI,  1976), are


1.90,  0.19, and 0.019  p.g/&, respectively,  if exposure is assumed to be  from the


consumption of  drinking water and  fish and shellfish products and at  157 Hg/&,


15.7  \ig/&,  and 1.57 ng/Ji,, respectively, if exposure is assumed to be  from the


consumption of  aquatic organisms only.


3.8.3.   Food.   Chloroform has been approved by the Food and Drug Administration


(FDA)  as a component of articles intended for  use  in contact with  food (i.e., an
                                      3-41

-------
indirect  food  additive).    The  use  of  chloroform in  the  food industry  is

summarized as follows:


          Component of  adhesives                 21 CFR  175.105

          Adjuvant substance required           21 CFR  177.1580
            in the production of
            polycarbonate resins


     Chloroform also has  been  exempted from the  requirement  of  tolerance when

used as a  solvent in pesticide formulations that  are  applied to growing crops (40

CFR 180.1001), or when used as a  fumigant after harvest for barley, corn, oats,

popcorn, rice, rye, sorghum (milo),  or wheat (40 CFR 180.1009).

3.8.4.   Drugs and Cosmetics.   The  positive  NCI  carcinogenicity  bioassay  of

chloroform (NCI,  1976)  has prompted  the FDA to restrict the use of chloroform in

drug  (21 CFR 310-513) and cosmetic  (21 CFR 700.18) products.

3-9. RELATIVE SOURCE CONTRIBUTIONS

      The  sum  of  all  the environmental releases  of  chloroform from all sources

listed in Section  3-^-3 amounts to  a  total of  10,438 metric  tons.  All sources

are  summarized  in Table  3-8  with  the percent  of the  total  emissions.   Total

emissions from all sources  constitute about  5.6% of production (184,000 metric

tons).  Table 3-8 does not include estimated  emissions  from laboratory  use.  Rem

et al.  (1982) suggested that these  are potentially large but gave no numerical

estimate.
                                      3-42

-------
                                   REFERENCES








ACGIH  (American  Conference  of  Governmental  Industrial  Hygienists).    1930.




Documentation  of  the Threshold Limit  Values.   4th  ed.   Cincinnati, OH.   pp.




90-91 .








ACGIH  (American  Conference  of  Governmental  Industrial  Hygienists).    1981.




Threshold  Limit Values  for  Chemical  Substances  and  Physical  Agents   in  the




Workroom Environment with Intended  Changes  for  1931.   Cincinnati,  OH.  pp.  13,




42.









Ahlstrom, R.C., Jr.  and J.M.  Steele.   1979.  Methylchloride.  In:   Kirk-Othmer




Encyclopedia of Chemical Technology,  3rd  ed.,  M. Grayson and  D. Eckroth,  eds.




New York:  John Wiley and Sons, Inc.  vol. 5.  pp. 677-685.








Altshuller.  1980.  Lifetimes of organic molecules in the troposphere and lower




stratosphere.  In:  Advances in Environmental Science and Technology,  vol.  10.




Pitts and Metcalf, eds.   New York:   John Wiley and Sons,  Inc.  p. 181-219.








Anonymous.   1983.   Chemical  Profile.  Chloroform.   Chemical Marketing Reporter,




June 25,  1979,  p.  54.








Anthony,  T.    1979.   Methylene  chloride.   In:    Kirk-Othmer Encyclopedia  of




Chemical  Technology,  3rd ed.,  M.  Grayson and D. Eckroth, eds., New  York:  John




Wiley and Sons, Inc.  vol.  5.  p.  686-693.
                                     3-43

-------
Atkinson, R., K.R. Darmall, A.C. Lloyd, A.M. Winer  and  J.N.  Pitts, Jr.   1979.




Kinetics and mechanisms of  the  reactions  of the hydroxyl radical  with  organic




compounds in the gas phase.   In:   Ado.  Photochem.   11:  375-^8.









Beale, J.  (n.d.).  Letter from J. Beale, Dow Chemical U.S.A.,  Midland,  Michigan,




to L. Evans, Emissions Standards and Engineering Division, U.S.  EPA concerning




Dow facility at Freeport,  TX.   (Cited in  Rem et al.,  1932).









Bentley, R.E.,  et al. 1975.  Acute toxicity of chloroform to  bluegill (Lepomis




macrochirus),  rainbow  trout,   (Salmo  gairdneri),  and  pink  shrimp   (Penaeus




durorarum).  Contract No.  WA-6-99-1 U14-B.  U.S. Environmental Protection Agency




(Cited in U.S.  EPA, 1930h.)









Birge, W.J. et al.  1979-   Toxicity of organic  chemicals to embryo-larval stages




of fish.  EPA-560-79-007.  U.S. Environmental Protection Agency.  (Cited in U.S.




EPA, 1930h.)









Bouever,  E.J.,  B.  Rittman and P.L.  McCarty.   1981.  Anaerobic  degradation  of




halogenated  1-  and 2-carbon organic compounds.   Environ. Sci.  Technol.   15:




596-599.








Bozzelli, J.W. and B.B. Kebbekus.  1979.   Analysis  of selected volatile organic




substances  in  ambient air.  State  of New  Jersey  Department  of Environmental




Protection.

-------
 Brown, R.H. and C.J. Purnell.  1979.   Collection  and analysis of trace organic




 vapour  pollutants  in  ambient  atmospheres.   The  performance  of a  Tenax-GC




 adsorbent tube. J.  Chromatog.   173: 79-90.








 Callahan, M.A., M.W. Slimak,  N.W.  Gabel,  I.P.  May,  C.F. Fowler, J.R. Freed, P.




 Jennings, R.L. Durfee, F.C. Whitmore,  B.  Maestri, W.R.  Mabey,  B.R. Holt and C.




 Gould.  1979.  Water related fate of 129 priority pollutants, vol.  II.  U.S. EPA




 -  440/4-79-029B.









 Chemical Marketing Reporter.  1981.  Chemical Profile: Trichloroethylene.  April




 6, 1981 .









 Gherkin, A. and J.F. Catchpool.  1964.  Temperature dependence of anesthesia in




 goldfish.  Science.  144: 1460.  (Cited in U.S. EPA, 1980h.)









 Clayberg, H.D.  1917.  Effects of ether and  chloroform on certain fishes.  Biol.




 Bull. 32: 234.  (Cited in U.S. EPA, 1980h.)









 Cox, R.A.,  R.G. Derwent,  A.E.J. Eggleton and J.E. Lovelock.  1976.  Photochemical




 oxidation of halocarbons in the troposphere.  Atmosph.  Environ.  10:  305-308.









 Cronn, D.R. and  D.E.  Harsch.  1979.   Determination of atmospheric  halocarbon




concentrations by  gas  chromatography-mass spectrometry.   Anal.  Letters.    12:




1489-1496.
                                     3-45

-------
Crutzen, P.J.  and  J. Fishman.   1977.   Average  concentrations  of  OH  in  the




troposphere and budgets  of  CH^,  CO,  H2, and CH^Cd    Geophys. Res. Lett.   4:



321-324.









Davis, D.D., G. Machado,  B.  Conaway,  Y.  Oh and B.  Watson.   1976.   A temperature




dependent kinetics study  of  the  reaction of OH with CH..C1,  CHClo,  and  CH,,Br.




Jour. Chera. Phys.   65:  1268-1274.








DeShon, H.D.   1979-   Chloroform.   In:   Kirk-Othmer Encyclopedia  of  Chemical




Technology, 3rd ed.,  M.  Grayson and D. Eckroth, eds.   New York:   John Wiley and




Sons, Inc.  vol 5.   pp.  693-703.








Billing, W.L.  1977.   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:  405-409.








Dilling, W.L., N.B. Tefertiller  and G.J. Kallos.   1975.   Evaporation rates and




reactivities of methylene chloride, chloroform, 1,1,1-trichloroethane,  trichlo-




roethylene,  tetrachlorocthylene,  and  other  chlorinated   compounds  in  dilute




aqueous solutions.   Environ. Sci. Technol.   9: 833-838.








Environmental  Reporter.   1982.  State  Air Laws.   Bureau  of  National  Affairs.




Washington, DC.  (Cited in Rem et al., 1982).








Graedel, T.E.  1978.   Chemical  compounds in the atmosphere.  Academic Press, NY.




p.  331, 334.
                                      3-46

-------
Grimsrud, E.P. and R.A. Rasmussen.  1975.  Survey and analysis of halocarbons in




the atmosphere by gas  chromatography-mass spectometry.  Atmos. Environ.  9:1 01 4-




101 7.









Harapson,  R.F.    1930.   Chemical,  kinetic,  and  photochemical data  sheets  for




atmospheric reactions.  Washington,  D.C.   Department of  Transportation Report




FAA-EE-80-17.









Harsch,  D.E.,  R.A. Rasmussen,  and D.  Pierotti.   1977-   Identification of  a




potential source of chloroform in urban air.   Chemosphere.  11:  769-775.









Harsch, D.E., D.R. Cronn and W.R.  Slater.  1979.   Expanded list of halogenated




hydrocarbons measureable  in ambient air.  Jour.  Air Pollut. Control Assoc.   29:




975-976.









Harsch, C. and A.J. Leo.   1979.  Substituent  constants for correlation analysis




in chemistry and biology.   New York:   John Wiley and Sons.  p. 172-173-









Hobbs, F.D.  1978.  Trip Report for Vulcan Materials Company, Geismar, Louisiana.




Prepared  for  U.S.  Environmental  Protection  Agency.    Emissions Standards  and




Engineering Division,  Research Triangle Park,  NC.   (Cited in Rem et al., 1932).









Holzer, G., H. Shanfield,  A. Zlatkis,  W.  Bertsch, P. Juarez, M. Mayfield and  H.M.




Liebich.  1977.  Collection and analysis of trace organic emissions from natural




sources.  Jour. Chromatog.   142: 755-764.
                                     3-47

-------
Howard, C.J.  and  K.M.  Evenson.   1976.    Rate  constants for  the reactions of




hydrozyl with methane and fluorine, chlorine, and bromine substituted methanes




at 296°  Kelvin.  Jour.  Chem.  Phys.   64:  197-202.








Jensen, S. and R.  Rosenberg.   1975.  Degradibility of some chlorinated aliphatic




hydrocarbons in sea water and  sterilized  water,  water res.  9:  659-661.








Jones, J.R.E.   V}47a.   The oxygen consumption of Gasterosteus  aculeatus  L. to




toxic solutions.  Exp.  Biol.   23:  298.   (Cited  in  U.S. EPA,  1980h.)








Jones,  J.R.E.   1947b.    The  reactions  of Pygosteus  gungitius L.  to  toxic




solutions.  Jour.  Exp.  Biol.   24:  110.   (Cited  in  U.S. EPA,  1980h.)








Lant,  S.G.   1979.   Diamond Shamrock.   Personal communications to D. Goodwin,




Office of Air Quality Planning and Standards.  April 3,  1978.   (Cited in Wagner




et al., 1930).








Lillian, D., J.B.  Singh, A. Appleby,  L.  Lobban,  R.  Aruts, R. Gompert, R. Hague,




J. Tommey, J. Kazazis,  M. Antell,  D.  Hansen  and B.  Scott.   1975a.   Atmospheric




fates of halogenated hydrocarbons.  Environ. Sci.  Technol.   9:  1042-1048.








Lillian, D., H.B.  Singh, A. Appleby,  L.  Lobban,  R.  Arnts, R. Gompert, R. Hague,




J. Toomey,  J.  Kazazis,  M. Antell, D. Hansen and  B. Scott.   I975b.  Fates  and




levels of ambient halocarbons.  Amer. Chem. Soc.  Symp. Series.   17:  152-158.
                                      3-48

-------
Lowenbach  and  Schlesinger  Associates.    1979.    Chloroform:   A  Preliminary




Materials Balance.  Prepared for the Office of Toxic Substances, U.S.  Environ-




mental Protection Agency.  Washington, DC.  (Cited in Rem et al., 1982).







Mabey, W. and T. Mill.   1978.  Critical review of hydrolysis of organic compounds




in  water under  environmental  conditions.  Jour.  Phys. Chem.  Ref.  Data.   7'




383-41 5.








Merck  Index:   An Encyclopedia  of Chemicals and Drugs.   1976.  M. Windholz, ed.




9th ed.  Rahway, New Jersey.  Merck and Co., Inc.  p. 272.








Metcalf  and  Eddy,  Inc.   1972.   Waste Water  Engineering,  Collection Treatment




Disposal.  McGraw-Hill Book Co, NY.  (Cited in Wagner et al., 1980).








Murray,  A.J.  and J.P.  Riley.   1973.   Occurrence of  some chlorinated aliphatic




hydrocarbons in the environment.   Nature.  242: 37-38.








WAS (National Academy of Sciences).  1978.  Chloroform, carbon tetrachloride, and




other  halomethanes: An environmental assessment.   National  Research  Council.



Washington,  D.C.  29U  pp.








NCASI  (National  Council for Air and  Stream Improvement).   1977.  Analysis  of




Volatile Halogenated Organic Compounds in Bleached Pulp Mill Effluent.   NCASI




Technical Bulletin No.  298.   New  York,  NY.  (Cited in Rem et al., 1982).
                                     3-49

-------
NCI (National Cancer Institute).  1976.   Report on Carcinogenesis Bioassay  of




Chloroform.  Available  from National Technical Information Service, Springfield,




Virginia. (NTIS PB-26M-018).









NIOSH (National Institute for Occupational Safety and  Health).   1974.   Criteria




for a  recommended standard  -  occupation exposure  to chloroform.   Publ.  No.




(NIOSH)  75: 114.  Dept. Health, Education,  and Welfare.   Washington,  B.C.









NIOSH (National Institute for Occupational Safety and  Health).   1977.   Criteria




for a recommended stnadard -  occupational  exposure to waste anesthetic gases and




vapors.  Publ. No. (NIOSH) 77-140. Department of Health, Education, and Welfare,




Washington, D.C.









Okuno, T., M. Tsuji and K. Shintani.   1974.  On the chlorination of hydrocarbons




in air.  Jour. Japan. Soc. Air Pollution.  9:  211.









Pearson, C.R. and G. McConnell.  1975.  Chlorinate C  and C? hydrocarbons in the




marine environment.  Proc. R. Soc. Land B.  189: 305-332.








Pellizzari,  E.D.   1977.   The measurement of  carcinogenic vapors in ambient




atmosphere.  U.S. Environmental Protection Agency.  EPA-600/7-77-055.









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




for making a preliminary assessment of halogenated organic compounds in man and




environmental media.  U.S. Environmental Protection Agency.  EPA-560/1 3-79-01 0.




Available from NTIS PB80-109168.
                                      3-50

-------
 Rasmussen,  R.A., D.E.  Harsch,  P.H.  Sweany,  J.P. Krasnec and D.R. Crown.   1977.




 Determination of atmospheric halocarbons by a temperature-programmed gas chroma-




 tographic freezeout concentration method.  Jour. Air Pollut. Contr.  Distr.  27:



 579-581 .









 Rem,  R.M.,  M.E. Anderson,  S.A. Duletsky,  D.C. Misenheimer and  H.F.  Rollius.




 (1982).   Chloroform Materials  Balance,  Draft Report.   EPA contract  68-02-3168,




 Task  69.









 Sax,  N.I.  (ed.).   1981.  Dangerous  Properties of Industrial Materials Report.




 Van Nostrand Reinhold  Co.  New York, NY.  March/April,  p.46.  (Cited in Rem et




 al.,  1982).









 Singh,  H.B.   1977.    Atmospheric  halocarbons:  Evidence  in favor  of reduced




 average hydroxyl radical concentration in the  troposphere.  Geophys. Res. Lett.




 >4: 101-104.









 Singh, H.B., L.J.  Salas, H. Shigeishi, A.J. Smith, E. Scribner and L.H. Cavanagh.




 1979.   Atmospheric distributions,  sources and sinks  of  selected  halocarbons,




 hydrocarbons, SFg, and  NjO.  U.S.  Environmental Protection  Agency.   EPA-600/3-




 79-107.









 Singh, H.B., L.J.  Salas, A. Smith and H. Shigeishi.  1980.  Atmospheric measure-




ments of selected  toxic organic chemicals-interim report- 1979.   U.S.  Environ-




mental Protection  Agency, EPA-600/3-80-072.   Avail.  NTIS PB80-198989.
                                     3-51

-------
Singh, H.B., L.J. Salas, A.J. Smith and  H.  Shigeishi.   1981.  Measurements  of




some  potentially  hazardous  organic  chemicals in  urban  environments.   Atmos.




Environ.  1 5: 601-612.








SRI International.  1933-  1933 Directory of chemical producers:   United  States




of America.   Menlo Park, California.   SRI International,   p.  502,  503, 58 H.








Stevens, A. A.,  C.J. Slocum, D.R.  Seeger and  G.G. Robeck.  1976.  Chlorination of




organics drinking water.  Jour.  Amer.  Water Works  Assoc.   68: 615-620.








Symons, J.M., T.A. Bellar, J.K.  Carswell, J.  DeMarco,  K.L. Kropp,  G.G.  Robeck,




D.R. Seeger, C.J.  Slocum, B.L. Smith and A. A. Stevens.  1975.   National Organics




Reconnaissance Survey for  Halogenated  Organics.   J.  Am. Water Works  Assoc.   p.
TAPPI  (Technical  Association of  the Pulp  and Paper  Industry).    1963.    The




Bleaching of Pulp.  TAPPI Monograph No. 27.  New York, NY.  (Cited in Rem et al.,




1982).








U.S. EPA.  n.d.  National Organic Monitoring Survey.  Technical Support Division,




Office of Water Supply.  Washington,  DC.  (Cited in Rem et al.,  1982).









U.S. EPA.   1976.  Atmospheric Freons  and  Halogenated  Compounds.   EPA-600/3-78-




108.   Environmental  Sciences  Research Laboratory,  Research Triangle  Park,  NC.




(Cited in Rem et al., 1932).
                                     3-52

-------
U.S. EPA.   1978a.   In-depth studies  on health  and  environmental  impacts  of




selected water pollutants.  U.S. Environmental Protection Agency.  Contract No.




68-01-4646.  (Cited in U.S. EPA, 1980h.)








U.S. EPA.   1978b.   Statement  of basis  and  purpose for  an amendment  to the




national-interim primary drinking water regulations on trihalomethanes.  Office




of Water Supply, Washinton, D.C.  (Cited in U.S. EPA,  198Oh.)








U.S. EPA.  1979a.  Fate of Priority Pollutants in Publicly Owned Treatment Works.




EPA-400/1-70-301 .   Office of Water Regulations and Standards,  Washington, DC.




(Cited in Rem et al., 1982).








U.S. EPA.  I979b.  Compilation of  Air Pollution Emissions Factors, Third Edition




-Supplement 12.   AP-42.   Transportation and  Marketing of  Petroleum  Liquids.




Office of Air Quality Planning  and Standards,  Research  Triangle Park,  NC.   pp.




4.4-1 to 4.4-13.  (Cited in Rem et al., 1982).








U.S. EPA.  1980a.  Organic Chemical Manufacturing Volume 8:  Selected Processes.




Report 5:  Chloromethanes  by Methane  Chlorination Process.   EPA-450/3-80-028c.




Office of  Air  Quality  Planning  and  Standards.   Research Triangle Park,  NCP,




(Cited in Rem et al., 1982).








U.S. EPA.  I930b.  Organic Manufacturing Volume 8:  Selected  Processes.  Report 6:




Chloromethanes by Methanol  Hydrochlorination  and Methyl Chloride  Chlorination




Process.    EPA-450/3-80-028c.    Office  of Air  Quality Planning  and  Standards.




Research Triangle Park,  NC.  (Cited  in Rem et  al.,  1982).
                                     3-53

-------
U.S. EPA.  19800.  Synthetic Organic Chemical Manufacturing Industry - Background




Information for Proposed Standards, VOC  Fugitive  Emissions.   EPA-450/3-80-33a.




Office of  Air Quality  Planning  and  Standards.   Research  Triangle Park,  NC.



(Cited in Rem et al., 1982).









U.S. EPA.   1980d.   Organic Chemical Manufacturing Volume 8:  Selected Processes.




Report 1:    Ethylene  Bichloride.   EPA-450/3-80-028c.    Office  of  Air  Quality




Planning and Standards.   Research Triangle Park, NC.  pp. III-1 to  III-9.  (Cited




in Rem et al., 1982).








U.S. EPA.   I930e.   Development Document for Effluent Limitations Guidelines and




Standards for  the Steam Electric  Point Source  Category.   EPA-440/1-80-029b.




Office of Water Regulations and  Standards.  Washington,  DC.   (Cited in Rem et




al., 1982).








U.S. EPA.   198Of.   Development Document for Effluent Limitations Guidelines and




Standards for the  Pulp, Paper and  Paperboard and Builders'  Paper and Board Mills.




EPA-440/1-80-025b.  Office of Water Regulations and Standards.  Washington, DC.



(Cited in Rem et al., 1982).








U.S. EPA.   1980g.   Organic  Chemical Manufacturing  Volume 8:  Selected Processes.




Report 3:   Fluorocarbon  (Abbreviated Report).   EPA-U50/3-80-028c.  Office of Air




Quality Planning and Standards, Research Triangle Park, NC.   p.  III-1 to III-6.




(Cited in Rem et al., 1982).
                                      3-54

-------
U.S. EPA.   I980h.   Ambient Water Quality  Criteria  for Chloroform.   Available




from:   National Technical Information  Service,  Springfield,  Virginia.   (NTIS




PBS 1-11 7442).








U.S. EPA.    1931 a.   The  determination of  halogenated chemical indicators  of




industrial contamination in water by the purge and trap method.  Method  502.1.




U.S. Environmental Protection Agency,  EPA-600/4-81-059 •








U.S. EPA.  1981b.   Air Pollution Emission Factors,  Third  Edition Supplement 12.




AP-42.    Storage  of  Organic Liquids.    Office  of  Air  Quality Planning  and




Standards.   Research Triangle Park,  NC.  (Cited in Rem  et  al.,  1982).








U.S. EPA.   1931C.   The  1980  Needs Survey.   EPA-430/9-81-008.   Office  of  Water,




Washington, DC.   (Cited in Rem et al.,  1982).








U.S. EPA.    1982a.   Projections  of Total  Lead  Usage Under  Alternative  Lead




Phasedown Programs.  Washington, DC.  (Cited in Rem et  al.,  1982).








U.S. EPA.  I982b.   Lead Phasedown Regulations.  47 FR 49322.  November  1 .  (Cited




in Rem et al.,  1982).








U.S. EPA.   I982c.   Chloroform  Position  Document  2.   Office of  Pesticides and




Toxic Substances.   Washington,  DC.   pp. 1-3.  (Cited  in Rem et  al.,  1982).








USITC.   1932. Synthetic organic chemicals.  United States production and  sales,




1981.  U.S. International  Trade Commission,  Washington, D.C.  USITC  Publication




1292.  p.  245.
                                     3-55

-------
Utidjian,  H.M.D.   1976.   Jour.   Occup.   Med.   18:  253.   (Cited in ACGIH, 1980.)








Wagner, K.,  H.  Bryson,  G.  Hunt,  A.  Shcohet.   1980.   Draft Report  Level I




Materials balance  chloroform.  U.S. Environmental Protection Agency  Contract  No.




68-01-5793.
                                      3-56

-------
                 4.   DISPOSITION AND RELEVANT PHARMACOKINETICS








4.1.  INTRODUCTION







     Considering that chloroform was the major anesthetic agent in use during the




hundred years from its introduction by Simpson in 1847 (Waters,  1951; Snow, 1858;




Simpson,  1817)  until after  the Second World  War,  there is  relatively little




detailed  information  about  its pharmacokinetics and metabolism in man.   This




undoubtedly is due to the fact that until recently,  specific and sensitive ana-




lytical methods were  unavailable for the measurement  of CHCl, and its metabolites




at the concentrations in which they were likely to be present  in vivo.  Although




chloroform as an anesthetic agent has  been replaced by  drugs  with  less cardiac




and hepatic  toxicity,  it is  still  widely  used in large bulk as an industrial




solvent,  as a  chemical  intermediary,  and as a grain  fumigant.  Chloroform is




present in  the water supplies  of  many United States cities  in concentrations




reaching  311 |ig/&, and  also  has been indentified as a  contaminant of  the  air




(U.S.   Occupational   Safety  and  Health Administration  (OSHA), 1978;  National




Institute for Occupational Safety and Health  (NIOSH), 1977b; Dowty et al., 1975;




Symons  et al.,  1975).   Ordinary exposure  to chloroform,  therefore,  includes




occupational, food,  drinking water,  and ambient air (NIOSH, 1977b;  Dowty et al.,




1975;  McConnell et al.,  1975), hence,  exposure can  be chronic by  both oral and




pulmonary routes, but  at levels far  below anesthetic  concentrations  (5000 to




10,000 ppm;  24.85 to 49.70 g/nr) .  Nonetheless, chloroform has been detected in




the breath  of  healthy  people  living  in  non-industrial environments  (Conkle




et al., 1975) and in post-mortem human tissue samples  (McConnell et al., 1975).
                                      4-1

-------
4.2. ABSORPTION





     Chloroform is rapidly and extensively absorbed through the  lungs  and  from


the gastrointestinal  tract.    Inhalation is  considered the  primary route  of


entrance into man for  occupational exposure and air pollution.  Absorption after


oral  ingestion  is of  particular interest  for chloroform  as a  contaminating


component  of  drinking water  and foodstuffs.   Significant absorption  through


intact skin occurs only with liquid  contact  or submersion.





4.2.1.  Dermal Absorption.  Absorption of chloroform through the skin from direct


liquid contact  (immersion of hands or arms)  is a slow  process.   Early studies


(Torkelson et al., 1976;  Schwenkenbecher,  1904; Witte, 187*0 showed that chloro-


form does  penetrate the  skin and  can  be  absorbed into the  body  by  this route.


Tsurata  (1975,  1977)  has  studied the percutaneous  absorption of  a series of


chlorinated organic solvents applied to a standard area of shaved abdominal mouse


skin for  15 minute periods.  Absorption was quantitated by presence of compound


in  total  mouse  body plus expired  air,  as determined by GC.   For all solvents,


percutaneous  absorption  linearly increased  with time over  the  short exposure


period and was directly related to water  solubility.  For chloroform the absorp-

                              '->
tion rate was 329 moles/min/cm  skin,  third highest of 8 solvents  measured.   This


investigator  extrapolated  this absorption rate to a  calculation of the amount


absorbed  into the human body  as the result  of  1 min  immersion of both hands

       p
(800  cm   area).   The  estimated amount absorbed,  19.7 mg/min, was equated to an


inhalation exposure concentration of 2429 ppm for 1 min. Tsurata concluded  that


skin absorption  from  liquid contact  could be a significant route of entry  into


the body  for chloroform.   More  recently Jakobson  et  al.  (1983)  carried out


similar experiments with guinea pigs for  10  chlorinated organic  solvents  (which
                                       4-2

-------
however did not  include  chloroform).   Liquid contact  (skin  area,  3.1 cm ) was




maintained for up  to  12  hours and solvent concentration in blood was monitored




during, and for  some  solvents after exposure.   For these solvents,  the blood



elimination curves following dermal exposure were  non-linear, corresponding to a




kinetic  model   involving   at  least  two   body   compartments.     Furthermore




percutaneous absorption  of  these  solvents,  as reflected by blood concentration




profiles, showed three different  patterns  that were related to water solubility.




For solvents which were relatively hydrophilic [300 to 900 rag/100 ml water] the




blood  concentration  increased  steadily  during  the   entire  dermal  exposure,




indicating  that  absorption  occurs  faster  than  elimination  by metabolism  or




pulmonary excretion.  Chloroform with a water solubility of about 750 mg/100 ml




water might be expected to be in this group.









4.2.2. Oral.   The  kinetics  of gastrointestinal absorption of  chloroform after




oral ingestion have not been specifically studied; however, transmucosal diffu-




sive passage occurs readily,  as expected from  its neutral and lipophilic proper-




ties (Tables 4-1  and 4-2), and as demonstrated by its biological effects produced




by peroral administration of a wide  range of dosages  and  dosing schedules  for




toxicity studies  in rats, mice,  guinea pigs, and  dogs (Fishbein,  1979;  Hill




et al., 1975;  Brown et al.,  1974;  Kimura et  al.,  1971;  Klaassen and Plaa, 19&7;




Miklashevskii  et  al.,  1966;  Plaa  et al.,  1958; Eschenbrenner  and Miller, 1945),




teratolog)^ studies in  rats and rabbits  (Thompson  et al.,  "974),  and metabolism




studies in mice,  rats,  rabbits,  monkeys, and man (Brown  et al.,  1974; Taylor




et al., 1974; Fry et al.,  1972; Rubinstein  and Kanics, 1964; Paul and Rubinstein,




1963).  Accidental  and intentional ingestion of chloroform with rapid appearance




of clinical symptoms also has  been reported  in man (Storms,  1973;  Schroeder,




1965;  Piersol  et  al.,  1933).
                                      4-3

-------
                                     TABLE 14-1

            Physical Properties  of Chloroform and Other Chloromethanes*
Ostwald Solubility, 37°C

Dichloromethane
Chloroform
Carbon tetrachloride
Vapor Pressure
at 25°C, torr
400
250
100
Water/
air
7.6
4.0
0.25
Blood/
air
9.7
10.3
2.4
Olive Oil/
air
152
401
361
*Source:   Sato and Nakajima, 1979
Conversion factors:

20°C; 750 mmHg      1 ppm in air =4.97 |ig/& =4.97
37°C; 760 mmHg      1 ppm in air = 4.69 Hg/fc = 4.69 rag/iir
                                        4-4

-------
                                TABLE 4-2




             Partition Coefficients for Human Tissue at 37°C
Tissue
Blood
Brain
Grey matter
White matter
Heart
Kidney
Liver
Lung
Mus cl e
Fat tissue
Coefficient
8.0

16
24
8
11
17
7 -
12
280
Relative
to
blood


2.0
3.0
1.0
1.4
2.1
0.9
1.5
35.0
Source:   Steward et al.,  1973
                                     4-5

-------
     Brown et al. (1974) and Taylor et al.  (1974)  found  that  l4C-chloroform  in



olive oil given perorally to mice,  rats,  and monkeys (60  mg/kg)  was  essentially



completely absorbed by virtue of a 93 to 9856 recovery of radioactivity in exhaled



air, urine,  and  carcass (Table 4-9).   Absorption was rapid,  with peak  blood



levels at 1 hour in mice and monkeys.   In man,  Fry et  al.  (1972)  observed that


1 3
 JC-chloroform (0.5 g) in olive oil swallowed in a  gelatin  capsule  resulted  in



rapid appearance of the stable isotope in exhaled breath  (Table  4-8),  with peak



blood levels at 1 hour.



     Withey et al.  (1982) have investigated the  effect of dosing vehicle on the



intestinal  absorption  of  chloroform  in  fasting  rats  (400   gm)  following



intragastric intubation of equivalent  doses (75 mg/kg) in about 4 ml of water or



corn oil.  The postabsorptive peak blood concentration averaged 6.5 times higher



for water than  corn  oil (39 vs 6 |ig/ml), while the time to initial peak blood



concentration  was  essentially  the same   (5.6  vs.  6.0  min).    Although  the



absorption  from water  vehicle exhibited  one  blood concentration peak,  the



absorption from corn  oil showed two peaks  in blood concentration at 6  and  40



minutes.  The ratio of the areas under the blood  concentration curves for 5 hours



after dosing (AUC,  5 hr) was 8.7;  water,  corn oil.  These results suggest that



the absorption  of  chloroform with  both vehicles  is  rapid; but,  the rate  and



extent of absorption  may be  diminished, and the pattern of absorption altered  by



intragastric  intubation  of  high  volumes  (for a rat)  of  corn  oil  vehicle.   A



slower partitioning of lipophilic  compounds  dissolved in corn oil with  mucosal



lipids  can  be expected  in  comparison with  a water vehicle.    Furthermore,  in



contrast to aqueous absorption into the portal system and thence  to the liver,



corn oil  and other liquids are extensively transported  via mucosal  lymphatic



system which slowly drains by way  of the  left lymphatic  thoracic  duct  into the



systemic circulation via the superior  vena cava.  While these considerations are
                                      4-6

-------
unlikely to affect the pharmaeokinetics  of chloroform in man  in  any practical




way, they are of importance in relation to  the modes of dosing employed in long-




term carcinogenicity tests of chloroform and other lipophilic compounds.








4.2.3. Pulmonary  Absorption.   Chloroform has a relatively  high vapor pressure




(250 torr at 25°C; Table 4-1) and a  high blood/air  partition coefficient (8 to




10.3 at 37°C; Table 4-2), and hence, its vapor in ambient air is a primary mode of




exposure and  the  lungs a  principal  route  of entry into  the body.   The total




amount absorbed via the  lungs  (as  for all  vapors) is  directly proportional to:




(1) the concentration  of  the inspired air,  (2) the duration in time of exposure,




(3)  the  blood/air Ostwald  solubility coefficient, (4)  the solubility  in the




various body tissues,  and (5) physical activity, which increases pulmonary ven-




tilation rate  and cardiac  output.    Hence,  the  basic  kinetic parameters of the




pulmonary absorption of chloroform and its equilibration in the  body are as valid




for low concentrations expected in the ambient environment as for the high vapor




concentrations  associated  with its  use  as an  anesthetic (5000 to 10,000 ppm;




24.85  to  49.70 g/m3)  (Smith et al.,  1973;  Morris,  1951;  Waters,  1951).  These




parameters have not been  as well studied as  they have for modern anesthetics like




halothane (Fiserova-Bergerova and Holaday,  1979) or even for other common halo-




genated  hydrocarbon  solvents  like   triohloroethylene,  methylene  chloride,  or




methylchloroform.




     The  earliest attempt  at  controlled  studies  of pulmonary  absorption  of




chloroform in man were conducted by Lehmann and Hasegawa (1910).  These investi-




gators calculated retention values for chloroform (% inspired air concentration




of chloroform retained in the body) from differences  between inspired and expired




air concentrations  (analyzed  by alkali hydrolysis with chloride titration). As




expected, initial retention values were high, and decreased with exposure dura-
                                      4-7

-------
tion as total  body  equilibrium  with inspired air  concentration  was  approached



(Table 4-3).  The rate of uptake to equilibration and the final retention value



achieved is related  to the solubility of chloroform  in blood (blood/air partition



coefficient).  Figure 4-1 illustrates  for  chloroform  and other vapors that the




greater the  Ostwald  solubility  coefficient for  a  vapor agent,  the  less rapid



equilibrium occurs.   From the data of Lehmann and Hasegawa (Table  4-3) and recent



data of Smith et al.  (1973) of  blood levels during anesthesia shown in Figure



4-2, total  body  equilibrium with inspired  chloroform  concentration requires at




least >2 hours  in normal man at resting  ventilation  rate  and cardiac output.  The



retention value at equilibrium suggested  by the Lehmann  and Hasegawa  (1910) data




is  ^65%, and is  67% as calculated  from  the data of Smith et  al.  (1973).   The



difference,  33 to 3&%, represents body  elimination  of chloroform  by routes other



than  pulmonary  (primarily  by  metabolism).   The  percent  retention value  is



independent  of the inspired air concentration at equilibrium.



     The magnitude  of chloroform  pulmonary uptake into the body  (dose,  body



burden) is directly  related to  the  concentration of  chloroform in the inspired



air and to the duration  of exposure.   The  total  amount retained in  the body



during inhalation exposure can be estimated by multiplying percent retention (R)



by the volume of air inspired during the exposure period, or:



                       Amount uptake = (CT - C.) • V •  T
                                         JL    A


where V is  ventilation rate (Jl/minute),  T is exposure period  (minute), and C,. and



C   are  inspired   air   concentration   and  end  alveolar  air  concentration,



respectively.  Physical activity increases uptake by increasing the  ventilation



rate,  V, and the cardiac  output which influences rate of  distribution to the



various tissues of the body.
                                      4-8

-------
                                     TABLE 4-3

                   Retention and Excretion of Chloroform in Man
                      During and After Inhalation Exposure  to
                            Anesthetic Concentrations*
      Subject
Inspired air cone, (ppra)
Exposure period (min)

0 to 5
5 to 10
10 to 15
15 to 20
20 to 25
25 to 30
Postexposure (min)
4448
   4920

Retention (%)
4407
   74.5
   72.4
   68.6
   67.6
    NR
    NR
     68.4
     61.6
     51 .2
     50.2
      NR
      NR
  80.0
  74.2
  76.9
  74.6
  74.2
  73.8
                                             Excretion, mg/g,  expired  air
0 to 10
10 to 20
20 to 30
NR
NR
NR
NR
NR
NR
1.70
0.97
0.85
*Source:   Lehmann and Hasegawa,  1910

NR = Not  reported;  min =  minutes
                                        4-9

-------
    100
                       10       15       20

                            TIME, min
Figure 4-1.  Rate of rise of alveolar (arterial) concen-
tration toward inspired concentration for five anesthetic
agents of differing Ostwald solubilities (blood/air par-
tition coefficients): nitrous oxide,  0.47;  forane, 1.4;
halothane,  2.4;  chloroform, 8;  and  methoxyflurane,  11.
Note rate of alveolar chloroform rise  is  less than that
of  halothane  with  a  smaller  Ostwald  coefficient  and
greater than that of methoxyflurane with  a  larger coef-
ficient .

Source:  Munson (1973)
                           4-10

-------
 c
 V
 o
Z
Z
LLI
U
o
o
5
cc
o
LL
o
cc
o
I
CJ
    12
10
     CHLOROFORM
       O VENOUS BLOOD
       D ARTERIAL BLOOD
        BASE
        EXCESS
        pH
                +1.5
                40.0
                 7.44
                  •2.0
                  36.0
                   7.39
-2.0
33.0
 7.40
-3.0    -4.0
33.0   30.0
 7.39    7.39
-5.0    -1.5
27.0   36.0
 7.40    7.41
         0        1.2        3       4     |   5       6

                     POST INDUCTION TIME, hours


Figure  4-2.    Arteriovenous  blood  concentrations  of a
patient  during'anesthesia with  chloroform.   Note anes-
thetic blood  concentration for chloroform,  the decreasing
difference  between arterial  and  venous concentrations at
2 to 3 hours, indicating whole-body equilibrium,  and the
rapid  fall of  blood  concentration  with  termination  of
chloroform  exposure.

Source:  Smith  et al.  (1973).
                           '4-11

-------
     During  inhalation  of  chloroform  (and  in the  post exposure  elimination




phase), the arterial blood concentration of chloroform is directly proportional




to inspired air concentration (and end alveolar air concentration).  This fixed



relationship is defined by the blood/air partition coefficient in comparison to




other solvents (Sato and Nakajima, 1979) (Table 4-1), and hence, for equivalent




ambient air  exposure  concentrations,  the blood concentration of  chloroform Is




proportionally higher.   For  inspired air concentration required  for surgical




anesthesia  (8,000  to  10,000  ppm;  39/76  to  49.70 g/m3),  Smith  et  al.  (1973)




observed  a  mean arterial  blood  chloroform  concentration  for  10 patients of




9.8 mg/d& with a range of 7 to 16.5 mg/dil (Figure 4-2), and Morris (1951) found




similar values  for his  patients.   For  inspired  air  concentrations  less  than




anesthetic levels,  for example low vapor concentrations of 10 to  100 ppm (49.7 to




497 mg/nr), blood chloroform concentrations are lower in direct proportion.




     The amount of pulmonary absorption of chloroform  is also influenced by total




body weight and by  the total fat content  of the body (average body  fat  content is




8$ of  body  weight)  (Geigy Scientific Tables,  1973).   The capacity  of adipose




tissue  to absorb chloroform  rn vivo is  determined  by the product  of adipose




tissue  weight  and  lipid  solubility  of  chloroform.    The  lipid  solubility of



chloroform is relatively high  for this haloalkane (olive oil/air, 401;  Tables 4-1




and 4-2), and also, the adipose tissue/blood partition coefficient is high  (280




at 37°C); therefore, the uptake and storage of chloroform in adipose tissue can




be substantial, and it is increased with excess body weight and obesity.








4.3. TISSUE  DISTRIBUTION








     Chloroform, after pulmonary or peroral absorption, is distributed into all




body  tissues.   The  compound  crosses  the placental  barrier,   as  indicated by
                                      4-12

-------
embryotoxicity and  teratogenicity in  mice,  rats, and  rabbits after  oral  and




inhalation  dosing  (Murray et al.,  1979;  Dilley et al.,  1977;  Schwetz et al.,




1974; Thompson et al.,  1974).  It has been found in fetal liver (von Oettingen,




1964).  Chloroform can be expected to also appear in human colostrum and mature




breast milk, since it  has been found in fresh cow's milk and in high content in




cheese and butter (Table 4-4).




     As to  be  expected  from its lipophilic nature and  modest water solubility




(Table  4-1),  highest  concentrations  are found  in tissues  with  higher  lipid




content; relative tissue concentrations are reflected by individual tissue/blood




partition coefficients.   Coefficients for human  tissues,  given in  Table 4-2,




indicate that relative tissue concentrations are expected in the order of adipose




tissue  >  brain  > liver  > kidney  >  blood.   The absolute amounts  of chloroform




found in these tissues at  any given time are proportional to the body dose (i.e.,




to the  concentration in the  inspired  air  and duration of  inhalation or  to  the




oral dose, partition coefficient, and to the tissue compartment size).




     Gettler  (1934) and Gettler  and  Blume  (1931), using a  modified  Fujiwara




analytical method, determined the  chloroform content  of the  brain,  lungs,  and




liver of nine patients who died  during surgical  anesthesia (presumably 5000 to




10,000 ppm;  24.85 to 49.50 g/m3  inspired air)  as:  brain,  120 to 182; lung, 92 to




145; and liver, 65 to 88  mg/kg tissue wet weight.   Even higher  values (372 to 480



mg/kg in  brain tissue)  were found in  seven cases of  death due  to excessive




administration of chloroform (Gettler,  1934).   The blood  concentration  during




surgical anesthesia has  been recently  determined  (by  GC) to range  from  70 to




165 mg/JJ, in  10  patients  (average, 98)  by  Smith et al.  (1973).    These  tissue




concentrations  are  in  general  agreement  with  the    tissue/blood  partition




coefficients summarized  from  the literature by Steward et  al. (1973) and given in



Table 4-2.
                                     4-13

-------
                                    TABLE 4-4

                 Chloroform Content in United Kingdom Foodstuffs
                           and in Human Autopsy Tissue*
Chloroform in U.K. Foodstuffs
Foodstuff
Dairy produce
Fresh milk
Cheshire cheese
English butter
Hens eggs
Meat
English beef (steak)
English beef (fat)
Pig's liver
Oils and Fats
Margarine
Olive oil (Spanish)
Cod liver oil
Vegetable cooking oil
Beverages
Canned fruit drink
Light ale
Canned orange juice
Instant coffee
Tea (packet)

Fruit and vegetables
Potatoes (S. Wales)
Potates (N.W. England)
Apples
Pears
Tomatoes
Fresh bread
Chloroform
Hg/kg

5
33
22
1.4

4
3
1

3
10
6
2

2
0.4
9
2
18


18
4
5
2
2
2
Chloroform in Human Autopsy
Age of
Subject Sex Tissue

76 F Body fat
Kidney
Liver
Brain
76 F Body fat
Kidney
Liver
Brain
82 F Body fat
Liver

48 M Body fat
Liver
65 M Body fat
Liver

75 M Body fat
Liver

66 M Body fat

74 F Body fat





Tissue
Chloroform
Hg/kg
(Wet tissue)

19
2
5
4
5
5
1
2
67
8.7

67
9.5
64
8.8

65
10.0

68

52





"Source:   McConnell  et  al.,  1975
                                     4-14

-------
     In contrast  to  the high tissue levels  of  chloroform found in response to




inspired  air  concentrations  required  for  anesthesia, McConnell  et al.  (1975)




recently  analyzed  post-mortem  tissue  from  eight  persons, four males  and four



females, with an age  range  of 48 to 82, living in the United Kingdom in ordinary




non-industrial  circumstances,  for chloroform  and other  halogenated  compounds




(carbon   tetrachloride,   trichloroethylene,   perchloroethylene,   hexachloro-




butadiene).   Significant tissue  levels of  three  chlorinated  hydrocarbons were




found.  Chloroform level,  [ig/kg wet  tissue weight,  were:   body fat, 5  to  68




(average of 51); liver,  1 to  10 (average of 7.2); kidney, 2 to  5; and brain, 2 to




4 (Table 4-4).  Presumably, these tissue levels of chloroform were derived from




air, foodstuff  (Table 4-4), and drinking water contamination (OSHA, 1978; Dowty




et al., 1975; Symons et al.,  1975).




     There have  been few controlled exposure studies  in  animals investigating




the distribution  of  chloroform  in body tissues and  determining dose-dependent




tissue concentrations.  Chenoweth  et  al.  (1962) determined blood and tissue con-




centrations of chloroform in  two normal fasted dogs after 2.5 hours of surgical




anesthesia.  Concentration  of chloroform  in  the inhaled stream during anesthesia




was not determined, but anesthesia was judged to be satisfactory at an arterial




level of 45 to 50 mg/d£.  Blood and tissue chloroform was determined by infrared




spectroscopy after tissue extraction in cold carbon disulfide  and distillation.




Table 4-5 shows the  relative concentration  of  chloroform  in  body tissues.  The




highest concentrations were found in fat tissue, some  10-fold greater than  blood,




and in adrenals (4-fold  greater than  blood);  the concentrations in brain,  liver,




and kidney were similar  to  blood.




     Cohen and  Hood  (1969) used low-temperature whole-body autoradiography  to




study  the distribution of    C-chloroform  in mice.    Individual  mice  were




administered  2.4 (i£  of    C-labeled  chloroform  by  inhalation  over  a  10-minute
                                     4-15

-------
                                  TABLE 4-5

                Concentration  of Chloroform in Various Tissues
                    of  Two Dogs After  ?.5  Hours Anesthesia3

Arterial blood
Brain
Adrenal (total)
Fat , omentum
Right ventricle
Skeletal muscle
Lung
Liver
Spleen
Kidney
Bile
Thyroid
Pancreas
Urine
Dog A
|ig/gm wet
275
298
1185
2820
21*4
189
147
282
237
225
209
460
296
57
Dog B
tissue weight + 5%
397
392
1305
1450
314
155
336
290
255
226
205
760
350
73
 Source:  Chenoweth et al.,  1962

 Chloroform concentration was determined by infrared spectrometry after tissue
extraction.
                                      4-16

-------
 period.   The  animals  were sacrificed  0,  15,  or  120 minutes  after exposure.




 Autoradiography  of mice killed immediately after inhalation showed the highest




 concentration  of radioactivity in body fat and liver,  while lesser and relatively



 uniform  amounts  were seen  in blood,  brain,  lung, kidney,  and  muscle.   By  120




 minutes after  exposure, a considerable decrease in total radioactivity occurred,




 now  principally  confined to liver,  duodenum, and fat.  A  mottled appearance in




 the  liver suggested a segmental or localized distribution.   Biopsy specimens were




 taken  from selected tissues in each animal  and  radioactivity determined by scin-




 tillation counting.  Table  4-6 shows  the distribution of radioactivity (chloro-




 form and  metabolites)  in these tissues, and  tissue/blood concentration ratios.




 Following sacrifice, after  10 minutes of exposure, most tissues approach a unit




 concentration  with blood.  However,  in both  fat and  liver,  the concentration




 exceeds unity.  By  15 minutes, the  ratio of  radioactivity in  brown fat reaches




 its  peak at  15 times that found in blood.  The relative concentration of radio-




 activity in the liver continues to increase until the termination of the experi-




 ment at 120 minutes, when it reaches a final value 6.7 times in excess of that in




 the blood.  Kidney  and lung tissues also  increased in relative concentration over




 the  2-hour  period  to  a value of  1.53 and  1.43  of  blood, respectively.   The




 increasing ratios of liver and kidney/blood radioactivity represent a continued




 accumulation of metabolites within these organs.  High body fat/blood concentra-




 tions shows that  adipose tissue represents an important storage site, prolonging




 retention of chloroform in the body.




     Whole-body autoradiography was also carried out by Brown  et al.  (1974)  on




male and female  Sprague-Dawley rats  and squirrel monkeys  given    C-chloroform




perorally (60 mg/kg).  Male and female rats killed 3 hours after dosage showed no




apparent sex  difference  in distribution of  radioactivity.   Radioactivity  was




greatest in body  fat and liver, while lesser  amounts  were  seen in blood,  brain,
                                     4-17

-------
2
H-
3

II

3

3
C
d-
CD
M



































































cr ju
0.5=. GO
C 3 O
•OH- C
i 	 | g k-rf
H- CD o
Oh-1 CD
(1) • •
d- W
CD [B
0 0
D. T O
CD H- 3"
d- >-i) CD
CD H- 3
•"3 O
3 CD P
H- W 3
3 O.
P £
d- CD CC
H- "J O
0 CD O
3 Q.
Cfl p
ci-
H- — '
3 o vo
CD vo
P ->
O vJl
cr-
0 0
>-*> -3

s; ro
0 0
W 3
3 H-
p ci-
M CD
M W

P >-*>
d- 0
CD M
P O
0 S
cr H-
3
d- fjq
H-
3 _»
CD O
w g
CD H-
,o £3
C C
CD d-
3 CD
O
CD H-
3
14-^ P"
1— '
GO P
KJ H-
• O
O
•"$
O
cr
i— •
o
*~3
o


3

O
P
d-
P

CD
"O
*-^
CD
W
CD
3
d-
CO

O
2£
3

H-t)
P
d-





uo
	 k
V_J1
OO

1+

U)
CO
-tr





_k
-Cr
vO
O
1+
VO
CD
jr




ro
1 +
-Cr
O
UJ










— k
1°
w





— k
-^
-j
o





^
Ig
o






tx]
P
d-











	 k
cr.
— j
jrr

1 +

ro
0
— k






vO
UJ
1 +
vO
ro
— j




ro
CTv
CT>
1 +
UO
o
_»











ON
-Cr






vO
ro
Ul





^
Lk
CO






f-«
H-

CD
1










jr
O
— J

1+


uo
CTi
•
VO




ro
o
oo
1 +

vo
UJ




ro
VJI
O
1 +
— i
VO











— *
VJl
cr>






ro
_k
o





cr.
1^1
CT»






p>vj
H-
O.
3
CO
<
Ul vo -q
— ' vO Ul





ro ro uo
O> UO — 3
l-f 1+ 1+
cr. ro -cr
vo — q o











o o ->
CO CO O
-d -Cr 0






_k _k _k
O — » O
-~j ro o





0 0 _>
— a cr. o
o uo o








^~3
H-
CO
W
C
CD









O

S
H-











Ul
g
H-






ro
o
3
H-
•-S











0
3
H-
3




	 k
Ul
3
H-
3



	 k
ro
o
3
H-
3





























t— j
O
d-
P
»
CL
J-J.
O

O
d-
L-J.

^<
(T
O


d-
W
3
H-
\
3
W
















^
H-
W
W
C
CD
n~t
UJ
|_i
8
Qu
33
P
^-f-
H-
O

































0
3
O
CD
3
^
P
^J"
O
< W
p
-5 0
l-j. i-t,
0
Cfl P
D.
t— j H"
H- O
M P
W O
C d-
CD H-
W <
O d-
f~N ^^d
cr o
CD 3"
M
s o
O T
c o
M *~^
CD O
J
^^ 3
~?
S h- '
M C
•— ' M
J" s
V CD
P
cr
o
1— "
H-
d-
CD
Cfl
^~~~^

H*

























































H
W
-Cr
1
CT.





































-------
lung, kidney, and muscle.   Squirrel monkeys showed a similar distribution, with




the exception that high concentrations of radioactivity were present in the bile




and increased with time.   Examination  of bile extract by  gas-liquid  chromato-



graphy showed  the bile radioactivity was  unchanged chloroform,  indicating an




excretion of chloroform by the biliary  route in the monkey.




     Brown and his colleagues  (Taylor et al.,  1974) also investigated the tissue




distribution of chloroform in 3 strains  of mice (CF/LP,  CBA, and C57)  by whole-




body autoradiography after oral dosing  (60 mg/kg    C-chloroform).   In the male




mice of the three strains  examined  3 hours after dosing,  the greatest amounts of




radioactivity appeared in liver and kidneys, and lesser  amounts in renal cortex




but not medulla.   Female mice showed greatest radioactivity in liver, intestine,




and bladder, with much less radioactivity  in  kidney and  little differentiation




between renal cortex  and medulla.   The same general patterns were observed 5, 7,




and  24  hours after dosing.   Biopsy samples  of  these  tissues were taken,  and




radioactivity  determined  by  scintillation  counting.    Table  4-7  shows  the



                 14
distribution of   C-chloroform  radioactivity  in male and  female mice  killed  5




hours after dosing.   There is a 3.5-fold difference between the activity present




in the male and female  kidneys of each strain.  Male mice had greater activity in




the kidneys, but  female mice  showed relatively greater  activity in the liver.




This sex difference in distribution was  abolished by castration or testosterone




administration to female  mice.   The sex  difference in  tissue  distribution of



chloroform and its metabolites may relate to the nephrotoxic effect of chloroform




that occurs  in male  mice  but  not  in  female mice  (Bennet and Whigham,  1964;




Culliford  and   Hewitt,   1957;  Hewitt,   1956;   Shubik   and   Ritchie,   1953;




Eschenbrenner and Miller,  1945a,  1945b).   The pattern of  tissue distribution of




chloroform in  mice also  depends on mode of exposure.   Cohen and  Hood (19&9),




after chloroform  inhalation, found  highest levels in body fat (Table 4-6), while
                                     4-19

-------
                                  TABLE 4-7

                                    ill
             Tissue Distribution of   C-Chloroform Radioactivity

            in CF/LP Mice After Oral Administration (60 mg/kg)a'b
Mean DPM/100 mg
Tissue
Liver
Kidney
Brown fat
Blood
Male
18,157 (
13,759 (
1011
2910
(6)
1898)
1047)
(80)
(423)
wet weight (SEM)
Female
21,535 (
3920
1074
2906

(6)
2097)
(533)
(54)
(457)
Source:  Taylor et al.,  1974


Similar results were obtained for CBA and C57 strains.
                                    4-20

-------
 Brown  et  al.  (1974)  and Taylor et al.   (1974) observed lower  levels  in  fat  and




 highest levels  in liver and kidney following oral dosing  (Table 4-7).  The high




 liver  levels of chloroform after oral administration may  be  due,  in part, to



 first  passage and extraction by the liver  after this route of administration, to




 differences  of  time after exposure  (2 versus  5  hours),  and to  metabolism  and




 covalent  binding of metabolites to cellular macromolecules  (see below).




     It is worth re-emphasizing that the  sex difference in tissue distribution




 and  binding of  chloroform  (and metabolites) in kidney and liver, noted by Brown




 and  his  colleagues  (Taylor et al.,  1974), appeared to  be peculiar  to mice  and




 that these workers did not observe such differences in  male and female rats or




 squirrel  monkeys (Brown et al., 1974).









 4.4. EXCRETION









     Elimination of chloroform  from  the body  is  perforce  the sum of metabolism




 and  excretion  of  unchanged   chloroform  via   pulmonary  and  other  routes.




 Unmetabolized chloroform is excreted almost exclusively through the  lungs; how-




 ever,  metabolism  of  chloroform  is  extensive,  with  the  proportion  excreted




 unchanged dependent  on body  dose.   Surprisingly,  considering  its  historical




 importance,  its longtime use as  an industrial chemical and anesthetic agent, few




 controlled experimental studies in man have been  made  on the kinetics of excre-




 tion of chloroform.









 4.4.1.  Pulmonary Excretion. Figure 4-3 shows the time-course of pulmonary elim-




ination of chloroform after accidental  inhalation exposure to a mixture of sol-




vents,  including chloroform, carbon  tetrachloride, trichloroethylene,  and per-




chloroethylene.   Stewart et al. (1965) determined, post-exposure, the  alveolar
                                     4-21

-------
 z
 o
 e
 •-
 UJ
 u
 O
 U
 or
O
UJ
                                OPERCHLOROETHYLENE   —
                                D CARBON TETRACHLORIDE —

                                A CHLOROFORM         —

                                OTETRACHLOHOETHYLENE ~~
       0 12 24 36 48
96   120
                                    144   168  192  216  240  264   288   312   336  360

                                       TIME, hr
Figure   H-3.      Exponential   decay   of   chloroform,   carbon   tetrachloride,
perchloroethylene  and  trichloroethylene   in  exhaled  breath  of  48  year-old
male  accidentially exposed  to vapors  of  these solvents.    The  alveolar  air
(end tidal) was collected  in a  50-mJl  glass  pipette,  and the sample was analyzed
by infrared spectroscopy and gas-liquid chromatography.  Initial values (30 rain.
post  exposure)   were   chloroform,  7  ppm;  carbon  tetrachloride,   9.5  ppm;
perchloroethylene, 11 ppm; and  trichloroethylene,  4  ppm.

Source:  Stewart  et al. (1965).
                                        4-22

-------
air concentrations of  these  solvents  by infrared and gas-liquid chromatography


analysis.  The kinetics of pulmonary excretion of the solvents are independent of


one another.  However, all, including chloroform, demonstrate the typical kine-


tics   of  gaseous  vapor   pulmonary   elimination,  that   has   been  observed


experimentally  for  relatively  hydrophobic,  volatile  gaseous  anesthetics  and


industrial  solvents  (Eger, 1963; Fiserova-Bergerova et  al.,  1974,  1979, 1980;


Droz  et  al., 1977).   At  termination  of exposure  with  zero concentration in


inspired  air,  chloroform  (and  the other  solvents)  immediately  begins  to be


eliminated from the body into the lungs, with blood and alveolar air concentra-


tions  describing  parallel  exponential  decay  curves with  three major components


(Figure 4-3).  These exponential  components  have  been  related by  many investi-


gators (Eger, 1963; Fiserova-Bergerova et  al.,  1974-,  1979,  1980;  Droz et al.,

1977) to  first-order kinetics  of pulmonary elimination associated with desatura-


tion of physiological  compartments in accordance with a blood flow-limited model


in which  the  rate constants are determined  predominately  by tissue perfusion,


volume of tissue distribution and by partition coefficients:



     Tissue uptake and                               FT
     desaturation of    =  F  •  X    .   exp  (	-	  • t )
     compartment,T               bl/air        -VT A T/bl



where F is blood flow through tissue compartment, V  is volume of the compartment,

X is partition coefficient, and  exp is base of natural logarithm.



                                  0.693 V
     Since three exponential  components  are  typically observed experimentally,


three physiological  tissue  compartments are included in the model described by a


three-term exponential function of the form:


              Uptake, Desaturation =  A e ~at + B e "^ + C e ~yt
                                     4-23

-------
where A, B, C are macrocoefficients  and a,  (3,  Y»  are hybrid constants  (defined




above).  These terms represent three flow-limited  major body compartments  (1)  a




vessel-rich group of tissues  (VRG) with high blood flow and high diffusion rate




constant  (VRG:    brain,  heart,   kidneys,  liver  and  endocrine  and  digestive




systems), (2)  lean body mass  (MG:  muscle and skin), and  (3) adipose tissue  (FG).




More  recently,   Fiserova-Bergerova  and  coworkers  (1974,  1979,   1980)   have




mathematically re-formulated  this  physiological, first-order model to accomodate




the effect of metabolism on uptake, distribution and clearance of inhaled  vapor




compounds.




     The half-time  (t  1/2) of elimination from the  physiological  compartments




(VRG  < MG  <  FG)  are  independent of  the  body  dose,  but  are  dependent  on




tissue/blood partition coefficients and blood/air  partition coefficients .  Since




these solvent compounds have high  solubility in body fat  (Table 4-1),  they are




eliminated  slowly  from  fat  depots with long  half-time  of elimination,  as




illustrated by  Stewart's  patient  (Stewart et al.,  1965)  in Figure 4-3.   From




Figure  4-3, it  can  be  estimated graphically that  chloroform has a half-time of




elimination from the fat compartment (FG) of =36 hours, with similar long half-




times  for the highly fat soluble compounds,  perchloroethylene and carbon tetra-




chloride.




     There  is  little  information  available  for  the half-times  of  pulmonary




elimination of chloroform  from the VRG and MG.  From the  early data of Lehmann




and Hasegawa  (1910) given  in Table H-3,  the half-time of  pulmonary elimination




from the VRG appears to be =30 minutes.  A similar estimate  can be made from the




data  of Smith et al.  (1973)  and  Morris  (1951)  at termination of anesthesia in




man;  these workers  found that   blood  chloroform  concentration  rapidly fell




exponentially from 7 to 3.5 mg/d£  within  30 minutes  (Figure 4-2).
                                      4-24

-------
      Pulmonary  elimination  of chloroform was investigated by Fry  et  al.  (1972)


                                       1 3
 in  male and  female  volunteers given   JC-chloroform  in  olive  oil  orally  (by



 gelatin capsule).  Chloroform was determined in expired air by GLC.  Their data,



 summarized in Table 4-8, show that the  amount of chloroform excreted  through  the




 lungs  within 8 hours  (expressed  as a percentage  of  the dose,  0.1  to  1.0  g),



 increased (0 to 65%) in proportion  to the dose.  Following a peak  blood  concen-



 tration (0.5 mg/dS, for a 500 mg dose)  1 hour after oral dosage,  absorption,  and



 distribution,  the  blood chloroform concentration declined  exponentially with



 three  components:  (1) a very rapid disappearance, with half-time  of  14 minutes



 possibly corresponding to VRG compartment kinetics,  (2) a slower disappearance,



 with  half-time  of  90 minutes  corresponding  to MG kinetics,  and  (3) a very slow



 disappearance, with very long half-time from adipose  tissue.  This  half-time  was




 undetermined, but  chloroform  was  detected in blood  and  breath 24 hours  later.



 Fry and coworkers (1972) noted a linear relationship for their subjects between



 pulmonary excretion of chloroform and body weight deviation from  ideal, an index




 of excessive leanness or excessive  body fat  from normal.   Their data in Figure



 4-4 show that  for  both male and  female subjects given a  standard oral dose of



 chloroform,  lean subjects eliminate via  the lungs a greater  percentage of  the



 dose, while  overweight  subjects eliminate  less chloroform.  The different slopes



 of the linear relationship  for men and women presumably  reflect the different



 proportion of  adipose  tissue in  the  two sexes.   The  bodies of  women  tend to



 contain higher proportions  of  fat than those of men  (Geigy  Scientific  Tables,



 1973).  These observations reinforce the role of adipose tissue  as a storage site



 for chloroform.




     Brown  and  his  coworkers  (Brown  et  al., 1974;  Taylor  et al.,  1974)  have




demonstrated  an  animal species difference in  the  amount  of pulmonary excretion of


 14
  C-chloroform from a standard oral dose  (60 mg/kp, body weight)  given in ol.ive
                                     4-25

-------

Subjects
8 M and F
1
1
1
Pulmonary
Dose
(g)
0.5
1.0
0.25
0.10
TABLE 4-8
Excretion of 1 ^CHCl
Dose: Percent of
Mean
for 8 Hours b
40.3
64.7
12.4
nil
Following Oral
Dosea
Range
17.8 to 66.6
NA
NA
NA
                                    -I -3
             Pulmonary excretion  of   JCO?  following  0.5  g oral  dose

                       1 O                                   V\
                    of   CHC1  •   Cumulative  percent  of dose


                       	Time after dose  (hours)	


Subjects               0.5       1.75      2.5       5.5      7.5


Male (1)               2.1      24.1      35.9      49.2     50.6


Female (62.7 kg) (1)   0.5      10.7      28.3      47.5     48.5


Recalculated from the data of Fry et al.,  1972


 vfithin 4ft of value calculated for infinite time


NA = Not applicable
                                        4-26

-------
       -6-4-2      0     +2      +4      +6     +8

       BODY-WEIGHT DEVIATION FROM CALCULATED NORMAL, kg


Figure  4-4.    Relationship  between  total 8-hour  pulmonary
excretion of chloroform  following  0.5-g oral  dose in man and
the deviation of body weight  from ideal.   The different slopes
of  the  linear  relationship  for  men  and women  reflect the
different proportion of adipose tissue in  the two sexes.
                         4-27

-------
oil.  Mice  (three strains), rats,  and squirrel monkeys excrete chloroform via  the




lungs  (6,  20, and  79%,   respectively,  of the  standard dose).   This  species




difference is primarily related to the capacity to metabolize  chloroform rather




than differences  in  pulmonary kinetics, since, as shown in Table 4-9, the percen-




tage  of  the  dose metabolized  to   CO-  is  inversely  proportional  to that  of




pulmonary  excretion.   The  mice,  48  hours  after dosing,  retained only 2%  of




chloroform radioactivity (Table 4-9).




     Withey  and  Collins  (1980)  determined  the kinetics  of distribution  and




elimination of chloroform from blood  of  Wistar rats  after  intravenous adminis-




tration  of  3,   6,   9,   12  or   15 m/kg  of  chloroform  given in   1  ml  water




intrajugularly.   For all doses, the blood decay  curves exhibited three components




of exponential disappearance of chloroform (a,  (3> YcomP°nen^s) anc* "best" fitted




a first-order three  compartment model.   Table  4~'ib summarized the values obtained




for  the  kinetic  parameters.   For  volatile,  lipophilic compounds,  for  which a




major route of elimination is  pulmonary, experiments utilizing dose administra-




tion  via relatively large intravenous bolus injections  (relative  to  rat total




blood volume), and which measures only blood  chloroform disappearance, provide a




number  of  problems  for  data  interpretation  of  elimination   (pulmonary  and



metabolism)  and/or   tissue  distribution.    In these  experiments,  pulmonary




elimination,  which  is rapid  for organic solvents,  occurs  simultaneously with




distribution  and  metabolism;  in contrast, experiments in  which the  animal is




preloaded  by  oral  or  inhalation administration, distribution is  more readily




separable  from pulmonary  elimination.   After intravenous administration (Table




4-tlb),  the rate constant  k   (for elimination of chloroform from the central




compartment  blood out  of the  body (principally via  pulmonary excretion and/or




metabolism) was dose-dependent and consistent with a  half-time of elimination of




only 3.6 min for the lowest dose and only  6.2 rain for the highest dose.  Since the
                                      4-28

-------
                                      TABLE 4-9
                                                        14,
               Species Difference in the Metabolism of  ' C-Chloroform
                               (Oral Dose of  60 mg/kg)*
14
C-radioactivity 48 hours after dose
Species No. Mean values as percent dose
Expired 14CHC1
or metabolites
Mice
CF/LP, 19 6.1
CBA, C57
strains
Rats
S-D 6 19.7
Squirrel 6 78.7
Monkeys
Expired Urine +
14
C00 Feces Carcass Total
85.1 2.6 1.8 95.6
65.9 7.6 NR 93.2
17.6 2.0 NR 98.3
"Recalculated from the data of Brown et al.,  1974

NR = Not recorded
                                          4-29

-------
      Table  4-/0.  Kinetic Parameters for Chloroform After I.V. Administration to Rats
Dose* Vd
mg/kg ml
a
6
Y
k
e
kiz
k2i
kis
k3i
min"1 ± S.E.
3.0 45.07
± 0.04
6.0 53.57
± 12.61
f 9.0 64.46
o ±14.26
12.0 80.62
± 20.20
15.0 89.13
± 12.83
15.0; t, min
"2
0.72
± 0.11
0.64
± 0.13
0.64
± 0.084
0.32
± D.20
0.35
± 0.048
2.1

0.135
± 0.001
0.081
± 0.01
0.095
± 0.0001
0.060
± 0.028
0.070
± 0.006
9.9

0.0287
±0.0064
0.0158
± 0.0019
0.0189
± 0.0009
0.0074
± 0.0056
0.0134
± 0.0005
51.7

0.1907
± 0.0239
0.1874
±0.0356
0.1284
±0.0217
0.1035
± 0.0232
0.1124
± 0.0110
6.2

0.2346
± 0.0651
0.2681
± 0.0631
0.2529
± 0.0489
0.1071
± 0.0774
0.1104
± 0.0256
6.3

0.2575
± 0.0316
0.1862
± 0.0188
0.2421
± 0.0064
0.1192
± 0.0676
0.1523
± 0.0188
4.6

0.0921
± 0.0008
0.0730
± 0.0246
0.0594
± 0.0034
0.0395
± 0.0253
0.0396
± 0.0104
17.5

0.0421
± 0.0098
0.0233
± 0.0032
0.0281
± 0.0042
0.0106
± 0.0084 '
0.0193
± 0,0018
35.9

2 to 4 rats/dose  .  From Withey et al.,  1982.

-------
half-times  for distribution into  other tissue  compartments from  the central
compartment  blood  have  longer  half-times,  it is  likely  that,  of  the  dose
introduced  into  the blood, a  major portion  (depending  on  dose)  was excreted
within a few minutes by the lungs.  Further indication that  the dose and/or mode
of administration influenced the distribution and elimination of chloroform was
shown by the proportional increase of the apparent volume of distribution, V, (45
ml;  3 mg/kg  to 89  ml;  15 mg/kg)  and the decrease with dose  in  values for rate
constants of  transfer  from blood to other  tissue  compartments.   The volume of
distribution (Vd) of chloroform was 89 ml for the highest  dose or about 22% b.w.,
surprisingly low for a lipid soluble compound that is known to diffuse into all
the major organ systems (Tables 4-5 and 4-6).   Adipose tissue is  known to  be  a
major tissue compartment for chloroform [Section 4.3]; the clearance of chloro-
form from perirenal  fat was found  to  be slow with a half-time of  106 min,  and
since the  rate constant,  k    was  given  as 0.0193  min~  (for  15  mg/kg  dose)
indicating a  half-time of 36  min, the  adipose tissue  appears to  be a  deep
compartment.   These investigators  believe  that  their  kinetic  data  shows  no
evidence in the rat  of  nonlinear or  dose-dependent Michaelis-Menten kinetics and
they suggest that a dose of 15 mg/kg is below hepatic metabolism saturation.
4.4.2. Other Routes of Excretion.  Chloroform is not  eliminated  in significant
amounts from the body by  any route other than pulmonary.  Studies  of chlorinated
compounds in the urine after  chloroform inhalation exposure or peroral dosage to
animals and  humans  have  failed to detect  unchanged chloroform  (Brown  et  al.,
1974; Fry  et al.,  1972).   Brown  et  al.  (1974) identified  chloroform in  high
concentration in the bile of squirrel monkeys  after  oral    C-chloroform  dosage,
and suggested an active enterohepatic circulation in this species.  For monkeys,
                                     4-31

-------
they found only 2% of dose radioactivity in combined urine and  feces  collected




for 48  hours  after dosing (Table  4-9),  and only 8  and 3%  for rats  and mice,




respectively.







4.4.3. Adipose Tissue Storage.  There is no definitive  experimental  evidence  in




the  literature concerning  bioaccumulation  after chronic  or  repeated  daily




exposure to chloroform.   However,  there are practical  reasons  to  believe  that




extended residence in body fat occurs.  In man, chloroform has  a relatively  high




fat tissue/blood  partition coefficient  of  35 (Table 4-2), a  long  half-time  of




elimination from adipose  tissue compartments of =36  hours (Figure  4-3),  and  it




has been  detected in blood and breath  24  to 72 hours  after  a  single exposure




(Stewart, 1974; Fry  et al.,  1972;  Stewart  et al., 1965.).  Figure  4-5  clearly




shows  the   slow  elimination  of chloroform  from  the  adipose  tissue of  dogs




following a  3-hour  anesthesia  (Chenoweth   et al.,  1962).  Despite  the  rapid




exponential  decline  of   blood  levels   of  chloroform  within  3  hours  after




termination of anesthesia, significant levels of  chloroform were still  present




20 hours later.




     Fry et al. (1972) has provided  indirect evidence  in  man  of the storage  of




chloroform in  body fat (Figure 4-4), while analysis of body fat of  animals given




a single inhalation exposure  or oral dosage demonstrated marked accumulation  of




chloroform in  this  tissue  (Taylor  et al.,  1974; Cohen  and Hood,  1969)  (Tables




4-5, 4-6, 4-7).  Particularly pertinent are the observations of McConnell et al.




(1975), who demonstrated  the occurrence of  significant amounts of  chloroform




(and other chlorinated hydrocarbons) in autopsy tissues (highest concentrations




in body fat) of humans exposed only to ordinary ambient air  (Table  4-4), and  of




Conkle  et al.  (1975), who  analyzed the GC-MS alveolar  air  of eight  fasting




healthy men  working  in  a nonindustrial  environment  and  found in  three men
                                     4-32

-------
         "!;HCI?
 PENTOBARBITAL

TIME.hr
Figure  4-5.    Blood  and  adipose  tissue  concentrations  of
chloroform during and  after anesthesia in a dog.  Note the high
prolonged levels of chloroform in adipose tissue (broken line)
for 20 hours even  after  rapid exponential fall in blood  con-
centration  (solid  line)  with termination of chloroform anes-
thesia after three hours.

Source:  Chenoweth et al. (1962).

-------
significant rates  of pulmonary excretion of chloroform  (Table  4-10),  as  well  as




other halocarbons  (for  example,  methylene  chloride,  dichlorobenzene,   methyl-




chloroform) in all eight men.








4.5. BIOTRANSFORMATION OF CHLOROFORM









4.5.1. Known Metabolites. The haloforms, and chloroform in particular, long have




been  known  to undergo  extensive  mammalian  biotransformation.   Zeller  (1883)




clearly demonstrated  an increased  daily urinary inorganic chloride excretion




representing 25 to 60% of the  dose in dogs given oral  doses  of chloroform (7  to




10 g  in gelatin capsules).   Eighty years  later,  Van Dyke et  al.  (1964),  using




  Cl-chloroform,   confirmed  in rats that  the  extra  urinary inorganic  chloride




originated from the metabolism of chloroform.  Zeller  (1883)  also found, in the




urine of his dogs  given  chloroform,  a levo-rotatory oxidative metabolite that  he




suggested to  be the glucuronide of  trichloromethanol, a  compound  only  recently




postulated as an  intermediate  of  Pnc-n oxidative metabolism (Figure  4-6).  Van




Dyke and coworkers (1964) also found evidence for   (^metabolites (-2%  dose)  in




the urine of  their rats  given  chloroform.   However, other  investigators (Brown




et al., 1974; Fry et  al.,  1972;  Paul and  Rubinstein,  1963; Butler,  1961)  with




newer methodologies have not been able to identify lesser chloromethanes in the




urine or breath of mouse, rat, or man after chloroform exposure.




     In addition to the chloride ion, it has been established from both in vivo




and  in vitro studies  that  the major  end-product metabolite of  chloroform  is




carbon  dioxide  (COp)   (Brown  et al.,  1974;  Fry  et al.,  1972;  Rubinstein  and




Kanics, 1964; Van Dyke  et al.,  1964;  Paul and Rubinstein,  1963),  with  phosgene




identified as the immediate precursor metabolite  from  In vitro studies (Mansuy




et al., 1977;  Pohl  et al.,  1977;  Ilett et al., 1973)  (Figure 4-6).   C02 from
                                      4-34

-------
                              TABLE 4-11




      Levels  of Chloroform in Breath of Fasted Normal Healthy Men*
Subject
A
B
C
D
E
F
G
H
Age
34
28
33
38
47
28
38
23
Chloroform excretion, |ig/hr
2.0
ND
ND
ND
11.0
ND
ND
0.22
*Source:   Conkle et  al.,  1975




ND = Not  detected
                                   4-35

-------
MAJOR AEROBIC PATHWAY
       H-C C13
                   P450, O2
                   NADPH
                 MICROSOMES
               [HOCCI3]
      ACCEPTOR
      PROTEIN
          I
         CO —*
     H2C - CH - CX- OH
      S  NH
       \ /
        C
         it
        0

 2-OXOTHIAZOL1DINE-
   4-CARBOXYLIC ACID
                                       -HCI
                O=C CI2
                                 PHOSGENE
                                         >
                                 CYSTEINE
                               CONDENSATION
 H2O
	•»- 2 HCI + CO2
                                GLUTATHIONE
                                CONJUGATES?
 MINOR ANEROBIC PATHWAY
           CHCI3
                 ANEROBIC
                  NADPH
           -i*~ P450 - Fe2+ : C CI2 + HCI

                     I  +H2O


P450 - Fe2+ CO —*	  CO + 2 H Cl
                  REDUCED
                MICROSOMES
  Figure  4-5.  Metabolic pathways of chloroform biotransformation.
  (Identified  CH  Cl., metabolites are underlined.)
                              U-36

-------
chloroform metabolism is primarily excreted through the  lungs,  but  a small per




centage (<1%) is incorporated into endogenous metabolites and excreted into the




urine as  bicarbonate,  urea,  methionine, and  other amino acids  (Brown et al.,



1974).  Carbon monoxide (CO)  has  also  been identified as  a very minor metabolite




of anaerobic  chloroform metabolism (Figure  4-6),  both  from In  vitro studies




(Ahmed  et  al.,  1977;  Wolf  et al.,  1977)  and _in  vivo  animal  studies (Anders




et al.,  1978; Bellar et al.,  1974).




     In addition to chloroform metabolites that are excreted, phosgene and other




"reactive intermediates" of chloroform metabolism  interact  with  and covalently




bind  to  tissue  acceptors such as protein  and lipids  (Docks  and  Krisna,  1976;




Uehleke and Werner,  1975; Brown et al., 1974; Ilett  et al., 1973; Cohen and Hood,




1969; Reynolds, 1967; Cessi  et al.,  1966).




     The liver is the principal site of chloroform metabolism, although Paul and




Rubinstein (1963) and Butler  (1961)  found that rat kidney,  adipose  tissue,  and




skeletal muscle  also converted  chloroform  to CO  (25,  0.8,  and 8%  of  liver,




respectively) .









4.5.2.  Magnitude  of Chloroform  Metabolism.    Chloroform   is  metabolized  to




differing extents in man and other animal species.  Since chloroform and  other




halogenated  hydrocarbons  are  thought   to   produce   pathological  effects   by




metabolism in target tissues to reactive intermediates that  covalently bind  to




macromolecules (Chapter 5),  the  total  capacity to metabolize chloroform as well




as individual tissue sites of metabolism are important  determinants  of expected




interspecies  differences in  toxic susceptibility.  This interdependence between




intensity  of  toxic response  and  metabolism,  and  interspecies  differences  in




magnitude  of  metabolism, are  important considerations  in  extrapolation  from




experimental  animal to man (Reitz et al.,  1978).
                                     4-37

-------
     Few studies have been made of the capacity of man to metabolize  chloroform;




virtually no studies have been made  of the  phamacokinetic,  endocrine,  genetic,




and environmental  factors  modifying metabolism in man.   The early studies  of




Lehmann and Hasegawa  (1910)  on  the retention  of  chloroform from inspired  air




inhaled  by  three  volunteers  (4500  to  5000 ppm,  average of  64$)   (Table  4-3)




suggest  that  36%  of  pulmonary  uptake  of   chloroform  in man  is metabolized.




Similarly, a retention  value  of  67$,  calculated from the data of Smith  et  al .




(1973) for patients  inhaling  10,000 ppm  chloroform during  surgical  anesthesia




(Figure 4-2),  indicates  33% chloroform metabolism during inhalation exposure.  A




similar estimate of  the extent  of chloroform metabolism during  anesthesia  has




been  made  by Feingold  and Holaday (1977).   These  workers simulated, with  a




computer, chloroform inhalation  kinetics  using a non-linear  whole-body  compart-




ment al model, and  found that  the percent of  chloroform  uptake metabolized  was




30%.  This rate of metabolism  remained  constant during 8 hours of anesthesia,  and




continued for several days following termination of anesthesia, presumably from




chloroform stored during anesthesia  in adipose tissue.




     Fry et al.  (1972)  have  investigated the  metabolism of chloroform  in  man




after  single  oral  doses.   Isotopicall y  labeled  ^C-chloroform dissolved  in




1.0 m£ olive oil/gelatin capsule was given  to 12 healthy male and female  volun-




teers  (58 to  60 kg body  weight)  at doses  of 0.1 to  1.0 g.   For  two  of  the




volunteers,  pulmonary excretion of  13CO  in  expired  air  from  the metabolism of




chloroform was serially  connected over a 7.5 hour period and analyzed by mass




spectrometry.  The  results given in Table 4-8 show that 49 and 51$  of a 0.5 g dose



                     13                                           1 ?
was metabolized  to   COp.    Pulmonary  excretion  of unchanged   JC-chloroform




during a  comparable period (8 hours)  in a separate experiment  with these  two




subjects were 67 and  40$,  respectively.  No metabolites other than  C02  (e.g.,




methylene dichloride, tetrachloroethane)  were found in expired  air,  and chloro-
                                     4-38

-------
form was not found in the urine.  These results indicate that:   (1) virtually all




of an oral chloroform dose (0.5 g) can be accounted for by pulmonary excretion of



C02 and unchanged chloroform,  (2) metabolism of chloroform to CC^ is  =50% of this



dose, and (3) absorption and metabolism are rapid and virtually complete within 5




hours as shown in  Table 4-8,  possibly because of  first pass through the liver.



In this respect, the kinetics  of metabolism  of oral doses  may  differ substan-



tially from inhalation doses.   Furthermore, the data  of Table 4-8 suggest that



the fraction of the dose metabolized  is dose-dependent.   Thus,  an oral dose of



0.1 g was completely metabolized  (100/0,  with no  chloroform excreted unchanged



through the lungs;  but  for a  1.0 g  dose,  65% was  excreted and only  35%  meta-



bolized.  These results  suggest that metabolism is rate-limited in man, since a



diminishing proportion of dose is metabolized with increasing dose (Table 4-8) .



     In man, Chiou (1975)  has shown that  up to 3&% of an oral dose of chloroform



is metabolized in  the liver,  and up  to about M%  is  excreted  intact  from  the



lungs before the  chloroform reaches the  systemic circulation,  an  example of a




first-pass effect.



     Animal experiments demonstrate  a marked  species difference in  the metabo-



lism  of  chloroform.   Early experiments  in the 1960's  by Paul and  Rubinstein



(1963), Van  Dyke et al .  (1964), and  Cohen and Hood (1969) with mice  and rats



given   C-chloroform indicated a minimal metabolism of chloroform (~W) occurred
in these species.  More recent  studies  by Brown and his coworkers (Brown et al . ,



1974; Taylor et  al . ,  1974) have shown that mice and rats metabolize chloroform to



CO  extensively (65 to 85%),  and to a greater extent than non-human primates or



man.  These  investigators gave equivalent oral doses  (60 mg/kg)  of   C-chloro-


                                                                    14
form to mice (3 strains),  rats, and squirrel monkeys, and determined   C-labeled




chloroform or volatile metabolites and 1J4C02 in  expired  air,  1  C-radioactivity




in urine, and   C-radioactivity remaining in animals  at  sacrifice  4 to 8 hours
                                     4-39

-------
after dosing.   Total  recovery  of    C-chloroform radioactivity was excellent and




accounted for  93 to 98%  of the administered  dose.   Their results are summarized




in Table 4-9.   Using 1  C02 as  a measure of the fraction of the chloroform dose



metabolized (intermediate chloromethane metabolites were not found  in  breath or




urine), mice metabolized 8555 of the dose,  rats, 66%,  and squirrel monkeys, 18%.




A  further  2 to 8% of    C-radioactivity (CO  incorporated into  urea, bicar-




bonate, and amino acids) were found in  urine.  They found no strain  difference in




mice,  or  sex  difference  in mice,  rats, or monkeys  in capacity to metabolize




chloroform, or  in  tissue distribution  and  binding of metabolites, except for




mice, where kidney radioactivity concentration was  greater in males than females




and lesser in  livers of  males  than  females (Table  U-7).  These  findings of  Brown




and coworkers  of large  interspecies differences for metabolism of chloroform and




marked sex  differences in mice  (but not other species) for tissue distribution




and  covalent  binding of  intermediate metabolites  to  tissue macromolecules  in




liver and kidney emphasize the difficulties  and dangers of extrapolating studies




in lower animals to man  (Reitz et al.,  1978).









4.5.3- Enzymic  Pathways  of Biotransformation.  It  has been postulated for many




years  (Docks  and Krishna,  1976; Uehleke and Werner,  1975; Brown  et al.,  1974;




Ilett  et al.,   1973;  Reynolds,  1967;  Paul  and  Rubinstein,  1963)  that a reactive




metabolite  of CHC1-  is  responsible for  its liver and renal toxicity in man (von




Oettingen,  1964; Conlon,  19&3) and experimental animals (Bhooshan  et al.,  1977;




Pohl et al., 1977; Ilett et al.,  1973; Klaassen and Plaa, 1966), and possibly the




production  of  liver tumors in mice  (Eschenbrenner and  Miller,   1945a).    For




example, when rats or mice are treated with   C-chloroform, the extent  of hepatic




necrosis  parallels  the  amount of    C-label  bound irreversibly to  liver protein




 (Docks and  Krishna,  1976; Brown et al., 1974; Ilett et  al., 1973).   Both necrosis
                                      4-40

-------
and  binding are potentiated  by  pretreatment of animals  with phenobarbital, a



known inducer of liver microsomal metabolism, and inhibited  by pretreatment with


the  inhibitor piperonyl butoxide.  Chloroform administration  also decreases the


level  of  liver glutathione  in  rats  Pretreated  with  phenobarbital,  further



suggesting that a reactive metabolism is produced (Docks and  Krishna,  1976; Brown



et al.,  197*0.   The results of in vitro studies with rat and  mouse liver micro-

                                                              iti
somes support  the in vivo observations by  establishing that   C-chloroform is


metabolized  to  a reactive  metabolite which  binds  covalently  to  microsomal



protein  (Bhooshan  et  al.,  1977;  Sipes et al.,  1977; Uehleke  and Werner,  1975;



Ilett et al., 1973).  This metabolic process is oxygen dependent and appears to



be mediated  by a  cytochrome  P..™ which  is inducible   by  phenobarbitol (Sipes



et al.,  1977; Uehleke and Werner, 1975; Ilett et al.,  1973).


     The  demonstrations   by  Pohl et al.  (1977) and  Mansuy  et al.   (1977)  of


carbonyl chloride  (phosgene)  formation from chloroform  by  rat  microsomal  pre-


parations  suggest  that  phosgene may be  the key causal  agent for these  toxic



effects.   The  finding of Weinhouse and collaborators  (Shah et  al.,  1979)  that



phosgene is also a reactive metabolic  intermediate  in  the metabolism of carbon



tetrachloride emphasizes basic similarities in the  metabolism and toxicities of



these two chloroalkanes.   Figures 4-6 and  4-7 show for  comparison the currently


proposed pathways of metabolism of chloroform and carbon tetrachloride.


     Figure 4-6 indicates that the initial step in  the  metabolism of chloroform


involves the oxidation of the aliphatic carbon  (H-C) to trichloromethanol  by a



phenobarbital inducible cytochrome Ph,-0 (Sipes et al.,  1977; Uehleke and Werner,


1975; Ilett  et  al.,  1973).    This metabolic step has  been  suggested by Mansuy


et al. (1977) and Pohl et al.  (1977)  as the precursor of phosgene formed by rat



microsomes  in vitro from  chloroform.   Phosgene was  confirmed as  a metabolite by



reaction with cysteine  to give  2-oxothiazolidine-4-carboxylic  acid which  was
                                     4-'41

-------
  ACCEPTOR
     1
     CCI
  C CI3C C CI3
    CHCI3
                      CCL4
                                  REDUCTIVE
                                  DECHLORINATION
                                  ANAEROBIC
                                  MICROSOMES
                                  NADPH
P450 - Fe2+ • C CI3 + C\'
                      lP450-Fe3+.CI3COH]

ACCEPTOR
PROTEIN i
1 '

•HCI

•CCL2
                                            H2O
                              • P450 - Fe2+C CI4
                           P450 - Fe2+ : C CI2 + CT

                                 I  H20

                                 CO + HCOOH
                                                     LIPOPEROXIDATION
                                                     CONJUGATION
                                                     MALONALDEHYDE
                                                    2 H Cl
                           PHOSGENE
H,C-CH-COOH

  I  I
  S NH
   V
    CYSTEINE
    CONDENSATION
  2-OXOTHIAZOLIDINE
   4 • CARBOXYLIC ACID
 Figure H-7.   Metabolic pathways  of  carbon tetrachloride  bio-
 transformation.  (C Cljj metabolites  identified are underlined).

 Source:   Shah et al. (1979).
                              4-42

-------
identified by GC-CIMS.  Trichloromethanol  is  highly  unstable and spontaneously




dehydrochlorinates to produce phosgene (Seppelt,  1977).  The  electrophilic phos-



gene reacts  with water to  yield  COp,  a  known  metabolite  of CHCl^  in  vitro



(Rubinstein and  Kanics,  1964;  Paul  and Rubinstein,  1963)  and  i_n  vivo (Brown




et al., 1974;  Fry et  al.,  1972), with  protein  to form a covalently bound product




(Pohl et al.,  1977; Sipes et al.,  1977; Uehleke and Werner,  1975; Brown et al.,




1974; Ilett et al.,  1973),  or with cysteine  (Pohl et al.,  1977),  and possibly




with glutathione (Docks  and Krishna,  1976;  Brown et al.,  1974).  The finding that




deuterated chloroform (CDC1.J  depletes  glutathione in the  livers  of rats less




than CHC1  supports this notion  (Docks and Krishna, 1976).




     The postulated oxidation of the  C-H bond of  chloroform by P45Q to produce




trichloromethanol  which  spontaneously yields  phosgene is  further  supported  by




the observations of Pohl and Krishna (1978). These workers  found that chloroform




metabolism to phosgene by rat liver microsomes  is oxygen  and NADPH dependent, and


                                                                           1 R

inhibited by CO and SKF 525-A.   Moreover,  in  the  presence of cysteine and    0^


            1 R
atmosphere,   0  is incorporated into the 2-oxo position of 2-oxothiazolidine-4-




carboxylic acid.   Oxidative cleavage of the  C-H  bond  appears to  be the rate-




determining step,  since  deuterium  labeled  chloroform (CDC1,) is biotransformed




into phosgene slower than CHC1  • CDCl, appears also to be less hepatotoxic than




CHC1    Pohl (1980) has  further  characterized the metabolism of chloroform  in rat



liver  microsomes  by measuring  the covalent  binding of    CHC1., and C HC1-,  to




microsomal  protein.   Chloroform does  not  appear to be  activated  by reductive




dechlorination  to the radical  •CHC12>  because the  3H-label does  not  bind  to



                              14
microsome protein  as does the    C-label.




     Figure 4-7 summarizes  current knowledge of the biotransformation of carbon




tetrachloride.  The first step  is  a rapid  reductive formation of the trichloro-




methyl  ( -CCl-)  radical  by complexing with one or  more  of the PJICQ cytochromes
                                      4-43

-------
(Shah et al . ,  1979;  Foyer et al . ,  1978;  Recknagel  and  Glende,  1973).    This




radical  undergoes  several  reactions in addition to binding to lipids  (Villarruel




and Castro, 1975;  Uehleke and Werner,  1975;  Villarruel  et al . ,  1975;  Gordis ,



1969; Reynolds, 1967)  and protein (Uehleke  et  al . ,  1977;  Uehleke  and  Werner,




1973), although not to nucleic acids  (Uehleke et al., 1977;  Uehleke  and Werner,




1975; Reynolds, 1967).   Anaerobically, the  addition of a  proton and electron




yields  chloroform  (Glende et al . , 1976;  Uehleke et al . ,  1973;  Fowler,  1969;




Butler,  1961), dimerization  to  hexachloroethane (Uehleke et al . ,  1973; Fowler,




1969), or further  reductive  dechlorination to CO via the carbene,  : CC   (Wolf
et al., 1977).  Aerobically, the «CC1  radical  is  oxidized by the P^Q system to




trichloromethanol  (Cl-COH) ,  which is  the  precursor  of  phosgene  (ClpCO)  that




Weinhouse and colleagues (Shah et al . ,  1979)  have  shown to be an intermediate in




carbon  tetrachloride  metabolism  by   rat  liver   homogenates.     Hydrolytic




dechlorination of phosgene yields C0? (Shah et al . ,   1979).




     Under normal physiological conditions  (i.e.,  aerobic conditions), a minimal




formation of chloroform might be expected to  occur.   Carbon tetrachloride yields




a chloroform most readily i_n vitro under anaerobic conditions and its formation




is inhibited by oxygen (Uehleke et al . , 1977; Glende  et al . ,  1976).  Shah et al .



(1979) observed that chloroform does  not compete successfully with carbon tetra-




chloride for  initial  binding to Ph,-n cytochrome  (Sipes et al . ,  1977; Recknagel



and Glende, 1973).  Wolf et  al.  (1977)  also  found that binding of chloroform to




reduced cytochrome  Pj.™ was  very slow  compared to that of carbon tetrachloride.




     Anders and coworkers  (Anders et al . ,  1978; Ahmed et al . , 1977) have shown,




as  they have  for  di ha lorn ethanes , that  trihal omethanes ,  including chloroform,




also yield CO as  a metabolite.  Intraperitoneal administration of  haloforms  ( 1 to




4 mmoles/kg) to rats led to dose-dependent  elevations in blood CO levels. Treat-




ment of the rats  with  either  phenobarbital  (but not 3-methyl-cholanthrene) or SKF

-------
525-A, respectively,  increased or decreased metabolism to CO.  The order of yield




of CO from the iodoforms was greatest for iodoform > bromoform > chloroform for




the same dose.  Thus, chloroform was minimally metabolized to CO (i.e., to less



than one-tenth of  that  for iodoform  or bromoform).  Similar findings were made by




these workers with rat  liver microsomes  (Ahmed et al., 1977).  Metabolism of the




haloforms to CO by rat  liver microsomes required NADPH,  could proceed anaerob-




ically but was increased 2-fold  by 0~, was increased by pretreatment with pheno-




barbital and inhibited by SKF 525-A  or COC12 pretreatment, and was stimulated by




glutathione or cysteine addition in  both anaerobic (3-fold) or aerobic (8-fold)




conditions.  These results suggested that haloforms  were metabolized to CO via a




cytochrome  PMJ-Q  dependent system.   However,  chloroform  was  a  poor  substrate




compared to iodoform or bromoform, yielding <2% of its quantity of CO as formed




from equimolar concentrations  of these  halomethanes.  Wolf  et  al.   (1977)  also




found that chloroform,  to a  very limited extent, was metabolized to CO by reduced




rat Pjjj-Q preparations.   These  workers investigated the spectral and biochemical




interactions of a series of  halogenated  methanes with rat liver microsomes under




anaerobic  reducing  conditions.    Tetra-  (e.g.,  CClj.)  and trihalogens  (e.g.,




CHC1 ) all formed complexes  with reduced cytochrome  PJ,™ with absorption peak at




460 to 465.   A  shift  to 454 occurred  with CO formation and subsequent  complexing




of  CO  to Pjj50.   CO formation  required NADPH, was  higher in microsomes  from




phenobarbital and 3-methylcholanthrene-treated rats, and was not found  at  high




oxygen concentrations (<8%) . Figure 4-8 shows the relative  rates of CO formation




from carbon tetrachloride and other  polyhalomethanes.  Chloroform, in  comparison




to  carbon  tetrachloride, was  a very poor  reaction substrate, and  binding  of




chloroform to  reduced  cytochrome  P^   was extremely  slow  compared to that  of




carbon tetrachloride.  Wolf et  al. (1977) proposed the reduction  sequence shown




in Figures 4-6 and  4-8 for the reductive  dechlorination  of chloroform  and  of
                                     4-45

-------
      120
      100
   o
   D.
   01
   "o
   c
   2
   O
   t-
   DC
   O
   LL
   O
   u
       20  -
                                                       20
Figure U-8.  Rate of carbon monoxide formation after  addition  of
various halomethanes  to sodium  dithionite-reduced liver  micro-
somal preparations from phenobarbitol-treated rats.  Note the low
rate of metabolism of chloroform to CO compared  to carbon  tetra-
ohloride.
Source:  Wolf et al . (1977).

-------
carbon tetrachloride to  yield  CO via a carbene (CClg) intermediate.  The physio-




logical importance of this pathway of metabolism appears to be more significant




for carbon tetrachloride than for  chloroform.








4.6. COVALENT BINDING TO CELLULAR MACROMOLECULES









4.6.1. Proteins and Lipids.  Reactive intermediates of the metabolism of chloro-




form  (phosgene,  carbene,  *C1) and carbon tetrachloride  (*CC1   phosgene,  car-




bene,  »C1) that irreversibly bind to cellular macromolecules (covalent binding)




are generally  believed  to result  in  an  alteration of cellular integrity, which




leads  to  centrolobular  hepatic  necrosis and renal proximal  tubular epithelial




damage.  Chloroform, mole for mole,  is generally accepted to be less hepatotoxic




than  carbon  tetrachloride  (Brown,  1972; Klaassen and Plaa,  1969;  Plaa et al.,




1958).




      Chloroform in non-lethal  doses produces renal damage in mice, dogs, and man;




(Bhoosan et al.,  1977; Pohl  et al.,  1977; Ilett  et al., 1973; Klaassen and Plaa,




1966,  1967;  Bennet  and  Whigham,  1964;  von Oettingen,  1964;  Conlon, 1963;  Plaa




et  al., 1958; Culliford  and Hewitt, 1957; Hewitt, 1956; Shubik and Ritchie, 1953)




whereas, in  experimental animals, carbon tetrachloride  does  not  do so (Storms,




1973;  Klaassen and Plaa, 1966; Plaa and Larson,  1965; Bennet and Whigham, 1964;




Culliford and Hewitt, 1957;  Hewitt,  1956; Shubik and Ritchie,  1953), although it




does  in man  (New  et  al.,  1962; Guild  et al.,  1958).   To explain  these species




differences  in toxicity  as  well  as known  intraspecies  (Hill  et al.,  1975;




Deringer  et  al.,  1953;  Shubik and Ritchie,  1953) and sex  differences (Taylor




et  al., 1974; Ilett et al.,  1973; Bennet and Whigham,  1964; Culliford and Hewitt,




1957;  Hewitt,   1956;   Deringer   et  al.,   1953;  Shubik  and  Ritchie,   1953;




Eschenbrenner and Miller, 1945b),  prevailing concepts implicate (1) differences
                                      4-47

-------
in the rater, of metabolism  and  organ system  capacities  for metabolism, which in


turn  determines  the  amount of  irreversible  macromolecular  binding  and  (2)


differences in the enzyme pathways for metabolism of the two haloalkanes  (Figures


4-6 and 4-7).


     Carbon tetrachloride,  by  a reductive  dechlorination  via complexing with


reduced P^,  yields  the trichloromethyl free radical  (-CCl^)   (Recknagel and


Glende, 1973; Slater,  1972)  (Figure 4-7), which can covalently bind to lipid and


protein (Shah et  al.,  1979; Villarruel et al.,  1975; Castro and Diaz Gomez,  1972;


Reynolds,  1967), and  can also initiate  peroxidation  of polyenoic  fatty  acids


(Slater, 1972; Recknagel and Ghoshal, 1966).  Chloroform does not  appear  to  be


activated to free radicals («CC1  or  »CHC1 ),  but  does bind  covalently to  liver


lipid  and  protein  (Sipes et al.,  1977;  Docks and Krishna,  1976;  Uehleke  and


Werner, 1975; Brown et al., 1974; Ilett  et al., 1973), and initiates lipid peri-


oxidation  in  some  circumstances  (Koch et al.,  1974;  Ilett,  1973;  Brown,  1972;


Slater, 1972).  Several investigators  have shown that diene conjugates (products


of lipoperoxidation) are not increased .in vivo in normal rats when chloroform is


inhaled or  injected intraperitoneally (Brown et al.,  1974; Brown,  1972; Klaassen


and  Plaa,  1969), but  only  when rats  are pretreated with  phenobarbital  and  the


metabolism  of  chloroform is greatly enhanced (Brown et al.,  1974; Brown, 1972).


Studies iri  vitro show diene conjugation  and  malonaldehyde formation  (an index of


lipoperoxidation)  by  microsomes  of  phenobarbital  pretreated  rats  were  not


increased  but  decreased by the  addition  of chloroform  (Brown, 1972;  Klaassen and


Plaa,  1969),  suggesting  that with isolated  microsomes, metabolism  of chloroform


is  too small  for sufficient quantities  of reactive intermediates to accumulate


and  initiate  lipoperoxidation.   Howeve~,  Rubinstein and Kanics  (1964)  found


chloroform to  be more rapidly metabolized by rat microsomal fractions than carbon

                  AjM
tetrachloride (see^Table  4-11).   These findings indicate  that differences in
                                      4-48

-------
                                    TABLE 4-12.




             Covalent Binding of Radioactivity From   C-Chloroform and

            14                                                        i
              C-Carbon Tetrachloride in Microsomal Incubation In Vitro

Incubation Microsomal
Condition Protein Lipid
nmol/mg in
1 C-CC14 N2 20.0 76.0
1J4C-CHC13 N2 5.1 4.1
02 8.5 7.0
To Added
serum albumin
60 minutes
1.4
0.9
1.7
 Source:   Uehleke et al.,  1977



T^licrosomes from phenobarbital pretreated rabbits
                                          4-49

-------
metabolic  activation  [carbon tetrachloride  to produce  free radicals  (Figure



4-7),  but  chloroform  primarily to phosgene  (Figure 4-6)] explain the  greater



potential  of carbon tetrachloride for initiating  lipoperoxidation.   Table  4-11


shows the data of Uehleke et  al. (Uehleke et al.,  1977;  Uehleke and Werner, 1975)



for  the  covalent binding of rabbit  microsomes following incubation  with  14C-



labeled chloroform and carbon tetrachloride.  Both protein and lipid  binding of

14
  C-radioactivity  are  4-fold  and  20-fold,  respectively, more  extensive  for



carbon  tetrachloride  than  chloroform;   lipids  are  labeled  preferentially  by



carbon tetrachloride but are not by  chloroform.   Furthermore,  covalent  binding


from  chloroform  metabolism  occurs  mainly with anaerobic conditions (a  minor



metabolic  pathway)  and  is  not greatly  increased  with aerobic metabolism,  the


major pathway for metabolism of chloroform, which  is 0  dependent (Figure 4-6).



     Covalent binding occurs preferentially to lipids and proteins  of the endo-


plasmic reticulum proximate to P^    system  for metabolism.   However,  consider-


able covalent binding  from chloroform metabolites occurs in other cell fractions



of liver and kidney, particularly to mitochondria  (Uehleke and Werner,  1975; Hill


et al., 1975).   Hill  et al. (1975)  found when C57BL  male  mice were injected



interperitoneally with 0.07 raJl/kg    C-chloroform  in oil and sacrificed 1? hours


later, that  in the  liver,  50% of the radioactivity was  irreversibly bound  to


rnicrosome, 23% to mitochondria, 25%  to  cytosol, and <2%  to nuclei;  for  kidney,


38% of radioactivity was bound to microsome,  39% to mitochondria, 22% to cytosol,


and <2% to nuclei.   A similar distribution was found in male NMRI mice by  Uehleke



and Werner (1975), who observed minimal binding to microsomal RNA but significant


binding to nicotine-adenine  nucleotides.  The  data of Ilett et al. (1973), shown


in Figure  4-9, demonstrates that in  C57 BL/6 mice, the amount of covalent  binding


in  liver  and  kidney  microsomal   fractions increases  proportionally with  the


chloroform dose.
                                     4-50

-------
   c
   1
   Q.
   U>
   "o
   E
   c
   O
   z
   o
   z
   ffi
  o
  o
                       23456
                     CHLOROFORM DOSE, mmol/kg
Figure  4-9.     Effect  of   increasing   dosage   of   i.p.-injected
14
  C-chloroform on extent of covalent binding of radioactivity  in
vivo  to  liver and  kidney  proteins  of male mice  6 hours  after
administration.

Source:  Ilett et al. (1973).
                         4-51

-------
     4.6.1.1. GENETIC STRAIN DIFFERENCE — Hill  et  al.  (1975)  described  in mice




two genetic variations  in chloroform toxicity paralleling genetic differences  in




covalent binding in  liver  and  kidney.   In one inbred strain  (DBA/2), the male




animals were  4  times more  sensitive  to  the  lethal  effects  of oral doses  of




chloroform  (LD^  of  0.08  m£/kg)  than  the second strain (C57  EL/6,  LD^  of




0.33 mK,/kg).  Males of the  F.) hybrid strain (B6D2F1/J)  had an  intermediate LD™




of 0.2 mjl/kg, midway between those of  the  two parental  strains.  The suscepti-




bility of  DBA mice was related to  a  dose-dependent necrosis  of the proximal




convoluted renal tubules.   However,  mice  of all three  genotypes  that received




>0.17 mjl/kg chloroform  exhibited both renal tubular necrosis and hepatic  centro-




lobular necrosis.  Males and females of the same strain exhibited similar  dose




thresholds to hepatic  damage,  but females  died  of chloroform-induced  hepatic




damage without developing renal lesions.   This  sex-related absolute  difference




is  dependent  on androgen  profile  of  the  mice;  testosterone-treated  females




become sensitive  to renal   toxicity  (Bennet and Whigham,  1964;  Culliford  and




Hewitt, 1957; Eschenbrenner and Miller, 1945b).




     Table 4-12 shows the extent of covalent binding in  liver and kidney of these




three  strains  after  a  single  intraperitoneal  injection  of    C-chloroform




(0.07 m£/kg)  to  the males.   Kidney  homogenates  from  DBA/2J male  mice,  more




sensitive to renal  necrosis,  contained  more than 2-fold as much radioactivity as




those from resistant C57BL/6J;  covalent  binding in the F   hybrid was  interme-




diate, as  expected.   A significant  difference  was also  noted in labeling  of




kidney subcellular  fractions.   While all subcellular  fractions of susceptible




male DBA mice were labeled to a greater  extent than  F1 or C57BL  strains,  the




greatest increase was in labeling  of the mitochondrial  fraction.
                                      4-52

-------
                                    TABLE 4-13

           Mouse Strain Difference in Covalent Binding of Radioactivity
                              From    C-Chloroform
                                                  a,b
Tissue
  Liver

  Kidney
  Liver
     Nuclei
     Mitochondria
     Microsoraes
     Cell sap
Specific Activity
Relative to C57 BL
DBA
0.82
2.41
0.67
1.14
0.64
0.98
F1
Tissue homgenates
0.96
1.64
Subcellular fractions
0.76
1.14
0.73
1.09
C57BL
1.00
1.00
1.00
1.00
1 .00
1.00
Kidney
Nuclei
Mitochondria
Microsoraes
Cell sap

2.20
3.67
1.74
1.65

1.67
1.97
1.44
1.23

1.00
1.00
1.00
1.00
 Source:  Hill et al.,  1975
                                        14,
 'Adult male mice of each genotype given  ' C-chloroform (0.07 mJl/kg) intra-
 peritoneally and sacrified 12 hours later.  Genotype comparisons are given as
 ratio of radioactivity to C57BL = 1.
                                         4-53

-------
     In the liver, the distribution of covalent  binding was generally opposite to




that observed in the  kidneys  (Table 4-12),  but  neither liver homogenates  nor




subcellular fractions  showed significant  strain differences.








     4.6.1.2. SEX DIFFERENCE -- Kidneys of male  mice are known to covalently bind




more   C-chloroform radioactivity than do those of females, but females bind more




in the liver than males (Taylor et al.,  1974;  Ilett et al., 1973) (see Tables 4-7




and 4-13).   Table  4-14 shows  that  pretreatment of male mice  with phenobarbital




increases  covalent  binding  in the liver but not  in  the kidney  (Ilett  et al. ,




1973).  A similar observation  has been made by Kluwe et al.  (1978) in male mice.




They  found that  phenobarbital increased liver but not  kidney  microsomal  acti-




vity;  3-methylcholanthrene,  dioxin, and PCGs  increased both liver  and kidney




microsomal  enzyme activities.   From the renal and hepatic toxicity  profile to




chloroform displayed by mice treated with these various  inducers, these investi-




gators  concluded that  the chloroform metabolite(s)  responsible  for  hepatic




damage is  probably generated in the liver,  and  the metabolite(s) responsible for




renal  damage is generated in  the kidney.








      4.6.1.3.  INTER-SPECIES DIFFERENCE — In  addition to  intra-species  strain




 (mice)  differences  in covalent  binding noted  above,  Uehleke and Werner  (1975)




have  also  observed an apparent inter-species  difference. Figure 4-10 shows  the in




 vitro  binding of radioactivity  from   C-chloroform  by microsomal preparations




 from  rat,  mouse, rabbit, and  man.   Human and rabbit microsomes have  the highest




 rate  of covalent binding from  chloroform, with  the mouse followed  by  the  rat




 considerably lower.   Inter-species differences in the covalent  binding rates  for




 carbon tetrachloride were small. These species differences in binding of chloro-




 form  metabolites to  protein  and lipid in vitro  do  not, however, parallel  the
                                      4-54

-------
                              TABLE 4-14

       In Vivo Covalent Binding  of Radioactivity From   CHClo in

         Liver and Kidney of Male and Female Mice (C57BL/6)a'b
                              Covalent Binding of   C-Chloroform
                                   nmoles/mg protein + S.E.

Liver
Kidney
Male
2.92 + 0.35
2.34 + 0.16
Female
3.66 + 0.39
0.39 + 0.02
aSource:   Ilett et al.,  1973

 Mice were sacrificed 6  hours after intraperitoneal administration
of 3.72 nmoles/kg of   CHC1.,.
                                    4-55

-------
                                  TABLE 4-15"


           In Vitro  Covalent Binding of Radioactivity from   CHC1  to
                                                                3
          Microsomal Protein from Liver and Kidney of Male  and Female

                                Mice  (C57BL/6)*

Male
Male
Female
Pretreatment

NA
Phenobarbital
NA
Covalent Binding of
p moles/mg protein/5
Liver
572 ± 54
1454 + 143
419 + 20
C-Chloroform
minutes + SEM
Kidney
44.6 + 4.1
41.0 + 3.2
14.6 + 2.5
"Source:   Ilett et al.,  1973.


NA = Not  applicable
                                        4-56

-------
0)   T
cr  o
cr  3
                H-
                W
                C
                T
                CD
                                                             RADIOACTIVITY, nmol 14C from 14CHCI3 /mg
o
c
O
CD
CD
3"
CD
?r
CD
3
Q.
0)
T
3
CD
        03
        cf
O
c
w
CD
3
O.
        3
        H-
        3
        O
        C
        cr
        CD
        Q.
       IS'
       H-
       cf
       fu
       cf
       UJ
       -J
            cf
            o
            •D
            CD
            H-
    3
    Q.
           •o
            H-
            CL
    3
    H-
    O

    O
    00
    O

    CD
    GO
    O


    3
    O
    T

    &>
 O
 O

TO
 W
 O
 3

 O
 "3
         CD  ^


         |  I
         00  "-,
         H-  =
         cr  5'
cr
H-

a

OQ
tu
Q-
I-"
o
03
o
cf
H-
<
H-
cf
               IS3
               O
               U)
                       CJ!
                       O

-------
species differences  in  metabolism of chloroform  in vivo,   as  measured by  the



conversion of chloroform to  CO   (Table 4-9);  in vivo, the mouse has the greatest
                             ^                .1..  i  i


capacity to metabolize chloroform (Q0%) , followed  by rats  (65%),  nonhuman  pri-


mates (20%), and man (30 to  50%).







     4.6.1.4. AGE DIFFERENCE -- Uehleke  and Werner (1975) have  shown that  irre-



versible protein  binding of radioactivity  from   C-chloroform and    C-carbon



tetrachloride to liver microsomes of  newborn rats  (18 hours old) is low compared



to that of microsomes from adult rats (32 days); however, the binding was  shown



to be proportional to PH™  content of the  microsomes,  which was  proportionally



low in microsomes from newborn rats.







4.6.2. Nucleic Acids. PJICQ  systems activate chloroform and  carbon tetrachloride



in vivo and in_ vitro  to reactive metabolites that extensively covalently bind to



proteins and lipids, but do so  only  minimally  to nucleic acid  (Uehleke et  al.,



1977; Wolf  et al., 1977; Uehleke and Werner,  1975; Fowler, 1969;  Reynolds, 1967),



unlike many other carcinogens that bind  DNA.   Reitz et al. (1980) measured DMA



alkylation  in liver  and kidneys  of mice  after  an  oral  dose   of  240  mg/kg


14                                                                    _4
  C-chloroform (specific activity not given) and found  values of 3 x 10   and  1 x


  -4
10   mol %  for liver and kidney  DNA respectively.   These  workers  judged  that



chloroform has very  little  direct interaction with DNA when compared to  known



carcinogens, as reported in the  literature  for dimethylnitrosamine (3.5 x  10~


                                                             p

mol % alkylation, liver  DNA),  dimethylhydralazine  (2.6 x  10  ,  colon DNA) and



N-methyl-N-nitrosourea (1.5 x  10,  brain  DNA) but given  by parenteral routes



(Pegg and  Hui,  1978; Cooper et  al.,  1978;  Kleihues and Margison,  1974).   The



failure of chloroform or carbon tetrachloride reactive  species  to significantly



bind DNA has been ascribed to  their short half-life compared  to epoxides, and to
                                      4-58

-------
their lack  of  nuclear penetration.   Recently,  however, Diaz Gomez  and Castro



(1980) have shown that highly purified rat liver nuclear preparations  are able to



anaerobically  activate  carbon  tetrachloride,  and  to  aerobically  activate



chloroform  to  reactive metabolites that  bind to nuclear  lipids  and proteins.




Their data, given in  Table 4-15, show that  activity in nuclear  preparations is



smaller  than  in  microsomes,  but within  the same  order  of magnitude.   These



results might  be  relevant  to the hepatocarcinogenie  effects  of chloroform and



carbon  tetrachloride  in mice  and  rats,  since  the nuclear targets  (DMA,  RNA,



nuclear proteins) are  in the  immediate  neighborhood sites  of  activation,  thus,



making unnecessary the present assumption  that the highly reactive intermediates



(•CCl-j,  phosgene, malonaldehyde,  or  carbene),   produced   at  the  endoplasmic



reticulum, must travel to the nucleus.







4.6.3- Role of Phosgene.  Phosgene is a prominent  intermediate of both chloroform



and  carbon  tetrachloride metabolisms  (Figures 4-6,  4-7).   It  is known to  be



highly reactive and toxic to cells and tissues (Pawlowsky and  Frosolono,  1977),



and  its  two highly reactive  chlorines  suggest  that  it  could  act on  cellular



macromolecules  similar to  bifunctional  alkylating agents.    Reynolds  (1967)


            14
showed that   C-phosgene, given to intact rats, labeled liver protein (and lipids



to a smaller extent).   The  pattern of  labeling was quite different from that of



  C-carbon tetrachloride and  more  similar to   C-chloroform.    Moreover,  ^ Cl-



ear bon tetrachloride radioactivity was also  stably incorporated into liver lipid



and protein, pointing  to the  •CC1,,  radical rather than phosgene  as the  reactive



form  for  carbon  tetrachloride  that   labels  lipid.    Cessi  et al. (1966)  also


               14
reported that    C-phosgene  labeled terminal  amino  group  of  polypetides in  a



manner similar to in vivo protein labeling produced  by carbon  tetrachloride.
                                     4-'5 9

-------
                                    TABLE 4-16

       Covalent Binding of Radioactivity from   C-Chloroform and    C-Carbon
      Tetrachloride in Rat Liver Nuclear and Microsomal Incubation In Vitro*
                            Incubation
                            Condition
Protein
Lipid
      p mol/mg + S.D.
  C-CC1,
     Nuclear
     Microsomal
21.9 + 2.5
50.3 + 4
14? + 12
190 +11
  C-CHC1,
     Nuclear
     Microsomal
27.0 + 3
68.0 + 9
 20+3
 57+8
*Source:   Diaz Gomez and Castro,  1980
                                       4-60

-------
4.6.4. Role of Glutathione.  Ekstrom and Hogberg  (1980) found that chloroform, in


freshly  isolated  rat liver  cells,  induced depletion of  cellular  glutathione.


Brown et al. (1974) demonstrated that exposure of rats to an atmosphere of 0.5%


chloroform for 2 hours markedly decreased glutathione (GSH) in the liver when the


animals were pretreated with  phenobarbital  to stimulate  metabolism.   GSH liver


content of untreated rats was  not decreased.  Phenobarbital pretreatment has been


shown to  markedly potentiate  toxicity  of  both  chloroform  and carbon   tetra-


chloride in rats  (Docks and  Krishna,  1976;  Cornish et al.,  1973;  McLean, 1970;


Scholler, 1970).  However,  it has not been possible to detect a decrease in the


liver glutathione levels  following  administration of carbon  tetrachloride  or


trichlorobromomethane (Docks  and Krishna,  1976; Boyland  and  Chasseaud,  1970).


Sipes et al.  (1977)  have  shown that the addition  of  GSH liver microsomes from

                                                       ill
phenobarbital pretreated rats  incubated in  vitro with    C-labeled  halocarbons,


chloroform, carbon  tetrachloride,  and  trichloromomethane, inhibited  covalent


binding =80$ for all three  compounds.  Their results,  given in Table  4-16, also


show the effects  of  anaerobic and aerobic  conditions  on  covalent  binding.  The


reduction in  binding of chloroform  by  an  atmosphere of  N?  suggests that  its


bioactivation  is  mediated  by a cytochrome  P.    oxidative pathway to  phosgene


(Figure 4-6),  while the  enhanced  binding of carbon  tetrachloride in N~  reflects


Pj,j50 mediated  reductive  pathways  (Figure  4-7)  and formation of free  radical.


These investigators  suggest  that  in phenobarbital-treated animals,  chloroform


depletes  liver  GSH  by  the  formation  of  conjugate   between  the  reactive


intermediate phosgene and GSH (Docks and Krishna, 1976; Brown et al.,  1974).   In


the case of carbon tetrachloride, they suggest that GSH addition in vitro (Table


4-16) also  conjugates with  the  phosgene  metabolite of  carbon tetrachloride


produced when incubated in air,  but,  in addition,  GSH decreases the levels  of


•CC1, by reducing the free radical to  chloroform.  In vivo, it is postulated that
                                     4-61

-------
                                    TABLE 4-17
            Effect of Glutathione, Air, N2 or CO:  02 Atmospere on the



           IH Vitro Covalent Binding of C  Cl^,  CHC1,,  and C Br Cl~  to Rat
                             Liver Microsomal Protein
                                                 Substrate
Incubation Conditions
 CC1,
CHC1-
CBrCl.
Air





N2



C0:02(8:2)




SKF 525 A (0.5 mM), Air




Glutathione, Air




NADPH omitted, Air
         ill

 p moles   C-bound/mg microsomal protein/minute




 97+10             59+5              1456+66




310 +51             21+1              1370 + 143




 18+1             20+1               853 +  62




109+5              7+7              2105 + 159




 17+2             15+2               218 +  25




  6+1              3+0                65+13
 Source:  Sipes et al.,  1977



bl4                                                      -^
   C-labeled substrate is a final concentration of 1 x 10 JM incubated at 37°C

 Microsomes were from phenobarbital pretreated rats.
                                       4-62

-------
the  oxidized  glutathione may  be  reduced  back  to  reduced  GSH by  glutatione




reductase; this would explain the lack of fall of liver GSH content with carbon




tetrachloride .in vivo (Gillette, 1972).  Thus,  the  toxic  effects of chloroform




and  carbon  tetrachloride  may  be  mediated  through  different mechanisms  of




covalent binding, and GSH may  play different roles for these  chlorocarbons  in




preventing covalent binding of reactive intermediates of metabolism.




     Chloroform and carbon tetrachloride  are  known to cause greater liver damage




in fasted animals than in fed animals  (Diaz  Gomez  et al.,  1975; Jaeger et al.,




1975; Krishnan and Stenger,   1966;  Goldschmidt  et al.,  1939;  Davis  and Whipple,




1919).  For chloroform,  a decreased content of  hepatic GSH from  fasting has been




postulated to  be  responsible for  the  increased susceptibility  of  fasted mice




(Docks and Krishna,  1976; Brown et al.,  1974).  Nakajima and  Sato (1979) have




recently offered an  additional  explanation.    These investigators studied  the




metabolism of the chlorocarbons In vitro with  microsomes  from  livers of fasted




rats, and found that the disappearance of  chloroform from incubation increased  3-




fold for a 24 hour fast, although fasting produced no significant increase in the




microsomal protein and  cytochrome  Pnc-n  liver  contents  (Table   4-17).   Similar




results were obtained for carbon tetrachloride.   These observations suggest that




the  increased  toxicity  of  chloroform  and  carbon  tetrachloride  from  food




deprivation may  be  due not  only  to  decreased GSH,  but  also  to  a  greater




production of reactive  intermediates and covalent  binding  to   cellular  macro-




molecules .









SUMMARY




     At ambient temperatures, chloroform is  a  volatile liquid  with  high  lipid




solubility and appreciable solubility  in water.   Hence, chloroform  is  readily




absorbed into the body through the lungs and intestinal mucosa;  the  portals  of
                                     4-63

-------
                                                    TABLE 4-1$

                    Effects of 24-Hour Food Deprivation on Chloroform and Carbon Tetrachloride
                     .In Vitro  Microsomal Metabolism, Protein, and P-450 Liver Contents of Rats*


Male
Female
Fed Fasted Ratio Fed
Metabolism, nmole/g/min
Chloroform
Carbon
tetrachloride
19.7 + 2.6 55.1 + 7.5 2.8
1.9 + 0.2 5.9 + 0.8 3.1
Protein content
27.7 + 3.7 23.0 + 2.7 NR
P-450, nmol/mg
0.842+0.123 0.823+0.03 NR
15.3 + 6.8
1.1 + 0.5
, mg/kg liver
22.5 + 1.5
protein
0.638 + 0.051
Fasted Ratio
39.3 + 2.5 2.6
4.5 + 0.3 4.1
23.7+1.7 NR
0.673 + 0.044 NR
*Source:   Nakajiraa and Sato,  1979

NR = Not  reported

-------
entry with exposure  from air, water and food.   Few data are available  on the
pharmacokinetics of absorption and excretion of chloroform in man, particularly
at the low exposure  concentrations expected in  ambient  air  and  drinking water.
However,  studies  show   absorption  from  the  gastrointestinal   tract  in  man,
monkeys, rats and mice is rapid  and  complete,  occurring by  first-order passive
absorptive  processes.    A   dose-dependent  first-pass   effect  with  pulmonary
elimination  of  unchanged chloroform occurs  with oral  ingestion in man,  thus
decreasing the amount of  chloroform  reaching the systemic circulation.  In rats,
the kinetics  of  peroral absorption are also influenced by the dosing vehicle; the
absorption rate is decreased for  chloroform given in corn oil vehicle as compared
to an aqueous solution.   Pulmonary uptake and  elimination occur  also  by first-
order diffusion processes  with  three  distinct  components with  rate  constants
corresponding to  tissue  loading  or  desaturation of  at  least three major  body
compartments.   Half-times  in man have  been found  to  be approximately  14-30
minutes, 90  minutes  and  24-36 hours, respectively.   The longest half-time  is
associated with the lipids and the adipose tissue compartment.  During inhalation
exposure,  at  equilibrium  with  inspired   air  concentration,   the  blood/air
partition coefficient is  about 8 at  37°C and  the adipose tissue/blood  partition
coefficient is 280 at 37°C.    The quantity  of chloroform absorbed is  dependent
also on body weight and  fat  content  of  the  body.
     Tissue distribution  of  chloroform  is consistent with its  lipophilic  nature
and modest water solubility.  This chloroalkane readily  crosses  the  blood brain
and  placental   barriers  and  distributes into  breast  milk.    Concentrations
occurring  in  all major tissue organs  are  dose  related  to  inspired  air
concentrations or to  oral dosage.  Relative tissue  concentrations occur  in the
order of adipose tissue  > brain > liver  > kidney >  blood.
                                     4-65

-------
     Elimination of chloroform from  the  body occurs by two major  and  parallel




occurring processes:  1) pulmonary elimination of unchanged  chloroform  by  first




order kinetics, and 2) metabolism of chloroform.   Chloroform  is  metabolized  in



the liver, and to a lesser extent in the  kidneys and other tissues.  Metabolism




is dose-dependent and saturable,  with a greater proportion of  small doses  being




metabolized.   There  are  striking  differences  in  the  pharaacokinetics  and




quantitative metabolism of chloroform in man as compared to other animals.  For




large steady-state  body  burdens,  30-40/& is  metabolized by  man,  20% by  the




nonhuman primate, > 65% by the rat,  and > 85%  by the mouse.  Metabolism produces




phosgene  and   other   putative   reactive  metabolites  that   covalently   bind




extensively to cellular lipids and proteins, although not significantly  to DNA or




other nucleic acids.  The intensity  of metabolite  binding and organ localization




parallel the acute cellular toxicity of chloroform in liver and kidney observed




in experimental  animals.  Both  binding  and  toxicity  are highly  dependent  on




animal  species  and genetic strain, as well  as on sex and age.   An additional




variable is the tissue level of reduced glutathione which plays an important role




in protecting against both binding and toxicity.  Conversely, inducers of hepatic




and renal P450 metabolizing systems increase  binding and toxicity.
                                      4-66

-------
                                4.S REFERENCES







Ahmed, A.E.,  V.L.  Kubic and M.W.  Anders.   1977.   Metabolism of haloforms  to



carbon monoxide.  I. In vitro studies.   Drug Metab.  Dispos.   5: 198-204.







Anders, M.W.,  J.L. Stevens,  R.W.  Sprague, Z.  Shaath and  A.E.  Ahmed.    1978.



Metabolism of haloforms to carbon  monoxide.  II.  In vivo studies.   Drug Metab.



Dispos.  6: 556-560.







Bellar, T.A.,  J.J.  Lichtenburg and R.C.  Kroner.   1974. Occurrence  of  organo-



halides  in  chlorinated drinking  waters.    Amer.  Water   Works  Assoc.  Jour.



66: 703-706.







Bennet, R.A.  and  A.  Whigham.   1964.   Chloroform sensitivity  of  mice.   Nature.



204: 1328.







Bhooshan,  B.,  L.R. Pohl and  G.  Krishna.   1977.   A  biochemical  mechanism for



chloroform activation.  Fed. Proc. (Abstr).  36: 396.







Boyland, E. and L.F. Chasseaud.  1970.   The effect of  some carbonyl compounds on



rat liver glutathione levels.  Biochem. Pharmacol.   19: 1526-1528.







Brown,  D.M.,  P.F. Langley, D.  Smith  and  D.C.  Taylor.   I974a.   Metabolism of


                                       1 4
chloroform.   T.   The   metabolism  of    C-chloroform  by   different  species.




Xenobiotica.  4: 1 51-163-
                                      4-67

-------
Brown, B.R., K.G. Sipes and A.M. Sagalyn.  I974b.   Mechanisms  of acute hepatic




toxicity:  Chloroform, halothane and glutathione.  Anesthesiol.  41 :  554-561 .








Brown, S.S.  1972.  In Foreign compound metabolism in mammals,   vol. 2.  Mathway,




D.C., ed.  London, The Chemical Society,  p.  461.








Butler, T.C.  1961.   The  reduction of  carbon tetrachloride in vivo and reduction




of  carbon tetrachloride  to  chloroform  in  vitro  by  tissues  and  tissue cons-




tituents.  Jour. Pharmacol. Exp. Ther.  134:  311-319-








Castro,  J.A.  and M.I. Diaz  Gomez.   1972.  Studies on  irreversible  binding of




  C-CHC1_  to microsomal  lipids in rats  under  varying experimental conditions.




Toxicol.  Appl. Pharmacol.  23:  541-552.








Cessi,  C.,  C. Columbini and L. Mameli.   1966.   The reaction of liver proteins




with  a  metabolite of  carbon  tetrachloride.  Biochem.  Jour.  101: 46c.








Chenoweth,  M.B., D.N. Robertson, D.S.  Erly and  K.  Goekhe.   1962.   Blood and




tissue  levels  of ether,  chloroform,  halothane  and  methoxyflurane  in dogs.




Anesthesiol.   23: 101-106.








Chiou,  W.L.   1975.  Quantitation of hepatic and pulmonary first-pass  effects and




its implications in pharmacokinetic  study. I.  Pharmacokinetics of chloroform  in




man.   J.  Pharmac. Biopharmaceut.  3: 193-201.
                                      4-68

-------
Cohen,  E.N. and W. Hood.   1969.  Application of low-temperature autoradiography




studies of the uptake  and metabolism  of volatile  anesthetics in  the  mouse.




Anesthesiol.  30:  306-314.








Conkle, J.P.,  B.J.  Camp  and  B.E.  Welch.   1975.   Trace composition  of human




respiratory gas.  Arch.  Environm. Health.  30:  290-295.








Conlon M.F.  1963.  Addiction to chlorodyne.   Brit.  Med. Jour.   2:  1177-1178.








Cooper, H.K.,  J. Buecheler and P. Kleihues.   1978.   DNA alkylation  in mice with




genetically different  susceptibility  to  1 ,2-dimethylhydrazine-induced  colon




carcinogenesis.  Cancer  Res.   38: 3063-3065.








Cornish, H.H.,  B.P.  Ling  and M.L. Barth.   1973.   Phenobarbital  and  organic




solvent toxicity.   Amer.  Ind.  Hyg.   34:  487-492.








Culliford,  D.  and  H.B. Hewitt.  1957.  The influence  of sex hormone status on the




susceptibility of mice to chloroform-induced necrosis  of renal  tubules.   Jour.




Endocr.  14: 381-393.








Davis,  N.  and  C. Whipple.   1919.   The  influence of  fasting  and  various diets on




the  liver  injury  effected  by  chloroform   anesthesia.    Arch.   Int.   Med.




23: 612-633.








Deringer,  M.K., T.B.  Dunn  and  W.E.  Heston.  1953.  Results of strain C3H mice to




chloroform.  Proc. Soc.  Exp.  Biol.  Med.   83:  474-479.
                                     4-69

-------
Diaz Gomez, M.I. and J.A. Castro.   1980.   Nuclear activation of  carbon  tetra-




chloride on chloroform.   Res.  Comm.  Chem.  Path.   Pharmacol.   27:  191-194.









Diaz Gomez, M.I.,  C.R. DeCastro, E.G. DeFerreyra, N. D'Acosta, O.M. DeFerros and




J.A. Castro.  1975.  Mechanistic studies on carbon tetrachloride  hepatotoxicity




in fasted and fed rats.   Toxicol.  Appl. Pharmacol.  32: 101-^08.









Dilley, J.V., N. Chernoff, D.  Kay, N. Winslow and G.W.  Newell.  1977.  Inhalation




teratology  studies of  five  chemicals in  rats.   Toxicol.  Appl.  Pharmacol.




41 :  196.









Docks, E.L. and G.  Krishna.  1976.  The role of glutathione on chloroform-induced




hepatoxicity.  Exp. Mol. Pathol.  24: 13-22.









Dowty, B.J., D.R. Carlisle and L.L.  Laseter.  1975.  New Orleans drinking water




sources tested  by  gas chromatography-mass  spectrometry.   Occurrence and origin




of aromatics and halogenated  aliphatic hydrocarbons.   Environm.   Sci. Technol.




9: 762-765-








Droz, P.O. and J.G. Fernandez.  1977.  Effect of physical workload on retention




and metabolism of  inhaled organic solvents.  A comparative theoretical approach




and  its application with  regards to  exposure  monitoring.    Int.  Arch.  Occup.




Environ. Hlth.  38: 231-246.









Eger, E.I.   1963-  Mathematical model  of uptake and distribution.  In:  Papper,




E.M.    and  Kitz,   R.J.  eds.   Uptake  and  Distribution  of  Anesthetic  Agents.




McGraw-Hill, New York.  pp. 72-103.
                                      4-70

-------
Ekstrom, T.  and J. Hogberg.  1980.  Chloroform induced glutathione depletion and




toxicity in freshly isolated hepatocytes.  Biochem. Pharmacol. 29: 3059-3064.








Eschenbrenner, A.B. and E. Miller.   1945a.   Induction of hepatomas  in  mice by




repeated oral administration of  chloroform with observations on sex differences.




Jour. Natl.  Cancer Inst.  5: 251-255.








Eschenbrenner, A.B.  and E.  Miller.  1945b.  Sex differences in kidney morphology




and chloroform necrosis.  Science.   102:  302-303.








Feingold, A.  and  D.A.  Holaday.   1977.   The pharmacokinetics of  metabolism of




inhalation anaesthetics.  Brit.  Jour. Anaesth.   49: 155-162.








Fiserova-Bergerova,   V.,  J.   Vlach   and  J.C.  Cassady.    1980.    Predictable




"individual  differences"  in uptake  and  excretion of  gases  and  lipid  soluble




vapours simulations study.   Br.  J.  Ind.  Med.  37:  42-49.








Fiserova-Bergerova,  V.  and D.A.  Holaday.   1979.   Uptake and  clearance of inhal-




ation anesthetics in man.   Drug. Metab.  Rev.   9:  43-60.








Fiserova-Bergerova,  V., J.  Vlach and  K.  Singhal.   1974.  Simulation and  predic-




tion of uptake, distribution, and exhalation of  organic  solvents.   Br.  J.  Ind.




Med.  31 : 45-52.








Fishbein, L.  1979.  Potential halogenated industrial carcinogenic and mutagenic




chemicals.  II.    Halogenated  saturated  hydrocarbons.   Sci.  Total  Environm.




11 : 153.
                                     4-71

-------
Fowler, S.J.L.   1969.   Carbon tetrachloride metabolism  in  the rabbit.   Brit.




Jour.  Pharmacol.  37:  733-737.








Fry, J., T. Taylor and  D.F. Hathaway.  1972.  Pulmonary elimination of chloroform




and its metabolite in man.  Arch. Int. Pharmacodyn.  196: 98-111.








Geigy  Scientific Tables.   1973.    Diem,  K.  and C.  Lentner, eds.    7th  ed.




Ciba-Ceigy, Basle.








Gettler, A.O.  1934.  Medicolegal aspects of deaths associated with chloroform or




ether.  Araer. Jour. Surg.  26: 168-17**.








Gettler,  A.O.  and H.  Blume.  1931.   Chloroform in the  brain,  lungs and liver.




Quantitative recovery  and determination.   Arch. Pathol.  11: 554-560.








Gillette,  J.R.  1972.   Pharmacology and the future of man.  In:  Proceeding of the




 5th Intern.  Cong. Pharmacol.  San Francisco, Karger, Basel.








Glende, E.A. Jr.,  A. M.  Hruszkewycz and H.O. Recknagel.   1976.   Critical role of




 lipid  peroxidation in  carbon tetrachloride-induced loss  of  aminopyridine de-




 methylase, cytochrome  P450  and glucose  6-phosphatase.   Biochem. Pharmacol.




 25: 2163-21 70.








 Goldschmidt, S.,  N.M.  Vars and I. Ravdin.  1939.  The influence of the foodstuffs




 upon the  susceptibility of the  liver  to  injury by chloroform and  the  probable




 mechanism of their action.  Jour.  Clin.  Invest.  18: 277-289.
                                      4-72

-------
Gordis,  E.   1969.   Lipid metabolites  of carbon  tetrachloride.   Jour.  Clin.




Invest.   48: 203-209.








Guild, W.R., J.V.  Young  and  J.P. Merrill.  1958.   Anuria  due  to  carbon tetra-




chloride intoxication.  Ann.  Int. Med.  48: 1221-1227.








Hewitt,  H.B.  1956.  Renal necrosis in mice after accidental exposure to chlor-




oform.  Brit. Jour. Exp.  Path.  37: 32-39.








Hill, R.N., T.L.  Clemens,  O.K. Liu, E.S. Vesell and W.D. Johnson.  1975.  Genetic




control of chloroform toxicity in mice.  Science.  190: 159-160.








Ilett, K.F., W.D. Reid, I.G.  Sipes  and G. Krishna.  1973-  Chloroform toxicity in




mice:  Correlation of renal and hepatic necrosis with covalent binding of metab-




olites to tissue macromolecules.   Exp. Mol. Pathol.  19:  215-229.








Jaeger,  R.J.,  R.B.  Connolly  and S.D. Murphy.   1975.   Short-term  inhalation




toxicity of halogenated hydrocarbons.   Effects on fasting rats.  Arch. Environm.




Health.   30: 26-30.








Jakobson,  I., J.E. Wahlberg,  B. Holmberg and G. Johansson.  1982.  Uptake via the




blood and elimination of 10 organic solvents following epicutaneous exposure to




anesthetized guinea pigs.  Toxicol. Appl.  Pharmacol.  63: 181-187.








Kimura,  E.T., D.M. Ebert  and  B.W.  Dodge.   1971.   Acute  toxicity and  limits of




solvent  residue  for sixteen  organic  solvents.    Toxicol.  Appl.  Pharmacol.




19: 699-704.
                                     4-73

-------
Klaassen, C.D. and G.L.  Plaa.   1969.   Comparison of the biochemical alterations




elicited in livers from rats treated  with carbon tetrachloride,  chloroform,  and




metahylchloroforms.  Biochem.  Pharmacol.   18:  2019-2027.








Klaassen, C.D. and  G.L.  Plaa.   1967.  Relative  effects of  various chlorinated




hydrocarbons on liver and  kidney  function in mice.  Toxicol. Appl.  Pharmacol.




10: 119-131 .








Klaassen, C.D. and  C.L.  Plaa.   1966.  Relative  effects of  various chlorinated




hydrocarbons on liver and  kidney  function in mice.  Toxicol. Appl.  Pharmacol.




9: 139-151 .








Kleihues, P.  and  G.P.  Margison.   1974.   Carcinogenicity of N-methyl-N-nitros-




urea.  Possible role of excision repair  of 0  -raethylguanine  from  DNA.   J.  Nat.




Cancer Inst.  53:  1839-1841 .








Kluwe, W.M., K.M.  McCorraack and J.B.  Hook. 1978.  Selective modification of the




renal and hepatic toxicities  of chloroform  by induction of  drug metabolizing




enzyme systems in  kidney and liver.  Jour. Pharmacol. Exp. Therap.  207:  566-573-








Koch, R.R.,  E.A. Glende and R.O. Recknagel.  1974.  Hepatotoxicity of bromotri-




chloromethane-bond   dissociation   energy  and   lipoperoxidation.     Biochem.




Pharmacol.   23: 2907-2915.








Krishnan, N. and R. Stenger.  1966.   Effects of starvation on the hepatotoxicity




of carbon tetrachloride.   Amer. Jour. Pathol.  49: 239-246.
                                      4-74

-------
Lehman, K.B.  and  Hasegawa.   1910.   Studies of  the absorption  of chlorinated




hydrocarbons in animals and humans.  Arch. Hyg.  72: 327-342.  (Ger.)








Mansuy, D., P. Beaune, T. Cresteil, M. Lange and J.P.  Lerous.   1977.  Evidence




for  phosgene  formation  during  liver  microsomal  oxidation  of  chloroform.




Biochem.  Biophys. Res. Commun.  79: 513-517.








McConnell, G., D.M. Ferguson and C.R. Pearson.   1975.  Chlorinated hydrocarbons




and the environment.  Endeavor.  34: 13-18.








McLean, A.E.   1970.   The  effects of protein  deficiency  and microsomal enzyme




induction by DDT and phenobarbitone on the active toxicity of chloroform.  Brit.




Jour. Exp. Pathol.  51: 317-321.








Miklashevskii, V.E., V.N. Tugarinova,  N.L. Rakhmanina and G.P. Yakovleva.  1966.




Toxicity of chloroform administered perorally.   Hyg. Sanit.  31: 320-322.








Morris,  L.E.V.    1951 .   Chloroform  in   blood  and  respired  atmosphere.   In:




Chloroform, A Study After 100  Years.  R.M. Waters,  ed.,  Madison, University of




Wisconsin Press,  pp. 95-119.








Munson, E.S.  1973«  Discussion:   Chloroform, halothane and regional anesthesia.




Anesth. Analg.  52: 8-9.








Murray, F.J., B.A.  Schwetz,  J.G.  McBride  and R.E.  Staples.  1979.  Toxicity of




inhaled  chloroform  in pregnant  mice and  their  offspring.    Toxicol.  Appl.




Pharmacol.  50: 515-522.
                                      4-75

-------
Nakajima T. and A. Sato.   1979.   Enhanced activity of  liver  drug-metabolizing




enzymes for  aromatic  and chlorinated hydrocarbons following  food  deprivation.




Toxicol. Appl. Pharmacol.  50:  549-556.









New, P.S., G.D.  Lubash,  L. Scherr and A.L. Rubin.  1962.   Acute renal failure




associated with  carbon  tetrachloride intoxication.   Jour.  Amer.  Med.  Assoc.




181 : 903-906.









NIOSH  (National  Institute  for  Occupational Safety).  1977b.   National Occupa-




tional  Hazard Survey.   Vol.  III.   Survey Analyses  and  Supplemental Tables.




Cincinnati, USDHEW.  pp. 2864-2866.









OSHA (U.S. Occupational  Safety and Health Administration).  1978.  Air contami-




nants.  U.S.  Code Fed. Regl., Title 29, part 1910" 1000.  pp. 612-614.









Paul,  G.G. and D.  Rubinstein.   1963.  Metabolism  of carbon tetrachloride and




chloroform by the rat.   Jour. Pharmacol. Exp. Ther.   1 41 : 141-148.









Pawlowsky, R.  and M.F. Frosolono.  1977.  Effect of phosgene  on  rat  lungs after




single  high  exposure.  Arch. Environm. Health.  32: 278-283-









Pegg,  A.E. and G. Hui.   1978.   Formation and subsequent removal  of 0 -methyl-




guanine from deoxyribonucleic acid in rat  liver and kidney  after small doses  of




dimethylnitrosamine.   Biochem. J.  173:  739-748.









Piersol,  G.M., H.J. Tiemen and L.S.  Kau.   1933.  Fatal  poisoning  following  the




ingestion of chloroform.  Med. Clin. North  Amer.  17: 587-601.
                                      4-76

-------
Plaa, G. L. and  R.E.  Larson.   1965.  Relative  nephrotoxic  properties of  chloro-




inated  methane,  ethane,  and  D24 ethylene derivatives  in  mice.   Toxicol. Appl.




Pharmacol. 7:  37-44.








Plaa, G.L., E.A.  Evans  and C.H. Hine.   1958.   Relative hepatotoxicity of seven




halogenated hydrocarbons.  Jour. Pharmacol. Exp. Ther.  123: 224-229.








Pohl , L.R.  1980.   Investigations  of the mechanism of  activation of chloroform.




Toxicol. Res.  Proj.  Director.   5: No. 8.0463.








Pohl,  L.R.,  B.  Bhooshan,  N.F.  Whitaker and  G.  Krishna.   1977.   Phosgene:   A




metabolite of chloroform.  Biochem. Biophys.  Res. Commun.  79: 684-691.








Pohl,  L.R.  and G.  Krishna.   1978.  Deuterium isotope  effect  in  bioactivation




and  hepatotoxicity of chloroform.  Life Science 23: 1067.








Poyer,  J.R., R.A.  Floyd, P.B. McCoy, E.G. Janzen  and E.R. Davis.   1978.  Spin-




trapping  of the  trichloromethyl  radical produced during enzymic NADPH oxidation




in  the  presence  of  carbon  tetrachloride  or  bromotrichloroethane.   Biochem.




Biophys.  Acta.  539:  402-409.








Recknagel, R.O. and A. K.  Ghoshal.  1966.  Quantitative  estimation of peroxidative




degeneration  of  rat  liver  microsomal  and  mitochondrial  lipids  after  carbon




tetrachloride poisoning.   Exp. Molec. Pathol.   5: 413-426.








Rechnagel, R.O. and E.A.  Glende, Jr.   1973.  Carbon tetrachloride hepatotoxicity:




An   example  of   lethal   cleavage.    CRC  Crit.   Rev.   Toxicol.    2:   263-297.
                                    4-77

-------
Reitz, R.H.,  J.F.  Quast,  W.T. Scott,  P.G.  Watanabe and  O.J.  Behring.   1980.




Pharmacokinetics and  macromolecular effects  of chloroform  in  rats and  mice:




Implications for carcinogenic  risk estimation.  Water Chlorinat.  Environ. Impact




Health Eff.  3: 983-993.









Reitz, R.H., P.J. Gehring  and  C.N. Park.  1978.  Carcinogenic risk estimation for




chloroform:    An alternative  to  EPA's  procedures.    Food  Cosmet.  Toxicol.




16: 511-514.









Reynolds,  E.S.   196?.   Liver parenchymal cell  injury.  IV.   Pattern of incor-




poraton of carbon and chlorine from carbon tetrachloride into chemical constit-




uents of liver in vivo.  Jour. Pharraacol. Exp. Ther.  155: 117-126.









Rubinstein, D. and  L.  Kanics.  1964.  The conversion of carbon tetrachloride and




chloroform  to carbon dioxide by rat  liver  homogenates.   Can.  Jour.  Biochem.




42: 1 577-1 585.









Sato,  A.  and  T. Nakajima.    1979.   A structure-activity  relationship of some




chlorinated hydrocarbons.  Arch. Environm. Health.  34: 69-75.









Scholler,  K.L.  1970.   Modification of the  effects  of chloroform  on the rat




liver.  Brit.  Jour. Anaesth.  42: 603-605.









Schroeder, H.G.  1965.  Acute and delayed chloroform poisoning.  A case report.




Brit. Jour. Anaesth.  37: 972-975.
                                      4-78

-------
Schwenkenbecher.   1904.   The  absorption capability of  the skin.   Arch.  Anat.




Physiol.  pp. 121-165.








Schwetz, B.A.,  B.K.J. Leong  and P.J. Gehring.  1974.  Embryo- and fetotoxicity of




inhaled chloroform in rats.   Toxicol.  Appl. Pharmacol.  28: 442-451.








Seppelt,  K.    1977.   Trifluromethanol,  CF OH.   Angew. Chem.  Int. Ed.  Engl.




16: 322-323.








Shah, H., S.P.  Hartman  and S.  Weinhouse.  1979.  Formation of carbonyl chloride




in  carbon  tetrachloride  metabolism  by  rat  liver  in vitro.    Cancer  Res.




39: 3942-3947.








Shubik, P. and  A. L. Ritchie.  1953.  Sensitivity of male DBA mice to the toxicity




of chloroform as a laboratory hazard.   Science.  117:  285.








Simpson, J.Y.  1847.  On  a  new  anaesthetic agent,  more efficient than sulfuric




ether.  Lancet.  2:  549-550.








Sipes,  I.G.,  G.  Krishna  and  J.R.  Gillette.   1977.   Bioactivation of  carbon




tetrachloride,  chloroform and bromotrichloromethane:  Role of cytochrome P-450.




Life Sci.  20:  1 541-1 548.








Slater,  T.F.   1972.  Free  Radical  Mechanisms in  Tissue Injury.   Pion.  Ltd.,




London.
                                      4-79

-------
Smith,  A.A.,  P.O. Volpetto, Z.W. Gremling, M.B. DeVore and A.B.  Classman.   1973.



Chloroform, halothane and regional  anesthesia.   A comparative study.   Anesth.



Analg.   52: 1-11.







Snow, J.  1858.  On chloroform and other anesthetics.  London,  John Churchill.








Steward,  A.,  P.R.  Allot, A.L.  Cowles and  W.W.  Mapleson.   1973.    Solubility



coefficients  for  inhaled anaesthetics  for water,  oil,  and biological  media.



Brit. Jour. Anaesth.  45: 282-293.







Stewart,  R.D.   197^.   The use of breath  analysis  in clinical  toxicology.  In:



Essays in Toxicology.  New York,   Academic Press.  Chapter 5.







Stewart, R.D., H.C. Dodd, D.S.  Erly  and B.B. Holder.  1965. Diagnosis of solvent



poisoning.  Jour. Amer. Med. Assoc.  193: 1097-1100.








Storms, W.W.   1973-  Chloroform parties.  Jour. Amer. Med. Assoc.  225: 160.







Symons  J.M.,  T.A.  Bellar,  J.K. Carswell, J.  DeMarco,  K.L.  Kropp,  G.G. Robeck,



D.R. Seeger,  C.J. Slocum, B.L. Smith and A.A.  Stevens.  1975-   National Organics



Reconnaissance Survey for Halogenated Organics.  Amer. Water Works Assoc. Jour.



67:  634-647.







Taylor,  D.C.,  D.M.  Brown,  R. Kuble  and P.P.  Langley.   1974.   Metabolism of


                                                         1 4
chloroform.  II.  A  sex  difference in the metabolism of   C-chloroform in mice.



Xenobiotica.   4: 165-174.
                                      4-80

-------
Thompson, D.J., S.D. Warnet  and  V.V. Robinson.   1974.   Teratology  studies on



orally administered chloroform in the rat and rabbit.  Toxicol. Appl. Pharmacol.



29: 348-357.






Torkelson, T.R., F.  Oyen  and V.K. Rowe.  1976.   The  toxieity of chloroform as



determined by single and  repeated exposure  of laboratory  animals.   Amer. Ind.



Hyg.  Assoc. Jour.   37: 697-704.







Tsurata, H.  1975.   Percutaneous absorption of organic sclent.  1.  Comparative



study of  the  iri  vivo percutaneous absorption of  chlorinated  solvents in mice.



Ind. Health.  13: 227-236.







Tsurata,  H.   1977.   Percutaneous absorption of organic solvents.  2.  A method



for measuring the  penetration  rate  of chlorinated solvents through excised rat



skin.  Ind. Health.  15: 131-140.







Uehleke,  H.  and T.  Werner.    1975.   A  comparative study  on the irreversible



binding  of labeled  halothane,  trichlorofluoromethane,  chloroform,  and  carbon



tetrachloride to hepatic protein  and lipids  in vitro and in vivo.   Arch. Toxicol.



34: 289-303.





                                                              14
Uehleke,  H., K.H. Hellmer and S.  Tarabelli.   1973-   Binding of  C-carbon tetra-



chloride  to microsomal proteins iri vitro and formation of CHC1., by reduced liver



microsomes.  Xenobiotic.  3: 1-11.







Uehleke,  H., T. Werner, H. Greim and M. Kramer.  1977.  Metabolic activation of



halothanes and tests in vitro for mutagenicity.  Xenobiotica.  7: 393-400.
                                      4-81

-------
Van Dyke,  R.A.,  M.B. Chenoweth and A. van Poznak.  1964.   Metabolism of volatile
                                                            1 4
anesthetics.  I. Conversion in vivo of several anesthetics to   CCL and chloride.
                              	                              d
Biochem.  Pharmacol.   13:  1239-124?.
Villarruel, M.C. and J.A.  Castro.   1975.   Irreversible binding of carbon tetra-
chloride  to  microsomal phospholipids.   Free  radical nature  of the  reactive
species and alterations in the physico-chemical  properties  of  the target fatty
acids.  Res. Commun. Chera. Pathol.  Pharmacol.   10: 105-115.


Villarruel, M.C., M.I.   Diaz-Gomes and J.A. Castro.  1975.   The nature of the
irreversible binding  of carbon-tetrachloride to  microsomal lipids.   Toxicol.
Appl. Pharmacol.  33: 106-116.


von  Oettingen,  W.F.   1964.    The Halogenated  Hydrocarbons of  Industrial  and
Toxicological Importance.   Amsterdam, Elsevier Publ. Co.   pp. 77-108.


Waters, R.M.  1951.  Chloroform.   A Study After 100 Years.  Madison, University
of Wisconsin Press.

Withey, J.R.,  B.T.  Collins and P.G. Collins.  1982.   Effect of vehicle on the
pharmacokinetics and  uptake  of four halogenated hydrocarbons  from  the gastro-
intestinal  tract of the rat.  Pre-print paper  submitted to J. Appl. Toxicol.,
December, 1982.
                                      4-82

-------
Withey,  J.R.  and  B.T. Collins.  1980.  Chlorinated aliphatic hydrocarbons used in




the foods  industry:   The comparative  pharmacokinetics of methylene  chloride,




1,2-dichloroethane,  chloroform and  trichloroethylene  after  i.v.  administration




in the rat.  J. Environ.   Path. Toxicol.  3: 313-332.








Witte, C.  1874.   XXXIV.  Experiments  concerning  the effects  of chloroform on




blood circulation.  Dtsch. Z Chir.  4: 5*48-566.








Wolf, C.R., D.  Mansuy, W.  Nastainczyk, G. Deutschmann and V.  Ullrich.  1977.  The




reduction  of polyhalogenated  methanes  by  liver  microsomal cytochrome  P-450.




Molec. Pharmacol.  13:  698-705.








Zeller,   A.   1883.  On  the  fate  of  iodoforms  and chloroform  in the  organism.




Hoppe-Seyler's Z. Physiol. Chem.   8: 78-79.

-------
                                 5.  TOXICITY




5.1  EFFECTS OF ACUTE EXPOSURE TO CHLOROFORM



     In both humans  and  experimental  animals,  characteristic effects  of  acute




exposure to chloroform are depression of the central nervous system and hepatic




damage.  Renal and cardiac effects also  occur.   The systemic toxic  effects  of




chloroform appear to  be similar regardless of whether exposure or administration




occurred by inhalation, oral, or  parenteral routes.  The only  systemic  effect




documented for dermal administration,  however,  is renal  damage.




5.1.1   Humans




     5.1.1.1    ACUTE  INHALATION  EXPOSURE  IN  HUMANS  —  Information  on  the




effects of acute inhalation exposure of  chloroform  on humans  has  been obtained




primarily during its use as an inhalation anesthetic.   The  relationship of the




concentration of chloroform in inspired air  and blood to anesthesia is described




in Table 5-1 (Goodman and Oilman,  1980).   Concentrations of  chloroform used for




the induction of anesthesia were in  the  range  of 2-3 volumes % (20,000-40,000




ppm), followed by lower maintenance  levels (NIOSH,  1974; Adriani,  1970).




     Chloroform  inhalation  has a  depressive  effect on the  central  nervous




system.   Excitement  due  to release  of  inhibitions is  followed  by progressive




depression of the cortex, higher centers,  medulla  and spinal cord  (Wood-Smith and




Stewart,  1964).   Centers  controlling  temperature  regulation,  respiration,




vomiting, vasomotor,  and  vagal activity are  all depressed (Adriani, 1970).




     The cardiovascular system is also affected by anesthetic use of chloroform.




The myocardium is directly depressed in  deeper planes of anesthesia.   A blood




level sufficient to cause respiratory failure may also cause cardiac arrest.  In




addition,  chloroform  sensitizes  the   autonomic   tissues  of  the   heart  to




epinephrine,  causing arrhythmias.   It  has  been found  that  under  chloroform
                                      5-1

-------
                                    TABLE 5-1

Relationship of Chloroform Concentration in Inspired Air and Blood to Anesthesia*
                                  In Inhaled Air           In Blood
                                    Volumes %
 *Source:  Goodman and Oilman, 1980
Not sufficient for anesthesia
Light anesthesia
(after induction)
Deep anesthesia
Respiratory failure
<0.15
0.15 to 0.20
0.20 to 1.50
2.0
<2
2 to 10
10 to 20
20 to 25
                                   5-2

-------
anesthesia regarded  as  normal,  the heart is  subject  to arrythmias  and  extra-




systoles (Kurtz et al.,  1936; Orth et al.,  1951).  Orth et al.  (1951) found a high




incidence of ventricular arrhythmias, 20 of 52 cases investigated, and four cases



of temporary cardiac arrest.  Blood pressure is lowered by chloroform as a result




of a 3-fold action:  cardiac slowing due to vagal stimulation, depression  of the




vasomotor  center,  and  dilation of splanchnic  blood  vessels  (Krantz and  Carr,




1965).




     Respiratory  effects  of  chloroform inhalation  include   increased rate  and




depth of  respiration during  induction and  in light  anesthesia, and decreased




minute volume exchange  in  deeper planes of anesthesia.  The Hering-Breuer  reflex




remains active.  Bronchial smooth muscle is relaxed and secretions are increased.




Laryngeal spasms are caused by  high concentrations  (Adrian!,  1970).




     In the gastrointestinal  tract, chloroform markedly  stimulates the flow  of




saliva during  induction and recovery, but  salivation  is  inhibited in  deeper




planes of anesthesia  (Goodman and Oilman, 1980).  The pharyngeal or gag reflex is




depressed.    Under  anoxic  conditions,  pharyngeal  muscle  spasms  result  in




stertorous respiration  and thick  mucus is  excreted  (Adriani, 1970).  Stomach




movements  are  decreased or abolished  as  tone is  reduced.    Gastric secretory




activity is  inhibited  or  abolished.    Post-anesthetic  dilation of the stomach




occurs in nearly all  cases.  Nausea and vomiting often occur during recovery from



anesthesia.  The mechanism is central  rather than local,  but  may  be  due in  part




to irritation of  the stomach  by  swallowed vapor  (Goodman  and Gilman,  1980).




Intestinal tone,  motility, and  secretory  activity are  inhibited or abolished



(Adriani, 1970).




     In the urinary tract,  chloroform  anesthesia  results  in a decrease in urine




flow,  possibly due to the release of antidiuretic hormone and renal vasoconstric-




tion,  leading  to  a  decrease  in  renal  blood flow and glomerular  filtration.
                                      5-3

-------
Polyuria  occurs  after  recovery  (Goodman  and  Oilman,   1980).     Chloroform




anesthesia may be followed by albuminuria and glycosuria.   Post-operative  urine




retention occurs frequently.   Renal tubular necrosis has been found in  cases  of



severe poisoning (Wood-Smith and Stewart, 1964).




     During obstetric use of chloroform, uterine contractions are  only  slightly




decreased  by  light anesthesia,  but  are markedly  inhibited  in deeper  planes.




Chloroform rapidly crosses the placental barrier,  and respiratory  depression  in




the infant is likely to occur (Wood-Smith and Stewart, 1964).




     Chloroform anesthesia also has metabolic effects in  humans.  A rise  in  blood




glucose accompanies anesthesia.  Levels  may rise >2-fold and remain elevated for




several hours.  The liver glycogen falls  coincident  with the rise in blood sugar.




This, in turn,  is a result of the release of epinephrine  from the adrenal medulla




during the period of excitation.  There is also a decrease in glucose utilization




in the  periphery (Goodman and Oilman,  1980; Krantz  and Carr,  1965).   Acidosis




occurs, characterized by a fall in plasma biocarbonate and phosphate.




     Chloroform  is acutely toxic to  the  liver,  although  in  so-called  delayed




chloroform poisoning, the full effects  of  damage done during  and  shortly  after




administration are not seen for 24-48 hours.  The  glycogen content of the  liver




is rapidly depleted;  three-fourths in the first half-hour and less rapidly there-




after.   There is  centrilobular  and,  in  severe  cases,  mid-zonal  and  massive




necrosis.   Cells  which  survive  show  fatty  degeneration.   Symptoms   include




progressive weakness,  prolonged  vomiting, delirium,  coma,  and  death.   They




develop from the first  to the third  day  after  exposure.   Jaundice, increased




serum  bilirubin, bile  in the urine,  reduction in  liver function, increased




nitrogen excretion, lowered blood  prothrombin and fibrinogen, and the appearance




of leucine, tryosine, acetone,  and diacetic acid  in the  urine  are  some of the




more prominent findings.  The hemorrhagic tendency is due to reduced prothrombin
                                      5-4

-------
formation by the injured liver.  Death usually occurs on the fourth or fifth day,




and autopsy reveals degeneration and necrosis of liver tissue, most marked around




the central veins (Goodman and Oilman,  1980;  Wood-Smith and Stewart,  1964).



     Hematologic effects  due to  acute chloroform inhalation  are seen  during




anesthesia.  Erythrocytes are increased in number as the  spleen  is  constricted




and red blood cells are extruded into the circulation.  Leukocytes are increased




in number during the  post-anesthetic period,  reaching a maximum within 24 hours




and returning to normal  in 48 hours.  There is  an increase in polymorphonuclear




cells.  Platelets remain  unchanged.  One  half-hour  after exposure,  there is  a




decrease in clotting  time. Prothrombin time is  increased.  Prothrombin synthesis




is impaired by liver  toxicity as previously noted (Adriani,  1970).




     The effects of  chloroform  on the  eye  include dilation of  the pupils, with




reduced reaction to  light  as  well as  reduced intraocular  pressure  (Sax,  1979;




Winslow and Gerstner, 1978).



     Signs of chloroform  poisoning include  a characteristic sweetish odor on the




breath, cold and clammy skin, and dilated  pupils  (Winslow and  Gerstner,  1978).




Nausea and  vomiting  commonly occur.   Ketosis,  due  to  incomplete oxidation of




fats,  as well  as  a  rise  in  blood sugar,  accompanies  chloroform  intoxication.




Initial  excitation  alternating  with apathy   is   followed  by   prostration,




unconsciousness, and possible death due  to cardiac  and central nervous  system




depression  (Winslow and  Gerstner,  1978).




     The above discussion  presents  observations made on the  effects  of chloro-




form inhalation during general anesthesia.   Information  on  the effects of experi-




mental acute inhalation  exposure of chloroform  in humans is limited to the work




of Lehman and Hasegawa (1910) and Lehman  and Schmidt-Kehl  (1936)  as  reviewed by




NIOSH  (1974).  The duration of exposure was <30 minutes and only the subjective
                                      5-5

-------
responses of  the  subjects  were measured.   The dose-response relationships  as




tabulated by NIOSH (1974)  are presented in Table 5-2.




     5.1.1.2    ACUTE  ORAL  EXPOSURE  IN  HUMANS --  Case reports  of  suicides



(Piersol et  al.,  1933;  Schroeder, 1965) and of recreational abuse (Storms,  1973)




of chloroform present  some  information on the effects of acute  imbibition.   A




fatal dose of Ingested chloroform may be as little  as  one-third  of  an  ounce  (10




m£)  (Schroeder,  1965).   The  initial  effect  is  usually unconsciousness  and




possibly death (within 12 hours without treatment) due to respiratory or cardiac




arrest.  If the patient survives, delayed  effects  are observed  within 48  hours




after  regained  consciousness.   These symptoms  include  vomiting,  anorexia,




jaundice, liver  enlargement,  albimuria,   ketosis,  ketonuria  and  glucosuria,




hemorrhage  due  to  lowered  blood  fibrinogen  and  prothrombin,  reduced  serum




bicarbonate,  increased  blood  sugar,  coma  and possible  death.   Upon  autopsy,




extensive hepatic centrilobular necrosis is evident.




     5.1.1.3    ACUTE  DERMAL  AND OCULAR  EXPOSURE  IN  HUMANS  — Chloroform  is




absorbed through  the intact  skin (von Oettingen, 1964).  Application of chloro-




form to  the skin is  followed after  3 minutes by a  pungent and  burning pain




reaching its maximum after 5 minutes,  associated with erythema,  hyperemia,  and



finally vesication (Oettel,  1936).   Exposure of the eye to concentrated chloro-




form vapors  causes a stinging sensation.  Splashing the substance into the eyes




evokes  burning,  pain, and  redness   of  the conjunctival  tissue.    The  corneal




epithelium is  sometimes impaired; however, regeneration starts rapidly  and  leads




to full recovery  within 1-3  days (Winslow  and Gerstner, 1978).




5.1.2   Experimental Animals




     5.1.2.1    ACUTE INHALATION EXPOSURE  IN ANIMALS  — Tolerance of animals to




chloroform has been summarized by Lehmann  and Flury  (1943)  and by  Sax  (1979).




Similar central nervous system  effects  are seen in animals at approximately the
                                      5-6

-------
                                   TABLE 5-2

                         Dose-Response  Relationships*
160 ppra (0.8 mg/fl,) for unspecified time - no odor

205 ppra (1.0 mg/il) for unspecified time - light transient odor

390 ppm (1.9 mg/&) for 30 minutes -light transient odor

920 ppm (4.5 mg/3,) for 7 minutes - stronger, lasting odor; dizziness, vertigo
     after 3 minutes

680 ppm (3.3 mg/£) to 1000 ppm (5.0 mg/£) for 30 minutes - moderately strong
     odor; taste

1100 ppm (5.4 mg/&) for 5 minutes - still stronger, permanent odor; dizziness,
     vertigo after 2 minutes

1400 ppm (6.6 mg/£) to 1800 ppm (8.57 mg/£) for 30 minutes - stronger odor,
     tiredness, salivation, giddiness, vertigo, headache, taste

3000 ppm (14.46 mg/£) for 30 minutes - all above plus pounding heart, gagging

4300 ppm (20.8 mg/£) to 5000 ppm (25 mg/S,) for 20 minutes - dizziness and
     light intoxication

5100 ppm (25 mg/]i) for 20 minutes - dizziness and light intoxication

7200 ppm (35.3 mg/&) for 15 minutes - dizziness and light intoxication as
     above but more pronounced

•Source:  Lehman and Hasegawa (1910) and Lehman and Schmidt-Kehl (1936).
                                  b-/

-------
same magnitude  of  exposure  that produced  these  effects in humans.   In  mice,




exposure to 2500 ppm  for 2 hours produced no obvious effects, 3100 ppm for 1 hour




produced slight narcosis,  while 4000 ppm induced deep narcosis  within one-half



hour.   Only slight  symptoms  are  seen at  2000-6000  ppm for longer  exposures.




Fatal  exposures were  4100-8200 ppm  for  mice,  12,300  ppm  for  rabbits,  and




16,300-20,500 ppm for guinea pigs  (duration of fatal  exposures  not  specified).




In cats, exposure to  7200 ppm resulted in disturbance of the equilibrium  after  5




minutes, light narcosis after 60 minutes, and deep narcosis after 93 minutes of




exposure.   Exposure to 21,500 ppm  produced disturbances  in equilibrium  after  5




minutes, light  narcosis  after  10  minutes, and  deep narcosis in cats after 13




minutes of  exposure.




     Kylin  et al.  (19&3) described the effects of  a single exposure of  mice to




100, 200, 400, or 800 ppm of chloroform for 4 hours.  The  mice exposed to 100 ppm




did  not  develop demonstrable  liver necroses,  although moderate  fatty infiltra-




tion of  the liver  was  noted.   In  mice exposed to 200 ppm, some necrotic  areas




appeared  in the liver  and  there  was  an  increase in  serum  ornithine-carbamyl




transferase.  Exposure to chloroform  at  400 and 800 ppm resulted  in increased




hepatic necrosis and serum enzyme  activity.



     More  recent  data  regarding toxic effects of  acute  inhalation  exposure to




chloroform  were presented  by Wood  et al.  (1982), although the study was designed




primarily   to  investigate  the  role  of  hydrogen  bonding in  the  anesthetic




mechanism.   Groups of  mice  in a rotating cage were given  a  single  exposure of




upto 3  hours of varying concentrations  of chloroform or deuterated chloroform,




each concentration being held  constant for about 20 minutes  and then being raised




until  the  mice had  lost their righting  reflex.   The  concentration was then




lowered to  1/2 the ED,-n (=1500 ppm) where it remained  until  the mice had regained




their  righting  reflex.    The  duration  of these  manipulated  exposures  never
                                      5-8

-------
exceeded 3 hours.  Only 4 of 47 mice given chloroform gained the righting reflex;




indeed,  some mice  died  or  were comatose.   Upon histological  examination of




animals sacrificed 3-6 hours after  exposure, mild hepatic centrilobular necrosis



and  very mild  renal  tubular  necrosis  was  observed.   The  animals receiving




deuterated chloroform survived for  24  hours,  after which they were sacrificed.




The liver and kidney lesions in these mice were more severe,  perhaps owing to the




longer survival time, which may have allowed these lesions to develop.




     5.1.2.2    ACUTE ORAL EXPOSURE IN ANIMALS




     Kimura et al. (1971) performed acute oral toxicity studies in newborn (5-8




g),  14-day-old  (16-50  g),  young adult (80-160 g), and older adult (360-470 g)




rats.  The chloroform was  given in undiluted form to unfasted rats.  LDp... values




in mS,/kg  (95% confidence  limits)  were  reported  as  follows:   14-day-old,  0.3




(0.2-0.5); young adult, 0.9  (0.8-1.1), and older adult,  0.8 (0.7-0.9).  The young




and  older  adult rats were  males;  the  other  two groups  contained  rats  of both




sexes.  When  compared with  15  other solvents  included in this study,  the LDj-0




values for chloroform were the lowest in  the two adult  groups and next to lowest




in the 14-day-old  rats.  Only a  rough approximation  of the LD^ could be obtained




for the newborn rats; volumes of  0.01 m&/10 kg body weight were generally fatal.




Lower volumes could  not  be measured with any  degree  of accuracy  and were  not



attempted.




     Torkelson et al.  (1976)  reported  an oral LD   of 2.0  g/kg  (1.05-3.80) in




male rats.   Animals  receiving  as little  as 0.25  g/kg showed adverse  effects.




Other recent studies of acute oral  toxicity have  reported LD   values (with 95/t




confidence limits)  of  1120 mg/kg  (789-1590)  in  ICR  male mice  and  1400  mg/kg




(1120-1680) in females (Bowman et al.,  1978),  and  908  mg/kg  (750-1082)  and 1117




mg/kg (843-1514),  respectively, in  male and female rats (Chu et  al., 1980).
                                      5-9

-------
     In a  study that compared  the toxicities  of  halogenated hydrocarbons,  a




single oral dose of 60 mg/kg chloroform to mice had no toxic  effect  (Hjelle  et




al., 1982).




     Hill (1978) performed experiments in mice designed  to study variability  in




susceptibility to chloroform toxicity from single  oral doses  based  upon  genetic




sex differences.   For  three  strains  of  mice, the LD    values (mS,/kg)  were:




DBA/2J, 0.08; B6D2F1/J,  0.20; and C57BL/6J, 0.33.  The  animals more  sensitive  to




chloroform-induced death  were  also  found  to  be  more   susceptible  to  renal




toxicity.  Males were found to  be more sensitive to renal  damage  and  death than




were females.   This  difference  was related to  testosterone  and  it was  further




noted that C57BL/10J males are relatively testosterone deficient in  comparison




to DBA  males.   The  C57BL/6J strain used in  the  Hill  (1978) study  is  closely




related  to  the  C57BL/10J  strain   and  may,   therefore,   also  be  testosterone




deficient.




     In male B6C3F1 mice,  severe diffuse renal necrosis  occurred  after a single




oral dose of 240 mg/kg  and  focal tubular regeneration occurred after a single




dose of 60 or 240 mg/kg.   These effects were  not  seen after  15 mg/kg (Reitz  et




al., 1980).  Liver damage  (hepatocellular necrosis  and swelling  with inflamma-




tory cell infiltration)  occurred only at the highest doses.




     Chu et al.  (1982a)  studied the  effects of acute oral exposure of  chloroform




on Sprague-Dawley rats.  Groups  consisted  of 10 male and 10 female animals given




a single oral dose of 0, 546, 765,  1071, 1500, or  2100 mg/kg  of  chloroform in a




volume of 5 m£/kg corn oil.  Clinical signs of toxicity  included  depression and




coma,  but the authors did  not specify whether these signs  occurred at all dose




levels.  Treated rats surviving  for 14 day.3 consumed less food and had depressed




growth rates.  Gross  examination revealed increased liver  and kidney  weights  at




1071 mg/kg.  Upon comprehensive  histological examination,  only mild to moderate
                                     5-10

-------
lesions, even at  high  doses,  were observed in these  organs.   No  changes  were




noted in other organs,  including brain and  heart.  The hepatic and renal lesions




were  characterized  by  hepatocyte variations  and occasional  vesiculation  of



biliary epithelial  nuclei in  the liver, and  by  bilateral focal  interstitial




nephritis and fibrosis  in the kidney.  Changes in  hematological and biochemical




parameters were also observed  in the  1071 and 1500  rag/kg treated groups.  Choles-




terol levels  increased  while  lactate dehydrogenase  activity and liver  protein




levels decreased.  In female rats, the activity of microsomal aniline hydroxylase




was induced by chloroform exposure.  The numbers  of  lymphocytes were reduced in




both males and females, as were hemoglobin and hematocrit values.   Upon longer




exposures of 5, 50, or 500 ppm chloroform in drinking water for 28 days, the only




toxic effect  observed   was  a  decreased  number of  neutrophils in  the  highest




dose-exposed rats.   Examinations were performed  as  in the  single-dose  experi-




ment .



     5.1.2.3    ACUTE DERMAL AND OCULAR EXPOSURE  IN  ANIMALS — Torkelson et al.




(1976)  found  that chloroform, when  applied to the skin  of rabbits,  produced




slight to moderate irritation and delayed healing of abraded skin.  When applied




to the uncovered  ear  of  rabbits,  slight hyperemia and exfoliation occurred after




one to four treatments.  No greater injury was  noted after 10 applications.  One




to two 24-hour applications, on  a cotton  pad bandaged on  the shaven  belly of the



same rabbits, produced  a slight hyperemia with moderate necrosis and a resulting




eschar  formation.   Healing  appeared  to  be  delayed  on the  site  as well  as  on




abraded areas which were  also  covered  for  24  hours  with  a cotton pad soaked in




chloroform.




     Single application of either  1.0,  2.0, or 3.98 g/kg  for  24 hours under an




impermeable plastic cuff held tightly around the  clipped  bellies  of each of two




rabbits did not result  in any  deaths.   However, extensive  necrosis of the  skin
                                     5-11

-------
and considerable weight loss occurred at all levels.  All animals were sacrificed
for study  2 weeks  after  exposure.  All treated  rabbits  exhibited  degenerative
changes in the kidney tubules graded in intensity  with dosage levels.  The livers
were not grossly affected.
     In the same study (Torkelson  et al., 1976),  liquid chloroform,  dropped into
the eyes of three rabbits, caused  slight irritation of the conjunctiva which was
barely  detectable  1  week after  treatment.    In  addition, slight  but  definite
corneal injury occurred,  as  evidenced by staining with fluorescein.  A purulent
exudate occurred for >2 days after treatment.  Although started 30 seconds after
instilling the  chloroform,  thorough  washing  of one eye  of  each rabbit  with  a
stream of running water for  2 minutes did not  significantly alter the response in
the washed eyes from that of the  unwashed  eyes.
     5.1.2.4    INTRAPERITONEAL AND  SUBCUTANEOUS ADMINISTRATION IN  ANIMALS —
The toxicity of chloroform in mice after subcutaneous administration (Kutob and
Plaa,  ig62b) and  intraperitoneal  administration (Klaassen and  Plaa,  1966)  has
been compared with  that of other halogenated hydrocarbons  (Pohl, 1979).  In these
studies,  the  LD   values for  carbon tetrachloride, chloroform, and dichloro-
methane were 200, 27.5, and  76 mraol/kg  after subcutaneous administration and 20,
14, and 23 mmol/kg when given intraperitoneally.   When the  relative  hepato-
toxicity of these compounds  was compared,  a  subcutaneous dose of 0.5 mmol/kg of
carbon  tetrachloride produced approximately the  same  degree  of liver damage as
6.2 mmol/kg of  chloroform.   After intraperitoneal  administration,  these values
dropped to 0.01  mmol/kg for  carbon tetrachloride  and 2.3 mmol/kg for chloroform.
Dichloromethane did not  cause significant histological  changes  in  the liver by
either route of administration.  At doses  that produced liver toxicity,  chloro-
form caused kidney lesions which  ranged from  the presence of hyaline droplets,
                                      5-12

-------
nuclear pycnosis,  hydropic degeneration, and increased eosinophilia, to necrosis
with karyolysis and loss of epithelium of the convoluted tubules.
     Ilett et al.  (1973) found that intraperitoneal administration of chloroform
caused  centrilobular  hepatic necrosis  in mice  of  both  sexes, whereas  renal
necrosis was observed only in male mice.
5.2. EFFECTS OF CHRONIC EXPOSURE TO CHLOROFORM
     A  characteristic  effect  of chronic  exposure  to chloroform  is  hepatic
damage; this effect has been documented primarily  in studies  with experimental
animals.  As was  the case for  acute  exposure to chloroform,  hepatic  damage  in
chronic studies results from  either  inhalation or oral administration  of this
chemical.  Effects on the kidneys and thyroids have  also  been observed  in some
experiments.  This section will  discuss both  subchronic (=90  days)  and  chronic
exposure  studies, because  many  of  the  subchronic   studies  were  preliminary
range-finding tests for the chronic studies.
5.2.1.  Humans
     5.2.1.1    CHRONIC  INHALATION   EXPOSURE  IN  HUMANS  —  Only  two  chronic
inhalation studies that  reported measurements  of  exposure concentrations,  as
well  as  effects  on human health, were  found.  Neither study  (Challen  et al.,
1958; Bomski et al.,  1967) is particularly adequate or recent.
     Challen et al. (1958) investigated complaints of  workers (mainly women) in a
plant  manufacturing  lozenges   that   contained   chloroform  as  a   principle
ingredient.   Before  exhaust  ventilation was  installed,  9 of the  10  exposed
workers had  complained of symptoms  of tiredness,  dull-wittedness,  depression,
gastrointestinal distress,  and frequent and scalding  urination.  Breathing-zone
monitoring during simulation of  "pre-ventilation" working  conditions  suggested
that the emplyees had been exposed to =77-237 ppm.   Discussions with management
revealed that some of these workers had occasionally been observed to behave in a
                                     5-13

-------
silly manner or to stagger about during  the workday.  Another group of workers (N




= 10) had been exposed primarily to  concentrations of 22-71 ppm.  Eight of these




workers complained of less severe symptoms.   Apparently,  both  groups of workers



had been exposed  to  occasional peak concentrations of =1163  ppm lasting 1.5-2




minutes. At least four workers in each  group worked half-time.  None of the five




controls reported symptoms  similar  to those  reported  by the  exposed  workers.




Eight of the higher exposure employees (77-237  ppm for 3-10 years, followed by =2




years without exposure),  nine of the  lower  exposure employees  (22-77  ppm  for




10-24 months), and five unexposed anployees  submitted to  physical examinations,




including liver  function tests  (thymol  turbidity,  serum bilirubin,  and  urine




urobilinogen tests).   These examinations and  tests revealed no evidence of  any




organic lesion,  including liver damage, attributable to exposure to chloroform.




     In humans,  hepatic damage is the most common toxic effect  of acute exposure




to chloroform, as noted previously.   According to Pohl (1979),  only one report of




liver abnormalities in humans after chronic exposure to chloroform has been found




in the literature and no additional  reports were found in  the more recent litera-




ture.  In this study  (Bomski et al., 1967),  17 cases of hepatomegaly were found




in a group of 68 industrial  workers  who were exposed to chloroform in concentra-




tions ranging from  2-205 ppm  for 1-4 years.    These  were unknown  rather  than




breathing zone concentrations.   Three  of  the  17 workers  with  hepatomegaly were




judged  to have  toxic hepatitis  on  the basis of  elevated serum  enzymes.   The



frequency of viral hepatitis among the 68 chloroform-exposed workers was higher




(4.4? versus 0.38?) than the frequency  among a group of inhabitants of the city,




XI8 years of age.  This phenomenon  also  occurred in the 2 previous years.  Ten




cases of splenomegaly were also diagnosed among the 68 workers.  There appears to




be no comparison with incidences of these conditions in nonexposed workers.
                                      5-14

-------
     5.2.1.2    CHRONIC ORAL  EXPOSURE  IN  HUMANS —  There are  relatively  few




reports of toxic effects following chronic ingestion of chloroform (Pohl, 1979),




and more recent reports were not located in  the literature.  In one case, it was



estimated  that  a male  patient  ingested  1.6-2.6 g  of chloroform  in  a  cough




medicine daily for =10 years (Wallace, 1950).  Blood and urine analyses, as well




as liver  function tests,  indicated  the individual suffered from hepatitis  and




nephrosis.   Another  report  described three  patients addicted to chlorodyne,  a




tincture containing  chloroform and morphine  (Conlon, 1963).  Liver biopsy showed




severe cellular  damage  in one of these individuals  who had ingested  21  rnS, of




chloroform  daily for an  undetermined  period  of time.    All  three  displayed




evidence  of serious mental  and physical  deterioration,   including  peripheral




neuropathy.  It is not possible to  determine if  the  adverse effects were due to




chloroform, morphine, or ethanol.



     More recently,  the safety of a dentifrice containing 3>^%  chloroform and a




mouthwash containing 0.43^ was assessed in  studies lasting  >_1  year (DeSalva et




al., 1975).  The subjects using the  dentifrice were exposed to =70 mg (0.0^7 mfi,)




of chloroform each day, whereas  the  groups  using  the  mouthwash were  exposed to




=178 mg (0.12 m&).  The results of  liver function tests and blood urea nitrogen




determinations  showed  no  statistical  differences  between control and experi-




mental subjects.



     Epidemiologic studies  of humans  exposed to chloroform in  their drinking




water have  focused on carcinogenic  endpoints, and, hence,  are  discussed in the




chapter on  carcinogenicity.



5.2.2   Experimental Animals




     5.2.2.1    CHRONIC  INHALATION   EXPOSURE  IN  ANIMALS   —  Experiments  with




several  species of  animals  (Torkelson  et   al.,  1976)  give  some  information




regarding potential  effects of long-term inhalation exposure to chloroform.  The
                                      5-15

-------
animals were exposed  to  chloroform  5 days/week for 6 months.  Exposure to 25 ppm
of chloroform for up  to  4 hours/day had no adverse effects in male rats as judged
by organ and body weights, and by  gross  and  histological examination of livers
and  kidneys.   Exposure  to 25 ppm for  7  hours/day,  however, produced histo-
pathological changes  in the livers and  kidneys of male  but  not  female  rats.
These  changes  were  characterized  as  lobular  granular degeneration  and  focal
necrosis throughout the liver and  cloudy swelling of  the kidneys.   The  hepatic
and renal effects appeared to be reversible because rats exposed according to the
same protocol, but given a  6 week  recovery  period  following exposure, appeared
normal by the criteria tested.  Increasingly  pronounced changes were observed in
the  livers  and  kidneys of  both  sexes of rats  exposed to 50 or  85 ppm  for  7
hours/day.   Hematologic  indices,  clinical  chemistry,  and urinalysis  values,
tested at higher levels of exposure,  were within normal  limits.   Each exposure
group  and  control group had 10-12 rats/sex  except  for the 25 ppm,  4 hour/day
group, which had 10 male rats and no females.
     Similar experiments with guinea pigs (N = 8-12 sex/group) and rabbits (N =
2-3/sex/group) gave  somewhat  inconsistent  results.   Histopathological  changes
were observed in livers  and kidneys of both species at  25 ppm but not at 50 ppm in
either species,  nor even at 85 ppm  in guinea pigs.  The results of these studies
are summarized in Table 5-3.
     Other  reports of  effects  of chronic inhalation  exposure to  chloroform in
experimental animals  were not found in the more recent literature.
     5.2.2.2    CHRONIC  ORAL  EXPOSURE IN  ANIMALS  — The  data  from  several
studies on the effects of chronic and  subchronic oral exposure to  chloroform are
summarized  in Table 5-4.  Low levels of exposure (15-64 mg/kg/day, 6 days/week)
have  been  reported to  increase  survival in mice and  rats (Roe  et  al.,  1979;
Palmer et al., 1979)  and to be associated with possible transient CNS depression
                                      5-16

-------
                                                                 TABLE 5-3

                             Effects  of Inhalation  Exposure of Animals to  Chloroform, 5 Days/Week for 6 Months*
 Species
             Sex
                    ppm
       Exposure
       hours/day
      Number in Group
Started            Survived
                                                                                                      Effects
rats
85
                                                    10
              F       85
                                                     10
                                                                       10
                       50       7
                                 10
                       50       7
                                 10
                                                                        10
                       25       7
                       25       7
                                 12
                                 12
                                                                        12
                                 Excess mortality attributed to pneumonia on basis of gross
                                 and microscopic appearance of lungs;  slight depression of final
                                 body weight;  increase (p<0.05) in relative but not absolute
                                 weights of liver, kidneys, and testes;  no effect on spleen
                                 weight; histological findings included  marked centrilobular
                                 granular degeneration of the livers and cloudy swelling of the
                                 kidneys but no histopathological changes in testes; hematologic
                                 values (including differential count),  urinalysis values
                                 and SGPT,  SUN, and SAP values all "within normal limits"

                                 No evidence of pneumonia;  final body weights and weights of
                                 liver and spleen unaffected, relative and absolute kidney
                                 weights increased; histological findings included marked
                                 centrilobular granular degeneration of  the livers and cloudy
                                 swelling of the kidneys; hematologic, urinal ysis, and
                                 SGPT, SUN, and SAT values  all "within normal limits"

                                 Depression of final body weights (p<0.05); increases
                                 in relative (p<0.05) but not absolute weights of kidneys,
                                 spleen, and testes; histopathological changes in livers
                                 and kidneys similar to those seen at  85 ppm; hematologic
                                 values, urinalysis values, and SGPT,  SUN,  and SAP values
                                 all "within normal limits"

                                 Final body weights and weights of liver and spleen unaffected;
                                 increase in relative kidney weight (p <0.05);  histopatho-
                                 logical changes in livers  and kidneys similar to
                                 those seen at 85 ppm but somewhat less  marked;  hemato-
                                 logic values, urinal ysis values, and  SGPT, SUN,
                                 and SAP values all "within normal limits"

                                 No effect  on  final body weights or weight  of liver,
                                 spleen, or testes; increased relative kidney weight
                                 (p <0.05); lobular granular degeneration with
                                 focal areas of necrosis throughout the  liver;  cloudy
                                 swelling of renal tubular  epithelium

                                 No statistically significant effect on  body
                                 weight; increased relative but not absolute
                                 kidney and spleen weights;  no histopathologic
                                 changes in kidneys and spleen;  microscopic appearance
                                 of livers  not specified

-------
                                                                       TABLE 5-3

                               Effects of Inhalation Exposure of Animals  to Chloroform,  5  Days/Week for  6 Months* Ccont.)
Exposure Number in Group
Species
rats



Sex ppm hours/day Started
M,F 25 7 12/sex
plus 0 ppm for
6 weeks (recovery
period)
Survived Effects
8 M, "Normal" by the criteria tested at this dosage level
10 F (see 25 ppm above)


guinea
Pigs
 I
CD

rabbits
dogs
                       25
             M,F    85,50,  or 25
             M,F    85,  50,  or 25
                                       1,2,  or  4         10, 10, 10           7,8,4,      No evidence of adverse effects by the criteria
                                                                          respectively   tested (i.e., final body weight; weights of livers, kidneys,
                                                                                         spleen, testes; and probably gross and microscopic
                                                                                         appearance of at least the liver and kidneys)

                                          7            8 to 12/sex/        50 to 92%      No adverse effects at 50 or 85 ppm other than
                                                      exposure level      (mortality     marked pneumonitis in F at 85 ppm; some histopatho-
                                                                          not related    logical changes in livers of both sexes and
                                                                          to exposure)   kidneys of M at 25 ppm (criteria tested were body weights,
                                                                                         organ weights, and gross and microscopic appearance
                                                                                         of organs)

                                          7            2 to 3/sex/         0 to 1 death/  No adverse effects at 50 ppm; some histopatho-
                                                      exposure level      group, not     logic changes in kidneys and liver and pneumonitis
                                                                          related to     in lungs at 25 and 85 ppm.  Hematologic and
                                                                          exposure       clinical chemistry values within normal
                                                                                         limits at 85 ppm (criteria tested were same as for rats)

                                          7               1/sex             1/sex         No adverse effects in M; marked cloudy swelling of
                                                                                         renal tubular epithelium and increase in
                                                                                         capsular space in glomeruli of kidneys in F
                                                                                         (criteria tested were same as for rats and included
                                                                                         clinical chemistry and hematological studies)

•Source:  Torkelson et al.,  1976

M = male; F = female;  SGPT = serum glutamic  pyruvic transaminase; SUN = serum urea nitrogen; SAP = serum alkaline phosphatase

Controls for each species and sex included at least one unexposed and one air-exposed group, each comparable in number of animals to the exposed
groups.  In statistical comparisons of organ and  body weights, values for the control group (unexposed or air-exposed) closer in body weight to the
test group were used.   Mortality in control  groups was similar to mortality in treated groups, with the exception of excess mortality in male rats
exposed to 85 ppm for  7 hours/day or to 25 ppm  for 4 hours/day.  No explanation was given by Torkelson et al. (1976) for the high mortality in the
4 hours/day group.  Strains  of animals and age  or weight at the start of the experiment were not specified.  Purity of the chloroform used was 99.3%
(0.4} ethyl alcohol and <0.3% of an unknown).
             M,F
25

-------
                                  TABLE  5-4

Effects of Subchronic or Chronic  Oral  Administration of Chloroform to Animals
Species, Strain
Age/Weight at
Start sex
Rats, Sprague-Dawley M,F
weanling, 101 g M,
94 g F
No. at
Start
20/sex/dose
level
Vehicle
drinking
water
Dosage
0, 5, 50, 500,
or 2500 ppm in
drinking water
Duration
90 days, after
which 10 rats/
group were killed
Response
Increased mortality,
decreased growth rate,
and decreased food
Reference
Chu et al., 1982b
                          ad lib. corre-
                          sponding  to
                          intakes   of  0,
                          0.11-0.71, 1.2-
                          1.5, 8.9-14, or
                          29-55 mg/rat/day
                          The highest  dosec
                          corresponds  to
                          =291 mg/kg/day F,
                          310 mg/kg/day M
and 10 rats/group
observed for
additional 90 days
intake at highest dose;
increased frequency of
mild to moderate
liver and thyroid
lesions at highest
dose, including
increases in cytoplasmio
homogeneity, hepato-
cyte density, and
cytoplasmic volume,
vacuolization due to
fatty infiltration, some
vesiculation of biliary
epithelial nuclei, and
hyperplasia in livers;
and reduced follicle
and colloid density,
increased epithelial
height, some focal
collapse of follicles in
thyroid; no histopatho-
logical effects in
kidney, brain, and heart;
after the 90-day recovery,
the lesions were very
mild and similar to
those seen in controls

-------
                                                                              TABLE 5-4


                                        Effects of Subchronic  or  Chronic  Oral  Administration of Chloroform to Animals (cont.)
un
 i
Species, Strain
Age/Weight at
Start Sex
Rats, Osborne-Mendel M
6 weeks/ 190 g



















No. at
Start Vehicle Dosage Duration
30/group drinking 0, 200, 400, 600 10 rats/group
except water 900, or 1800 ppm killed at 30,
HO ad lib. in drinking water 60, and 90 days
controls ad lib. cor- of exposure
responding to
intakes3 of 0,
20, 38, 57, 81, or
160 mg/kg/day,
plus 0 ppra group
matched with
1800 ppm group
for water con-
sumption








Response Reference
Dose-related signs Jorgenson and
of depression during Rushbrook, 1980
1st week only; dose-
related reduction in
water consumption;
decreased weight gain
in 160 mg/kg group;
increased incidence of
"hepatosls" in livers
of treated rats at 30
and 60 days but not
90 days (not dose-
related) ; no other
treatment-related
effects on serum
clinical chemistry
values or urinalysis
values, or gross and
microscopic appearance
of tissues, including
kidney.

-------
                                                                              TABLE 5-4

                                         Effects of Subchronic or Chronic Oral Administration of Chloroform to Animals  (cont.)
Species, Strain
Age/Weight at
Start

No. at
Sex Start Vehicle Dosage Duration


Response


Reference
       rats, Osborne-Mendel
        52 days/240 g M
        175 g F
M,F       50/sex/dose    corn  oil,
          level;  20/sex   gavage
          matched con-
          trols;  =99/
          sex colony
          controls
M: 0, 90, or 180
 mg/kg/day;
F: 0, 100, or 200
 mg/kg/day (TWA);
5 days/week
78 weeks treat-
ment plus 33 weeks
observation
 i
t-o
Treated animals had      NCI, 1976
dose-related decrease
in survival and weight
gain, slight decrease
in food consumption,
increased severity and
incidence of pulmonary
lesions characteristic
of pneumonia; necrosis of
hepatic parenchyma NS in
controls, 3/50 low dose M,
4/50 high dose M, 3/49 low
dose F, 11/48 high dose F;
hyperplasia of urinary
bladder epithelium 1/18
control M, 7/45 low dose
M, 1/45 high dose M, NS
for control F, 6/43 low
dose F, 2/40 high dose
F; increased splenic
hematopoiesis in 1/18
control M, 3/45 low dose
M, 6/45 high dose M; both
control and treated
animals had chronic
nephritis; low but
statistically signi-
ficant increased
incidence of renal
epithelial tumors
in treated M (see
Carcinogenic!ty
section)

-------
                                                                     TABLE 5-4

                                Effects  of Subchronic or Chronic Oral Administration of Chloroform to Animals  (cont.)
Species, Strain
 Age/Weight at
 Start
                       Sex
           No. at
           Start
                                                Vehicle
                                                                Dosage
                                                                                  Duration
                                                                                   Response
                                                                                                   Reference
rats, Sprague-Dawley
  NS
M,F
10/sex/dose
  level
toothpaste,
 gavage
0,  15, 30, 150,
or 110 mg/kg/day
6 days/week
13 weeks
rats,  Sprague-Dawley
  SPF.  180  to 240  g  M
  130  to 175 e F
 M,F
50/sex/group   toothpaste,
                gavage
               0 or  60 mg/kg/day
               6 days/week
                    80 weeks exposure
                    plus 15 weeks
                    observation
At 410 mg/kg/day,
increased liver weight
with fatty change
and necrosis,
gonadal atrophy,
increased cellular
proliferation in
bone marrow; at
150 mg/kg/day,
changes less pro-
nounced but effect
(NS) on relative
liver and kidney
weights; pre-
sumably no effects at
lower dosage levels
Palmer et al.,  1979
                    Survival  of treated
                    animals slightly
                    better than that of
                    controls  (32$  treated M,
                    22$  control M,  26$
                    treated F,  14$ control
                    females survived to 95
                    weeks); body weights of
                    treated rats slightly and
                    progressively  depressed;
                    intercurrent respiratory
                    and  renal disease in all
                    groups; minor  histologi-
                    cal  changes in livers but
                    no evidence of "treatment-
                    related toxic  effect" in
                    livers; decrease
                    (p<0.01)  in relative
                    liver weight in
                    treated females;  no
                    gross or  histologic
                    treatment-related
                    changes in  brain;
                    possible  effect  (NS)  on
                    incidence of severe glomerulo-
                    nephritis;  decrease
                    in plasma cholines-
                    terase in treated females
                         Palmer et al., 1979

-------
                                                                   TABLE 5-1.  (cent.)

                                  Effects  of Subchronic or Chronic Oral Administration of Chloroform to Animals  (cont.)
Species , Strain
Age/Weight at
Start

No. at
Sex Start Vehicle Dosage Duration


Response


Reference
mice, B6C3F1,
 6 weeks/19 g
         30/group
         except 40
         ad lib.
         controls
               drinking
               water
               0,  200, 400, 600,
               900,  1800, or
               2700  ppm in
               drinking water
               ad lib. cor-
               responding to
               intakes"
               of =0,  32, 64,
               97, U5, or 290
               mg/kg/day; addi-
               tional  0 ppm
               group matched
               with  2700 ppm
               group for water
               consumption
                    10 mice/group
                    killed at 30, 60,
                    and 90 days of
                    exposure
                    Dose-related signs
                    of depression during
                    first week only; marked
                    reduction in
                    water consumption
                    in higher-dose
                    groups during first
                    2 weeks; body
                    weight losses (p<0.05)
                    in 97, 145, and 290 mg/kg
                    groups and in matched
                    controls during first week;
                    mild hepatic centrilobular
                    fatty change in 64, 97,
                    145, and 290 mg/kg groups
                    at 30 days, but only in 2
                    highest dosage groups at
                    60 and 90 days; increase
                    in liver fat/liver
                    weight (p<0.05) for
                    290 mg/kg group at
                    all 3 periods;
                    no other treatment-
                    related changes in serum
                    enzyme levels, urinalysis
                    values, or gross or
                    microscopic appearance of
                    tissues including kidney
                         Jorgenson and
                         Rushbrook, 1980
mice, B6C3F1
 35 days/l8g M,
 17 g F
M,F
50/sex/dose
level; 20/sex
matched
controls
corn oil,
gavage
M: 0, 138, or 227
mg/kg/day;
F: 0, 238, or 447
mg/kg/day;
5 days/week
78 weeks treat-
ment plus 14 to 15
weeks observation
Survival decreased       NCI, 1976
in high dose F,
unaffected in
other treated groups;
high incidences of
hepatocellular car-
cinoma in treated mice
(see Carcinogenicity sec-
tion); renal inflamma-
tion in 10/18 control M,
2/50 low dose M,
1/50 high dose M

-------
                                                                            TABLE  5-4

                                       Effects of Subchronic or Chronic Oral Administration of Chloroform to Animals (cont.)
Species, Strain
Age/Weight at
Start
mice, Schofield
No. at
Sex Start Vehicle Dosage Duration
M,F 10/sex toothpaste, 0, 60, 150, 6 weeks
per dosage gavage or 425 mg/kg/
day; 6 days/
week
Response Reference
At 425 mg/kg, 100? Roe et al . , 1979
mortality; at 150 mg/kg,
8/10 M died and
weight gain of F
markedly retarded; at
60 mg/kg, weight gain
of both sexes moder-
ately retarded; no
other observations men-
tioned
     mice, ICI (expt. 1)
     ICI (SPF) (expt. 2)
     ICI, CBA, C57BL,
     CF/1 (expt. 3)
      <10 weeks old
\_n
 i
expt. 1:
 M,F
expt. 2
and 3:
  M
treated
and con-
trol,
plus
some F
control
treated:
52/sex/dose
level;  con-
trol: 52  to
206/sex/
strain/
vehicle plus
untreated
toothpaste in
all 3 expt.
for all
strains and
sexes plus
arachis oil
in expt. 3
for ICI,
gavage
0, 17 (expt.1),
or 60 mg/kg/day,
6 days/week
80 weeks treat-
ment; 16 to 24
weeks observa-
tion according
to numbers of
survivors
Survival generally
better in 60 mg/kg
groups than in con-
trols except for
CF/1 animals or when
chloroform given in
arachis oil; slight
retardation of weight
gain in 60 mg/kg
groups; no effect
on hematologic values
(tested in expt. 2
only); no treatment-
related adverse effect
on liver or other
tissues except in kid-
neys as follows:
60 mg/kg in tooth-
paste - increased inci-
dence of moderate
to severe renal
changes (p <0.001) in
CBA and CF/1 M,
60 mg/kg in oil-
increased incidence of
moderate to severe kid-
ney disease (p <0.05) in
1C1 M; increased incidence
of benign and malig-
nant kidney tumors
in ICI M treated with
60 mg/kg in tooth-
paste or oil (see
Carcinogenicity section)
Roe et al.,  1979

-------
                                                                        TABLE 5-4

                                    Effects of Subchronic or Chronic Oral Administration of Chloroform to Animals (cont.)
Species, Strain
Age/Weight at
Start
dogs , beagle
18 to 24 weeks
7 to 8 g
No. at
Sex Start
M,F 1 /sex/dose
level for
90 and 120
mg/kg; 2/sex/
dose level
for lower
dosages
Vehicle
toothpaste
in gelatin
capsule,
orally
Dosage
30, 45, 60, 90,
or 120 mg/kg/day,
7 days /week
Duration
13 weeks for
30 and 45 mg/kg,
18 weeks for
60 mg/kg, 12 weeks
for 90 and 120
mg/kg
Response Reference
No deaths; occasional Heywood et al., 1979
vomiting; marked
weight loss in all
dogs and poor
general condition in
some at 60 mg/kg or
higher ; apparent
 I
M
vn
suppression of appe-
tite initially at all
dosages and through-
out at 60 mg/kg and
higher; jaundice and
increased SAP, SCOT,
SGPT, bllirubin, and
ICD values in male at
120 mg/kg; increased
SGPT values in 1/4 and
increased SAP and SCOT
values in 2/4 at 60
mg/kg; hepatocyte
enlargement and vacuo-
1ation with fat depo-
sition at 60 mg/kg and
higher; discoloration of
liver, increased liver
weight, and slight fat
deposition in hepato-
cytes at 45 mg/kg; no
effect on any of these
clinical chemistry or
histological parameters
at 30 mg/kg

-------
                                                                        TABLE 5-4

                                   Effects of Subohronio or Chronic Oral Administration of Chloroform to Animals  (cont.)
Species, Strain
Age/Weight at
Start
dogs, beagle,
8 to 24 weeks,
7 to 8 kg
No. at
Sex Start
M,F 8 /sex/ dose
level; 16/sex
vehicle
controls;
plus other
controls
Vehicle
toothpaste
in gelatin
capsule,
orally
Dosage
0, 15, or 30
mg/ kg/day;
6 days/week
Duration
7.5 years treat-
ment plus 20 to
21 weeks observa-
tion
Response
No effect on survival ,
growth, organ
weights , hematologic
or urinalysis values
(checked at intervals
thr ougho ut ) ; moderat e
Reference
Heywood et al . , 1979
 I
ro
dose-related elevation
of SGPT reaching peak
in sixth year of study,
reverting to normal
levels after treatment
discontinued; other
serum enzyme indica-
tors of hepatic damage
(checked during the
latter portion of the
study) followed
pattern similar to SGPT,
but BSP retention
and ICD values were
unaffected; aggregation
of vacuolated histio-
cytea ("fatty cysts")
in livers of all groups
but cysts were larger
and more numerous in
treated dogs and
persisted after treat-
ment ended; fat depo-
sition affected more
renal glomeruli in
30 mg/kg group than
in other groups
  Calculated by Jorgenson and Rushbrook (1980) from measured  average  body weights  and water  consumption.

  Calculated from Jorgenson and Rushbrook's statement that  the  mice had actual  intake levels  of  from  148 to  175J  of  the
  intended levels of 20, 40, 60, 90, 180,  and 270 mg/kg/day.

  Calculated by Chu et al. (1982b) by multiplying the fluid intake volume by  the concentration of  chloroform

  Growth rate data were given only for the highest dose and mg/kg/day were  calculated from this  information  by  taking average weights over the 90-day
  period of exposure.
  SPF  =  specific pathogen-free; SGPT = serum glutamic-pyruvic  transaminase;  SAP  = serum  alkaline  phosphatase; BSP  = bromsulphthalein;
  ICD  =  isocitric dehydrogenase (serum)
  Purity of the chloroform samples used in all these studies was generally high  and is discussed  in the section  on Carcinogenic!ty.
  M  =  male; F = female; TWA = time-weighted average dose for days on which chemical was  administered;  NS = Not specified;

-------
and mild hepatic changes in mice and rats (Jorgenson and Rushbrook, 1980; Palmer
et al., 1979), hepatic damage in dogs  (Heywood et al.,  1979),  and renal damage in
male mice of some sensitive strains (Roe et al., 1979), and in dogs (Heywood et
al.,  1979).    In addition  to  hepatic damage,  ingestion  of  =300  mg/kg/day of
chloroform produced thyroid lesions in rats (Chu et  al.,  1982b).  In one study, a
decreased incidence  of renal  inflammation occurred  in male  mice  treated  with
chloroform at 138 or 227 mg/kg/day, 5 days/week (NCI,  1976).   A  similar effect
may have occurred in rats (Palmer et al.,  1979), but the authors did not specify
whether the effect of  chloroform  was  to increase or decrease the  incidence of
intercurrent renal disease.  The NCI (1976) report stated that hepatic necrosis,
hyperplasia of  the  urinary bladder epithelium, and  increased splenic  hemato-
poiesis in rats may have been related  to chloroform treatment, but incidences in
controls for some of these effects were not  reported and the data, shown in Table
5-4, are difficult  to  interpret.   Many of the  studies  summarized  in Table 5-4
were at  least  partially  designed as  investigations  of  carcinogenicity  and,
hence,  are also discussed in the  carcinogenicity section of this  document;  some
experimental details are discussed more  fully in that  section.
     From the  data  presented in  Table  5-4,  it appears that  rats  and mice can
tolerate higher daily intakes of  chloroform when it is  given  in  their  drinking
water or  in a toothpaste base (by gavage) than they can when the chemical is
administered in  corn or  arachis  oil  (by gavage).  In  the subchronic study of
Jorgenson and  Rushbrook (1980), rats and mice appeared to adapt to low levels of
chloroform intake (up to =100 aig/kg/day); signs  of depression and  mild  hepatic
damage  that  occurred  initially had disappeared by 90 days of treatment.  Elevated
indices of  liver damage  (e.g.,  SGPT  levels)  in dogs  chronically exposed  to
chloroform  reverted  to  "normal"  after  treatment  was  discontinued,  although
histological changes persisted.   Similarly, the  mild liver and thyroid  lesions
                                     5-27

-------
seen in rats  exposed to high doses of chloroform via their drinking water for 90




days were no  longer apparent in rats allowed to recover for an additional 90 days




(Chu et al.,  1982b).



5.3  INVESTIGATION OF  TARGET ORGAN TOXICITY IN EXPERIMENTAL ANIMALS




5.3-1   Hepatotoxicity.   An extensive review  of  the early literature  dealing




with  chloroform-induced  liver  damage by  von  Oettigen  (196M)  notes  studies




beginning in 1891.   More recently,  Groger and Grey  (1979) summarized  reports




describing chloroform-induced liver hepatotoxicity as follows:   typical  effects




of  chloroform  on  liver  cells  are extensive  vacuolization,  disappearance  of




glycogen,  fatty  degeneration,  swelling,   and  necrosis,  all  starting  in  the




centrilobular areas.  There is also often  hemorrhaging into the parenchyma and




infiltration of  polymorphonuclear cells and  monocytes.   Electron-microscopic




observations of  liver parenchymal  cells  from  chloroform-intoxicated  rats  as




carried out by  Scholler (1966, 196?) revealed deposition  of lipid droplets in the




cytoplasm, partial  destruction  of the  mitochondrial matrix, proliferation of




smooth  endoplasmic reticulum, and swelling of the rough  endoplasmic  reticulum




with detachment of ribosomes.




     Kylin et al. (1963)  conducted a study  of the hepatotoxic effects of inhaled




trichloroethylene, tetrachloroethylene, and chloroform  in mice  with  the objec-




tive of finding  the  lowest  concentration of the  substances producing signs of




liver  damage after a  single U-hour exposure period.   Histological examination




showed that a concentration  of 100 ppm caused moderate fatty  infiltration in mice




killed  1 day  after exposure.  At  >200 ppm, the extent of  the alteration increased




with concentration and was  more pronounced after  1 day than  3-   Thus,  judging




from  the  histological  picture,   the smallest  concentrations   (ppm)  of  the




different agents to produce  more severe alterations in the exposed group than in



the controls were as follows:
                                      5-28

-------


1 day after
exposure
3 days after
exposure
Trichloro-
ethylene
(ppm)
1600-3200

>3200

Tetrachloro-
ethylene
(ppm)
<200

200 to 400


Chloroform
(ppm)
<100

100 to 200

On  this  basis,  the  hepatotoxic  effects  of  trichloroethylene,  tetrachloro-




ethylene, and  chloroform  are  in  the  approximate  ratios 1:10:20.  The amount of




liver  fat  was  raised at  400  ppm of  chloroform.    A  third indicator  of liver




toxicity  was  an  increase  in  serum  ornithine   carbamyl  transferase  (S-OCT)




activity at 24 hours in animals exposed to 200, 400, and 800 ppm of chloroform.




     A  study of  the effect  of oral doses  of chloroform on the extent of liver




damage in white mice ("of  a  Swiss strain") was conducted by Jones et al. (1958).




Minimal  changes  characterized by midzonal  fatty  infiltration  were observed 72




hours  after the administration  of 30  mg  (0.02  m£)/kg.   When  the  dose  was




increased to 133 mg (0.09  m£)/kg, a massive fatty infiltration of the total liver




lobule was found.  At a level of 355 mg (0.24 m£)/kg,  massive fatty infiltraton




occurred along with severe central  lobular necrosis.  Information on the hepato-




toxicity of long-term chloroform administration has been  presented in the sec-




tions on Effects of Chronic  Exposure to  Chloroform.  Inhalation exposure of rats




to  25,  50,  or  85  ppm chloroform  for 7  hours/day,  5  days/week  for  6  months




produced centrilobular granular degeneration and focal  necrosis in their livers.




In subchronic studies, ingestion  of up to =100  mg/kg/day of chloroform produced




mild, transient histological changes  in the livers  of  rats and mice (Jorgenson




and Rushbrook,  1980),  ingestion of 145 or  190 mg/kg/day produced fatty change in




the livers of  mice  (Jorgenson and  Rushbrook,   1980), and  administration  of 410




mg/kg/day by gavage produced fatty change  and  necrosis in  the  livers  of rats.




Dogs treated subchronically  with  45 mg/kg/day by the oral route had histological
                                     5-29

-------
evidence of slight hepatic fatty change, with increasingly severe changes noted




at dosages of 60 and 120 mg/kg/day.




     In chronic oral studies,  rats  had minor histological change in their livers




and a decrease in relative liver weights  when  given 60 rag/kg/day of chloroform, 6




days/week, while the  livers  of mice were unaffected at  this  dosage.   Dogs had




some evidence of liver damage  (clinical  chemistry parameters) and an increase in




the number and size of fatty cysts in their lifetime when administered 15 or 30




mg/kg/day orally for 6 days/week.  The mechanism by which chloroform exerts its




hepatotoxic  effects has been widely investigated and efforts have been made to




identify the responsible metabolite(s).




     As  long ago as  1928,  it was suspected  that  the liver  damage  induced by




chloroform may be due not only to the chemical itself, but might be caused by a




degradation  product (Lucas,   1928).   The concept  that  chloroform  is excreted




unchanged was  disproved since a large number of studies  (Butler, 1961; Paul and




Rubenstein,  1963; Van  Dyke et al.,  1964; and Reid and Krishna, 1973) indicated




that the  tissue  necrosis  induced  by  chloroform is  associated with the covalent




binding of toxic metabolites and alkylation of tissue proteins.  Autoradiograms




have revealed that this binding occurs predominantly in the necrotic areas (Illet




et al., 1973).  It was also shown  by McLean (1970) that pretreatment of rats with




phenobarbital   (a   microsomal  enzyme   inducing  agent)  greatly  enhances  the




lethality of  chloroform.




     Brown  et  al.  (197*4) proposed  a  mechanism of  chloroform hepatotoxicity




implicating  a  free  radical metabolite which can react with glutathione  (GSH) (a




tripeptide which protects against hepatotoxicity),  diminishing GSH levels in the




liver.   According  to  this  hypothesis,   once  GSH levels are  depleted,  further




metabolism  would  result  in  the reaction  of  the  metabolite  with  microsomal




protein, and hence, necrosis.  This proposal  was  based on observations  in pheno-
                                      5-30

-------
barbital pre-treated rats anesthetized with chloroform, that hepatotoxicity was

enhanced and GSH levels were decreased  by  the  induction of microsomal enzymes.

Covalent binding of chloroform metabolites to  microsomal  proteins  In vitro was
also enhanced by enzyme induction,  an  effect prevented  by GSH.  Similar findings

were reported in mice by Ilett et  al. (1973), who found severe chloroform-induced

centrilobular necrosis in phenobarbital  pretreated mice, but only slight centri-

lobular damage in mice exposed only to chloroform.

     Thus, the hepatotoxicity of chloroform appears to depend on 1) the rate of

its biotransformation to produce  reactive metabolite(s), and 2) by the amount of

GSH available to conjugate with and thus inactivate the metabolite(s).

     The role of GSH in chloroform-induced hepatotoxicity was further studied by

Docks and Krishna  (1976),  who  found that only thoses  doses  of chloroform that

decreased liver  GSH caused liver  necrosis when  administered  to  phenobarbital

pretreated rats.

     More recently, Ekstrom et  al.  (1982) studied the mechanism of GSH depletion

by chloroform  in rats  pretreated  with phenobarbital.   The  synthesis of  GSH

proceeds via two enzymatic steps, the  first of which is rate limiting:
        glutamate + cysteine I-Slutamyl-cysteine,  dipeptide
                                 synthetase

In the presence of glycine, the reaction continues via GSH synthetase to produce

GSH.  When the soluble fraction from livers of rats sacrificed at various  times

after chloroform exposure was incubated  in  the  presence of these amino acids, it

was found that GSH synthesis was inhibited within 4-6 hours, while liver necrosis

was evident  only after 6 hours.   When  glycine was  eliminated  from the initial

part of the  incubation, the  dipeptide accumulated,  but at a lower  rate in the

presence of chloroform than in its absence.  Later  addition of glycine resulted

in GSH synthesis at a rate similar to  control values.  Thus, it  appears  that

chloroform, or rather a reactive  metabolite, inhibited GSH synthesis at the rate
                                     5-31

-------
limiting step  (i.e., the formation of  dipeptide  by Y-glutamyl-cysteine synthe-



tase).



     The biotransformation of chloroform (as discussed in Chapter i|) depends on



the activity of the microsomal drug metabolizing enzymes.  Substances that induce



these enzymes were shown to,  indeed, enhance the hepatotoxicity of chloroform as



evidenced by  increased  serum glutamic-pyruvic transaminase (SGPT)  levels  and



decreased hepatic glucose-6-phosphatase  activity  (Lavigne  and  Marchand, 1974).



An  inhibitor  of  the  drug  metabolizing  enzymes  SKF-525A,   however,  while



increasing   the   excretion  of    C-carbon   monoxide   in   rats  administered


14
  C-labelled chloroform, failed to diminish the  hepatotoxicity  of chloroform,



leading  these  authors to  conclude  that factors  other than metabolism  may be



involved.



     McMartin  et  al. (1981)  demonstrated that  altering the  cytochrome P-450



concentrations in the livers  of chloroform-exposed  rats also altered the hepato-



toxicity, as measured by the incidence  of  hepatic  lesions  and  by serum alanine



aminotransferase  activities.   Both  fasting  and  phenobarbital  pretreatment



increased the cytochrome P-USO content and liver  damage,  while  cadmium produced



the opposite effect.



     Theories  of  chloroform hepatotoxicity involve the  formation  of reactive



intermediates by liver enzymes.  How these intermediates exert their hepatotoxic



effect has been the subject of several  studies.   It has been suggested by Masuda



et al.  (1980) that,  based on  the chloroform-induced  indices of hepatotoxicity of



decreased microsomal glucose-6-phosphatase  activity and cytochrome P-'450 content



with increased hepatic malondialdehyde levels, the lipid peroxidation hypothesis



proposed for  carbon tetrachloride may  also apply  to the  case  of chloroform.



Qualitative  and  mechanistic  differences  of   hepatotoxicity  between  the  two



chemicals were noted, however.
                                      5-32

-------
     The interactive hepatotoxicity of  chloroform  and carbon tetrachloride was


studied by Harris  et  al.  (1982) who found  that,  while neither  chemical  alone


given at subthreshold dose altered SGPT activity,  hepatic triglyceride content,


or hepatic calcium content, when  given  together, these  chemicals increased the


toxic response in rats.  Administration  of either or both chemicals had no effect


on GSH levels or conjugated  diene  formation, but ethane expiration was increased


in rats  given  both  chemicals.   Diene conjugation  and ethane  expiration are


indices of lipid peroxidation.  Histopathological  changes were more severe from


the combinati^i than from either chemical alone.   Although the mechanism of the


hepatotoxic interaction  between chloroform  and carbon tetrachloride is unclear,


the authors suggested that there might be a  combined effect of phosgene formation


and lipid peroxidation initiation.


     It should be noted that  the  prevailing theories  implicate phosgene as the


major metabolite responsible  for  chloroform hepatotoxicity  (Reynolds  and Yee,


1967;  Sipes  et  al.,  1977;  Mansuy et  al., 1977;  Pohl  et  al.,  1977).   Other


potential toxic metabolites  discussed by Pohl  (1979) in a review of this subject


are a trichloromethyl  radical  and  dichlorocarbene;  however,  they are considered


less important than phosgene in this  regard.


     A  study  by Stevens  and  Anders   (1981) supports  the  phosgene-mediated

                                                                            in
mechanism.  The time course of  changes in SGPT  levels and covalent binding of   C


to proteins  was examined in  microsomal and  soluble fractions  from  phenobar-


bital-pretreated  rats  sacrificed  at  various   times   after   chloroform  or


  C-chloroform administration.  It was  found  that    C binding was  maximal at 6


hours while indices of liver damage peaked at 18 hours after chloroform exposure.


Further experiments were  performed in  which  diethyl  maleate  (a GSH  depletor)


treatment caused increased   C-binding  to soluble  and microsomal  fractions and


increased SGPT levels, perhaps by inhibiting the metabolism of phosgene to carbon
                                     5-33

-------
monoxide or stable conjugates.  Cysteine, which reacts with phosgene to produce




2-oxothiazolidine-M-carboxylic acid, had a  protective  effect.   Diethyl maleate




also diminished, but did not  eliminate the  deuterium  isotope  effect  on the GSH



dependent chloroform metabolism to carbon monoxide, which  would be  expected if




carbon monoxide formation occurred subsequent to phosgene production.  Thus, the




hepatotoxicity  of  chloroform can be altered  by altering  the  various  reaction




pathways  of  phosgene,  strongly  indicating  that phosgene  is  the toxic  inter-




mediate.




     From the above discussion,  it  appears  that, to be hepatotoxic, chloroform




must first be metabolized by microsomal  drug metabolizing  enzymes to an active




intermediate, probably phosgene,  which  in turn can react  by  various pathways,




depending on GSH levels, one  of which is the covalent binding to liver proteins




resulting in necrotic lesions.




5.3.2    Nephrotoxicity




     As  noted by Watrous and  Plaa (1972), the extensive body of research on the




hepatotoxicity  of  halogenated hydrocarbons has tended to  overshadow  the  fact




that some of  these agents  are also  nephrotoxic.  Earlier reports of chloroform




nephrotoxicity  include those of Heller  and Smirk  (1932),  who  found that  rats



anesthetized with  chloroform showed  a diminished ability to excrete a water load




given  prior  to  anesthesia,   and Knocher  and Mandelstam  (1944),  who  noted  that




chloroform injection produced a fatty infiltration of  the kidney.




     Renal  necrosis produced by the  oral  administration  of  chloroform  was




described by  Eschenbrenner  and  Miller  (1945a).  The necrosis, observed only in




male mice, involved portions of both proximal and distal convoluted tubules.  The



nuclei  of the epithelial cells were  often absent or  fragmented and the cytoplasm




was  coarsely granular  and  deeply eosinophilic.   The  glomeruli and collecting




tubules  appeared normal.
                                      5-34

-------
      Sex and strain differences in the sensitivity of mice  to chloroform nephro-




 toxicity were further studied by Deringer et  al.  (1953).  Exposure of strain C3H




 mice  to air containing =5 mg/£  of  chloroform  for 1,  2,  or 3 hours resulted in



 lesions of the kidneys of all of the males but in none of the females.  In animals




 dying within  1  day  after exposure,  epithelium of  the proximal  tubules  and




 portions of the  distal tubules were  generally  necrotic.  The  lumens  of those




 segments of tubules were dilated. The glomeruli were relatively unaffected.  The




 mice  dying  or sacrificed at later time intervals exhibited calcification in the



 necrotic area.




      Similar  lesions  were  found in  males  of  strains  C3H, C3Hf,  A,  and  HR.




 However, strains  C57BL,  C57BR/cd,  C57L,  and  ST were resistant  to  chloroform-



 induced nephrotoxicity.




      Comparable results  were reported  by Krus  and  Zaleska-Rutczynski (1970).




 Subcutaneous administration of chloroform to C3H/He male mice resulted in renal




 tubular necrosis, with death ensuing U-9  days later.   The lesions were calcified




 with  no evidence  of regeneration.  Female mice of this strain,  males and females




 of the C57BL/6JN and BN strains, and FI generation males  of the cross of female




 C3H/He with male C57BL/6JN mice survived the administration of chloroform (0.1 m?-




 of 0.05 g chloroform in 1 mS, ethyl laureate).  Additional studies were performed




 with  males  and females  of  this F   generation and the resistant BN  strain,  in



which animals were sacrificed at various time  after  chloroform administration.




All mice survived, while  all  female mice were  resistant, showing no  kidney




 lesions at  any time in the experiment.  Renal damage was morphologically apparent




in all male mice  by 12 hours,  but regeneration  developed  by day 4 and continued




until  the end of the  experiment.  It  was concluded  that  although all male mice




had tubular lesions,  the  ones surviving had tubules that did not  calcify  and  a




large  degree of renal  regeneration.
                                     5-35

-------
     Several investigators  have studied ]the influence of testosterone on chloro-

form-induced renal damage.   Eschenbrenner and Miller (1945b) performed  experi-

ments in which they saw extensive necrosis of portions of the proximal and distal

renal tubules  in  normal  male and  in  testosterone-treated castrated male  mice

following the acute oral  administration of chloroform.  However, no necrosis was

found after chloroform was administered to female mice or  castrated  male mice.

Continuing this line  of investigation, Culliford and Hewitt (1957)  reported the

following results:


     1) Adult male mice of  two strains  (CBA and WH) developed extensive  necrosis
        of the renal  tubules after exposure to low concentrations  of chloroform
        vapor  (7-10 mg/2, for 2 hours).   Adult females showed no  renal damage
        after equivalent  exposure.

     2) Adult females became fully susceptible to necrosis after treatment with
        androgens. The susceptibility of males was greatly reduced by treatment
        with estrogens.

     3) Castration removed  the susceptibility of the males  of  one strain, but did
        not  completely remove  it  in  another.   The residual susceptibility of
        castrates was abolished by adrenalectomy.

     *() Male mice under 11  days old were not susceptible to necrosis even after
        massive doses  of androgen.  Between 11 and 30 days, they were susceptible
        if given androgen.   Thereafter, they became spontaneously susceptible.

     5) Liver damage  occurred in nearly all exposed mice and was not correlated
        with sex hormone  status.

     6) Susceptibility could be  induced in gonadectomized mice by methyl testos-
        terone, testosterone propionate,  dehydroepiandrosterone,  progesterone,
        and large doses of cortisone acetate.

     Hill (1978) also  performed experiments demonstrating similar strain and sex

differences in chloroform-induced  renal toxicity.  The renal  toxicity of a fixed

oral dose  of chloroform  to castrated male  mice was increased  with increasing

doses of administered testosterone.  Plasma levels of testosterone in resistant

strains  tended to be lower  than  levels In  susceptible  strains.    Hill (1978)

conjectured that a testosterone may act  by sensitizing the renal proximal convo-

luted  tubules  to  chloroform  through   a   testosterone   receptor  mechanism.
                                      5-36

-------
Eschenbrenner and Miller (1945b), however, linked susceptibility to the nephro-




toxic action of  chloroform  to differences in kidney morphology  and physiology




induced by testosterone.



     Information on  the  nephrotoxicity of long-term chloroform  administration




has been presented in the section on Effects  of Chronic Exposure  to Chloroform.




Inhalation of  25,  50,  or 85 ppm  of  chloroform  7 hours/day, 5 days/week  for  6




months produced cloudy swelling of the renal tubular epithelium  in  rats.   Male




mice of certain sensitive strains  had increased incidences of moderate to severe




renal disease when treated orally  with chloroform at a dosage of 60 mg/kg/day,  6




days/week in a chronic study (Roe et al.,  1979).  Chronic oral administration of




30 mg/kg/day of  chloroform,  6 days/week, to  dogs  produced  an increase in  the




numbers of renal glomeruli affected by  fat deposition (Heywood et al.,  1979).




     The mechanism of the nephrotoxicity  of chloroform has been less extensively




studied than has that of hepatotoxicity.   Ilett et al. (1973) suggested that the




hepatotoxic metabolite produced in the  liver  is  transported  via the circulation




to the  kidney  where  it  exerts its nephrotoxic  effects.   More  recent  studies




(McMartin et al.,  1981; Kluwe and Hook,  1981) suggest that chloroform may also be




metabolized in the kidney, but by a different mechanism.




     McMartin et  al.  (1981)  altered  the  concentrations of cytochrome  P-450 by




fasting, by phenobarbital pretreatment, or by administration of cadmium to rats




given chloroform as a challenge.  Fasting increased cytochrome P-^50 concentra-




tions in both  liver  and  kidney, and chloroform-induced damage was  enhanced in




both organs of fasted animals.  Pretreatment with  cadmium decreased cytochrome




P-450 in livers but not kidneys  and significantly diminished liver damage due to




chloroform while having no effect on kidney damage due  to chloroform exposure.




Phenobarbital pretreatment resulted in increased liver but not kidney cytochrome




P-450 and, likewise,  chloroform-induced  damage  was enhanced in  livers  but  not
                                     5-37

-------
kidneys.  Thus, pretreatments that altered hepatic cytochrorne P-450 levels had no




effect  on chloroform-produced  renal effects,  suggesting that a metabolite  of




chloroform, which is responsible for kidney damage,  is produced in  the kidney.



     The mechanism of chloroform nephrotoxicity was  also investigated by Kluwe




and Hook  (1981), who found  no difference  between  nephrotoxicity and  hepato-




toxicity with respect to the  effects of microsomal enzyme inhibitors and diethyl




maleate.   Mice  were injected  intraperitoneally  with chloroform either  before




piperonyl  butoxide  or   SKF-525-A   administration  or  after  diethyl  maleate,




piperonyl  butoxide or  SKF-525A exposure.    Although it  inhibits  microsomal




enzymes,  SKF-525-A, administered either  before or  after  chloroform administra-




tion,  did not  reduce the  hepatotoxicity or the nephrotoxicity  of chloroform.




This  is  consistent  with  a  similar finding  by Lavigne and Marchand  (1974).




Piperonyl butoxide, when given before chloroform, protected against toxicity in




both organs, but  when given  after  chloroform,  did  not.   This finding indicated




that  an  enzymatic  step in  the metabolism of chloroform by  both  organs  was




inhibited.  The effect of diethyl maleate was  to  enhance  the toxicity of chloro-




form in both organs.  Thus, the mechanism  of  chloroform nephrotoxicity appears to




be similar to that of hepatotoxicity with respect to these substances.




5.4  FACTORS MODIFYING THE TOXICITY OF CHLOROFORM




     From the preceding  discussion, it is evident that the alterations of micro-




somal  enzyme  activity  or  hepatic  GSH levels  influence  the severity of toxic




effects  induced  by a given  amount of  chloroform.    It  follows then  that  many




factors could alter chloroform toxicity by affecting these parameters or acting




through other mechanisms.  These substances are of  interest  because they fall




into  categories   of  accidental or  intentional exposure  to  humans.   Alcohol,




dietary components, pesticide, and steroids  are some  of the substances which are




discussed below.
                                      5-38

-------
5.^.1   Factors that Increase the Toxicity.  The effect of ethanol pretreatment




on  chloroform-induced  hepatotoxicity in  mice was  studied by  Kutob and  Plaa




(1962a).   An intoxicating dose (5 g/kg) of  ethanol  was  administered orally to



mice daily for 15 days initially, with a systematic shortening of the duration to




a single exposure.   A  challenging  dose of  chloroform (0.08 mil/kg) was adminis-




tered subcutaneously either 12, 15, or 24 hours after ethanol  treatment.   Liver




dysfunction  was  measured  by  prolongation  of  phenobarbital  sleeping  time,




bromsulphalein  (BSP)  retention,  liver  succinic dehydrogenase  activity,  and




histological  examination.   Regardless of the ethanol  treatment  period,  pheno-




barbital sleeping  time was significantly increased  in mice receiving  ethanol




followed by chloroform  when compared with mice receiving either substance  alone.




Similar findings were found for BSP retention.   The  iri vitro  succinic dehydro-




genase activity was  significantly  reduced  by ethanol  pretreatment  followed by




chloroform administration 12 or 24  hours, but not 48 hours, later, when compared




with activities from mice receiving only  chloroform.   Histological changes were




seen in the livers of mice given ethanol  15 hours to  4 days prior to chloroform




challenge,  while mice receiving either chemical alone had morphologically nonnal




livers.   It was  also  determined  that the  ethanol  treatment increased  liver




triglyceride content, with a maximum at  15  hours, and that ethanol pretreatment




significantly increased  the  concentration  of chloroform  in the livers  with  a



maximum at  12 hours after chloroform challenge.  From  these results, it was noted




that a  single dose  of ethanol was  just  as effective  as multiple  doses.    A




mechanism  was proposed  for  the  ethanol  enhanced  chloroform-induced hepato-




toxicity in which  ethanol  increases  liver  lipid  content   (as  evidenced  by




increased triglycerides) resulting  in increased concentrations  of chloroform to




be metabolized in  the liver.
                                     5-39

-------
     In support  of this  mechanism is  the observation  that  oral  isopropanol




pretreatment for 5 days (0.3 m£/100 g for 2 days and 0.15 mil/100  g  for  3 days)




followed 12 hours later by  5 daily inhalations of chloroform  (5000 ppm first day,



2500 ppm on the  next four days), 2 hours/day led to severe fatty infiltration of




the liver.  Chloroform alone increased the pool of triglycerides (Danni  et al.,




1981).




     In contrast,  Sato et  al.  (1980) studied  the mechanism by  which  ethanol




enhances hydrocarbon metabolism, including that  of  chloroform.   Rats  ingesting




ethanol in their drinking  water for  3 weeks  were sacrificed 10 hours  after the




final exposure.  Control  rats  were given isocaloric glucose solutions.   Liver




microsomal enzyme systems  were prepared  and  liver  protein  and  cytochrome P-450




contents analyzed, the increased contents being indicative  of microsomal enzyme




synthesis in response to alcohol.   When chloroform  was added as a substrate, its




metabolism was  enhanced  by 6 times, much more  than could  be accounted  for by




enzyme induction alone.  Microsomes prepared from rats that were withdrawn from




ethanol 24  hours prior  to sacrifice  did  not  show enhanced activity.    In a




subsequent study (Sato et al.,  1981), rats receiving a single gavage dose of 0,




2,  3,  4,  or  5  g/kg  ethanol  were  sacrificed  18  hours  later.    The ln_ vitro




metabolism of chloroform by rnicrosomes  prepared from these rats was enhanced very




little at 2 g/kg, slightly more at  3 g/kg, and dramatically at 4 g/kg.  At  5 g/kg,




however, enhanced enzyme activity  was  no greater than  at 3 g/kg ethanol.  When



ethanol was added directly  to the incubation system,  the metabolism of chloroform




was inhibited.  Rats receiving 5 g/kg ethanol retained relatively large amounts




in the blood and liver, while those receiving  4  g/kg retained  almost  none.  If



the  ethanol  remaining in  the rats  exerted  an inhibitory  effect on  enzyme




activity,  then  microsoraal  enzymes prepared  from 5 g/kg ethanol-treated rats,




mixed with the soluble fraction from control rats, should  show increased activity
                                      5-40

-------
when compared with  both  microsomal and soluble  fractions  from ethanol-treated




rats.  This was found to be the case.




     Based on these results and studies  on metabolism of other hydrocarbons in



vitro and in vivo, Sato et al.  (1981)  suggested  that ethanol is both a stimulator




and an  inhibitor  of drug metabolizing  enzymes,  depending on  how  much ethanol




remains in the body and thus how much time has  elapsed since ethanol ingestion.




Thus, when  ethanol is first  ingested,  it  acts  as  a competitive  inhibitor of




microsomal enzyme activity,  but as it disappears  from the  body,  an optimum for




stimulation may be reached and metabolism enhanced.  It was postulated that since




the metabolism of chloroform was enhanced to a much  greater  extent than can be




explained  from  enzyme  induction   alone,  perhaps ethanol  modifies the  enzyme




activities  by  other mechanisms such as modification of membrane  properties,




allosteric effects, or by displacement of substrate  already bound.




     Polybrominated biphenyls  (PBBs) have  also  been found  to  potentiate  the




toxicity of chloroform (Kluwe and Hook, 1978) .  Mice were fed diets containing 0,




1, 25,  or 100 ppm PBB for 14 days.   One day  before sacrifice, the mice were given




a single intraperitoneal  injection of 0,  0.5, 2.5, 5.0,  or 50 ^1/kg chloroform.




PBB enhanced  the  toxicity of chloroform  in both the liver  and the kidney as




evidenced by results of blood urea  nitrogen  (BUN) and  serum glutamic oxaloacetic




transaminase (SCOT) determinations and  by  inhibition of p-aminohippuric  acid




(PAH) uptake by renal slices.   PBB also reduced the LDp... of chloroform  in these




mice and the deaths were  attributed to hepatic  necrosis.  Since PBBs were known




to induce the drug metabolizing enzymes,  their effects on  chloroform were assumed




to be due to enhanced chloroform metabolism.




     Steroids appear to play a  role in the potentiation  of chloroform toxicity,




especially in the kidney as seen from  the sex-related differences in  the  response




of mice (Eschenbrenner and Miller,  1945a;  Deringer et al.,  1953)  and by experi-
                                     5-41

-------
rnents   involving   testosterone   administration   to   castrated   male   mice




(Eschenbrenner  and Miller,  19l*5b;   Culliford and  Hewitt,  1957;  Hill,  1978)




discussed previously.  Clemens et al. (1979) further studied this phenomenon in



castrated male and  intact  female mice.   Dose-dependent  testosterone sensitiza-




tion of renal tubules to a fixed  dose of chloroform was observed  in castrated




males with the response ranging from kidney dysfunction to death at high doses.




The  androgenic   progestin,  medroxyprogesterone acetate,  enhanced  chloroform-




induced kidney damage in both castrated males and intact females.  Progesterone




or hydrocortisone potentiated chloroform  toxicity in DBA/2J castrated male mice,




but not in the C57BL/J6  strain males nor  in any of the females.   The mechanism by




which the androgens exerted their  potentiation may have  been  mediated through




strain specific androgen receptors of  the proximal convoluted tubular  cells.  The




mechanism for the potentiating action of the other steriods was less clear.




     The potentiation of chloroform  toxicity by ketones and ketogenic substances




has been studied extensively in recent years.  Hewitt et al. (1979) and Cianflone




et al. (1980) found that while pretreatment  of mice with the insecticide, kepone




(a ketone), enhanced the liver damage caused by chloroform exposure, the struc-




turally related  mirex (a non-ketone)  did not.  Other ketones were compared for




their ability to enhance the hepatotoxic  and  nephrotoxic action of chloroform in




rats with the following results:  methyl  n-butyl ketone and 2, 5-hexanedione were




the most potent  enhancers,  followed by acetone, followed by n-hexane (a ketogenic




chemical)  (Hewitt  et al.,  1980).    Jernigan and  Harbison  (1982)  studied the




potentiation  by  2,5-hexanedione  of  chloroform hepatotoxicity in  mice  with




specific reference  to sex differences.  Female mice were more susceptible to the




dose-dependent  enhancement  of  chloroform  hepatotoxicity  as  well  as to  the




dose-related increase in hepatic microsomal  enzymes.  Pretreatment of male mica




with 2>5-hexanedione potentiated the toxicity of deuterated chloroform but to a
                                      5-42

-------
lesser extent than  chloroform;  however,  this deuterium isotope  effect  did  not
occur in female mice in_ vivo.   Phenobarbital  pretreatment did elicit a deuterium
isotope effect in female mice iri vivo, suggesting that 2,5-hexanedione pretreat-
ment  altered  chloroform metabolism  by  a different  mechanism than  did  pheno-
barbital.  Jernigan  and Harbison (1982) speculated that perhaps female mice have
greater microsomal enzyme activities, different membrane properties, or perhaps
produce a different  reactive metabolite of chloroform than do males.
     The mechanism  of  ketone  potentiation  of  chloroform-induced  hepato-  and
nephrotoxicity was  also investigated  by Branchflower  and  Pohl (1981)  using
methyl n-butyl  ketone  (MBK).   Male  rats were pretreated with MBK  followed  by
chloroform administration.   The metabolism  of  chloroform by liver  and  kidney
microsomal enzymes  and the toxicity to these  organs  were examined.   Control
experiments were conducted in which rats were  either not  pretreated, not given
chloroform, or given deuterated chloroform (CDC1 ) instead of chloroform.  MBK
increased  cytochrome  P-450  levels  and  NADPH-dependent  cytochrome  reductase
activity in liver microsomes,  while  having no effect on  renal  levels of these
microsomal components.   MBK  pretreatment doubled  the rate  of metabolism  of
chloroform to  diglutathionyl   dithiocarbonate  (GSCOSG)  in microsomal  prepara-
tions, and more  GSCOSG  was excreted into the bile of the pretreated animals when
compared with rats receiving only  chloroform.  The  amount  of  GSCOSG in bile  was
less in MBK-CDC1,. treated animals.   GSH  levels were  significantly decreased  by
MBK treatment and this decrease was enhanced  following chloroform exposure,  and
to a lesser extent,  following CDC1,, exposure.   Rats  pretreated with MBK followed
by chloroform had greatly elevated  levels of SGPT associated with liver necrosis
and  significantly greater BUN  levels associated  with renal cortical  tubule
lesions  over  the control  groups.    A  mechanism was proposed whereby MBK,  by
increasing cytochrome  P-450 levels,  enhanced the metabolism  of chloroform  to
                                     5-43

-------
phosgene.  Furthermore, according to the hypothesis,  the phosgene was converted




to GSCOSG through GSH,  levels of which were diminished by MBK,  because the more




phosgene formed, the more GSH was depleted  in the reaction.  The  results with



CDC1.J indicated that C-H bond was involved in the mechanism.  Although MBK also




potentiated chloroform  toxicity  to  the kidney, a different mechanism may have




been  involved  since  renal  cytochrome  P-450  and renal  GSH  levels were  not




affected.




5.4.2   Factors that Decrease the Toxicity.   As discussed above, the experiments




of Sato  et  al.  (1980,  1981)  indicate that ethanol is both  a stimulator  and an




inhibitor  of  microsomal  enzymes,  and  hence,  of   chloroform  metabolism  and




toxicity, depending upon the length of time after ingestion.




     Disulfiram  and its metabolites have  also  been studied with respect to their




protective  effects of chloroform'-induced hepatotoxicity (Scholler, 1970;  Masuda




and Nakayama,  1982).   Disulfiram is used in treating chronic alcoholism and is




metabolized to  carbon disulfide and diethyldithiocarbamate  (a herbicide)  (IARC,




1976).    Disulfiram,  a known  inhibitor  of  the  microsomal drug  metabolizing




enzymes,  given  to rats  prior to  chloroform anesthesia completely prevented the




elevated SGPT activity and liver necrosis observed in rats administered chloro-



form alone  (Scholler et al.,  1970).   More recently,  Masuda  and Nakayama  (1982)




studied the effects of diethyldithiocarbamate  and carbon disulfide pretreatment




in mice challenged with chloroform.  Both substances had a protective effect as




measured  by SGPT activity,  liver calcium content, and centrilobular necrosis.




Diethyldithiocarbamate and carbon disulfide  decreased the activity of the drug




metabolizing  enzymes rn vivo and iii vitro,  but  only in the presence of NADPH,




indicating  that these substances must first  be metabolized before exerting their




inhibitory  effect on chloroform metabolism.  Gopinath and Ford (1975) also found




that dithiocarbamate and  carbon  disulfide protected  against chloroform hepato-

-------
toxicity in rats, and the  effect was  presumed  to  be due to suppression  of  the




drug metabolizing enzymes.




     Dietary components  can also alter  the toxicity of  chloroform.   It is  a



widely held opinion that  low  protein  content of the  diet  decreases  microsomal




enzyme activities, while a high protein diet increases  the activities (McLean and




McLean, 1969; Nakajima et al., 1982).   If this is the case, a  low  protein  diet




should  protect  against  chloroform  hepatotoxicity  by  inhibiting  the  enzymes




responsible  for  chloroform metabolism.   It was  found,  however, that  protein




depletion did not alter the toxicity  of  chloroform  in rats given a single  oral




dose (McLean and McLean, 1969; McLean, 1970).  If  pretreated with phenobarbital




or NDDT to induce microsomal enzymes,  rats maintained on a standard  diet were no




more  susceptible  to  chloroform-induced  liver damage  that  were   pretreated,




protein-depleted rats  (McLean, 1970).




     More recently, Nakajima  et  al.  (1982)  studied the individual effects  of




protein, fat, and  carbohydrate on the metabolism of chloroform in relation to its




toxicity in  male  rats.    Test  diets were varied with respect  to carbohydrate,




protein, or fat  while  maintaining isocaloric contents.  Microsomal  enzymes were




prepared and chloroform was  added as  a substrate.  The following  results were




obtained:  decreased food  intake increased  liver microsomal enzyme  activities;




decreased sucrose content in the  diet  increased the  metabolic  rate;  varying  the



protein and  fat  content,  while holding  the sucrose  content  constant, had  no




effect on  the metabolic rate; a carbohydrate-free  diet,  which  contained  high




protein and high  fat,  accelerated the rate of chloroform metabolism  almost  as




much as  1  day  of food  deprivation.    The authors  concluded that it  is  a  high




carbohydrate content, rather than a low protein  content, which is responsible  for




the  decreased  microsomal  enzyme metabolism  of  chloroform  and,   hence,   its




toxicity.

-------
5.5  SUMMARY;  CORRELATION OF EXPOSURE  AND EFFECT




     The purpose of  this  section is to delineate dose-response relationships for




the systemic toxicity of chloroform.



5.5.1   Effect of Acute Inhalation Exposure.   The adverse  effects  on  humans of




inhaling high  concentrations  of chlorof^opi  have been well  documented in  the




course of its use as an anesthetic.  Studies  that define the threshold region of




exposure for such effects in humans are, however, sparse at  best.   Experiments




involving subchronic exposure of several species of  animals  give  some informa-




tion on toxicity thresholds for renal and hepatic  effects, but little for CNS and




none for cardiovascular effects.




     The only experimental studies  conducted with humans  (Lehmann  and  Hasegawa,




1910; Lehmann  and Schmidt-Kehl,  1936)  involved relatively short exposures  and




subjective responses.   The  results  of these studies  indicate  that  the odor of




chloroform can be perceived at about 200 ppm.  Subjective CNS effects (dizziness,




vertigo) apparently  did not  occur at 390 ppm during a  30-minute exposure but were




perceived at about 900 ppm after 2-3 minutes of  exposure.   Subjects exposed to




1400 ppm for  30 minutes experienced  tiredness and headache  in addition to the




above CNS symptoms.   The threshold for  "light  intoxication"  was about 4300 ppm




(20  minutes).   An exposure duration of 30 minutes  or less  is  insufficient to




achieve  pulmonary  steady state  (or  total   body  equilibrium,  Section  4.2.3).




Hence, longer exposures at these concentrations would be  expected  to cause more




severe effects.




     Chloroform concentrations used for the induction of anesthesia ranged from




about 20,000-40,000 ppm (NIOSH, 1974;  Adriani,  1970)  and for the maintenance of




anesthesia  ranged  from  1500  ppm  (light   anesthesia)  to  15,000  ppm  (deep




anesthesia) (Goodman and Oilman, 1980).  Continued exposure to 20,000 ppm could




result  in  respiratory  failure,  direct  depression of  the myocardium,  and  death
                                      5-46

-------
(Section 5.1.1).  Levels of exposure sufficient to produce anesthesia have also
caused cardiac  arrythmias  and  extrasystoles (Kurtz et al.,  1936;  Orth et al.,
1951) and  hepatic  necrosis and fatty  degeneration (Goodman and Oilman,  1980;
Wood-Smith and Stewart, 1964).
     Data from acute animal exposures tend to show similar CNS effects at roughly
the same levels of exposure that produced these  effects  in humans  (Lehmann and
Flury, 19^3).   In addition, some data  on the threshold for hepatic effects has
been obtained for mice. Kylin et al. (1963), in experiments with female mice of
an  unspecified  strain,  found  that  single,  4-hour   exposures  to  chloroform
produced mild hepatic  effects (increased incidence of  moderate  fatty infiltra-
tion) at  100 ppm.    At  200 ppm,  in addition  to fatty  infiltration,  hepatic
necrosis and increased serum ornithine carbamyl  transferase activity occurred.
(An elevation in serum levels of this enzyme indicates liver damage  according to
Divincenzo and  Krasavage,   1974.)   Further  increases in fatty infiltration,
necrosis, and serum  enzyme activity were observed at  400 and 800 ppm.   These
effects appeared to  be reversible because the extent of change  was less severe 3
days after exposure than it was  1 day after  exposure.
     Damage to  the kidneys  of  male mice of  sensitive  strains (e.g.,  C3H)  has
occurred at exposure levels as  low  as  5 mg/£ (1025 ppm) for 1  hour  (Deringer et
al., 1953).  The damage consisted of necrosis of the epithelium  of  the proximal
tubules.
5.5.2   Effects of  Acute Oral Exposure.  Dose-response data for acute oral  expo-
sure of humans  to  chloroform is  limited to case  reports.   A  fatal dose  of as
little as 1/3 ounce (10 m£) was reported (Schroeder,  1965).
     A variety of dose-response data is  available for acute oral  administration
of chloroform to animals.   Single doses that were sufficient to adversely affect
kidney function (measured as excessive  loss of  glucose  and/or  protein in  the
                                     5-47

-------
urine in male mice) ranged from 89-149 ing/kg in sensitive and relatively Insensi-


tive strains (Hill,  1978).  At single oral  doses of 1071 mg/kg,  but not 756 or 546


ing/kg chloroform, increases in organ weights and mild  to moderate  lesions  were

observed in the livers and kidneys  of  Sprague-Dawley  rats  (Chu et  al.,  1982a).


In male  B6C3F1 mice, renal necrosis occurred  after 20 mg/kg  and focal  tubular


regeneration occurred after 60 or  240 mg/kg but not after 15 mg/kg (Reitz et al.,


1980).  A low observed adverse-effect  level (LOAEL) for hepatic effects  in mice


can be identified from the study of Jones  et al., 1958, in which 30 mg/kg caused


midzonal fatty infiltration.   Doses  in  the range of 133-355 mg/kg (Jones  et al.,


1958; Reitz et al.,  1980) represent a PEL (Frank-Effect-Level)  for hepatic damage


(including  centrilobular  necrosis) in mice.    According  to  Torkelson   et  al.


(1976), rats given "as little as" 250  mg/kg chloroform "showed adverse effects"


on liver and kidney  as determined by  gross  pathological  examination.   Reported


oral LD   values for mice ranged from   119-1400 mg/kg,  depending on  sex,  strain,


and age  (Kimura et al., 1971; Hill,  1978;  Bowman et al.,  1978).  For rats,  LD™


values of 908-2000 mg/kg have been reported (Chu et al., 1980; Torkelson et al.,


1976).  The lethal dose studies included both 24-hour and delayed deaths.


5.5.3    Effects of Dermal  Exposure.   Chloroform is  irritating to the  skin.  It


has been reported to cause degenerative changes in the renal tubules of rabbits


exposed dermally to high doses under extreme conditions (1-3-98 g/kg body weight


for 24  hours  under  an  impermeable  plastic cuff)  (Torkelson et  al., 1976).   In


humans,  toxicity  from  dermal exposure is probably not  important  in comparison


with other routes.


5.5.4    Effects  of  Chronic  Inhalation Exposure.    Limited information  on the


effects of long-term intermittent  exposure of humans or animals to chloroform is


available.  A  study  involving a  small number  of workers (Challen et al., 1958)

                                                             o
indicates that  long-term  exposure to  20-71 ppm (98-346  mg/m ) for a 4-8 hour
                                      5-48

-------
workday, with occasional brief excursions to =1163 ppm,  may represent a LOAEL for
symptoms of  CNS  toxicity.   No  evidence of liver damage or other organic lesion
was  detected  by  physical examination  and  clinical  chemistry tests.   A single
report linking liver enlargement and viral  hepatitis to occupational exposure to
10-200 ppm chloroform  (Bomski et  al.,  1967)  is flawed by the  apparent  lack of
suitable  controls.    The available  data  do  not  define a  NOAEL (no-observed-
adverse-effect-level) or NOEL (no-observed-effect- level) for humans.
     Experiments with several species  of  animals  (Torkelson et al.,  1976)  give
some information regarding the threshold region for hepatic and renal effects of
inhalation exposure  to  chloroform  (Table 5-3).   The  animals were  exposed to
chloroform 5 days/week for 6 months.  Exposure  to 25 ppm of chloroform for up to 4
hours/day had no adverse effects in male rats  as judged by organ and body weights
and  probably  the gross and  microscopic appearance  of at  least  the  liver  and
kidneys, although the authors were not  explicit about the latter.   Exposure to 25
ppm for 7 hours/day,  however, produced  histopathologic  changes in the livers  and
kidneys of male but  not  female rats.  These changes were characterized as lobular
granular  degeneration  and  focal  necrosis throughout  the  liver  and  cloudy
swelling of the kidneys.  The hepatic and renal effects  appeared to be reversible
because rats  exposed in  the same way, but given a 6-week "recovery period" after
the  exposure  period, appeared  normal   by  the  criteria  tested.    Increasingly
pronounced changes were  observed in the livers  and kidneys of both sexes  of  rats
exposed to 50 or  85 ppm.  Hematologic, clinical chemistry, and urinalysis values,
tested at the two higher levels  of exposure, were "within normal  limits."
     Similar  experiments with guinea pigs and  rabbits gave somewhat inconsistent
results.   Histopathological lesions were observed in liver  and kidneys  of  both
species at  25 ppm but not at 50 ppm in either species and  not in guinea pigs  even
at 85 ppm (Torkelson et al.,  1976).
                                     5-49

-------
     The experiments of Torkelson et al.  (1976) indicate  that  subchronic  expo-




sure to  25  pprn (123 mg/m^),  U  hours/day, 5  days/week  represents  a NOAEL  and




exposure  to  ?5 ppm,  7 hours/day,  5  days/week respresents  a  LOAEL for  rats.




Guinea pigs and rabbits may be slightly less  sensitive.




5.5.5   Effects of  Chronic  Oral  Exposure.  Little dose-response data  for oral




exposure of humans to chloroform appear to be  available  (Chapter 5).   A single




controlled study has been performed.  In this  study, subjects were exposed to 70




or 178 mg of chloroform/day  (-1 or 2.5 mg/kg/day assuming 70 kg body weight) for




at least  1 year (DeSalva et al.,  1975).  Neither liver function tests nor blood




urea  nitrogen  determinations  (a measure  of  kidney function)  revealed statis-




tically  significant differences  between  exposed  and  control  subjects.   Case




reports  involving  abuse  of  medicines  containing  chloroform  (Wallace,  1950;




Conlon,  1963) are not adequate for risk assessment because of the small numbers




of patients, exposure to other agents, and imprecise estimates of intake.




     Subchronic and chronic toxicity experiments with rats, mice, and dogs, when




considered  together (Table 5-4),  do not  clearly establish a NOAEL  or  NOEL.




Although  no adverse  effects  were  observed  in four strains  of  mice  given 17




mg/kg/day  of chloroform,  6 days/week  for 2  years  (Roe  et al.,  1979),  at the




lowest dose level tested (i.e.,  15 mg/kg/day,  6 days/week for 7.5  years) in dogs,




chloroform treatment was associated  with an elevation in SGPT  in some  but not all




of  the  other tested clinical chemistry indicies  of hepatic  damage  (Heywood et




al.,  1979).  The livers of dogs treated  with chloroform at this dosage  level had




larger and more numberous "fatty cysts"  than were  found in controls.  These fatty




cysts  consisted  of aggregations  of  vacuolated   histiocytes.   No effect on




survival,  growth,  organ  weights, gross  and  histological appearance  of other




organs,  or  hematologic  or  urinalysis values  was observed at this dosage  level.




Hence  15 mg/kg/day  (6  days/week) represents a LOAEL for dogs for effects  on the
                                      5-50

-------
liver.  Chronic oral administration of 60 mg/kg/day of chloroform (6 days/week)




was associated with slight  hepatic changes in rats (Palmer et al., 1979) and with




increased incidences of moderate to severe renal disease in male mice of sensi-



tive strains (Roe et al.,  1979).




     None of  the  three species  tested in long-term  experiments  appeared to be




markedly more sensitive to the toxicity  of chloroform  than any  other;  dogs may




have been slightly more  sensitive.   There were  considerable  differences among




strains of mice in the  sensitivity of the males to chloroform  nephrotoxicity, as




had also been observed in acute toxicity experiments.




     As would  be  expected, dosages that produced little  or  no histologic  or




clinical chemistry evidence  of toxicity when  given  subchronically  (15  and  30




mg/kg/day;  rats, dogs) resulted in greater evidence  of toxicity  when given for




longer periods of time  (Palmer et al., 1979; Heywood et al., 1979).  The response




to chloroform in the long-term studies may have been modified  by the presence or




absence of intercurrent respiratory and renal disease,  but no consistent pattern




is obvious from an inspection of the data in Table 5-1*.




5.5.6   Target  Organ  Toxicity.    Target organs  characteristic of the  acute




toxicity of chloroform  are  the central nervous system,  liver,  kidney, and heart.




For chronic exposure to chloroform, characteristic target organs are the liver




and kidney,  and possibly the central nervous system.  Some dose-response data are



available for the  toxicity  of chloroform  to the  liver, kidney,   and central




nervous system; these  data are  summarized in Table  5-5 by target   organ.   The




studies from  which these  data are drawn are discussed more  fully  elsewhere in




Chapter 5,  but  a  comparison  on the basis of  endpoinb  (target organ)  was also




considered to be useful.




     Manifestations of liver  damage include  centrilobular necrosis, vacuoliza-




tion,  disappearance  of glycogen,  fatty  degeneration  and swelling  (Groger and
                                      5-51

-------
                                                     Table  5-5

                                        Target Organ  Toxicity of  Chloroform
Target
Organ
Route and
Type of
Exposure
                                   Species
                                                 Dose or
                                                 Exposure
Effect on
Target Organ
                                                                                                   Reference
liver     inhalation,  acute
          (surgical anesthesia)
                                   human
liver     inhalation,  acute
liver     inhalation,  acute
liver     inhalation, chronic
liver     inhalation, chronic
liver     inhalation, chronic
                                   mice
                                   mice
                         rats
                         rats
                         rats
                                        induction =
                                        20,000-40,000  ppm x
                                        a  few minutes,  plus
                                        maintenance  =  1500-
                                        15,000 ppm x variable
                                        duration
                                        100  ppm x 4 hours,
                                        single exposure

                                        200  ppm x 4 hours,
                                        single exposure
                                                 25 ppm,  4 hours/day,
                                                 5 days/week  x
                                                 6 months

                                                 25 ppm,  7 hours/day,
                                                 5 days/week  x
                                                 6 months

                                                 50 or  85 ppm,  7  hours/
                                                 day, 5 days/week x
                                                 6 months
                                                                          necrosis,  fatty degen-
                                                                          eration in some
                                                                          patients
                                                                           fatty  infiltration
                                                                          necrosis,  fatty  infil-
                                                                          tration,  increase  in
                                                                          SOCT

                                                                          no  effect
                                                                           lobular  granular
                                                                           degeneration,  focal
                                                                           necrosis

                                                                           marked centrilobular
                                                                           granular degeneration
                         NIOSH,  1974;
                         Goodman and
                         Gilman,
                         1980;
                         Wood-Smith
                         and Stewart,
                         1964

                         Kylin et al.,
                         1963

                         Kylin et al.,
                         1963
                         Torkelson
                         et al., 1976
                         Torkelson
                         et al., 1976
                         Torkelson
                         et al., 1976

-------
                                                     Table  5-5

                                   Target Organ Toxicity of  Chloroform  (cont.)


£
oo
Target
Organ
liver
liver
liver
liver
Route and
Type of
Exposure
oral, acute
oral, acute
oral, acute
oral, acute
Dose or
Species Exposure
mice 30 mg/kg bw;
single dose
mice 133 mg/kg, single
dose
mice 355 mg/kg, single
dose
mice 60 mg/kg, single
Effect on
Target Organ Reference
fatty infiltration Jones et al . ,
1958
massive fatty infiltra- Jones et al.,
tion and severe necrosis 1958
massive fatty infiltra- Jones et al.,
tion and severe necrosis 1958
no effect Reitz et al.,
liver     oral, acute
mice
liver     oral (drinking water)    rats
          subchronic
liver     oral (drinking water)    mice
          subchronic
dose

240 mg/kg, single
dose
               20, 38, 57, 81, or
               160 mg/kg/day x
               90 days
               64 or 97 mg/kg/day x
               90 days
hepatocellular necrosis
and swelling; inflamma-
tion

transient hepatosis
(at 30 and 60, but
not at 90 days)
                         transient centrilobular
                         fatty change  (at 30
                         and 60 but not at
                         90 days)
1980

Reitz et al.,
1980
Jorgenson
and
Rushbrook,
1980

Jorgenson
and
Rushbrook,
1980

-------
                                                        Table 5-5


                                       Target Organ Toxicity of Chloroform  (cont.)
ui
-Cr
Target
Organ
liver
liver
liver
liver
liver
liver
liver
liver
Route and
Type of
Exposure Species
oral (drinking water) mice
subchronic
oral (gavage) rats
subchronic
oral (capsule), dogs
subchronic
ora] (capsule), dogs
subchronic
oral (capsule), dogs
subchronic
oral (capsule), dogs
subchronic
oral (gavage), rats
chronic
oral (gavage), mice
chronic
Dose or
Exposure
145 or 190 mg/kg/day x
90 days
410 mg/kg/day, 6 days/
week x 13 weeks
30 mg/kg/day, 7 days/
week x 13 weeks
45 mg/kg/day, 7 days/
week x 13 weeks
60 mg/kg/day, 7 days/
week x 18 weeks
120 mg/kg/day, 7 days/
week x 12 weeks
60 mg/kg/day, 6 days/
week x 80 weeks
60 mg/kg/day, 6 days/
week x 80 weeks
Effect on
Target Organ
fatty change
fatty change and
necrosis
no effect
slight fatty change
fatty degeneration,
increase in SCOT and
SGPT
fatty degeneration,
jaundice, increase
in SCOT, SGPT, bili-
rubin
minor histological
changes and decrease
in relative liver weight
no effect
Reference
Jorgenson
and
Rushbrook,
1980
Palmer et al . ,
1979
Heywood
et al., 1979
Heywood
et al., 1979
Heywood
et al., 1979
Heywood
et al., 1979
Palmer et al . ,
1977
Roe et al . ,
1979

-------
                                                     Table 5-5

                                    Target Organ Toxicity of Chloroform (cont.)
Target
Organ
liver
Route and
Type of
Exposure
oral (capsule),
Species
dogs
Dose or
Exposure
15 or 30 mg/kg/day,
Effect on
Target Organ
increases in SGPT and
Reference
Heywood et al . ,
          chronic
liver     oral,  chronic
liver     dermal, acute
kidney    inhalation, acute
kidney    inhalation, chronic
kidney    oral, acute
                                   humans
                                   rabbits
                                   mice,  males
                                   of sensitive
                                   strains

                                   rats
                                   mice,  males
                                   of sensitive
                                   strains
                                                  6 days/week x 7.5
                                                  years
2.5 mg/kg/day for
>_1 year

3.98 g/kg x 24 hours
under plastic cuff,
single exposure

5000 ppm, 1 hour,
single exposure
25, 50, or 85 ppm,
7 hours/day, 5 days/
week x 6 months

89 mg/kg, single dose
                                                                           other serum indicators
                                                                           of hepatic damage,
                                                                           increase in size and
                                                                           number of fatty cysts
                                                                           (vacuolated histiocytes)

                                                                           no effect
                                                                           no macroscopic
                                                                           pathologic changes
                                                                           necrosis and calcifi-
                                                                           fication of tubular
                                                                           epithelium

                                                                           cloudy swelling of
                                                                           tubular epithelium
                                                                           loss of glucose or
                                                                           protein in urine
                                                   1979
De Salva
et al.,  1975

Torkelson
et al.,  1976
Deringer
et al.,  1953
Torkelson
et al., 1976
Hill, 1978

-------
ON
                                                         Table 5-5




                                        Target  Organ Toxicity of Chloroform (cont.)
Target
Organ
kidney


kidney

kidney


kidney


kidney



kidney

kidney



Route and
Type of
Exposure
oral, acute


oral, acute

oral, acute


oral, acute


oral (drinking water)
subchronic


oral (drinking water)
subchronic
oral (drinking water)
subchronic


Dose or
Species Exposure
mice, males 149 mg/kg, single dose
of sensitive
strains
mice, male 15 mg/kg, single dose

mice, male 60 mg/kg, single dose


mice, male 240 mg/kg, single dose


rats 160 mg/kg/day x
90 days


rats =300 mg/kg/day x
90 days
mice 290 mg/kg/day x
90 days


Effect on
Target Organ
loss of glucose or
protein in urine

no effect

focal tubular
epithelial regenera-
tion
severe diffuse cortical
necrosis, focal tubular
epithelial regeneration
no effect



no effect

no effect



Reference
Hill, 1978


Reitz et al. ,
1980
Reitz et al . ,
1980

Reitz et al. ,
1980

Jorgenson
and
Rushbrook,
1980
Chu et al . ,
1980b
Jorgenson
and
Rushbrook,
1980

-------
                  Table 5-5



 Target Organ Toxicity of Chloroform (cont.)
Target
Organ
kidney
kidney
v* kidney
kidney
kidney
kidney
Route and
Type of
Exposure
oral (capsule),
subchronic
oral, chronic
oral (gavage),
chronic
oral (gavage),
chronic
oral (gavage),
chronic
oral (gavage)
chronic,
Species
dogs
humans
rats
mice
mice , males
of sensitive
strains
mice , males
of sensitive
Dose or
Exposure
120 mg/kg/day,
7 days /week x
12 weeks
2.5 mg/kg/day,
7 days /week, for
>1 year
200 mg/kg/day,
5 days /week x
78 weeks
138 or 227 mg/kg/day,
5 days /week x
78 weeks
17 mg/kg/day,
6 days /week x
80 weeks
60 mg/kg/day,
6 days /week x
Effect on
Target Organ
no effect
no effect on BUN
no effect
decreased incidence
of renal disease
no effect
increased incidence
of moderate to severe
Reference
Jorgenson
and
Rushbrook,
1980
De Salva
et al., 1975
NCI, 1976
NCI, 1976
Roe et al . ,
1979
Roe et al . ,
1979
strains
80 weeks
renal disease

-------
CNS
                                                     Table  5-5
                                    Target Organ  Toxicity of  Chloroform  (cont.)
Target
Organ
kidney
kidney
kidney
CP,
vil
OO
kidney
central
nervous
system
(CNS)
Route and
Type of
Exposure
oral (gavage)
chronic
oral (capsule) ,
chronic
oral (capsule),
chronic
dermal, acute
inhalation, acute
Species
mice, males
of insensi-
tive strains,
females
dog
dog
rabbits
humans
Dose or
Exposure
60 mg/kg/day,
6 days /week x
80 weeks
15 mg/kg/day
6 days/week x
7.5 years
30 mg/kg/day,
6 days /week x
7.5 years
1.0, 2.0, and
3.98 g/kg x 24 hours
under plastic cuff,
single exposure
900-1400 ppm for
>30 minutes, single
exposure
Effect on
Target Organ
no effect
no effect
increase in fat
deposition in glomeruli
degenerative changes
in tubules
dizziness, tiredness,
headache
Reference
Roe et al. ,
1979
Heywood et al . ,
1979
Heywood
et al., 1979
Torkelson
et al., 1976
Lehman n
and
Hasegawa, 1910;
Lehmann and
inhalation,  acute
humans
M300-5100 ppm x
20 minutes, single
exposure
dizziness, light
intoxication
Schmidt-Kehl,
1936

Lehmann and
Hasegawa,  1910;
Lehmann and
Schmidt-Kehl,
^936

-------
                                                        Table 5-5

                                       Target Organ Toxicity of Chloroform (cont.)
ui
Target
Organ
CNS
Route and
Type of
Exposure
inhalation, acute
Species
humans
Dose or
Exposure
1500-2000 ppm, si
Effect on
Target Organ
ngle maintenance of light
Reference
Goodman and
CNS       inhalation,  acute



CNS       inhalation,  acute



CNS       inhalation,  acute


CNS       inhalation,  acute


CNS       inhalation,  acute


CNS       inhalation,  acute



CNS       inhalation,  acute
                                      humans
                                      humans
                                      mice
                                      mice
                                      mice
                                      cats
                                      cats
                                                     exposure
15,000 ppm, single
exposure
20,000-40,000 ppm x
a few minutes, single
exposure

2500 ppm x 12 hours,
single exposure

3100 ppm x 1 hour,
single exposure

4100 ppm x 0.5 hours,
single exposure

7200 or 21,500 ppm x
5 minutes, single
exposure

7200 ppm x 60 minutes
single exposure
anesthesia (after
induction)

maintenance of heavy
anesthesia (after
induction)
                                                                                                    Gillman,  1980
Goodman and
Gillman, 1980
induction of anesthesia  NIOSH, 197H
                         Adriani,  1970
no obvious effects
light narcosis
deep narcosis
disturbance of equilib-
rium
light narcosis
Lehmann and
Flury, 1943

Lehmann and
Flury, 1943

Lehmann and
Flury, 1943

Lehmann and
Flury, 1943
Lehmann and
Flury, 1943

-------
CNS
                                                     Table  5-5

                                    Target Organ  Toxicity of  Chloroform  (cont.)
Target
Organ
CNS

CNS

CNS

CNS
Route and
Type of
Exposure
inhalation, acute

inhalation, acute

inhalation, acute

inhalation, chronic
Species
cats

cats

cats

humans
Dose or
Exposure
7200 ppm x 93 minutes
single exposure
21,500 ppm x 10 minutes
single exposure
21,500 ppm x 13 minutes
single exposure
20-71 ppm (with excur-
Effect on
Target Organ
deep narcosis

light narcosis

deep narcosis

tiredness
Reference
Lehmann and
Flury, 19^3
Lehmann and
Flury, 19^3
Lehmann and
Flury, 1943
Challen
inhalation,  chronic
humans
sions to 1163 ppm
lasting 1.5-2 minutes)
for 4-8 hours/day,
5 days/week

77 to 237 ppm (with
excursions to =1163
lasting 1.5-2 minutes)
for 4-8 hours/day,
5 days/week
                                                                                                    et  al.,  1958
tiredness, depression,
occasional silliness
or staggering during
the workday
Challen
et al., 1958

-------
01
                                                        Table 5-5




                                       Target Organ Toxicity of Chloroform  (cont.)
Target
Organ
CNS
CNS
CNS
CNS
CNS
Route and
Type of
Exposure Species
oral, acute rats
oral (drinking water), rats
subchronic
oral (drinking water), mice
subchronic
oral (drinking water) rats
subchronic
oral (gavage), rats
subchronic
Dose or
Exposure
350 mg/kg
single dose
20-160 mg/kg/day x
90 days
32-290 mg/kg/day x
90 days
=300 mg/kg/day x
90 days
60 rag/kg/day
6 days /week x
80 weeks
Effect on
Target Organ
minimum narcotic dose
(MND )
dose-related signs
of depression during
1st week only
dose-related signs of
depression during 1st
week only
no histopathologic
changes in brain
no effect on gross
or histologic appear-
ance of brain
Reference
Jones et al . ,
1958
Jorgenson
and
Rushbrook,
1980
Jorgenson
and
Rushbrook ,
1980
Chu et al . ,
1982b
Palmer et al . ,
1977

-------
Grey,  1979).   In  the kidney,  chloroform exposure  produces necrosis  of  the




proximal and distal convoluted tubules  (Eschenbrenner and Miller,  1945a).   The




mechanism  by  which  chloroform produces  these  effects  has  been  extensively



studied in experimental animals.  From  the studies  summarized  in Section 5.3.1




(Brown et al.,  1974;  Ilett  et al., 1973; Docks and Krishna,  1976; Ekstrbm et al.,




1981; Lavigne and Marchand, 1974; McMartin et al.,  1981; Masuda  et  al.,  1980;




Harris et  al.,  1982;  Stevens  and Anders, 1981),  it  appears  that chloroform is




first metabolized in the  target organ by microsomal drug metabolizing enzymes to




a reactive  intermediate, probably phosgene, which in turn  can  react by various




pathways, depending on glutathione levels, one of which is  the  covalent binding




to liver proteins resulting in necrotic lesions.   A similar mechanism may or may




not  occur in the kidney (Kluwe and Hook, 1981).




5.5.7   Factors  that  Modify the  Toxicity of Chloroform.    Several  substances




alter the toxicity of  chloroform, most  probably  by  modifying the metabolism of




chloroform to a reactive  intermediate  (see Section 5.4).  These substances are of




interest  because  humans  may be accidentally  or  intentionally exposed  to them.




Factors  that  potentiate the  toxic  effects  induced  by  exposure  to chloroform




include ethanol  (Kutob and Plaa, 1962;  Sato et al.,  1980, 1981), polybrominated




biphenyls  (Kluwe  and  Hook,  1978), steroids  (Clemens  et al.,  1979), and ketones




(Hewitt et al., 1979; Jernigan and Harbison,  1982; Branchflower and Pohl, 1981).




Disulfiram  and its metabolites (Scholler et al., 1970; Masuda  and  Nakayama, 1982;




Gopinath  and Ford,  1975)  and  high carbohydrate  diets  (Nakajima  et al., 1982),




appear  to  protect against  chloroform toxicity.
                                      5-62

-------
                                 5.   REFERENCES








Adrian!, J.  1970.  The Pharmacology of Anesthetic Drugs.  Springfield, Illinois,




Charles C. Thomas,  pp. 57-60.








Bomski, H., A. Sobolweska and A. Strakowski.  1967.  [Toxic damage of the liver




by chloroform in  chemical  industry workers.]  Arch. Gewerbepathol.  Gewerbehy.




24: 127-131*.  (Ger.)   (Cited in Utidjian, 1976;  NIOSH,  1974.)








Bowman, F.J.,  J.  Borzelleca  and A.E. Munson.    1978.   The  toxicity  of  some




hale-methanes in mice.  Toxicol. Appl. Pharmacol.   44: 213-215.








Branchflower,  R.V. and L.R. Pohl.  1981.   Investigation of the mechanism of the




potentiation of chloroform-induced hepatotoxicity and nephrotoxicity by methyl,




n-butyl ketone.   Toxicol. Appl. Pharmacol.  61:  407-413.








Brown,  B.R.,  Jr.,  I.G.  Sipes  and A.M.  Sagalyn.   1974.   Mechanisms  of acute




hepatic toxicity:    Chloroform,  halothane  and   glutathione.    Anesthesiology.




41: 554-561.








Butler, T.C.  1961.   Reduction of carbon  tetrachloride  in vivo and reduction of




carbon tetrachloride  and  chloroform in vitro by tissues and tissue constituents.



Jour. Pharmacol. Exp. Ther.  134: 311-319.








Challen, P.J.R., D.E.  Hickish and J. Bedford.  1958.  Chronic  chloroform intoxi-




cation.  Brit. Jour.  Ind. Med.   15: 243-249.
                                     5-63

-------
Chu,  I.,  V. Secours, I. Marino and D.C.  Villeneuve.   1980.  The acute toxicity of




four  trihalomethanes   in  male  and female  rats.    Toxicol.  Appl.  Pharmacol.




52: 351-353.







Chu,  I.,  D.C.  Villenuve,  V.E.  Secours  and  G.C. Becking.   1982a.   Toxicity of




trihalomethanes:   I.   The  acute and  subacute  toxicity of  chloroform,  bromo-




dichloromethane, chlorodibromomethane  and bromoform in  rats.   J.  Environ. Sci.




Health.  B17: 205-224.








Chu,  I.,  D.C. Villenuve,  V.E.  Secours, G.G.  Becking  and V.E. Valli.   1982b.




Toxicity  of   trihalomethanes:   II.     Reversibility of  toxicological  changes




produced by chloroform, bromodichloromethane, chlorodibromomethane  and bromoform




in rats.  J.  Environ.  Sci.  Health.  B17: 225-240.








Cianflone, D.J., W.R.  Hewitt, D.C. Villenueve  and G.L. Plaa.  1980.   Role of




biotransformation in the  alterations  of chloroform  hepatotoxicity produced by




Kepone and Mirex.  Toxicol. Appl. Fharmacol.  53: 140-149.








Clemens,  T.L.,   R.N.   Hill,   L.P.   Bullock,  W.D.   Johnson,  L.G.  Sulatos  and




E.S. Vesell.   1979.  Chloroform toxicity in the mouse:   Role of genetic factors




and steroids.  Toxicol. Appl. Pharmacol.  48: 117-130.








Conlon, M.F.   1963.  Addiction to chlorodyne.  Brit. Med. Jour.  2: 1177-1178.








Culliford, D. and H.B. Hewitt.  1957.  The influence of sex hormone  status on the




susceptibility of  mice to  chloroform-induced  necrosis  of the renal  tubules.



Jour. Endocrinol.  14: 381-393.
                                     5-64

-------
Danni, 0., 0. Brossa, E. Burdino, P. Milillo and G. Ugazio.  1981.  Toxicity of
halogenated hydrocarbons in pretreated rats:  An experimental model for the study
of  integrated  permissible  limits of environmental poisons.   Int. Arch. Occup.
Environ. Health.  49:  105-112.


Deringer, M.K., T.B. Dunn and W.E. Heston.  1953.   Results  of exposure of strain
C3H mice to chloroform.  Proc. Soc. Exp. Biol. Med.  83:
DeSalva, S. ,  A. Volpe,  G.  Leigh and T. Regan.  1975.  Long-term safety studies of
a  chloroform-containing  dentifrice  and  mouth-rinse  in  man.   Food  Cosmet.
Toxicol.   13: 529.

Divincenzo, G.D. and W.J.  Krasavage.  1971*.  Serum ornithine carbamyl transferase
as liver  response  test for exposure to organic  solvents.  Amer.  Indust.  Hyg.
Assoc. J.  35: 21-29.

Docks, E.L. and G.  Krishna.   1976.  The role of glutathione in chloroform-induced
hepatotoxicity.  Exp. Molec. Pathol.  24:  13-22.

Ekstrb'm,  T. ,  J. Hoegbers  and B. Jernstroem.    1982.   Inhibition of  hepatic,
gamma-glutamyl-cysteine synthetase  by chloroform.   Biochem.  Pharmacol.    31:
3510-3513.

Eschenbrenner , A.B. and E.  Miller.   19l5a.   Induction of hepatomas  in mice by
repeated oral administration  of chloroform, with  observations on sex  differ-
ences.  Jour.  Natl. Cane. Ins tit.   5:  251-255.
                                     5-65

-------
Eschenbrenner,  A.B.  and  E. Miller.  1945b.  Sex differences in kidney morphology




and chloroform necrosis.  Science.  102: 302-303.








Goodman, L.S. and A. Oilman.  1980.   The Pharmacological Basis of Therapeutics.




6th ed.  New York, MacMillan and Co.








Gopinath, C. and E.J.H.  Ford.  1975.   The role of microsomal hydroxylases in the




modification of  chloroform  hepatotoxicity  in rats.   Brit. Jour.  Exp.  Pathol.




56: 412-422.








Groger, W.K.L. and T.F.  Grey.  1979.   Effect of chloroform on the activities of




liver enzymes in rats.  Toxicology.   14: 23-28.








Harris,  R.N.,  J.H.  Ratnayake,  V.F.  Garry and M.W.  Anders.   1982.  Interaction




hepatotoxicity  of  chloroform  and   carbon  tetrachloride.     Toxicol.   Appl.




Pharmacol.  63: 281-291.








Heller, H. and F.H.  Smirk.   1932.  The influence of anaesthetics  and hypnotics on




the absorption and excretion of water.  Jour. Physiol. (London).  76: 292-302.








Hewitt, W.R., H. Mujajima, M.G. C6te and G.L. Plaa.  1980.  Acute alteration of




chloroform-induced hepato- and  nephrotoxicity by n-hexane, methyl n-butyl ketone




and 2,5-hexanedione.  Toxicol. Appl. Pharmacol.  53: 230-248.








Hewitt, W.R., H. Miyajima, M.G. C6t6 and G.C. Plaa.  1979.  Acute alteration of




chloroform-induced  hepato-  and nephrotoxicity by  Mirex and Kepone.   Toxicol.




Appl.   Pharmacol.  48:  509-527.
                                      5-66

-------
Heywood, R., R.J. Sortwell,  P.R.B.  Noel,  A.E. Street, D.E. Prentice, F.J.C. Roe,
P.F. Wardsworth,  A.N.  Worden and N.J. Van Abbe'.   1979.   Safety evaluation of
toothpaste containing chloroform.  III.  Long-term study in beagle dogs.  Jour.
Environm. Pathol. Toxicol.  2: 835-851.

Hill, R.N.  1978.  Differential toxicity of chloroform in the mouse.  Ann.  New
York Acad. Sci.  298: 170-175.

Hjelle, J.J.,  A.S. Gordon  and D.R. Peterson.   1982.   Studies  on carbon tetra-
chloride-ethanol interactions in mice.  Toxicol. Lett.  10: 17-24.

IARC (International  Agency  for  Research  on Cancer).   1976.   Disulfirams.  In:
Some Carbamates, Thiocarbamates and Carbazides.  IARC Monographs on the Evalua-
tion of the Carcinogenic Risk of Chemicals to Man.  WHO: Lyon, France.  IARC, p.
85-96.

Ilett,  K.F., W.D. Reid,  I.G.  Sipes  and G. Krishna.  1973.  Chloroform toxicity in
mice:  Correlation of renal and hepatic necrosis with covalent binding of metabo-
lites to tissue macromolecules.  Exp.  Molec.  Pathol.   19:  215-229.

Jernigan,  J.D.  and  R.D. Harbison.    1982.   Role of  biotransformation in  the
potentiation  of  halocarbon  hepatotoxicity  by  2,5-hexanedione.   J.  Toxicol.
Environ. Health.  9:   761-781.

Jones,  W.M., G. Margolis and C.R. Stephen.  1958.   Hepatotoxicity of inhalation
anesthetic drugs.  Anesthesiology.   19: 715.
                                     5-67

-------
Jorgenson, T.A.  and C.J. Rushbrook.  1980.  Effects of chloroform in the drinking




water of rats and mice:  Ninety-day subacute toxicity study.  SRI International,




Menlo  Park,  California.   Publication  No.  EPA-600/1-80-030.   Available  from



National  Technical   Information  Service,  Springfield,  Virginia  (NTIS  PB80-




219108).








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.  Relative effect  of various chlorinated




hydrocarbons on liver and  kidney  function in mice.   Toxicol.  Appl.  Pharmacol.




9: 139-151.








Kluwe, W.M. and J.B. Hook.  1978.   Polybrominated biphenyl-induced potentiation




of chloroform toxicity.  Toxicol.  Appl. Pharmacol.   45: 861-869.








Kluwe,  W.M.  and  J.B.  Hook.   1981.   Potentiation  of  acute  chloroform nephro-



toxicity  by  the  glutathione  depletion diethyl maleate  and protection  by the




microsomal enzyme inhibitor piperonyl butoxide.  Toxicol. Appl. Pharmacol.  59:



457-466.








Knocker, P. and J. Mandelstam.  1944.  Rhythmic changes in  the rat produced by a




single injection of chloroform.   Nature (London).  154: 148-149.








Krantz, J.C.,  Jr. and  C.J.  Carr.   1965.  Pharmacologic  Principles  of Medical




Practice.  6th ed.  Baltimore, Williams and Wilkins.
                                      5-68

-------
Krus, S. and Z.  Zaleska-Rutczynski.   1970.  Morphological  counterparts  of the




genetically  determined  resistance  of   mice   to   chloroform.     Experientia.




26: 101-102.







Kurtz, C.M., J.H. Bennett and H.H.  Shapiro.  1936.  Electrocardiographic studies




during surgical anesthesia.  J. Amer. Med. Assoc.  106: 434-441.








Kutob, S.D. and G.L. Plaa.   1962a.  The effect of acute ethanol intoxication on




chloroform-induced liver damage.  Jour. Pharmacol.  Exp. Teratol.  135: 245-251.








Kutob, S.D. and G.L. Plaa.   1962b.  A procedure of estimating the hepatotoxicity




potential of certain industrial solvents.  Toxicol.  Appl. Pharmacol.  4:  354.








Kylin,  B.,  H.  Reichard, I.  Sumegi  and  S.  Yllner.   1963.   Hepatotoxicity of




inhaled trichloroethylene,  tetrachloroethylene and chloroform.  Single exposure.




Acta. Pharmacol.  Toxicol.  20: 16-26.








Lavigne, J.G.  and C. Marchand.   1974.   The  role  of metabolism  in chloroform




hepatotoxicity.  Toxicol. Appl. Pharmacol.  29:  312-326.








Lehmann, K.B. and F.F.  Flury.  1943.   Chlorinated hydrocarbons.   In; Toxicology




and Hygiene  of Industrial  Solvents,  1943.  K.B. Lehmann and F.F.  Flury,  Eds.




Williams and Wilkins Co., Baltimore,   p.  138-115, 191-196.








Lehmann, K.B. and Hasewaga.   1910.   [Studies  of the absorption  of  chlorinated




hydrocarbons in animals  and humans.]  Arch. Hyg.   72:  327-342. (Ger.)  (Cited in



NIOSH, 1974.)
                                     5-69

-------
Lehmann, K.B. and L. Schmidt-Kenl.  1936.   [The  thirteen  most  important chlor-




inated aliphatic hydrocarbons from the standpoint of industrial hygiene.] Arch.




Hyg.  116: 131-200.  (Ger.)  (Cited in NIOSH,  1974.)







Lucas, G.H.W.   1928.   A  study  of the  fate and toxicity of bromine and chlorine




containing anesthetics.  Jour.  Pharmacol. Exp. Ther.  3**:  223.








Mansuy, D.,  P.  Beaune, T.  Cresteil, M.  Lange  and J.P. Lerous.   1977.  Evidence




for  phosgene  formation  during  liver  microsomal  oxidation  of  chloroform.




Biochem. Biophys. Res.  Commun.  79: 513-517.








Masuda, Y. and N. Nakayama.  1982.  Protective effect of diethyldithiocarbamate




and carbon disulfide against liver injury induced by various hepatotoxic agents.




Biochem.  Pharmacol.  31: 2713-2725.








Masuda, Y., I. Yano and T.  Murano.   1980.  Comparative studies on the hepatotoxic




actions  of chloroform and  related halomethanes in normal and  phenobarbital-




pretreated animals.  Jour. Pharmacobio. Dyn.  3: 53-6M.








McLean, A.E.M.  and  E.K.  McLean.   1969.   Diet and  toxicity.   Brit.  Med. Bull.




25: 278-281.








McLean, A.E.M.   1970.   The effect  of  protein deficiency  and microsomal enzyme




induction by  DDT and phenobarbitone on  the acute toxicity  of  chloroform and a




pyrrolizidiue alkaloid, retrorsine.  Brit. Jour. Exp. Pathol.   51: 317-321.
                                      5-70

-------
McMartin, N.D., J.A. O'Connor, Jr.,  L.S. Kaminsky.  1981.  Effects of differen-
tial changes in rat hepatic and  renal cytochrome P-450 concentrations on hepato-
toxicity  and  nephrotoxicity  of  chloroform.    Res.  Commun.  Chem.  Pathol.
Pharmacol.  31: 99-110.

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

NCI  (National  Cancer  Institute).   1976.   Report  on Carcinogenesis Bioassay of
Chloroform.  Available from National Technical Information Service, Springfield,
Virginia.   (NTIS PB-26H-018).

NIOSH (National Institute  for Occupational Safety and Health).   1974.  Criteria
for  a Recommended Standard-Occupational Exposure to Chloroform.  Available from
National   Technical  Information   Service,   Springfield,   Virginia.     (NTIS
P8-246-695).

Oettel,  H.  1936.  Arch. Exp. Pathol. Pharmacol.  183: 655.

Orth, O.S., R.R. Liebenow and R.T. Capps.  1951.   III-The  effect of chloroform on
the  cardiovascular system.   In:  R.M. Waters (Ed).  Chloroform - A  Study After 100
Years.   University of Wisconsin Press, Madison, 1951.  p. 39-75.

Palmer,  A.K.,  A.E.  Street, F.J.C. Roe, A.N.  Worden and  N.J.  Van Abbe'.  1979.
Safety evaluation of toothpaste  containing  chloroform.  II.  Long-term  studies in
rats.  Jour. Environm. Pathol.  Toxicol.  2: 821-833-
                                      5-71

-------
Paul,  B.B. and  D.  Rubinstein.   1963-   Metabolism  of carbon  tetrachloride  and




chloroform by the rat.  Jour.  Pharmacol. Exp.  Ther.  141:  141-1U8.








Piersol, G.M., A.J. lumen and L.S. Kau.   1933-   Fatal poisoning  following  the




ingestion of chloroform.  Med.  Clin.  North Am.  17: 587-601.








Pohl,  L.R.,  B.  Bhooshan,  N.F.  Whittaker and G. Krishna.   1977.   Phosgene:   A




metabolite of chloroform.  Biochem. Biophys. Res.  Commun.   79: 68U-691.








Pohl,  L.R.  1979.   Biochemical toxicology of chloroform.  Rev.  Biochem.  Toxicol.




1: 79-107.








Reid,  W.D.  and  G. Krishna.   1973.  Centrilobular hepatic necrosis  related to




covalent  binding  of metabolites  of  halogenated aromatic  hydrocarbons.    Exp.




Molec.  Pathol.   18: 80-95.








Reitz,  R.H.,  J.F. Quast,  W.T.  Stott,  P.G. Watanabe  and  P.J. Gehring.   1980.




Phannacokinetics  and  raacromolecular  effects  of chloroform in  rats and  mice:



Implications  for  carcinogenic  risk estimation.    Water  Chlorinat.  Environm.




Impact Health Eff.  3: 983-993.








Reynolds, E.S. and A.G.  Yee.  1967.  Liver parenchymal cell  injury. V.  Relation-



                                          1M
ships between patterns of chloromethane  -   C incorporation into constituents of




liver in vivo and cellular injury.  Lab. Invest.  16: 591-603.
                                     5-72

-------
Roe,  F.J.C.,  A.K. Palmer,  A.N.  Worden and N.J.  Van Abbe' .  1979.  Safety evalua-




tion of toothpaste containing chloroform.  I.  Long-term studies in mice.  Jour.




Environm. Toxicol.  2: 799-819.








Sato, A., T.  Nakajima and Y. Koyama.   1980.  Effects of chronic ethanol consump-




tion on hepatic metabolism of aromatic and chlorinated hydrocarbons in rats.  Br.




J. Indust.  Med.  37: 382-386.








Sato, A., T.  Nakajima and Y. Koyama.   1981.  Dose-related effect of a single dose




of ethanol  on the  metabolism  in rat liver of  some  aromatic  and  chlorinated




hydrocarbons.  Toxicol. Appl. Pharmacol.  60: 8-15.








Sax, N.I.  1979.   Dangerous Properties  of Industrial  Materials, Fifth edition.




New  York, Van Nostrand Reinhold Company, Inc.  p. 193.








Scholler, K.L.   1966.  Elektronmikroskopische untersuchungen an leberzellen der




Ratte nach narkosen mit verschiedenen inhlalationsnarkotica.  Der Anaesthetist.




15:  145-148.  (Ger).








Scholler, K.L.   1967.  Effect of haloethane and chloroform on the fine structure




and  protein synthesis of rat liver.   Experimentia.  23: 652-655.  (In German).








Scholler,  K.L.    1970.   Modification of  the  effects  of  chloroform  on  the rat




liver.   Brit. Jour. Anaesth.  42: 603-605.








Schroeder,  H.G.    1965.    Acute and  delayed chloroform  poisoning.    Brit.  J.




Anaesth.  37: 972-975.
                                      5-73

-------
Sipes, I.G.,  G.  Krishna and  J.R.  Gillette.   1977.   Bioactivation  of  carbon




tetrachloride, chloroform,  and bromotrichloromethane:  Role of cytochrome P-450.




Life Sci.  20: 1541-1548.







Stevens,  J.L. and M.W. Anders.  1981.  Effect of cysteine, diethyl maleate,  and




phenobarbital treatment on the hepatoxicity  of  (1H)- and (2H)-chloroform.  Chem.




Biol. Interact.  37: 207-217.








Storms, W.W.   1973.  Chloroform parties.  J. Amer. Med. Assoc.  225:  160.








Torkelson,  T.R., F.  Oyen and V.K.  Rowe.   1976.  The  toxicity  of  chloroform as




determined  by  single and repeated  exposure  of  laboratory animals.  Amer.  Ind.




Hyg.  Assoc.  Jour.  37: 697-705.








Utidjian,  H.M.D.   1976.   Recommendations  for  a  chloroform  standard.   Jour.




Occupat. Med.  18: 253-








Van  Dyke,  R.A.,  M.B. Chenoweth and A.V.  Poznak.  1964.  Metabolism of volatile




anesthetics.   I.    Conversion iri  vivo  of  several  anesthetics  to     C0~  and




chloride.   Biochem.   Pharmacol.  13: 1239-1247.








von  Oettigen,  W.F.   1964.  The halogenated hydrocarbons of industrial and toxi-




cological  importance.  Amsterdam,  Elsevier.








Wallace,  C.J.    1950.   Hepatitis and nephrosis  due to  cough  syrup  containing




chloroform.   Calif.  Med.   73:  442.
                                      5-74

-------
Watrous, W.M.  and G.L.  Plaa.    1972.   Effect  of halogenated hydrocarbons  on




organic ion accumulation by renal cortical  slices of rats and mice.   Toxicol.




Appl. Pharmacol.  22: 528-543.








Winslow, S.G. and H.B. Gerstner.  1978.  Health aspects of chloroform - A review.




Drug Chem.  Toxicol.  1:  259-275.








Wood, S.,  B.  Wardley-Smith, M.  J. Halsey and C.J. Green.  1982.  Hydrogen bonding




in mechanisms of  anesthetic  tested  with chloroform and  deuterated  chloroform.




Br. J.  Anaesth.  5M: 387-391.








Wood-Smith,  F.G.  and H.C.  Stewart.    196U.    Drugs  in  Anaesthetic  Practice.




Washington, D.C.  Butterworth,  Inc.   pp. 131-135.
                                     5-75

-------
                United States
                Environmental Protection
                Agency
                Office of Health and
                Environmental Assessment
                Washington DC 20460
EPA-600/8-84-004A
March 1984
External Review Draft
&EPA
                Research and Development
Health Assessment
Document for
Chloroform

Part 2 of 2
Review
Draft
(Do Not
Cite or Quote)
                               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.

-------
Review
Draft
(Do Not
Cite or Quote)
EPA-600/8-84-004A
March 1984
External Review Draft
              Health Assessment
        Document for Chloroform
                    Part 2  of 2
            External Review  Draft
                          NOTICE
This document is a preliminary draft. It has not been formally released by the U.S. Environmental
Protection Agency 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 AGENCY
                Office of Research and Development
             Office of Health and Environmental Assessment
             Environmental Criteria and Assessment Office
                 Research Triangle Park, NC 27711

                        March 1984

-------
                                 PREFACE
     The Office of Health and Environmental Assessment has prepared this health
assessment to serve as a "source document" for EPA use.  This health assessment
document was developed for use by the Office of Air Quality Planning and
Standards to support decision-making regarding possible regulation of chloroform
as a hazardous air pollutant.
     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 chemical's toxicity and related characteristics
are qualitatively identified.  Observed effect levels and other measures of
dose-response relationships are discussed, where appropriate, so that the
nature of the adverse health response are placed in perspective with observed
environmental levels.
     This document will  be  subjected to a  thorough  copy editing and proofinq
following the revision based  on the  EPA's  Scientific  Advisory Board review
comments.

-------
                               TABLE OF CONTENTS
LIST OF TABLES	     vi

LIST OF FIGURES	       x

1 .    SUMMARY AND CONCLUSIONS	     1 -1

2.    INTRODUCTION	     2-1

3.    BACKGROUND INFORMATION 	     3-1

     3.1    INTRODUCTION 	     3-1
     3.2   PHYSICAL AND CHEMICAL PROPERTIES  	     3-2
     3.3   SAMPLING AND ANALYSIS 	     3-4

           3.3.1    Chloroform in Air	     3-4
           3.3.2   Chloroform in Water	     3-5
           3.3.3   Chloroform in Blood	     3-6
           3.3.4   Chloroform in Urine	     3-6
           3.3.5   Chloroform in Tissue	     3-6

     3.4   EMISSIONS FROM PRODUCTION AND USE  	     3-6

           3.4.1    Emissions from Production	     3-6
           3.4.2   Emissions from Use	    3-20
           3.4.3   Summary of Chloroform Discharges from Use	    3-26

     3.5   AMBIENT AIR CONCENTRATIONS 	    3-26
     3.6   ATMOSPHERIC REACTIVITY 	    3-32
     3. 7   ECOLOGICAL EFFECTS/ENVIRONMENTAL PERSISTENCE 	    3-33

           3.7.1    Ecological Effects	    3-33
           3.7.2   Environmental Persistence	    3-36

     3.8   EXISTING CRITERIA, STANDARDS, AND  GUIDELINES 	    3-39

           3.8.1    Air	    3-39
           3.8.2   Water	    3-41
           3.8.3   Food	    3-41
           3.8 .4   Drugs and Cosmetics	    3-42

     3.9   RELATIVE SOURCE CONTRIBUTIONS 	    3-42
     3.10  REFERENCES 	    3-43

4.    DISPOSITION AND RELEVANT PHARMACOKINETICS  	     4-1

     4.1    INTRODUCTION 	     4-1

-------
                           TABLE OF CONTENTS (cont.)
5.

4.2



4.3
4.4



4.5



4.6




4.7
4.8

ABSORPTION 	 ,
4.2.1 Dermal Absorption 	 	
4.2.2 Oral Absorption 	 ,
4.2.3 Pulmonary Absorption 	 ,
TISSUE DISTRIBUTION 	
EXCRETION 	
4.4.1 Pulmonary Excretion 	
4.4.2 Other Routes of Excretion 	
4.4.3 Adipose Tissue Storage 	
BIOTRANSFORMATION OF CHLOROFORM 	
4.5.1 Known Metabolites 	
4.5.2 Magnitude of Chloroform Metabolism 	
4.5-3 Enzymic Pathways of Biotransformation 	
COVALENT BINDING TO CELLULAR MACROMOLECULES 	
4.6.1 Proteins and Lipids 	
4.6.2 Nucleic Acids 	
4.6.3 Role of Phosgene 	
4.6.4 Role of Glutathione 	
SUMMARY 	
REFERENCES 	
TOXICITY 	
5.1


EFFECTS OF ACUTE EXPOSURE TO CHLOROFORM 	
5.1 .1 Humans 	
5.1 .2 Experimental Animals 	
Page
	 4-2
	 4-2
	 4-3
	 4-7
	 4-12
	 4-21
	 4-21
	 4-31
	 4-32
	 4-34
	 4-34
	 4-37
	 4-40
	 4-47
	 4-47
	 4-53
	 4-59
	 4-61
	 4-63
	 4-67
	 5-1
	 5-1
	 5-1
	 5-6
     5.2   EFFECTS OF CHRONIC EXPOSURE TO CHLOROFORM  	    5-13

           5.2.1    Humans	    5-13
           5.2.2   Experimental Animals	    5-15

     5.3   INVESTIGATION OF TARGET ORGAN TOXICITY  IN  EXPERIMENTAL
           ANIMALS 	    5-28

           5.3.1    Hepatotoxicity	    5-28
           5.3.2   Nephrotoxicity	    5-34
                                      IV

-------
                           TABLE OF CONTENTS (cont.)

                                                                           Page

     5.4   FACTORS MODIFYING THE TOXICITY OF CHLOROFORM  	   5-38

           5.4.1   Factors that Increase Toxicity	   5-39
           5.4.2   Factors that Decrease Toxicity	   5-44

     5.5   SUMMARY; CORRELATION OF EXPOSURE AND EFFECT 	   5-46

           5.5.1   Effects of Acute Inhalation Exposure	   5-46
           5.5.2   Effects of Acute Oral Exposure	   5-47
           5.5-3   Effects of Dermal Exposure	   5-48
           5.5.4   Effects of Chronic Inhalation Exposure	   5-48
           5.5.5   Effects of Chronic Oral Exposure	   5-50
           5.5.6   Target Organ Toxicity	   5-51
           5.5.7   Factors that Modify the Toxicity of Chloroform	   5-62

     5.6   REFERENCES	   5-63

6 .   TERATOGENICITY AND REPRODUCTIVE EFFECTS	    6-1

     6.1    REFERENCES	   6-16

7.   MUTAGENICITY	    7-1

     7.1    INTRODUCTION	    7-1
     7.2   COVALENT BINDING TO MACROMOLECULES	    7-1
     7. 3   MUTAGENICITY STUDIES IN BACTERIAL TEST SYSTEMS	    7-3
     7.4   MUTAGENICITY STUDIES IN EUCARYOTIC TEST SYSTEMS	    7-9
     7.5   OTHER STUDIES INDICATIVE OF DNA DAMAGE	   7-14
     7.6   CHROMOSOME STUDIES	   7-19
     7.7   SUGGESTED ADDITIONAL TESTING	   7-21
     7.8    REFERENCES	   7-23

8.   CARCINOGENICITY 	    8-1

     8.1    ANIMAL STUDIES 	    8-1

           8.1.1   Oral Administration (Gavage): Rat	    8-2
           8.1.2   Oral Administration (Gavage): Mouse	   8-11
           8.1.3   Oral Administration (Capsules):  Dog	   8-21
           8.1.4   Intraperitoneal Administration:  Mouse	   8-24
           8.1.5   Evaluation of Chloroform Carcinogenicity by
                     Reuber (1979)	   8-26
           8.1.6   Oral Administration (Drinking Water):  Mouse:
                     Promotion of Experimental Tumors	   8-26

     8 .2   CELL TRANSFORMATION ASSAY 	   8-32

           8.2.1   Styles (1979)	   8-32

-------
                      TABLE  OF  CONTENTS  (cont.)
8 . 3   EPIDEMIOLOGIC STUDIES  	    8-34

      8.3.1    Young et al.  (1981)	    8-37
      8.3-2   Hogan et al.  (1979)	    8-40
      8.3.3   Cantor et al.  (1978)	    8-42
      8.3.4   Gottlieb et al.  (1981)	    8-46
      8.3.5   Alavanja et al.  (1978)	    8-48
      8.3.6   Brenniman et  al.  (1978)	    8-50
      8.3.7   Struba (1979)	    8-51
      8.3-8   Discussion	    8-53

8.4   QUANTITATIVE ESTIMATION  	    8-56

      8.4.1    Procedures for the  Determination of Unit Risk	    8-59
      8 .4.2   Unit Risk Estimates	    8-70
      8.4.3   Comparison of Potency  with  Other Compounds	    8-80
      8.4.4   Summary of Quantitative  Assessment	    8-80

8.5   SUMMARY 	    8-85

      8.5.1    Qualitative	    8-85
      8.5.2   Quantitative	    8-88

8.6   CONCLUSIONS 	    8-89
8.7   REFERENCES 	    8-91
8.8   APPENDIX A:   Comparison Among Various Extrapolation
                     Models	     A-l
                                 VI

-------
LIST OF  TABLES
Table
3-1
3-2
3-3
3-4

3-5
3-6
3-7
3-8
3-9
3-10

3-11
3-12
4-1

4-2
4-3

14-4

4-5

4-6

4-7


Physical Properties of Chloroform 	
Chloroform Producers, Production Sites, and Capacities 	
Chloroform Discharges from Direct Sources 	
Ethylene Dichloride Producers, Production Sites and
Capacities 	
Chloroform Discharges from Indirect Sources 	
Chlorodifluoromethane Producers and Production Sites 	
Chloroform Discharges from Use 	
Relative Source Contribution for Chloroform 	
Ambient Levels of Chloroform 	
Acute and Chronic Effects of Chloroform on Aquatic
Organisms 	
Values for knu 	
Un
Summary of EXAMS Models of the Fate of Chloroform 	
Physical Properties of Chloroform and Other
Chloromethanes 	
Partition Coefficients for Human Tissue at 37°C 	
Retention and Excretion of Chloroform in Man During and
After Inhalation Exposure to Anesthetic Concentations 	
Chloroform Content in United Kingdom Foodstuffs and
in Human Autopsy Tissue 	
Concentration of Chloroform in Various Tissues of Two
Dogs After 2 . 5 Hours Anesthesia 	
Concentration of Radioactivity (Chloroform Plus
Metabolites) in Various Tissues of the Mouse (N MRI) 	
1 4
Tissue Distribution of C-Chloroform Radioactivity
in CF/LP Mice After Oral Administration (60 mg/kg) 	
Pajje
3-3
3^
3-1 3

3-15
3-21
3-23
3-27
3-28
3-29

3-34
3-38
3-40

4-4
4-5

4-9

4-14

14-16

4-18

4-20
     vn

-------
                            LIST OF TABLES  (eont.)

Table                                                                      Page

                               1 3
4-8     Pulmonary Excretion of   CHCL_ Following Oral
           Dose:   Percent of Dose	   4-26

4-9     Species Difference in the Metabolism of   C-Chloroform	     4-29

4-4b    Kinetic Parameters for Chloroform After I.V. Administration
          to Rats	   4-30

4-10    Levels of Chloroform in Breath of Fasted Normal
          Healthy Men	   4-35
                                                1 ii
4-11     Covalent Binding of Radioactivity From   C-Chloroform and
            C-Carbon Tetrachloride in Microsomal Incubation
          In Vitro	   4-49

4-12    Mouse Strain Difference in Covalent Binding of Radioactivity
          From   C-Chloroform	     4-53

                                                       1 4
4-13    In Vivo Covalent Binding of Radioactivity From   CHC1
          in Liver and Kidney of Male and Female Mice (C57BL/6)	   4-55

                                                        1 4
4-14    In Vitro Covalent Binding of Radioactivity from   CHC1_
          to Microsomal Protein from Liver and Kidney of Male and
          Female Mice (C57BL/6)	   4-56

                                               1 4
4-15    Coyalent Binding of Radioactivity from   C-Chloroform and
            C-Carbon Tetrachloride in Rat Liver Nuclear and
          Microsomal Incubation I_n Vitro	   4-60

4-16    Effect of Glutathione, Air, N  or CO:  0- Atmosphere
          on the In Vitro Covalent Binding of C Cl^ and C Br Cl
          to Rat Liver Microsomal Protein	   4-62

4-17    Effects of 24-Hour Food Deprivation on Chloroform and
          Carbon Tetrachloride In Vitro Microsomal Metabolism,
          Protein, and P-450 Liver Contents of Rats	   4-64

5-1     Relationship of Chloroform Concentration in Inspired
          Air and Blood to Anesthesia	     5-2

5-2     Dose-Response Relationships	     5-7

5-3     Effects of Inhalation Exposure of Animals to Chloroform,
          Five Days/Week for Six Months	   5-17

5-4     Effects of Subchronic or Chronic Oral Administration of
          Chloroform to Animals	   5-19
                                      vm

-------
                             LIST OF TABLES (cont.)

 Table                                                                      Page

 5-5     Target Organ Toxicity  of Chloroform ............................     5-52

 7-1      Genetic Effects  of Chloroform  on Strain D7 of
           S.  Cerevisiae ................................................     7-1 o

 3-1      Effect of Chloroform on  Kidney Epithelial  Tumor  Incidence
           in  Osborne-Mendel Rats .......................................     8-5

 8-2     Effect of Chloroform on  Thyroid Tumor  Incidence  in  Female
           Osborne-Mendel  Rats ..........................................     8-7

         Toothpaste Formulation for  Chloroform  Administration ...........     8-9

         Effects of Chloroform on Hepatocellular Carcinoma Incidence
           in  B6C3F1  Mice ...............................................    8-13

 8-5     Kidney Tumor Incidence in Male ICI Mice Treated  with
           Chloroform [[[    8-17

 8-6      Liver and  Kidney  Necrosis and  Hepatomas in  Strain A Mice
           Following  Repeated Oral Administration of Chloroform
           in  Olive Oil .................................................    8-19

 8-7      SGPT  Changes in Beagle Dogs Treated with Chloroform ............    8-23

 8-8      Effect  of  Oral Chloroform Ingestion on  the  Growth of Ehrlich
           Ascites  Tumors ...............................................    8 -29

 8-9      Effect  of  Oral Chloroform Ingestion on  Metastatic Tumor Takes
           with  B1 6 Melanoma ............................................    8 -30

 8-10     The Effect of Oral Chloroform  Ingestion on  the Growth and
           Spread of  the Lewis Lung Tumor ...............................    8-31

 8-11     Correlation  Coefficients Between Residual Mortality
           Rates in White Males and THM Levels in Drinking Water
           by  Region  and by Percent of  the County Population
           Served in  the United  States ..................................    8 -4^4
8-12    Correlation Coefficients Between Bladder Cancer
          Mortality Rates by Sex and BTHM Levels in Drinking
          Water by Region of the United States
8-13    Risk of Cancer of the Rectum Mortality Associated

-------
                            LIST OF TABLES  (cont.)

Table                                                                      Pat:e
8-14
8-15
8-16
8-17
8-18
Cancer Risk Odds Ratios and 95? Confidence Intervals
(Chlorinated Versus Unchlorinated) 	
Incidence of Hepatocellular Carcinomas in Female and Male
B6C3F1 Mice 	
Incidence of Tubular-Cell Adenocarcinomas in Male
Incidence of Malignant Kidney Tumors in Male ICI Mice 	
Upper-Bound Estimates of Cancer Risk of 1 mg/kg/day.
8-54
8-71
8-72
8-72

          Calculated by Different Models on the Basis of Different
          Data Sets	   8-75

8-19    Relative Carcinogenic Potencies Among 53 Chemicals Evaluated
          by the Carcinogen Assessment Group as Suspect Human
          Carcinogens	   8-82

-------
                                LIST  OF FIGURES
 1-1      Rate  of  Rise  of Alveolar  (Arterial)  Concentration Toward
           Inspired  Concentration  For Five Anesthetic Agents  of
           Differing Ostwald Solubilities	   1-10

 1-2      Arteriovenous Blood Concentrations of a Patient During
           Anesthesia  with Chloroform	   1-11

 1-3      Exponential Decay of Chloroform, Carbon Tetrachloride,
           Perchloroethylene and Trichloroethylene in Exhaled
           Breath of 18 Year-old Male Accidentally Exposed to
           Vapors of These Solvents	   1-22

 1-1      Relationship  Between Total 8-Hour Pulmonary Excretion of
           Chloroform  Following 0.5-g Oral Dose in Man and the
           Deviation of Body Weight From Ideal	   1-27

 1-5      Blood and Adipose Tissue  Concentrations of Chloroform During
           and After Anesthesia in a Dog	   1-33

 1-6      Metabolic Pathways of Chloroform Biotransformation.
           (Identified CH Cl_ metabolites are underlines)	     1-36

 1-7      Metabolic Pathways of Carbon Tetrachloride
           Biotransformation	   1-12

 1-8      Rate of Carbon Monoxide Formation After Addition of Various
           Halomethanes to Sodium Dithionite-reduced Liver Microsomal
           Preparations From Phenobarbitol-treated Rats	   1-16

 1-9      Effect of Increasing Dosage of i.p.-Injected
            C-Chloroform on Extent of Covalent Binding of
           Radioactivity In Vivo to Liver and Kidney Proteins of Male
           Mice 6 Hours After Administration	   1-51

 1-10     Comparison  of irreversible binding of radioactivity from
            C-CHC1- to protein and lipid of microsomes from
           normal rabbit, rat, mouse, and human liver incubated
           in vitro  at 37°C in 0_	     1-57
                               fL

8-1      Survival curves for Fisher 311 Rats in a Carcinogenicity
           Bioassay  on Chloroform	    8-5

8-2      Negative Result in Transformation Assay of Chloroform
           which was also Negative in the Ames Assay	   8-35

8-3      Frequency distribution of CHC1  levels in 68 U.S.
           drinking water supplies.  The abscissa is linear in the
           logarithm of the level	   8-13
                                       XI

-------
                            LIST OF FIGURES (cont.)

                                                                           Page

8-4     Point and Upper-Bound Estimates of Four Dose-Response Models
          Over Low-Dose Region on the Basis of Liver Tumor Data
          for Female Mice	     8-76

8-5     Point and Upper-Bound Estimates of Four Dose-Response Models
          Over Low-Dose Region on the Basis of Liver Tumor Data
          for Male Mice	     8-77

8-6     Histogram Representing the Frequency Distribution of the
          Potency Indices of 53 Suspect Carcinogens Evaluated by
          the Carcinogen Assessment Group....	     8-81
                                     XII

-------
     The Office of Health and Environmental  Assessment  (OHEA),  U.S.  EPA,
is responsible for the preparation  of this  health  assessment  document.  The
Environmental  Criteria and Assessment Office (ECAO/RTP),  OHEA,  had  the
overall  responsibility for coordination  and  the  document  production  effort.

                             Project Manager

                            Si  Duk  Lee,  Ph.D.
               Environmental  Criteria and Assessment Office
              U.S. EPA,  Research  Triangle Park,  N.C.  27711
                              (919)  541-4159
                          Authors  and  Reviewers

               The  principal  authors of  this  document are:

                           Larry Anderson Ph.D.
                       Carcinogen  Assessment  Group
                        U.S.  EPA,  Washington, D.C.

                           David  Baylis, M.S.
                       Carcinogen  Assessment  Group
                       U.S.,  EPA,  Washington, D.C.

                           Chao W. Chen, Ph.D.
                       Carcinogen  Assessment  Group
                       U.S.,  EPA,  Washington, D.C.

                           Carol  Sakai, Ph.D.
                 Reproductive Effects Assessment Group
                       U.S.,  EPA,  Washington, D.C.

                        Sheila Rosenthal, Ph.D.
                 Reproductive Effects Assessment Group
                       U.S.,  EPA,  Washington, D.C.

                          I.W.F. Davidson, Ph.D.
                     Bowman  Gray  School of Medicine
                           Winston Salem, N.C.

                          D.  Anthony Gray, Ph.D.
                        Syracuse  Research Corp.
                             Syracuse, N.Y.

                          Sharon B. Wilbur, M.A.
                        Syracuse  Research Corp.
                             Syracuse, N.Y.

                         Joan P.  Coleman, Ph.D.
                        Syracuse  Research Corp.
                             Syracuse, N.Y.
                                      xm

-------
     The following individuals provided peer-review  of  this  draft or earlier
drafts of this document:

U.S. Environmental  Protection Agency

Joseph Padgett
Office of Air Quality Planning and Standards
U.S. EPA

Karen Blanchard
Office of Air Quality Planning and Standards
U.S. EPA

Jerry F. Stara, D.V.M.
Office of Health and Environmental  Assessment
Environmental Criteria and Assessment Office
U.S. EPA

Lester D. Grant, Ph.D.
Office of Health and Environmental  Assessment
Environmental Criteria and Asssessment Office
U.S. EPA

Participating Members of  the Carcinogen Assessment Group

Roy E. Albert, M.D. (Chairman)
Elizabeth L. Anderson, Ph.D.
Larry D. Anderson,  Ph.D.
Steven Bayard, Ph.D.
David L. Bayliss, M.S.
Chao W. Chen, Ph.D.
Margaret M. L. Chu, Ph.D.
Bernard H. Haberman, D.V.M., M.S.
Charalingayya B. Hiremath, Ph.D.
Robert E. McGaughy, Ph.D.
Dharm W. Singh, D.V.M., Ph.D.
Todd W. Thorslund,  Sc.D.

Participating Members of  the Reproductive  Effects Assessment Group

Peter E. Voytek, Ph.D. (Director)
John R. Fowle, III, Ph.D.
Carol Sakai, Ph.D.
Ernest Jackson, M.D.
K.S. Lavappa, Ph.D.
Sheila Rosenthal, Ph.D.
Vicki Vaughn-Dellarco, Ph.D.
                                           xiv

-------
External  Peer Reviewers
Dr. Karim Ahmed
Natural Resources Defense Fund
122 E. 42nd Street
New York, N.Y.   10168

Dr. Eula Bingham
Graduate Studies and Research
University of Cincinnati  (ML-627)
Cincinnati, Ohio  45221
(513) 475-4532

Dr. James Buss
Chemical Institute of Industrial
  Toxicology
Research Triangle Park,  N.C.   27709

Dr. I.W.F. Davidson
Wake Forest University
Bowman Gray Medical School
Winston Salem,  N.C.

Dr. Larry Fishbein
National Center for Toxicological
  Research
Jefferson, Arkansas  72079
(501) 542-4390

Dr. Derek Hodgson
Department of Chemistry
University of North Carolina
Chapel Hill, N.C.  27514

Dr. Marshall Johnson
Thomas Jefferson Medical  College
Department of Anatomy
1020 Locust Street
Philadelphia, Pennsylvania  19107

Dr. Trent Lewis
National Institute of Occupational
  Safety and Health
26 Columbia Parkway
Cincinnati, Ohio  45226
(513) 684-8394

Dr. Richard Reitz
Dow Chemical, USA
Toxicology Research Laboratory
1803 Building
Midland, Michigan  48640
Dr. Marvin A. Schneiderman
Clement Associates, Incorporated
Arlington, Virginia  22209
(703) 276-7700

Dr. Bernard Schwetz
National Institute of Environmental
  Health
Research Triangle Park, NC  27709
(919) 541-7992

Dr. James Selkirk
Oak Ridge National Laboratory
Oakridge, Tennessee  37820
(615) 624-0831

Dr. Samuel Shibko
Food and Drug Administration
Division of Toxicology
200 C Street, S.W.
Washington, D.C.  20204
Telephone:

Dr. Robert Tardiff
1423 Trapline Court
Vienna, Virginia  22180
(703) 276-7700

Dr. Norman M. Trieff
University of Texas Medical Branch
Department of Pathology, UTMB
Galveston, Texas  77550
(409) 761-1895

Dr. Benjamin Van Duuren
Institute of Environmental Medicine
New York University Medical Center
New York, New York  10016
(212) 340-5629

Dr. James Withey
Health and Protection Branch
Department of National Health &
  Welfare
Tunney's Pasture
Ottawa, Ontario
CANADA, KIA 01Z
                                         xv

-------
                  6.  TERATOGENICITY AND REPRODUCTIVE EFFECTS





     Schwetz et al. (1974) evaluated the effects of reagent grade chloroform



 (lot no. 9649, Burdick and Jackson Laboratories, purity not reported) on the



maternal and fetal well-being of Sprague-Dawley rats.  Twenty female rats were



exposed, by inhalation (7 hr/day), to 30, 100, and 300 ppm chloroform on days



6 through 15 of gestation.  The authors analyzed the results statistically



using Fisher's Exact Probability test, analysis of variance, Dunnett's test,



or Tukey's test to compare the frequency of anomalies, resorptions, maternal



and fetal weights, body lengths, liver weights, or serum glutamic pyruvic



transaminase (SGPT) activity in the exposed versus the control  groups.  The



level of significance was chosen at p < 0.05, and the litter was used as the



experimental unit.



     When the animals were exposed to the highest dose of chloroform (300 ppm),



there was a significant increase in the number of resorptions and a decrease in



the conception rate (Schwartz et al., 1974).   At the lower doses (30 and 100



ppm), no alterations in resorption rate, fetal body weight, conception rate,



number of implantations,  or average litter size were observed.   Fetal crown-



rump length was significantly decreased at 30 and 300 ppm, but  not at the 100



ppm level.  At 100 ppm, an increase in the incidence of acaudia (absence of



tail),  short tail, imperforate anus,  subcutaneous edema,  missing ribs, and



delayed ossification of sternebrae were observed.  At 300 ppm,  subcutaneous



edema and abnormalities of the skull  and sternum were observed, but the



incidence of these was not statistically significant.   The authors pointed out



that the small  numbers of survivors in the 300 ppm group  (4+7 versus the



control  10+4 live fetuses/litter)  may have prevented adequate statistical



evaluation of this effect.
                                      6-1

-------
     In this study (Schwetz et al., 1974),  chloroform also produced maternal



toxicity, such as a decrease in the rate of maternal  weight gain  at all  dose



levels and a decrease in food consumption during pregnancy at  the 100 and 300



ppm level.  Other maternal  effects observed were changes  in liver weight gain



during pregnancy (no change in absolute liver weight  gain at 30 ppm,  but an



increase at 100 ppm, possibly due to the concomitant  anorexia  at  this dose).



No significant changes in SGPT activity were observed in  groups exposed  to



chloroform at the 100 and 300 ppm levels (the only two doses evaluated).



Since the developmental  effects observed in the 300 ppm group  were associated



with anorexic effects in the mother, a starvation control  group was included



in this study.  This starvation control group was restricted to food consumption



comparable to the 300 ppm chloroform group.  Animals  on a starvation diet



(allowed 3.7 g/day of food  on days 6 through 15 compared  with  control  animals



whose food consumption average 19-25 g/day  on days 6  through 15)  had a signifi-



cant decrease in the absolute weight of the liver and an  increase in the relative



weight of the liver.  The effects of 300 ppm chloroform on the increase  in the



relative weight of the liver were much greater than starvation alone.  Addition-



ally, and perhaps most importantly, exposure to 300 ppm chloroform resulted  in



a dramatic decrease in the  number of animals pregnant at  sacrifice (15%  pregnant



versus 88% in air control), a decrease in the number  of live fetuses per litter



(4 versus 10 live fetuses/litter), and an increased percentage of resorptions



(100% vs. 57%).  Examination of the uteri indicated that  the conceptus had



been completely resorbed very soon after implantation.  These  effects appeared



to result primarily from chloroform exposure, and not maternal  toxic influence,



since anorexia and liver weight changes associated with starvation were  not



accompanied by embryotoxic  and teratogenic  effects.
                                      6-2

-------
     Murray et al.  (1979) evaluated the effects of chloroform (spectral  grade,



Mai 1inckrodt, lot  CSZ, code 4434, purity not reported) administered by inhalation



(7 hr/day) to CF-1  mice.  Thirty-five animals per group were exposed on  gesta-



tion  days 1 through 7 and 6 through 15; forty animals per group were exposed



on gestation days  8 through 15.  Only one dose, 100 ppm, was tested during



three different time periods (days 1 through 7, 6 through 15, 8 through  15 of



gestation) in CF-1  mice.  The varying exposure periods were designed to  evaluate



the effects of chloroform in very early pregnancy, organogenesis,  and somewhat



later in pregnancy.  Special sodium sulfide staining of the uteri  was used to



detect very early  pregnancies.



     The authors (Murray et al., 1979) analyzed the results statistically



using the Fisher's  Exact Probability test to evaluate pregnancy incidence; the



modified Willcoxan  test for fetal outcomes; the Mann-Whitney signed rank test



for SGPT activity;  and one-way  analysis of variance for fetal body weights and



body  measurements,  maternal body weights, liver weights, food consumption, and



number of implantations and resorptions.  The level  of significance was  chosen



at p  < 0.05.



     Murray et al.  (1979) reported that 100 ppm chloroform resulted in a decrease



in the total number of pregnancies when the animals were exposed on days 1-7 or



6-15  of gestation  but not on days 8-15 (see Table 6-1 for summary  of data).  In



the pregnant animals, however,  there was no significant effect on  the average



number of implantation sites.  In animals exposed on days 1-7 of gestation, but



not in those exposed on days 6-15 or 8-15, there were significant  increases in



resorptions per litter.  This effect was accounted for by the loss of two



entire litters.  Mean fetal body weight and crown-rump length were significantly



decreased in the groups exposed on days 1-7 and 8-15, but not in those exposed




on days 6-15.  Maternal toxicity (slight decrease in body weight gain during





                                      6-3

-------
                         TABLE 6-1.  SUMMARY OF EFFECTS
                             (Murray et al., 1979)


Chemical:  Chloroform, spectral  grade, Lot CSZ, Code 4434, Mallinckrodt,  Inc.
Animal:  CF-1 mice, 35 animals in groups exposed on days 1-7 and 6-15;
         40 animals in groups exposed on days 8-15
Route of exposure:   Inhalation,  100 ppm (one dose only)
Duration of exposure:   7 hr/day, days 6-15 of gestation
               Pregnant
Days      (Implantation sites)
Additional  pregnancies
   (special  stain)
 (number of animals)
Total pregnancies
(implantation sites
and special  stain)
          Exposed      Control       Exposed

1-7     11/34 (32%)  22/35 (63%)       4

6-15    13/35 (37%)  29/34 (85%)       2

8-15    18/40 (45%)  25/40 (62%)       6
           Control       Exposed      Control

              4       15/34 (44%)  26/35 (74%)

              2       15/35 (43%)  31/34 (91%)

              1       24/40 (60%)  26/40 (65%)
                                                (continued on the following page)
                                6-4

-------
                            TABLE 6-1.  (continued)
Days
     1-7
     6-15
    8-15
Fetal  effects
resorption

fetal  body weight
and crown-rump
length

   	*

delayed skeletal
ossification
                                                               fetal body weight
                                                               and crown-rump
                                                               length

                                                               cleft palate

                                                               delayed skeletal
                                                               ossification
Maternal effects
body weight gain
body weight gain

liver weight

SGPT (only one
dose)
body weight gain

1iver wei ght
*Not significantly different.
                                      6-5

-------
pregnancy) was seen in groups  exposed  on  days  6-15,  with  a  more  severe  decrease



in groups exposed on days 1-7  and  8-15.   Less  food  and  water1  were  consumed  in



all  experimental  groups as compared  to controls.  Absolute  and  relative weight



of the liver were increased in groups  exposed  on  days  6-15  and  8-15,  hut not



in those exposed  on days 1-7.   Serum glutamic  pyruvic  transaminase (SGPT)



activity was increased in mice exposed on days 6-15, which  was  the only time



period evaluated  for this measurement.



     A summary of this study (Murray et  al.,  1979)  is  presented  in Table 6-1.



One major result  was that chloroform,  under the conditions  of this experiment,



caused a decrease in pregnancy.  The authors  concluded  that chloroform  affected



the stages either prior to, or in, early  implantation.   However, because of the



small numbers of  animals and the lack  of  dose-response  evaluation  (only one



dose was tested,  100 ppm), this conclusion must be  considered tentative until



future studies confirm this observation.   Other results of  this  study (Murray



et al., 1979) indicated that the incidence of  cleft palates increased in pups



exposed in utero  on days 8-15 of gestation, but not on  days 1-7 or 6-15.  The



authors (Murray et al., 1979)  suggested three  possibilities to  explain  this last



result.  The first was that earlier exposure on days 6-15 prevented susceptible



concept! from implanting.  The second possibility was  that  the  number of litters



available (11 in  the group exposed on days 6-15) was insufficient  to detect



this effect.  The third was that the teratogenic effect (cleft  palate)  did not



occur in concepti exposed on days 1-7 since they were  exposed before organo-



genesis.  The number of offspring coming to term was consistently less  in all



exposure groups than in the controls  (days 1-7, 9 litters versus 22 in  control;



days 6-15,  11 litters versus 29 in control; days 8-15, 18 litters  versus 24 in



control).   Therefore it was not clear whether chloroform produced teratogenicity




separate from embryotoxicity.  Since the pups with cleft palate in groups





                                      6-6

-------
were also retarded in growth, it was suggested that the ability of chloroform



to cause malformations was indirect and not direct.  However, this was not



experimentally determined, and it is not known if the tendency for lowered



fetal weight and/or delayed skeletal growth is correlated with a higher



incidence of malformations.



     In conclusion, this study (Murray et al., 1979) indicated that chloroform



administration (100 ppm) by inhalation (7 hr/day) produced teratogenic and



embryotoxic effects and interfered with pregnancy in addition to causing



maternally toxic effects (changes in liver weight and decreases in weight gain



during pregnancy).  Exposure in the early stages of pregnancy appeared to



produce a decreased incidence of conception, but the results of this study did



not conclusively determine which days of pregnancy were most susceptible to the



effects of chloroform.  To answer this question, it would be necessary to use



a greater number of doses and larger numbers of animals.



     Thompson, Warner, and Robinson (1974) investigated the effect of chloroform



administered orally, using Sprague-Dawley rats and Dutch-Belted rabbits.  The



rats were intubated with chloroform (Mai 1inckrodt, Batch ZJL dissolved in corn



oil, purity not reported) twice a day in divided doses  of 20 to 501 mg/kg/day.



The rabbits were intubated once a day in doses of 20 to 398 mg/kg/day.  Each



study was divided into two parts, a range-finding portion designed to establish



the proper dose range (six rats were administered 79, 126, 300, 316,  and 501



mg/kg/day of chloroform; five rabbits were administered 63, 100,  159, 251, and



398 mg/kg/day), and a teratology study,  using  greater numbers of  animals and



three doses (25 rats were administered 20, 50, and 126  mg/kg/day).   The rats



were exposed to chloroform on days 6 through 15 of gestation, while the rabbits



were exposed on days 6 through 18 of gestation.
                                      6-7

-------
     Statistical  evaluations of maternal  body weight gains,  food consumption,



implantations, corpus luteum, resorptions,  litter size,  and  fetal  weights were



made by an analysis of variance and Dunnett's Two-Tailed Multiple Range test.



Sex ratios and frequency of anomalies among the fetal  population and among



litters were analyzed by the chi-square test.  In all  analyses, the level of



significance chosen was p < 0.05.



     The data for the range-finding portion of this study (Thompson et al.,



1974) was not presented; however,  the authors reported that  rats treated orally



with greater than 126 mg/kg/day chloroform had signs of maternal toxicity, such



as a decrease in food consumption, acute toxic nephrosis, hepatitis, and gastric



erosion.  Fetal development was adversely affected in rats receiving 316 and



501 mg/kg/day with a decrease in fetal viability, litter size, and fetal weight



and an increase in the number of resorptions.  Only two rats survived when



given 501/mg/kg/day; one was not pregnant, the other had complete early



resorptions.  In the rat teratology study, animals receiving 50 and 126 mg/kg/day



displayed signs of maternal toxicity  (lowered body weight gain, lowered food



consumption, fatty changes in the liver).  No overt toxic effects were observed



in animals given 20 mg/kg/day, and no malformation was noted at any dose level.



     In the range-finding portion of  the study by Thompson et al.  (1974) using



rabbits, severe acute hepatitis and nephrosis were observed in animals given



63 mg/kg/day and higher.  No overt signs of toxicity were observed at the



25 mg/kg/day level.  In the two surviving dams given 100 mg/kg/day, one had four



resorption sites with no viable concepti, while the other was not  pregnant.  No



other embryotoxic or teratogenic effect was observed.  In the teratology study



of rabbibs, maternal toxicity  (depressed weight gain) was observed at the  50



mg/kg/day level.   In the fetus, mean  body weight was depressed at  the 20 and




50 mg/kg/day levels, but no abnormalities were observed.  Incomplete skeletal





                                      6-8

-------
ossification at the 25 or 35 mg/kg/day level  was observed amongst fetuses, but



not amongst litters when analyzed statistically.



     In this study (Thompson et al., 1974), adverse effects resulting from



chloroform exposure, such as skeletal  deformities and deficiencies in pregnancy



maintenance that were observed by Schwetz et  al. (1974) and Murray et al. (1979),



were not observed except at doses acutely toxic to the dam.  The authors



suggested that the greater maternal toxicity  observed in this study compared



to that of others (Schwetz et al., 1974; Murray et al., 1979) could be explained



by the route or duration of exposure.   In the study by Thompson et al. (1974),



rats and rabbits were exposed orally to chloroform or twice a day, while Schwetz



et al. (1974) and Murray et al. (1979) administered chloroform by inhalation



7 hr/day.  The lack of information on the pharmacokinetic interaction of chloro-



form,  however, prevents an evaluation of the role of exposure in producing



adverse reproductive outcome.



     Burkhalter and Ralster (1979) evaluated the potential of chloroform to



adversely affect behavior in developing ICR mice.  This study was designed as a



preliminary screening study, with the parental generation of male and female



mice exposed 21 days prior to mating, during mating, and for an additional 21



days.  The offspring were exposed starting on day seven until day 21 after



birth.  Only one oral dose of chloroform, 31.1 mg/kg/day, was administered to



five control and five experimental animals.  Each litter was reduced to eight



pups,  and three pups were randomly selected for behavorial teratogenic testing.



The chloroform  (Mai 1inckrodt, nanograde purity) was administered by gavage and



delivered  in a solution of 1 part polyoxyethylated vegetable oil, Emulphor



(EL-620, GAP Corp., New York), and 8 parts saline.  A  variety of behavorial



responses were evaluated which included:  righting reflex, forelimb placing




response,  forepaw  grasp, rooting  reflex, cliff  drop aversion, auditory startled





                                      6-9

-------
response, bar-holding ability, eye opening,  motor performance,  and learning



ability.  The scoring for these responses was based upon predesignated criteria.



These criteria established behavioral  ability by measuring both objective



standards (time to complete test) and  subjective standards ("weak" or "complete"



grasp of paws).



     Analysis of variance was used to  statistically evaluate tests of passive



avoidance.  Screen test latencies were analyzed by t-test.  The data from the



neurobehavioral developmental scale were analyzed using a Mann-Whitney U test.



The level of statistical  significance  was chosen at p < 0.05.



     Burkhalter and Ralster (1979) reported  that the sizes of  litters were



similar for both the control  and experimental groups; however,  fetal body weight



gain of pups during the 14 days of exposure  (days 7-21 following birth) was



decreased.  Forelimb placement response was  reduced in the exposed group on



day 5 and 7 of birth, but not on day 9.  The significance of this reduction is



not known, although the recovery on day 9 suggested that the effect was



reversible.  The other behavorial responses  were not significantly different



in the exposed groups.  Burkhalter and Balster (1979) concluded that 31.1



mg/kg/day of chloroform produced no significant adverse behavorial effects in



pups exposed both in utero and after birth (days 7-21).  However, since this



study was designed as a preliminary screening study, using one dose and small



numbers of animals, there was no attempt to  evaluate the full  range of dose-



related effects.  In future studies, it would be desirable to evaluate at



least three doses, including doses high enough to produce some maternal toxicity,



in addition to using a larger sample size.



     In an abstract, Dilley et al. (1977) reported the effects of exposure to



chloroform on pregnant rats  (strain not reported, number of animals not reported),




The animals were exposed by inhalation (20 +_ 1.2 gm3/day) on days 7-14 of





                                      6-10

-------
gestation.  Two lower concentrations were administered in the study, but the



doses were not reported in the abstract.   Dilly reported that chloroform



increased fetal mortality and decreased fetal  weight gain; however, there were



no malformations.



     In another abstract, Ruddick et al.  (1980) exposed 15 Sprague-Dawley rats



to 100, 200, and 400 mg/kg/day of chloroform by gavage on days 6-15 of gestation.



Chloroform (purity not reported) was reported to cause maternal  toxicity



(changes in weight gain, biochemical and  hematological parameters, and liver or



kidney changes).  Chloroform also produced adverse effects in fetal development



(type not specified), but the authors attributed these effects primarily to



maternal toxicity and not directly to chloroform exposure.  Without the



presentation of this data, however it is  not possible to fully evaluate



these results.



6.1  SUMMARY



     In summary, the results of four articles and two abstracts indicated that



under the conditions of the experiments,  chloroform has the potential for



causing adverse effects in pregnancy maintenance, delays in fetal  development



and  production of terata in laboratory animals.  The adverse effects on the



conceptus were observed in association with maternal toxicity.  The type and



severity of effects appeared to be specific to the conceptus, affecting the



fetus to a much greater degree than the mother.  Therefore, it was concluded



that chloroform has the potential for causing embryotoxicity and teratogenicity



(Schwetz et al., 1974; Murray et al., 1979).  The results of other studies



indicated that chloroform has no significant effect on neonatal  behavior



(Burkhalter and Balster, 1979) and does not cause adverse fetal  effects except



at maternally toxic levels (Thompson et al., 1974).  The two abstracts did not




contain enough detail for critical scientific review  (Oil ley et al., 1977;





                                      6-11

-------
Ruddick et al.,  1980).



     Studies administering chloroform by inhalation for 7 hr/day (Schwetz  et



al., 1974; Murray et al., 1974) reported more severe outcomes than other studies



which administered chloroform by intubation once or twice a day (Thompson  et



al., 1974; Burkhalter and Balster, 1979).  However, since the pharmocokinetic



relationship associated with route or duration of exposure have not been studied,



it is not possible to evaluate the importance of the route of exposure in



causing adverse  reproductive outcome.  To evaluate more fully the influence of



these factors,  additional investigations would have to be conducted.
                                      6-12

-------
6.2  REFERENCES

Burkhalter, J., and R.L. Balster.  1979.  Behavioral  teratology evaluation of
     chloroform in mice.  Neurobehavioral  Toxicol. 1:199-205.

Dilley, J.V., N. Chernoff, D. Kay, N.  Winslow, and E.W. Newell.  1977.  Inhalation
     teratology studies of five chemicals  in rats.  Toxicol.  Appl.  Pharmacol.
     41:196.

Murray, F.A., R.A. Schwetz, J.G. McBride,  and R.E. Staples.   1979.   Toxicity of
     inhaled chloroform in pregnant mice and their offspring.   Toxicol.  Appl.
     Pharmacol. 50:151-522.

Ruddick, J.A., D.C. Villenouve, I. Chu, and V.E.  Balli.  1980.   Teratogenicity
     assessment of four halomethanes.   Teratology 21.-66A.

Schwetz, B.A., B.K.J.  Leong,  and P.J.  Gehring.  1974.   Embryo-  and  fetotoxicity
     of inhaled chloroform in rats.  Toxicol.  Appl.  Pharmacol.  28:442-451.

Thompson,  D.O., S.O.  Warner,  and V.B.  Robinson.   1974.   Teratology  studies  on orally
     administered chloroform  in the rat and rabbit.   Toxicol. Appl.  Pharmacol. 29:
     348-357.
                                     6-13

-------
                               7. MUTAGENICITY
7.1   TNTROniJCTION
    The nutagenic potential of chloroform (CHC^) has been assessed by
evaluation of the results from five jn yjtro bacterial studies, one
host-mediated assay using Sajnonella as the indicator organism, one yeast
study, one Drosophila sex-linked recessive lethal test, one in vitro mammalian
cell  mutagenicity assay, two sperm head abnormality tests, three chromosome
aberration studies, and six DNA damage studies (sister chromatid exchange and
unscheduled DNA synthesis).  These reports are discussed below.  Also, several
assays from a recently published screening study, in which 42 chemicals were
tested in various short-term protocols, are briefly discussed.  The majority
of the above studies were negative.  Information relating to the binding of
metabolically activated CHC13 to cellular macromolecules is presented before
the sections assessing the genetic damage caused by CHC13.  This was done to
set the stage for the discussion of the negative mutagenicity studies
described below and to support the suggestion  that CHC13 may be a weak
mutagen.  In addition, suggestions for further testing are presented.
7.?.   CnVALENT RINDING TO MACROMOLECULES
    As mentioned in Chapter 4, the primary reactive metabolite of CHC13 is
phosgene, COCl^.  Phosgene is a crosslinking agent and may covalently bind
to and crosslink macromolecules.   Thus, the  toxicity and carcinogenicity of
CHC13 nay be related to its metabolism to phosgene.   The DNA binding
potential  and carcinogenicity of  phosgene are  currently  under investigation in
Or. R.L. Van Duuren's  laboratory  at New York University  Medical  Center.
Preliminary  evidence indicates that phosgene binds to DNA (Dr.  Sipra Ranerjee,
New York University Medical  Center,  personal communication).   The  binding
potential  of metabolically activated CHC13 has  been  assessed  in  several
                                     7-1

-------
studies.  Some of these studies have been described in Chapter 4, section 4.6.



Additional studies and studies not adequately covered in Chapter 4 for the



purposes of this chapter on mutagenicity are described below.



      Diaz Gomez and Castro (1980a) assessed the CHC13 activation potential



of purified rat liver nuclei by measuring covalent binding of nuclear-



activated CHC13 to nuclear protein and lipid.  The results were compared to




results obtained from similar incubation mixtures containing microsomes



instead of purified nuclei.  The incubation mixtures containing either nuclei



(1.3 mg protein/ml) or microsomes (1.56 mg protein/ml) were incubated for 30



min in 6.4 nM 14CHC13 (5.4 Ci/mol) and an NADPH generating system.  The



authors observed that the extent of binding to protein in the nuclear



preparation was approximately 40% of that observed for microsomes (nuclei, 27




pmol/mg; microsomes, 68 pmol/mg).  Rinding to nuclear lipid was approximately



35% of that observed for microsomes (nuclei, 20 pmol/mg; microsomes, 57




pmol/mg).  Thus, isolated nuclei were less efficient than microsomes in



metabolizing CHC13, but the results were within the same order of magnitude.



     This study suggests that metabolism of CHC13 to a reactive



intermediate(s) can occur in nuclear membranes, as may be the case with other



xenobiotics (Weisburger and Williams, 1982).  Tt should be mentioned, however,



that the nuclear preparations were contaminated with trace amounts of



endoplasmic reticulum,  which may have been sufficient to result in at least



part of the nuclear activation observed.




     In a subsequent study, niaz Gomez and Castro (1980b) exposed rat or mouse



liver HNA or RNA to l^CHCl 3 in_ vjj/o or JJT_ vitro without finding any




significant binding of  14C to the nucleic acids.  However, the specific



activity of the l^CHC^ was only 5.4 Ci/mol, which may have been too low




to allow for observation of binding to DNA (Brookes and Lawley, 1971),
                                     7-2

-------
especially with a background count of 160 dpm.



    Reitz et_ aj_. (1980) published results of DNA alkylation studies in livers



and kidneys of male mice that were exposed orally to CHC13 at 240 mg/kg.  A



very small amount (1-3 x 10~6 alkylations/deoxynucleotide) of alkylation was



observed.  However, these results cannot be interpreted because the specific



activity of the l^CHC^ used was not specified and because the



experimental procedures used were not described.



     In summary, binding of metabolically activated CHC13 to liver



microsomal and nuclear protein and lipid has been observed.  The only studies



that attempted to measure binding to nucleic acids were inconclusive.



7.3  MUTAGENICITY STUDIES IN BACTERIAL TEST SYSTEMS



    Uehleke et_ a_L  (1977) tested CHd3 for mutagenicity in suspension assays



with _S. typhimurium strains TA1535 and TA1538.  No nutagenic activity was



observed.  About 6-9 x 10^ bacteria  were incubated for 60 min under N£ in



tightly closed test tubes with 5 mM  CHC13 and microsomes (5 mg protein)  plus



an NADPH generating system.  The mutation frequencies (his+ colony forming



units/10^ his" colony forming units)  were less than 10 for both strains



and the spontaneous mutation frequencies were 3.9 +_ 3.7 for strain TA1535 and



4.4^ 3.5 for strain TA1538.  At this concentration of CHC13,  survival  of



the bacteria was at least 90%.   Additional  higher concentrations should  have



been tested, because the mutagenic range can  occur at higher toxicities.



Dimethylnitrosamine (50mM),  cyclophosphamide  (0.5mM),  3-methylcholanthrene



(0.1 mM), and benzo[ajpyrene (0.1 mM) were the positive controls used  in  this



study.   Information was not  provided  on  the  survival  of the bacteria  at these



concentrations of  the positive  control  chemicals.   Although these chemicals



were mutagenic in  the presence  of the S9 activation  system,  they may  not  be



appropriate as controls for  CHC13 because they  are  not  halogenated  alkanes
                                     7-3

-------
and therefore are not metabolized like them.



    Studies demonstrating that rnetabolically  activated CHC13 hinds  to




protein and lipid in the presence of rabbit nicrosomes were also described in



the paper by Uehleke et_ aj_. (1977) and are mentioned in Chapter 4 of this




document.  However, it is not clear from the description provided in this




report by Uehleke et_ aj_. (1977) whether rat,  mouse,  or rabbit microsomes were



used in the mutagenicity studies.  If mouse or rat microsomes rather than



rabbit microsomes were used for the mutagenicity experiments, it cannot be



assumed that CHC13 was sufficiently activated, since activation sufficient



for binding of l^CHClg ^0 macronolecules was shown in this paper only with




rabbit microsomes.  Another deficiency in this study is that the Ames strains



TA98 and TA10D were not used.  These strains contain an R factor plasmid that



increases the sensitivity of the tester strains to certain mutagens.



    The mutagenicity of CHC13 was also tested in a study designed to




evaluate the mutagenic potential of chemicals identified in drinking water



(Simmon et_ a_K , 1977).  No mutagenic activity was detected with CHC13.  The



authors tested 71 of the 300 chemicals that had been identified in public



water supplies.  CHC13 was tested at 10% by volume  (1.24 M) in a suspension



assay with Salmonella strains TA1S3S, TA1537, TA1538, TA9R, and TA100 and S9



mix prepared from Aroclor  1254-treated rats.  This  concentration of CHC13



exceeds  its solubility.  Mutagenic activity was not observed.  However,



information on toxicity was not  provided.



    CHC13 was also  tested  in this study  in a desiccator to  assess



mutagenicity due to  vapor  exposure  (Simmon et_ aj_.,  1977).   Agar plates were



placed uncovered  in  a desiccator above a  glass  petri dish  containing the



CHC13.   The desiccator  contained a magnetic stirrer which  acted as  a fan  to



aid in evaporation  of the  measured  amount  of CHC13  and to  maintain  an even
                                      7-4

-------
distribution of the vapors. Plates were exposed to the vapors for 7-10 hr and



then removed from the desiccators, covered, and incubated approximately 40 hr



before scoring.  As in the suspension assay described above, mutagenic



activity was not observed and no information on toxicity was provided.



     This study by Simmon et_ aj_. (1977), although lacking some specific



details of the CHC13 assay, clearly identifies other trihalomethanes



(CHBT3, CHBT2C1, CHBrCl?) as mutagens in the vapor assay in desiccators.



Methyl bromide, methyl chloride, methyl  iodide, and methylene chloride were



also found to be mutagenic in the desiccator assay.  However, these seven



halogenated compounds did not require metabolic activation to exhibit



mutagenic activity.  It may be that CHC13 itself is not mutagenic and the



rat liver S9 was not sufficient to metabolize CHC13 to a potential  mutagenic



reactive intermediate (? phosgene), even though the demonstration of



mutagenicity of three of the chemicals tested [bis(2-chloroisopropyl)ether,



vinyl chloride, and vinylidene chloride] required or was enchanced by this S9



mix.  Because CHC13 is a liver carcinogen in the mouse and not in the rat



(NCI bioassay, 1976), mouse liver microsomes may be more appropriate than rat



liver microsomes as a component of an activation system for CHC13



mutagenesis.  It may also be that a reactive intermediate was formed, but it



was too reactive or short-lived to be detected in a test system that uses



exogenous metabolic activation.



    Kirkland et_ aj[. (1981)  studied the mutagenicity of CHC13 in Escherichia



coli strains WP2p and WP2uvrA"p, using reversion to tryptophan prototrophy



as the endpoint.  The bacteria were treated with CHC13 in plate



incorporation and preincubation tests both with and without rat liver



microsomes (plus cofactors)  prepared from Aroclor 1254-induced rats.  The



concentration of protein in  the microsomal  suspension  was not given.  CHC13
                                     7-5

-------
was added at 10,000, 1,000, 100, 10, 1, or 0.1 ug/plate.   Negative results



were obtained in both tests.  However, there was no indication that



volatilization of CHC13 was prevented in the preincubation test.   Also,  it



appears from the description of the protocol for the plate incorporation test



that the procedure used to prevent loss of CHC13 was inadequate.   CHC13



was added to suspensions of the bacteria in molten agar,  and each mixture was



rapidly mixed on a Whirlimixer and poured onto agar plates.  The  plates  were



then incubated in gas-tight containers.  Excessive evaporation of CHC13  may



have occurred during the mixing of the molten agar/bacteria/CHCl3



suspension.  2-Aminoanthracene was used as a positive control  requiring



netabolic  activation and N-methyl-N1-nitro-N-nitrosoguanidine was used  as the



positive control not requiring activation.  These chemicals are not volatile



and are therefore inadequate positive controls for CHC13.  Also,  it cannot



be assumed that a microsomal activation system that metabolizes



2-aminoanthracene is sufficient to metabolize CHC13.



    Gocke eit_ a_U (1981) assessed the mutagenicity of 31 chemicals (including



CHC13) used as ingredients in European cosmetics.  Three test systems were



used:  the Salmonel1 a/mi crosome test, the sex-linked recessive lethal test in



Orosophija, and the micronucleus test for chromosome aberrations  in mice.  The



latter two tests will be discussed in the following sections.



    For the Sajmonel1 a/microsome assays at least five doses of each compound



were tested, usually up to 3.6 mg/plate for nontoxic and soluble  compounds.



Salmonella strains TA1535, TA100, TA1538, TA98, and TA1537 were used with and



without activation by S9 mix prepared from Aroclor-pretreated rats.  Because



this was a screening study in which 31 chemicals were tested,  details of the



assay protocols were not given.  Three halogenated aliphatic hydrocarbons were



tested (1,1,1-trichloroethane, dichloromethane, and  CHC^).  Because of the
                                     7-6

-------
 volatility of these compounds, the bacteria were exposed  in airtight
 desiccators for 8 hr.  The composition and purity of these chemicals were not
 specified.
    The first two substances exhibited mutagenic activity with and without
 metabolic activation.  CHC13 was  inactive.  However, as discussed above in
 the evaluation of the Simmon et^ aj_.  (1977) paper, CHC13 may require
 metabolic activation to be mutagenic, and the rat liver S9 preparation may not
 have been sufficient.  Also, if a reactive metabolite was formed, it may have
 been too reactive to be detected  under conditions of exogenous activation.
 Phosgene, the primary reactive metabolite of CHC13, is very reactive and
 unstable (Kirk-Othmer, 1971).
    In a recently published screening study of 42 chemicals (de Serres and
 Ashby, 1981), CHC13 was evaluated in 38 jn vivo and jn vitro short-term
 tests designed to assess potential genotoxicity.  Results from bacterial
 assays carried out in 18 laboratories using Salmonejla (Ames reversion test)
 or 1E_. coli  (forward mutation test) were essentially negative.   However, these
 results are inconclusive, because nowhere was it mentioned in  the protocols
 that excessive volatilization and escape of CHC13 from the culture dishes
 was prevented.  In addition, the problems with external  activation systems
 mentioned above also apply to the bacterial  assays  carried out in this
 screening study.
    Agustin and Lim-Sylianco (1978)  investigated the mutagenicity of CHC13
 in a host-mediated assay using Salmonella strains TA1535 and TA1537  as the
 indicator organisms  that were injected  into  male and female  mice.   The authors
 found that  male mice metabolized the  CHC13 to a  mutagen  active in strain
TA1537.   However,  they  reported only  the  ratios  of  mutation  frequencies for
treated  vs.  control  animals  and gave  no indication  of  the  actual  colony counts
                                     7-7

-------
observed.  The mutation frequency (tested/control)  for  strain  TA1537  in  male
mice was 36.75 and that for female mice, 2.30.   The mutation frequencies for
strain TA1535 were 0.61 and 0.12, respectively.   Details  of  the  procedures
used (i.e., doses of CHC13, numbers of bacteria  injected  and recovered,
route of exposure, and time of exposure before  the  animals were  killed)  were
not presented.  Thus, although there is suggestive  evidence  of a positive
response, definitive conclusions cannot be reached  because of  inadequacies  in
the way the data were reported and because details  of the procedures  used were
not provided.
     Agustin and Lim-Sylianco (1978) also studied the mutagenicity of
ether-extracted urine concentrates from 10 male mice in a bacterial  spot test,
using strain TA1537.  The mice were exposed to  CHC13 at 700  mg/kg.  Urine
concentrates from CHCl3-treated mice yielded 302 revertant  colonies and  a
zone of inhibition of 29 mm, whereas urine from control animals  yielded  10
revertant colonies with no zone of inhibition.   Details of  the ether
extraction procedure were not provided, but the likelihood  of  a  false positive
result due to the presence of histidine in the  extracted urine is unlikely,
because the urine concentrate from the control  animals yielded only 10
colonies and was presumably subjected to the same extraction procedure.
    In summary, the results from the above bacterial studies are inconclusive,
because false negative results could have been  obtained due  to a number  of
factors, including:
    1.  The activation systems used may have been inadequate for metabolism of
        CHC13.
    2.  Phosgene, the primary reactive metabolite of CHC13,  is unstable  and
        highly reactive.  Because exogenous activation systems were used in
        many of these studies, any phosgene generated  (assuming an adequate
                                     7-8

-------
         activation  system) may  have  been  scavenged  by microsomal  protein  or



         lipid  before  reaching the  DNA.



     3.   Adequate  exposure to CHC13 may  not  have  occurred  if  appropriate



         precautions were not taken to prevent the evaporation of  CHC13.



 It  is, of  course, also  conceivable that the negative results reflect the



 possibility that  CHC13  is not a mutagen.



     The  positive  result in the  host-mediated study  utilizing jn_ vivo



 metabolism of  CHC13 could not be evaluated because  the details of the



 procedures used and the appropriate  data were not reported.  This result may



 be  suggestive  of  a positive response and indicates  the need for additional



 testing.  The  results from the  urine spot test are  suggestive of a positive



 response.



 7.4  MUTAGENICITY STUDIES IN EUCARYOTIC TEST SYSTEMS



     Callen et_  a_K (1980) carried out a study on the mutagenicity of CHC13 in



 the  D7 strain  of  Saccharomyces cerevjsi ae, which contains an endogenous



 cytochrome P-450 dependent monooxygenase metabolic activation system.  By



 using this strain of yeast,  Callen and his coworkers eliminated the need for



 the  exogenous type of metabolic activation system used  in the above bacterial



 studies.  Three different  genetic endpoints can be examined with this system:



 gene conversion at the trp5  locus, mitotic crossing over at the adej? locus,



 and  gene reversion at the  i1v1  locus.  The effect of CHC13 on these



endpoints was measured by  exposing cells in suspension to 2.5,  5.0,  and 6.3  g



of CHC13 per liter of buffer (21 mM,  41  mM, and 54 mM,  respectively).   The



purity of the CHC13  sample (from J.T. Baker) was  not provided.   Escape  of



volatilized CHC13 is not expected  to  have  occurred to any significant



extent,  because the  incubations  were  carried out  in  screw-capped  glass  tubes.



Results  of the Callen  et al.  study are presented  in  Table 7-1.   A  1-hr
                                     7-9

-------
                        TABLE 7-1.  GENETIC EFFECTS OF CHLOROFORM ON STRAIN D7 OF S. CEREVISIAE
—I
I
o

Concentration, mM

Survi val
Total colonies
% of control
trp5 locus (gene conversion)
Total convertants
Convertants/105 survivors
ade2 locus (nitotic crossing over)
Total twin spots
Mitotic cross-overs/10^ survivors
Total genetically altered colonies
Total genetically altered colonies/
lf)3 survivors
ilvl locus (gene reversion)
Total revertants
Revertants/10" surviviors
0

1423
100

246
1.7

1
1.6
6
1.0


61
4.3
21

1302
91

274
2.1

1
1.7
11
1.9


46
3.5
41

982
69

450
4.6

2
4.1
43
8.9


81
8.2
54

84
6

278
33.1

4
44.8
47
52.7


50
60.0

-------
treatnent of cells with 54 mM CHC13 resulted in an increased convertant to



survivor ratio with a  marginal  increase in the observed number of convertant



colonies.  Similar results were obtained for mitotic crossing over and gene



reversion.  Toxicity at this concentration was high (6% survival).




    At the lower concentrations of CHC13 (21 mM and 41 mM),  a small



dose-related increase  (1.2-fold and 2.7-fold, respectively)  in gene



convertants was observed.  In addition, a 9-fold increase in the frequency of




genetically altered colonies, which are due to gene conversion and mitotic



crossing over, was observed at 41 mM CHC13.  For gene reversion, a 2-fold



increase was observed.  Toxicity was low at these levels.  These results are




suggestive of a positive response, but additional studies are needed before it



can be conclusively stated that CC14 causes genetic effects  in yeast.




    Sturrock (1977) tested the mutagenic effects of CHC13 at the



8-azaguanine locus in  Chinese hamster lung fibroblast cells  (V-79 cells) in




culture.  The cells were grown to a monolayer and exposed for 24 hr to an



atmosphere containing  1 to 2.5% CHC13.  Cells were then plated onto media



with or without 8-azguanine.  After incubation, the plates were examined for



mutations and survival.  No significant increase in the frequency of mutants



was observed in treated cultures as compared with untreated  controls.



However, the xenobiotic biotransformation capability of the  cells used in this



study is unknown.



     fiocke et_ a_K (1981) evaluated the mutagenicity of CHC13 by carrying out



tests in ProsopMl a to detect sex-linked recessive lethal mutations.  The



flies were exposed by the adult feeding method to 25 mM CHd3.  Three



successive broods (3-3-4 days) of flies were examined for sex-linked recessive




lethal mutations.  Over 40DO chromosomes per brood were tested.  In  two of the



broods, small increases in mutations were observed.  Results (sex-linked
                                     7-11

-------
recessive lethal s/chromosomes) were as follows:   Brood 1, 20/4616 (0.430/,);



Brood ?., 13/4349 (0.29%).  Controls were 0.27% and 0.14%, respectively.   These



increases were not significant.  To obtain significant increases more



chromosomes would have to he tested.




    Because sperm head abnormalities are thought by Hyrobek and Rruce (1978)



to arise from mutations in the genes that code for spermatogenesis,  it is



possible that assays for sperm head abnormalities may be used as an  indication



of the mutagenic potential of chemicals.  In a screening study in which  54



chemicals were tested for induction of sperm head abnormalities, Topham  (1980)



reported that CHC13 did not induce sperm head abnormalities when injected in



mice.  Groups of five male mice received five daily intraperitoneal  injections



of corn oil  alone (5 ml/kg/day) or CHC13 in corn oil  at 0.025, 0.05, 0.075,



0.1, and 0.25 mg/kg/day.  Topam (1980) reported  that  the highest of  these



doses (0.25 mg/kg/day) was lethal.  However, in  Chapter 5 of this document,  it



is stated that the LO^n, of CHC13 for ICR male mice is 789-1590 mg/kq.  The



reason for the discrepancy in reported topic response to CHC13 is not clear.



Five weeks after the last dose, caudal sperm smears were examined for head



abnormalities. The raw data for the experiments  with  CHC13 were not



presented in the paper. Topham stated that CHC13 induced sporadic small



increases in abnormal sperm heads at a high dose, but this result could  not  be



repeated.




    Another sperm head abnormality study was carried  out by Land et  al.



(1981).  This study was designed to determine whether certain anesthetics



affect mouse sperm morphology.  Groups of five male mice (11 weeks old)  were



exposed by inhalation to CHC13 at 0.04 and 0.08% (vol/vol) for 4 hr/day  for



5 days in glass exposure chambers.  Control  mice (N = 15) were exposed to



compressed air under similar conditions.  Twenty-eight days after the first
                                     7-12

-------
exposure, the nine survivors from each exposure level  were killed and the



caudal  spern were examined for head abnormalities.   The results were reported



as % abnormal sperm (+ SEM) and were as follows:   control, 1.42 (n.08);  0.08%



CHC13,  3.48 (0.66); and 0.04% CHC13, 2.76 (0.31).   The authors  concluded



that exposure of mice to CHC13 resulted in a significant increase in sperm



head abnormalities compared to the control  (P < 0.01).  However, significance




was calculated by the t-test and by the F test.  Use of these tests may  not be



appropriate in this case because of the non-homogeneity of the  variance  in



CHCl3-treated and control  groups (Dr. Chao Chen,  Carcinogen Assessment



Group,  U.S. EPA, personal  communication).  This study  is suggestive of a



positive response, but a more appropriate statistical  analysis  of the data is



needed.  However, the data necessary to carry out  such a statistical analysis




were not provided.



    Testing of the mutagenic potential  of CHC13 in  eucaryotic systems was



carried out in the same screening study edited by  de Serres and Ashby (1981)



that was discussed in the previous section of this  chapter for  bacterial



assays.  Seven yeast assays, two jn vitro mammalian ONA damage  assays



(unscheduled DNA synthesis and sister chromatid exchange), and  three



whole-animal tests (Drosophila sex-linked recessive lethal, mouse bone marrow



micronucleus, and mouse sperm abnormality) were described for CHC^.  The



ONA damage studies will be discussed in the next  section of this chapter.



    The seven yeast assays involved both forward  and reverse point nutations,



mitotic crossing over, mitotic gene conversion, and induction of aneuploidy in



mitotic cells.  The latter three assays test for  ONA damage.  A positive




result  was obtained only in the forward mutation  assay utilizing



S c h i20s a c c haromycej potnbe as the test organism.  In the reverse mutation



assay,  CHC13 was tested only with stationary cells  in  the presence of rat S9
                                     7-13

-------
mix.  Exposure was for 24 hr.  Growing cells are none sensitive to the



mutagenic effects of several  chemicals than are stationary cells,  possibly



because log-phase yeast cells contain an endogenous cytochrome P-450



metabolizing system (Callen et_ aj_., 1980).  The ONA damage studies in yeast




yielded negative results.  The negative results in the mitotic qene conversion



assay, which was carried out in strain 07, are in conflict with the weakly



positive results reported for CHC13 by Callen et_ a_L (1980) as described




above.



    The three whole-animal tests on CHC13  (Drosophila sex-linked recessive



lethal, mouse bone marrow micronucl eus, and mouse sperm abnormality) described



in  the de Serres and Ashby (1981) report yielded negative  results.  However,



de  Serres and Ashby, in their overview of  the performance  of the assay systems



used  in this  study, state that the whole-animal tests had  low  sensitivity and



that  a negative  result has "very little significance."  It was recommended




that  CHC13  be tested further in irj \nvo short-term tests.



    In summary,  the results  from the  eucaryotic test systems suggest that



CHC13 may be  a weak mutagen.   Results  indicative of  a positive response were



obtained only in studies  using test organisms possessing endogenous  activation



systems  (i.e., yeast and  mice).  More studies,  particularly with  organisms



possessing  endogenous  activation,  are needed  before  a definite conclusion on



the mutagenicity of CHC13 can  be  reached.   Suggested studies  are  described



later in this chapter.



7.5  OTHER  STUDIES INDICATIVE  OF ONA  DAMAGE



    Two  types of DNA damage  studies,  sister  chromatid exchange (SCE)  and




unscheduled DNA  damage (HDS),  are  described  in  this  section.



    Sister  chromatid exchange  is  thought  to  involve  DNA breakage.   For  this



 reason,  assays  for SCE have  been  used as  an  indicator of  primary  DNA damage.
                                      7-14

-------
 White et_ aK  (1979)  studied the induction of SCEs  by  CHC13 and other
 anesthetics.   Information  about the  purity of these compounds  was  not  given.
 Exponentially  growing  Chinese hamster ovary cells  were  exposed to  the  gases JJT_
 vitro with  and without S9  mix (10% by volume) prepared  from Aroclor-induced
 rat livers.   Exposure  was  at 0.71% (vol/vol)  CHC13 (88  mM)  for 1 hr  in
 closed screw-capped  culture flasks.   The  cells were then  incubated for  24  hr
 in medium containing 10 uM 5-bromo-2-deoxyuridine.  SCEs  per chromosome were
 0.544 _+ 0.018  for  CHCl3-exposed cells  and 0.536 _+  0.018 for  controls.   The
 short exposure time  and low concentration raise serious questions  concerning
 the conduct of this  assay.   Also,  even though the  rat liver  S9 was sufficient
 for activation of  vinyl-containing compounds  to derivatives  (presumably
 epoxides) that induced  SCE,  it  may not have been adequate for activation of
 CHC13 and the  other  haloalkanes  that  tested negative.
     Another SCE study was  carried out  by  Kirkland et_ aj_.  (1981) using human
 lymphocytes.   The  cells were  treated with  CHC13 at 25,  50, 75, 100, 200, and
 400 ug/ml (0.2, 0.4, 0.6,  0.8,  1.6, and 3.3 mM, respectively) for 2 hr in the
 presence of S9 mix from Aroclor-induced rats.  Metaphase spreads from
 approximately  100 cells per treatment were examined.  Acetone was the negative
 control, and a positive control was not included in the assay because the same
 donor's lymphocytes had previously shown a dose-related increase in SCE after
 treatment with benzo[a]pyrene in the presence or absence of S9 mix.  A small
 increase in  SCE occurred at 50 ug of CHC13 per ml,  but no  dose-response
 trend was observed.
    There are several problems with this study.   First,  since a positive
 dose-related increase in SCE after treatment of lymphocytes with
benzo[a]pyrene  was  not  dependent on S9 mix, this  positive  control  is
 inadequate for  substances that are likely  to require metabolic  activation.   In
                                     7-15

-------
addition, this control  was not concurrent and is therefore not  an  appropriate



positive control.   Second, there is no indication that  volatilization  and



escape of CHC13 was prevented.  Third, the maximal  dose was only 3.3 mM.



Fourth, infornation about the toxicity of CHC13 for the lymphocytes  was  not



provided.




    Two jrj vitj-Q mammalian DNA damage studies on CHC13  (tins and SCE) were



described in the volume edited by de Serres and Ashby (1981).   The SCE assay



(Chapter 51) utilized an exogenous activation system and yielded negative




results.  The Chinese hanster ovary cells were exposed  to CHOI 3 in the



presence of rat liver S9 mix for only 1 hr.  This length of time may be



insufficient, particularly since a positive response for 2-acetylaninofluorene



was obtained in the presence of S9 after 2 hr of exposure and  not  after  a 1-hr




exposure.  Also, as mentioned above, exogenous activation may  be inappropriate



when testing CHC13 for genotoxic activity.  Thus, the negative  results




obtained in this study are inconclusive.  In addition,  as in the bacterial



tests described above, there was no indication that precautions were taken  to



prevent evaporation and loss of CHC13 from the culture  flasks.



    Unscheduled DNA synthesis (UDS), measured as repair of chemically  induced



DNA damage, is an additional method of testing for genetic damage.  Mirsalis



et_ aj_. (198?) measured IJDS in primary rat hepatocyte cultures  following  rn_



vivo treatment of adult male Fischer-344 rats (175-275  g) with  CHC13 at  40



and 400 mg/kg by gavage.  Control rats received corn oil (the  vehicle  for



CHC^) by gavage.   Several additional chemicals were also tested in  this



study.  At 2 or 12 hr after treatment, the livers were  perfused in sjtu  and




hepatocytes were isolated.  Approximate"!/ 6 x 105 viable cells  were  seeded



in 35-mm culture dishes containing coverslips and allowed to attach  to the




coverslips for about 90 min.  After the coverslip cultures were washed,  they
                                     7-16

-------
 were  incubated  in  a  medium  containing  in  uCi  [^H]thymidine  (40-50 Ci/mmol)
 per ml  for  4  hr.   The  cultures  were  washed  again  and  incubated  in medium
 containing  0.25 mM cold  thymidine  for  14-16 hr.   The  extent  of  LIDS was
 assessed  by autoradiography.  Net  grains/nucleus  were  calculated  as the silver
 grains  over the nucleus  minus the  highest grain count  of  three  adjacent
 nuclear-sized areas  over the cytoplasm.
    Cells from  negative  control  animals  (given vehicle  only)  ranged from  -3.0
 to -5.1 net grains/nucleus.  Several chemicals tested  positive  in this assay
 (<_ 5  net  grains/nucleus  was considered positive),  including  methyl
 methanesulfonate,  dimethylnitrosamine, 2-acetylaminofluorene, benzidine,  and
 others.   CHC13  at  40 and 400 mg/kg yielded  a  negative  response  (-2.7 to -4.4
 net grains/nucleus).   However,  rats  are not susceptible to CHCl3-induced
 hepatocarcinogenesis (NCI bioassay,  1976).  The negative  response observed in
 this  study is consistent with this fact.  Renzo[a]pyrene  and
 7,12-dimethylbenzCaJanthracene, carcinogens that, like CHC13, are  not rat
 liver carcinogens, also tested negative in this assay.  These chemicals tested
 positive in the jn_ vitro rat hepatocyte (IDS assay (Williams e>t_ a]_., 1981).
 This discrepancy suggests that the jrj^ vitro test  may be more sensitive than
 the in vivo assay.  CHC13 has not been tested in  the jn vitro rat  hepatocyte
 Uns assay.  The mouse is susceptible to CHCl3-induced liver tumors (NCI
 bioassay, 1976)  and therefore may be a more appropriate test animal for the j_n_
 vjvQ DOS assay.
    Also, it is  uncertain whether the method of assaying for UHS used  in  this
 study  (subtraction of cytoplasmic grain counts from nuclear  grain counts)  will
 allow  for detection of  a weak  response.  In  a  recent article discussing the
validity of  the  autoradiographic procedure for detecting UPS in  rat
hepatocytes,  Lonati-Gal ligani  et_ aU  (1983)  describe some  potential problems
                                     7-17

-------
with this method.  First, they found that it is difficult to obtain hepatocyte



preparations of reproducible quality.  Preparations can differ in their



metabolic capabilities.  In order to avoid false negative results with



potential weak UDS inducers due to poor hepatocyte preparations,  they suggest



that test chemicals should be studied in conjunction with a potent



UnS-inducing analog and that negative results should be accepted  only in tests



in which the analog is strongly positive.  No known positive analog of CHC13




was tested in the study of Mirsalis et_ aj_. (1982).  Second, the cytoplasmic



layer covering the nucleus is thinner than the cytoplasmic area next to the



nucleus.  Therefore, a variable overcorrection is probably applied, as



witnessed by the usually higher cytoplasmic than nuclear counts observed in



control  cells (e.g., see above results for cells from control  animals).  This



effect would tend to obscure a weakly positive UDS response.  Lonati-Galligani



et_ £l_. (1983) suggest that an alternative endpoint be determined.  Instead of



subtracting cytoplasmic grain counts from nuclear grain counts, the grains



over the nucleus and over a cytoplasmic area should be scored  and




dose-response curves plotted separately.  Both dose-response curves should be



considered before a decision is reached on whether exposure to a  certain



chemical results in DOS.



    Results of an additional DNA repair study were published by Reitz et al.



(1980).   Mice were exposed orally to CHC13 at 240 mg/kg and repair in the



livers was assayed.  Negative results were obtained.  However, these results



cannot be interpreted because no information on the methodology used to assay



for HNA repair was given.  In order to interpret these results one would have



to know the length of time the mice were exposed to CHC13.  Some  compounds



require a longer exposure than others (e.g., 2-acetylaminofluorine in Mirsalis



et_ aj_.,  1982).  Also, the opportunity for false negative results, as discussed
                                     7-18

-------
above, exists in this study as well.
    The uns assay discussed in Chapter 48 of the de Serres and Ashby (1981)
volume was carried out with Hela cells, which do not contain a P-450
activation system.  An exogenous rat  liver S9 activation system was employed.
Although CHC13 tested positive in this assay in the absence of the
activation system, the discussion of  this assay in the de Serres and Ashby
book suggests that this result is misleading because of an inadequacy in the
statistical method employed.  In the  presence of rat liver S9, CHC13 tested
negative.  However, as already discussed, exogenous activation is probably
inappropriate when testing compounds  such as CHC13 for their ability to
cause genetic damage.
    In summary, because of various deficiencies in the above studies, the
determination of the DNA-damaging potential  of CHC13 requires additional
studies.  Indication of the DNA-damaging potential of CHC13 was suggested by
the small increases in conversion and mitotic crossing over observed in yeast
(Callen et_ a_K, 1980), as described in the previous section.
7.6  CHROMOSOME STUDIES
    Kirkland et_ aj_. (1981) studied the ability of CHC13 to induce
chromosome breakage in cultured human lymphocytes.  The cells from one donor
were treated with CHC13 at 50, 100, 200, and 400 ug/ml for 2 hr in the
presence of an S9 activation system derived from Aroclor 1254-induced rats.
The positive control compound, benzo[a_]pyrene, in a separate experiment with
the same donor's lymphocytes induced  chromosome breakage with or without S9
treatment.  The response of this donor's lymphocyte chromosomes to CHC13 was
a random variation around the control value.  The highest breakage level was
at 200 ug/ml with 8 breaks/100 cells  compared with 5.5 breaks/100 cells in the
control.  However, this difference was not significant according to the
                                     7-19

-------
chi-square test. The same four problems  discussed in  the  previous  section  for
the SCE study carried out by Kirkland et_ aK  (1981) apply to  their study on
chromosome breakage.
    According to Schmid (1976), the bone marrow micronucleus  test  can  be used
to detect clastogens and spindle poisons.  Micronuclei  are small  elements  that
contain either pieces of chromosomal  fragments originating from clastogenic
events or whole chromosomes resulting from malfunction  of the spindle
aparatus.  Gocke et_ aj_. (1981) used this test to study  chromosome  aberrations
in mice exposed to CHC13.  The animals were treated with  CHC13 at  0 and 24
hr, and bone-marrow smears were prepared at 30 hr. The purity of  the  CMC!3
(purchased from Merck) was not provided.  Four mice  (two  males and two
females) were used for each of three doses and one control.  The animals were
given two intraperitoneal injections of CHC13, each  at  238, 476, and 952
mg/kg (2, 4, and 8 mmol/kg, respectively).  The authors state that the assay
was performed according to Schmid (1976);  thus, it  can be assumed that the
doses chosen included the highest tolerable dose.  Slides were coded,  and  1000
polychromatic erythrocytes were scored per mouse.
    The results were as follows (dose, % micronucleated polychromatic
erythrocytes):  0 mg/kg, 1.2%; 2 x 238 mg/kg, 2.2%,  2 x 476 mg/kg, 2.6%;  2 x
952 mg/kg, 2.2%.  Thus, a dose-related increase was  not observed.   Three
halogenated alkanes were tested (dichloromethane, 1,1,1-trichloroethane,  and
CHC13) and all yielded negative results.  Of 30 chemicals tested, only two
(pyrogallol and hydroquinone) yielded positive results  in the micronucleus
test.  Positive controls were not included in the assay,  but the positive
results for pyrogallol and hydroquinone  indicate that the assay system was
working.
                                     7-20

-------
     The  micronucleus  test  was  also  used  by  Agustin  and  Lim-Sylianco  (1978)  to



 study  the  clastogenic potential  of  CHC13.   The  authors  tested  seven



 concentrations  of  CHC13  up to  900 mg/kg  in  the  mouse.   The  number  of  mice



 used and their  sex was not specified.  The  CHC13 was purchased  from



 Mallinckrodt  and was  redistilled before  use.   Information on purity of  the



 CHC13  was  not provided.  For each slide, 1000 polychromatic erythrocytes



 were scored.  The  authors  reported  that  CHC13 was clastogenic.   Results were



 as  follows  (dose in mg/kg,  number of micronucleated polychromatic  erythrocytes



 per 1000 polychromatic erythrocytes _+ SE):  0, 4 +_ 1; 100, 3 +_ 1; 200, 5 _+ 1;



 400, 5 4. 1; 600, 9 _+  2;  700, 17 _+ 4; 800, 9 ± 2; 900, 10 _+ 2.   The authors



 stated that these  data indicate  that CHC13 must be metabolized  to  a



 clastogenic substance, because a straight-line dose-response relationship was



 not observed.   However,  the data provided by the authors are not sufficient to



 support  this interpretation.




    In the same paper, Agustin and Lim-Sylianco demonstrated that vitamin E



 administered 1  hr after  CHd3 reduced the number of micronucleated cells



 observed at 700 mg of CHCl3/kg ( 17 +_ 4) to the control  level  (4 _+ 1).  The



 significance of this  result is not clear.



    This study  by Agustin and Lim-Sylianco (1976)  is difficult to interpret



 because  details  of the experimental  procedures  necessary to  permit  an



 evaluation of the results were not provided (e.g.,  number and  sex of the



animals and positive and  negative controls).  This  study suggests that CHC13



may affect chromosomes.  However, corroborative studies  are  needed  to  confirm



or refute this suggested  response.



    In  summary,  based  on  the results of  the three  chromosome aberration



studies described above,  there  are  suggestive  but not  conclusive data  that



CHC13 is  clastogenic.   Negative results  were reported  by Kirkland et al.
                                     7-21

-------
(1981) and by Gocke et_ aj_. (1981), while Agustin and Lim-Sylianco (1976)
reported a positive result.   More studies are needed before it can be
conclusively determined whether or not CHC13 is clastogenic.
7.7  SUGGESTED ADDITIONAL TESTING
    More studies are needed  on the covalent binding of ^CHCl3 to DNA.
Such studies should be done  with 14CHCl3 at a higher specific activity
[about 40 Ci/mol (Brookes and Lawley, 1971)] than used by Diaz Gomez and
Castro (1980b).
    Additional studies on the ability of CHC13 to cause DNA damage are
needed.  Examples of such tests include measurement of unscheduled DNA
synthesis jrj vi^/o in mice and in both rat and mouse hepatocytes after in vitro
exposure (Williams, 1981), measuring endpoints suggested by Lonati-Gal1igani
et_a_l_. (1983).
    Additional studies are needed to corroborate and extend or refute the
Cal len et_ aj_.  (1980) study in yeast, which possesses an endogenous activation
system and is capable of assaying for point mutations, mitotic crossing over,
and gene conversion.
    Further testing for the ability of CHC13 to cause chromosome aberrations
is needed, particularly in jn vivo systems.
    In addition to the study by Gocke et^ aj_. (1981), further testing of the
ability of CHC13 to cause sex-linked recessive lethal mutations in
Drosophjla is needed.  However, in order to detect a weak  response, a larger
number of chromosomes than analyzed by Gocke et_ aJN  (1981)  should be scored.
7.8   SUMMARY AND CONCLUSIONS
    It has been demonstrated that chloroform (CHC13) can be metabolized j_n_
vivo  and in vitro to a substance(s)  (presumably phosgene)  that interacts with
protein and lipid.  However, the  only experiment measuring interaction  of
                                     7-22

-------
metaholically activated CHC13 with DNA, in which adequate information was



given concerning the experimental procedures used, yielded a negative result



(Diaz Gomez and Castro, 1980h).  This result was judged as inconclusive



because the specific activity of the 14CHCl3 may have been too low.



    The majority of the assays for mutagenicity and genotoxicity have also



yielded negative results.  However, many of these results are inconclusive



because of various inadequacies in the experimental protocols used.  The major



problem is with those bacterial, sister chromatid exchange, and chromosome



aberration studies that used reconstituted exogenous activation systems (i.e.,



S9 mix).  In none of these studies was it shown that CHC13 was activated or



metabolized by the activation system used.  Metabolism of 2-aminoanthracene or



vinyl compounds (used as positive controls) is probably an inadequate



indication that the activation system can metabolize CHC13, because these



substances are not halogenated alkanes and are therefore not metabolized like



them.  A better indication that the activation system is sufficient for



metabolism of CHC13 may be to show that it metabolizes l^CHC^ to



intermediates that bind to macromolecules.  A second problem with experimental



protocols utilizing exogenous activation systems relates to the possibility



that any reactive metabolic intermediates formed may react with microsomal or



membrane lipid or protein before reaching the DNA of the test organism.  A



third potential problem occurs in those jn yjtro protocols in which



precautions were not taken to prevent escape of volatilized CHC13.



    Studies in which endogenous or in yi^p activation systems were used



include those reported by Callen et_ aj_. (1980) in yeast, Gocke et_ aj_. (1981)



in Drosophija (sex-linked recessive lethal test) and mice (bone-marrow



micronucleus test), Topham (1980) and Land et_ a]_. (1981) in mice (sperm head



abnormalities), and Agustin and Lim-Sylianco (1978) in mice (bone marrow
                                     7-23

-------
micronucleus test and host-mediated assay).  The results from several  of these



studies suggest that CHC13 may be a weak mutagen.




    Tn summary, with the present data, no definitive conclusions can be



reached concerning the mutagenicity of CHC13.  However, there is some



indication (from the binding studies and from the mutagenicity tests that



utilized endogenous or ui  vivo metabolism) that CHC13 may have the potential



to be a weak mutagen.  In  order to substantiate this, only certain



wel1-designed in vivo mutagenicity studies or studies with organisms



possessing endogenous eucaryotic P-450 activation systems are recommended.
                                     7-24

-------
 7.9  REFERENCES

 Agustin, J.S., and C.Y. Lim-Sylianco.  1978.  Mutagenic and clastogenic
    effects of chloroform.  Bull. Phil. Biochem. Soc. 1:17-23.

 Brookes, P., and P.O. Lawley.  1971.  Effects on DNA:  Chemical methods.
    In: A. Hollaender (ed.) Chemical Mutagens.  Vol 1, pp. 121-144.  New
    York:  Plenum Press.

 Callen, n.F. C.R. Wolf, and R.M. Phil pot.  1980.  Cytochrome P-450 mediated
    genetic activity and cytotoxicity of seven halogenated aliphatic
    hydrocarbons in Saccharomyces cereyjsjae.  Mutat. Res. 77:55-63.

 de Serres, F.J., and J. Ashby (eds.)  1981.  Evaluation of Short-Term Tests
    for Carcinogens.  Progress in Mutation Research, Vol I.  Elsevier/North
    Hoi land.

 Diaz Gomez, M.I., and J.A. Castro.  1980a.  Nuclear activation of carbon
    tetrachloride and chloroform.  Res. Commun. Chem. Pathol. Pharmacol.
    27:191-194.

 Diaz Gomez, M.I. and J.A. Castro.  1980b.  Covalent binding of chloroform
    metabolites to nuclear proteins-no evidence for binding to nucleic acids.
    Cancer Lett. 9:213-218.

 Gocke, E., M.-T. King, K. Eckhardt, and n. Wild.  1981.   Mutagenicity of
    cosmetics ingredients licensed by the European communities.  Mutat.  Res.
    90:91-109.

 Kirkland, O.J., K.L. Smith, and N.J. Van Abbe.   1981.  Failure of chloroform
    to induce chromosome damage or sister-chromatid exchanges in cultured
    human lymphocytes and failure to induce reversion in Esc her ichi a cojj .
    Fd. Cosnet. Toxicol. 19:651-656.

 Kirk-Othmer Encyclopedia of Chemical Technology, Second  Edition, Supplement
    volume.  1971.   pp.  674-683.   Interscience  Publishers.

 Land,  P.C., E.L. Owen, and H.W.  Linde.   1981.   Morphologic changes in mouse
    spermatozoa after exposure to inhalational  anesthetics  during early
    spermatogenesis.  Anesthesiol.  54:53-56.

 Lonati-Galligani, M., P.H.M Lohman,  and F. Berends.   1983.   The validity of
    the autoradiographic method  for detecting DNA repair synthesis in rat
    hepatocytes in  primary culture.   Mutat. Res. 113:145-160.

Mirsalis, J.C., C.K. Tyson, and  B.E. Butterworth.   1982.   Detection  of
    genotoxic carcinogens  in  the  jn  yiy°, ~ j.n. vitro  hepatocyte  DNA repair
    assay.   Environ. Mutagen.  4:553-562.

National  Cancer Institute (NCI).   1976.   Report  on  Carcinogenesis Bioassays of
    Chloroform.  National  Technical  Information  Service,  Springfield, Virginia
    (NTIS PB-264-018).
                                     7-25

-------
Schmid, W.  1976.  The micronucleus test for cytogenetic analysis.  In:  A.
    Hollaender (ed.).  Chemical Mutagens.  Vol. 4, pp. 31-53.   New York:
    Plenum Press.

Simmon, V.F., K. Kauhanen, and R.G. Tardiff.  1977.  Mutagenic activity  of
    chemicals identified in drinking water.  In:   0. Scott, B.A. Bridges,  and
    F.H. Sobels (eds.), Progress in Genetic Toxicology, pp. 249-258.  New
    York:  Elsevier/North Holland Biornedical Press.

Sturrock, J.  1977.  Lack of mutagenic effect of halothane or chloroform on
    cultured cells using the azaguanine test system.  Br. J. Anaesth.
    49:207-210.

Topham, J.C.  1980.  Do induced sperm-head abnormalities in mice specifically
    identify mammalian mutagens rather than carcinogens?  Mutat. Res.
    74:379-387.

Uehleke, H. T. Werner, H. Greim, and M. Kramer. 1977.  Metabolic activation of
    haloalkanes and tests in vjtro for mutagenicity.  Xenobiotica. 7:393-400.

Weisburger, J.H. and G.M. Williams. 1982.  Metabolism of chemical  carcinogens.
    In: F.F. Becker (ed.).  Cancer, a Comprehensive Treatise, 2nd Edition.
    vol. 1, pp. 241-333.  New York:  Plenum Press.

Reitz, R.H., J.F. Quast, W.T. Stott, P.G, Watanabe, and P.J. Gehring.   1980.
    Pharmacokinetics and macromolecular effects of chloroform in rats  and
    mice: Implications for carcinogenic risk estimation.  In: R.L. Jolley,
    W.A.  Brungs, and R.B. Cumming (eds.).  Water chlorination:   Environmental
    Impact and Health Effects, Vol. 3, pp. 983-992.

White, A.E., S. Takehisa, E.I. Eger, S. Wolff, and W.C. Stevens.  1979.
    Sister chromatid exchanges induced by inhaled anesthetics.  Anesthesiol.
    50:426-430.

Williams, G.M. 1981.  Liver culture indicators for the detection of chemical
    carcinogens. In: Short-Term Tests for Chemical Carcinogens,  H.F. Stich and
    R.H.C. San (eds.), pp. 275-289.  New York:  Springer Verlag.

Wyrobek, A. and W.R. Bruce.  1978.  The induction of sperm-shape abnormalities
    in mice and humans.  In:  A. Hollaender (ed.).  Chemical Mutagens.
    Vol. 5, pp. 257-285.  New York:  Plenum Press.
                                     7-26

-------
                             8.  CARCINOGENICITY








8.1.  ANIMAL STUDIES







     The carcinogenicity of chloroform has been evaluated in mice, rats, and



dogs.  Evidence for carcinogenic activity by chloroform includes induction of



renal epithelial tumors in male Osborne-Mendel  rats (National Cancer Institute



[NCI] 1976), hepatocellular carcinomas in male and female B6C3F1 mice (NCI



1976), kidney tumors in male ICI mice (Roe et al.  1979), and hepatomas in



female strain A mice (Eschenbrenner and Miller 1945) and NIC mice (Rudali



1967).  Cape! et al. (1979) demonstrated an ability of chloroform to promote



growth and metastasis of murine tumors.   Chloroform was not shown to be



carcinogenic in (C57 x DBA2 Fl) mice (Roe et al.  1968), female Osborne-Mendel



rats (NCI 1976), female ICI mice, and male mice of the CBA, C57BL, and CF/1



strains (Roe et al. 1979), male and female Sprague-Dawley rats (Palmer et al.



1979), and male and female beagle dogs (Heywood et al. 1979).  Chloroform was



negative in a pulmonary tumor induction bioassay  in male strain A/St mice



(Theiss et al.  1977).   Chloroform in liquid solution did not induce trans-



formation of baby Syrian hamster kidney (BHK-21/C1 13) cells in vitro.   Under



the conditions  of the carcinogenicity bioassays showing carcinogenic activity



for chloroform  specifically in kidney and liver of mice and rats, the conclu-



sion can be made, by applying the IARC classification  approach for carcinogens,



that there is sufficient evidence for the carcinogenicity of chloroform in



experimental animals.
                                      8-1

-------
8.1.1.  Ural Administration (Gavage):  Rat








     8.1.1.1.   NATIONAL CANCER INSTITUTE (1976) --  A carcinogenesis bioassay



on chloroform in Osborne-Mendel rats was reported by the NCI (1976).  The




chloroform product (Aldrich Chemical Company, Milwaukee, Wisconsin) was shown



to be 98% pure chloroform and 2% ethyl alcohol (stabilizer)  by gas-liquid



chromatography, flame ionization detection, and infrared spectrometry at the



carcinogenesis bioassay laboratory.   Chloroform solutions in corn oil were



prepared fresh each week and stored under refrigeration.



     Fifty animals of each sex were assigned to each of two  dose groups.



Treated animals were compared with matched vehicle-control  groups (20 males and



20 females) and with vehicle colony control groups (99 males and 98 females) that



included the matched control group and three other controls  groups put on study



within 3 months of the matched control group.  Matched control  and treated



animals were housed in the same room, and colony controls were housed in two



different rooms.



     Doses selected for the main study in rats were estimated as those maximally



and one-half maximally tolerated based on survival, body weights, clinical signs,



and necropsy examinations in a preliminary toxicity test in  which chloroform was



given by gavage for fa weeks with a subsequent observation period of 2 weeks



without treatment.  The chronic study began when rats were 52 days old and



ended with sacrifice of survivors at 111 weeks.  Chloroform  was administered in



corn oil by gavage 5 days each week during the initial 78 weeks.  Doses of 90



and 180 mgAg/day were administered to male rats throughout  the chronic study;



however, since initial doses of 125 and 250 mg/kg/day were reduced to 90 and 180



mgAy/day at 22 weeks, doses given to female rats were expressed as time-weighted



averages of 100 and 200 mg/kg/day.





                                     8-2

-------
     Decedents and survivors were necropsied, and tissues and organs were ex-



amined microscopically.   Body weights and food consumption were monitored weekly



for the first 10 weeks and monthly thereafter.  Animals were observed twice daily.




     In matched control  and both dose groups, at least 50% of the male arid



female rats survived as  long as 85 and 75 weeks, respectively.   Seven matched



control, 24 low-dose, and 14 high-dose males and 15 matched control, 22 low-dose,



and 14 high-dose females survived until  the end of the study.  Only one control



male rat died before 90  weeks; the increase in death rate of control males after



90 weeks was, according  to the NCI (1976) report, "probably due to respiratory



and renal  conditions."  Overall survival  was less in treated animals than in



controls (Figure 8-1).



     Decreased body weight gain was evident in both sexes of rats in both



treatment groups.   Initial mean body weights for all groups were about 175 g



for females and 250 g for males.  By 50 weeks, mean body weights were approxi-



mately 400 g in control, 350 g in low-dose, and 330 g in high-dose females;



by 100 weeks, mean body  weights were about 375 g in all groups  of females.  In



males, mean body weights were about 640 g in the control  group, 550 g in the



low-dose group, and 500  g in the high-dose group by 50 weeks; by 100 weeks,



mean body weights  were approximately 500 g in all groups.   Food consumption



was reported as slightly lower in treated animals,  but data were not provided.



Appearance and behavior  among groups were generally similar, but hunching,



urine stains on the lower abdomen, redness of eyelids, and wheezing were noted



in treated animals early in the study.



     A statistically significant (P < 0.05) increase in renal epithelial



tumors of tubular  cell origin was found  in treated  male rats (Table 8-1).   The



epithelial  tumors  were described as follows:  Of 13 tumors in high-dose males,




10 were carcinomas and three were adenomas; two carcinomas and  two adenomas





                                     8-3

-------
                                    PROBABILITY OF SURVIVAL
  CO
   I
t— ' <
UD O)
  CO
  (/I
O_

CO


oo

-pi


CU
(-1-
c/>


Z3

CU

O
OJ
  rs
  o
  (D
  13
  O


  (-1-
  O
  OJ
  l/l
  O
  rs

  n
  o
  -5
  O
  -h
  O
 m
 O
         CD
         CD
         7T
         C1
m
O
C
O
CD
CD
                p  p
                b  ->•
                o  o
                 p
                 ro
                 o
                               p
                               GO
                               o
                                        o
p
en
o
p
en
o
p
bo
o
p
CD
o
o
o
             o
             fo
             o
                    I
                                        >  O D
         C/5  O
C
D
             O)
             o
    00
    o
             o
             o
             o
                       —
                       CD
              0)

              CO
                                    PROBABILITY OF SURVIVAL

-------
                           TABLE 8-1.   EFFECT OF  CHLOROFORM  ON  KIDNEY EPITHELIAL TUMOR INCIDENCE
                                                  IN OSBORNE-MENDEL RATS
                                                        (NCI  1976)
                                      Male
                                                                      Female
Treatment3
     Controls'3

Colony      Matched
             Dose  (mgAg/day)c

             90            180
                                 Controlsb

                            Colony      Matched
                                     Dose  (mgAg/day)c

                                      100          200

CO
1
cr

Kidney tumor 0/99(0%)
incidence0'
P value"
Time to first
tumor (weeks) —
0/19(0%) 4/50(8%)e
0.266

102
12/50(24%)f
0.0141

80
0/98(0%) 0/20(0%) 0/49(0%) 2/48(4%)9
0.495

102
Survival  at
 terminal
 sacrifice
 (111 weeks)
   26%
37%
48%
28%
51%
75%
45%
29%
aChloroform in corn oil administered by  gavage  5 times  per week  for 78 weeks.
^Colony controls consist of four vehicle-control  groups,  including matched controls,  given corn oil.
cDoses are time-weighted averages.
dAnimals with tumor/animals examined.
eTwo with tubular cell adenocarcinoma  and two with  tubular cell  adenoma.
^Ten with tubular cell adenocarcinoma  and two with  tubular cell  adenoma.
90ne with tubular cell adenocarcinoma  and one with  squamous cell  carcinoma in  the renal  pelvis.
^Fisher's Exact Test, compared with matched controls.
iFor adenocarcinomas alone, P value is 0.03.

-------
 comprised the tumors found in four low-dose males; one renal epithelial car-
 cinoma and  one squamous cell carcinoma from renal pelvic transitional epithelium
 were noted  in two high-dose females.  One low-dose male had both a malignant
 mixed tumor and a tubular cell adenoma in the left kidney, and a high-dose
 male had a tubular cell carcinoma and a tubular cell  adenoma in the right
 kidney.  Renal epithelial carcinomas were large and poorly circumscribed, and
 they infiltrated surrounding normal tissue.   Renal epithelial  adenomas were
 circumscribed and well-differentiated.  Additional kidney tumors included
 malignant mixed tumors in two low-dose and two colony control  males and
 hamartomas in one low-dose male, one high-dose male,  and one colony control
 male.
     Although a statistically significant (P < 0.05)  increase  in thyroid tumors
 in both treatment groups of female rats as compared with colony controls was
 reported, the toxicologic significance of this finding appears questionable in
 that C-cell  tumors and follicular cell tumors, which  have different embryonic
 origins and different physiologic functions, are combined in the incidences
 described in Table 8-2; the majority of tumors were adenomas;  the spontaneous
 incidence of thyroid tumors in Oshorne-Mendel  females is variable as stated,
without presentation of historical  data,  in  the NCI (1976) bioassay report;
and the increased incidence of thyroid tumors  in treated females is not
significant  (P > 0.05)  when compared with data for matched controls.
     No significant  (P < 0.05) differences for other  tumor types among groups
were apparent.   Four rats were lost (missing or autolyzed) for pathology.
     Non-neoplastic  lesions described as  treatment-related include necrosis of
liver parenchyma, epithelial  hyperplasia  in  the urinary bladder, and
hernatopoiesis  in  spleen.   Inflammatory pulmonary lesions characteristic of
                                     8-6

-------
                             TABLE 8-2.   EFFECT  OF  CHLOROFORM  ON  THYROID TUMOR INCIDENCE
                                            IN  FEMALE  OS60RNE-MENDEL RATS
                                                     (NCI  1976)
Dosea>b
(mg/kg/day)
0 (matched)"
o° 0 (colony)h
100
200
Fol licular cell
tumors0
Incidence^
1/19(5%)
1/98(1%)
2/49(4%)
6/49(12%)
C-cell
tumorsd
Incidence
0/19(0%)
0/98(0%)
6/49(12%)
4/49(8%)

Incidence
1/19(5%)
1/98(1%)
8/49(16%)
10/49(20%)
Total tumors6
Time to first tumors
P value9 (weeks)
110
110
0.216 73
0. 121 49
aChloroform in corn oil administered by gavage 5 times per week.
^Time-weighted average doses.
cAdenomas except for carcinoma in one low-dose and two high-dose animals.
^Adenomas except for carcinoma in one high-dose animal.
eSee text.
fAnimals with tumors/animals examined.
SFisher's Exact Test, compared with matched controls.
nColony controls consist of four vehicle-control groups, including matched controls, given corn oil.

-------
pneumonia were found in all groups, but the severity and incidence of these



lesions were stated (data not reported) to have been greater in treatment groups.



     Under the conditions of this bioassay, chloroform treatment significantly



(P < 0.05) increased the incidence of renal epithelial tumors in male Osborne-



Mendel  rats.   Although the number of matched vehicle controls was low, the



use of pooled colony controls gives additional  support for treatment-related



effects.   Moreover, historical  control  incidence of renal  epithelial  tumors in



Osborne-Mendel  rats was reported as rare.



     Lower survival rates and body weights in rats than in matched controls



provide evidence that the chloroform doses used were toxic to the rats used in



this study, and a more precise  estimate of dose-response perhaps could have



been obtained if additional lower doses had been given, and if constant doses



rather than time-weighted averages had  been used.   Treated animals were housed



in the same room as rats treated with other volatile compounds (1,1,2,2-tetra-



chloroethane, 3-chloropropene,  ethylene dibromide, carbon  tetrachloride);



however,  since controls were in the same room as treated animals and  oral



chloroform doses probably would have been  much  higher than ambient levels of



other volatiles, the likelihood that the other  volatile compounds were responsible



for the observed results is considered  to  be low.   Additionally, these other



volatile compounds did not induce kidney tumors in Osborne-Mendel rats (NCI



1976; Weisburger 1977).  It should be noted that ambient levels of volatiles in



the animal quarters were not measured.








     8.1.1.2.  PALMER ET AL. (1979) --   Palmer  et  al. (1979) reported carcin-



ogenicity studies on chloroform in Sprague-Dawley  rats.  Chloroform was prepared



in toothpaste,  as described in  Table 8-3 herein for the Roe et al. (1979) study,



and administered by gavage.  Dose levels of 15, 75, and 165 mg CHCl3A9/day were

-------
selected for the carcinogenicity study based on results of a preliminary range-

finding study showing the lowest toxic dose, indicated by liver and kidney

changes, as 150 mg/kg/day.
        TABLE 8-3.  TOOTHPASTE FORMULATION FOR CHLOROFORM ADMINISTRATION
                               (Roe et al. 1979)
     Ingredient                                        Percentage w/w


     Chloroform                                             3.51

     Peppermint oila                                        0.25

     Eucalyptoia                                            0.50

     Glycerol                                              39.35

     Carragheen gum                                         0.45

     Precipitated calcium carbonate                        48.53

     Sodium lauryl sulphate                                 1.16

     Sodium saccharin                                       0.03

     White mineral oil                                       1.10

     Water                                                  5.12

            Total                                          100.00

    aEssential  oil flavor components.




     An initial carcinogenicity study  was done in which 25 rats of each sex per

group received one of the selected doses in toothpaste containing essential

oils (flavor components), indicated in Table 8-3, 6 days  per week.   A concurrent

control group of 75 males and 75 females was administered toothpaste without

chloroform and essential  oils.   A second carcinogenicity  study was done in which
                                      8-9

-------
50 male and 50 female specific  pathoyen-free  (SPF)  Sprague-Dawley  rats were



dosed with 60 mg CHCl3/kg/day  in toothpaste with  essential  oils b  days per week,



and 50 control rats of each  sex were given toothpaste without  chloroform but



with essential oils.



     Body weights were measured weekly,  and food  consumption was  recorded.



Body weights were initially  180 to 240 g for  males  and  130  to  175  g  for females.



Blood and urine analysis were  performed  in the  first study, and serum and



erythrocyte cholinesterase activities were monitored in  the second study.



All animals were necropsied, and tissues and  organs were examined  histopatho-



logically.  Adrenals, kidneys,  liver, lungs,  and  spleen  were weighed.



     Chloroform was not carcinogenic in  these studies.   Significant  (P < O.Ob)



body weight loss in high-dose  males in the first  study  (data not  reported) and



maximal body weight gain of  approximately 370 g in  control  males,  330 g in



treated males, 220 g in control females, and  180  g  in  treated  females in the



second study suggest an effect  from chloroform  treatment,   Other  than a 40%



reduction of plasma cholinesterase levels and slight decreases in  serum



glutamic-pyruvic transaminase  and serum alkaline  phosphatase in treated females,



additional toxic effects from  chloroform treatment  were  not evident.



     Low survival, attributed  to respiratory  disease,  was  apparent in both



studies.  The initial study  was terminated at 52  weeks;  50% of the animals



in all groups had died by 52 weeks in the second  study,  which  was  ended at 95



weeks.  Except for 48 control  females in the  initial study, no more  than  18



animals were alive in each group at the  conclusion  of  either study.  Although



carcinogenic activity for chloroform was not  observed,  these studies on



Sprague-Dawley rats are weakened by the  high  early  mortality in control and



treated animals.
                                    8-10

-------
8.1.2.  Oral  Administration (Gavage):   Mouse

     8.1.2.1.  NATIONAL CANCER INSTITUTE (1976) --  The carcinogenicity of
chloroform in B6C3F1 mice was evaluated by the NCI (1976).   The chloroform
product and chloroform solutions in corn oil were those used in the NCI (1976)
carci nogenesis bioassay in rats discussed herein.
     Each of two dose groups was composed of 50 males  and 50 females.   Treated
mice were compared with matched vehicle control groups (20 males and 20 females)
put on study 1 week earlier and with vehicle colony control  groups (77 males and
80 females), which included the matched control group  and three other  control
groups put on study within 3 months of the matched control  group.  All control
mice were housed in the same room with treated mice.
     Maximally and one-half maximally  tolerated doses  for the main study in
mice were estimated from a preliminary toxicity test done as described for the
NCI (1976) rat study.  Mice were started in the chronic study at 35 days of
age, and the study was concluded with  sacrifice of survivors at 92-93  weeks.
Chloroform in corn oil was administered by gavage 5 days per week during the
first 78 weeks.  Initial  dose levels of 200 and 100 mg/kg/day for males and
400 and 200 mg/kg/day for females were raised to 300 and 150 mg/kg/day for males
and 500 and 250 mg/kg/day for females  at 18 weeks.  Thus, doses expressed as time-
weighted averages for the entire study were 138 and 277 mg/kg/day for  males and
238 and 477 mg/kg/day for females.
     Survival in mice was similar among groups except  for high-dose females.
At least 50% of the animals in each group survived as  long as 85 weeks.  Ten
matched control, 33 low-dose, and 30 high-dose males,  and 15 matched control,
34 low-dose, and 9 high-dose females survived for the  duration of the  study.
All but two deaths in high-dose females occurred after 70 weeks.
                                    8-11

-------
     Body weight gain among groups was comparable.   Male  and  female  mice  in-



itially weighed about 18 and 15 g, respectively.  Mean  body weights  at  50 weeks



were approximately 35 g in males and 28 g in females, and these  levels  were



generally sustained throughout the remainder of the study. Food consumption



was stated to have been equivalent among groups.  Appearance  and behavior among



groups were similar except for bloating and abdominal distension noted  in



treated animals beginning after 42 weeks of treatment.



     Statistically significant (P < 0.05) increases in  hepatocellular  carcinomas



in both sexes in both treatment groups of mice were observed  (Table  8-4).



Various histopathologic types of hepatocellular carcinomas were  observed.



Hepatocellular carcinoma metastasized to the lung in two  low-dose males and



two high-dose females, and to the kidney in one high-dose male.   Twenty animals



were reported as missing or autolyzed, and therefore were not  included  in the



pathology report.



     Non-neoplastic lesions in mice attributed to treatment include  nodular



hyperplasia of the liver in 10 low-dose males, six  low-dose females, and  one



high-dose female;  and liver necrosis in one low-dose male, four  low-dose  females,



and one high-dose  female.   Nine high-dose females with  hepatocellular  carcinoma



had cardiac atrial  thrombosis.  Kidney inflammation was diagnosed in 10 matched



control males, two low-dose males, and one high-dose male.



     Under the conditions  of this bioassay,  chloroform  treatment significantly



(P < 0.05) increased the incidence of hepatocellular carcinoma in male  and



female B6C3F1 mice.  Although the number of matched vehicle controls was  low,



the use of pooled  colony controls gives additional  support for treatment-related



effects.  Moreover, historical  control incidence of hepatocellular carcinomas



in B6C3F1 mice was reported as 5-10% in males  and 1% in females.
                                    8-12

-------
                     TABLE 8-4.  EFFECTS OF  CHLOROFORM ON HEPATOCELLULAR CARCINOMA INCIDENCE IN B6C3F1 MICE
                                                           (NCI  1976)
   Treatment3
                                       Male
                     Controls5

                 Colony     Matched
                         Dose  (mg/kg/day)c

                         138         277
                                                                                      Female
                Controls'5

            Colony    Matched
                       Dose (mg/kg/day)c

                       238         477
00
Hepatocellular
 carcinoma
 i ncidence^
P value8

Time to first
 tumor (weeks)
5/77(8%)   1/18(6%)    18/50(36%)   44/45(98%)


                       0.011       3.13xlO-13
                                                                       1/80(1%)  0/20(0%)
                                   36/45(80%)  39/41(95%)


                                   4xlO-10     3.7xlO-14
                      72
                              72
                          80
54
90
66
   aChloroform in corn oil administered by gavage 5 times per week.
   bColony controls consist of four vehicle-control  groups, including matched controls, given corn oil
   cDoses a^e time-weighted averages.
   dAnima1s with tumor/animals examined.
   9Fishea's Exact Test, compared with matched controls.
67
Survival at
terminal
sacrifice
(92 weeks)


48%


50%


65%


65%


81%


75%


75%


20%

-------
     A more precise estimate of dose-response  perhaps  could  have  been  obtained

if additional  lower doses had been used and if constant  doses  rather than  time-

weighted averages had been used.  Treated  animals  were housed  in  the same  room

as animals treated with other volatile  compounds*; however,  since 1) controls

were in the same room as treated animals,  2)  oral  chloroform doses  probably

would have been much higher than ambient levels of other volatiles,  3) the

cages had filters to limit the amount of chemical  released  into  the  ambient

air, 4) the total room air was exchanged 10 to 15  times  per  hour, and  5) dosing

was done in another room under a large  hood,  the likelihood  that  the other vol-

atile compounds were responsible for the observed  results is considered to be

low.  It should be noted that ambient levels  of volatiles in the  animal  quarters

were not measured.



     8.1.2.2.   ROE ET AL. (1979) --  Roe et al . (1979) studied the carcinogenicity

of chloroform in toothpaste in four strains of mice (C57BL,  CBA,  CF/1, and ICI).

The toothpaste formulation used is presented in Table  8-3.   The  chloroform

product was described as British Pharmacopoeia grade which  was not contaminated

with other haloalkanes or phosgene.  Toothpaste was prepared fresh each month.

Chloroform in arachis oil was also tested in ICI mice.

     Dose levels for the carcinogenici ty studies were  selected based on results

of a 6-week preliminary range-finding study in male and  female Schofield mice

which indicated moderate weight gain reduction at  the  lowest toxic dose of

60 mg CHCl3/kg/day.  Three different carci nogenicity studies were conducted  in
     *l,l,2,2-tetrachloroethane, 3-chloropropene, chloropicrin,
1,1-dichloroethane, trichloroethylene, sulfolene, iodoform, ethylene dichloride,
methyl chloroform, 1,1,2-trichloroethane, tetrachloroethylene, hexachloroethane,
carbon disulfide, trichlorofluoromethane, carbon tetrachloride, ethylene
dibromide, dibromochloropropane.


                                    8-14

-------
which mice, initially no more than 10 weeks old, were given chloroform by



gavage  6 days per week for 80 weeks followed by observation for 13 to 24 weeks.



In one  study, 52 male and 52 female ICI mice per dose group were given 17 or



60 mg/kg/day of chloroform in toothpaste and compared with 100 ICI mice of each



sex concurrently given toothpaste without chloroform, peppermint oil, or



eucalyptol.  A second study, confined to male ICI mice, included 52 untreated



mice, 260 mice given toothpaste alone without chloroform, eucalyptol, or



peppermint oil, and groups of 52 mice each given, in toothpaste, 60 mg



CHCl3/kg/day, eucalyptol up to 32 mg/kg/day, or peppermint oil  up to 16 mg/kg/day;



treatment with chloroform, eucalyptol, or peppermint oil  was done in the



absence of the other two compounds.  In the third study,  groups of 52 male mice



of each of the C57CL, CBA, CF/1, and ICI strains were given 60 mg CHCl3/kg/day



in toothpaste and compared with concurrent vehicle-control  groups of 52 mice



each, and with 100 untreated ICI mice.  Fifty-two ICI male mice given 60 mg



CHCl3/kg/day in arachis oil, and concurrent control  mice  given arachis oil



alone, were also evaluated in the third study.



     Body weights were recorded in each study,  and food consumption was



estimated in the second and  third studies.   In  each  study,  the  animals were



necropsied, and tumors and lesions as well  as routine tissues  and organs were



examined histopathologically.   Adrenals, kidneys,  livers,  lungs,  and  spleens



were wei ghed.



     Although the authors  stated (data were not  reported)  that  body weight  gain



was poorer in  each  treatment group than in  controls  in  the  third  study on  four



mouse strains,  differences  in  survival, body weights,  and  food  consumption



between  control  and treatment  groups  were  not statistically (P  <  0.05)  signifi-



cant, either as  shown  with data or as  stated by  the  authors  without data.



Median survival  was ^>  73 weeks  for all  groups in  the two  studies  on  ICI  mice





                                    8-15

-------
alone; by terminal  sacrifice in the study on four strains  (survival  patterns



were not reported), 52 to 79% of the C57BL and CBA mice and  12 to 31% of the



CF/1 and ICI mice were alive.  Liver and kidney weights were slightly lower



(data not reported) in male ICI mice given chloroform in toothpaste.   Incidences



of tumors and lesions between control  and chloroform-treated animals  were not



significantly (P < 0.05) different except for:  1) increased kidney  tumor



incidences in treated male ICI mice, as shown in Table 8-5,  and 2) a  significantly



(P < 0.001, chi-square test) higher incidence of moderate  to severe  kidney



"changes" in treated CBA and CF/1 males than in corresponding controls and of



moderate to severe kidney disease (P < 0.05, chi-square test) in ICI  males



given CHC13 in arachis oil than in arachis oil controls, as  described by the



authors without presentation of data.   Results in Table 8-5  indicate  more



effective induction of kidney tumors by chloroform in arachis oil than by



chloroform in toothpaste.  Kidney tumors were not found in C57BL, CBA, and



female ICI mice, and malignant kidney tumors were diagnosed  in two control



and one treated CF/1 mice.  Malignant kidney tumors were identified  as hyper-



nephromas, and benign kidney tumors were characterized as  cortical adenomas.



Eucalyptol and peppermint oil were not toxic in male ICI mice in these studies.



     Results of the studies by Roe et al. (1979) show the  ability of  chloroform



to induce kidney tumors in male ICI mice.  The stronger induction of  kidney tu-



mors by chloroform in arachis oil compared with chloroform in toothpaste may  re-



flect an effect of dosing vehicle on chloroform absorption,  since Moore et al.



(1982) demonstrated greater severity of acute toxicity and regenerative changes



in kidneys of male CFLP mice given single gavage doses of  60 mg CHCl3/kg when



corn oil rather than toothpaste was the dosing vehicle.  Kidney pathology was



noted in treated animals in the study with four strains of mice; however,



although poorer body weight gain reported for treated mice in each of the





                                    8-16

-------
              TABLE 8-5.  KIDNEY TUMOR INCIDENCE IN MALE ICI MICE
                            TREATED WITH CHLOROFORM
                         (adapted from Roe et al. 1979)
Dose group
 Numbers of mice
examined histo-
   logically
 Number of mice with kidney tumors
Benign       Malignant         Total
First study

Vehicle-control3
17 mg CHCl3Ag/dayb
60 mg CHCl3/kg/dayb

Second study

Untreated control
Vehicle-control3
60 mg CHCl3/kg/dayc

Third study

Untreated control
Vehicle-control^
60 mg CHCl3/kg/daye

Vehicle-control^
60 mg CHCl3/kg/day9
     72
     37
     38
     45
    237
     49
     83
     49
     47

     50
     48
  0
  0.
  0
  1
  2

  1
  3
0
0
31
                 0
                 0.
                 21
0
0
3


&
                0
                0
                8J
                1
                6
                9j
                0
                1
                5

                0
               12J
toothpaste base vehicle without chloroform, eucalyptol, and peppermint oi
bChloroform given in toothpaste base with eucalyptol and peppermint oil.
cChloroform given in toothpaste base without eucalyptol and peppermint oil,
toothpaste base vehicle without chloroform.
eChloroform given in toothpaste base.
fArachis oil.
9Chloroform given in arachis oil.
"Statistically significant versus  vehicle-control (P < 0.05).
""Statistically significant versus  vehicle-control (P < 0.01).
^Statistically significant versus  vehicle-control (P < 0.001).
                                      8-17

-------
strains would suggest that a maximun tolerated dose was being approached,  the



observation that survival, body weights,  and other pathology between control



and treated mice in each of the four strains were not  significantly (P < 0.05)



different also suggests that higher doses could have been tested to more strongly



challenge the mice for carcinogenicity.   Since mice were as old as  10 weeks at



the start of the studies, it is evident  that treatment could have been started



when the mice were younger to cover a greater portion  of their lifespan during



growth.  A fuller evaluation of chloroform carcinogenicity could have been made



if female mice had also been included in  each study.







     8.1.2.3.  ESCHENBRENNER AND MILLER  (1945) — An early study on hepatoma



induction by chloroform in mice was described by Eschenbrenner and  Miller



(1945).  Strain A mice, initially 3 months old, with a historical spontaneous



hepatoma rate of < 1% at 16 months of age were selected for treatment.



"Chemically pure" chloroform was used, but chemical analysis was not indicated.



Dose groups of five males and five females each were treated with doses of



2.4, 1.2, 0.6, 0.3, or 0.15 g/kg of chloroform in olive oil by gavage.  Controls



received olive oil alone.



     In the study of hepatoma induction,  mice were dosed every 4 days for a



total of 30 doses.  When 8 months old, mice were examined for hepatomas at one



month after the last dose; however, these animals were given an additional dose



of chloroform 24 hours before necropsy.   Tissues and organs were examined



histopathologically.  Liver necrosis also was examined in mice given a single



gavage treatment of one of the indicated doses of chloroform (one male and two



females per group) 24 hours before removal of liver for microscopic evaluation.



     Incidences of liver and kidney necrosis and hepatomas are shown  in Table 8-6.



Liver necrosis was noted in both sexes in the three highest-dose groups.  Males in





                                    8-18

-------
all  treatment groups developed kidney necrosis, whereas kidney necrosis was not

apparent in females.  No males in the three highest-dose groups and no females

in the highest-dose group survived the study.   All  deaths occurred by 48 hours

after the second administration of chloroform.   All  surviving females dosed

with 0.6 or 1.2 g CHCl3/kg had hepatomas.
      TABLE 8-6.  LIVER AND KIDNEY NECROSIS AND HEPATOMAS IN STRAIN A MICE
       FOLLOWING REPEATED ORAL ADMINISTRATION OF CHLOROFORM IN OLIVE OIL
                  (adapted from Eschenbrenner and Miller 1945)
Observation
Liver necrosis

Kidney necrosis

Deaths3

Hepatomas in
survi ving
animals receiv-
ing 30 dosesa

Sex
F
M
F
M
F
M



F
M
Dose (g/kg)
2.4 1.2 0.6
+ + +
+ + +
000
+ + +
5/5 1/5 2/5
5/5 5/5 5/5



4/4 3/3
— — — — *" —
0.3
0
0
0
+
0/5
2/5



0/5
0/3
0.15
0
0
0
+
0/5
0/5



0/5
0/5
Control
0
0
0
0
0/5
0/5



0/5
0/5
aNumerator is positive occurrences.Denominator is animals observed.
     In the experiment on the ability of a single dose of chloroform to produce

tissue necrosis, there was sharp distinction between normal  and necrotic cells

in liver.  Doses of 2.4 and 1.2 g/kg produced extensive necrosis in all liver

lobules, and the 0.6 gAg dose produced necrosis in some lobes.   Mice given

30 doses of chloroform in the hepatoma study had moderate liver cirrhosis and
                                      8-19

-------
necrosis, however, animals given 3U doses  that did  not  result  in  necrosis had



livers that appeared normal.   Necrosis  was not found  in  hepatoma  cells, and



hepatomas contained cords of  enlarged  liver-like  cells which formed  disorganized



anastamosing columns.  The hepatomas did not  appear invasive,  and metastasis was



not found.



     Renal necrosis in males  was localized in the areas  of the  proximal and



distal tubules.  Glomeruli and collecting  tubules appeared normal.   The severity



of renal  necrosis was less with lower  doses.  The different kidney responses



by males  and females to chloroform treatment  may  be due  to the  unique  lining



of the Bowman's capsules with flat and  cuboidal epithelium in  females  and males,



respectively (an anatomic sexual  dimorphism in mice).  Although  few  animals



were available  for pathologic examination, the Eschenbrenner and  Miller



study (1945) indicates that hepatomas  in female mice  were induced at chloroform



doses that also produced liver necrosis.   Early mortality precluded  all animals



given chloroform doses that produced liver necrosis from developing  heoatomas,



Hepatomas were  not induced by non-necrotizing doses of chloroform; however, a



lifetime  study  perhaps could  have given a  stronger  indication  of  the carcinogenic



potential of chloroform at these lower  doses.  The  observation  of kidney necrosis



in males  without tumor formation and lack  of  necrosis in hepatomas suggests



that liver in strain A mice was uniquely sensitive  to tumor induction  at



necrotizing doses, or that there might  have been  additional factors  in liver



tumor formation besides necrosis.  Furthermore, since a  dose of  chloroform was



given 1 day before sacrifice—a factor  which  in itself could have been responsible



for producing necrosis, as supported by liver necrosis  found in mice which died



after one or two treatments with chloroform—it is  not clear what the  extent of



necrosis  was during the last  month of  observation,  when  the animals  were un-



treated.





                                      8-20

-------
     8.1.2,4.  RUDALI (1967) --  Rudali  (1967)  reported  a carcinogenicity  study



on chloroform in NLC mice.  Details such as  age and sex  of the mice were not



given.  The mice received twice-weekly doses of 0.1 ml  of a 40% solution of



chloroform in oil  by force-feeding for an unspecified treatment period.



Twenty-four animals were initially on study, but only five "sound mice"  were



evidently given a pathologic examination. An average survival period of 297



days was reported, but it is not clear if this  period applied  to the total



group of 24 or to the smaller group of five.  The observation  period for the



study was not mentioned.  The use of a control  group was not indicated,  nor was



a chemical  analysis of the chloroform sample provided.



     Three of the five mice examined in pathology were diagnosed with hepatomas



and hepatic lesions; however, details of the pathologic  observations were  not



reported.  Although the study by Rudali  (1967)  gives evidence  for carcinogenic



activity by chloroform in NLC mice, it is weakened by a  lack of experimental



details, the absence of a control group, and the small  number  of animals examined



in pathology.







8.1.3.  Oral Administration (Capsules):   Dog







     8.1.3.1.  HEYWOOD ET AL. (1979) --  The carcinogenicity of chloroform in



toothpaste was evaluated in beagle dogs by Heywood et al . (1979).  The tooth-



paste formulation used was that previously described in  Table  8-3 herein except



for reduced amounts of carragheen gum and glycerol.  Chloroform in toothpaste



was transferred from a syringe to gelatin capsules immediately before dosing.



     Doses were selected from results of a preliminary range finding-study in



which one or two dogs of each sex per group  were given oral chloroform doses  7



days per week for 13 (30 and 45 mg/kg/day),  18 (60 mg/kg/day), or 12 (90 and





                                    8-21

-------
120 my/kg/clay) weeks.  Because 45 mg/kg/day,  which  produced  pathologic  changes  in



the liver, was the lowest toxic dose,  dose levels of 0,  15,  and  30 mg CHCl3/kg/day



were chosen for the carcinogenicity study.



     In the carcinogenicity study,  chloroform was given  orally in capsules  6  days



per week for over 7 years.  Eight males  and eight females were assigned to  each



treatment group and to an untreated control group,  and  16 dogs of each  sex



composed a vehicle-control group.  The dogs were initially 18 to 24 weeks old.



All of the dogs were clinically examined before treatment, and had been receiving



medication annually for common diseases.  Dogs were fed  200 g of diet twice



daily until week 300, when obese dogs  received reduced  daily rations of 300 g.



Body weights, food consumption, and water intake were estimated  during  the



study.  Hematology, serum biochemistry,  and urinalysis were  included in the



evaluation of chloroform toxicity.   Treatment was stopped at 376 weeks, and



survivors were sacrificed for macroscopic examination at 395 to  399 weeks.



Major organs were weighed.  Tumors, lesions,  and routine tissues and organs



were evaluated microscopically.  Liver and kidney specimens  from control  and



high-dose dogs were also examined by electron microscopy.



     Survival, body weights, food and  water consumption, and appearance of  the



eyes were unaffected by chloroform treatment.  Mean body weights increased  from



7 to 8 kg initially to a maximum of 14 to 15  kg; however, reduction of  diet



portions for obese dogs complicated the  body  weight results.  Results of blood



and urine analyses were unremarkable except for dose-related increases  in SGPT



levels (Table 8-7), which could reflect  liver pathology.



     No treatment-rel ated carcinogenic effects were found in necropsy and



microscopic examination of tissues  and organs.  Non-neoplastic diagnoses  showed



that fatty cysts in the livers of all  groups  were larger and more numerous  in



treated dogs.





                                    8-22

-------
                                  TABLE 8-7.  SGPia CHANGES IN BEAGLE DOGS TREATED WITH CHLOROFORM
                                                 (adapted from Heywood et al. 1979)
CO

ro
CJ
Group mean SGPT (MU/ml )
Treatment
(mg CHCl3/kg/day)

30 mg
15 mg
Vehicle-control
Untreated
Pretreatment
6 26
24 34b 58C
22 29 33
22 29 30
24 30 30
Treatment stage (weeks)
52 104 156
52C 64C 76C
32 45 46d
29 40 30
27 37 29
208
91C
55d
40
30
260 312 372
147C 128C 102C
95C 89C 66
33 47 51
32 50 50
Post-treatment (weeks)
14
105d
53
56
53
19
111
48
128
56
         aberum giutamic-pyruvic transaminase.
         ^Comparison with untreated group; P < 0.05.
         cComparison with untreated group; P < 0.01.
         ^Comparison with untreated group; P < 0.001.

-------
     The study by Heywood et al.  (1979) did not show a carcinogenic  effect  of



chloroform in toothpaste given to beagle dogs.  Range-finding tests  and SGPT



and liver fatty cyst diagnoses in the carcinogenicity study suggest  that a



maximally tolerated dose was approached in the carcinogenic!ty study.   It is



not certain if 7 years was long  enough for carci nogenicity testing with respect



to the lifespan of the beagle dog (13 to 14 years),  but by 7  years spontaneous



tumor formation was becoming evident.








8.1.4.  Intraperitoneal Administration:  Mouse








     8.1.4.1.  ROE ET AL. (1968)  --  Roe et al . (1968) investigated  the carcino-



genicity of chloroform in newborn (C57 X DBA£ Fl) mice.  ChToroform  was subcu-



taneously injected into one group of mice as a single dose of 200 ug when the



animals were less than 24 hours  old, and into another group of mice  as  eight



daily doses of 200 ug, each beginning when the animals were 1 day old.   Control



groups were given the dosing vehicle, a*rachis oil, alone.   Survivors were



sacrificed for necropsy at 77 to  80 weeks.



     No carcinogenic effect of chloroform was found.  However, since the study



was reported as an abstract, experimental  details were not provided.  ChToroform



doses were rather low, and the use  of newborn mice given one  or a few  doses of



chloroform is not equivalent to  lifetime treatment of animals given  doses as



high as those maximally tolerated.   Additionally, there may be differences  in



chloroform metabolism between newborn and adult (C57 x DBA? Fl) mice.   Hence,



it is concluded that the study by Roe et al. (1968)  does not  present sufficient



evidence  for an absence of carcinogenicity in chloroform.
                                      8-24

-------
     8.1.4.2.  THEISS ET AL. (1977) --  The carcinogenicity of chloroform was



evaluated by Theiss et al. (1977) by means of the pulmonary tumor induction



bioassay in strain A mice.



     Test animals were male strain A/St mice initially 6 to 8 weeks old.   Pre-



liminary toxicity tests were performed for selection of maximum tolerated doses;



in these tests, mice received six intraperitoneal  injections of chloroform for



2 weeks and were observed for another 4 weeks.  Results of the preliminary test



were not reported.  In the bioassay, chloroform doses in tricaprylin were 80



and 200 mg/kg, administered 3 times weekly for a total  of 24 intraperitoneal



injections; and 400 mg/kg,  which was injected only twice.  Fifty control  mice



were given tricaprylin alone.  Each treatment group contained 20 animals.  Mice



were sacrificed 24 hours after the last dose, and  lungs were removed for  counting



and examining surface adenomas microscopically.  The chloroform product was



reagent grade (Aldrich Chemical  Company),  but its  chemical  composition was not



reported.  A positive control group of 20  mice was given one injection of 1 g/kg



of urethan in saline, and compared with 50 controls given saline alone.



     Chloroform treatment did not produce  a pulmonary adenoma response in this



study.  The average number  of lung tumors  per mouse was 0 to 0.39 in each



group, except for the positive controls, which had an average of 19.6 lung tumors.



At least 90% of the mice in each group survived,  except for the mice given 400



mg CHCl3/kg, where there was  45% survival.   However,  since  this type of bioassay



is basically a screen for carcinogens, a negative  result does not  necessarily



indicate a lack of carcinogenic  potential.   Evidence for the carcinogenic



activity of chloroform is available in other studies  described  in  this document,



and,  according to the authors, there is evidence  for carcinogenic  activity in



other compounds,  e.g., 2-chloroethyl  ether  and hexachlorocyclohexane in liver,



which also tested negative  in the Theiss et  al.  (1977)  study.  Carcinogenic





                                    8-25

-------
effects of chloroform have been shown in  the  liver and  kidney,  whereas  the  lung



was apparently not a target organ in the  Theiss  et al.  (1977)  study and in



other studies,







8.1.5.  Evaluation of Chloroform Carcinogenicity by Reuber (1979).   Reuber



(1979) evaluated the carcinogenicity of chloroform based  on  his review  of slides



in the NCI (1976) bioassay, and his review of data in  other  carcinogenicity



studies described in this document.  Reuber concurred  with reported findings  of



rat kidney tumors and mouse hepatocel lular carcinomas  in  the NCI  (1976) study,



mouse hepatomas in the Eschenbrenner and  Miller  (1945)  and Rudali  (1967) studies,



and mouse kidney tumors in the Roe et al . (1979) study.   However, Reuber also



concluded that there were treatment-related neoplasms  in  the NCI  study  in addition



to those reported.  In rats, Reuber concluded that chloroform treatment induced



liver tumors (hepatocellular carcinomas and neoplastic  nodules) and cholangio-



fibromas and cholangiocarcinomas in addition  to  kidney  tumors.   Besides hepato-



cellular carcinomas, malignant lymphoma was also concluded by Reuber to have



been induced by chloroform treatment in mice. Reuber  also noted  that treated



rats and mice did not exhibit liver cirrhosis, that treated  rats  with thyroid



tumors generally did not have liver or kidney tumors,  and that liver necrosis



was apparent only in high-dose female mice.  The differences in histopathologic



interpretation of tissue specimens in the NCI bioassay  between the  Reuber study



and the NCI report, outside of a difference of opinion  between pathologists,



are not clear.








8.1.6.  Oral Administration (Drinking Water): Mouse:   Promotion  of Experimental



        Tumors
                                    8-26

-------
      8.1.6.1.  CAPEL ET AL.  (1979)  —  The effect of chloroform on the growth of



murine tumors was assessed by Capel et al. (1979).  Redistilled analar chloroform



was used, but chemical analysis of  the product was not indicated.  Test animals



were  male C57CL/105cSn/01a and male Theiller-Original (TO) mice, 20 to 22 g body



weight.  A cage of 20 mice drank 80 to 100 ml of water each day; hence, chloroform



was added to yield doses of  0.15 or 15 mg CHCl3Ag/day for two dose groups, with



each  mouse drinking 4 ml water per day.  Fresh chloroform solution was given



daily and was protected from light.



      In one experiment, the authors stated that 100 TO mice in each dose group



were  divided into three "approximately equal" subgroups.   One subgroup



(pretreated) was treated with chloroform for 14 days before and after inoculation



of Ehrlich ascites tumor cells.   Another subgroup (post-treated) was given



chloroform only after inoculation of tumor cells.   The third subgroup, also



inoculated with tumor cells, served as untreated controls.  Tumor cells had



been maintained in the peritoneal cavity of male TO mice by weekly passage of



10^ cells.   Peritoneal  fluid was collected 7 days after inoculation of



cells and diluted with buffered saline.  All  mice in the three subgroups were



given intraperitoneal  injections of 0.1 ml  diluent (10^ cells).   At the end



of exponential  growth at 10 days following inoculation of cells, animals were



sacrificed for removal  of peritoneal fluid.   The peritoneal  cavity was washed



with heparinized buffered saline.   Fluid and washings were combined and diluted



with buffered saline.   Cells were disrupted by sonication for estimation of



DNA levels  per ml  cell  suspension as a measure of total  cell  content.



     A second experiment was done in which  100 C57BL mice in  each dose group



were subdivided into three subgroups (pretreated,  post-treated,  control) each



of which was  treated with chloroform according to  the protocol  in the  first
                                      8-27

-------
experiment.  Each mouse received a subcutaneous injection of 10^ B16 melanoma



cells suspended in 0.1 ml  buffered saline.   Inoculum was obtained from a  C57BL



mouse which had received a transplant of syngeneic 816 melanoma cells maintained



by intramuscular passage every 14 days.   Animals were sacrificed at  21 days



after inoculation, and spleen, mesenteric lymph nodes, and lungs were examined



for metastases.



     In the third experiment, Lewis lung tumor cells were maintained by



serial  intramuscular transplantation in  C57BL mice.   A group of 100  mice



was divided into three approximately equal  subgroups (pretreated, post-treated,



control) to investigate the effect of 15 mg CHCl3Ag/day on tumor growth  and



spread according to the protocol  used in the first experiment.   Each mouse



received intramuscular injections of 2 x 10^ cells suspended in 0.1  ml  buffered



saline into a thigh.   Animals were killed 14 days after administration of



tumor cells, and both the  tumor-bearing  and the normal  thighs were skinned and



severed at the knee and hip.   Tumor weight  was estimated as the difference



between the weights of the thighs.  Pulmonary tumor foci  were also counted.



     For estimation of the effect of 0.15 mg CHCl3Ag/day in the third experi-



ment, 100 mice were divided into subgroups  of 20 animals each and were pretreated



with chloroform before (for 8, 6, 4, or  2 weeks) and after injection of the



Lewis cells.  Mice were sacrificed at 16 days after inoculation of tumor  cells,



and tumor weights and numbers of lung foci  were determined.  In these animals,



homogenates of primary tumors were prepared for B-glucuronidase estimation



and protein content.



     The results of these  experiments are summarized in Tables  8-8,  8-9,  and



8-10.  Body weights and survival  were not affected by chloroform treatment.



Ehrlich ascites tumor cells,  as equated  with DNA content, were  significantly



(P < 0.05) increased in high-dose mice,  and slightly though not significantly




                                    8-28

-------
                       TABLE 8-8.  EFFECT OF ORAL CHLOROFORM  INGESTION
                                                           (Cape!  et  al.
)N  THE  GROWTH OF EHRLICH ASCITES TUMORS
 1979)
CO
I


Dose

0. 15 mg /kg /day


15 mg/kg/day


Treatment
group
Control
Post-treated
Pretreated
Control
Post-treated
Pretreated
Number of
animals per
group
33
33
33
43
37
30

Average body
weight (g)a
38.3 + 3.7
39.4 + 2.9
37.9 _+ 3.2
39.4 + 3.4
37.5 + 3.0
37.0 + 3.9

Tumor DNA
(ug/ml )a
661 + 222
724 + 254
770 _+ 283
637 + 221
1143 + 324
827 + 245


Si gnif icant

NS
NS

P < 0.001
P < 0.001
                                     mean + S.D.
          NS = Not significant; P  > 0.05.

-------
                     TABLE  8-9.   EFFECT OF ORAL CHLOROFORM INGESTION ON METASTATIC  "TUMOR TAKES" WITH B15  MELANOMA
                                                         (Capel et al. 1979)
CO

CO
o


Dose

0. 15 mg /kg /day


15 mg/kg/day



Treatment
Control
Post-treated
Pretreated
Control
Post-treated
Pre-treated

Number of
animals per group
26
31
28
30
32
32
Animals

Spleen
15
35
36
15
31
31
with B16 melanoma invasion in organs (%)
Mesenteri c
lymph nodes
13
10
29
12
25
32
Lung
(a)a
12
10
18
6
19
20
(b)b
3
5
10
4
6
20
         aNumbers in column  (a) refer to the percentage of animals with tumor foci on the lungs.
         ^Numbers in column  (b) refer to the average number of lung metastases.

-------
         TABLE  8-10.   EFFECT OF ORAL CHLOROFORM INGESTION ON THE GROWTH AND  SPREAD  OF  THE LEWIS LUNG TUMORa
                                                  (Capel et al. 1979)
Number of Average Tumor
animals per body weight weight
Dose Treatment group
0.15 mg/ 8d
kg /day 6
4
CO o
1 L.
- 0
(control )
15 mg/ Control
kg /day Post-treated
Pretreated
20
20
20
20
20

33
33
33
(g)
30.6 +
30.6 +
29.0 +
29.0 +
29.3 +_

23.6 +
24.5 T
24.4 +_
— — : — f — R 	
3.8
3.8
3.6
3.5
2.7

1.3
2.3
2.2
(g)
3.5 + 0.81
3.3 + 0.72
3.3 + 0.54
3.2 + 0.72
3.1 +_ 0.11

1.6 + 0.31
1.7 + 0.51
1.8 + 0.12
Lung
metastases
165 + 56
170 + 41
154 + 39
147 + 44
142 _+ 34

44 + 26
57 + 19
61 + 19
B-glucuronidase
Si gni ficance
NS
NS
NS
NS



P < 0. 05
P < 0.01
acti vityb
0.33 +
0.27 +
0.38 +
0.49 +
0.58 +




0.56
0.79
0.073
0.070
0.094




Protei
n
contentc
78.2
66.8
60.1
60.3
50.8




+
+
+
T
T




4.2
2.7
5.1
4.7
6.2




 1 » ^- -J 'Jl I V «J \~ f\ ^S \ *_ .J ,J ^_ V4 lull V^  Lrlll,  \\t\f \A I I J  '  *J % U/ •
^Expressed as mole product/ng protein/fain.
cMi11igrams of protein  after extraction  mg/g  wet  weight,
^Duration of treatment  (weeks).
NS = not significant,  P >  0.05.

-------
increased in low-dose animals.   Invasion by B16 melanoma cells, especially in



the spleen, was augmented by both doses, and the numbers of lung foci  were



also greater in both treatment  groups.   Metastasis of Lewis cells was  increased



only by treatment with 15 mg CHCl3/kg«   There was no change in B-glucuronidase



levels based on tumor protein content in the low-dose group.   The increased



tumor protein levels appear to  reflect  tumor growth which was not evident by



weighing.



     The study by Capel  et al.  (1979) shows an ability of chloroform to enhance



the growth of three types of murine tumors in mice.  A dose of 15 mg CHC13A9



was effective in each experiment, whereas a dose of 0.15 mg CHCl3Ag was effective



only in the test with 1316 melanoma cells.  Although this study does not evaluate



the ability of chloroform to induce primary tumors, it does give evidence for a



promoting effect of chloroform  on the growth and spread of experimental tumors



at low doses.  However,  the mechanism by which chloroform enhanced tumor growth



in the study by Capel et al. (1979) is  not certain, and the relevance  of this



study to the evaluation  of the  carcinogenic potential of chloroform is not



clear.







8.2.  CELL TRANSFORMATION ASSAY







8.2.1.  Styles (1979).  Styles  (1979) reported an investigation on chloroform



in a cell transformation system with BHK cells, using growth in semi-solid agar



as an endpoint, as part  of a larger study (Purchase et al. 1978) done  to screen



chemicals for carcinogenic potential.  The BHK-agar transformation assay tech-



nique used has been described by Styles  (1977) and Purchase et al. (1978).



In the study reported by Styles (1979), baby Syrian hamster kidney (BHK-21/C1 13)



cells were exposed to five different doses of test substance in vitro in serum-




                                      8-32

-------
free liquid tissue culture medium in the presence of rat liver microsomal



fraction and cofactors  (S-9 mix; Ames et al. 1975).  The liver microsomal



fraction was obtained from Sprague-Dawley rats induced with Arochlor 1254.



     Cells were grown and maintained in Dulbecco's modification of Eagle's



medium in an atmosphere of 20% COg in air.  Cells were maintained at 37°C until



confluent, and then were trypsinized and resuspended in fresh growth medium.



Resuspended cells were grown until  90% confluent for transformation assays or



100% confluent for stock.  Only cells with .normal morphology were used for



assays.  To minimize spontaneous transformation frequency, cells were obtained



at low passage, grown to 90% confluency, and frozen in liquid nitrogen.  Cells



were thawed at 37°C in growth medium for further use.



     Test compounds were dissolved in DMSO or water as appropriate.   Each dose



was tested in replicate assays.   Cells incubated until  90% confluency were



trypsinized and resuspended in Medium 199 at a concentration of 10^ cells/foil.



Resuspended cells (10^) were incubated with test chemical  and S-9 mix at 37°C



for 4 hours.   After treatment, cells were centrifuged and resuspended in growth



medium containing 0.3% agar.   Survival  after treatment was estimated by incubating



1,000 cells at 37°C for 6 to 8 days before counting colonies.   Transformation



was evaluated by counting colonies  after cells were plated and incubated for 21



days at 37°C.   The dose-response for transformation was compared with that for



survival.   Styles (1977) accepted a fivefold increase in transformation frequency



above control  values at the 1X50 as a positive result.   The spontaneous trans-



formation frequency of BHK cells (72 experiments) in this  study was  50 _+ 16 per



10° survivors.   The suitability  of  the soft agar medium for colony growth was



checked by assays with polyoma-transformed BHK-21/C1 13 cells  or Hela cells.



     Cell  transformation results were negative with exposure to chloroform so-



lution  in DMSO added to culture  medium in a dose range  that included levels at





                                    8-33

-------
which toxicity was observed (Figure 8-2).  Although chloroform doses high



enough to produce toxicity did not induce transformation, exposure of cells to



chloroform as a vapor could have provided a comparison of the transformation



potential of chloroform as a vapor and chloroform in liquid solution.



     The study by Purchase et al. (1978), which was done on 120 chemicals of



various classes, showed that the BHK-agar transformation assay system was about



90% accurate in discriminating between compounds with demonstrated carcinogenic



or noncarcinogenic activity, and was in approximately 83% agreement with the



results of assays done by the authors with S.  typhimurium  (TA 1535, TA 1538,



TA 98, TA 100).   Styles (1979) indicated, without presenting numerical  data,



that the results obtained in Salmonella assays on chloroform in liquid solution



were similar to the findings of the transformation assays.   Purchase et al.



(1978) also observed that metabolically activated agents transformed BHK cells



more strongly in the presence of S-9 mix, thus suggesting that BHK cells have



limited intrinsic metabolic capability.








8.3  EPIDEMIOtOGIC STUDIES








     In the last decade there has appeared in  the literature a host of epi-



demiologic and statistical studies of cancer and exposure to the constituents



of drinking water, of which chloroform is one  (Harris 1974, Page et al.  1976,



Tdrone and Gart  1975, Buncher 1975,  Vasilenko  and Magno 1975,  De Rouen  and  Diem



1975,  McCabe 1975, Kruse 1977, Alavanja et al. 1978, Rafferty  1979, Kuzma et al.



1977,  Harris et  al.  1977,  Salg 1977,  Mah et al.  1977, Brenniman et al.  1978,



Tuthill  et al.  1979,  Wilkins 1978).   These studies have been  subjected  to



several  critical  reviews (Wilkins et  al.  1979, U.S.  Environmental  Protection
                                      8-34

-------
           g  100
           DC

           D
  50



   0




1100
           c/)
           tr

           §   900

           >
           DC
           2   700

           QC
           uu
           Q_
               500
          DC

          O
          DC

          h-
              300
               100 -
                  0
                                CHCU
            0.25       2.5       25       250



                  CONCENTRATION (//I/ml)


                  ALSO AMES-VE
                                                                2500
Figure  8-2.  Negative  result in transformation assay of chloroform which

            was also  negative in the Ames assay.  (Styles  1979)
                                  8-35

-------
Agency 1979, National  Academy of Sciences 1978)  and have been discussed  in some



detail.  Some very general  relationships have been noted by the reviewers.  Of



particular importance  is the appearance of some  consistency in the finding of



cancer of the large intestine, rectum, and bladder associated with the consti-



tuents of drinking water.



     It must be emphasized  that none of the studies discussed in this section



implicates chloroform  directly as the sole or dominant constituent of drinking



water responsible for  the excess of cancer at these sites.   Over 300 volatile



organic contaminants have been identified in drinking water, and many of these



have been identified as carcinogens (Wilkins et  al . 1979).



     However, chloroform at a peak concentration of 266 ug/1  has been shown to



exceed peak concentrations  of other detected carcinogens by levels 37 times



higher than those of the next highest carcinogen,  vinyl  chloride (Wilkins  et al .



1979).



     Chloroform measurements appear to range largely between 1 and 112



ug/1, according to a survey of 76 drinking water supplies  (Cantor et al . 1978).



     Although a direct association cannot be made, the possibility still  exists



that since chloroform  is apparently the predominant component in chlorinated



drinking water, it could be a contributing factor  in the etiology of the cancer



associated with the consumption of drinking water.



     Almost all  of the above-referenced studies  were ecological  correlation



investigations, and only a  few utilized case-control methods.  The studies



varied by sample size, cancer sites considered,  control  variables, and the types



of endpoints used as indicators.  Among the problems posed  by the data in  these



studies are the following:   1) a lack of data measuring the quantity of  chlorine



and chloroform in drinking  water; 2) the limited nature of  recently acquired



data on the quality and quantity of organics in  drinking water;  3) the limited





                                      8-36

-------
 amount of  information given  regarding personal consumption of drinking water;
 4) the long  latency periods  associated with most cancers (current cancer rates
 reflect exposures received decades earlier); and 5) the demographic effects of
 migration, which adds another dimension of difficulty to the quantification of
 personal consumption of drinking water over time.
     Since publication of the three reviews referred to above, several additional
 studies of cancer and exposure to trihalomethanes have been published.  The
 following  pages discuss each of these studies in detail.

 8.3.1.  Young et al. (1981).  Young et al. (1981) conducted a case-control  study
 in which cancer deaths in 8,029 white females were matched with non-cancer  deaths
 in some 8,029 white females for county of residence, year of death, and age
 recorded on  death certificates in 28 counties in the State of Wisconsin from
 1972 through 1977.  Information about the chlorine content of the drinking
 water of the 16,058 cases and controls was derived from mail-back questionnaires
 recently submitted to the superintendents of 202 waterworks encompassing the
 counties sampled.  The questions pertained to prechlorination and postchlorina-
 tion dosages used over the past 20 years  (average daily dose in ppm).   For  14%
 of the sample who were not served by a waterworks,  decedents were assigned
 chlorine dosages of  zero.  The assignment was on the basis of water supplied to
 decedent's usual place of residence.
     Odds ratios were calculated from a  logistic  regression model.   This model
 provided estimates of the relative risk  of site-specific cancer deaths for
exposure of the previous  20 years to  high,  medium,  and  low chlorine doses,  as
compared with no chlorination.   Urbanicity,  marital  status, and site-specific
high-risk  occupation were controlled  in  the  model.   Only colon  cancer  showed a
significant (P < 0.05)  association with  chlorine  intake in  all  three dosage

                                      8-37

-------
categories.   However, no gradient of increasing risk  with increasing dosage  was



apparent.   For the high, medium, and low dosage categories,  the odds ratios



were 1.51, 1.53, and 1.53, respectively.  All  were significant  at  P < 0.05.   In



those counties where the drinking water supplies were exposed to rural  runoff,



the odds ratios for colon cancer increased to 3.43, 3.68, and 2.94 for high,



medium, and low average daily chlorine doses when controlled for water source



depth and purification.  These were statistically significant at the P = 0.025



level.   Colon cancer mortality was not related to chlorination  in  counties not



exposed to rural runoff.  This finding is consistent  with the hypothesis that



trihalomethanes are formed through the action of chlorine on organic substances



in drinking water.



     Nonsignificant risks were evident at the remaining sites,  i.e., esophagus,



stomach, rectum, liver, pancreas, kidney, bladder, lung, brain, and breast.   The



average daily chlorine dose categories were designated by the authors as follows:



none (less than 0.01 ppm), low (0.01-0.99 ppm), medium (1.00-1.70 ppm), and high



(1.71-7.00 ppm).



     The authors made a number of assumptions regarding exposure of subjects



and controls to chloroform.  They assumed that chlorine in drinking water would



represent a good surrogate for exposure of cases and controls to chloroform,



reasoning that trihalomethanes such as chloroform are believed to result from



the reaction of chlorine with naturally occurring organics in water.  Although



drinking water at the tap was not analyzed for chloroform or other trihalomethanes,



the authors assumed that the measured levels of chlorine at the respective water-



works would correlate well with presumed exposure to chloroform in drinking water.



     Such implicit assumptions appear questionable for several  reasons.  First,



the latent period for the development of several, if not most,  of the cancer sites
                                      8-38

-------
is most probably greater than 20 years.   This is longer than the period covered



by the exposure data on chlorination of water supplies used by the authors.



     Second, migration within and around the 28-county area could have masked



any real risk that was related to exposure.   A diagnosis of colon cancer,  which



has a 5-year survival  rate of better than 46%, could have induced victims  to



migrate to urban areas (where chlorine levels were higher) in order to obtain



better medical  care, thus leading to a false positive association.



     Third, the amount of chloroform that is formed from the addition  of



chlorine is a function of several  important  variables:  the quantity of organics



in the water supply, treatment practices, and chlorine dosages.   The quantity of



organics in the water supply is, in turn, determined by the nature of  the  water



supply source.   Surface water (rivers and streams) receives large quantities of



organics from land runoff, whereas groundwater contains little or no organic



material; hence, the likelihood of chloroform formation from the addition  of



chlorine to a groundwater supply is minimal.



     Fourth, liquids intake rates and amounts vary considerably from person  to



person.  It is  clear that most people satisfy their liquid requirements through



a variety of drinks besides tap water, (e.g., milk, orange juice, coffee,



soda).  It is conceivable that many may drink little water because of  these



competing sources of liquid refreshment.   Therefore, it is probable that many



persons who were ranked as having been exposed to chloroform may in fact have



had little exposure to it.  The resulting miselassification of cases and controls



by exposure category would tend to mask any  gradient of increasing risk with



exposure if one existed.



     Another possibly confounding variable not controlled for in this  study  1s



the dietary intake of meat and foods low in  fiber content (Reddy et al. 1980),



both of which have been hypothesized as being related to colon or rectal cancer.





                                      8-39

-------
The dietary intake of such foods, however, is not known to be correlated with



the quantity of chlorine in drinking water, although the possiblity of a spurious



correlation cannot be ruled out.   In more urbanized counties where chlorine



levels are higher, residents may  consume a diet of more meat and less fiber.



     In summary, a definite association of chlorine or chloroform in drinking



water with an increased risk of colon cancer should not be made for the reasons



stated.







8.3.2.  Hogan et al.  (1979).  Hogan et al. (1979) conducted an ecological study



of site-specific cancer rates based on National Cancer Institute (NCI) cancer



mortality data by county for the  years between 1950 and 1969 (Mason and McKay



1974) and on chloroform levels in finished drinking water, as determined by the



U.S. Environmental Protection Agency (EPA) in two separate surveys (U.S. EPA



1975).  The first survey, known as the National Organics Reconnaissance Survey



(NORS), consisted of samples from 80 water treatment facilities across the



country.  The second survey covered 83 utilities in the states of Illinois,



Indiana, Michigan, Minnesota, Ohio, and Wisconsin.  Linear multiple regression



analyses were done for each set of data separately.  The dependent variable was



county site-specific cancer mortality.  Weighted and unweighted regression



coefficients were determined for  a number of independent variables selected by



the author based on a study by Hoover et al. (1976).  A variety of demographic



characteristics related to cancer mortality were used in addition to the variable



"chloroform levels" as determined from the NORS and regional surveys to explain



cancer mortality.  These characteristics were:  county population density, per-



cent of urbanization per county,  percent of nonwhite people, percent of foreign-



born, county median family income, educational level, percent of workforce



employed in manufacturing, chloroform level in drinking water samples, and





                                      8-40

-------
county population.   According to  the  authors,  the weighting  was  based  on  the



inverse of the square root  of the population  of  the  race-sex county  stratum,



and was done chiefly to  improve the precision  of the  regression  estimates.



     Significant  positive statistical  correlations were  found between  chloroform



levels in treated drinking  water  and  cancer mortality  specific for bladder,



rectum, and large intestine in the "weighted"  regression  for white females.  On



the other hand,  only stomach cancer appeared to  be positively correlated  sig-



nificantly with  chloroform  levels in  white males.  Without weighting,  cancers



of the bladder,  rectum,  thyroid,  and  breasts were significantly  correlated with



chloroform levels in white  females.   In  white  males,  cancers of  the  pancreas



and rectum were  significantly correlated with  chloroform  without weighting.



Only estimated regression coefficients were provided  with their  corresponding P



values.  The study  contained no information on actual  levels of  chloroform



observed in drinking water.  Nonwhites were not  considered because of  the small



sizes of the populations from which rates were derived.



     Ecological  studies  such as this  one are necessarily  weak because  their infor-



mation is based  on  aggregate rather than individual  data. The evidence for an



association is indirect  and definite  conclusions cannot  be drawn, although



hypotheses may be formulated.  It is  not certain whether  a multiple  linear



regression technique is  the proper method for  analyzing  such data, since the



assumption of linearity  implied in its selection may  not  be  warranted.  Also,



since the model  contains no interaction  terms, it is  implicit that the chosen



control variables are independent of  each other, and  such an assumption may also



be unwarranted.   Furthermore, as  was  mentioned in the  Young  et al. (1981)



study, these data are weakened because it was  assumed  that the subjects were



actually exposed  to the  levels of chlorine (or chloroform) indicated.  Another



limitation is that  since the chloroform  data were collected  in 1975, the more





                                      8-41

-------
relevant exposure data (assuming a  general  cancer  latency  of  1U  to  30 years)



should be those of 1920 to 1959, given  that the  site-specific  cancer mortality



data covered the period 1950-1969.







8.3.3.  Cantor et al.  (1978).  Cantor et  al .  (1978),  in  an ecological study of



cancer mortality and halomethanes in drinking water,  used  age-standardized cancer



mortality rates by site and sex in  whites  for the  years  1968-1971,  but  only in



the 923 U.S. counties  that were more than  50% urban  in  1970.   This  study  was



similar to the Hogan et al. (1979)  study  with respect to  its  design; i.e., a



weighted linear regression model was used  with sex-  and  site-specific cancer



rates as the dependent variable.  The weight was directly  proportional  to the



square root of the counties'  person-years  at  risk  and thus inversely proportional



to the standard deviation of  the estimated mortality  rate. Chloroform  (CHC13),



bromochloromethane (BTHM), and total trihalomethane  (THM)  levels were obtained



from the two EPA surveys  (U.S. EPA  1975)  used in the  Hogan et  al . study.



Demographic variables  used in the regression model on a  county-wide basis were:



percent of urbanization (1970); median  school years  completed  by persons  over  age



25; population size (ratio of 1970  to 1950 population);  percentage  of the work



force in manufacturing; and percentage  of  foreign-born.  Although a predicted,



age-adjusted, site-specific cancer  rate was calculated  for each  county  based on



this regression technique, only the data  for 76  counties,  where  more than half



of the population of the  counties was served  by  a  sampled  water  supply, were



actually used in this  correlation analysis of THM  levels with  residual  mortality



rates.  Figure 8-3 gives  a frequency distribution  of  the  chloroform levels in  these



76 U.S. drinking water supplies. The three indicators,  chloroform, bromochloro-



methane, and total trihalomethane,  were highly correlated  with one  another.



    Positive nonsignificant correlations  with THM  levels were evident with





                                    8-42

-------
         25
         20 -
     u
     z
     UJ
     cc
     DC
     D
     U
     O
     O
     u.
     O
     o
     LU
     CC
15 -
         10 —
         0
—
1


0.001

, 1



I




0.01 0.1 1.0
	 1
100
          5 —
                         MICROMOLES CHCI3/LITER
Figure  8-3.   Frequency distribution of CHC13  levels in 76 U.S. drinking
             water supplies.  The abscissa  is  linear in the logarithm of
             the level.  (Cantor et al.  1978)
                            8-43

-------
respect to several  forms of cancer, including lymphoma and kidney cancer

in males (Table 8-11).   But according to the authors, bladder cancer mortality

rates gave the strongest and most consistent association with THM exposure

after controlling for differences in social  class, ethnic group, urbanicity,

region, and extent of county industrialization (Table 8-12).   However, the

association appeared to be greatest with respect to BTHM and  not chloroform.

The corresponding correlations for chloroform were positive but nonsignificant.

The authors report that although other sites appeared to be positively correlated

with THM levels, the inconsistencies "outweigh the consistencies," thus casting

doubt on the reliability of these correlation coefficents; i.e., the direction

anc} strength of the correlations bear little relationship to  the percent of

population served by treated drinking water and/or by region.
   TABLE 8-11.  CORRELATION COEFFICIENTS BETWEEN RESIDUAL MORTALITY RATES IN
         WHITE MALES AND THM LEVELS IN DRINKING WATER BY REGION AND BY
          PERCENT OF THE COUNTY POPULATION SERVED IN THE UNITED STATES
                              (Cantor et al.  1978)
Site of THM
cancer Indicator
Kidney CHC13

Lymphoma BTHM
(non-
Hodgkins )



Kidney CHC1 3

Lymphoma BTHM
(non-
Hodgkins)
Correlation
coefficients
for regions
• North South Mountain Pacific
0.11 -0.
(0.54)a (o.
0.06 0.
(0.74) (0.

Correlati on
the percent
50-64%
-0.16
(0.44)
-0.33
(0.11)

11 0
73) (0
08 0
79) (0

coefficients
of the popul
65-84%
-0.11
(0.60)
-0.19
(0.36)

.66
.11)
.05
.92)

for counties
ation served
85-100%
0.42
(0.04)
0.36
(0.08)

of the U.S.
All regions
0.14
(0.33)
0.06
(0.70)

in which
was:
50-100%
0.07
(0.55)
-0.08
(0.81)

aP value for two-tai1edt-test is snown in parentheses.
                                      8-44

-------
TABLE 8-12.  CORRELATION COEFFICIENTS BETWEEN BLADDER CANCER MORTALITY RATES BY
      SEX AND BTHM LEVELS IN DRINKING WATER BY REGION OF THE UNITED STATES
                              (Cantor et al. 1978)
Correlation coefficients by region
Bladder cancer
Number
Male white
Female white
North
31
0.529
(0.002)
0.30
(0.11)
South
13
0.04
(0.90)
0.20
(0.51)
Mountain
7
-0.02
(0.96)
0.63
(0.13)
Total
51
0.30
(0.03)
0.33
(0.02)
aP value for two-tailed t-test is shown in parentheses.   Counties with at
 least 65% of their populations served by one water supply were included in
 this analysis.
     The authors noted an association of kidney cancer with chloroform exposure

that was restricted to males, but was significant only in counties where at least

85% of the public was served by treated drinking water.   In counties where less

than 85% was served by treated drinking water, the correlation coefficients

were actually negative.  Combining all  counties with greater than 50% served by

treated drinking water, the correlation coefficent was nonsignificant and close

to zero.  One interesting observation was that without controlling for ethnicity,

the authors found a "fairly strong" association of THM levels with colon cancer

and lung cancer rates in both sexes, and even a dose-response relationship

between these tumor sites and the proportion of the population exposed.   However,

when ethnicity was added to the regression model, these  relationships disappeared.

     Again, this is a descriptive study from which hypotheses can be

formulated only for future in-depth study.   It cannot be concluded that  even the

significant positive correlations in the study indicate  any evidence of  real

associations.  As the authors point out, potential  sources  of error (i.e.,


                                      8-45

-------
control  of confounders such as cigarette smoking and  diet)  are  particularly



difficult since no direct information  is available  on the  individuals  studied.



The main problem with such studies,  as mentioned earlier,  is  that  the  data  are



aggregate rather than individual.   Such data frequently  include large  numbers



of individuals who never received  the  exposure in question.   Associations



derived from such data may be misleading and are often unreliable.








8.3.4.   Gottlieb et al. (1981).   Gottlieb et al. (1981)  completed  a  case-control



study of the relationship between  Mississippi  River drinking  water and the  risk



of rectum and colon cancer.  The study was based on mortality data gathered  from



20 parishes in southern Louisiana.   Rectal and colon  cancer deaths (692 and  1,167,



respectively) from 1969 to 1975 were matched one-to-one  to  non-cancer  deaths by



age at death, year of death, sex,  and  race, within  the same parish group.   A



parish group consisted of similar  parishes with respect  to  industrial  and  urban-



rural characteristics and were defined so that each parish  included  nearly



equal populations using groundwater and surface water supply  sources,  based  on



information from the 1970 census.



     Three different estimators of exposure were used.  The first, "source!ife,"



is defined as follows:  "mostly surface" (birth and death  in  a  surface-water-using



parish); "some surface" (some known surface water use at birth  or  death);  "pos-



sible surface" (death in a groundwater parish but had either unknown or out-of-



state birthplace); and "least surface" (birth and death  in  a  groundwater-using



parish).  Length of residence was  also considered,  if known and for more than



10 years.  The second index used was chlorine level (none,  low  [less than  1.09



ppm], or high [greater than 1.08 ppm]).  The third  index was  the level of



organics in the drinking water (low [less than 68 ppm] and  high [greater than



or equal to 68 ppm]).  Sourcelife could be determined for  99.2% of the entire





                                      8-46

-------
 group  of  3,718  cases  and  controls,  but  51%  had  no  data  for  length  of  residence



 or  had lengths  of residence  of  under  10 years.  For those with  lengths  of



 residence  of  less than  10 years, water  sources  during the possible carcinogenic



 period were unknown.  Chlorine  values were  available for some 78.9% of  the



 3,718  sources,  while  organics levels were available for only 50.1% of the sources.



 The analyses  using the  latter two variables were equivocal, possibly due to the



 lack of information on  these parameters.



     Colon cancer was found  not to  be related significantly to any water variable,



 although the  number of  colon cancer cases in this  study (1,167) was greater than



 the number of rectum cancer  cases (692).  The authors hint that the earlier



 correlation found in ecological  studies could have resulted from confounding



 with urban lifestyles.  Rectal cancer,  on the other hand, was found to be



 significantly elevated with  respect to  surface or Mississippi  River water



 consumption.  Based on sourcelife, the  odds ratio for rectal cancer for those



 who were born and died using groundwater sources was 2.07 (95% confidence interval



 [C.I.]  1.49-2.88) based on a multidimensional  contingency table analysis.



 Chlorination was significantly associated with rectal  cancer,  and for those who



 used river water, the risk decreased as the distance from the  mouth of the



 river  increased.  The odds ratio for cancer of the rectum at a  location below



 New Orleans versus one above the city was 1.82 (95% C.I. 1.01-3.26).   The authors



 noted that both sexes were at increased risk.   With respect  to  controlling for



 the effect of chlorination where adequate numbers  existed, the  surface water versus



 groundwater effect on rectal  cancer  was of only borderline significance (P =



 0.05),  implying a chlorine effect.



     With  respect to levels  of organics, information was available  for over  48%



of the  rectal  cancer group and their controls.   The odds ratio  calculated based



on these data  was nonsignificant (Table 8-13),  but  was probably  subject  to  some




                                      8-47

-------
      TABLE  8-13.   RISK OF  MORTALITY  FROM  CANCER  OF  THE  RECTUM  ASSOCIATED
                   WITH LEVELS  OF  ORGANICS IN  DRINKING WATER
                             (Gottlieb  et  al.  1981)

High (J> 68 ppm)
Low (< 68 ppm)
Total
Odds ratio
Cases
110
232
342

Controls
97
220
317
1.08
bias with respect to availability of exposure data as a function of date of death.

     With respect to colon cancer, the authors felt that since they had grouped

the parishes according to industry and urban characteristics (matching was done

within the parish group), they successfully eliminated urban lifestyle as a

confounder in their evaluation of colon cancer and drinking water.

     The results of this study suggest that cancer of the rectum is linked to

the consumption of surface water, and since chlorination appears to be an effect

modifier altering the risk ratio to only borderline significance, it would

seem that chlorination does contribute to the risk of rectal cancer.
8.3.5.  Alavanja et al. (1978).  Alavanja et al. (1978) reported on a case-

control study of 3,446 gastrointestinal and urinary tract cancer deaths

(1,595 females and 1,851 males) occurring during a 3-year period from 1/1/68

to  12/31/70 in seven counties of New York State.  Some 3,444 individually

matched noncancer deaths were also selected.  Independent variables were:

residence in an urban or rural area, residence in an area served by chlorinated

or  nonchlorinated water, residence in an area served by surface water or ground-


                                       8-48

-------
 water,  and  occupation.  Cases were  taken  from computer tapes of New York State



 death  certificates,  and were individually matched with an equal number of



 non-cancer  deaths  for the  same year.  Matching variables were  age, race, sex,



 foreign-  versus  United States-born, and county of usual residence.  If potentially



 confounding  variables could not be  controlled via the matching process, the



 cases  and controls were stratified  by these confounding variables.  The data



 were analyzed  by the chi-square test.  A statistically significant excess of



 gastrointestinal and urinary tract  cancer mortality occurred among women in the



 urban  county of Erie (odds ratio [OR] = 2.08), with nonsignificant excesses



 in  Schenectady County (OR  = 2.98) and Allegany County (OR = 4.13).  Likewise,



 among  men a  statistically  significant excess of gastrointestinal  and urinary



 tract  cancer mortality occurred in Erie County (OR = 2.15) and Rensselaer



 County  (OR = 1.98), and a  nonsignificant excess occurred in Schenectady County



 (OR =  1.96) and Allegany County (OR = 2.85).  Although the study encompassed a



 seven-county area, almost  two-thirds of the deaths occurred in Erie County.



 The combined overall  odds  of dying from gastrointestinal  and urinary tract



 cancer for all seven counties  combined (including Erie),  were only 1.79 based



 on  3,446 cases, whereas in Erie County alone they were 3.15 based  on 2,177  cases.



 The authors concluded that males and females residing  in  the chlorinated  water



 areas of the counties noted above were at  a greater  risk  of gastrointestinal



 and urinary tract cancer mortality not due to  age,  race,  ethnic distribution,



 urbanicity, occupation,  inorganic  carcinogens  (Cd, As,  Be,  Pb,  Ni, N03),  or



 surface/groundwater difference.   No  environmental  data  are  provided, however,



 to  characterize quantities of  chlorine (or chloroform)  exposure.   "Inadequate



water quality data" prevented the  authors  from  making a  "definitive claim that



the process of chlorination is  directly  or indirectly  responsible  for  the
                                    8-49

-------
greater risk of gastrointestinal  and urinary  tract  cancer mortality"  in



chlorinated cancer areas.   No description is  given  of how residence was



classified into chlorinated versus nonchlorinated water areas  or surface  water



versus groundwater areas through  the use of water distribution maps,  a  practice



which can result in misclassification on the  basis  of exposure.   Again,  because



of the lack of individual  dosage  data on chloroform exposure and the  low  signifi-



cance of the risks described, this study can  only be regarded  as suggestive for



gastrointestinal and urinary tract cancer mortality.







8.3.6.  Brenniman et al. (1978).   Brenniman et al.   (1978) attempted  to  confirm



the findings of Alavanja et al.  (1978) in a case-control study of gastrointestinal



and urinary tract cancer mortality among whites in   70 Illinois communities using



both chlorinated and nonchlorinated groundwater.  The authors  limited the study to



groundwater because of the possible introduction of confounding effects due to



agricultural runoff and industrial sewage in surface water.   The 3,208 cases and



43,666 controls used were those of  Illinois deaths   occurring between  1973 and



1976.  Controls were selected from a pool of non-cancer deaths after the elimina-



tion of certain special types of deaths, such as perinatal deaths.



     Chlorinated groundwater communities were matched with nonchlorinated



groundwater communities that were similar with  respect to urbanicity and



Standard Hetropolitan Statistical Area  (SMSA) description.  To ensure a minimum



follow-up period, water supplies were categorized as chlorinated or nonchlorinated



according to a  "1963 inventory of municipal water facilities."  Additionally,



questionnaires were sent to water treatment plants  in the communities to verify



the  1963  data.  The beginning dates  for  chlorination were obtained for many of



the  plants.  Based on an EPA survey, it  was found that  14 chlorinated groundwater
                                       8-50

-------
 supply sources  in Illinois  had  chloroform  concentrations  ranging  from  less  than



 1 ug/1 to  50 ug/1,  with a mean  concentration  of  10.8  ug/1.



      In females,  statistically  significant  increased  relative  risks  of cancer



 of the large intestine  and  rectum  (OR  =  1.19,  P £ 0.05),  as well  as  total



 digestive  tract cancer  (excluding  liver) (OR  = 1.15,  P £0.05), were found  for



 chlorinated  versus  nonchlori nated  Illinois  groundwater supplies.  With respect



 to total gastrointestinal and urinary  tract cancer, the risk was  significantly



 increased  in females living  within standard metropolitan  statistical areas  (OR



 = 1.28,  P  <_  0.025)  and  within urban areas  (OR  = 1.24, P £ 0.025)  between



 chlorinated  and nonchlorinated  groundwater  communities.



     Where evidence was available concerning a history of chlorination, the



 authors  noted that  the  relative risk of total   gastrointestinal  and urinary



 tract  cancer tended to  increase with time from initial chlorination, although



 the  change was small.   The greatest increase occurred in urban nonstandard



 metropolitan areas  (OR  =  1.14 if chlorinated since 1963 and nonsignificant, but



 OR = 1.28  if chlorinated since 1953 and significant,  P £0.025).  Although



 several  significant findings were observed in  this study,  it is surprising that



 the authors  so readily dismissed the results of their own  study on the  basis that



 confounding  factors such as diet, smoking,  and occupation  were  not controlled.



 These  authors felt that the findings were tenuous and did  not  confirm the findings



 of Alavanja et al.  (1978) either in strength or in consistency.  They state that



 "chlorination of groundwater does  not  seem  to  be  a major factor in the  the



etiology of site-specific gastrointestinal  and urinary tract cancers."








8.3.7.   Struba (1979).   Struba  (1979),  as part of  his  Ph.D.  thesis,  completed a



case-control  study of mortality  in  North  Carolina  on  individuals  who  died at



age 45  or under  during  the period  1975-1978.  The  cancer sites  studied  were  the





                                    8-51

-------
rectum, colon, and urinary bladder.   Between 700 to 1,500 cases per site were

matched with controls by age, race,  sex, and geoeconomic region (coastal,

piedmont, or mountain).   Non-cancer  deaths were excluded if cancer was listed

as a contributory or underlying cause of death.  For colon and rectal  cancer,

certain precancerous colonic disorders were excluded (ulcerative colitis,

familial polyposis, and adenomatous  polyposis).  Water data were classified by

source, treatment, and previous use.  "Source" was defined as ground,  surface

uncontaminated, or uncontaminated by upstream pollution.  "Treatment"  was

defined as none, prechlorinated, post-chlorinated, or both.  "Previous use"

included the following 15 categories of upstream pollution for contaminated

water only:

                       (1)  Tobacco manufacturing
                       (2)  Textile manufacturing
                       (3)  Textile bleaching and dyeing
                       (4)  Furniture manufacturing
                       (5)  Pulp and paper mills
                       (6)  Chemical industries
                       (7)  Petroleum refining
                       (8)  Rubber and plastics manufacturing
                       (9)  Leather tanning and finishing
                      (10)  Abrasives, asbestos, minerals
                      (11)  Primary metals industries
                      (12)  Electroplating
                      (13)  Electric power generation
                      (14)  Urban areas >  50,000
                      (15)  Out-of-state upstream discharges

     The author found small but significant odds ratios  (1.3 to 2.0)  for all

three  sites  in  rural areas, as well as significant odds  ratios for each  of the

water  quality  variables in  many stratified or  combined  analyses.   Odds  ratios for

urban  areas  (population over  10,000) were  generally not  significant.  Urbanization

was  shown  to be an  effect modifier  for colon cancer and  a  likely  confounder  for

rectal  and bladder  cancers.   The author considered socioeconomic  status  to be a

likely  confounder  for cancer  of the  rectum and  bladder.  Multivariate analyses

showed  no  evidence  that occupation  acted as a  confounder for bladder  cancer  in


                                       8-52

-------
this study.  To estimate migration effects, cases and controls were stratified



by place of birth and death (birth and death in the same county; birth and



death in North Carolina; and death in North Carolina, birth unspecified) and



substratified by region, age, race, sex, and urbanization.  Odds ratios for



treatment  (chlorinated and nonchlorinated) were computed for all of the strata.



For all three cancer sites, the group with least migratory influence had the



highest odds ratio, thus lending support to the author's supposition that an



increasing migratory effect is associated with a decreasing risk of cancer of



all three sites.



     Additionally, Struba found an increasing gradient of risk from the coastal



regions of North Carolina to the mountains, a finding that he maintains is



consistent with a stronger contrast between surface water and water from deep



wells than between surface water and water from shallow wells, which are known



to be susceptible to contamination by surface water seepage into groundwater



aquifers.   However, the author notes that this difference could be due to



differences in water treatment practices or confounding by uncontrolled factors



such as dietary habits or lifestyles.







8.3.8.   Discussion.  These later ecological and case-control  studies of chlorine



exposure and cancer risk from water supplies consistently support the finding



of an increased risk of bladder, colon,  and rectal  cancer from exposure to



chlorinated water.   This association is  at best weak,  although significant,  as



evidenced  by odds risk ratios  that range up to 3.6  in  the Young et al.  (1981)



study,  but generally fall  between  1.1  and 2.0 (see  Table 8-14) in the remaining



case-control  studies.   The risk  ratios derived in these studies could be



explained  by the confounding effects of  uncontrolled influences such as smoking,



diet,  air  pollution, occupation, and lifestyle.   However, the  consistency  of





                                      8-53

-------
        TABLE 8-14.   CANCER RISK ODDS RATIOS AND 95% CONFIDENCE  INTERVALS
                       (CHLORINATED VERSUS UNCHLORINATED)
                             (Crump and Guess 1982)
Site
Rectum
Alavanja
et al.
1978a
1.93
(1.32, 2.83)
Brenniman
et al.
1978b
1.26 (crude)
(0.98, 1.61)
Young
et al.
1981°
1.39 high
(0.67, 2.86)
Gottlieb
et al.
1981d>e
1.41
(1.07, 1.87)
Struba
1979d,e
1.53
(1.24, 1.

89)
                       1.22 (ad-
                       justed)
                              1.16 medium
                             (0.58,  2.32)
                               1.13  low
                             (0.61,  2.08)
Colon 1.61
(1.28, 2.03)


1.08 (crude)
(0.96, 1.22)
1.1 (ad-
justed
1.51 high 1.05
(1.06, 2.14) (0.95, 1.18)
1. 53 medium
(1.08, 2.00)
1.30
(1.13, 1.50)


                                       1.53 low
                                     (1.11, 2.11)
Bladder
    1.69
(1.11,  2.56)
1.04 (crude)   1.04 high
1.07
1.54
0.98 (ad- 1.03 medium
justed) (0.42, 2.54)
1.06 low
(0.60,
3.09)
Calculated for both sexes and all  races combined.   Confidence intervals
 were not stated in Alavanja et al.  (1978).   Crump  (1979) calculated them by
 applying the method of Fleiss (1979) to data in Alavanja et al.  (1978).
^Calculated for Caucasians of both  sexes.   Adjusted values were adjusted for
 age, sex, urban/rural, and SMSA/nonSMSA.   Confidence intervals were not stated
 in the original report.   Crump (1979) calculated them by applying the method
 of Fleiss (1979) to data on total  cases and total  controls supplied by Dr.
 Brenniman in personal  communication.
Calculated for white females and for high,  medium, and low
 chlorine doses compared with no chlorination.   Odds ratios
 intervals computed by logistic regression,  controlling for urbanization,
 marital status, and site-specific  occupation.
Calculated for both sexes and all  races combined.
eStruba and Gottlieb et al. also computed  odds  ratios for surface water versus
 groundwater as follows.   Struba:  rectum 1.55  (1.26, 1.91); colon 1.27
 (1.10, 1.46); bladder 1.48 (1.22,  1.80).   Gottlieb et al.:  rectum 1.51 (1.21,
 1.90); colon 0.95 (0.88, 1.03).
                                                    average  daily
                                                    and  confidence
                                      8-54

-------
 the finding across several independent and diverse study groups supports the
 finding of a definite risk.  Of course, all of the case-control studies use
 residence data and cause-of-death information from death certificates, and
 thus are not strictly incidence studies.  Bias can creep in from several
 sources: differential survivorship rates due to proximity to better medical
 care and treatment facilities, higher socioeconomic status, and the possibility
 of migration of newly diagnosed cancer patients to major medical care centers
 where chlorination is used to a greater extent.  Underestimates of risk can
 result from failure to control for migration before to diagnosis, misclassifica-
 tion of cause of death, and use of chlorination as a surrogate variable in
 place of more direct measurements of chloroform, especially if the chlorinated
 source contains few organic contaminants.   Hence, the association is weak but
 significant with regard to the three cancer types and exposure to chlorinated
 drinking water.  Since exposure to chlorine in water is not the same as exposure
 to chloroform, the most that can be said is that there is a suggestion of an
 increased risk of cancer of these three sites from exposure to chloroform.   If
 this risk truly exists, it may be due to an intermediate in the natural synthesis
 of chloroform (communication with Dr.  Kenneth P. Cantor, NCI).
     In summary, it appears that there may be a suggestion  of an increased risk
 of certain forms of cancer (bladder,  large intestine,  and rectum)  due to the
 presence of tri'halomethanes in drinking water.   Beyond this, little more can be
 said.   The significant excess risk of colon cancer from chlorine in drinking
water does not constitute evidence of an association  of colon/rectal  cancer with
 chloroform.   The statistically significant positive correlation of  bladder  can-
cer and BTHM levels in drinking water is not  readily  attributable to  chloroform.
The evidence of a significant association  of  kidney cancer  with chloroform
exposure in  drinking  water is even more questionable,  since it  was  based  on
                                      8-55

-------
the findings of only one study,  which  was  confined to males  residing in counties



where more than 85% of the population  was  served  by  treated  drinking water.  A



statistically positive correlation  was seen  only  in  males  residing  in  counties



with over 85% treated drinking water.   No  association was  observed  in  females



in these same counties, and the  correlations  were actually negative for both



males and females in counties with  less than  85%  treated  drinking water.



     It appears that these case-control and  ecological  studies  in humans  suggest



a weak association of certain forms of cancer with trihalomethanes  and chloroform



in drinking water, but further epidemiologic  research should be  performed to



confirm these findings.







8.4.  QUANTITATIVE ESTIMATION







     This section evaluates the  unit risk  for chloroform  in  air  and water and



the potency of chloroform relative  to  other  carcinogens that the Carcinogen



Assessment Group (CAG) has evaluated.   The unit risk is defined  as  the lifetime



cancer risk to humans from daily exposure  to  a concentration of  1 ug/1 in water



by ingestion or daily exposure to 1 ug/m-^  in  air  by  inhalation.  If the unit



risk is calculated from a model  that is linear at low doses, then the  unit  risk



could be used as the slope for calculating risk at low  doses.



     The unit risk estimate for  chloroform represents an  extrapolation below the



dose range of experimental data.  There is currently no solid scientific  basis



for any mathematical extrapolation  model that relates exposure to cancer  risk



at extremely low concentrations, including the unit  concentration given above.



For practical reasons, such low levels of  risk cannot be  measured directly, either



by animal experiments or by epidemiologic  studies.   Low-dose extrapolation must



therefore be based on current understanding  of the mechanisms of carcinogenesis.





                                      8-56

-------
 At the  present  time, the  dominant  view of the carcinogenic process  involves the
 concept that  most  cancer-causing agents  also cause  irreversible damage to DNA.
 This  position is based  in part on  the fact that a very large proportion of agents
 that  cause  cancer  are also mutagenic.  There is reason to expect that the quantal
 response, which  is characteristic  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 breast and thyroid cancer,  skin
 cancer  induced by arsenic in drinking water,  liver cancer induced  by aflatoxin
 in the  diet).  Some supporting evidence  also  exists from animal  experiments  (e.g.,
 the initiation stage of the two-stage carcinogenesis model  in rat  liver and  mouse
 skin).  Linearity is also supported when  the  mode  of action  of the carcinogen
 in question is similar to  that of  the background cancer production in  the  exposed
 population.
     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  to  low levels  of  the
dose-response  relationship.  The  risk estimates  made with  such  a model  should
be regarded  as conservative,  representing the plausible upper limits for the
risk;  i.e.,  the true  risk  is  not  likely to be higher than the estimate, but it
                                    8-57

-------
could be lower.  For several  reasons, the unit risk  estimate based  on animal



bioassays is only an approximate indication of the 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.   Second,  the concept  of  equivalent



doses for humans compared with animals on a mg/surface area basis is  virtually



without experimental verification as to carcinogenic response.   Finally, human



populations are variable with respect to genetic constitution and diet, living



environment, activity patterns, and  other cultural factors.



     The unit risk estimate can give a rough indication of the  relative potency



of a given agent compared with other carcinogens.  The comparative  potency of



different agents is more reliable when the  comparison is based  on studies in



the same test species, strain, and sex, and by the same route of exposure.



     The quantitative aspect  of the  carcinogen risk  assessment  is included here



because it may be of use in the regulatory  decision-making process  (setting



regulatory priorities, evaluating the adequacy of technology-based  controls, etc).



However, it should be recognized that the estimation of cancer  risks  to humans



at low levels of exposure is  uncertain.  At best, the linear extrapolation



model used here provides a rough but plausible estimate of the  upper  limit of



risk; i.e., it is not likely  that the true  risk would be much more  than the



estimated risk, but it could  very well be considerably lower.  The  risk estimates



presented in subsequent sections should not be regarded as an accurate representa-



tion of the true cancer risks even when the exposures are accurately  defined.



The estimates presented may be factored into regulatory decisions to  the extent



that the concept of upper risk limits is found to be useful.
                                    8-58

-------
 8.4.1.   Procedures  for  the  Determination  of  Unit  Risk.

      8.4.1.1.   LOW-DOSE  EXTRAPOLATION  MODEL  --  The mathematical  formulation
 chosen  to  describe  the  linear  nonthreshold dose-response  relationship  at  low
 doses  is the  linearized  multistage  model.  This model employs enough arbitrary
 constants  to  be  able  to  fit  almost  any monotonical ly increasing dose-response
 data, and  it  incorporates a  procedure  for estimating the  largest  possible
 linear  slope  (in the  95% confidence  limit sense)  at low extrapolated doses that
 is  consistent with  the  data  at  all  dose levels of  the experiment.
      Let P(d) represent  the  lifetime risk (probability) of cancer at dose d.  The
 multistage model has  the form

                P(d)  = 1 - exp  [-q0 +  qjd + qxd2 + ... + qkdk)]
 where
                          qi _>  0, i  =  0, 1, 2, ..., k

 Equivalently,

                    Pt(d) = 1 - exp [qjd + q2d2 + ...  + qkdk)]
 where
                              Pt(d)  = P(d) - P(0)
                                        1  -  P(0)
 is the extra risk over background rate  at  dose d.
     The point estimate of the coefficients  q-j,  i  = 0,  1,  2,  ...,  k,  and con-
 sequently,  the extra risk function,  Pt(d), at any  given  dose  d,  is calculated
by maximizing the likelihood function of the  data.
     The point estimate and  the 95%  upper  confidence  limit of the  extra risk,
Pt(d), are  calculated  by using the computer  program GLOBAL79, developed by
Crump  and Watson (1979).   At low doses, upper 95%  confidence  limits on  the

                                    8-59

-------
extra risk and lower 95% confidence limits  on the  dose  producing  a  given  risk



are determined from a 95% upper confidence  limit,  q*, on  parameter  q  .  Whenever



qi > 0, at low doses the extra risk P-t(d)  has approximately  the  form  Pt(d)  =



q* x d.  Therefore, q* x d is a 95% upper  confidence  limit  on  the extra risk



and R/q* is a 95% lower confidence limit  on the dose, producing  an  extra  risk



of K.  Let Lg be the maximum value of the  log-likelihood  function.  The upper-



limit q* is calculated by increasing q^ to  a value q* such  that  when  the  log-



likelihood is remaximized subject to this  fixed value q*  for the  linear



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



equation



                           2 (L0 - L!) =  2.70554





where 2.70554 is the cumulative 90% point  of the chi-square distribution  with one



degree of freedom, which corresponds to a  95% upper limit (one-sided).  This



approach of computing the upper confidence limit for  the  extra risk P^(d) is  an



improvement on the Crump et al. (1977) model.  The upper  confidence limit for



the extra risk calculated at low doses is  always linear.   This is 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, Pt(d)  will be



abbreviated as P.]



     In fitting the dose-response model,  the number of  terms in  the polynomial



is chosen equal to (h-1), where h is the  number of dose groups in the experiment,



including the control group.



     Whenever the multistage model does not fit the data  sufficiently well, data



at the highest dose are 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
                                      8-60

-------
whether a fit is acceptable, the chi -square statistic
                              X  =
is calculated where N^ is the number of animals in the i-t  dose group, X^ is



the number of animals in the ith dose group with a tumor response, P^ is the



probability of a response in the itn dose group estimated by fitting the multi-



stage model to the data, and h is the number of remaining groups.   The fit is



determined to be unacceptable whenever X^ is larger than the cumulative 99%



point of the chi-square distribution with f degrees of freedom, where f equals



the number of dose groups minus the number of non-zero multistage  coefficients.







     8.4.1.2.   SELECTION OF DATA —  For some chemicals, several  studies in



different animal species, strains, and sexes, each run at several  doses and



different routes of exposure, are available.   A choice must be made as to which



of the data sets from several studies to use in the model.   It may also be



appropriate to correct for metabolism differences  between species  and for



absorption factors via different routes of administration.   The procedures used



in evaluating these data are consistent with the approach of making a maximum-



likely risk estimate.   They are 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 statistically significantly higher than the control  for



at least one test dose level  and/or where the tumor incidence rate shows a statis-



tically  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




                                      8-61

-------
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 identical

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 geo-

metric mean of numbers Aj, A2, ..., Am  is defined as
                                x A2 x ...  x AJ
                                                1/fo
     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.



     8.4.1.3.   CALCULATION OF HUMAN EQUIVALENT DOSAGES --   Following the  sugges-

tion of Mantel and Schneiderman (1975),  it is assumed that mg/surface area/day

is an equivalent dose between species.   Since, to a close  approximation,  the

surface area is proportional  to the two-thirds power of the weight,  es  would be

the case for a perfect sphere,  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
                                      8-62

-------
Then, the lifetime exposure is
                                I e x m
                                Le x W
                                      2/3
     8.4.1.3.1.   Oral  --  Often exposures are not given in  units  of nig/day,  and


it becomes necessary to convert the given exposures  into mg/day.   Similarly, in


drinking water studies, exposure is expressed as ppm in the water.   For  example,


in most feeding  studies exposure is given in terms of ppm in the  diet.   In these


cases, the exposure in mg/day is




                                m = ppm x F x r




where ppm is parts per million of the carcinogenic agent in the diet or  water,  F


is the weight of the food or water consumed per day  in kg,  and  r  is the  absorption


fraction.   In the absence of any data to the contrary, r is assumed to be equal


to one.  For a uniform diet, the weight of the food  consumed is proportional  to


the calories required, which in turn is proportional  to the surface area, or two-


thirds power of  the weight.   Water demands are also  assumed to  be  proportional  to


the surface area, so that




                                m « ppm x


or



                                    m    a ppm.
                                     o /o
                                   rW2/3





As a result, ppm in the diet or water is often assumed to be an equivalent expo


sure between species.   However, this is not justified for the present study,


since the ratio  of calories  to food weight is  very different in the diet of  man
                                      8-63

-------
as compared with laboratory animals,  primarily  due  to  differences  in the moisture


content of the foods eaten.   For the  same  reason, the  amount  of  drinking water


required by each species also differs.   It is therefore  necessary  to use an


empirically derived factor, f = F/W,  which is the fraction  of an organism's  body


weight that is consumed per day as  food, expressed  as  follows:



                                            Fraction  of  body
                                            weight  consumed as


                  Species         W         ^food       ^water
Man
Rats
Mice
70
0.35
0.03
0.028
0.05
0.13
0.029
0.078
0.17
Thus, when the exposure is given as a certain dietary or water  concentration  in


ppm, the exposure in mg/W2/3 is
* F =
                                       x f x W = ppm x f x w1/3
                 _
                 rW2/3     w2/3        w2/3




When exposure is given in terms of mg /kg/day = m/Wr = s,  the conversion  is


simply



                                JD _ = s x W1/3
                                rW2/3




     8.4.1.3.2.  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 proportion-


ally to the amount of air breathed in, and 2) the carcinogen is a poorly water-


soluble gas which reaches an equilibrium between the air breathed and the body
                                      8-64

-------
 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.



      8.4.1.3.2.1.  Case 1.  Agents that are in the form of particulate matter

 or 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.

     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 1/day and 113-g rats breathe 105

 1/day.  For mice and rats of other weights, W (in kilograms), the surface 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 rn3/day


 For humans, the value of 30 m3/day* is adopted as a standard breathing rate

 (International  Commission on  Radiological  Protection 1977).   The equivalent

exposure in mg/W2/3 for these agents  can be derived from the air intake data  in a
     *From "Recommendation of the 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.


                                      8-65

-------
way analogous to the food intake data.   The empirical  factors  for  the  air  intake

per kg per day, i  = I/W, based on the previously stated relationships,  are

tabulated as follows:


                      Species          W          i  =  I/Id
Man
Rats
Mice
70
0.35
0.03
0.29
0.64
1.3
Therefore, for particulates or completely absorbed gases,  the equivalent  exposure

in mg/W^/3 -js


                        d =   m  = Ivr  = iUvr = -jwl/3vr
                            w2/3   w2/3   w2/3



     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.



     8.4.1.3.2.2.  Case 2.  The dose in mg/day of partially soluble vapors is

proportional to the Q^ consumption, which in turn is proportional  to W*-/3 and

is also proportional to the solubility of the gas in body  fluids,  which can be

expressed as an absorption coefficient, r, for the gas.   Therefore, expressing

the 02 consumption as 02 = k W2/3, where k is a constant independent of species,

it follows that


                               m = kW2/3xvxr
or
                                 d =   m  = kvr
                                     W2/3
                                      8-66

-------
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 ug^n3 in experimental animals is equivalent to the same concentration in

humans.  This is supported by the observation that the minimum alveolar concen-

tration 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.



     8.4.1.4.   CALCULATION OF THE UNIT RISK FROM ANIMAL STUDIES —  The risk

associated with d mgAg^/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 simply 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 ug/m3 in the air, then for

case (1),  d = 0.29 x  701/3 x 10~3 mgAg2/3/day, and for case (2),  d = 1, when

ug/fa3 is the unit used to compute parameters in animal  experiments.

     If exposures are given in terms of ppm in air, the following  calculation

may be used:


                   1  ppm = 1.2 x mo1ecu1ar weight (gas) mg/m3
                                 molecular weight (air)




Note that  an  equivalent method of calculating unit risk would be  to use mg/kg

for the animal  exposures,  and then  to increase the j^*1  polynomial  coefficient by


                                      8-67

-------
an amount


                         (wh/wa)J/3  j  = l,  2,  ....  k,



and to use mgAg equivalents for the unit risk  values.




     8.4.1.4.1.   Adjustments for Less Than Lifespan  Duration  of  Experiment  --


If the duration  of experiment Le is less than the natural  lifespan  of  the test


animal L, the slope q*, or more generally the exponent  g(d),  is  increased by

                     1     o
multiplying a factor (L/Le) .  We assume that if the average  dose  d is continued,


the age-specific rate of cancer will continue to increase  as  a  constant function


of the background rate.  The age-specific rates for  humans increase at least  by


the third power of the age and often by a considerably  higher power, as demon-


strated by Doll  (1971).  Thus, it is expected that the  cumulative  tumor rate


would increase by at least the third power of age.  Using  this  fact, it is  assumed


that the slope q*, or more generally the exponent g(d), would also increase by


at least the third power of age.  As a result, if the slope q* [or g(d)] is


calculated at age Le, it is expected 1 that if the experiment had  been continued


for the full lifespan L at the given average exposure,  the slope q* [or g(d)]

                                            O
would have been increased by at least  (L/Le) .


      This adjustment is conceptually consistent with the proportional  hazard


model proposed by Cox  (1972) and the time-to-tumor model considered by Daffer et


al.  (1980), where the probability of cancer by age t and at dose d is given by




                        P(d,t) = 1  - exp  [-f(t) x g(d)]




      8.4.1.5.  MODEL FOR ESTIMATION  OF UNIT RISK  BASED ON HUMAN DATA  —  If human


epidemiologic studies and sufficiently valid exposure  information are available
                                      8-68

-------
 for  the  compound, they are always used in some way.   If they show a carcinogenic
 effect,  the data are analyzed to give an estimate of the linear dependence of
 cancer rates on lifetime average dose, which is equivalent to the factor B^.
 If they  show no carcinogenic effect when positive animal evidence is available,
 then it  is assumed that a risk does exist, but is smaller than could have been
 observed in the epidemiologic study, and an upper limit to the cancer incidence
 is calculated assuming hypothetically that the true incidence is below the level
 of detection in the cohort studied, which is determined largely by the cohort
 size.  Whenever possible, human data are used in preference to animal  bioassay
 data.
     Very little information exists that can be used to extrapolate from high-
 exposure occupational  studies to exposures at low environmental  levels.   However,
 if a number of simplifying assumptions are made, it is possible to construct a
 crude dose-response model  whose parameters can be estimated using vital
 statistics, epidemiologic studies,  and estimates of worker exposures.
     In human studies,  the response is measured in terms of the relative risk
 of the exposed cohort  of individuals as compared with the control  group.   The
 mathematical  model  employed for the low-dose extrapolation assumes that  for low
 exposures the lifetime  probability  of death  from cancer, PQ,  may be represented
 by the linear equation

                                  P0 = A + BHx

where A is  the lifetime probability in the absence of the agent,  and x is  the
average lifetime exposure  to  environmental levels  in  units  such  as ppm.   The
factor BH is  the increased probability of cancer  associated with  each unit
increase  of x,  the  agent  in air.
                                      8-69

-------
     If it is assumed that R, the relative risk of cancer for exposed workers

as compared with the general  population, is independent of length of exposure

or age at exposure and depends only on average lifetime exposure, it follows

that


                          R = JL  =  A + BH (XT = x?)
                              P0       A + BH x 1

or

                             RP0 = A + BH (xi + x2)
where xj = lifetime average daily exposure to the agent for the general  popula-

tion, X2 = lifetime average daily exposure to the agent in the occupational

setting, and PQ = lifetime probability of dying of cancer with no or negligible

exposure.

     Substituting PQ = A + BH xj and rearranging gives




                               BH = P0 (R - 1)A2



To use this model, estimates of R and X2 must be obtained from epiderr.i ologi c

studies.  The value PQ is derived by means of the life table methodology from

the age- and cause-specific death rates for the general population found in

U.S.  vital statistics tables.



8.4.2.  Unit Risk Estimates.



     8.4.2.1.  DATA AVAILABLE FOR POTENCY CALCULATION —  Evidences of carcino-

genic activity of chloroform from lifetime treatment studies in laboratory

animals include:  1) significantly (P < 0.05) increased incidences of hepato-


                                      8-70

-------
cellular carcinomas in female and male B6C3F1 mice (Table 8-15) and kidney

tumors in male Osborne-Mendel rats (Table 8-16) in an NCI (1976) bioassay, where

animals were given chloroform in corn oil by gavage; and 2) kidney tumors in

male ICI mice given chloroform in arachis oil by gavage (Roe et al. 1979, Table

8-17).   These data sets are used to estimate the carcinogenic potency of chloro-

form using the linearized multistage model and the dose conversion procedure

as described previously.   For comparison, three other low-dose extrapolation

models, the probit, Weibull, and one-hit, are also used to calculate the carcino-

genic potency of chloroform.
TABLE 8-15.  INCIDENCE OF HEPATOCELLULAR CARCINOMAS IN FEMALE AND MALE B6C3F1 MICE
                                   (NCI 1976)
                        Human (animal)
                      dose (mg/kg/day)a                 Incidence rate


Female                       0                              0/20 (0%)

                        11.59 (238)                        36/45 (80%)

                        23.23 (477)                        39/41 (95%)



Male                         0                              1/18 (6%)

                         6.72 (138)                        18/50 (36%)

                        13.49 (277)                        44/45 (98%)


aHurnan equivalent dose is  calculated by d x (5/7) x (78/90) x (0.034/70)1/3
 = 4.87 x 10-2 x d, where  d is the animal  dose given 5 days per week for 78
 weeks (out of a lifespan  of 90 weeks).  The body weights  are assumed to be
 34 g for mice and 70 kg for humans.
                                      3-71

-------
TABLE 8-16.  INCIDENCE OF TUBULAR-CELL ADENOCARCINOMAS IN MALE OSBORNE-MENDEL RATS
                                   (NCI 1976)
Human (animal)
dose (mg/kg/day)a                                       Incidence rate
    0                                                        0/19

 8.62 (90)                                                   2/50

17.24 (180)                                                 10/50
aHuman equivalent dose is calculated by d x (78/104) x (5/7) x (0.4/70)1/3
 = 9.58 x 10~2 x d, where d is the animal dose given 5 days per week for 78
 weeks (out of a lifespan of 104 weeks).  The body weights are assumed to be
 400 g for rats and 70 kg for humans.
       TABLE 8-17.  INCIDENCE OF MALIGNANT KIDNEY TUMORS IN MALE ICI MICE
                               (Roe et al. 1979)
Human (animal)
dose (mg/kg/day)a                                Incidence rates
       0                                              0/50

  3.79 (60)                                           9/48
dHuman equivalent dose is calculated by d x (6/7) x (80/90) x (0.04/70)1/3
 = 6.32 x 10~2 x d, where d is the animal dose given 6 days per week for 80
 weeks (out of a lifespan of 90 weeks).  The body weights are assumed to be 400
 g for mice and 70 kg for humans.
                                      8-72

-------
     8.4.2.2.  CHOICE OF LOW-DOSE EXTRAPOLATION MODELS —  In addition to the



multistage model currently used by the CAG for low-dose extrapolation, three



more models, the probit, the Weibull, and the one-hit models, are used for



comparison (Appendix A).  These models cover almost the entire spectrum of risk



estimates that could be generated from existing mathematical  extrapolation



models.  Generally statistical in character, these models are not derived from



biological arguments, except for the multistage model, which  has been used to



support the somatic mutation hypothesis of carcinogenesis (Armitage and Doll



1954, Whittemore 1978, Whittemore and Keller 1978).  The main difference among



these models is the rate at which the response function P(d)  approaches zero or



P(0) as dose d decreases.   For instance, the probit model would usually predict



a smaller risk at low doses than the multistage model because of the difference



of the decreasing rate in  the low-dose region.   However, it should be noted



that one could always artificially give the multistage model  the same (or even



greater) rate of decrease  as the probit model  by making some  dose transformation



or by assuming that some of the parameters in the multistage  model are zero.



This, of course, would not be reasonable if the carcinogenic  process for the



agent were not known a priori.  Although the multistage model appears to be



the most reasonable or at  least the most general  model  for low-dose extrapola-



tion, the point estimate generated from this model  is of limited value because



the shape of the dose-response curve beyond the experimental  exposure levels



remains in question.   Furthermore, point estimates  at low doses extrapolated



beyond the experimental  doses could be extremely unstable and could differ



drastically, depending on  the amount of lowest  experimental dose.   Since upper-



bound estimates from the multistage model  at low doses  are relatively more



stable than point estimates, it is suggested that the upper-bound estimate for



the risk (or the lower-bound estimate for the dose) be  used in evaluating the





                                      8-73

-------
carcinogenic potency of a suspect carcinogen.   The upper-bound estimate can be



taken as a plausible estimate if the true dose-response curve is actually



linear at low doses.  The upper-bound estimate means that the risks are not



likely to be higher, but could be lower if the compound has a concave upward



dose-response curve or a threshold at low doses.   Another reason one can, at



best, obtain an upper-bound estimate of the risk  when animal  data are used is



that the estimated risk is only a probability conditional on  the assumption



that an animal  carcinogen is also a human carcinogen.  Therefore, in reality,



the actual risk could range from a value near zero to an upper-bound estimate.







     8.4.2.3.  CALCULATION OF THE CARCINOGENIC POTENCY OF CHLOROFORM ~  Using



the incidence data in Tables 8-15 to 8-17 and the corresponding human equivalent



doses, the maximum likelihood estimates of the parameters were calculated for



each of the four models referred to above (see Table A-l in Appendix A).   These



models can be used to calculate either point estimates of risk at a given dose,



or the virtually safe dose for a given level of risk.  The upper-bound estimates



of the risk at 1 rng/kg/day, calculated from each  of these models on the basis



of different data sets, are presented in Table 8-18.  From this table, it is



observed that the multistage model predicts a comparable risk on the basis of



four different data sets, while the probit and Weibull models are very unstable



and predict a wide range of risk depending on which data base is used for the



risk calculation.  Figures 8-4 and 8-5 present the point and  upper-bound estimates



of risk over the low-dose range, calculated from  the four models.  The carcino-



genic potency of chloroform is represented by the geometric mean of the risk



estimates calculated from the linearized multistage model, on the basis of liver



tumor data for female and male mice.  Although the risk calculated from the data



for female mice is greater than that calculated from the data for male mice, the





                                      8-74

-------
                   TABLE 8-18.  UPPER-BOUND ESTIMATES OF CANCER RISK OF 1 mg/kg/day, CALCULATED BY DIFFERENT MODELS
                                                 ON THE BASIS OF DIFFERENT DATA SETS3
CO
I
en
Data base
Liver tumors in female
mice (NCI 1976)
Liver tumors in male mice
(NCI 1976)
Kidney tumors in male rats
(NCI 1976)
Kidney tumors in male mice
(Roe et al. 1979)
Multistage
1.6 x 10-1
3.0 x 10-2
1.3 x 10-2
9.0 x lO-2
Probit Weibull One-hit
1.5 x lO'1 4. 2 x 10-1 1.6 x 10-1
7.3 x ID'12 2.4 x 1Q-3 1.5 x 1Q-1
3.0 x ID'5 1.3 x 10-3 1.6 x 1Q-2
NA NA 9.0 x ID'2
          aUpper-bound estimates are calculated by the one-sided 95% confidence limit.


          NA = not applicable.  Models are not applicable because there is only one dosed group.

-------
           0.58
                                      Dose in mg/kg/day
          Multistage/One-hit. —•
          Probit:
          Weibull:
M.L.E. (Maximum likelihood estimate)

upper-bound estimate

M.L.E.

upper-bound estimate
                                        —•   upper-bound estimate
Figure 8-4.   Point and upper-bound estimates of  four dose-response models over
              low-dose region  on  the basis of liver tumor data  for female mice.
              (NCI 1976)
                                        8-76

-------
0.25




0
c
o
Q.
0
cc.
•| 0.125
0
O)
o
c
o
CD
CJ

n
1 X'
X
/
/ X
xx /
X /
/ /
- // -
x /
X / ^XD
X
X / _DX
-•^ X X0-
x / ^-°
// ^" ^
'^<^^^--

                0
Multistage:  —
                                        1
                                       Dose in mg/kg/day
                                      M.L.E.
                    n— — D— — G—    upper-bound estimate



                    0    0    ^     upper-bound estimate

        Probit:       (Risk is too small to show in the graph)

        Weibull:     — -- — ---    M.L.E.
                    ^r— -*^— — &— —    upper-bound estimate
Figure 8-5.   Point and upper-bound  estimates of  four dose-response  models over
              low-dose region on  the basis of liver  tumor data for male mice.
              (NCI  1976)
                                         8-77

-------
estimates from both data sets  are combined because the  data  for  males  includes

an observation at a lower dose,  and the response at this  dose  does  not appear

to be inconsistent with the female data, if the linear  dose-response  relation-

ship is assumed.

     Thus, the risk at 1 mg/kg/day is



                  P = (1.6 x 10-1 x 3.0 x 10-2)l/2 = 7  x  iQ-2
     This number differs little from the geometric mean of the q* (upper-bound
                                                                1
of the linear parameter) calculated from the two data sets, and thus  is  used

herein as the slope for calculating risk at low doses.
     8.4.2.4.   RISK ASSOCIATED WITH 1 ug/to3 OF CHLOROFORM ,IN AIR --  A paucity

of information presently exists on the retention of inhaled chloroform.   The

only available estimate of pulmonary absorption of chloroform is a  study by

Lehmann and Hasegawa (1910), who estimated an approximate average of 64.1%

(67.6%, 50.2%, 74.6%) retention of a mean chloroform exposure level of 4,592

ppm breathed by three humans for 20 minutes.   The relationship is not certain

between these early data on short exposures to high chloroform levels and data

on pulmonary absorption of chloroform by humans for longer periods  at lower

exposure levels.   In the absence of additional data, absorption rates of 65%

by inhalation and 100% orally are assumed for the purpose of a unit risk estimate.

Under this assumption, 1 ug/ni3 of chloroform in the air would result in an

absorbed effective dose of 1.7 x 10-4 mg/kg/day or
        d - 0.65 x (10-3 mg/m3 x 20 m3/day)/70 kg = 1.7 x 10-4 mg /kg/day
                                      8-78

-------
      Therefore,  the  risk,  P,  associated with  1 ug/m3 of chloroform in air is

                       P  =  7 x  ID'2 x  1.7 x  10-4  =  1 x  10'5

      8.4.2.5.  RISK  ASSOCIATED WITH 1 ug/LITER OF  CHLOROFORM IN DRINKING WATER
 For drinking water exposure,  it is assumed that  100% of the chloroform in
 drinking water can be  absorbed, and that water intake  is 2 1/day.  Under these
 assumptions, the daily dose from consumption of water  containing 1 ug/1
 (1 ppb) of chloroform  is calculated as follows:

       d = 1 ug/1 x  2  I/day x 10~3 mg/ug x 1/70 kg = 2.9 x 10~5 mg/kg/day

 Therefore, the risk associated with 1 ug/1 of chloroform in water is

                       P = 7 x lO'2 x 2.9 x 10-5 =  2 x  10'6

      This estimate appears consistent with available epidemiologic data  such as
 the odd ratios for bladder cancer, which were estimated to range from 1.04 to
 1.69  (Table 8-14).   According to a survey of 76 water supply systems  in  the
 United States, the chloroform measurements ranged from 1 ug/1  to 112  ug/1.
 A rough estimate of the cancer risk on the basis  of these  statistics  ranges
 from:

                B = (1.04 - 1) x 7 x 10-4/112 = 3 x 10-7/(ug/1 )
 to
                B = (1.69 - 1) x 7 x 10-4/1 = 5 x 10-4/(Ug/l )

where 7 x ID'4 is the estimated background  bladder cancer  mortality rate  in  the
United States.
                                      8-79

-------
8.4.3.   Comparison of Potency with Other Compounds.   One  of the  uses  of the



quantitative potency estimate is to compare the relative  potencies  of carcinogens.



Figure 8-6 is a histogram representing the frequency distribution  of  potency



indices for 53 suspect carcinogens evaluated by the  CAG.   The actual  data



summarized by the histogram are presented in Table 8-19.   The potency index  is



derived from q*, the 95% upper bound of the linear component in  the multistage



model, and is expressed in terms of (mMol /kg/day)"1.  Where no human  data



were available, animal oral studies were used in preference to animal inhalation



studies, since oral studies constitute the majority  of animal studies.



     Based on the available data concerning liver tumors  in female and male



mice (NCI 1976), the potency index for chloroform has been calculated as 8 x



10°. • This figure is derived by multiplying the slope q* = 7 x 10~2 mgAg/day



and the molecular weight of chloroform,  119.4.  This places the potency index



for chloroform in the fourth quartile of the 53 suspect carcinogens evaluated



by the CAG.



     The ranking of relative potency indices is subject to the uncertainties



involved in comparing estimates of potency for different chemicals based on



varying routes of exposure in different  species, by means of data from studies



whose quality varies widely.  All of the indices presented here are  based on



estimates of low-dose risk, using linear extrapolation from the observational



range.  These indices may  not be  appropriate for the comparison of potencies  if



linearity does not exist at the low-dose range, or  if comparison is  to be made



at the high-dose  range.  If the latter  is the  case, then an  index other than



the one calculated above may be more appropriate.







8.4.4.  Summary  of Quantitative Assessment.  Four data sets  that contain



sufficient  information are used to estimate the carcinogenic potency  of





                                       8-80

-------
                                                           >
J2>


-------
TABLE 8-19.  RELATIVE CARCINOGENIC  POTENCIES  AMONG  53  CHEMICALS  EVALUATED
    BY THE CARCINOGEN ASSESSMENT GROUP  AS  SUSPECT  HUMAN CARCINOGENS1.2'3
Compounds
Acryloni tri le
Aflatoxin B^
Aldrin
Allyl Chloride
Arsenic
B[aJP
Benzene
Benzi dine
Beryl 1 ium
Cadmium
Carbon tetrachloride
Chlordane
Chlorinated ethanes
1 ,2-dichloroethane
hexachloroethane
1,1,2,2-tetrachloroethane
1 ,1 ,1-trichloroethane
1 ,1 ,2-trichloroethane
Chloroform
Chromium
DDT
Dichlorobenzi dine
1 , 1-dichloroethy lene
Di el drin
SI ope
(mg/kg/day)"1
0.24(W)
2924
11.4
1.19x10-2
lb(H)
11.5
5.2xlO-2(W)
234(W)
4.86
6.65(W)
1.30x10-1
1.61
6.9x10-2
1.42x10-2
0.20
1.6xlO--3
5.73x10-2
7x10-2
41 (W)
8.42
1.69
1.47x10-1(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
100
354.5
253.1
97
380.9
Potency
i ndex
1x10+1
9xlO+5
4x10+3
9x10-!
2x10+3
3x10+3
4x100
4x10+4
4x10+1
7x10+2
2x10+1
7x10+2
7x10°
3x100
3x10+1
2x10-1
8x100
8x10°
4x10+3
3x10+3
4x10+2
1x10+1
1x10+4
Order of
magnitude
(logic
index )
+ 1
+6
+4
0
+3
+3
+1
+5
+2
+3
+1
+3
+ 1
0
-1
+1
+1
+4
+3
+3
+1
+4
                                                 continued  on  the  following  page]
                                  8-8Z

-------
TABLE 8-19.   (continued)
Compounds
Dini trotol uene
Diphenylhydrazine
Epichlorohydrin
Bis(2-chloroethyl )ether
Bis(chloromethyl )ether
Ethylene dibromide (EDB)
Ethylene oxide
Heptachlor
Hexachlorobenzene
Hexachlorobutadiene
Hexachlorocyclohexane
technical grade
alpha isomer
beta isomer
gamma isomer
Methylene chloride
Nickel
Nitrosamines
Dimethylnitrosamine
Diethylnitrosamine
Dibutylni trosamine
N-ni trosopyrrol i di ne
N-ni troso-N-ethy 1 urea
N-ni troso-N-methyl urea
N-nitroso-diphenylamine
PCBs
Slope
(mg/kg/day)"1
0.31
0.77
9.9x10-3
1.14
9300(1)
8.51
0.63(1)
3.37
1.67
7.75xlO-2
4.75
11.12
1.84
1.33
6.3xlO-4
1.15(W)
25.9(not by q*)
43.5(not by q*)
5.43 1
2.13
32.9
302.6
4.92xlO-3
4.34
Molecular
weight
182
180
92.5
143
115
187.9
44.0
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
324
Potency
index
6x10+1
1x10+2
9x10-1
2x10+2
1x10+6
2x10+3
3x10+1
1x10+3
5x10+2
2x10+1
1x10+3
3x10+3
5x10+2
4x10+2
5x10-2
7x10+1
2x10+3
4x10+3
9x10+2
2x10+2
4x10+3
3x10+4
IxlQO
1x10+3
Order of
magnitude
(login
index)
+2
+2
0
+2
+6
+3
+1
+3
+3
+1
+3
+3
+3
+3
-1
+2
+3
+4
+3
+2
+4
+4
0
+3
                       (continued on the following page)
         8-83

-------
                               TABLE  8-19.   (continued)

Compounds
Phenols
2,4,6-trichlorophenol
Tetrachlorodioxin
Tetrach loroethylene
Toxaphene
Trichloroethylene
Vinyl chloride
Remarks:
1. Animal slopes are 95%

Slope
(mg/kg/day)"1
1.99xlO-2
4.25xl05
3.5xlO-2
1.13
1.9x10-2
1.75x10-2(1)

upper-limit slop

Molecular
weight
197.4
322
165.8
414
131.4
62.5

es based on the

Potency
index
4x10°
lxlO+8
6x10°
5x10+2
2.5x100
1x10°

linearized multis
Order of
magnitude
(logio
index)
+1
+8
+1
+3
0
0

tage 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-tnreshold
    model.

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

3.  Not all  the carcinogenic potencies  presented  in this table  represent  the  same
    degree  of certainty. All  are subject  to  change as new evidence becomes  available.
                                      8-84

-------
chloroform.  They are liver tumors in female mice (NCI 1976), liver tumors in
male mice (NCI 1976), kidney tumors in male rats (NCI 1976),  and kidney tumors
in male mice (Roe et al. 1979).  The unit risks at 1 mg/kg/day,  calculated by
the linearized multistage model on the basis of these four data  sets,  are
comparable.  The geometric mean, q* = 7 x 10-2/(mg/kg/day), of the potencies
calculated from liver tumors in male and female mice, is taken to represent
the carcinogenic potency of chloroform.  The upper-bound estimate of the cancer
risk due to 1 ug/m3 of chloroform in air is P = 1 x 10-5.   The upper-bound
estimate of the cancer risk due to 1 ug/1  in water is P =  2 x 10-6. The
carcinogenic potency of chloroform is in the fourth quartile  among the 53 suspect
carcinogens evaluated by the CAG.
     The unit risks given above are calculated under the assumption that mg per
unit of body surface area is equivalent between mice and humans.   If the dose
in mg/kg/day is assumed to be equivalent,  then these unit  risks  would  be reduced
approximately by a factor of 12.

8.5.  SUMMARY

8.5.1.  Qualitative.  Chloroform in corn oil administered  at  estimated maximally
and one-half maximally tolerated doses by  gavage for 78 weeks produced a
statistically significant increase in the incidence of hepatocellular  carcinomas
in male and female B6C3F1 mice and renal  epithelial  tumors (malignant  and
benign) in male Osborne-Mendel rats; a carcinogenic response  of  female Osborne-
Mendel rats to chloroform was not apparent in this  study.   Use of more than
two doses in these studies might have given a more  precise estimate of dose-
response.
                                      8-85

-------
     A statistically significant  increase in  the  incidence  of  renal  tumors



(benign and malignant)  was found  in  another study in  male  ICI  mice  treated  with



chloroform in either toothpaste or arachis oil  by gavage  for  80 weeks;  however,



treatment with a gavage dose of chloroform in toothpaste  for  80 weeks  did not



produce a carcinogenic  response in female ICI mice and  male mice of the CBA,



C57BL, and CF/1 strains.   Induction  of malignant  kidney tumors in male  ICI  mice



was greater when chloroform was administered, at  the  same  dose, in  arachis  oil



instead of toothpaste.   A carcinogenic response was not observed in male and



female Sprague-Dawley rats given  chloroform in toothpaste  by  gavage for 80



weeks, but early mortality was high  in control  and treatment  groups.   Gavage



doses of chloroform in  toothpaste did not show a  carcinogenic  effect in male



and female beagle dogs  treated for over 7 years.   The results  of preliminary



toxicity tests and the  carcinogenicity studies suggest  that doses of chloroform



in toothpaste given to  mice, rats, and dogs in the carcinogenicity  studies



approached those maximally tolerated.  However, chloroform doses given  to mice



and rats in toothpaste  or arachis oil were lower  than those above given in  corn



oi 1.



     Hepatomas were found in NIC  mice given chloroform in  oil  by force-feeding



twice weekly for an unspecified period of time, and in  female  strain A mice given



chloroform in olive oil by gavage once every  4 days for a  total of  30 doses at



a level which produced  liver necrosis; however, small numbers  of animals were



examined for pathology, the duration of these studies was  either uncertain  or



appeared to be below the lifetime of the animals, and no  control group of  NIC



mice was apparent.  Although a carcinogenic effect of chloroform was not evident



in newborn (C57 x DBA2  - Fl) mice given single or multiple subcutaneous doses



during the initial 8 days of life and observed for their  lifetimes, the dose



levels used appeared well below a maximum tolerated dose  and  the period of





                                      8-86

-------
treatment after birth was quite short compared to lifetime treatment.   Chloroform



was ineffective at maximally tolerated and lower doses in a pulmonary  adenoma



bioassay in strain A mice; however, other chemicals that have shown carcinogenic



activity in other tests were also ineffective in this pulmonary adenoma bioassay.



Although an ability of chloroform to promote the growth and spread of  Lewis lung



carcinoma, Erlich ascites, and B16 melanoma cells in mice has been shown, the



mechanism by which chloroform produced this effect is uncertain and the relevance



of this study to the evaluation of the carcinogenic potential of chloroform is



presently not clear.  Chloroform in liquid solution did not induce transforma-



tion of baby Syrian hamster kidney (BHK - 21/C1 13) cells in vitro at  doses



high enough to produce toxicity; additional testing of chloroform as a vapor



could have provided a comparison of cell  transformation potential between



chloroform as a vapor and chloroform in a liquid solution.



     An additional carcinogenicity study  on female B6C3F1 mice and male Osborne-



Mendel  rats given chloroform in drinking  water over a wide range of dose levels



is in progress at the Stanford Research Institute.



     There are no epidemiologic cancer studies dealing with chloroform per se.



Chlorinated drinking water can contain significant amounts of chloroform.   There



appears to be a weak but statistically significant risk of cancer of the bladder,



large intestine, and rectum with the presence of chlorine in drinking  water.



The odd ratios calculated in the latter ecological  and case-control  studies



range up to a high of 3.68 for cancer of  the colon in the Young et al. study,



but most fell between 1.1 and 2.0.  The risk ratios derived in each study



could be explained by the confounding effects of several  factors; i.e., smoking,



diet, air pollution, occupation, or lifestyle.  However,  the consistent finding



of a statistically significant excess of  cancer across several  independent



and diverse study populations supports the finding of a definite risk.  Bias





                                      8-87

-------
can creep into these studies  from differential  surrounding  rates  due  to



proximity to better medical  care and  treatment  facilities,  higher socioeconomic



status, and the possibility  of migration  of  cancer  patients  to  medical care



facilities in areas where chlorination is used  to a greater  extent.   Under-



estimates of risk may result  from failure to control  for  migration effects



prior to diagnosis, misclassification of  cause  of death,  and use  of chlorination



as a surrogate variable for  chloroform, especially  if few organic contaminants



are in the water.  Exposure  to chlorinated drinking water will  not necessarily



result in exposure to chloroform if organic  contaminants  are not  present.



Many contaminants found in drinking water other than chloroform are carcinogenic,



but they generally are found in much  smaller quantities as compared with



chloroform levels found in water sources  containing large quantities of



organics.  The presence of these other substances,  some carcinogenic, makes  it



impossible to incriminate chloroform directly as the cause of the excess cancer



at the three sites.  Hence,  there appears to be an  increased risk of cancer  of



the bladder, rectum, and large intestine from chlorinated water and, by  inference,



from chloroform.







8.5.2.   Quantitative.  Four data sets that contain  sufficient information are



used to  estimate the carcinogenic potency of chloroform.   They are liver tumors



in female mice  (NCI  1976), liver tumors in male mice (NCI 1976), kidney  tumors



in male  rats  (NCI  1976), and kidney tumors in male mice  (Roe et al.  1979).



The unit risks at  1 mg/kg/day, calculated by the linearized multistage model



on the   basis of these four data sets, are comparable.    The geometric mean,  q*



=  7 x  10"2/(mgAg/day), of the potencies calculated from liver tumors in



male and female  mice, is taken to represent the carcinogenic potency of



chloroform.  The upper-bound estimate  of the cancer risk due to  1 ug/m3 of





                                       8-88

-------
 chloroform  in air is  P  =  1 x  10"5.   The upper-bound estimate of the cancer risk



 due to  1 ug/liter in water is P = 2  x  10'6.  This estimate appears consistent



 with the limited epidemiologic data  available for humans.







 8.6.  CONCLUSIONS








     Evidence that chloroform has carcinogenic activity is based on increased



 incidences  of hepatocellular carcinomas in male and female B6C3F1 mice, renal



 epithelial   tumors in male Osborne-Mendel rats, kidney tumors in male ICI mice,



 and hepatomas in NLC and female strain A mice.   As concluded elsewhere in this



 document, no definitive conclusions can be reached concerning the mutagenicity



 of chloroform based on present evidence.  The toxicity of chloroform in liver



 and kidney, as noted in this document, is  considered to occur through covalent



 binding of  a reactive metabolic intermediate, possibly phosgene, with cellular



 macromolecules; the evidence discussed in  the metabolism section herein



 indicates that reactive metabolites of chloroform can react extensively with



 proteins and lipids, but minimally with nucleic acids.   Applying the Interna-



 tional  Agency for Research on Cancer (IARC) criteria for animal  studies, the



 level  of evidence for carcinogenicity would be  sufficient for concluding that



 chloroform  is carcinogenic in experimental  animals.



     There  are no epidemiologic  studies of cancer and chloroform per se.   There



 appears to  be an increased risk  of cancer  of the bladder,  rectum,  and  large



 intestine from chlorinated drinking  water  and,  by inference,  from  chloroform.



Applying the IARC criteria to assess the carcinogenicity of chloroform  in



humans,  there is limited evidence  for the  carcinogenicity  of  chlorinated



drinking water in humans,  and inadequate evidence for  the  carcinogenicity  of



chloroform  in humans.





                                      8-89

-------
     The overall  IARC classification  for  chloroform  is  2B, which by definition



designates chloroform as  probably  carcinogenic to humans.  In the  IARC scheme,



Group 2 chemicals are divided into higher (Group A)  and lower (Group B) degrees



of evidence depending on  whether evidence for their  carcinogenicity in humans



is concluded to be limited (Group  A)  or  inadequate (Group B).
                                     8-90

-------
                                    REFERENCES







Alavanja, M.,  I. Goldstein, and M.  Susser.   1978.  A  case-control  study  of



gastrointestinal and urinary tract  cancer mortality and drinking water



chlorination.   Pages 395-409 in Water chlorination:   Environmental  impact



and health effects, Vol. 2.  R. Jolley, H. Gorchev, D.H. Hamilton,  Jr.,  eds.



Ann Arbor, Michigan:  Ann Arbor Science Publishers.







Ames, B.N., J. McCann, and E. Yamasaki.  1975.  Method for detecting carcino-



gens and mutagens with the Salmonella/mammalian-microsome mutagenicity test.



Mutat. Res.  31:347-364.








Armitage, P., and R. Doll.   1954.   The age distribution of cancer and a  multi-



stage theory of carcinogenesis.  Br. J. Cancer 8:1-12.







Brenniman, G.R., J. Vasilomanolakis-Lagos, J. Amsel, T. Namekata, and A.M. Wolff.



1978.  Case-control study of cancer deaths in Illinois communities  served  by



chlorinated or nonchlorinated water.  Pages  1043-1057 in Water chlorination:



Environmental impact and health effects.   R.  Jolley, H. Gorchen, and H.  Hamilton,



Sr., eds.   Ann Arbor,  Michigan:  Ann Arbor Science Publishers.







Buncher, C.R. 1975.  Cincinnati drinking water:   An epidemiplogic study  of



cancer rates.  University of Cincinnati  Medical  Center, Cincinnati, Ohio.







Cantor,  K.P., R. Hoover, T.J.  Mason, and L.J. McCabe.  1978.   Association of



cancer mortality with  halomethanes in drinking water.   J.  Natl. Cancer Inst.



61(4):979-985.





                                      8-91

-------
Capel , I.D., H.M.  Dorrell, M. Jenner, M.H. Pinnock, and D.C. Williams.  1979.



The effect of chloroform ingestion on the growth of some murine tumors.   Eur.



J. Cancer 15:1485-1490.







Cox, C.B. 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., and H.A. Guess.   1982.  Drinking water and cancer:   Review of



recent epidemiological findings  and  assessment  of  risks.   Ann. Rev.  Public



Health 3:339-357.







Crump, K.S., and W.W. Watson.   1979.  GLOBAL79:   A FORTRAN program  to extra_



polate dichotomous animal carcinogenicity data  to  low  dose.   Natl.  Inst.  Environ.



Health Science, Contract  No.  l-ES-2123.







Crump, K.S., H.A. Guess,  and  L.L. Deal.   1977.   Confidence intervals and test



of  hypotheses  concerning  dose-response  relations  inferred  from animal



carcinogencity data.   Biometrics 33:437-451.







Daffer,  P.,  K. Crump,  and M.  Masterman.   1980.   Asymptotic theory for analyzing



dose-response  survival data  with application  to low-dose  extrapolation problem.



Mathematical Biosciences  50:207-280.
                                       8-92

-------
De Rouen, T.A., and J.E. Diem.  1975.  The New Orleans drinking water  contro-



versy:  A statistical perspective.  Am. J. Public Health  65:1060-1062.







Doll, R. 1971, Weibull distribution of cancer:   Implications  for models  of



carcinogenesis.  J. Roy Stat. Soc. A 13:133-166.







Dripps, R.D., J.E. Eckenhoff, and L.D. Vandam.   1977.  Introduction to anesthesia



the principles of safe practice.  5th ed.  Philadelphia,  PA:  W.B. Saunders



Company, pp.  121-123.







Eschenbrenner, A.B., and E. Miller.  1945.   Induction of  hepatomas in  mice by



repeated oral administration of chloroform with  observations  on sex differences.



J. Natl. Cancer Inst. 5:251-255.







Federation of American Societies for Experimental Biology  (FASEB).  1974.  2nd



ed.  Vol. III.  Philip L. Altman and Dorothy S.  Dittmen,  eds.  Bethesda,



Maryland.  Library of Congress No 72-87738.







Fleiss, J.L.  1979.  Confidence intervals for odds ratio  in case-control  studies:



The state of  the art.  J. Chronic Dis. 32:69-77.







Gottlieb, M.S., J.K. Cam, and D.T. Morris.  1981.  Cancer and drinking  water



in Louisiana:  colon and rectum.  Int. Epidemiol. 10(2):117-125.







Harris, R.H.  1974.  Implications of cancer-causing substances in Mississippi



River Uater.  Environmental Defense Fund, Washington, D.C.
                                      8-93

-------
Harris, R.H.,  T.  Page, and N.A. Reiches.  1977.  Carcinogenic hazards of



organic chemicals in drinking water.   Pages 309-330 in Origins of Human Cancer.



H.H.  Hiatt, J.D.  Watson, and J.A. Winsten, eds.  Cold Spring Harbor Laboratory.








Heywood, R. ,  R.J. Sortwell, P.R.6. Noel, A.E. Street, U.E. Prentice, F.J.C.



Roe,  P.P.  Wadsworth, A.N. Worden, and N.J. Van Abbe.  1979.  Safety evaluation



of toothpaste  containing chloroform.   III. Long-term study in beagle dogs.



J. Environ. Pathol. Toxicol. 2:835-851.








Hogan, M.D.,  P. Chi, D.G. Hoel, and T.J. Mitchell.  1979.  Association  between



chloroform levels in finished drinking water supplies and  various site-specific



cancer mortality rates.  J. Environ.  Pathol. Toxicol. 2:873-887.








Hoover, R.N.,  F.W. McKay, and J.F. Fraumeni.   1976.  Fluoridated drinking



water and the occurrence of cancer.  J.  Natl. Cancer Inst. 57:757-768.








Howe, R.B., and K.S. Crump.   1982.  A computer program to  extrapolate quantal



animal toxicity data to  low dose.  OSHA Contract No. 41USC252C3.








International  Commission on Radiological  Protection  (ICRP).   1977.   Recommenda-



tion of the ICRP, Pub. No.  26, adopted  Jan.  17,  1977.  Oxford,  England:



Pergamon Press.








Kovar, J., and D. Krewski.  1981.  RiskSl:   A  computer program  for  low-dose



extrapolation of quantal response toxicity data.   Department  of  Health  and




Welfare, Canada.
                                        8-94

-------
Kruse, C.W.  1977.  Chlorination of public water  supplies  and  cancer  - Washington



County, Maryland experience.  A final  report, Johns Hopkins  University,  School



of Hygiene and Public Health, to the Office of Research and  Development,  Health



Effects Research Laboratory, Cincinnati, Ohio.  Unpublished  draft.







Kuzrna, R.J., C.M. Kuzma, and C.R. Buncher.  1977.  Ohio drinking water source



and cancer rates.  Am. J. Public Health 67:725-729.







lehmann, K.B., and 0. Hasegawa.  1910.  Studies of the absorption  of  chlorinated



hydrocarbons in animals and humans.  Arch. Hyg. 72:327-342.  (German).







Mah, R.A., G.H. Spivey, and E. Sloss.  1977.  Cancer and chlorinated  drinking



water.  Final report on EPA Contract No. CA-6-99-3349-J to Health  Effects



Research Laboratory, Office of Research and Development, U.S.  Environmental



Protection Agency, Cincinnati, Ohio.







Mantel, N., and M.A. Schneiderman.   1975.   Estimating safe levels, a  hazardous



undertaking.  Cancer Res. 35:1379-1386.







Mason, T.J., and F.W. McKay.  1974.  U.S.  cancer mortality by  county:  1950-1969.



Washington, D.C., U.S. Government Printing Office.







McCabe, L.J.  1975.   Association between trihalomethanes in  drinking  water



(NORS data) and mortality.   Health Effects Research Laboratory, U.S.



Environmental Protection Agency, Cincinnati, Ohio.
                                      8-95

-------
Moore, D.H., L.F. Chasseaud, S.K. Majeed, D.E. Prentice, and F.J.C.  Roe.   1982.



The effect of dose and vehicle on early tissue damage and  regenerative  activity



after chloroform administation to mice.  Fd. Chem. Toxicol. 20:951-954.








National Academy of Sciences.  1978.  Epidemiological studies of  cancer



frequency and certain organic constituents of drinking water.   A  review of



recent literature published and unpublished.  Washington,  D.C.







National Cancer  Institute (NCI).  1976.  Report on carcinogenesis  bioassay  of



chloroform.  Natl. Tech. Inf. Serv.  PB-264018.  Springfield, VA.







National Institute for Occupational  Health and Safety (NIOSH).   1974.



Criteria for a recommended standard:  Occupational exposure to  chloroform.



UHEW Publication No. 75-114.








Page, T., R.H. Harris, and S.S. Epstein.  1976.   Drinking  water and  cancer



mortality in Louisana.  Science 193:55-57.







Palmer, A.K., A.E. Street, F.J.C. Roe, A.N. Worden, and  N. J. Van  Abbe.   1979.



Safety evaluation of toothpaste containing chloroform.   II. Long-term



studies in rats.  J. Environ. Pathol. Toxicol. 2:821-833.







Purchase, I.F.H., E. Longstaff, J. Ashby, J.A. Styles, U.  and P.A. Lefevre,  and



F.R. Westwood.   1978.  An evaluation of 6 short-term tests for  detecting organic



chemical carcinogens and recommendations for their use.  Br. J. Cancer  37:873-959.
                                      8-96

-------
Rafferty,  P.  1979.   Drinking water  quality  and  cancer  mortality patterns.



Master's Thesis.  University of  North  Carolina,  Chapel  Hill,  North Carolina.



Unpublished.





Reddy, B.S.,  L.A. Cohen, G.D. McCoy, P.  Hill, J.H.  Weisburger,  and E.L.



Wynder.  1980.  Nutrition and its relationship  to  cancer.   Naylor Dana  Institute



for Disease Prevention, American Health  Foundation,  Valhalla,  New York.



Advances in Cancer Research, Vol. 32.








Reuben, M.D.   1979.  Carcinogenicity of  chloroform.  Environ.  Health  Perspect.



31:171-182.








Roe, F.J.C.,  R.L. Carter, and B.C.V. Mitchley.   1968.   Test of  chloroform  and



8-hydroxyquinoline for carcinogenicity using newborn mice.  Br.  Emp.  Cancer



Campgn. 46:13.







Roe, F.J.C.,  A.K. Palmer, A.N. Worden, and  N.J.  Van Abbe.   1979.   Safety



evaluation of toothpaste containing chloroform.  I.  Long-term  studies in mice.



J. Environ. Pathol. Toxicol. 2:799-819.







Rudali, G.  1967.  Oncogenic activity of  some halogenated hydrocarbons used



in therapeutics.  UICC Monogr. Ser.  7:138-143.







Salg, J.  1977.  Cancer mortality and drinking water in  346 counties of the



Ohio River Valley Basin.   Ph.D.  Thesis.  Department of  Epidemiology, University



of North Carolina.  Also final report, EPA Contract No. 5-03-4528-J to Health



Effects Research Laboratory, Office of Research and Development,  U.S. Environ-



mental  Protection Agency, Cincinnati, Ohio.



                                      8-97

-------
Struba, R.J.  1979.  Cancer and drinking water quality.   Ph.D.  Thesis.   Univer-



sity of North Carolina, Chapel Hill, North Carolina.   156  pp.   Available  from:



University Microfilms International, Ann Arbor, Michigan,  Publication  No.



8022514.








Styles, J.A.  1977.  A method for detecting carcinogenic organic chemicals  using



mammalian cells in culture.  Br. J. Cancer 36:558-563.








Styles, J.A.  1979.  Cell transformation assays.   Pages 147-164 in  G.E.  Pagett,



ed.  Topics in Toxicology:  Mutagenesis in sub-mammalian systems:   Status and



significance.  Baltimore, MD:  University Park Press.








Tarone, R.E., and J.J.  Gart.  1975.  Review of "The  implications of cancer-



causing substances in Mississippi River water" by  R.H. Harris.  Bethesda,



Maryland, Mathematical  Statistics and Applied Mathematics  Section,  National



Cancer Institute.








Theiss, J.C., G.D. Stoner, N.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.








Tuthill, R.W., G.S. Moore, and D.W. Polakoff.   1979.  A statistical model for



predicting chloroform levels in chlorinated surface water  supplies.  J. Arn.



Water Works Assoc. 71:37-39.








U.S.  Environmental Protection Agency.   1975.   Preliminary  assessment of suspected



carcinogens in drinking water:  Report to Congress, December  1975.





                                      8-98

-------
U.S. Environmental Protection Agency.   1979.   National  interim  primary  drinking



water regulations; control of trihalomethanes  in drinking water;  final  rule.



Federal  Register 44(231)29:68624-68707  (November 29).   Part  III.







Vasilenko, P., and L. Magno.  1975.  Factors relating to the incidence  of



cancer mortality in New Jersey.  Princeton University,  Princeton, New Jersey.



Unpublished.







Weisburger, E.K.  1977.  CArcinogenicity studies on halogenated hydrocarbons.



Environ. Health Perspect. 21:7-16.








Whittemore, A.  1978.  Quantitative theories of carcinogenesis.   In:  Advances



in cancer research, Vol.  27.  New York, NY:  Academic Press, Inc., pp.  55-58.








Whittemore, A., and B. Keller.  1978.   Quantitative theory of carcinogenesis.



Society for Industrial and Applied Mathematics Review 20:1-30.








Wilkins, J.R.   1978.   Chlorination of public water supplies and cancer  in



Washington County, Maryland.  Ph.D. Thesis.  Department of Environmental



Health Sciences, Johns Hopkins University School of Hygiene and Public  Health.



Also final report with C.W.  Kruse on EPA Contract No. R805198-01-0 to Health



Effects  Research Laboratory, Office of  Research and Development,  U.S.



Environmental  Protection  Agency, Cincinnati, Ohio.







Wilkins, J.R., N.A. Reiches, and C.W. Kruse.   1979.  Organic chemical



contaminants  in drinking  water and cancer.   Am. J.  Epidemiology 110:420-448.
                                       8-99

-------
Young, T.B. , M.S. Kanarek, and A.A. Tslatis.  1981.  Epidemiologic study of



drinking water chlorination and Wisconsin female cancer mortality.  J. Natl.



Cancer Inst. 67(6):1191-1198.
                                     8-100

-------
                                   APPENDIX A


                COMPARISON AMONG DIFFERENT EXTRAPOLATION MODELS




     Four models used for low-dose extrapolation, assuming the independent


background, are:



Multistage:              P(d) = 1 - exp [-(q^d + ... + qkdk)]



where q^ are non-negative parameters



                                 A + Bln(d)
Probit:                  P(d) =  /   f(x) dx
                                — 00



where f(.) is the standard normal  probability density function



Weibull:                 P(d) = 1 - exp [-bdk]



where b and k are non-negative parameters



One-hit:                 P(d) = 1 - exp [-bd]



where b is a non-negative parameter.



     The maximum likelihood estimates (MLE) of the parameters in the multistage

and one-hit models are calculated by means of the program GL08AL82, which was

developed by Howe and Crump (1982).   The MLE estimates  of the parameters in the


probit and Weibull  models are calculated by means of the program RISK81, which

was developed by Kovar and Krewski  (1981).


     Table A-l presents the MLE of parameters in  each of the four models that


are applicable to a data set.
                                        A-l

-------
 TABLE A-l.  MAXIMUM LIKELIHOOD ESTIMATE OF THE PARAMETERS FOR EACH OF THE FOUR EXTRAPOLATION MODELS,
                                          BASED ON DIFFERENT DATA BASES
Data base
Liver tumors in female
mice (NCI 1976)

Liver tumors in male mice
(NCI 1976)
3»
I
rsi
Kidney tumors in male rats
(NCI 1976)

Kidney tumors in male mice
(Roe et al. 1979)
Multistage
qj = 1.35 x
q2 = 0
(q* = 1.7 x
qi = o
q2 = 1.34 x
(q* = 3.0 x
qi = 0
q2 = 7.07 x
(q* = 1.3 x
Probit
10'1 A = -2.03
B = 1.17
ID'1)3
9 A = -7.15
10"^ B = 3.51
ID'2)
. A = -4.58
ID'4 B = 1.31
10-2)
q1 = 5.5 x 10"2 NA
(q* = 9. 1 x 10-2)
Weibull One-hit
b = 1.68 x 10"1 b = 1.35 x 10'1
k = 0.92

b = 7.95 x 10~4 b = 1.21 x 10'1
k = 3.25

b = 2.08 x 10'4 1.56 x 10"2
k = 2.45

NA 5. 5 x 10~2
a q* is the 95% upper-bound confidence limit of the linear parameter in the multistage model,

NA = not applicable.  The models are not  applicable since  there is  only one dose group.

-------