BROMOFORM

BROMOFORM

BROMOFORM

CAS Number: 75-25-2

Synonyms: Methyl tribromide; Tribromomethane; Methenyl tribromide

Molecular Formula: CHBr3

TLV–TWA, 0.5 ppm (5.2 mg/m3)

A3 — Confirmed Animal Carcinogen with Unknown Relevance to Humans

 

TLV® Recommendation

A TLV–TWA of 0.5 ppm (5.2 mg/m3) is recommended for occupational exposure to bromoform to minimize the potential for possible liver and kidney damage reported in rodent studies (Bowman et al., 1978; Chu et al., 1980; Condie et al., 1983; Kroll et al., 1994a, b; Aida et al., 1992), as well as for upper respiratory tract irritation and lacrimation reported in humans (von Oettingen, 1955). The dose of bromoform from exposure to the recommended TLV–TWA value is well below the lowest-observed-adverse-effect level (LOAEL) for liver effects in orally-exposed rats (56 mg/kg) (Aida et al., 1992). An A3, Confirmed Animal Carcinogen with Unknown Relevance to Humans, notation is designated based on increased incidences of adenomatous polyps and adenocarcinomas of the large intestine of rats treated with bromoform (NTP, 1988).

Sufficient data were not available to recommend a Skin or sensitization (SEN) notation or a TLV–STEL. Although bromoform is absorbed through the skin (Xu et al., 2002), the low systemic toxicity and lethality of bromoform (NLM, 2007; Chu et al., 1982a) suggests that dermal exposure is unlikely to lead to toxic effects. The reader is expected to be familiar with the section on Excursion Limits in the Introduction to the Chemical Substance TLVs of the current edition of the Documentation of the TLVs® and BEIs® for the guidance and control of excursions above the TLV–TWA, even when the eight-hour TWA is within the recommended limit.

TLV® Basis

Liver damage; upper respiratory tract irritation; eye irritation.

Chemical and Physical Properties

Bromoform is a colorless, non-flammable, heavy liquid, similar to chloroform in odor and taste. Chemical and physical properties include (NLM, 2007):

Molecular weight: 252.73
Specific gravity: 2.89 at 20°C
Freezing point: 8°C
Boiling point: 149.1°C
Vapor pressure: 5 torr at 20°C
Vapor density: 8.7
Percent in saturated air: 0.7 at 25°C
Solubility: 3.1 g/l water at 25°C; miscible with organic solvents
Reactivity: liquid bromoform will attack some forms of plastics, rubber, and coatings
Conversion factors at 25°C and 760 torr: 1 ppm = 10.35 mg/m3; 1 mg/m3 = 0.097 ppm

Major Uses

Bromoform has been used as a chemical intermediate; in the synthesis of pharmaceuticals; and as a solvent for waxes, greases, and oils. It has also been used to make high-density liquids for geologic analysis and in the shipbuilding, aircraft, and aerospace industries (NLM, 2007).

Animal Studies

Acute

Dogs exposed to 7000 ppm bromoform or more became deeply anesthetized after eight minutes and died after one hour (Irish, 1963). The lowest lethal four-hour inhalation concentration of bromoform was reported to be 4500 mg/m3 for rats (Izmerov et al., 1982). Oral LD50 values for bromoform were greater than 1 g/kg in rats and mice (NLM, 2007; Chu et al., 1982a). Liver and kidney damage were observed in rats and mice (Bowman et al., 1978; Chu et al., 1980; Condie et al., 1983). The subcutaneous LD50 was 1820 mg/kg in mice (Kutob and Plaa, 1962). Narcosis and hepatotoxic effects were observed, and bromoform was rated as 38 times less hepato-toxic than tetrabromomethane (carbon tetrabromide). A single intraperitoneal dose of 758 mg/kg produced kidney dysfunction in rats (Kroll et al., 1994a, b). Bromoform was a less potent renal toxicant than the other trihalomethanes investigated (dichlorobromomethane, chloroform, and dibromochloromethane). Unlike other halomethanes, the hepatotoxicity of bromoform was not potentiated by prior treatment with chlordecone (Agarwal and Mehendale, 1983).

Undiluted bromoform was moderately irritating to rabbit eyes, and healing was complete in one to two days. Repeated skin contact caused moderate irritation to rabbit skin (Torkelson and Rowe, 1981).

Subchronic

Groups of 10 male rats were given 0, 5, 50, or 500 ppm bromoform in their drinking water for 28 days. Growth rate and food intake were unaffected and a slight increase in kidney weight was observed in the 500 ppm group. No histopathological changes were seen in the tissues examined (Chu et al., 1982a). Groups of Wistar rats (seven males and seven females) were given a diet containing micro-capsules of bromoform at 0.07, 0.2, and 0.6% of the diet for one month. Suppression of body weight gain was seen in the high-dose males. Vacuolization and swelling of the liver were noted in all groups given bromoform. Renal lesions were not observed. The LOAEL for bromoform was 56.4 mg/kg (Aida et al., 1992).

Groups of 20 male and female rats were given 0, 5, 50, 500, or 2500 ppm bromoform in their drinking water for 90 days (equivalent to approx 0, 0.3, 3, 30, and 150 mg/m3). Ten rats from each group were killed at 90 days and the remaining animals were given tap water for an additional 90 days. Mild histological changes in livers and thyroids were observed at 90 days in all treatment groups but not in the 90-day recovery groups (Chu et al., 1982b). The reversibility of these effects and the types of changes seen suggest that they are adaptive changes rather than adverse effects. Statistical analysis of the data was not presented. Male ICR mice given various concentrations of bromoform in drinking water for up to 90 days did not exhibit significant behavioral toxicity (Balster and Borzelleca, 1982).

Chronic/Carcinogenicity

Groups of 50 F-344/N rats of each sex and 50 female B6C3F1 mice were administered 0, 100, or 200 mg/kg bromoform by corn oil gavage five days per week for 103 weeks. Male B6C3F1 mice were administered 0, 50, or 100 mg/kg bromoform on the same schedule. Reduced survival of male rats in the high dose group (200 mg/kg) was noted. There was some evidence of carcinogenicity of bromoform in male rats and clear evidence in female rats based on increased incidences of adenomatous polyps and adenocarcinomas of the large intestine. There was no evidence for carcinogenic activity in mice (NTP, 1988). Administration of bromoform in drinking water (1.1 g/l) to male F344/N rats and B6C3F1 mice for 13 weeks induced aberrant crypt foci in the colons of rats but not mice (DeAngelo et al., 2002). Aberrant crypt foci are thought to be preneoplastic lesions involved in colon cancer. Feeding rats a folate-deficient diet for 26 weeks increased the incidence of aberrant crypt foci due to bromoform (0.5 g/l in drinking water) (Geter et al., 2005). Repeated intraperitoneal administration of bromoform produced pulmonary adenomas in strain A mice (Theiss et al., 1977).

Genotoxicity

Genotoxicity testing with bromoform has given mixed results. Bromoform has been reported to be positive, inconclusive, and negative in the Salmonella mutagenicity assay (Torkelson and Rowe, 1981; Haworth et al., 1983; Zeiger, 1990; Roldan-Arjona and Pueyo, 1993; Le Curieux et al., 1995; DiMarini et al., 1997; Landi et al., 1999; Kargalioglu et al., 2002). Bromoform was consistently mutagenic in Salmonella strain RSJ100, which contains the rat glutathione S-transferase theta-1 gene, suggesting that this form of glutathione S-transferase may be involved in bromoform bioactivation (DiMarini et al., 1997; Landi et al., 1999). However, there was no difference in the weak induction of DNA damage by bromoform in human whole blood cultures from individuals expressing glutathione S-transferase theta versus individuals lacking glutathione S-transferase theta as assessed by the Comet assay (Landi et al., 1999). Bromoform induced DNA damage in E. coli PO37 with or without a metabolic activation system and was clastogenic in the newt micronucleus assay (Le Curieux et al., 1995). Bromoform was positive in Drosophila sex-linked recessive lethal assays but negative in Drosophila reciprocal translocation assays (Woodruff et al., 1985).

Bromoform was positive in the mouse lymphoma cell mutation assay (Myhr et al., 1990) and produced sister-chromatid exchanges in Chinese hamster ovary cells (Galloway et al., 1985; Anderson et al., 1990), rat erythroblastic leukemia cells (Fujie et al., 1993), and human lymphocytes (Morimoto and Koizumi, 1983; Banerji and Fernandes, 1996). Administration of up to 253 mg/kg bromoform to rats intraperitoneally or orally produced chromosomal aberrations in bone marrow (Fujie et al., 1990); conflicting results were obtained with mice (Morimoto and Koizumi, 1983; Stocker et al., 1997). Oral gavage of male F-344 rats with up to 380 mg/kg bromoform for seven days did not produce DNA strand breaks (Potter et al., 1996).

Reproductive/Developmental Toxicity

No significant alterations in the number of resorption sites, fetuses per litter, fetal body weights, fetal malformations, or visceral anomalies were observed in the offspring of rats administered up to 200 mg/kg/day bromoform in corn oil by oral gavage on gestational days 6–15 (Ruddick et al., 1983).

Absorption, Distribution, Metabolism, and Excretion

Bromoform is metabolized by cytochromes P450 in vitro (Wolf et al., 1977; Ahmed et al., 1977) and in vivo (Anders et al., 1978) to ultimately form carbon monoxide. Free radicals were detected during incubation of bromoform with rat hepatocytes (Tomasi et al., 1985).

Detailed studies of the biotransformation of bromoform have not been reported. Oral administration of 14C-bromoform to Sprague-Dawley rats (100 mg/kg) and B6C3F1 mice (150 mg/kg) resulted in 60–90% absorption, wide organ distribution, and excretion primarily by exhalation of the parent compound or CO2 (Mink et al., 1986). Forty percent of the bromoform dose was converted to CO2 in mice while only 4% was converted in rats. A nonlinear relationship between oral dose (63 and 126 mg/kg) and area-under-the-curve (AUC) in rat blood suggested saturation of metabolic processes in this dose range (da Silva et al., 1999). Co-administration of bromoform and other trihalomethanes and related water disinfection by-products increased the AUC of bromoform, indicating inhibition of bromoform metabolism by these compounds (da Silva et al., 1999, 2000; St.-Pierre et al., 2003).

Human Studies

Exposure to bromoform vapor was reported to cause irritation of the respiratory tract, pharynx, and larynx, as well as lacrimation and salivation. Accidental ingestion of the liquid has produced central nervous system depression with coma and loss of reflexes. Smaller doses have produced listlessness, headache, and vertigo (von Oettingen, 1955). Quantitative dosimetry information was not available.

Bromoform is absorbed through human breast skin in vitro with a dermal absorption coefficient of 0.21 cm/hour at 25°C (Xu et al., 2002).

TLV® Chronology

1965: proposed: TLV–TWA, 5 ppm; Skin
1966: proposed: TLV–TWA, 0.5 ppm; Skin
1967–1996: TLV–TWA, 0.5 ppm; Skin
1995: proposed: A3, Confirmed Animal Carcinogen with Unknown Relevance to Humans
1996–2008: TLV–TWA, 0.5 ppm; Skin; A3
2008: proposed: TLV–TWA, 0.5 ppm; A3
2009: Adopted: TLV–TWA, 0.5 ppm; A3 Literature search current through mid-2008

References

Agarwal AK; Mehendale HM: Absence of Potentiation of Bromoform Hepatotoxicity and Lethality by Chlordecone. Toxicol Lett 15:251–257 (1983).
Ahmed AE; Kubic VL; Anders MW: Metabolism of Haloforms to Carbon Monoxide: I. In vitro Studies. Drug Metab Dispos 5:198–204 (1977).
Aida Y; Takada K; Uchida O; et al.: Toxicities of Microen-capsulated Tribromomethane, Dibromochloromethane and Bromodichloromethane Administered in the Diet to Wistar Rats for One Month. J Toxicol Sci 17:119–133 (1992).
Anders MW; Stevens JL; Sprague RW; et al.: Metabolism of Haloforms to Carbon Monoxide: II. In vivo Studies. Drug Metab Dispos 6:556–560 (1978).
Anderson BE; Zeiger E; Shelby MD; et al.: Chromosome Aberration and Sister-Chromatid Exchange Test Results with 42 Chemicals. Environ Mol Mutagen 16(Suppl 18):55–137 (1990).
Balster RL; Borzelleca JF: Behavioral Toxicity of Trihalo-methane Contaminants of Drinking Water in Mice. Environ Health Perspect 46:127–136 (1982).
Banerji AP; Fernandes AO: Field Bean Protease Inhibitor Mitigates the Sister-Chromatid Exchanges Induced by Bromoform and Depresses the Spontaneous Sister-chromatid Exchange Frequency of Human Lymphocytes in vitro. Mutat Res 360:29–35 (1996).
Bowman F; Borzelleca JF; Munson AE: The Toxicity of Some Halomethanes in Mice. Toxicol Appl Pharmacol 44:213–216 (1978).
Chu I; Secours V; Marino I; et al.: The Acute Toxicity of Four Trihalomethanes in Male and Female Rats. Toxicol Appl Pharmacol 52:351–353 (1980).
Chu I; Villeneuve DC; Secours VE; et al.: Toxicity of Trihalomethanes: I. The Acute and Subacute Toxicity of Chloroform, Bromodichloromethane, chlorodibromomethane, and bromoform in Rats. J Environ Sci Health B 17:205–224 (1982a).
Chu I; Villeneuve DC; Secours VE; et al.: Trihalome-thanes: II. Reversibility of Toxicological Changes Produced by Chloroform, Bromodichloromethane, chlorodibromomethane, and bromoform in Rats. J Environ Sci Health B 17:225–240 (1982b).
Condie LW; Smallwood CL; Laurie RD: Comparative Renal and Hepatotoxicity of Halomethanes: Bromo-dichloromethane, Bromoform, Chloroform, Dibromo-chloromethane, and Methylene Chloride. Drug Chem Toxicol 6:563–578 (1983).
da Silva LM; Charest-Tardif G; Krishnan K; et al.: Influence of Oral Administration of a Quaternary Mixture of Trihalomethanes on Their Blood Kinetics in the Rat. Toxicol Lett 106:49–57 (1999).
da Silva LM; Charest-Tardif G; Krishnan K; et al.: Evaluation of the Pharmacokinetic Interactions between Orally Administered Trihalomethanes in the Rat. J Toxicol Environ Health A 60:343–353 (2000).
DeAngelo AB; Geter DR; Rosenberg DW; et al.: The Induction of Aberrant Crypt Foci (ACF) in the Colons of Rats by Trihalomethanes Administered in the Drinking Water. Cancer Lett 187:25–31 (2002).
DiMarini DM; Shelton ML; Warren SH; et al.: Glutathione S-Transferase-mediated Induction of GC → AT Transitions by Halomethanes in Salmonella. Environ Mol Mutagen 30:440–447 (1997).
Fujie K; Aoki T; Wada M: Acute and Subacute Cytogenetic Effects of the Trihalomethanes on Rat Bone Marrow Cells in vivo. Mutat Res 242:111–119 (1990).
Fujie K; Aoki T; Ito Y; et al.: Sister-Chromatid Exchanges Induced by Trihalomethanes in Rat Erythroblastic cells and Their Suppression by Crude Catechin Extracted from Green Tea. Mutat Res 300:241–246 (1993).
Galloway S; Bloom A; Resnick M; et al.: Development of a Standard Protocol for in vitro Cytogenetic Testing with CHO cells: Comparison of Results for 22 Compounds in Two Laboratories. Environ Mutagen 7:1–52 (1985).
Geter DR; Moore TM; George MH; et al.: Tribromome-thane Exposure and Dietary Folate Deficiency in the Formation of Aberrant Crypt Foci in the Colons of F-344/N Rats. Food Chem Toxicol 43:1405–1412 (2005).
Haworth S; Lawlor T; Mortelmans K; et al.: Salmonella Mutagenicity Test Results for 250 Chemicals. Environ Mutagen 5(Suppl 1):3–142 (1983).
Irish DD: Aliphatic Halogenated Hydrocarbons. In: Indus-trial Hygiene and Toxicology, 2nd Rev ed, Vol II, pp 1262–1263. FA Patty (Ed). Interscience Publishers, New York (1963).
Izmerov NF; et al.: Toxicometric Parameters of Industrial Chemicals. Centre of International Projects, GKNT, Moscow (1982).
Kargalioglu Y; McMillan BJ; Minear RA; et al.: Analysis of the Cytotoxicity and Mutagenicity of Drinking Water Disinfection By-products in Salmonella typhimurium. Teratogen Carcinogen Mutagen 22:113–128 (2002).
Kroll RB; Robinson GD; Chung JH: Characterization of Trihalomethane (THM)-induced Renal Dysfunction in the Rat. I. Effects of THM on Glomerular Filtration and Renal Concentrating Ability. Arch Environ Contam Toxicol 27:1–4 (1994a).
Kroll RB; Robinson GD; Chung JH: Characterization of Trihalomethane (THM)-induced Renal Dysfunction in the Rat. II. Relative Potency of THMs in Promoting Renal Dysfunction. Arch Environ Contam Toxicol 27:5–7 (1994b).
Kutob SD; Plaa GL: A Procedure for Estimating the Hepatotoxic Potential of Certain Industrial Solvents. Toxicol Appl Pharmacol 4:354–361 (1962).
Landi S; Hanley NM; Warren SH; et al.: Induction of Genetic Damage in Human Lymphocytes and Mutations in Salmonella by Trihalomethanes: Role of Red Blood Cells and GSTT1-1 Polymorphism. Mutagenesis 14:479–482 (1999).
Le Curieux F; Gauthier L; Erb F; et al.: Use of the SOS Chromotest, the Ames-fluctuation Test, and the Newt Micronucleus Test to Study the Genotoxicity of Four Trihalomethanes. Mutagenesis 10:333–341 (1995).
Mink FL; Brown TJ; Rickabaugh J: Absorption, Distribution, and Excretion of 14C-Trihalomethanes in Mice and Rats. Bull Environ Contam Toxicol 37:752–758 (1986).
Morimoto K; Koizumi A: Trihalomethanes Induce Sister-Chromatid Exchanges in Human Lymphocytes in vitro and Mouse Bone Marrow Cells in vivo. Environ Res 32:72–79 (1983).
Myhr BC; McGregor D; Bowers L; et al.: L5178 Mouse Lymphoma Cell Mutation Assay Results with 41 Compounds. Environ Mol Mutagen 16(Suppl 18):138–167 (1990).
Potter CL; Chang LW; DeAngelo AB; et al.: Effects of Four Trihalomethanes on DNA Strand Breaks, Renal Hyaline
Droplet Formation and Serum Testosterone in Male F-344 Rats. Cancer Lett 106:235–242 (1996).
Roldan-Arjona T; Pueyo C: Mutagenic and Lethal Effects of Halogenated Methanes in the Ara Test of Salmonella typhimurium: Quantitative Relationship with Chemical Reactivity. Mutagenesis 8:127–131 (1993).
Ruddick JA; Villeneuve DC; Chu I; et al.: A Teratological Assessment of Four Trihalomethanes in the rat. J Environ Sci Health B 18:333–349 (1983).
Stocker KJ; Statham J; Howard WR; et al.: Assessment of the Potential in vivo Genotoxicity of Three Trihalome-thanes: Chlorodibromomethane, bromodichloromethane, and bromoform. Mutagenesis 12:169–173 (1997).
St.-Pierre A; Krishnan K; Tardif R: Evaluation of the Influence of Chloroacetic Acids on the Pharmacokinetics of Trihalomethanes in the Rat. J Toxicol Environ Health A 66:2267–2280 (2003).
Theiss JC; Stoner GD; Shimkin MB; et al.: Test for Carci-nogenicity of Organic Contaminants of United States Drinking Waters by Pulmonary Tumor Response in Strain A Mice. Cancer Res 37:2717–2720 (1977).
Tomasi A; Albano E; Biasi F; et al.: Activation of Chloro-form and Related Trihalomethanes to Free Radical Intermediates in Isolated Hepatocytes and in the Rat in vivo as Detected by the ESR-spin Trapping Technique. Chem Biol Interact 55:303–316 (1985).
Torkelson TR; Rowe VK: Halogenated Aliphatic Hydrocarbons. In: Patty’s Industrial Hygiene and Toxicology, 3rd Rev ed, Vol 2B, Toxicology, pp 3469–3470. John Wiley & Sons, New York (1981).
US National Library of Medicine (NLM): Bromoform. In: Hazardous Substances Data Bank. Toxicology Data Network (TOXNET). Online at: http://toxnet.nlm.nih.gov/ (2007).
US National Toxicology Program (NTP): Toxicology and Carcinogenesis Studies of Tribromomethane (Bromo-form) (CAS No 75-25-2) in F-344/N Rats and B6C3F1 Mice (Gavage Studies). Technical Report Series No 350. DHHS (NIH) Pub No 88-2805. NTP, Research Triangle Park, NC (1988).
von Oettingen WF: The Halogenated Aliphatic, Olefinic, Cyclic, Aromatic, and Aliphatic-Aromatic Hydrocarbons, Including the Halogenated Insecticides, Their Toxicity, and Potential Dangers, pp 65–67. USPHS Pub No 414. US Government Printing Office, Washington, DC (1955).
Wolf CR; Mansuy D; Nastainczyk W; et al.: The Reduction of Polyhalogenated Methanes by Liver Microsomal Cytochrome P450. Mol Pharmacol 13:698–705 (1977).
Woodruff RC; Mason JM; Valencia R; et al.: Chemical Mutagenesis Testing in Drosophila. V. Results of 53 Coded Compounds Tested for the National Toxicology Program. Environ Mutagen 7:677–702 (1985).
Xu X; Mariano TM; Laskin JD; et al.: Percutaneous Absorption of Trihalomethanes, Haloacetic Acids, and Haloketones. Toxicol Appl Pharmacol 184:19–26 (2002).
Zeiger E: Mutagenicity of 42 Chemicals in Salmonella. Environ Mol Mutagen 16(Suppl 18):32–54 (1990).

 

You cannot copy content of this page