Fact-checked by Grok 2 weeks ago

Bromoform

Bromoform (CHBr₃), systematically named tribromomethane, is a that exists as a colorless liquid at , exhibiting a sweet odor reminiscent of . With a of 2.89 g/cm³, it sinks in where it possesses limited (about 3.1 g/L at 25°C) and is nonflammable under standard conditions. Bromoform occurs naturally in marine environments, primarily produced by and macroalgae such as seaweeds, contributing to atmospheric loading as a volatile . Industrially, it serves as a , a in , and in laboratory applications like density gradient centrifugation for separating minerals or biological materials. It also forms as a disinfection during chlorination of bromide-containing , representing a key route of human exposure through . Despite historical use as a and expectorant in —replaced due to adverse effects—bromoform is acutely toxic via , , and dermal , targeting the liver, kidneys, and , with evidence of carcinogenicity in animal studies. Regulatory bodies monitor its levels in water supplies owing to potential health risks from chronic low-level exposure.

History

Discovery and development

Bromoform, or tribromomethane (CHBr₃), was first synthesized in by German chemist Carl Jacob Löwig, who obtained it by distilling bromal (tribromoacetaldehyde) with , in direct analogy to the contemporaneous preparation of from . This method highlighted bromoform's status as a haloform, a class of compounds featuring a tri-substituted with , and established its chemical kinship to , which had been isolated just a year prior. Löwig's work built on his earlier discovery of in 1825, underscoring the rapid exploration of bromine-based organics following the element's isolation. During the mid-19th century, bromoform's properties were further characterized, revealing its colorless, heavy liquid state with a chloroform-like odor and sweet taste, alongside reactivity patterns typical of haloforms, such as susceptibility to hydrolysis under alkaline conditions. By the late 19th century, its sedative potential was noted, paralleling chloroform's anesthetic applications; it was administered in small doses (1–2 drops, approximately 15–20 mg/kg) to children for whooping cough relief, though its use remained limited compared to chloroform due to observed higher toxicity, including reports of fatalities from overdosing. These early pharmacological trials, spanning into the early 20th century, ceased as safer alternatives emerged, shifting focus to bromoform's role as a chemical intermediate rather than a therapeutic agent. Key advancements in the mid-to-late involved refined analytical techniques like , which enabled detection of trace bromoform in environmental samples, uncovering its production by marine macroalgae and planktonic organisms through bromoperoxidase-mediated of organic precursors. This revelation, building on studies of volatile halocarbons, marked a milestone in recognizing bromoform not solely as a synthetic compound but as a biogeochemical player, with atmospheric implications traced to oceanic emissions. Such insights paralleled broader understandings of organohalogen cycles, distinct from sources.

Early industrial applications

In the late 19th and early 20th centuries, bromoform was employed as a for extracting waxes, greases, and oils, leveraging its high and non-polar properties to facilitate separations in industrial and settings. Its use extended to geological applications, where it served as a heavy medium for -based separation of minerals and ores, allowing differentiation based on specific differences in assays. Medically, bromoform saw limited adoption around the early 1900s as a to alleviate coughing in children with pertussis (), administered in small doses for its calming effects similar to but with observed greater risks of and overdose fatalities. Empirical reports documented multiple child deaths from accidental overdoses, highlighting its narrower therapeutic window and higher acute toxicity profile compared to , which curbed broader clinical uptake despite initial interest as an anesthetic alternative. By the mid-20th century, particularly post-1940s, bromoform's industrial roles diminished as safer, less toxic solvents and anesthetics emerged, confining it to niche reagent applications such as extraction solvents and geological testing where its density remained advantageous. This shift reflected accumulating evidence of its hazards, including liver damage, prompting replacement in broader commercial contexts.

Properties

Physical properties

Bromoform is a colorless to pale yellow liquid at , possessing a sweet similar to that of . It is denser than and nonflammable under standard conditions, exhibiting stability without decomposition at ambient temperatures and pressures. Key physical constants of bromoform include the following:
PropertyValueConditions
Density2.89 g/cm³25 °C
149–151 °C760 mmHg
8 °C-
1.60020 °C, n_D
5 mmHg20 °C
Water solubility0.1–3.1 g/L20–25 °C
Bromoform is sparingly soluble in water but fully miscible with common organic solvents such as ethanol, ether, and chloroform. Its relative vapor density of 8.7 indicates that vapors are heavier than air. For identification, bromoform displays characteristic spectroscopic features: in ¹H NMR (CDCl₃), a singlet at δ ≈7.4 ppm for the single proton; in IR, C-H stretch near 3000 cm⁻¹ and C-Br modes below 700 cm⁻¹.

Chemical structure and reactivity

Bromoform possesses the molecular formula CHBr₃, featuring a central carbon atom tetrahedrally coordinated to one and three atoms, resulting in C3v symmetry. The C-Br bond lengths measure approximately 1.93 , while the C-H bond length is about 1.07 , reflecting the sp³ hybridization of the carbon center and the larger of compared to lighter . The trihalide arrangement imparts distinctive reactivity, primarily through nucleophilic attack at the carbon due to the electron-withdrawing effects of the atoms, which polarize the C-Br bonds and stabilize departing ions. Bromoform undergoes base-catalyzed faster than under identical aqueous conditions, twice the rate, via a involving to form the trichlorobromomethyl anion , followed by stepwise elimination of to yield a dihalocarbene that reacts with to produce and ions. This process highlights the causal role of the acidic C-H proton (a ≈ 15-16) in initiating , contrasting with monohaloalkanes where direct SN2 displacement predominates. Relative to other trihalomethanes, bromoform exhibits intermediate stability: more resistant to thermal decomposition than iodoform, which readily releases iodine upon heating, but less inert than chloroform owing to weaker C-Br bonds (bond dissociation energy ≈ 285 kJ/mol versus 340 kJ/mol for C-Cl). Photochemically, exposure to natural sunlight induces decomposition at a rate of 0.21 h-1 at 30°C, generating bromine radicals through homolytic C-Br cleavage, which can propagate radical chain reactions in environmental contexts.

Synthesis

Laboratory methods

Bromoform is prepared in the laboratory primarily through the , which involves the successive alpha-bromination of acetone followed by cleavage of the resulting trihaloacetone under basic conditions. In this process, acetone reacts with in the presence of a base such as or to yield bromoform and . A representative procedure entails dissolving 30 mL of acetone in 150 mL of 20% aqueous solution at 50°C, followed by the slow addition of 75 mL of via a , with subsequent addition of 800 mL of 10% to neutralize excess . The reaction mixture is then steam-distilled to isolate the product, producing 100-110 g of crude bromoform, corresponding to yields of approximately 70-80% based on acetone consumption. The heavy organic layer of bromoform is separated from the aqueous phase, washed with to remove salts, dried over , and purified by under reduced pressure to avoid , collecting the fraction boiling at 149°C at . To optimize yields, an excess of (typically 3 equivalents) and is employed to drive complete tribromination, while controlling the below 60°C minimizes side reactions such as over-bromination or acetone . Variations using , generated from and oxidizing agents like , offer a safer alternative to direct handling for smaller scales, though they may introduce impurities if is present. Alternative laboratory routes include the of dissolved in , where anodic oxidation generates that reacts to form bromoform. This method requires a suitable with electrodes and controlled to favor haloform formation over dibromomethane. Another approach involves treating with aluminum bromide, leveraging exchange to produce bromoform, though this is less common due to the hygroscopic nature of AlBr3. All syntheses necessitate stringent safety measures, as is corrosive and generates toxic fumes, while bromoform vapors cause respiratory irritation and . Reactions must be conducted in a well-ventilated with splash-proof , nitrile gloves resistant to , and a lab coat; neutralization of waste with is essential to quench residual before disposal. Empirical data from handling protocols emphasize monitoring for spills, as bromoform's (2.89 g/mL) causes it to sink in , complicating cleanup.

Industrial and commercial production

Bromoform is produced commercially through the , in which acetone is treated with and a base such as , generating bromoform alongside and byproducts. Alternative industrial routes include the reaction of with aluminum tribromide or of in , though these are less commonly scaled for production. These methods leverage sourced from brines or oceanic bitterns, but the processes are optimized for batch rather than continuous flow due to bromoform's niche demand and handling requirements for the volatile, dense liquid. Annual production volumes in the United States have remained low, with manufacturers reporting 10,000 to 500,000 pounds (4.5 to 227 metric tons) under the Chemical Inventory Update Rule for , , and ; earlier data indicate less than 500 metric tons in 1975 and 50 to 500 metric tons in 1977. In the , registration under REACH places combined and import at 100 to 1,000 tonnes per annum as of recent assessments. Producers are primarily specialty firms such as and Geoliquids, with no evidence of large-scale commodity operations akin to facilities. Post-1970s environmental regulations, including those targeting trihalomethanes under the and broader restrictions on volatile halogenated organics, curtailed potential growth in output by limiting solvent and reagent applications, confining production to controlled specialty . Economic constraints arise from bromine's higher procurement costs—derived from limited global reserves—and purification demands to meet purity standards exceeding 99% for industrial grades, resulting in bromoform commanding premium pricing that sustains only modest volumes for targeted sectors.

Natural occurrence

Marine biosynthesis

Bromoform is biosynthesized in marine environments primarily by macroalgae and through enzymatic processes. These organisms employ vanadium-dependent haloperoxidases, particularly bromoperoxidases (BPOs), which catalyze the bromination of organic precursors such as fatty acids or short-chain hydrocarbons in the presence of bromide ions and . This pathway generates bromoform as a volatile , facilitating its release into and subsequent evasion to the atmosphere. Macroalgae, including like Corallina pilulifera, exhibit regulated BPO activity that correlates with bromoform production, often peaking under conditions of stress or seasonal changes. , such as diatoms, contribute in open ocean settings, with elevated concentrations linked to blooms where BPOs oxidize to (HOBr), which then reacts with organic substrates to form trihalomethanes like bromoform. Biochemical reconstitutions confirm that these enzymes obligately produce bromoform from marine-derived precursors, underscoring a dedicated biosynthetic route rather than incidental formation. Oceanic emissions of bromoform, estimated at approximately 214 Gg yr⁻¹ in recent coupled ocean-atmosphere models, represent the dominant flux, far exceeding anthropogenic inputs and driving bromine transfer in global cycles. As a key volatile organic (VOH), bromoform modulates tropospheric and stratospheric chemistry by serving as a bromine reservoir, with biosynthesis accounting for the majority of its environmental burden. These fluxes exhibit regional variability, with higher rates in productive coastal and temperate waters tied to algal activity.

Terrestrial and atmospheric sources

Trace amounts of bromoform are produced terrestrially by microorganisms, including fungi such as Curvularia species, which biosynthesize the compound through enzymatic pathways involving bromoperoxidases. These fungal isolates, derived from environments, have been cultured to yield bromoform, demonstrating a natural microbial capacity for its formation independent of marine influences. Additionally, bacterial communities possess genes for (de) processes that enable the natural production of bromoform, particularly in the presence of inorganic , as evidenced by detectable formation in enriched top layers. Atmospheric bromoform from terrestrial sources remains minimal, with concentrations typically below 3 parts per trillion (ppt) in background settings, reflecting limited direct emissions from land-based microbes relative to broader transport dynamics. Pre-industrial modeling indicates stable low-level atmospheric burdens under natural conditions, prior to significant anthropogenic inputs. Minor anthropogenic contributions arise from water chlorination processes, such as in industrial cooling systems, accounting for 12–28% of global emissions but overshadowed by natural dominance in overall atmospheric loading.

Applications

Laboratory and industrial uses

Bromoform serves as a dense medium in laboratory gradient for separating minerals, biological materials such as and dentine powders, and components, leveraging its of 2.89 g/cm³ to create gradients exceeding 2.8 g/cm³. In , it functions as a for extracting organic compounds from aqueous solutions and for applications in (NMR) spectroscopy and . Geological assays employ bromoform as a heavy for mineral separation by , enabling precise based on specific differences. It is also used as a certified reference standard in pharmaceutical and testing, including evaluations. Historically, bromoform found industrial application as an ingredient in fire-resistant chemicals and gauges for measurements, though such uses have largely been phased out in favor of limited production for reagents and specialized testing as of the early .

Agricultural and environmental applications

Bromoform-based feed additives have been developed in the 2020s to inhibit in , targeting enteric —a key agricultural source of gases. Synthetic formulations, such as Rumin8's investigational veterinary product containing bromoform in oil, reduced yield by 94-95% in during controlled trials at UC Davis in early 2025, with no significant effects on intake, average daily gain, or parameters. A of multiple studies reported average yield reductions of 43.3% across and at typical doses, with efficacy varying by animal type (higher in ) and composition, though higher doses amplified mitigation up to near-elimination levels without yield losses. These additives align with climate mitigation strategies by enabling producers to lower herd-level emissions, as evidenced by field trials integrating bromoform into systems yielding 37.7% average reductions in daily output. Regulatory advancements include product licensing in and provisional approval in by mid-2024, positioning bromoform as a viable tool for meeting national GHG targets in ruminant-heavy agricultural sectors. Environmental applications leverage bromoform's natural biosynthesis in marine ecosystems, where macroalgae like produce it as a halogenated compound, contributing to oceanic emissions that dominate global fluxes over inputs. This biogeochemical context tempers emphasis on unverified risks from low-dose agricultural use, as natural marine sources sustain ambient atmospheric levels without documented ecosystem collapse, prioritizing empirical efficacy data over precautionary biases in .

Toxicology and human health effects

Acute and systemic toxicity

Bromoform exhibits moderate via oral and routes in animal models, with oral LD50 values in rats ranging from 933 to 1,550 mg/kg, depending on strain and sex. LC50 values for rats over 4 hours approximate 3.1 mg/L, associated with (CNS) depression, dyspnea, and organ damage. Acute exposure in rodents primarily induces CNS effects such as narcosis and , alongside hepatic and renal lesions including fatty degeneration and , observed at doses exceeding 100 mg/kg. These outcomes stem from bromoform's metabolism by hepatic enzymes, particularly , yielding reactive intermediates like dihalomethyl radicals that bind to cellular macromolecules, precipitating and tissue injury. In humans, acute inhalation of bromoform vapors causes irritation to the , eyes, and mucous membranes, manifesting as lacrimation, salivation, and facial at concentrations above 850 , which represents the immediately dangerous to life or (IDLH) threshold derived from animal data. Empirical case reports from accidental oral exposures, often in children dosed historically as a (10–70 mg/kg), document CNS depression including drowsiness, convulsions, and , with hepatic and renal impairment in severe instances; fatalities occurred via , but survivors at lower doses recovered without sequelae following supportive care. Bromoform's rapid absorption and partial biliary excretion underscore the reversibility of low-dose effects, as unbound ions and unmetabolized parent compound are eliminated renally within hours to days.

Carcinogenicity and long-term effects

The U.S. Environmental Protection Agency (EPA) classifies bromoform as a probable (Group B2), based primarily on increased incidences of large intestinal tumors observed in male and female F344/N rats administered bromoform by gavage at doses of 40–125 mg/kg/day for 2 years, including adenomatous polyps and adenocarcinomas. In contrast, the International Agency for Research on Cancer (IARC) classifies bromoform as Group 3 (not classifiable as to its carcinogenicity to humans), citing inadequate evidence in humans and limited evidence in experimental animals, with the findings deemed insufficient for stronger categorization due to uncertainties in relevance. Animal carcinogenicity appears linked to non-genotoxic mechanisms involving and regenerative cell proliferation rather than direct DNA damage. National Toxicology Program (NTP) gavage studies in rats showed dose-dependent forestomach and intestinal epithelial preceding tumor formation, consistent with tissue injury from high bolus exposures rather than mutagenic initiation; no significant tumor increases occurred in mice at similar doses. assays for bromoform are largely negative or equivocal, with no consistent mutagenicity in bacterial tests (e.g., Ames strains TA98, TA100), chromosomal aberration studies, or micronucleus assays, though some DNA damage reports exist under cytotoxic conditions. This supports a mode-of-action threshold below typical environmental exposures, challenging linear low-dose extrapolations used in some risk assessments. Epidemiological data reveal no clear causal link between bromoform and human cancer, particularly in cohorts exposed via disinfection byproducts where bromoform constitutes a minor fraction of total trihalomethanes. Meta-analyses of trihalomethane-exposed populations show inconsistent associations with bladder or , often confounded by co-exposures like or bromodichloromethane, with bromoform-specific risks unresolvable due to measurement limitations and lack of dose-response gradients at ambient levels (<10 μg/L). Prioritizing mechanistic and exposure data over precautionary models, the absence of genotoxic potency and reliance on high-dose rodent artifacts suggest minimal human relevance absent confirmatory human evidence.

Exposure assessment and regulations

Human exposure to bromoform primarily occurs through ingestion of chlorinated drinking water, where it forms as a disinfection byproduct, with estimated daily intakes for the general U.S. population ranging from 0.1 to 5 μg per person based on typical concentrations of 1–10 μg/L in treated water supplies. Inhalation represents a secondary route, particularly during showering or occupational settings involving volatile emissions, though dermal absorption is minimal due to low solubility. Natural sources, such as marine emissions into seawater or air, contribute negligibly to human exposure, as direct consumption of untreated ocean water or significant inhalation from oceanic backgrounds is not a routine pathway for populations. In the United States, bromoform is regulated under the EPA's National Primary Drinking Water Regulations as part of total trihalomethanes (TTHM), with a maximum contaminant level (MCL) of 80 μg/L for the sum of chloroform, bromodichloromethane, dibromochloromethane, and bromoform, reflecting a zero MCLG due to its probable carcinogenic classification; individual bromoform levels are monitored but not separately capped beyond TTHM compliance. Occupationally, OSHA sets a permissible exposure limit (PEL) of 0.5 ppm (5 mg/m³) as an 8-hour time-weighted average with skin notation, aligned with NIOSH recommendations to prevent acute irritation and systemic effects. Bromoform is listed on the inventory as an existing chemical substance subject to risk management under the Toxic Substances Control Act. Globally, the World Health Organization provides a guideline value of 100 μg/L for bromoform in drinking water, derived from cancer potency estimates and assuming a 10⁻⁵ lifetime risk level, which exceeds some national thresholds but accounts for practical treatment feasibility. Regulatory approaches have faced scrutiny for imposing stringent limits on anthropogenic sources while overlooking ubiquitous natural production—estimated at thousands of tons annually from marine algae—potentially overemphasizing disinfection byproducts relative to total environmental flux, though human-relevant exposures remain dominated by water treatment rather than background levels. Despite its EPA probable carcinogen status and IARC Group 3 classification (not classifiable for human carcinogenicity due to equivocal data), bromoform-containing compounds like seaweed extracts (e.g., from ) have gained traction for regulatory approval as ruminant feed additives to suppress enteric methane emissions, with GRAS notices filed under low-dose scenarios (e.g., 0.2–0.5% of diet) prioritizing climate benefits over precautionary human residue concerns, as rumen metabolism minimizes transfer to milk or meat. This reflects a risk-benefit framework where veterinary applications proceed amid ongoing human safety evaluations, contrasting stricter potable water controls.

Environmental fate and impact

Persistence and bioaccumulation

Bromoform displays moderate persistence in surface waters, with overall degradation half-lives estimated at 1–6 months, influenced by volatilization, microbial biodegradation, and slow hydrolysis, though its pure hydrolytic half-life under neutral conditions exceeds 686 years, rendering hydrolysis negligible in environmental contexts. In anaerobic groundwater systems, half-lives shorten to 21–42 days due to reductive dehalogenation processes. Atmospheric persistence is shorter, with photooxidative degradation via hydroxyl radical reactions yielding half-lives of 20–30 days. Bioaccumulation of bromoform is limited, as evidenced by its octanol-water partition coefficient (log Kow) of approximately 2.4 and measured factors (BCF) in fish species ranging from 2 to 50. These values predict low steady-state accumulation in aquatic biota, with partition-based models confirming BCFs below 100 and negligible potential across trophic levels due to insufficient for efficient transfer. In soils, bromoform exhibits weak , characterized by organic carbon-normalized coefficients (Koc) of 50–300, which facilitate high mobility and into aquifers over retention on . Field and laboratory data indicate minimal adsorption to or minerals, with retardation factors near 1 in low-organic-carbon matrices, promoting advective transport in percolating water columns rather than diffusive binding.

Atmospheric and ozone effects

Bromoform (CHBr₃), a very short-lived substance, is emitted volatively into the primarily from marine sources such as and macroalgae, with secondary contributions from coastal and industrial activities. These emissions facilitate upward via deep , particularly in tropical regions, to the tropical layer (TTL) and subsequently into the lower . Modeling studies indicate that tropospheric losses through photolysis and reactions limit the efficiency of bromine release, with only approximately 10-20% of emitted bromine from bromoform and related compounds reaching the stratosphere as inorganic bromine (Bry). In the , photodecomposition of releases atoms that participate in catalytic cycles depleting , with each atom exhibiting roughly 60 times the ozone-destroying efficiency of . depletion potentials (ODPs) for bromoform vary seasonally and regionally, ranging from 0.10 to 0.72 for emissions in areas like the , reflecting differences in pathways and atmospheric processing. Global modeling incorporating observational constraints estimates bromoform's contribution to stratospheric loading at 0.5-1.6 , underscoring its role as a natural supplement to in chemistry. Empirical data from campaigns, observations, and ground-based measurements highlight the dominance of emissions in the atmospheric bromoform budget, with sources comprising 12-28% of global fluxes. This preponderance, validated by modeling of inventories, implies that incremental additions exert limited influence on total stratospheric and resultant loss, as natural variability in already drives substantial fluctuations. Debates persist over potential underestimations of discharges, yet integrated assessments affirm that bromoform's impact remains secondary to long-lived halocarbons regulated under the .

Water contamination and disinfection byproducts

Bromoform forms as a disinfection byproduct during when bromide ions, naturally present in source , react with disinfectants such as or in the presence of organic precursors like . This reaction is more pronounced in waters affected by intrusion, where concentrations increase, shifting speciation toward brominated forms including bromoform. or disinfection can also generate bromoform at levels of 1–10 µg/L in -containing waters, though chlorination remains the primary pathway in conventional . In treated , bromoform concentrations typically range from 0.1 to 12 µg/L, with most systems maintaining levels below 10 µg/L due to variability and treatment controls. Regulatory guidelines set maximums at 100 µg/L for bromoform individually (WHO) or within total trihalomethanes at 80 µg/L (EPA MCL), reflecting assessments where exceedances are rare in compliant systems. Naturally occurring bromoform in environments reaches 20–266 ng/L (0.02–0.266 µg/L) in coastal or polar seawaters and 0.03–0.15 ng/L in open oceans, indicating anthropogenic treatment contributions are higher but still low relative to total environmental exposure pathways. Health risk assessments position bromoform's contribution to overall disinfection byproduct exposure as minor, with lifetime cancer risks from typical levels falling below 10^{-6} in most models, far outweighed by chlorination's role in averting microbial outbreaks like , which historically caused millions of deaths annually before widespread adoption. Epidemiological data link mixed byproduct exposures to potential increments, but causal attribution to bromoform alone remains uncertain, as studies show species-specific differences limiting direct human extrapolation. Mitigation focuses on precursor removal via enhanced coagulation or adsorption to minimize before disinfection, alongside alternatives like chloramination, which reduces formation by 50–90% compared to free while preserving biocidal efficacy. or boiling can volatilize bromoform post-treatment, achieving up to 100% removal in household settings, though these do not address formation prevention and may concentrate non-volatile byproducts. Such strategies must balance against incomplete inactivation risks from reduced chlorine residuals, underscoring chlorination's net benefits in bromide-low source waters.

References

  1. [1]
    Bromoform | CHBr3 | CID 5558 - PubChem - NIH
    Denser than water (density: 2.9 g / cm3 ) and slightly soluble in water. Hence sinks in water. Nonflammable. Toxic by ingestion, inhalation and skin absorption.
  2. [2]
    [PDF] Bromoform | EPA
    Bromoform is a colorless to pale yellow liquid with a sweetish odor. (1,5). The chemical formula for bromoform is CBr. 3. H and the molecular weight is 252.75 g ...Missing: pubchem | Show results with:pubchem
  3. [3]
    BROMOFORM - CAMEO Chemicals - NOAA
    Denser than water (density: 2.9 g / cm3 ) and slightly soluble in water. Hence sinks in water. Nonflammable. Toxic by ingestion, inhalation and skin absorption.Missing: pubchem | Show results with:pubchem
  4. [4]
    Bromoform - an overview | ScienceDirect Topics
    Bromoform (CHBr3, also known as tribromomethane) is a short-lived trace gas, which represents a significant source of organic bromine to the atmosphere (Quack ...Missing: reliable | Show results with:reliable
  5. [5]
    Bromoform & Dibromochloromethane | Public Health Statement - CDC
    Bromoform and dibromochloromethane may form when chlorine reacts with other naturally occurring substances in water, such as decomposing plant material. Plants ...Missing: reliable | Show results with:reliable
  6. [6]
    [PDF] TOXICOLOGICAL PROFILE FOR BROMOFORM AND ...
    Therefore, the principal source of human exposure to bromoform and dibromochloromethane is chlorinated water supplied to homes, work, and public places.
  7. [7]
    [PDF] Tribromomethane (Bromoform) (CASRN 75-25-2) in F344/N Rats ...
    Tribromomethane (CAS 75-25-2), also known as bromoform, is a chemical intermediate and solvent. High doses caused death in rats and mice. 2-year studies were ...Missing: reliable | Show results with:reliable
  8. [8]
    BROMOFORM - ACGIH
    Bromoform (CHBr3) is a colorless, non-flammable liquid, similar to chloroform in odor and taste. It is a confirmed animal carcinogen.Missing: reliable | Show results with:reliable
  9. [9]
    [PDF] Trihalomethanes in Drinking Water: Chloroform, Bromoform - OEHHA
    Jun 10, 2019 · The toxicological profile is complete and reflects the somehow limited information about this chemical compound. The inclusion of a table ...<|separator|>
  10. [10]
    CHBr3 properties
    Bromoform was first prepared in 1832 by German chemist Löwig, who distilled a mixture of bromal (tribromoacetaldehyde) with potassium hydroxide. This ...
  11. [11]
    Toxicological Profile For Bromoform And Chlorodibromomethane
    Because of this property, bromoform was used as a sedative in the early part of this century. Doses of 1 to 2 drops (probably about 15 to 20 mg/kg) given 3 ...
  12. [12]
    Bromoform - an overview | ScienceDirect Topics
    Bromoform is used as a chemical intermediate in the synthesis of organic chemicals and pharmaceuticals. ... Bromoform was used in the late nineteenth and early ...Missing: first date
  13. [13]
    [PDF] TOXICOLOGICAL PROFILE FOR BROMOFORM AND ...
    Summaries of studies that provide reliable quantitative toxicity data are presented in Table 3-1 and Figure 3-1 for bromoform and in Table 3-2 and Figure 3-2 ...<|control11|><|separator|>
  14. [14]
    Air‐sea flux of bromoform: Controls, rates, and implications
    Penkett et al. [1985] identified bromoform as the greatest single source for atmospheric bromine. These atmosphere-based estimates generally rely on “background ...Missing: discovered | Show results with:discovered
  15. [15]
    Natural Formation of Chloroform and Brominated Trihalomethanes ...
    In situ enrichment of a soil top layer by KBr showed that soil has the potential to form chlorodibromomethane and bromoform naturally.
  16. [16]
    Bromoform in Mathura - Dealers, Manufacturers & Suppliers -Justdial
    By the late 19th century, bromoform began to be utilized in scientific research and industry. Its effectiveness as a solvent made it valuable in the extraction ...
  17. [17]
    PRODUCTION, IMPORT/EXPORT, USE, AND DISPOSAL - NCBI
    Bromoform may be prepared from acetone and sodium hypobromite, by treating chloroform with aluminum bromide, or by electrolysis of potassium bromide in ...
  18. [18]
    RELEVANCE TO PUBLIC HEALTH - Toxicological Profile for ... - NCBI
    At lower doses, the hepatotoxicity of bromoform and dibromochloromethane is characterized by fatty infiltration, cellular vacuolization and swelling, and ...
  19. [19]
    Effect of trihalomethanes (chloroform and bromoform) on human ...
    Feb 1, 2017 · The results indicated that the damage caused by bromoform was statistically more significant as compared to chloroform. Haemoglobin (HGB) ...
  20. [20]
    ICSC 0108 - BROMOFORM - INCHEM
    Solubility in water, g/100ml at 20°C: 0.1 (poor) Vapour pressure, kPa at 20°C: 0.67. Relative vapour density (air = 1): 8.7. Relative density of the vapour ...Missing: spectroscopic IR NMR
  21. [21]
    Bromoform - Optional[1H NMR] - Chemical Shifts - SpectraBase
    Bromoform · View entire compound with spectra: 21 NMR, 2 FTIR, and 6 MS (GC) · View the Full Spectrum for FREE! · View the Full Spectrum for FREE! · The full ...Missing: vapor pressure stability
  22. [22]
    Molecular Parameters - Carbon Compounds - Wired Chemist
    Molecular Parameters - Carbon Compounds ; CHBr · bromoform, C-H 1.068 ± 0.010. C-Br 1.930 ± 0.003, BrCBr 110.8 ± 0.3 ; CHCl · chloroform, C-H 1.073 C-Cl 1.767 ...Missing: tetrahedral | Show results with:tetrahedral
  23. [23]
    Kinetics of hydrolysis of chloroform and bromoform in aqueous ...
    Under identical conditions BF hydrolyzed twice as fast as CF. The mechanisms accounting for the base-catalyzed hydrolysis of haloforms are reviewed in short.Missing: decomposition | Show results with:decomposition
  24. [24]
    [PDF] Natural Sunlight Photodegradation of Halogenated Disinfection ...
    Bromoform, CHBr2Cl and. CHBrCl2 were found degraded at rates of 0.21, 0.156 and 0.098 h-1, respectively, under natural sunlight at 30 °C, whereas chloroform was ...
  25. [25]
    Decomposition of iodoform - Science made alive
    Iodoform is not a very stable compound. It decomposes on heating and gives iodine (purple vapor), most likely some hydrogen iodide (fuming at open end of test ...
  26. [26]
    Haloform Reaction - an overview | ScienceDirect Topics
    They are prepared by heating acetone with an alkaline hypohalite solution. Any methyl ketone yields a haloform under these conditions (see the haloform ...
  27. [27]
    Preparation of bromoform (tribromomethane; methane, tribromo-)
    The flask is immersed in water maintained at 50°. To the flask, 30 ml of acetone and 150 ml of 20% sodium carbonate solution are placed. 75 ml of bromine in the ...Missing: first | Show results with:first
  28. [28]
    "Large" Scale Bromoform Synthesis - Powered by XMB 1.9.11
    Aug 17, 2016 · The standard 'recipe' is: 1) Dissolve ~8g NaBr into 100mL distilled water. 2) Add 4mL of acetone to the solution. 3) Add 150mL household bleach.
  29. [29]
    NIOSH Pocket Guide to Chemical Hazards - Bromoform - CDC
    Skin:Prevent skin contact ; Eyes:Prevent eye contact ; Wash skin:When contaminated ; Remove:When wet or contaminated ; Change:No recommendation.
  30. [30]
    [PDF] Safety Data Sheet - Fisher Scientific
    Jul 5, 2015 · Precautions for safe handling: Avoid contact with skin, eyes, and clothing.Follow good hygiene procedures when handling chemical materials.
  31. [31]
    [PDF] TOXICOLOGICAL PROFILE FOR BROMOFORM AND ...
    Orrell and Mackie (1988) estimated that 6–9 kkg of bromoform were imported by Freeman Industries in the late 1980s. No current information was located on the ...Missing: discovered | Show results with:discovered
  32. [32]
    BROMOFORM | - atamankimya.com
    Bromoform is a brominated organic solvent with the formula CHBr3. Bromoform has an odor similar to chloroform and Bromoform density is very high (2,89).Missing: date | Show results with:date
  33. [33]
    Genetic and biochemical reconstitution of bromoform biosynthesis in ...
    To date, a hydrogen peroxide-producing enzyme has not been identified from seaweeds. This lack of insight also begets knowledge gaps in basic seaweed biology.
  34. [34]
    Bromoperoxidase and iodoperoxidase enzymes and production of ...
    Sep 15, 1996 · We describe laboratory experiments on the production of various brominated and iodinated compounds in cultures of marine diatoms, obtained from ...
  35. [35]
    Physiological function of bromoperoxidase in the red marine alga ...
    We demonstrated that bromoform produced by C. pilulifera played an important role in eliminating epiphytic organisms, especially microalgae on the surface. Such ...
  36. [36]
    Oceanic distribution and sources of bromoform and ... - AGU Journals
    Oct 6, 2007 · The tropical oceans are a source of reactive bromine to the atmosphere in the form of short-lived brominated methanes as bromoform (CHBr3) ...
  37. [37]
    Obligate Brominating Enzymes Underlie Bromoform Production by ...
    Feb 8, 2021 · Marine algae are prolific producers of bromoform (CHBr3). This naturally produced molecule is a potent environmental pollutant as it ...
  38. [38]
    Natural marine bromoform emissions in the fully coupled ocean ...
    Jun 21, 2024 · Moreover, the ratio of bromoform emissions (2.82) is higher than the ratio of bromoform production (2.38) and indicates an excess of 18 %.
  39. [39]
    Phytoplankton as a temperate marine source of brominated methanes
    Apr 20, 2016 · Bromoform (CHBr3) has an important role in transporting bromine from the ocean to the atmosphere, and released bromine catalyses ozone ...
  40. [40]
    [PDF] Marine sources of bromoform in the global open ocean - BG
    Mar 25, 2015 · Enhanced bromoform production during stress, as shown for macroalgae (e.g. Bondu et al., 2008), has not been demonstrated for phytoplankton.Missing: biosynthesis | Show results with:biosynthesis
  41. [41]
    Biosynthesis of bromoform by Curvularia fungi provides a natural ...
    Here we demonstrate that a Curvularia soil fungi isolate can be cultured to produce bromoform and subsequently inhibit methanogenesis.
  42. [42]
    Biosynthesis of bromoform by Curvularia fungi provides a natural ...
    Jan 14, 2025 · Here we demonstrate that a Curvularia soil fungi isolate can be cultured to produce bromoform and subsequently inhibit methanogenesis.Missing: terrestrial microbes
  43. [43]
    A metagenomic-based survey of microbial (de)halogenation ...
    Jun 29, 2016 · Furthermore it was demonstrated that in presence of inorganic bromide the formation of bromoform in soil is detectable. In all three soil ...
  44. [44]
    Correlations and emission ratios among bromoform ...
    Dec 13, 2005 · Typical values over the open ocean are 0.2–3 ppt for CHBr3, 0.1–0.5 ppt for dibromochloromethane (CHBr2Cl), and 0.5–2 ppt for dibromomethane (CH ...Missing: pre- | Show results with:pre-
  45. [45]
    Anthropogenic Bromoform at the Extratropical Tropopause - Jia - 2023
    May 10, 2023 · Bromoform (CHBr3) contributes to stratospheric ozone depletion but is not regulated under the Montreal Protocol due to its short lifetime ...Missing: volcanic | Show results with:volcanic
  46. [46]
    Simulations of anthropogenic bromoform indicate high emissions at ...
    Mar 18, 2021 · Bromoform is the major by-product from chlorination of cooling water in coastal power plants. The number of power plants in East and ...
  47. [47]
    Density Gradient - an overview | ScienceDirect Topics
    Halogenated liquids such as diiodomethane, bromoform, and tetrabromoethane can be used to prepare very dense solutions (>2.8 g cm−3). Problems associated with ...
  48. [48]
    Density gradient fractionation of dentine and bone powder
    The method employs ultracentrifugation of powdered material in density gradients made from acetone and bromoform. The density range of the fractions is ...Missing: centrifugation | Show results with:centrifugation
  49. [49]
    [PDF] Report (pdf)
    This makes it particularly useful in the density ranges for which heavy liquids such as bromoform (2.89 g/cm^) and methylene iodine (3.10 g/cm^) are used in the.
  50. [50]
    Bromoform – Reliable U.S. Supplier - Allan Chemical Corporation
    AllanChem provides Bromoform (CAS 75-25-2, CHBr3). High-purity for lab and industrial use. Consistent quality, in-stock, fast turnaround.
  51. [51]
  52. [52]
    [PDF] ATSDR Bromoform and Dibromochloromethane Tox Profile
    This is a toxicological profile for bromoform and dibromochloromethane, a compilation of information on a hazardous substance, including a public health ...
  53. [53]
    Bromoform - Chlorinated Drinking-Water - NCBI - NIH
    Bromoform has been identified as the major halogenated organic compound produced as a result of chlorinating seawater. Potential daily bromoform producation ...Missing: discovered | Show results with:discovered
  54. [54]
    effect of Rumin8 Investigational Veterinary Product—a bromoform ...
    The research demonstrates that Rumin8 oil, a synthetic bromoform-based feed additive, can reduce methane emissions in cattle by up to 95.0% without negativ.
  55. [55]
    Feeding a bromoform-based feed additive for methane mitigation in ...
    Bromoform (CHBr3) is a compound with strong methane-mitigating potential that binds to reduced vitamin B12, thereby inhibiting the cobalamin-dependent ...Missing: photobromination | Show results with:photobromination
  56. [56]
    A meta-analysis of effects of seaweed and other bromoform ...
    Research. A meta-analysis of effects of seaweed and other bromoform-containing feed ingredients on methane production, yield, and intensity in cattle.Missing: byproduct | Show results with:byproduct
  57. [57]
    A meta-analysis of effects of seaweed and other bromoform ...
    Jul 31, 2025 · At an average dose, bromoform feed reduced methane production by 47.3%, yield by 43.3%, and intensity by 39.0%, with greater mitigation in beef ...
  58. [58]
    Mitigating methane emissions in grazing beef cattle with a ... - PNAS
    This study's key finding—that bromoform-containing seaweed can reduce CH4 emissions (g/d) in cattle by an average of 37.7%—is an advancement in reducing the ...
  59. [59]
    Rumin8 achieves impressive 95.2% methane reduction in cattle trial ...
    Mar 11, 2025 · ... approved for use in Brazil and New Zealand. The study, titled “The effect of Rumin8 Investigational Veterinary Product – a bromoform-based feed ...
  60. [60]
    Study indicates synthetic bromoform feed additive is effective ...
    Mar 22, 2025 · New Zealand followed with provisional approval in July 2024. Additional trials are ongoing in Australia, New Zealand, Brazil, and the United ...
  61. [61]
    Natural and anthropogenic sources of bromoform ... - RSC Publishing
    Their atmospheric abundance is generally related to a common marine source, which is not well characterized.Missing: volcanic | Show results with:volcanic
  62. [62]
    The antimethanogenic efficacy and fate of bromoform and its ...
    Jul 11, 2025 · The half-life of bromoform was 26 min, coinciding with the production of dibromomethane accumulating to 22.1% of the initial bromoform amendment ...
  63. [63]
    Anthropogenic Bromoform at the Extratropical Tropopause - Jia - 2023
    May 10, 2023 · A new CHBr3 emission inventory based on natural and anthropogenic sources suggests that the latter account for 12%–28% of the global emissions.
  64. [64]
  65. [65]
    Transfer of Bromoform Present in Asparagopsis taxiformis to Milk ...
    During this metabolism, CHBr3 reacts with cytochrome p450 (superfamily of enzymes) into an intermediary dihalocarbonyl compound from which different ...
  66. [66]
    Bromoform - IDLH | NIOSH - CDC
    The revised IDLH for bromoform is 850 ppm based on acute inhalation toxicity data in animals.
  67. [67]
    [PDF] Provisional Peer Reviewed Toxicity Values for Bromoform
    Apr 14, 2005 · Some data on the acute or short-term effects of oral exposure are available because bromoform was used in the early 1900s to sedate children ...
  68. [68]
  69. [69]
    Drinking Water Disinfection Byproducts (DBPs) and Human Health ...
    Epidemiological studies have consistently observed an association between consumption of chlorinated drinking water with an increased risk of bladder cancer.Disinfection vs Disinfection... · Challenges for Epidemiology · Author Information
  70. [70]
  71. [71]
    National Primary Drinking Water Regulations | US EPA
    Maximum Contaminant Level ( MCL ) - The highest level of a contaminant that is allowed in drinking water. MCLs are set as close to MCLGs as feasible using the ...Chemical Contaminant Rules · Ground Water Rule · Surface Water Treatment Rules
  72. [72]
    [PDF] SAFETY DATA SHEET Bromoform - GeoLiquids
    May cause chemical conjunctivitis. Skin: Causes skin irritation. Harmful if absorbed through the skin. Inhalation: Causes irritation of the respiratory tract ...
  73. [73]
    Guidelines for drinking-water quality - NCBI Bookshelf - NIH
    Guideline values, Chloroform: 0.3 mg/l (300 µg/l). Bromoform: 0.1 mg/l (100 µg/l). Dibromochloromethane (DBCM): 0.1 mg/l (100 µg/l).
  74. [74]
    [PDF] Trihalomethanes in Drinking-water - World Health Organization (WHO)
    The most universally observed toxic effects of chloroform are liver and kidney damage. The severity of these effects per unit dose administered depends on the.
  75. [75]
    [PDF] a variety of assays on the genotoxicity of bromoform are equivocal ...
    Bromodichloromethane IARC has classified BDCM in Group 2B (possibly carcinogenic to humans). BDCM gave both positive and negative results in a variety of in ...
  76. [76]
    Is bromoform a food or a drug? | Sagentia Regulatory
    It's a rich, natural source of the active compound bromoform, which counteracts the formation of methane in the rumen. Efforts are underway to commercialize ...
  77. [77]
    Methane reduction, health and regulatory considerations regarding ...
    Historically, bromoform was used as a medicine in the early part of the last century as a sedative for children with whooping cough. Bromoform toxicity in ...<|separator|>
  78. [78]
    Health and Environmental Effects Document for Bromoform
    Bromoform was found in 100% of 34 samples of Arctic air from 8 sites near Alaska, Greenland, Norway and the North Pole; the samples were col- lected during ...
  79. [79]
    On the emissions and transport of bromoform: sensitivity to model ...
    Dec 21, 2015 · In a modelling study, we investigate the importance of the regional distribution of the emissions and of model resolution for the transport of ...
  80. [80]
    Bromoform as a source of stratospheric bromine - Sturges - 2000
    Jul 15, 2000 · Inclusion of dibromomethane and the mixed bromochloromethanes increased these figures to about 10% of Bro and 20% of Bry respectively.Missing: efficiency | Show results with:efficiency
  81. [81]
    The very short-lived ozone depleting substance CHBr3 (bromoform)
    Feb 17, 2021 · Seasonally dependent ozone depletion potentials (ODPs) for CHBr3 emitted in the Indian sub-continent were calculated to be 0.10, 0.34, 0.72, and ...
  82. [82]
    [PDF] a global modeling study of CHBr3 and CH2Br2
    Model sensitivity studies suggest that the contribution of CHBr3 to stratospheric bromine range from 0.5-1.6pptv when Bry is removed completely in con- vective ...
  83. [83]
    Oceanic bromoform emissions weighted by their ozone depletion ...
    Dec 10, 2015 · Here, we introduce a new approach for assessing the impact of oceanic halocarbons on stratospheric ozone by calculating their ozone depletion ...
  84. [84]
    Underestimation of Anthropogenic Bromoform Released into the ...
    Jan 15, 2022 · Anthropogenic CHBr 3 sources from industrial discharges may be a missing factor in global flux estimates or organic bromine to the atmosphere.
  85. [85]
    Effects of seawater intrusion on the formation of disinfection ...
    Jun 25, 2022 · Intrusion of seawater alters the chemistry of fresh water leading to the formation of additional brominated and iodinated disinfection byproducts (DBPs).
  86. [86]
    Formation of Halogenated Byproducts upon Water Treatment with ...
    The peracetic acid reagent contributes itself to hazardous bromoform formation when employed to oxidize/disinfect waters containing bromide: its use requires ...
  87. [87]
    [PDF] ATSDR Bromoform and Dibromochloromethane ToxGuide
    In chlorinated water samples, studies have reported mean concentrations ranging from 0.1–12 µg/L for bromoform and 1–. 28 µg/L for dibromochloromethane.
  88. [88]
    Vertical profiles of bromoform in snow, sea ice, and seawater in the ...
    Nov 15, 1997 · Elevated concentrations of bromoform were observed in both the ice (32–266 ng L−1 by liquid water volume) and seawater (∼20 ng L−1) at ...
  89. [89]
    [PDF] Drinking Water Criteria Document for Brominated Trihalomethanes
    Nov 15, 2005 · This document provides the health effects basis for setting Maximum Contaminant Level Goals (MCLGs) for brominated trihalomethanes in drinking ...<|separator|>
  90. [90]
    Assessing the Health Impact of Disinfection Byproducts in Drinking ...
    Mar 2, 2024 · This study provides a comprehensive investigation of the impact of disinfection byproducts (DBPs) on human health, with a particular focus on DBPs present in ...
  91. [91]
    Analysis of Cumulative Cancer Risk Associated with Disinfection ...
    Mar 24, 2020 · However, human health risks reported in epidemiological studies are likely due to mixtures of disinfection byproducts, marked by the presence of ...
  92. [92]
    [PDF] AWWA Ohio Section
    Nov 30, 2013 · The primary method for reducing DBPs is to reduce natural organic matter (NOM) via treatment processes at the water treatment plant prior to.
  93. [93]
    Innovative Approaches for Minimizing Disinfection Byproducts ...
    Techniques such as coagulation, activated carbon adsorption, and membrane filtration are standard methods employed in drinking water treatment plants (DWTPs) to ...<|control11|><|separator|>
  94. [94]
    Removal of Trihalomethanes from Drinking Water Via Heating Method
    Apr 20, 2016 · After boiling water for 1 minute, dibromochloromethane reduced from 8 to 4.2 µg/l (70% removal), and after 3 minutes it reached zero µg/l (100% ...Missing: mitigation strategies
  95. [95]
    Disinfection Byproducts Special Project - MN Dept. of Health
    Sep 27, 2024 · Exposure to DBPs can present some possible health risks, in addition to the clear benefits of reducing disease from pathogens. Hundreds of ...