Chloromethane
Chloromethane (CH₃Cl), also known as methyl chloride, is a colorless, flammable gas that serves as the simplest organochlorine compound, consisting of a single carbon atom bonded to three hydrogen atoms and one chlorine atom. It is the most abundant naturally occurring halocarbon in the Earth's atmosphere, primarily produced by oceanic sources and biomass burning, though smaller amounts are released from industrial activities.[1] With a faint, sweet odor detectable only at potentially toxic concentrations, chloromethane is highly volatile and slightly soluble in water (5.32 g/L at 25 °C), making it prone to rapid evaporation and atmospheric persistence of about one year.[2] Chloromethane is produced industrially on a large scale via reactions such as methanol with hydrogen chloride or chlorination of methane, serving mainly as an intermediate in silicone polymer production and other chemicals.[3][4] Historically used as a refrigerant and aerosol propellant, its applications have shifted due to safety concerns. It poses health risks primarily through inhalation, affecting the central nervous system, and contributes to stratospheric chlorine, though it is not bioaccumulative. Regulatory measures, including the U.S. EPA's reference concentration of 0.09 mg/m³, address exposure risks.[2][4]Properties
Physical properties
Chloromethane, with the molecular formula CH₃Cl, is the simplest chlorinated hydrocarbon and consists of a methyl group bonded to a chlorine atom. Its molecular weight is 50.49 g/mol. At standard temperature and pressure, it exists as a colorless gas characterized by a faint, sweet, ethereal odor detectable only at concentrations approaching toxic levels.[5] Key thermodynamic properties include a melting point of -97.4 °C and a boiling point of -24.2 °C, indicating its gaseous state under ambient conditions but liquidity when cooled or pressurized. The density of the liquid phase at the boiling point is 0.997 g/cm³, while the vapor density relative to air is 1.74. Chloromethane exhibits low solubility in water at 0.53 g/100 mL (5.3 g/L at 20 °C), reflecting its limited polarity, but it is highly miscible with organic solvents such as ethanol, diethyl ether, acetone, and benzene.[5][6][7][8][5] Its vapor pressure is notably high at 3800 mmHg (20 °C), contributing to its volatility and ease of evaporation. The critical temperature is 143.4 °C, above which it cannot be liquefied regardless of pressure. These properties make chloromethane suitable for applications requiring a readily compressible gas.[5][5]| Property | Value | Conditions | Source |
|---|---|---|---|
| Boiling point | -24.2 °C | 1 atm | PubChem |
| Melting point | -97.4 °C | 1 atm | PubChem |
| Liquid density | 0.997 g/cm³ | At boiling point | PubChem |
| Vapor pressure | 3800 mmHg | 20 °C | ChemicalBook |
| Critical temperature | 143.4 °C | - | PubChem |
| Water solubility | 0.53 g/100 mL (5.3 g/L) | 20 °C | Balchem SDS |
Chemical properties
Chloromethane, as a simple haloalkane, features a covalent C-Cl bond characterized by a length of 1.78 Å, determined through rotational spectroscopy measurements.[11] The bond dissociation energy for this C-Cl linkage is 351 kJ/mol at 298 K, reflecting the moderate strength typical of alkyl chlorides. The electronegativity difference between carbon (2.55) and chlorine (3.16) imparts polarity to the molecule, resulting in a dipole moment of 1.85 D, which influences its intermolecular interactions.[12] In terms of reactivity, chloromethane primarily undergoes nucleophilic substitution reactions via an SN2 mechanism, favored by the unhindered methyl group that allows backside attack by the nucleophile.[13] This is exemplified by its hydrolysis under basic conditions, where hydroxide ion displaces chloride: \ce{CH3Cl + OH^- -> CH3OH + Cl^-} This reaction proceeds through a concerted transition state without carbocation intermediates.[13] Additionally, chloromethane is susceptible to free radical halogenation, particularly chlorination, where UV light initiates abstraction of a hydrogen atom, leading to further substitution products like dichloromethane.[14] Chloromethane exhibits thermal stability under normal conditions but is highly flammable as a gas at room temperature, with explosive limits of 6.7–33.4% in air. At elevated temperatures above 400°C, it decomposes to hydrogen chloride and hydrocarbons via elimination or fragmentation pathways. Thermodynamically, its standard enthalpy of formation is -83.7 kJ/mol, indicating relative stability compared to dissociated fragments.[15] Isotopic variants of chloromethane, such as ¹³CH₃Cl or deuterated forms like CD₃Cl, are employed in labeling studies to trace metabolic pathways and biodegradation processes in environmental and biological systems, enabling precise tracking of carbon or hydrogen flux without altering reaction kinetics significantly.[16]Occurrence
Terrestrial sources
Chloromethane is emitted from marine environments primarily through biogenic processes involving phytoplankton and macroalgae such as seaweed. Laboratory studies have demonstrated that marine phytoplankton cultures produce methyl chloride via enzymatic mechanisms, including the action of chloroperoxidase enzymes that facilitate the methylation of chloride ions. For instance, species like the chlorophyte Dunaliella tertiolecta and diatom Phaeodactylum tricornutum have shown production rates in controlled settings, contributing to oceanic emissions estimated at a net global flux of approximately 0.66 Tg per year. Seaweed, particularly kelp, also releases methyl chloride as a metabolic by-product during degradation or direct biosynthesis, with field observations indicating sporadic high concentrations in coastal waters linked to algal blooms. Biogenic production of chloromethane occurs in terrestrial ecosystems through microbial activity, where fungi and bacteria methylate chloride ions in soils, decaying biomass, and organic matter. Wood-rotting fungi, such as those in the genus Phellinus (e.g., Phellinus pomaceus), are significant contributors, utilizing S-methyl groups from methionine as a methyl donor in secondary metabolism, leading to emissions during late growth phases. These fungi can convert up to 15% of available chloride to chloromethane in cellulose-based media, with global estimates attributing about 0.15 Tg per year to fungal sources worldwide, predominantly from tropical and subtropical forests. Bacterial production, often by methylotrophic species in soil microbiomes, supplements this through co-metabolism or direct utilization, though rates are generally lower and more variable in forest and grassland soils. Terrestrial plants, especially in tropical regions, emit chloromethane through metabolic processes involving halide methylation, potentially as a detoxification mechanism or volatile byproduct. Certain ferns (e.g., Nephrolepis biserrata) and trees in the Dipterocarpaceae family exhibit strong emissions, ranging from 3 to 100 nmol m⁻² h⁻¹ under natural conditions, driven by enzymatic activity similar to that in microbes. These plant sources are estimated to contribute around 0.6 Tg per year globally, with higher fluxes observed in humid tropical environments; for example, emissions from Australian tropical vegetation, including crops like sugarcane, have been measured at 200–500 kg ha⁻¹ year⁻¹ using enclosure techniques, linked to pectin demethylation or halide incorporation during growth. Volcanic and geothermal sources release trace amounts of chloromethane through hydrothermal vents, fumaroles, and degassing, likely formed abiotically from chloride interactions with organic precursors under high-temperature conditions. Measurements at sites like hydrothermal fields show concentrations in the parts-per-billion range, contributing negligibly to the global budget (less than 0.01 Tg per year), but serving as localized hotspots in geologically active areas. Overall, natural terrestrial and marine sources collectively account for an estimated global flux of 3–5 Tg of chloromethane per year, with oceans, plants, fungi, and biomass contributing the majority, while volcanic inputs remain minor. Flux chamber methods, involving sealed enclosures over soils, plants, or water surfaces to capture and quantify gas exchange, are commonly used for direct measurements, providing site-specific data that informs atmospheric models. These natural emissions play a key role in the tropospheric budget, with an atmospheric lifetime of about 1 year allowing for global dispersion.Extraterrestrial detections
Chloromethane (CH₃Cl) was first detected in the interstellar medium in 2017 toward the low-mass protostar IRAS 16293–2422 in the ρ Ophiuchi star-forming region, using the Atacama Large Millimeter/submillimeter Array (ALMA). The detection relied on millimeter-wave observations of rotational transitions, such as the J=13₁–12₁ line at 262.749 GHz for the ³⁵Cl isotopologue, confirming its presence at a column density of approximately 1.5 × 10¹⁴ cm⁻² in the warm inner envelope. The abundance of CH₃Cl relative to H₂ was estimated at around 10⁻¹⁰ to 10⁻⁹, based on radiative transfer modeling and comparisons with methanol abundances. In cometary environments, CH₃Cl was identified in the coma of comet 67P/Churyumov–Gerasimenko during the Rosetta mission (2014–2016) by the Rosetta Orbiter Spectrometer for Ion and Neutral Analysis (ROSINA). The molecule was detected via mass spectrometry, with an abundance ratio of CH₃Cl to CH₃OH ranging from 0.007 × 10⁻³ to 6 × 10⁻³, indicating levels comparable to those in protostellar envelopes and suggesting inheritance from the interstellar medium during solar system formation.[17] Formation of CH₃Cl in cold molecular clouds primarily occurs through gas-phase ion-molecule reactions, such as CH₃⁺ + Cl⁻ → CH₃Cl + e⁻, alongside contributions from ice mantle processes where HCl reacts with CH₃OH on dust grains during the pre-warm-up phase. These mechanisms operate efficiently at low temperatures (∼10 K), with models showing peak gas-phase abundances during the subsequent warm-up to ∼100–200 K, when ices sublimate and release the molecule into the gas. Identification in astronomical observations exploits the molecule's rotational transitions in the millimeter-wave regime, particularly the strong A-type (ΔK=0) and E-type (ΔK=±3n, n≠0) lines arising from internal rotation of the methyl group, which provide characteristic signatures for spectral line surveys. These transitions, observed between 100–300 GHz, enable precise abundance determinations and isotopic ratio measurements. The extraterrestrial detections of CH₃Cl highlight its abiotic formation in interstellar and cometary settings, underscoring the role of organohalogens in prebiotic chemistry by demonstrating that halogenated organics can arise naturally without biological influence, thus complicating biosignature interpretations in astrobiology and informing habitability assessments for early solar system bodies.[17] This presence links interstellar cloud chemistry to the delivery of complex molecules to nascent planets, potentially contributing to the primordial organic inventory essential for life's origins.Production
Industrial processes
The primary industrial process for chloromethane production is the free radical chlorination of methane, in which methane (CH₄) reacts with chlorine gas (Cl₂) to yield chloromethane (CH₃Cl) and hydrogen chloride (HCl). This exothermic reaction operates at temperatures of 400–500°C under light or thermal initiation, proceeding without catalysts to generate chlorine radicals that abstract hydrogen from methane. Yields for chloromethane typically range from 25–30% per pass, limited by sequential chlorination leading to byproducts like dichloromethane (CH₂Cl₂) and higher homologs.[18][19] Byproduct management is essential, as the reaction produces a mixture of chlorinated methanes; separation occurs through multistage fractional distillation, where chloromethane is isolated as an intermediate boiling fraction (boiling point -24°C), with unreacted methane recycled and heavier chlorides (e.g., CH₂Cl₂, boiling point 40°C) collected downstream. This distillation sequence achieves high purity (>99%) for commercial-grade chloromethane while minimizing waste.[20][21] Historically, production relied on the hydrochlorination of methanol (CH₃OH + HCl → CH₃Cl + H₂O), commercialized post-1930s using sulfuric acid catalysis, but shifted to methane chlorination in the 1970s amid rising methanol costs and abundant natural gas supplies, enhancing economic viability.[22] As of the 2020s, global production capacity stands at approximately 4–5 million metric tons annually, driven by demand in silicone and pharmaceutical sectors, with leading producers in China (accounting for over 50% of output) and the United States (via facilities operated by firms like Dow and INEOS).[23][24][25] The process demands significant energy for heating and compression—around 10–15 GJ per ton of chloromethane—but its catalyst-free nature avoids regeneration costs and simplifies reactor design, contributing to overall efficiency in large-scale operations.[18][21]Laboratory methods
Chloromethane can be synthesized in the laboratory through several controlled methods suitable for small-scale organic preparations. The classic approach involves the acid-catalyzed reaction of methanol with hydrogen chloride, facilitated by zinc chloride as a Lewis acid catalyst to promote the substitution. In a typical procedure, anhydrous zinc chloride is dissolved in excess methanol, and the mixture is refluxed while dry hydrogen chloride gas is bubbled through at around 150°C, yielding chloromethane gas along with water as a byproduct according to the equation: \ce{CH3OH + HCl ->[ZnCl2] CH3Cl + H2O} This method produces chloromethane in moderate yields, typically 60-80%, and is favored for its simplicity using readily available reagents. The reaction is exothermic and requires careful temperature control to minimize side products like dimethyl ether.[26][27] An alternative route, particularly valuable for isotopic labeling studies, employs diazomethane reacting with hydrogen chloride. Diazomethane, generated in situ from precursors like N-methyl-N-nitrosotoluene-p-sulfonamide (Diazald), is treated with anhydrous HCl in an ether solvent at low temperature (0-5°C) to afford chloromethane and nitrogen gas: \ce{CH2N2 + HCl -> CH3Cl + N2} This reaction proceeds rapidly and nearly quantitatively (>95% yield), making it ideal for incorporating carbon-13 or deuterium labels from labeled diazomethane precursors without isotopic dilution. However, diazomethane's explosive nature necessitates strict safety protocols, including small-scale operations and stabilization with ethanol. Another method utilizes the Grignard reagent methylmagnesium chloride, which can be chlorinated using sulfuryl chloride (SO₂Cl₂) under controlled conditions to generate chloromethane. The Grignard is prepared from methyl chloride and magnesium in dry ether, then reacted with SO₂Cl₂ at low temperature (-78°C) in a radical-mediated process, though yields are variable (40-70%) due to competing coupling reactions. This approach is less common but useful when avoiding protic acids.[28] Purification of laboratory-prepared chloromethane typically involves trap-to-trap vacuum distillation under an inert atmosphere (e.g., nitrogen or argon) to separate the volatile product (boiling point -24°C) from impurities like water, methanol, and hydrogen chloride. The crude gas is passed through cold traps cooled by liquid nitrogen or dry ice-acetone baths, with selective volatilization and collection in subsequent traps at intermediate temperatures (-80°C to 0°C). This fractional condensation achieves high purity (>99%) while preventing hydrolysis or polymerization.[29] Safety considerations are paramount given chloromethane's toxicity (irritant to eyes, skin, and respiratory system; LC50 ~5000 ppm in rats), flammability, and potential carcinogenicity. Preparations should be performed in a fume hood with explosion-proof equipment, using gas traps to capture effluents. In situ generation—directly bubbling the product into reaction mixtures for further use (e.g., in methylation)—is recommended to minimize storage and exposure risks. Personal protective equipment, including gloves, goggles, and respirators, is essential, and waste must be neutralized with base before disposal.[29]Uses
Modern applications
Chloromethane serves primarily as a chemical intermediate in modern industrial processes, with the majority of its production directed toward the synthesis of silicone polymers. In the Müller-Rochow direct process, chloromethane reacts with elemental silicon in the presence of a copper catalyst at elevated temperatures (typically 250–300°C) to yield dimethyldichlorosilane ((CH₃)₂SiCl₂), the foundational monomer for silicone elastomers, fluids, and resins widely applied in sealants, adhesives, lubricants, and medical devices. This application accounts for approximately 70–80% of global chloromethane consumption, underscoring its critical role in industries such as construction, automotive, electronics, and personal care.[30][31][32] A smaller but notable portion, around 5% of output, is utilized as a feedstock for methyl cellulose production. In this process, chloromethane reacts with alkali-treated cellulose under controlled conditions to introduce methoxy groups, forming methyl cellulose ethers that function as thickeners, stabilizers, and binders in pharmaceuticals, food products, cosmetics, and construction materials like cement additives. This application benefits from the compound's ability to precisely control the degree of substitution in the polymer chain.[29][19][33] Chloromethane also finds use in pharmaceutical manufacturing as a methylating agent for synthesizing active pharmaceutical ingredients and intermediates, including those for agrochemicals and certain analgesics. Its reactivity enables efficient alkylation steps in multi-stage organic syntheses, contributing to the production of compounds employed in treatments for various conditions.[31][34][35] Additional applications include the production of quaternary ammonium compounds, certain herbicides, and butyl rubber, accounting for roughly 4%, 4%, and 2% of global consumption, respectively.[5] Global demand for chloromethane in the 2020s approximates 4 million metric tons annually, predominantly in the Asia-Pacific region—particularly China and India—fueled by expanding silicone manufacturing and pharmaceutical sectors amid rapid industrialization and urbanization.[36][24][37]Historical and obsolete uses
Chloromethane, also known as methyl chloride, was first synthesized in 1835 by French chemists Jean-Baptiste Dumas and Eugène-Melchior Péligot by heating a mixture of methanol, sodium chloride, and sulfuric acid.[22] This marked the beginning of its recognition as an organohalogen compound, though industrial applications did not emerge until the early 20th century. By the mid-20th century, chloromethane's production and use had peaked, driven by its versatility in refrigeration and chemical synthesis, before many applications were discontinued due to toxicity concerns.[4] In the 1920s and 1930s, chloromethane served as a refrigerant in early household and commercial refrigerators under the name methyl chloride, valued for its low boiling point and heat transfer efficiency.[22] However, incidents of leaks causing fatal poisonings highlighted its high toxicity and flammability, leading to its replacement by less hazardous chlorofluorocarbons like Freon starting in the late 1930s.[38] This shift effectively ended its role in refrigeration by the mid-20th century.[4] Chloromethane also found use as a local and general anesthetic in the late 19th and early 20th centuries, particularly in mixtures for minor surgical and dental procedures. In 1901, French dentist Georges Rolland introduced Somnoform, a blend containing approximately 35% chloromethane, 60% ethyl chloride, and 5% ethyl bromide, which provided rapid induction for short operations.[39] Standalone chloromethane was applied topically for local numbing in the 1890s, but concerns over its narcotic effects, cardiac risks, and potential for overdose led to its discontinuation as an anesthetic by the 1940s.[40] From the 1920s through the 1970s, chloromethane acted as a key precursor in the production of tetramethyllead, an organolead compound used as an antiknock additive in gasoline to improve engine performance.[4] Tetramethyllead, introduced commercially in 1960, was synthesized by reacting chloromethane with sodium-lead alloys, though it represented a smaller fraction of leaded gasoline additives compared to tetraethyllead.[41] Its use declined sharply in the 1970s due to environmental regulations on lead emissions, culminating in the phase-out of leaded gasoline in most countries by the 1980s and early 1990s.[42]Environmental aspects
Atmospheric dispersion
Chloromethane enters the atmosphere from a combination of natural and anthropogenic sources, with predominantly natural emissions accounting for over 90% of the global budget of about 4–5 Tg per year, while anthropogenic contributions make up less than 10%.[13][43] Recent studies (as of 2024) confirm a total budget of ~4.7 Tg/yr, predominantly natural, though uncertainties persist regarding the exact contributions from vegetation and a previously identified missing natural source.[44][45] Natural sources include biomass burning, oceanic emissions, and terrestrial vegetation, particularly in tropical regions, whereas anthropogenic inputs primarily stem from chemical manufacturing processes and combustion activities. This source distribution results in a steady influx that maintains chloromethane as the most abundant organic chlorine compound in the atmosphere, contributing significantly to the natural chlorine loading in the troposphere.[43][13] Once released, chloromethane disperses globally and mixes rapidly in the troposphere due to its relatively short atmospheric lifetime of 0.8–1.0 years, governed mainly by its reaction with hydroxyl (OH) radicals:\ce{CH3Cl + OH ->[k = 3.6 \times 10^{-14} \, \mathrm{cm^3 \, molecule^{-1} \, s^{-1}} products}
This primary sink removes about 80% of chloromethane in the troposphere, producing hydrochloric acid (HCl), carbon monoxide (CO), and other products. The compound achieves a well-mixed distribution in the troposphere with an average concentration of 550 pptv, exhibiting minimal latitudinal gradients and seasonal variations of around 85 pptv in the Northern Hemisphere midlatitudes; however, stratospheric transport is limited, as only a small fraction reaches altitudes above 15–20 km before degradation.[13][46] Additional sink processes include stratospheric photolysis above 30 km, where ultraviolet radiation breaks down chloromethane into chlorine atoms and other radicals, and oceanic exchange, as surface waters are typically supersaturated, leading to net flux from ocean to atmosphere, with Henry's law constant of 3.4 M/atm governing solubility. Soil uptake and minor reactions with chlorine atoms also contribute marginally to removal. Long-term monitoring by the Advanced Global Atmospheric Gases Experiment (AGAGE) network, including sites like Mace Head and Cape Grim, indicates relatively stable global concentrations since the 1990s, but recent data (2000–2022) show a slight increasing trend of about +1.1 pptv per year, reflecting a balance between sources and sinks.[43][47][44]