Fact-checked by Grok 2 weeks ago

Halohydrin

A halohydrin, also known as a haloalcohol or β-halo alcohol, is a type of featuring a atom (typically , , or iodine) and a hydroxyl group (-OH) bonded to adjacent (vicinal) saturated carbon atoms. These molecules generally follow the structure R-CH(OH)-CHX-R' or similar variants, where the carbons bear only or hydrocarbyl groups otherwise, and common examples include 2-bromoethanol (ethylene bromohydrin) and 2-chloro-1-phenylethanol (styrene chlorohydrin). Halohydrins are primarily synthesized through the of a (X₂, where X = Cl or Br) to an in the presence of , or equivalently using hypohalous acid (HOX). This proceeds via a three-membered cyclic intermediate, first proposed in the context of by Roberts and Kimball in , where the alkene's π electrons attack the to form the bridged . then acts as a , attacking the more substituted carbon of the in a stereospecific anti-addition manner, followed by to yield the neutral product. In unsymmetrical , the addition is regioselective, with the hydroxyl group attaching to the more substituted carbon () and the to the less substituted one, due to the partial positive charge development on the more stable carbocation-like position during nucleophilic attack. These compounds play a crucial role in as versatile intermediates, notably for the preparation of through base-promoted intramolecular cyclization. Treatment of a halohydrin with a strong base (e.g., NaOH) displaces the via an SN2 mechanism, forming a three-membered oxirane ring while preserving the anti from the initial addition. This route is particularly valuable for regioselective synthesis from alkenes and has applications in producing pharmaceuticals, agrochemicals, and other fine chemicals, often as an alternative to peracid epoxidation methods. Halohydrins also exhibit reactivity in further transformations, highlighting their utility in stereocontrolled synthesis.

Definition and Nomenclature

Chemical Structure

A halohydrin features a core molecular framework consisting of a atom (X, where X = F, Cl, Br, or I) and a hydroxyl group (-OH) bonded to adjacent (vicinal) carbon atoms, defining its characteristic functionality. This refers to vicinal halohydrins, where the groups are on adjacent carbons, as opposed to halohydrins on the same carbon. The general formula is often expressed as R¹R²C(OH)–CR³R⁴X, with R¹–R⁴ representing hydrogen or alkyl/aryl substituents, allowing for a wide range of structural diversity. For chlorohydrins, a typical representation is R–CH(OH)–CH(Cl)–R', while analogous formulas apply to other halohydrins by replacing Cl with Br, I, or F. This vicinal positioning imparts distinct bonding characteristics, with the -carbon bond typically polar covalent and the carbon-hydroxyl bond enabling hydrogen bonding interactions. Halohydrins can exist in acyclic forms, such as linear chains in simple alkyl derivatives, or cyclic configurations, exemplified by six-membered like trans-2-chlorocyclohexanol, where the halogen and hydroxyl occupy adjacent positions across the ring. They also appear in complex molecules like carbohydrates, where the motif integrates into polyhydroxy frameworks with the halogen substituting a hydroxyl-bearing carbon adjacent to another -OH group.

Naming Conventions

Halohydrins are named systematically under IUPAC recommendations as substituted alcohols, with the serving as the principal that determines the suffix "-ol" for the parent chain. The carbon chain is selected and numbered to give the carbon atom bearing the the lowest possible locant, and halogen atoms are cited as prefixes (fluoro-, chloro-, bromo-, or iodo-) with their positions indicated. For instance, the compound with the formula HOCH₂CHClCH₃ is designated as 2-chloropropan-1-ol, prioritizing the 's position at carbon 1./Alcohols/Nomenclature_of_Alcohols) In addition to IUPAC , trivial or common names persist for certain halohydrins, particularly those derived from simple alkenes or featuring vicinal and hydroxy groups. The term "chlorohydrin" specifically denotes compounds where is the substituent, while "bromohydrin," "fluorohydrin," and "iodohydrin" apply analogously for the other . These names, such as chlorohydrin for HOCH₂CH₂Cl, emphasize the combination rather than the full structural detail and are widely used in synthetic contexts. Special naming conventions apply when halohydrins occur within complex molecules, such as carbohydrates. In sugar chemistry, these are typically designated as halo-deoxy derivatives, where the replaces a at a specific position, as in 3-chloro-3-deoxy-D-glucose. The follows carbohydrate-specific IUPAC rules, incorporating stereochemical descriptors (e.g., D or L) and locants that align with the standard numbering of the sugar ring or chain. The evolution of halohydrin traces back to 19th-century , when early discoveries of compounds like chlorohydrin in 1859 prompted the adoption of descriptive trivial names to reflect their hybrid nature. By the late 1800s, terms like "chlorohydrin" entered chemical literature around 1885, facilitating communication amid rapid advancements in reactivity studies. The shift to systematic IUPAC naming in the early standardized these conventions, integrating halohydrins into broader rules for alcohols and substituents while retaining trivial names for brevity in specialized applications.

Physical and Chemical Properties

General Properties

Halohydrins, particularly simple examples such as those derived from small alkenes, are typically colorless liquids or low-melting solids at room temperature. For instance, 2-chloroethanol exists as a colorless liquid with an ether-like odor. The presence of the hydroxyl group imparts high water solubility to halohydrins, often making them miscible with water, while the halogen substituent enhances overall molecular polarity. This solubility is exemplified by 2-chloroethanol, which is fully miscible in water. The vicinal arrangement of the halogen and hydroxyl groups contributes to this polarity in a single sentence. Boiling and melting points vary with molecular size and halogen type, but simple halohydrins generally have moderate boiling points due to hydrogen bonding from the OH group. 2-Chloroethanol, for example, boils at 128–130 °C and melts at –67 °C. In (IR) spectroscopy, halohydrins display characteristic absorption bands, including a broad O–H stretching vibration around 3400 cm⁻¹ indicative of hydrogen bonding in alcohols, and C–X stretching in the 850–515 cm⁻¹ range for the halogen (e.g., 850–550 cm⁻¹ for C–Cl). (NMR) spectroscopy reveals deshielded protons alpha to the halogen, typically appearing at 3.5–4.0 ppm; in 2-chloroethanol, the CH₂Cl protons resonate at approximately 3.87 ppm and the CH₂OH protons at 3.67 ppm in CDCl₃.

Stability and Reactivity

Halohydrins generally exhibit good thermal stability under ambient conditions but decompose at elevated temperatures, typically above 400°C for simple aliphatic examples. For instance, undergoes in the gas phase between 430–496°C, primarily yielding and via a process with an of approximately 230 kJ/mol. This decomposition pathway highlights the tendency of halohydrins to form carbonyl compounds under high heat, though more complex halohydrins may show lower onset temperatures depending on substituents and molecular structure. The hydroxyl group in halohydrins displays moderate acidity, with values typically in the range of 14–15, slightly lower than those of analogous simple alcohols ( ≈ 15.9–18). For , the of the OH group is reported as 14.02, reflecting the electron-withdrawing of the adjacent , which stabilizes the conjugate base by dispersing negative charge. This inductive withdrawal increases acidity compared to unsubstituted alcohols, with the effect strengthening as the 's rises (F > Cl > Br > I). Halohydrins show enhanced reactivity in acidic media, where of the hydroxyl group converts it into a more labile leaving group, facilitating subsequent transformations. This is a key step in reactivity but is amplified in halohydrins due to the neighboring halogen's potential to influence the through inductive or participation effects. In basic conditions, halohydrins are comparatively more stable, though the OH group remains deprotonatable. The nature of the significantly influences overall , with strength playing a dominant role: fluorohydrins are the most stable due to the robust C–F (dissociation energy ≈ 485 kJ/mol), followed by chlorohydrins (≈ 328 kJ/mol), bromohydrins (≈ 276 kJ/mol), and iodohydrins (≈ 238 kJ/mol), which are the least stable and prone to C–I cleavage. In specific polycyclic systems, the stability order mirrors this trend: chlorohydrin > bromohydrin > iodohydrin, underscoring the impact of weaker C–X bonds on and chemical endurance.

Synthesis

From Alkenes

Halohydrins are commonly synthesized from through the of a in the presence of . This method involves treating an with a dihalogen such as (Cl₂) or (Br₂) in an aqueous medium, leading to the formation of a intermediate that is subsequently attacked by as a . The mechanism proceeds in two key steps. First, the 's π electrons attack one end of the polarized X–X bond, forming a three-membered ring (chloronium or bromonium ion) and releasing the halide ion (X⁻). This cyclic intermediate shields one face of the , preventing syn addition. Second, water acts as the , attacking the more substituted carbon of the from the opposite face in an anti fashion, followed by proton loss to yield the halohydrin product. The partial positive charge in the is stabilized at the more substituted carbon, dictating the . This addition follows , where the hydroxyl group (OH) attaches to the more substituted carbon and the halogen to the less substituted one. For example, the reaction of propene with Br₂ in yields 1-bromopropan-2-ol as the major product. The general equation for a is: \text{R-CH=CH}_2 + \text{X}_2 + \text{H}_2\text{O} \rightarrow \text{R-CH(OH)-CH}_2\text{X} Yields are typically high under mild conditions, often at , with serving both as and reactant. The results in trans addition, producing a from achiral alkenes or meso compounds from symmetric ones like . This anti arises from the backside nucleophilic attack on the . This synthesis method was first applied industrially in for the production of ethylene chlorohydrin as a precursor to .

From Epoxides

Halohydrins are synthesized from s through nucleophilic ring-opening reactions, primarily under acidic conditions using hydrogen halides (HX, where X = Cl, Br, or I) in aqueous or alcoholic media. The reaction proceeds via an acid-catalyzed , where the oxygen is by H⁺ from HX, forming a protonated intermediate that enhances the electrophilicity of the adjacent carbons. This weakens the C-O bonds, allowing the (X⁻) to attack as the . In the acid-catalyzed pathway, the is such that the attacks the more substituted carbon of the protonated , as this position better stabilizes the partial positive charge in the , akin to an SN1-like mechanism for unsymmetrical epoxides. For symmetrical epoxides, such as cyclohexene oxide, the reaction with HBr yields trans-2-bromocyclohexanol, where the attaches to one of the equivalent carbons and the hydroxyl group to the adjacent carbon. Under basic conditions, ring opening with a (e.g., from NaX) inverts the , with the attacking the less hindered, less substituted carbon via an SN2 . This approach is less common for halohydrin formation compared to the acidic method but provides complementary selectivity for regioselective synthesis. The stereochemistry of ring opening is consistently , resulting in trans halohydrins due to backside attack by the , which inverts the at the attacked carbon. Epoxides employed in these reactions are frequently prepared upstream from alkenes via epoxidation.

From Other Precursors

Halohydrins can be synthesized from alpha-halo carbonyl compounds through of the , yielding vicinal halo-alcohols. For example, alpha-halo ketones or aldehydes are reduced using reagents such as , preserving the alpha-halogen while converting the carbonyl to a hydroxyl group. This method is particularly useful for preparing non-carbonyl halohydrins and is detailed in comprehensive reviews of alpha-halo carbonyl chemistry. Another route involves allylic of alkenes to form allyl halides, followed by enzymatic or addition processes that incorporate the hydroxyl group adjacent to the halogen. In biological systems, halohydrin dehalogenases or related enzymes facilitate the of allyl halides to halohydrins via intramolecular participation of the allylic halogen during functionalization. For instance, the reaction pathway can involve propene undergoing allylic halogenation to 3-bromoprop-1-ene (CH₂=CH-CH₂Br), followed by enzyme-mediated incorporation of the hydroxyl group to yield a vicinal halohydrin such as 1-bromopropan-2-ol through neighboring group migration. Electrochemical methods provide a clean alternative for halohydrin formation, particularly through anodic oxidation in media. In this approach, s are oxidized at the in the presence of sources like and or nucleophiles, generating iodohydrins without external oxidants or metal catalysts. Yields range from 19% to 90% depending on the alkene substrate, with the process involving generation of electrophilic iodine species. This technique is highlighted in recent advancements for sustainable synthesis. Less common chemical routes include selective substitution of vicinal dihalides, where one halogen is displaced by a hydroxyl group under controlled hydrolytic conditions. Enzymatic catalysis using halohydrin dehalogenases enables regioselective hydrolysis of compounds like 1,2-dibromoethane to 2-bromoethanol, offering high specificity in biological or biocatalytic contexts. Chemical analogs may employ mild aqueous conditions or silver-assisted displacement for similar transformations, though yields and selectivity vary. Recent biocatalytic advances include the enantiodivergent synthesis of halohydrins via engineered enzymes, which perform stereoselective C–H on alkyl to introduce the hydroxyl group adjacent to the (reported as of 2023). This method enables access to both enantiomers and complements traditional routes.

Reactions and Transformations

Epoxide Formation

The formation of from halohydrins proceeds via a base-promoted intramolecular cyclization, where the hydroxyl group of the halohydrin acts as a to displace the adjacent , forming a three-membered oxirane ring. This reaction is a classic example of an intramolecular adapted for ring closure. The mechanism begins with deprotonation of the hydroxyl group by a , generating an . This alkoxide then performs a backside nucleophilic attack on the carbon atom bearing the in an SN2 fashion, displacing the halide and closing the ring. The general equation for this transformation is: \ce{R-CH(OH)-CH2Cl + OH^- ->[base] R-CH\ -CH2 (epoxide) + Cl^- + H2O} where R represents an alkyl substituent, and the epoxide is depicted in its ring form. Typical conditions involve treatment with aqueous sodium hydroxide (NaOH) or potassium hydroxide (KOH), often at mild temperatures, leading to high yields. For instance, the chlorohydrin derived from propylene (1-chloro-2-propanol) cyclizes to propylene oxide, while the chlorohydrin from allyl alcohol yields glycidol (2,3-epoxy-1-propanol) in good efficiency. The stereochemistry of the cyclization features inversion of configuration at the carbon bearing the halogen due to the SN2 mechanism, which preserves the overall trans geometry of the starting halohydrin in the resulting epoxide. This cyclization serves as a key step in the industrial chlorohydrin process for production, where propylene chlorohydrin is treated with bases like (Ca(OH)2) or NaOH to generate the on a large scale.

Substitution and Elimination

Halohydrins serve as substrates for reactions primarily at the carbon atom bearing the , as the functions as a good in SN2 pathways, particularly when the carbon is primary or unhindered secondary. For instance, primary chlorohydrins can undergo halide exchange via the with in acetone, converting the to to facilitate subsequent transformations, as the acts as a strong displacing the in an SN2 mechanism. The general reaction for is represented as: \ce{R-CH(OH)-CH2Br + Nu^- -> R-CH(OH)-CH2Nu + Br^-} where \ce{Nu^-} is a nucleophile such as azide or cyanide, proceeding via backside attack with inversion of configuration at the substituted carbon. Leaving group ability influences the rate, with iodide being the best (I > Br > Cl) due to weaker C–I bond strength and better stabilization of the negative charge on the departing anion. Solvent effects are critical: polar aprotic solvents like acetone or DMF favor SN2 by solvating the nucleophile poorly, enhancing its reactivity, while polar protic solvents such as water or alcohols promote SN1 pathways for secondary or tertiary halohydrins through carbocation intermediates, potentially leading to rearrangements. The hydroxyl group in halohydrins can also undergo displacement under acidic conditions, where converts –OH to –OH₂⁺, a excellent leaving group (water), allowing substitution typically with hydrogen halides like HBr or to form dihalides. This SN1 or SN2 process (depending on the carbon substitution) follows the reactivity order > HBr > HCl, with primary halohydrins favoring SN2 and secondary favoring SN1 via formation. For example, treatment of a secondary chlorohydrin with HBr in acetic acid yields the corresponding vicinal chlorobromo compound with bromide replacing the hydroxyl. Elimination reactions of halohydrins are base-induced and proceed via E2 mechanisms, yielding alkenes when a β-hydrogen trans to the is available. The E2 pathway requires anti-periplanar alignment of the β-hydrogen and in the for efficient orbital overlap, leading to stereospecific anti elimination and favoring trans alkenes when applicable. formation can compete as an intramolecular pathway, but elimination predominates with bulky bases or in non-cyclizing geometries.

Other Key Reactions

Halohydrins can undergo oxidation of the hydroxyl group to yield α-halo ketones, a transformation that preserves the substituent while converting the secondary to a carbonyl. This reaction is typically achieved using mild oxidizing agents such as () in or (MnO₂) in neutral conditions, which selectively target the alcohol without displacing the adjacent halogen. For example, the oxidation of a chlorohydrin proceeds as follows: \ce{R-CH(OH)-CH2Cl ->[PCC or MnO2] R-C(O)-CH2Cl} This method is particularly useful for preparing α-halo carbonyl compounds, which serve as versatile intermediates in organic synthesis due to their reactivity in nucleophilic substitutions and enolate formations. In polymer chemistry, halohydrins are employed in coupling reactions to introduce vicinal functionalities into polymer chains, often through ring-opening of epoxide precursors followed by halogenation or direct incorporation via nucleophilic substitution. This approach allows for the synthesis of halohydrin-containing polymers with tailored properties, such as enhanced reactivity for further cross-linking or functionalization. For instance, polymer-bound epoxides can be converted to halohydrins using halide sources, enabling subsequent coupling steps that build complex macromolecular structures. Halohydrins also participate in C-C bond formation via organometallic routes, where the acts as a in cross-coupling reactions with organozinc or organoboron reagents under . These transformations extend the carbon skeleton while retaining stereochemical control from the original halohydrin, facilitating the construction of chiral building blocks. Such methods are valuable in , where the hydroxyl group can be protected or leveraged for directing selectivity. Under (UV) irradiation, halohydrins can exhibit photochemical halogen migration, involving 1,2-shifts facilitated by homolytic of the C- and intermediates. This process is particularly observed in iodohydrins or under conditions promoting hypo-halite formation, leading to rearranged products with potential applications in stereoselective synthesis. The migration is influenced by and , with visible or UV inducing rearrangements in hypoiodite derivatives of halohydrins. Recent developments since 2020 have advanced the use of halohydrins in asymmetric synthesis through chiral catalysts, particularly biocatalysts like halohydrin dehalogenases (HHDHs) and complexes. For example, evolved HHDHs enable stereoselective ring-opening of or epoxides with nucleophiles, yielding enantioenriched halohydrins for synthesis compatible with Cu(I) catalysis. Additionally, Ir-catalyzed of α-halo enones using dynamic kinetic resolution with chiral f-phamidol ligands produces vicinal halohydrins with high enantioselectivity (up to 99% ee), addressing limitations in scalable chiral pool access. In 2023, engineered enzymes (P450DA) enabled enantiodivergent synthesis of γ-halohydrins and β-haloallyl alcohols as building blocks for pharmaceuticals. In 2025, electrochemical of halohydrins was shown to enable cascading reactions for CO2 capture and formation of cyclic carbonates, offering sustainable routes to value-added chemicals. These methods highlight the growing role of halohydrins in enantiopure compound production.

Derivatives and Examples

Polyhalogenated Halohydrins

Polyhalogenated halohydrins are a class of halohydrins featuring multiple atoms attached to the carbon chain, typically adjacent to or on the same carbon as the hydroxyl group, exemplified by 1,2-dichloroethanol (ClCH₂CH(OH)Cl). These compounds differ from monohalohydrins by incorporating additional halogens, which can be vicinal, , or distributed along the chain, leading to distinct structural and reactive profiles. Synthesis of polyhalogenated halohydrins often involves the of halogenating agents in to polyhalogenated alkenes, analogous to standard halohydrin formation but adapted for substrates with pre-existing . For instance, undergoes reaction with in to produce 1,2-dichloroethanol: \ce{CH2=CHCl + Cl2 + H2O -> ClCH2CH(OH)Cl + HCl} This process proceeds via a chloronium intermediate, with the hydroxyl group adding to the more substituted carbon influenced by the existing chlorine . Alternatively, ring-opening of halogenated epoxides with halides can yield these structures, where the epoxide's inherent facilitates nucleophilic attack by the , incorporating additional during the process. The presence of multiple halogens imparts unique properties to polyhalogenated halohydrins, notably increased acidity of the hydroxyl group due to the inductive electron-withdrawing effects of the , which stabilize the conjugate and lower the compared to unsubstituted alcohols. This enhanced acidity facilitates under milder conditions, promoting further reactivity. Additionally, these compounds exhibit heightened instability and reactivity relative to monohalohydrins, often undergoing spontaneous elimination of to form carbonyl derivatives, such as aldehydes or ketones, driven by the or vicinal arrangement of the functional groups. Physical properties, such as (approximately 166.5°C for 1,2-dichloroethanol) and (1.398 g/cm³), reflect the polar nature augmented by the . Polyhalogenated halohydrins demonstrate elevated toxicity compared to their monohalogenated counterparts, with the additional exacerbating cellular damage through increased and reactivity toward biomolecules. In environmental contexts, they arise as disinfection byproducts during chlorination of water containing organic precursors, contributing to the overall health risks associated with halogenated disinfection byproducts, including potential carcinogenicity and .

Notable Compounds

One of the earliest notable halohydrins is ethylene chlorohydrin, also known as (HOCH₂CH₂Cl), first prepared in 1859 by Charles-Adolphe Wurtz via the reaction of with . This compound, isolated by distillation under reduced pressure, served as a key precursor in the development of production processes. Another significant example is 2-bromoethanol (HOCH₂CH₂Br), a simple bromohydrin prepared by the addition of to in aqueous medium, followed by with and . Alternatively, it can be synthesized from and , with isolation via to yield a colorless liquid boiling at 150–151°C. This compound is valued in for its reactivity as a bifunctional alkylating agent. Chloral hydrate (Cl₃CCH(OH)₂), discovered in 1832 by Justus von Liebig through chlorination of ethanol, is often regarded as a halohydrin analog despite being a geminal diol rather than a vicinal halohydrin. Its structure features three chlorine atoms adjacent to the hydrated carbonyl, and it was isolated as colorless crystals from aqueous solutions. Glycerol chlorohydrins, such as α-monochlorohydrin (HOCH₂CH(OH)CH₂Cl), are important derivatives obtained by controlled chlorination of glycerol with hydrochloric acid. These compounds, isolated by fractional distillation, find application in the manufacture of surfactants due to their amphiphilic properties.

Applications

In Organic Synthesis

Halohydrins serve as versatile intermediates in , particularly for constructing epoxides through base-mediated cyclization, which facilitates the introduction of oxygen functionality in complex molecules. This transformation is widely employed due to the ability of halohydrins to undergo regioselective ring closure, yielding epoxides that can be further functionalized via nucleophilic ring-opening reactions. In pharmaceutical synthesis, halohydrin-derived epoxides are crucial for producing chiral building blocks in antiviral drugs, such as the nelfinavir, where a chlorohydrin intermediate is generated via enzymatic reduction of an α-chloroketone and subsequently cyclized to form the epoxide core before incorporation into the final structure. Similarly, asymmetric organocatalytic methods have enabled the synthesis of epoxide intermediates for antiviral alkaloids like (+)-gliocladin C through kinetic resolution of spiro-epoxyindoles, achieving high enantiomeric excesses. Chiral halohydrins play a pivotal role in asymmetric induction during , enabling the stereocontrolled assembly of natural products and pharmaceuticals by leveraging their dual functionality for subsequent displacements or cyclizations. For instance, iridium-catalyzed of halogenated ketones produces enantioenriched halohydrins that serve as precursors for bioactive compounds, with selectivities up to 99% . In applications, chiral epoxides derived from halohydrins have been utilized in routes to antibiotics like monocillin I and hypoglycemic agents such as (R)-methyl palmoxirate, where the halohydrin step ensures stereochemical control. Specific examples include the enantioselective preparation of halohydrin precursors for β-blockers like , achieved through yeast-catalyzed reduction of α-haloketones, yielding the (R)-enantiomer with 95% and 85% yield using Pichia mexicana. In carbohydrate chemistry, halohydrins derived from inositols undergo hydride-mediated epoxide formation to access ring-modified derivatives, proceeding via an axial-rich conformation for efficient stereocontrol. The primary advantage of halohydrin formation lies in its regioselective functionalization of alkenes, where the reaction with X₂ in generates a that directs nucleophilic attack by to the more substituted carbon, following and yielding anti addition products with predictable regiochemistry. This contrasts with non-directed and enables precise control in polyfunctionalized systems. Modern organocatalytic methods have advanced this process, such as chiral ammonium salt-catalyzed asymmetric chlorocyclization of allylic alcohols, which proceeds through a chlorohydrin to afford chlorotetrahydrofurans with up to 99:1 dr and 95% ee, expanding access to enantioenriched motifs. Additionally, organocatalytic haloetherification of alkenes provides stereoselective halohydrins as synthetic equivalents for further elaboration in synthesis. These approaches update traditional methods by incorporating mild conditions and high stereocontrol, making halohydrins indispensable for efficient synthetic routes.

Industrial and Biological Roles

Halohydrins play a central role in industrial production, particularly through the chlorohydrin process for manufacturing (PO), a key precursor for polyurethanes, propylene glycols, and other chemicals. In this process, reacts with to form propylene chlorohydrin, which is then converted to PO using . As of 2024, the chlorohydrin route accounted for approximately 46% of global PO production, with overall PO demand reaching 10.1 million tonnes. This method dominates in regions like , where it supports large-scale facilities, though its share has been declining in some markets due to environmental concerns. As of 2025, the HPPO process continues to expand, with projections indicating it will capture a larger share of production due to its environmental benefits, potentially reducing the chlorohydrin dominance below 40% by 2030. Biologically, halohydrins are generated through enzymatic in marine organisms, serving roles in , signaling, and metabolite synthesis. Vanadium-dependent haloperoxidases in marine algae, such as those in and seaweeds, catalyze the oxidation of halides (e.g., or ) by to form hypohalites, which can add to unsaturated substrates like to produce halohydrins. Similarly, flavin-dependent halogenases in facilitate regioselective chlorination or bromination, leading to halohydrin intermediates in the of natural products, including antibiotics and pigments. These enzymes enable mild, stereospecific incorporation under aqueous conditions, contrasting with harsh chemical methods. In environmental contexts, halohydrins emerge as byproducts during chlorination for or disinfection, where reacts with bromide ions (abundant at ~65 mg/L in ) to form hypobromite, which can react with containing double bonds to yield bromohydrins. Such brominated species contribute to the suite of disinfection byproducts (DBPs) like and bromoacetic acids, posing challenges in coastal systems. Related to industrial scales, precursors like —derived from via halohydrin intermediates—see global production of around 2 million tonnes per year, supporting resin and synthesis. Sustainability issues with the chlorohydrin process include high salt waste (e.g., ) and consumption, prompting a shift to greener alternatives like the hydrogen peroxide-based PO (HPPO) process, which eliminates most inorganic byproducts and reduces water usage by up to 80%. Commercialized by Dow and since 2009, HPPO now represents growing capacity, with projections for further expansion to address eco-efficiency gaps in traditional routes.

Safety and Hazards

Health and Environmental Risks

Halohydrins exhibit through multiple exposure routes, acting as potent irritants to and eyes, with contact causing burns and inflammation. Inhalation leads to respiratory tract irritation, coughing, and potential , while or dermal can result in systemic effects including , , and . For instance, , a representative chlorohydrin, has an oral LD50 of approximately 71 mg/kg in rats, indicating high that may be fatal at low doses via , , or . Chronic exposure to halohydrins raises concerns for carcinogenicity and . Certain compounds, such as , are classified by the International Agency for Research on Cancer (IARC) as probably carcinogenic to humans (Group 2A), based on sufficient evidence from showing tumors in multiple organs and limited human data linking occupational exposure to cancer risks. is suspected of causing cancer and damaging the unborn child (EU CLP Category 2), though show no clear evidence of carcinogenicity. In the , while some halogenated organic compounds persist, halohydrins often exhibit moderate persistence due to their reactivity, including or microbial degradation, with half-lives typically ranging from days to weeks in (e.g., ~14 days for under neutral conditions). However, they can contribute to contamination if not degraded quickly. Simple bromohydrins show low potential in aquatic organisms due to low log Kow values (e.g., ~0.3 for 2-bromoethanol), though more lipophilic derivatives may pose higher risks. Human exposure to halohydrins occurs primarily through industrial accidents, where spills or leaks enable of vapors or dermal during and handling. Additionally, halohydrins form as minor disinfection byproducts in chloraminated , leading to low-level via treated supplies. Regulatory frameworks address these risks through occupational and environmental limits. The (OSHA) sets a (PEL) of 5 ppm (16 mg/m³) as an 8-hour time-weighted average for 2-chloroethanol, with skin notation due to absorption hazards. The U.S. Agency (EPA) monitors disinfection byproducts, including halogenated species, under the Stage 1 and Stage 2 Disinfectants and Disinfection Byproducts Rules (as of November 2025).

Handling Precautions

When handling halohydrins in laboratory or industrial settings, appropriate (PPE) is essential to minimize exposure risks. Workers should wear chemical-resistant gloves such as those made from Viton or rubber, safety goggles or face shields, and flame-retardant antistatic clothing to protect against skin contact, eye irritation, and potential ignition sources. Respiratory protection, including filters rated for organic vapors (e.g., A-P3 type), is recommended when working in areas with inadequate ventilation or during procedures that may generate mists or vapors. All manipulations should occur in a well-ventilated to prevent of volatile compounds. For storage, halohydrins must be kept in a cool, dry, well-ventilated area away from incompatible materials, particularly strong bases, which can induce to epoxides via intramolecular cyclization. Containers should be corrosion-resistant, tightly sealed, and stored under if prone to oxidation; access should be restricted to authorized personnel to prevent accidental exposure. Avoid proximity to heat sources, open flames, or ignition points due to the flammability of many halohydrins, such as . In the event of a spill, immediately evacuate non-essential personnel and ensure adequate to disperse vapors. Cover nearby drains to prevent environmental release, then absorb the liquid with inert materials like or commercial absorbents (e.g., Chemizorb), collecting the waste for proper disposal as hazardous material. Use non-sparking tools during cleanup to avoid ignition risks. For emergency exposure, first aid measures prioritize rapid decontamination. In cases of skin contact, immediately remove contaminated clothing and rinse the affected area with copious amounts of for at least 15 minutes, then seek attention. For eye exposure, flush with or saline for 15 minutes while holding eyelids open, removing contact lenses if present, and consult an ophthalmologist promptly. If inhaled, move the individual to fresh air and provide oxygen if breathing is difficult; call emergency services immediately. For , do not induce ; rinse the mouth with and seek urgent advice. Halohydrins exhibit general profiles similar to alkyl halides and alcohols, with potential for severe and systemic effects. Best practices include conducting operations in closed or contained systems, especially for volatile halohydrins, to limit airborne exposure and reduce fire hazards. Always label containers clearly, maintain an inventory of stored materials, and train personnel on emergency procedures; work should never be performed alone when handling these compounds. Wash hands and exposed skin thoroughly after use, and prohibit eating, drinking, or smoking in handling areas.

Nomenclature Issues

Common Misnomers

One common in the nomenclature of halohydrins arises with , a compound whose name suggests it is a chlorohydrin featuring a and hydroxyl group on adjacent carbons; however, it is actually an , specifically 2-(chloromethyl)oxirane, with a three-membered ring oxygen structure rather than a vicinal halo alcohol. Another frequent source of confusion is the term sulfuric chlorohydrin, which is an alternative name for chlorosulfonic acid (ClSO₃H), an inorganic oxoacid used in sulfonation reactions but lacking the characteristic C-C bonded and hydroxyl groups of organic halohydrins. Halohydrins are also sometimes conflated with alpha-halo ethers, such as (ClCH₂OCH₃), where a is attached to a carbon adjacent to an ether oxygen rather than a hydroxyl group, leading to misinterpretation of reactivity patterns in synthetic contexts. In addition, vicinal diols like (HOCH₂CH₂OH) can be mistakenly grouped with halohydrins due to structural similarity, though they lack the substituent entirely. A classic example of proper versus informal naming is ethylene chlorohydrin, commonly used but systematically designated as to distinguish it from other chlorinated derivatives and avoid implying a direct relation to (HOCl), which is an inorganic species not equivalent to organic chlorohydrins. These misnomers, including outdated historical designations like "oxychloride" for certain halo alcohols in early 19th-century literature, can propagate errors in chemical databases and patent documentation by misclassifying compounds and complicating literature searches. Early 19th-century literature often used terms like "oxychloride" for halo alcohols, stemming from incomplete understanding of structures before advanced. Modern IUPAC standardizes halohydrins as haloalcohols, e.g., , to avoid confusion. As of 2025, no major controversies persist, but green synthesis innovations highlight need for precise terminology in patents.

Historical Context

The direct formation of halohydrins through the addition of (generated from in aqueous media) to alkenes was first reported in 1863 by the German chemist Georg Ludwig Carius, yielding ethylene chlorohydrin from ethylene. This discovery laid the foundation for understanding to double bonds, as the compound was subsequently used by Charles-Adolphe Wurtz in to synthesize via base treatment. Industrial development of halohydrin chemistry accelerated in the early , particularly with the chlorohydrin for producing epoxides. Commercial production of via propylene chlorohydrin began in the early 1900s. Advances in the mid-20th century focused on mechanistic insights, with the of halohydrin formation—characterized by anti addition through a intermediate—elucidated through experimental studies in the 1950s, confirming trans product geometry via kinetic and product analyses. In the , environmental concerns over the traditional chlorohydrin route's salt-heavy have driven innovations in green synthesis, including biocatalytic resolutions using halohydrin dehalogenases for enantioselective production and electrochemical cascades that minimize reagents and byproducts. As of November 2025, these approaches address issues in traditional routes. Influential 19th-century chemists like contributed indirectly through foundational work on halogenated organics, including vapor density measurements aiding structural studies of haloforms, which share mechanistic parallels with halohydrin pathways.

References

  1. [1]
    halohydrins (H02727) - IUPAC Gold Book
    A traditional term for alcohols substituted by a halogen atom at a saturated carbon atom otherwise bearing only hydrogen or hydrocarbyl groups.Missing: definition | Show results with:definition
  2. [2]
  3. [3]
    The Halogenation of Ethylenes | Journal of the American Chemical ...
    Roberts, Navdeep Kaur, Fan Wu, Wei Li. Lewis Acid-Catalyzed Halonium Generation for ... Intermediate-catalyzed mechanism for the bromination of alkenes.Missing: halohydrin | Show results with:halohydrin
  4. [4]
    Sustainable Synthesis of Epoxides from Halohydrin Cyclization by ...
    Jul 4, 2022 · A facile synthesis of CaO- and MgO-based composite solid-based catalysts is demonstrated for a sustainable synthesis of epoxides from halohydrin cyclization.Introduction · Materials and Methods · Results and Discussion · Conclusions<|control11|><|separator|>
  5. [5]
    Halohydrins from Alkenes - Chemistry Steps
    Halohydrins are compounds containing a vicinal (adjacent) halogen and a hydroxyl groups, and they are prepared from alkenes via halonium ions.Missing: definition | Show results with:definition
  6. [6]
    Bromination of Alkenes - The Mechanism - Master Organic Chemistry
    Mar 15, 2013 · A halohydrin is a molecule containing C-OH and C-halogen bonds on adjacent carbons. The reaction also proceeds through a halonium ion ...
  7. [7]
    CHLOROHYDRIN Definition & Meaning - Merriam-Webster
    Word History ; Etymology. International Scientific Vocabulary chlor- + hydr- + -in entry 1 ; First Known Use. circa 1890, in the meaning defined above ; Time ...
  8. [8]
    CHLOROHYDRIN Definition & Meaning - Dictionary.com
    any of a class of organic chemical compounds containing a chlorine atom and ... Word History and Origins. Origin of chlorohydrin. First recorded in 1885 ...
  9. [9]
    [PDF] Principles of Chemical Nomenclature - iupac
    The growth of organic chemistry in the nineteenth century was associated with the development of more systematic nomenclatures, and chemists such as Liebig ...
  10. [10]
    2-Chloroethanol | ClCH2CH2OH | CID 34 - PubChem
    2-chloroethanol is a chloroethanol carrying a chloro substituent at position 2. It has a role as a xenobiotic metabolite. ChEBI. Ethylene Chlorohydrin is an ...Missing: angles ray
  11. [11]
    ICSC 0236 - 2-CHLOROETHANOL - INCHEM
    Flammable. Gives off irritating or toxic fumes (or gases) in a fire. Above 60°C explosive vapour/air mixtures may be formed. ; NO open flames, NO sparks and NO ...<|separator|>
  12. [12]
    IR: alkyl halides - Organic Chemistry at CU Boulder
    Alkyl halides show C-X stretches (850-515 cm-1), C-Cl stretches (850-550 cm-1), C-Br stretches (690-515 cm-1), and C-H wag (1300-1150 cm-1) in IR.
  13. [13]
    2-Chloroethanol(107-07-3) 1 H NMR - ChemicalBook
    2-Chloroethanol(107-07-3) 1H NMR · 90 MHz in CDCl3 · in CDCl3 · CCl4 solution · liquid film · 4880 A,200 M,liquid.
  14. [14]
    The thermal decomposition of 2-chloroethanol - CSIRO ...
    2-Chloroethanol decomposes at 430-496º into acetaldehyde and hydrogen chloride with first-order rate given by: k1 = 1012.8±1 exp(-229700 ± 4000/8.314T) s-l The ...
  15. [15]
    High-Temperature Thermal Decomposition of 2- Chloroethanol. Part ...
    Aug 10, 2025 · This paper reports single pulse shock tube and ab initio studies on thermal decomposition of 2-fluoro and 2-chloroethanol at T=1000-1200 K. Both ...
  16. [16]
    2-Chloroethanol | 107-07-3 - ChemicalBook
    Sep 25, 2025 · ... alcohol;2-CHLOROETHANOL GC STANDARD;CH2ClCH2OH;NCI-C50135;Ghlorohydrin ... pka, 14.02±0.10(Predicted). form, Liquid. color, Yellow. Odor, Faint ...
  17. [17]
    8.2: Properties of Alcohols and Phenols - Chemistry LibreTexts
    Jul 5, 2021 · The inductive effect is cumulative such that the acid strength of an alcohol becomes stronger (Lower pKa) as the number of halogens increases.Missing: halohydrins | Show results with:halohydrins
  18. [18]
    Reaction of Alcohols with HCl, HBr and HI Acids - Chemistry Steps
    For both mechanisms, the first step is the protonation of the alcohol to create the good leaving group H2O, after which the nucleophilic substitution occurs.
  19. [19]
    Cryogenic Organometallic Carbon–Fluoride Bond Functionalization ...
    This study demonstrates practical and high-yielding functionalization of the C–F bond under mild conditions, ie, at temperatures as low as −78 °C, in short ...Missing: fluorohydrins iodohydrins strength
  20. [20]
    Synthesis and Characterization of Bay Region Halohydrins Derived ...
    The relative stability of these derivatives is chlorohydrin > bromohydrin > iodohydrin. The chloro- and bromohydrins were isolated, but the iodohydrin ...
  21. [21]
  22. [22]
    [PDF] Organic Chemistry 1 Chapter 12. Reactions of Alkenes - CCC
    - Halohydrin formation involves treatment of alkenes with X2 and H2O (or HOX). Halonium ions are intermediates, and H2O attacks at the more substituted carbon ...
  23. [23]
    Alkene Reactivity - MSU chemistry
    In the case of hypochlorous and hypobromous acids (HOX), these weak Brønsted acids (pKa's ca. 8) do not react as proton donors; and since oxygen is more ...
  24. [24]
    Ethylene Oxide - Some Industrial Chemicals - NCBI Bookshelf
    Production. Ethylene oxide was produced from 1914 by the chlorohydrin process, in which ethylene chlorohydrin is prepared by reacting ethylene with hypochlorous ...
  25. [25]
    Opening of Epoxides With Acid - Master Organic Chemistry
    Feb 2, 2015 · Opening epoxides under acidic conditions occurs in two steps: first protonate epoxide, second nucleophile attacks, most substituted position, backside attack.
  26. [26]
    Epoxides Ring-Opening Reactions - Chemistry Steps
    Epoxides, due to high ring strain, undergo ring-opening reactions with strong nucleophiles (S_N2) and weak nucleophiles (acidic conditions, possibly S_N1-like).
  27. [27]
    Epoxide Ring Opening With Base - Master Organic Chemistry
    Feb 10, 2015 · Epoxide ring opening with base can be performed with hydroxide ion, alkoxides, Grignard reagents and several other nucleophiles.Epoxide Ring Opening With... · 1. Opening Epoxides With... · Quiz Yourself!
  28. [28]
  29. [29]
    Neighboring group migration in enzyme-mediated halohydrin ...
    Enzymatic halogenation of the double bond in allyl halides was influenced by intramolecular participation of the allylic halogen in the substrate molecule.
  30. [30]
    Halohydrin Dehalogenases Are Structurally and Mechanistically ...
    In most cases, specialized enzymes, designated dehalogenases, catalyze the cleavage of the carbon-halogen bonds, which is a key detoxification reaction.
  31. [31]
    15.7: Synthesis of Epoxides - Chemistry LibreTexts
    May 30, 2020 · Epoxides can also be synthesized by the treatment of a halohydrin with a base. This causes an intramolecular Williamson ether synthesis.
  32. [32]
    Nomenclature and preparation of epoxides (video) - Khan Academy
    May 11, 2014 · In the mechanism involving the halohydrin, Jay uses the base sodium hydroxide to depronate the halohydrin leading to an alkoxide.<|separator|>
  33. [33]
    Propylene Oxide - Some Industrial Chemicals - NCBI Bookshelf
    Commercial production of propylene oxide began in the early 1900s, by the chlorohydrin process involving the dehydrochlorination of propylene chlorohydrin with ...
  34. [34]
    On the Epoxidation of Alkenes
    The formation of epoxides by the halohydrin route is stereospecific. Each stereoisomer of 2-pentene gives a single epoxide. Because there are two SN2 inversions ...Missing: synthesis | Show results with:synthesis
  35. [35]
    The Production of Propene Oxide: Catalytic Processes and Recent ...
    This paper discusses the existing processes used for the production of propene oxide the chlorohydrin and hydroperoxide processes and their advantages and ...
  36. [36]
    Finkelstein reaction - Wikipedia
    It is an equilibrium reaction, but the reaction can be driven to completion by exploiting the differential solubility of various halide salts, or by using a ...
  37. [37]
    Leaving Group Ability Is Increased By Acid - Master Organic Chemistry
    Oct 23, 2024 · Adding Acid Increases Leaving Group Ability. Sure, it's one thing to recognize halogens (Cl, Br, and I) as having high leaving group ability ...
  38. [38]
    The Stereochemistry of Elimination Reactions Involving Halohydrin ...
    A Green Protocol for the Stereoselective Debromination of vicinal-Dibromides by [pmIm]BF4 under Microwave Irradiation. The Journal of Organic Chemistry 2005 ...
  39. [39]
    Reagent Friday: PCC (Pyridinium Chlorochromate)
    Sep 9, 2011 · PCC (pyridinium chlorochromate) is a useful oxidant in organic chemistry. It will oxidize primary alcohols to aldehydes, and secondary ...Missing: halo MnO2
  40. [40]
    Investigating Vicinal Functionalities by Ring-Opening of Epoxides on ...
    Sep 14, 2025 · This study synthesized and examined structure–property relationships of halohydrin-functionalized polymers derived from polycyclooctene (PCOE), ...
  41. [41]
    Enantioselective and Enantiospecific Transition-Metal-Catalyzed ...
    Aug 13, 2015 · Enantioselective and enantiospecific transition-metal-catalyzed cross-coupling reactions of organometallic reagents to construct C–C bonds.
  42. [42]
  43. [43]
    Asymmetric synthesis of chiral triazoles using a Cu(I)-compatible ...
    Jan 12, 2024 · In this study, the halohydrin dehalogenase HheG was used as the enzyme which was evolved to induce a stereoselective ring-opening reaction of ...
  44. [44]
    Highly efficient synthesis of enantioenriched vicinal halohydrins via ...
    A novel synthetic route was developed for the construction of chiral cis-vicinal halohydrins derivatives through Ir/f-phamidol-catalysed asymmetric ...
  45. [45]
    Ethanol, 1,2-dichloro-
    **Summary of CID 156449 - Ethanol, 1,2-dichloro-**
  46. [46]
    Haloalcohol - an overview | ScienceDirect Topics
    The chemistry of this functional group is dominated by the fact that it is inherently unstable relative to the corresponding ketone and hydrogen halide.Missing: solubility | Show results with:solubility
  47. [47]
    1,2-dichloroethanol | CAS#:74054-85-6 | Chemsrc
    Feb 27, 2024 · 1,2-dichloroethanol ; Molecular Weight, 114.95900 ; Density, 1.398g/cm3 ; Boiling Point, 166.5ºC at 760 mmHg ; Molecular Formula, C2H4Cl2O ; Melting ...
  48. [48]
    Fatty Acid Chlorohydrins and Bromohydrins Are Cytotoxic to Human ...
    This study has demonstrated the potential toxicity of the halohydrins, and implications for their formation in inflammation are discussed. Publication types.
  49. [49]
    Identification of Halohydrins as Potential Disinfection By-Products in ...
    Aug 6, 2025 · ... toxicity. Using gas chromatography coupled (independently) to low-resolution electron-impact mass spectrometry (LR-EI-MS), high-resolution ...Missing: polyhalogenated | Show results with:polyhalogenated
  50. [50]
    Ethylene oxide - American Chemical Society
    Nov 18, 2019 · Back in 1859, A. Wurtz treated ethylene chlorohydrin with potassium hydroxide to make ethylene oxide and potassium chloride. In 1914, BASF built ...Missing: origin | Show results with:origin
  51. [51]
    [PDF] A Study of the Reactions of Ethylene, Chlorine and Water in Bubble ...
    Wurz first prepared ethylene chlorohydrin in 1859 by reacting ethylene , glycol with hydrochloric acid(Jl). Carius, in 1863, reported a more direct synthesis( ...
  52. [52]
    Synthesis of Ethylene Chlorohydrin - ACS Publications
    Publication History. Published. online 1 May 2002. Published. in issue 1 April 1954.
  53. [53]
    Preparation of 2-bromoethanol - PrepChem.com
    The 2-bromoethanol is extracted by shaking twice with 100 ml of ether, dried over sodium sulfate, the ether is removed by evaporation yielding 85 g of bulk.Missing: isolation | Show results with:isolation
  54. [54]
    2-bromoethanol - Organic Syntheses Procedure
    In the 1-l. three-necked flask A (Fig. 8) is placed 550 cc. (4.56 moles) of 46 per cent hydrobromic acid (sp. gr. 1.46) (Note 1). Ethylene oxide (Note 2) is ...Missing: isolation | Show results with:isolation
  55. [55]
    Chloral hydrate | Podcast - Chemistry World
    Jul 31, 2012 · Chloral hydrate was first produced in 1832 by the great German chemist who first discovered nitrogen's role in fertilizers and made great ...
  56. [56]
    Preparation and Uses of Chlorinated Glycerol Derivatives - PMC - NIH
    The main focus of this review is the conversion of glycerol to chlorinated derivatives, such as epichlorohydrin and chlorohydrins, and their further use.
  57. [57]
  58. [58]
    Epoxides: Small Rings to Play with under Asymmetric Organocatalysis
    Mar 29, 2022 · Success stories from the synthesis of approved medicines and drug candidates using epoxide chem. in the development of robust and efficient ...
  59. [59]
    Iridium‐Catalyzed Asymmetric Hydrogenation of Halogenated ...
    Mar 30, 2018 · Iridium-catalyzed asymmetric hydrogenation of prochiral halogenated ketones was successfully developed to prepare various chiral halohydrins ...
  60. [60]
    [PDF] Applications of Chiral Three-membered Rings for Total Synthesis
    Apr 16, 2018 · Abstract: This review updates recent applications of asymmetric aziridination, azirination, thiirination, epoxi-.
  61. [61]
  62. [62]
    Lithium hydride as an efficient reagent for the preparation of 1,2 ...
    Jun 30, 2018 · Lithium hydride converts halohydrins to epoxides efficiently. · Inositol derived halohydrin flips to axial-rich conformation prior to reaction.
  63. [63]
    9.10: Formation of Halohydrins - Chemistry LibreTexts
    May 30, 2020 · The regioselectivity of halohydrin formation from an alkene reaction with a halogen in a nucleophilic solvent is analogous to the oxymercuration-demercuration ...Missing: paper | Show results with:paper
  64. [64]
    Absolute and relative facial selectivities in organocatalytic ...
    Mar 14, 2018 · Absolute and relative facial selectivities in organocatalytic asymmetric chlorocyclization reactions† ... The resulting halohydrin intermediate ...Results And Discussion · Synthesis Of Labeled... · Diastereoselectivity In...
  65. [65]
    Highly Stereoselective Intermolecular Haloetherification and ...
    Jun 25, 2015 · An organocatalytic and highly regio-, diastereo-, and ... [5k]); halohydrin synthesis: 84:16 e.r. (one example, Ref. [5l]) ...
  66. [66]
    Propylene Oxide Market Report | Global Forecast From 2025 To 2033
    The chlorohydrin process segment held the major share of the market in 2023, due to its extensive use in the production of propylene glycol, a key component in ...
  67. [67]
    Propylene Oxide Market Size, Share, Growth & Forecast 2035
    The global Propylene Oxide (PO) market demand stood at 10.1 million tonnes in 2024 and is expected to grow at a CAGR of 5.76% during the forecast period ...
  68. [68]
    Enzymatic Halogenation and Dehalogenation Reactions: Pervasive ...
    This review details the scope and mechanism of nature's halogenation and dehalogenation enzymatic strategies, highlights gaps in our understanding,
  69. [69]
    Structural Perspective on Enzymatic Halogenation - PMC - NIH
    Vanadium haloperoxidases have been found in fungi, marine algae, and bacteria. They require a vanadate cofactor, H2O2, and a halide to catalyze ...
  70. [70]
    Effect of Bromide on Chlorination Byproducts in Finished Drinking ...
    Other bromine-containing com- pounds can enter water through sewage and industrial effluents. Bromide is present in sea water at a concentration of 65 mg/L.
  71. [71]
    Effect of seawater intrusion on formation of bromine-containing ...
    This study quantified the formation and speciation of halogenated organic disinfection byproducts (DBPs) during chlorination as a function of seawater ...
  72. [72]
    Epichlorohydrin Market Size, Share, Analysis and Forecast 2035
    The global Epichlorohydrin market stood at approximately 1884 thousand tonnes in 2024 and is expected to grow at a CAGR of 4.52% during the forecast period ...
  73. [73]
    Presidential Green Chemistry Challenge: 2010 Greener Synthetic ...
    Oct 8, 2025 · Traditional PO production uses chlorohydrin or one of a variety of organic peroxides, which lead to coproducts such as t-butyl alcohol, styrene ...
  74. [74]
    How Propylene Oxide is Undergoing a Green Transformation?
    Oct 9, 2025 · For example, in 2023, the average annual profit for propylene oxide produced using the chlorohydrin process was approximately 786 yuan per ton, ...
  75. [75]
    Greener Synthetic Pathways Award | Dow Chemical and BASF
    The new HPPO process is considered a much greener reaction than the traditional routes to propylene oxide. For example, the chlorohydrin route requires ...Missing: issues alternatives
  76. [76]
    [PDF] Eco‐efficiency improvements in the propylene‐to‐epichlorohydrin ...
    Jun 27, 2023 · The novel process achieves 88% higher atom efficiency, 10% higher yield, 98% less wastewater, and lower energy use and reduced carbon dioxide ...
  77. [77]
    [PDF] Provisional Peer-Reviewed Toxicity Values for 2-Chloroethanol
    Nov 26, 2012 · The toxicity of 2-chloroethanol has not been reviewed by the Agency for Toxic Substances and Disease Registry (ATSDR, 2008) or the World Health ...Missing: halohydrin | Show results with:halohydrin
  78. [78]
    [PDF] 2-Chloroethanol - Publisso
    Jun 29, 2022 · In rats, 2-chloroethanol is oxidized to 2-chloroacetaldehyde in the liver by means of alcohol dehydrogenase and subsequently to 2-chloroacetic ...
  79. [79]
    [PDF] EPICHLOROHYDRIN 1. Exposure Data - IARC Publications
    Epichlorohydrin is probably carcinogenic to humans (Group 2A). In making the overall evaluation, the Working Group took into consideration the known ...
  80. [80]
    [PDF] Safety data sheet - ITW Reagents
    Sep 5, 2025 · Carcinogenicity Suspected of causing cancer. · Reproductive toxicity. Suspected of damaging the unborn child. 107-07-3 2-chloroethanol. NOAEL ( ...Missing: halohydrin | Show results with:halohydrin
  81. [81]
    Halohydrin dehalogenase immobilization in magnetic biochar for ...
    These compounds are structurally stable and do not decompose spontaneously, making them a widespread and persistent environmental ... bioaccumulation, as ...
  82. [82]
    Identifying Bioaccumulative Halogenated Organic Compounds ...
    May 28, 2015 · The presence of these compounds in Black skimmer eggs suggests they are persistent, bioaccumulative, potentially biomagnifying, and maternally ...
  83. [83]
    Organohalide respiration in pristine environments: implications for ...
    Dec 7, 2017 · Halogenated organic compounds, also termed organohalogens, were initially considered to be of almost exclusively anthropogenic origin.
  84. [84]
  85. [85]
    Identification of Halohydrins as Potential Disinfection By‐Products in ...
    Nov 24, 2011 · In 2001, two potential disinfection by-products (DBPs) were tentatively identified as 1-aminoxy-1-chlorobutan-2-ol (DBP-A) and its bromo ...
  86. [86]
    NIOSH Pocket Guide to Chemical Hazards - Ethylene chlorohydrin
    2-Chloroethanol, 2-Chloroethyl alcohol, Ethylene chlorhydrin. CAS No. 107-07 ... OSHA PEL. TWA 5 ppm (16 mg/m3) [skin] See Appendix G. Measurement Methods.
  87. [87]
    Stage 1 and Stage 2 Disinfectants and Disinfection Byproducts Rules
    Jul 1, 2025 · The Stage 1 Disinfectants and Disinfection Byproducts Rule (DBPR) reduces drinking water exposure to disinfection byproducts.Rule Summary · Rule History · Federal Register Notices
  88. [88]
    None
    ### Summary of Safety Information for 2-Chloroethanol (Aldrich - 23000)
  89. [89]
    Ch16: HO-C-C-X => epoxide - University of Calgary
    Halohydrins can be formed by the addition reaction X2 / H2O or HOX to alkenes. In the presence of a base, ring closure occurs via an intramolecular SN2 reaction ...<|control11|><|separator|>
  90. [90]
    General Properties of Epichlorohydrin - ChemicalBook
    Nov 8, 2019 · Despite its name, it is not a halohydrin. It is a colorless liquid with a pungent, garlic-like odor, moderately soluble in water, but miscible ...
  91. [91]
    Ethylene Oxide - NCBI - NIH
    The chlorohydrin process for the production of ethylene oxide was inefficient, because most of the chlorine that was used was lost as calcium chloride. Since ...
  92. [92]
  93. [93]
    Viktor Meyer - Wikipedia
    He is best known for inventing an apparatus for determining vapour densities, the Viktor Meyer apparatus, and for discovering thiophene, a heterocyclic compound ...Missing: halohydrin | Show results with:halohydrin