Fact-checked by Grok 2 weeks ago

Dehalogenation

Dehalogenation encompasses a range of chemical reactions that involve the removal of one or more halogen atoms (such as , , or iodine) from organic molecules by cleaving carbon-halogen bonds, serving as the inverse process to . This versatile transformation is fundamental in , where it enables the construction of unsaturated hydrocarbons, and in , where it aids in the degradation of persistent pollutants. In synthetic , dehalogenation often proceeds via eliminative mechanisms, such as , in which a (HX) is removed from adjacent carbons of an alkyl halide using a like alcoholic KOH, yielding alkenes or alkynes depending on the and conditions. For vicinal dihalides, treatment with in facilitates anti-elimination to stereospecifically form , a method particularly useful for controlling distribution in alkene synthesis. Reductive dehalogenation variants, employing agents like or catalytic with , are employed to fully reduce haloalkanes to alkanes or remove from aromatic systems without altering the carbon skeleton. Environmentally, reductive dehalogenation is a critical microbial process under conditions, where organohalide-respiring sequentially replace with using specialized enzymes called reductive dehalogenases, thereby detoxifying contaminants like polychlorinated biphenyls (PCBs), dioxins, and chlorinated ethenes in and soils. This pathway not only mitigates but also supports in microbes, as the process can couple with electron transport for . Abiotic reductive dehalogenation, facilitated by zero-valent metals like iron or catalyzed by nanoparticles, further enhances remediation strategies in engineered systems. Biologically, dehalogenation occurs via dedicated enzymes in and fungi, including dehalogenases that hydrolytically cleave C- bonds to produce alcohols and halides, or reductive dehalogenases that utilize corrinoid cofactors for stepwise removal from recalcitrant compounds. These enzymes play a natural role in the but are increasingly harnessed for of organohalogens, with applications extending to pharmaceutical , such as the defluorination of volatile anesthetics like in the liver. Overall, dehalogenation's mechanisms and applications underscore its importance across disciplines, from laboratory synthesis to global pollutant management.

Fundamentals

Definition and Classification

Dehalogenation refers to a chemical process that involves the of carbon- (C-X) bonds in compounds or, less commonly, metal- bonds in inorganic compounds, resulting in the removal of the atom and its replacement with , another , or through elimination to form unsaturated products. This process is distinct from , which entails the introduction of atoms into molecules via addition or reactions. Dehalogenation reactions are classified primarily based on the reaction pathway and the nature of the transformation. involves the replacement of the with through , often requiring electrons and protons; a generic representation is: \text{R-X} + 2\text{e}^- + 2\text{H}^+ \rightarrow \text{R-H} + \text{HX} where R is an group and X is a . proceeds via , where a (such as OH⁻ or CN⁻) displaces the from the carbon atom. , also known as , removes both a and an adjacent to form a double or , typically yielding alkenes or alkynes from alkyl or halides. is a rarer variant that couples removal with oxidation of the , often observed in specific transformations like the degradation of certain haloalkanes. Common substrates for dehalogenation include alkyl halides (R-X), vinyl halides, and aryl halides in organic contexts, where reactivity varies with the halogen (e.g., iodine being more labile than ) and the carbon framework. In inorganic contexts, dehalogenation applies to metal halides, such as salts in , where halogens are removed to form more stable compounds like phosphates or oxides. These reactions play key roles in and of halogenated pollutants.

Historical Context

The concept of dehalogenation emerged in the through observations of halogen removal from alkyl halides using metals such as . In 1849, English chemist Edward Frankland reported the reaction of ethyl iodide with zinc, yielding ethylzinc iodide and diethylzinc, which represented an early instance of reductive dehalogenation and laid the foundation for . This work, initially aimed at isolating free alkyl radicals, inadvertently demonstrated the potential of metal-mediated halogen abstraction in organic transformations. Advancements accelerated in the early with contributions from , who in developed the formation of organomagnesium reagents from alkyl halides and magnesium, enabling controlled reductions and further exploring dehalogenative processes in synthesis. By the 1920s and 1930s, catalytic hydrogenation emerged as a key milestone for dehalogenation, exemplified by Karl Wilhelm Rosenmund's 1921 introduction of a sulfur-poisoned on catalyst for selectively reducing acyl chlorides to aldehydes without over-reduction. These developments expanded dehalogenation's utility in , building on earlier metal reductions. The 1960s marked the discovery of biological dehalogenation, with studies identifying soil bacteria capable of enzymatically removing from compounds like chloroacetic acid, as reported in early investigations of microbial pathways. Influential figures like Grignard shaped the synthetic landscape, while environmental scientists such as Bruce Rittmann later advanced understanding of microbial processes in , including reductive dehalogenation of chlorinated phenols. Post-1970s, research evolved from synthetic applications to , spurred by the recognition of persistent pollutants like —banned in the U.S. in 1972—and PCBs, phased out by 1979, which highlighted the need for dehalogenative degradation strategies to mitigate contamination. This shift emphasized biological and catalytic methods to address ecological impacts.

Mechanistic Principles

Reductive Mechanisms

Reductive dehalogenation involves the removal of atoms from compounds through the addition of or , typically leading to the formation of less halogenated or fully dehalogenated products. This process is distinct from other dehalogenation types as it proceeds via , often under conditions to prevent oxidative side reactions. The primary pathway for reductive dehalogenation is the single (SET) mechanism, which generates intermediates. In this process, an (R-X) accepts an electron from a reductant, forming an alkyl (R•) and a anion (X⁻):
\ce{R-X + e^- -> R^\bullet + X^-}
The then undergoes or further to yield the dehalogenated product (R-H), often involving a second or :
\ce{R^\bullet + H^+ + e^- -> R-H}
This stepwise generates species consistent with the SET pathway in reactions with metals like .
Common reductants for these reactions include dissolving metals such as in (Zn/HCl) or (Na/Hg), which provide s under protic conditions. Catalytic using (Pd/C) with gas (H₂) is also widely employed, particularly for aryl halides or halides, as it facilitates :
\ce{R-X + H2 ->[Pd/C] R-H + HX}
In environmental contexts, such as microbial systems, zero-valent iron (Fe⁰) or other metals serve as electron donors, mimicking these chemical pathways under ambient conditions.
Specific examples illustrate the versatility of reductive dehalogenation. For vicinal dihalides, such as (Br-CH₂-CH₂-Br), treatment with leads to elimination of two halogens and formation of an (CH₂=CH₂) via sequential SET and radical coupling or . In contrast, geminal dihalides like CH₂Br₂ are converted to alkanes (CH₄) through stepwise , where each is replaced by . These transformations highlight the pathway's efficiency for polyhalogenated compounds. Stereochemistry in reductive dehalogenation varies with the and conditions. For vicinal dihalides, the process often proceeds with anti-elimination due to the trans arrangement of in the , preserving the observed in cyclic systems. However, in monohalides or under catalytic conditions, radical recombination can lead to retention or , depending on the lifetime of the radical .

Nucleophilic and Eliminative Mechanisms

Nucleophilic dehalogenation involves the direct displacement of a atom from an substrate by a , typically proceeding via SN1 or SN2 mechanisms for alkyl halides. In the SN2 pathway, a concerted backside attack by the inverts the configuration at the carbon center and is favored for primary and methyl substrates due to minimal steric hindrance. The general reaction is represented as: \text{R-X} + \text{Nu}^- \rightarrow \text{R-Nu} + \text{X}^- where R is an alkyl group, X is the halogen leaving group, and Nu is the nucleophile. A classic example is the Finkelstein reaction, where chloride or bromide in an alkyl halide is substituted by iodide using sodium iodide in acetone, leveraging the precipitation of NaCl or NaBr to drive the equilibrium. This SN2 process is particularly effective for primary alkyl chlorides and bromides, yielding alkyl iodides in high yields. For tertiary substrates or in polar protic solvents, the SN1 mechanism predominates, involving carbocation formation and racemization, though it is less common for dehalogenation due to competing rearrangements. Aryl halides resist direct SN1 or SN2 due to the sp²-hybridized carbon and poor overlap with the 's orbital, but under strong basic conditions and high temperatures, they undergo dehalogenation via an elimination-addition pathway involving a benzyne intermediate. In this mechanism, a strong base abstracts an ortho proton, leading to loss of the and formation of the transient benzyne species, which then adds the at either position, resulting in mixtures of ortho- and meta-substituted products. This was established through studies showing non-retention of the original pattern. Eliminative dehalogenation, or , removes both a and an adjacent hydrogen to form , primarily via E2 or E1 mechanisms. The E2 process is a concerted, bimolecular elimination requiring anti-periplanar between the leaving groups, promoted by strong bases like alkoxides in alcoholic solvents. For instance, treatment of an alkyl halide such as with alcoholic KOH yields via E2, following Zaitsev's rule to favor the more substituted . The reaction is depicted as: \text{R-CH}_2\text{-CHX-R'} + \text{B}^- \rightarrow \text{R-CH=CH-R'} + \text{HX} + \text{BH} E1 elimination, involving a carbocation intermediate, occurs under milder basic or neutral conditions with tertiary halides, but E2 dominates for synthetic dehydrohalogenation. In polyhalogenated compounds, alpha-elimination can generate carbenes; chloroform with a strong base like tert-butoxide undergoes dehalogenation to dichlorocarbene, a reactive singlet species used in cyclopropanation. Key factors influencing these mechanisms include ability, which follows the order I > Br > Cl > F due to decreasing C–X bond strength and increasing basicity of the ion, making the most facile leaving group. Substrate sterics hinder SN2 and favor E2 for secondary and halides, while polarity stabilizes ions in SN1 and E1 pathways.

Thermodynamic Aspects

Energetics of Dehalogenation

The energetics of dehalogenation are primarily governed by the dissociation energies (BDEs) of carbon- bonds, which determine the thermodynamic feasibility of . Typical BDEs decrease in the order C–F (485 /) > C–Cl (338 /) > C–Br (276 /) > C–I (238 /), reflecting the increasing atomic size and decreasing overlap efficiency of orbitals with carbon. This trend explains the higher reactivity of iodides and bromides in dehalogenation compared to chlorides and fluorides, as weaker bonds require less energy input for homolytic or heterolytic . Reaction enthalpies (ΔH_rxn) for dehalogenation can be estimated using the relationship ΔH_rxn = Σ BDE(reactants) – Σ (products), which accounts for the net energy change in breaking and forming. For reductive dehalogenation processes, such as the conversion of an alkyl bromide to the corresponding (R–Br + H₂ → R–H + HBr), these reactions are typically exothermic, with ΔH ≈ –75 kJ/mol, driven by the formation of strong C–H and H–X s that outweigh the C–X and H–H disruptions. In contrast, aryl halides exhibit less favorable energetics due to higher BDEs; for example, the C–Cl BDE in is approximately 393 kJ/mol, compared to 338 kJ/mol for alkyl chlorides, resulting in more endothermic or less exothermic ΔH_rxn values owing to stabilization of the aryl-halogen . Free energy changes (ΔG) further modulate thermodynamic favorability, given by the equation ΔG = ΔH – TΔS, where contributions (ΔS) play a key role in eliminative dehalogenation pathways. Eliminative processes, such as to form alkenes (e.g., R–CH₂–CHX–R' → R–CH=CH–R' + HX), often exhibit positive ΔS due to the net increase in molecular (from two reactants to three products), enhancing exergonicity (negative ΔG) at elevated temperatures as the –TΔS term dominates. Solvent effects influence these energetics through differential solvation of halide ions (X⁻), which are products in many heterolytic dehalogenations. Protic solvents strongly solvate small, basic anions like F⁻ via hydrogen bonding, stabilizing products and lowering ΔG more than in aprotic media, whereas larger I⁻ experiences weaker solvation, potentially shifting equilibria. In reductive cleavages, solvation modulates the of , altering ΔH by 10–20 kJ/mol depending on .

Factors Influencing Reactivity

The reactivity of substrates in dehalogenation reactions varies significantly with the structural features of the . In mechanisms such as SN2, primary alkyl halides exhibit higher reactivity than secondary, which in turn are more reactive than halides, primarily due to increasing steric hindrance at the reaction center that impedes back-side attack by the . and aryl halides display markedly lower reactivity in these processes compared to their alkyl counterparts, attributed to the sp² hybridization of the carbon atom bearing the halogen, which results in a stronger C-X bond with partial double-bond character and restricts nucleophilic approach. Reaction conditions and environmental factors profoundly impact dehalogenation . Temperature influences the rate through the Arrhenius relationship, where elevated temperatures reduce the effective barrier and accelerate the process across mechanisms. Solvent plays a pivotal role: polar protic solvents, such as or alcohols, stabilize charged states and ions, thereby favoring unimolecular pathways like SN1, whereas polar aprotic solvents, like (DMSO), solvate cations poorly and enhance reactivity to promote SN2 dehalogenation. In aqueous systems, modulates reactivity by altering the state of nucleophiles or bases, with acidic conditions often suppressing anionic nucleophiles and basic conditions accelerating eliminative pathways. (Ea) for dehalogenation differ among , generally decreasing in the order F > > Br > I due to the progressively weaker C-X bonds and better leaving group ability of heavier , facilitating lower energy barriers in substitution reactions. Steric and electronic effects further modulate dehalogenation rates. Beta-branching or bulky substituents adjacent to the carbon- bond sterically hinder the approach of nucleophiles, substantially slowing SN2 reactions, as exemplified by the sluggish reactivity of neopentyl halides compared to n-butyl halides. In contrast, electron-withdrawing groups or to the in SN1-prone substrates stabilize the developing intermediate, thereby lowering the and enhancing reactivity. Isotope effects provide insight into mechanistic details, particularly in eliminative dehalogenation. The (KIE) for versus at the beta position in E2 mechanisms typically ranges from 3 to 7 (kH/kD ≈ 3-7), reflecting partial C-H bond breaking in the rate-determining and confirming a concerted process. Representative examples illustrate these influences. In polar protic media like , tertiary alkyl bromides undergo rapid SN1 dehalogenation due to stabilization, whereas primary alkyl chlorides show minimal reactivity under the same conditions; conversely, in polar aprotic solvents such as acetone, primary alkyl iodides exhibit enhanced SN2 rates, outperforming bromides by factors related to ability.

Synthetic Applications

In Organic Synthesis

Dehalogenation plays a crucial role in as a interconversion strategy, converting organic halides to hydrocarbons or introducing unsaturation to build complex carbon skeletons. This process is particularly valuable for simplifying molecular frameworks after steps used in chain extension or cyclization reactions. Common reactions include the -mediated debromination of allylic halides, which proceeds under mild aqueous or alcoholic conditions to yield the corresponding alkenes with good . For instance, treatment of allylic bromides with dust in facilitates clean reduction, preserving nearby double bonds. Hydrodechlorination of aryl or alkyl chlorides using in acidic media is frequently employed in pharmaceutical synthesis to prepare intermediates by replacing with , as seen in the reduction of chlorinated aromatic precursors to bioactive scaffolds. A representative example is the synthesis of from vicinal dihalides, where in acetic acid or effects stereospecific elimination, often retaining the alkene geometry from the precursor. This method is widely used to generate cis- from anti-dibromides derived from alkene bromination. In , such as steroid chemistry, zinc-mediated dehalogenation removes alpha-halogen substituents from ketones, as demonstrated in the preparation of cholestenone derivatives by reducing 2-bromocholestan-3-one, streamlining access to natural frameworks. These approaches offer advantages like mild reaction conditions compatible with diverse functional groups and high selectivity for targeted , enabling efficient multi-step sequences. However, limitations include potential over-reduction of multiple or interference from acidic protons in the substrate. A specific application in involves reductive dehalogenation to remove halogen protecting groups, such as zinc- or palladium-assisted deiodination of iodinated residues during tritium labeling or side-chain modification, ensuring orthogonal deprotection without disrupting the peptide backbone.

Catalytic Methods with Transition Metals

Catalytic methods employing transition metals have revolutionized dehalogenation by enabling selective and efficient removal of halogen atoms from organic substrates, particularly aryl and alkyl halides, under mild conditions. These processes typically leverage low-valent metal centers, such as Pd(0), Ni(0), or (0/III), to activate the C–X bond, often proceeding via followed by reduction and elimination steps. This approach contrasts with stoichiometric reductions by allowing catalyst loadings as low as 1 mol%, minimizing waste and enhancing scalability for synthetic applications. Palladium catalysts are particularly effective for hydrodehalogenation of aryl halides, including challenging chlorides, due to the facility of of the C–X bond to (0) species. The general mechanism involves initial to form an aryl-(II)-X intermediate, followed by hydride delivery from a —such as H₂, silanes, or alcohols—via or direct insertion, and subsequent to yield the arene and HX byproduct. For instance, Pd-catalyzed hydrodehalogenation using silanes, as in methods employing Pd catalysts in THF, enables chemoselective reduction of aryl bromides and chlorides bearing sensitive groups like carboxylic acids or . Another variant utilizes alcohols as hydride sources; for example, Pd/YPhos complexes with in MeTHF provide mild conditions for hydrodehalogenation of aryl chlorides and polyhalogenated compounds, achieving high yields. approaches include Pd supported on ceria (Pd/CeO₂) for transfer hydrodehalogenation of halophenols using isopropanol, avoiding gaseous H₂ and tolerating functional groups like and carbonyls. Nickel catalysts offer a cost-effective alternative, especially for both aryl and alkyl halides, where to (0) generates an alkyl/aryl-(II)-X species, followed by β-hydride elimination from the reductant or to facilitate . A notable example is Ni-catalyzed hydrodehalogenation using isopropylzinc or tert-butylmagnesium as reductants, with O,N,O-ligated Ni complexes enabling good to excellent yields (70–95%) under mild conditions (, THF ) for unactivated alkyl bromides and iodides. These systems exhibit broad substrate scope, including primary and secondary alkyl halides, with low catalyst loadings (2–5 mol%) and high , avoiding over-reduction. Iron-based catalysts provide an economical and environmentally benign option, particularly for aryl halides, operating through or low-valent pathways initiated by Fe(III) to active Fe(0) species. A practical uses 1 mol% Fe(acac)₃ with t-BuMgCl as reductant in THF at 0°C, achieving hydrodehalogenation of aryl iodides, bromides, and chlorides in 1–1.5 hours with yields up to 98%, while tolerating halides like F, OR, CN, and CO₂R groups. The mechanism likely involves single-electron transfer to generate alkyl/aryl s, followed by , highlighting iron's utility in green catalysis for complex molecules. Overall, these methods excel in functional group tolerance, enabling selective dehalogenation in polyhalogenated systems, and have advanced toward sustainable practices, such as or alcohol-based , reducing reliance on pressurized H₂. Recent developments as of 2025 include visible-light-driven Cu-catalyzed dehalogenation for of polychlorinated compounds under mild conditions, and electrochemical methods combining s for efficient hydrodehalogenation, further expanding applications in sustainable .

Environmental and Biological Dehalogenation

Microbial Processes

Microbial dehalogenation refers to the biological transformation of halogenated compounds by microorganisms, primarily through enzymatic processes that cleave carbon-halogen bonds. These processes are crucial for the degradation of environmental pollutants such as chlorinated solvents and pesticides, occurring under both and aerobic conditions. Microorganisms employ specialized dehalogenase enzymes to facilitate these reactions, enabling the use of organohalides as carbon sources, acceptors, or mechanisms. Haloalkane dehalogenases (HLDs) represent a major class of hydrolytic enzymes involved in aerobic dehalogenation. HLDs catalyze the of carbon- bonds in , converting them to alcohols, halides, and protons via a mechanism. The reaction proceeds through an aspartate residue acting as a that attacks the carbon atom bound to the halogen, forming a covalent alkyl-enzyme , followed by water-mediated . A representative for this process is: \text{R-Cl} + \text{H}_2\text{O} \rightarrow \text{R-OH} + \text{HCl} where R denotes an alkyl group. HLDs are α/β-hydrolases found in various bacteria, such as Rhodococcus and Pseudomonas species, and have been extensively characterized for their substrate specificity and catalytic efficiency. Reductive dehalogenases (RDases), another key class, are corrinoid-dependent enzymes primarily active in anaerobic environments. These iron-sulfur cluster-containing proteins facilitate reductive dehalogenation by using low-potential electrons to replace the halogen with a hydrogen atom. RDases are membrane-associated and function in organohalide respiration, where organohalides serve as terminal electron acceptors. The mechanism involves electron transfer from a corrinoid cofactor (e.g., cobamide) to the substrate, often coupled with electron bifurcation to balance energy demands in low-redox environments. The general reductive reaction is: \text{R-Cl} + 2\text{e}^- + \text{H}^+ \rightarrow \text{R-H} + \text{Cl}^- RDases exhibit high specificity; for instance, those in Dehalococcoides species target polychlorinated ethenes. Anaerobic reductive dehalogenation pathways are exemplified by the sequential degradation of tetrachloroethene (PCE) to ethene by Dehalococcoides mccartyi strains. This bacterium uses multiple RDases, such as TceA for trichloroethene (TCE) to cis-dichloroethene (DCE) and VcrA for vinyl chloride (VC) to ethene, enabling complete detoxification under strictly anaerobic conditions with hydrogen as the electron donor. These pathways support microbial growth via dehalorespiration and are widespread in contaminated subsurface environments. Aerobic dehalogenation often involves oxidative pathways mediated by monooxygenases, which incorporate oxygen into halogenated substrates, leading to unstable intermediates that spontaneously release halides. Bacterial monooxygenases, such as those in Xanthobacter autotrophicus, initiate the degradation of compounds like by forming epoxides or diols, facilitating further . This cometabolic process does not yield energy directly but aids in pollutant breakdown. A notable example of microbial dehalogenation is the bacterial degradation of lindane (γ-hexachlorocyclohexane) by Sphingomonas paucimobilis, where the enzyme LinA catalyzes the initial dehydrochlorination. LinA, a periplasmic dehalogenase, converts lindane to pentachlorocyclohexene through the elimination of HCl, initiating a multi-step pathway that mineralizes the pesticide. This enzyme's unique stereospecificity highlights the diversity of microbial adaptations to halogenated xenobiotics.

Applications in Pollution Remediation

Dehalogenation plays a crucial role in by facilitating the breakdown of persistent halogenated pollutants, such as chlorinated solvents including trichloroethene (TCE) and tetrachloroethene (PCE), polychlorinated biphenyls (PCBs), and organochlorine pesticides like and . These compounds, widely used in industrial and agricultural applications, contaminate and , posing significant risks due to their and . Reductive dehalogenation transforms these contaminants into less harmful products, such as ethene from chlorinated solvents or lower-chlorinated congeners from PCBs and pesticides, thereby reducing environmental . Key methods for applying dehalogenation in remediation include enhanced reductive dechlorination (ERD), which involves injecting electron donors like to stimulate microbial communities under conditions, promoting sequential removal of chlorine atoms from contaminants. Bioaugmentation complements ERD by introducing specialized bacteria, such as Dehalobacter species, to accelerate dehalogenation in sites lacking sufficient native populations, particularly for transforming higher-chlorinated compounds into benign end products. These approaches are often implemented to treat plumes in aquifers, minimizing excavation and costs. Field applications of dehalogenation have been documented at sites since the 1990s, with notable success in reducing PCE concentrations in through ERD and . For instance, at the Grants Chlorinated Solvents Plume site in , ERD barriers established in the early 2000s have contributed to ongoing dechlorination of PCE and TCE, as part of broader remediation efforts. Similar outcomes, including significant reductions in PCE concentrations following injections in the mid-2010s, have been reported at dry-cleaning sites treated with ERD. These case studies highlight dehalogenation's effectiveness in large-scale restoration. Despite successes, challenges persist, including incomplete dehalogenation that can stall at toxic intermediates like , which is more carcinogenic than parent compounds and requires careful management to avoid rebound contamination. Monitoring via quantitative (qPCR) for dehalogenating genes, such as those in Dehalococcoides or Dehalobacter, enables early detection of microbial activity and potential stalling, guiding adjustments like additional dosing. These issues underscore the need for site-specific optimization to ensure full mineralization. Recent advances include abiotic dehalogenation using nanoscale zero-valent iron (nZVI), which rapidly reduces chlorinated solvents and pesticides through direct , often integrated into permeable reactive barriers (PRBs) for passive treatment of flows. nZVI-PRB systems, deployed since the 2010s, have shown up to 99% removal of PCE and in pilot tests by forming less chlorinated byproducts, offering a robust alternative or with biological methods for recalcitrant sites. As of , emerging research emphasizes bio-abiotic strategies and omics-based optimizations to improve complete and address incomplete pathways in microbial dehalogenation.

References

  1. [1]
    Dehalogenation - an overview | ScienceDirect Topics
    Dehalogenation is defined as a chemical reaction that removes halogen atoms from vicinal dihalides, commonly used in the synthesis of alkenes and can involve ...
  2. [2]
  3. [3]
    [PDF] Ground Water Issue: Reductive Dehalogenation of Organic ...
    For this reason research is currently underway to better define the basic mechanisms of reductive dehalogenation reactions. Such approaches may include: (1) ...
  4. [4]
    Microbial reductive dehalogenation - PMC - NIH
    A wide variety of compounds can be biodegraded via reductive removal of halogen substituents. This process can degrade toxic pollutants.
  5. [5]
    Reductive Dehalogenation | SpringerLink
    Reductive dehalogenation is widely accepted as a major degradative mechanism for highly chlorinated compounds in the environment.
  6. [6]
    Enzymatic halogenation and dehalogenation reactions - NIH
    Dehalogenating enzymes, on the other hand, are best known for removing halogen atoms from man-made organohalogens, yet also function naturally, albeit rarely, ...
  7. [7]
  8. [8]
    Reductive Dechlorination - an overview | ScienceDirect Topics
    Reductive dechlorination is a chemical process removing chlorine from organic compounds, often by electron reduction, and can be a biodegradation process.
  9. [9]
    Dehalogenation reactions between halide salts and phosphate ...
    Dehalogenation processes are useful for waste form production of chloride-based salt wastes because it allows for much higher waste loadings in the final waste ...
  10. [10]
    Zinc Alkyls, Edward Frankland, and the Beginnings of Main-Group ...
    Edward Frankland's reaction of zinc and ethyl iodide produced the first main-group organometallic compounds, leading to the birth of organometallic chemistry.
  11. [11]
    The Grignard Reagents | Organometallics - ACS Publications
    The Grignard reagents probably have been the most widely used organometallic reagents. Most of them are easily prepared in ethereal solution.
  12. [12]
    An Essay on the History of Catalytic Hydrogenation of Organic ...
    Apr 27, 2022 · The essay considers historical aspects of the appearance and development of methods for catalytic hydrogenation of organic compounds.
  13. [13]
    Microbial dehalogenation | Biodegradation
    Hardman DJ & Slater JH (1981) Dehalogenases in soil bacteria. J. Gen ... —— (1960) Decomposition of chloroacetates and chloropropionates by bacteria.
  14. [14]
    Reductive Dehalogenation and Conversion of 2-Chlorophenol to 3 ...
    ABSTRACTBiotransformation of 2-chlorophenol by a methanogenic sediment community resulted in the transient accumulation of phenol and benzoate.
  15. [15]
    From Industrial Toxins to Worldwide Pollutants: A Brief History ... - NIH
    Sep 28, 2018 · This article seeks to explore the history of knowledge about the toxicity of PCBs to elucidate the relationship between occupational and environmental health.
  16. [16]
    1.24: Nucleophilic Substitution, SN2, SN1 - Chemistry LibreTexts
    Sep 12, 2022 · These reactions are known as Nucleophilic Substitution Reactions, substitution reactions because one atom or group has been substituted for another.
  17. [17]
    E1 Reactions - Chemistry LibreTexts
    Jan 22, 2023 · Unimolecular Elimination (E1) is a reaction in which the removal of an HX substituent results in the formation of a double bond.Missing: eliminative dehalogenation
  18. [18]
    Carbon Dichloride as an Intermediate in the Basic Hydrolysis of ...
    Carbon dichloride as an intermediate in the basic hydrolysis of chloroform. A mechanism for substitution reactions at a saturated carbon atom.
  19. [19]
    Leaving Groups - Chemistry LibreTexts
    Jan 22, 2023 · In order for a leaving group to leave, it must be able to accept electrons. A strong bases wants to donate electrons; therefore, the leaving ...Missing: Cl sources
  20. [20]
    Covalent Bond Energies - gchem
    Single Bond Energies (kJ/mol). H, C, N, O, S, F, Cl, Br, I. H, 436. C, 413, 346. N, 391, 305, 163. O, 463, 358, 201, 146. S, 347, 272, -, -, 226. F, 565, 485 ...Missing: CI | Show results with:CI
  21. [21]
    [PDF] Bond dissociation energies in simple molecules
    Bond Dissociation Energies in Simple Molecules. B. deB. Darwent. Bond dissociation energy values (kcal/mol) and (kj/mol) of simple compounds are tabulated.
  22. [22]
    [PDF] Chemistry 471/671
    1) The average bond dissociation energies of C-F, C-Cl, C-Br and C-I are 473, 347, 293 and 238 kJ mol-1, respectively. a. Calculate the maximum wavelength of ...
  23. [23]
    [PDF] Bond Dissociation Energies
    TABLE 4.11 Bond Dissociation Energies (Continued). Bond. ΔHf298,. kJ/mol. Bond. ΔHf298,. kJ/mol. Beryllium (continued). BeCl9Cl. 540(63). Be9F. 577(42). Be9H.
  24. [24]
    [PDF] Theoretical Study of the C−Cl Bond Dissociation Enthalpy and ...
    Jun 27, 2009 · C−Cl bond in substituted chlorobenzene compounds is discussed ... rate C−Cl bond dissociation energy of 393.59 kJ/mol, only a 1.92 kJ ...
  25. [25]
    Elimination Reactions Are Favored By Heat
    Sep 10, 2012 · At some point, as temperature is increased, the ΔG for elimination will become more negative than delta G for substitution. In other words, more ...
  26. [26]
    The solvent effect on the electrocatalytic cleavage of carbon ...
    An exhaustive study has been carried out on the effect of solvent on the reductive cleavage of carbon-halogen bonds for two aromatic and two alkyl halides, ...
  27. [27]
    Medium Effects on the Reductive Cleavage of the Carbon−Halogen ...
    In the gas phase, the reaction driving force is determined by the difference in the dissociation energy of the C−X bond and the electron affinity of the halide ...
  28. [28]
    How Alkyl Halide Structure Affects E2 and SN2 Reaction Barriers
    Enthalpy vs Free Energy, Solvent Effects, and Choice of Base and Nucleophile. All of the patterns of reactivity described above were deduced from free energy ...
  29. [29]
    Nucleophilic Substitution in Vinylic Halides - Nature
    Nucleophilic substitution of a halogen atom directly attached to a double bond, as in the vinyl and phenyl halides, does not readily take place.
  30. [30]
    The dehalogenation reaction of organic halides by tributyltin radical
    The energy of activation (Ea) of the dehalogenation reaction induced by Bu3Sn radical dot was measured for some aryl, alkyl and benzyl halides.Missing: halogens | Show results with:halogens
  31. [31]
    Protic-dipolar aprotic solvent effects on rates of bimolecular reactions
    Modeling Protic to Dipolar Aprotic Solvent Rate Acceleration and Leaving Group Effects in SN2 Reactions: A Theoretical Study of the Reaction of Acetate Ion ...
  32. [32]
    Halogenated Organic Pollutant Degradation Driven by Fe(II) Redox ...
    Oct 21, 2025 · (103−106) In addition, pH is often the main factor affecting the reductive reactivity of various Fe(II) species, (40,66,95,107,108) and the ...
  33. [33]
    SN2 Mechanism - an overview | ScienceDirect Topics
    We attribute this order of reactivity to steric hindrance. Adding alkyl groups to the carbon atom of the carbon–halogen bond shields the carbon atom from attack ...
  34. [34]
    [PDF] Kinetic Isotope Effects in Organic Chemistry - Macmillan Group
    Sep 14, 2005 · This KIE is consistent with an E2 elimination in which the C - H/D bond is broken in the rate determining step. □ This KIE is consistent with an ...
  35. [35]
    Metal-Mediated Reductive Hydrodehalogenation of Organic Halides
    The metal-mediated hydrodehalogenation reaction is of paramount importance as a reducing method for the appropriate treatment of these chemicals.
  36. [36]
    Methods for the Preparation of Allenes Employing Indium- and Zinc ...
    Sep 7, 2011 · Simple and mild methods for the synthesis of allenes, employing indium- and zinc-mediated dehalogenation reactions of vicinal dihalides in an aqueous solvent, ...Missing: synthesis | Show results with:synthesis
  37. [37]
    Sterols. III. A Method for the Dehalogenation of Steroids
    A. S. K. Hashmi. Chromium(II) chloride: A reagent for chemo-, regio- and diastereoselective C?C-bond formation.Missing: total | Show results with:total
  38. [38]
    [PDF] Development of New Tritium Labelling Methods for Peptides ...
    Protecting Group and Coupling Strategies in Peptide Synthesis ... We first investigated the palladium mediated dehalogenation of the iodinated peptide using.
  39. [39]
    Pd-catalysed hydrodehalogenation of aryl chlorides: a mild method ...
    Here we report on the hydrodehalogenation of a series of chloro- and polyhalogenated compounds under mild reaction conditions using palladium complexes.
  40. [40]
    Expanding the scope of silane-mediated hydrodehalogenation ...
    A palladium-catalysed, silane-mediated hydrodehalogenation (HDH) reaction with increased substrate scope has been developed. Whereas previous attempts to ...
  41. [41]
    [PDF] efficient-transfer-hydrodehalogenation-of-halophenols-catalyzed-by ...
    ABSTRACT: The transfer hydrodehalogenation (THD) of halophenols is efficiently catalyzed by palladium supported on high surface ceria (Pd/CeO2) under mild ...
  42. [42]
    Nickel-catalyzed hydrodehalogenation of aryl halides - ScienceDirect
    Apr 1, 2013 · In the present study, the nickel-catalyzed dehalogenation of aryl and alkyl halides with iso-propyl zinc bromide or tert-butylmagnesium chloride ...
  43. [43]
    Applicability of Nickel-Based Catalytic Systems for ... - MDPI
    This review summarizes recent applications of nickel as a nonprecious metal catalyst in hydrodehalogenation (HDH) reactions of halogenated aromatic compounds ( ...
  44. [44]
    Practical iron-catalyzed dehalogenation of aryl halides
    An operationally simple iron -catalyzed hydrodehalogenation of aryl halides has been developed with 1 mol% Fe(acac)3 and commercial t-BuMgCl as reductant.
  45. [45]
    Pd/C Catalyzed Dehalogenation of (Hetero)aryls ... - Thieme Connect
    We found that the combination of Pd/C as catalyst and triethylsilane (TES) as hydrogen donor in THF resulted in a broadly applicable, easy to set up, and ...
  46. [46]
    Organohalide Respiring Bacteria and Reductive Dehalogenases
    This review focuses on RDases, concentrating on those which have been purified (partially or wholly) and functionally characterized. Further, the paper reviews ...
  47. [47]
    Dehalogenases: From Improved Performance to Potential Microbial ...
    Dehalogenases such as the haloalkane dehalogenases (HLDs) also play a role in decontamination of warfare agent such as sulphur mustard. With suitable HLDs ...
  48. [48]
    Haloalkane dehalogenases: Biotechnological applications
    Sep 11, 2012 · This review discusses the application of HLDs in the context of the biochemical properties of individual enzymes. Further extension of HLD uses ...
  49. [49]
    Reductive Dehalogenases Come of Age in Biological Destruction of ...
    A summary of characterized and identified reductive dehalogenases is reported in this review, revealing the diversity of substrates and preferred electron ...Review · Transcriptional Regulation... · Corrinoid Cofactors And...
  50. [50]
    A review in the current developments of genus Dehalococcoides, its ...
    This article reviews the current developments in genus Dehalococcoides as key dechlorinating bacteria in chlorinated ethene contaminated sites.
  51. [51]
    Complete Detoxification of Vinyl Chloride by an Anaerobic ... - NIH
    The 16S rRNA gene-based analysis implicated a Dehalococcoides population in VC reductive dechlorination, a finding that was supported by physiological evidence ...
  52. [52]
    Role of Dehalogenases in Aerobic Bacterial Degradation of ...
    Nov 13, 2014 · This review was conducted to provide an overview of dehalogenases involved in aerobic biodegradation of chlorinated aromatic compounds.
  53. [53]
    Monooxygenase-Mediated 1,2-Dichloroethane Degradation by ...
    We concluded that a monooxygenase is responsible for the first step in DCA degradation in strain DCA1. Oxidation of DCA probably results in the formation of the ...
  54. [54]
    Two Different Types of Dehalogenases, LinA and LinB, Involved in γ ...
    Among them, linA encodes an enzyme that catalyzes a unique dehydrochlorination reaction whose mechanism is still unknown. Dehalogenation is a key step in the ...
  55. [55]
    Novel LinA Type 3 δ-Hexachlorocyclohexane Dehydrochlorinase
    Oct 2, 2015 · LinA is the first enzyme of the microbial degradation pathway of a chlorinated insecticide, hexachlorocyclohexane (HCH), and mediates the ...
  56. [56]
    Reductive dechlorination for remediation of polychlorinated biphenyls
    Reductive methods for PCB remediation published in the last decade are reviewed. Catalytic HDC with H 2 and Fe-based reductive dechlorination are major methods.Missing: dehalogenation | Show results with:dehalogenation
  57. [57]
    Recent Strategies for Environmental Remediation of Organochlorine ...
    Shoifu et al. [42] used zero-valence iron and iron oxide to destroy organochlorine pesticides (aldrin, lindane and DDT) from water. The products obtained were ...
  58. [58]
    [PDF] Enhanced Reduction Dechlorination (ERD) and In Situ Chemical ...
    Enhanced Reduction Dechlorination (ERD) and. In Situ Chemical Reduction ... Common Electron Donors. • Biotic (organic). – Lactate. – Whey. – Emulsified ...
  59. [59]
    Growth of Dehalobacter and Dehalococcoides spp. during ... - NIH
    During dechlorination, both Dehalobacter and Dehalococcoides 16S rRNA gene copies increased 1 to 2 orders of magnitude, although there was considerable ...
  60. [60]
    [PDF] Grants Chlorinated Solvents Plume, Superfund Site ... - US EPA
    Enhanced reductive dechlorination is planned as a polishing step for the ISCO. • Enhanced reductive dechlorination (ERD) barriers in the shallow plume periphery ...
  61. [61]
    Enhanced reductive dechlorination showing positive returns at ...
    Case study highlights. Combined remedies approach using an Enhanced Reductive Dechlorination (ERD) process successfully remediated an Indiana drycleaning site ...
  62. [62]
    Microbial community response to a bioaugmentation test to degrade ...
    However, slow or incomplete biodegradation can lead to stalling at degradation byproducts such as 1,2-dichloroethene (cis-DCE) and vinyl chloride (VC). Over ...
  63. [63]
    Quantitative PCR Confirms Purity of Strain GT, a Novel ... - NIH
    ... incomplete microbial dechlorination of tetrachloro ethene (PCE) and trichloroethene (TCE) leads to dichloroethene and vinyl chloride (VG) accumulation. A ...
  64. [64]
    [PDF] Nanoscale zero valent iron and bimetallic particles for contaminated ...
    Feb 23, 2012 · The goal of this review is to present the current state of knowl- edge related to the use of nZVI/bimetallic nanometals for the in situ.
  65. [65]
    Nanoscale Metallic Iron for Environmental Remediation - NIH
    The use of metallic iron (Fe0) in subsurface reactive permeable barriers has been identified as such a technology (Gillham and O'Hannesin 1991, 1994; Matheson ...