Fact-checked by Grok 2 weeks ago

Dehydrohalogenation

Dehydrohalogenation is an in in which a and a atom are removed from adjacent carbon atoms in an alkyl halide substrate, typically forming an product and a byproduct. This β-elimination process involves the loss of HX (where X is Cl, Br, or I) from the α-carbon bearing the halogen and the β-carbon providing the . The reaction is most commonly promoted by treatment with a strong base, such as an alkoxide ion (RO⁻) or hydroxide ion (HO⁻), under conditions that favor elimination over substitution. Dehydrohalogenation can proceed via two primary mechanisms: the concerted, bimolecular E2 pathway, which is favored by strong bases and secondary or tertiary alkyl halides, and the stepwise, unimolecular E1 pathway, which involves carbocation intermediates and is promoted by weaker bases or polar protic solvents. In the E2 mechanism, bond breaking and formation occur simultaneously in a single transition state, leading to anti-periplanar geometry in the transition state for optimal orbital overlap. This reaction holds significant importance in synthetic as one of the primary methods for preparing s from readily available alkyl halides, enabling the construction of carbon-carbon double bonds essential for pharmaceuticals, materials, and synthesis. The regioselectivity follows Zaitsev's rule, favoring the more substituted (thermodynamically stable) , though conditions like bulky bases can promote the less substituted Hofmann product. Competition with reactions (SN1 or SN2) is common, particularly with primary alkyl halides, necessitating careful control of base strength, solvent, and temperature to optimize yields.

General Principles

Definition and Scope

Dehydrohalogenation is an in which a (HX, where X denotes a such as , , or iodine) is removed from a , typically yielding unsaturated compounds like alkenes, alkynes, or other functionalities with carbon-carbon multiple bonds. This process is fundamental in synthetic chemistry for introducing unsaturation into molecular frameworks. The scope of dehydrohalogenation extends across and inorganic domains. In contexts, it primarily involves substrates such as alkyl halides and halides, enabling the conversion of saturated to unsaturated hydrocarbons. In , the reaction applies to coordination compounds, where it facilitates ligand modifications or metal-hydride formations, and to ionic salts, often under basic conditions to generate reactive intermediates. Unlike , which eliminates (H₂O) from alcohols or similar compounds to form alkenes, dehydrohalogenation specifically targets HX removal, preserving distinct mechanistic and substrate requirements. Historically, dehydrohalogenation was first systematically explored in the 1850s using alcoholic KOH on alkyl bromides, with pioneering contributions from chemists like Alexander Butlerov, whose structural theories influenced subsequent developments. This early work laid the groundwork for rules, such as Zaitsev's rule established in the 1870s by Butlerov's student Aleksandr Zaitsev, and has since evolved into versatile modern synthetic strategies. A key prerequisite for the reaction is the availability of a β-hydrogen atom relative to the halogen-bearing carbon, ensuring proper alignment for elimination and underscoring its utility in for controlled unsaturation. Mechanisms often proceed via concerted E2 or carbocation-mediated E1 pathways, depending on conditions.

Reaction Mechanisms

Dehydrohalogenation reactions primarily proceed via two fundamental mechanisms: the bimolecular E2 elimination and the unimolecular E1 elimination. The E2 mechanism is a concerted, one-step process involving the simultaneous abstraction of a β-hydrogen by a base and departure of the halide , resulting in the formation of an . This pathway exhibits second-order , with the rate depending on both the substrate and base concentrations: rate = k[RX][B⁻]. In the , partial bonds form between the base and hydrogen, and between the α- and β-carbons, while the C-H and C-X bonds weaken concurrently. The general equation for the E2 mechanism is: \text{R-CH}_2\text{-CHX-R' + B: } \rightarrow \text{R-CH=CH-R' + BH}^{+} + \text{X}^{-} where B: represents the base, X is the halogen, and R/R' are alkyl substituents. This mechanism predominates under conditions favoring bimolecular processes, such as with strong bases. In contrast, the E1 mechanism is a stepwise, two-stage process initiated by the unimolecular ionization of the alkyl halide to form a carbocation intermediate, followed by deprotonation of a β-hydrogen. It follows first-order kinetics, rate = k[RX], independent of base concentration, and is common in polar protic solvents where carbocation stabilization occurs. Rearrangements, such as hydride or alkyl shifts, can arise due to the planar, achiral nature of the carbocation. A typical scheme for a tertiary halide is: \text{(CH}_3\text{)}_3\text{C-X } \rightarrow \text{ (CH}_3\text{)}_3\text{C}^{+} + \text{X}^{-} \rightarrow \text{ (CH}_3\text{)}_2\text{C=CH}_2 + \text{HX} This pathway is typical for halides or under conditions with weak . The choice between E2 and E1 mechanisms is influenced by several factors, including base strength, polarity, , and structure. Strong, non-bulky like ethoxide promote E2, while weak or nucleophiles favor E1 in polar that stabilize ions. Primary halides typically undergo E2 due to unfavorable formation, whereas halides lean toward E1. Higher generally enhance elimination over , shifting equilibrium toward E1 or E2. Stereochemistry plays a critical role, particularly in E2 eliminations, which require anti-periplanar between the β-hydrogen and for optimal orbital overlap in the . This leads to anti elimination, producing specific stereoisomers; syn elimination is rare and occurs only under constrained conditions. In cyclic systems, such as cyclohexyl halides, the diaxial conformation is necessary, enforcing —cis isomers react faster than trans due to easier attainment of the required . E1 eliminations lack such , as the allows attack from multiple directions. Regioselectivity in both mechanisms is governed by Zaitsev's rule, which predicts the major product as the more substituted (thermodynamically stable) , due to greater and inductive stabilization in the for E2 or the for E1. For instance, dehydrohalogenation of yields predominantly 2-butene over . Bulky bases can invert this, favoring Hofmann products (less substituted alkenes) via steric hindrance. Kinetic isotope effects provide mechanistic insight, particularly for E2, where replacing with at the β-position results in a primary isotope effect (k_H/k_D ≈ 6-7), indicating C-H bond cleavage in the rate-determining step. This arises from the higher of C-H bonds compared to C-D, lowering the difference (ΔE_a ≈ 5 kJ/mol) and slowing deuterated reactions. In E1, secondary isotope effects are observed, as follows formation.

Dehydrohalogenation of Alkyl and Vinyl Halides

Formation of Alkenes

Dehydrohalogenation of alkyl halides serves as a key method for synthesizing alkenes through base-promoted elimination of (HX). This process involves treating alkyl halides with strong bases in alcoholic solvents to favor the bimolecular E2 pathway, which is concerted and requires anti-periplanar alignment of the leaving groups. Typical conditions include alcoholic (KOH) or (NaOEt) at temperatures of 60–100°C, where the solvent reduces the base's nucleophilicity to minimize competing reactions. A representative for the is the dehydrohalogenation of : \ce{CH3-CHBr-CH2-CH3 + ^-OH ->[alcoholic KOH, 60-100°C] CH3-CH=CH-CH3 + H2O + Br^-} The product, 2-butene, forms as a mixture of (E)- and (Z)-, with the thermodynamically more stable (E)- predominating in equilibrated conditions. This elimination illustrates the scope for secondary alkyl halides, where the efficiently generates internal alkenes. Regioselectivity in alkene formation follows Zaitsev's rule, which predicts that the major product is the more highly substituted due to its greater thermodynamic stability from and inductive effects. For instance, treatment of with alcoholic KOH yields 2-butene as the major product (approximately 80%) and as the minor product (20%). However, using bulky bases such as tert-butoxide (t-BuOK) in tert-butanol shifts selectivity toward the less substituted Hofmann product, as steric bulk hinders approach to the more crowded β-hydrogen on the carbon adjacent to more alkyl substituents. This allows synthetic control over product distribution in cases with multiple β-hydrogens. The reaction's scope encompasses primary, secondary, and alkyl halides, with reactivity increasing in that order owing to enhanced transition-state stabilization by alkyl groups. Primary halides predominantly undergo (SN2) as a side reaction under these conditions, while secondary and tertiary halides favor elimination, though ones may proceed via competing E1 pathways if the base is weak. To optimize yields, conditions are adjusted to suppress substitution, such as using concentrated bases like NaOH or KOH in , which enhance elimination ratios to over 90% for secondary halides. Industrially, this method contributes to production, particularly for targeted syntheses where alkyl halides are readily available precursors.

Formation of Alkynes

The formation of alkynes through dehydrohalogenation involves the double elimination of from vicinal dihalides (where are on adjacent carbons) or geminal dihalides (where both are on the same carbon). This process proceeds via two successive E2 elimination reactions, the first generating a intermediate and the second yielding the . The reaction requires excess strong base, typically (NaNH₂) in liquid , to drive both eliminations to completion, as the intermediate is less reactive than alkyl halides due to the sp²-hybridized carbon. A representative equation for the conversion of (a vicinal dihalide) to illustrates the process: \mathrm{BrCH_2CH_2Br + 2 NaNH_2 \rightarrow HC \equiv CH + 2 NaBr + 2 NH_3} For dihalides, such as 1,1-dibromopropane (CH₃CH₂CHBr₂), treatment with excess NaNH₂ yields (CH₃C≡CH). In cases forming terminal s, an additional equivalent of base is often needed because the product is deprotonated by NaNH₂ to form the acetylide anion, which must be neutralized (e.g., via aqueous ) to isolate the neutral alkyne. This method is particularly valuable for laboratory synthesis of both and internal alkynes from readily available dihalides, offering a controlled route to these compounds that historically supported early acetylene production and remains key in for pharmaceuticals and materials.

Thermal Processes

Thermal dehydrohalogenation refers to the pyrolysis of alkyl and vinyl halides at high temperatures, typically 400–800 °C, to eliminate hydrogen halide and form unsaturated hydrocarbons without requiring a base. This process operates through a unimolecular molecular elimination mechanism involving a concerted four-centered cyclic transition state, where the C–H and C–X bonds break simultaneously to yield an alkene and HX. Unlike base-promoted eliminations, thermal processes favor this non-ionic pathway due to the gas-phase conditions and elevated temperatures that promote homolytic or heterolytic bond cleavage. A classic example is the gas-phase of ethyl chloride, which decomposes to and : \ce{CH3CH2Cl ->[~600^\circ\mathrm{C}] CH2=CH2 + HCl} This reaction follows first-order , with rates independent of pressure down to 0.2 mm Hg and temperatures ranging from 402–521 °C, confirming the unimolecular nature and minimal surface involvement. Similar thermal decompositions apply to other primary alkyl chlorides and bromides, yielding corresponding alkenes cleanly under controlled conditions. Industrially, thermal dehydrohalogenation is central to vinyl chloride monomer (VCM) production, where 1,2-dichloroethane (EDC) undergoes in tubular furnaces at around 500 °C to produce VCM and HCl, with conversions of about 50% and selectivities over 98%. This contrasts with catalytic cracking in refining, which uses acid catalysts at lower temperatures (450–550 °C) to break C–C bonds in hydrocarbons for olefin production, whereas thermal halide cracking relies on heat alone for C–H and C–X elimination, often integrated into balanced ethylene-chlorine processes. Despite its utility, thermal dehydrohalogenation demands high energy inputs to achieve activation energies of 40–60 kcal/mol, leading to operational costs and equipment stress. Side products, including , , and higher hydrocarbons from chain reactions or over-cracking, reduce yields and necessitate frequent maintenance, such as decoking. In comparison to base-mediated methods, thermal processes offer lower selectivity for specific alkenes, as secondary pathways and multiple β-hydrogen options produce a wider product distribution, though they avoid salt byproduct issues.

Dehydrohalogenation in Other Organic Reactions

Epoxide Synthesis

Dehydrohalogenation of s provides a versatile route to s through base-promoted intramolecular elimination of from β-halo alcohols, forming a three-membered oxirane ring. This process is particularly valuable in due to its ability to generate epoxides under mild conditions from readily available halohydrin precursors, which are often prepared via to alkenes. Common bases such as or facilitate the reaction by deprotonating the hydroxyl group, enabling nucleophilic displacement. A representative example is the conversion of 1-chloropropan-2-ol to propylene oxide: \ce{Cl-CH2-CH(OH)-CH3 + OH- ->[base] \overset{\ce{CH3}}{\underset{|}{\ce{CH-CH2}}} + H2O + Cl-} In this reaction, the base generates the alkoxide from the alcohol, which then undergoes intramolecular nucleophilic attack on the carbon bearing the chlorine, displacing chloride and closing the ring. The mechanism proceeds via an SN2-like pathway at the carbon-halogen bond, requiring anti-periplanar alignment of the oxygen and halogen for efficient backside attack. This intramolecular elimination contrasts with intermolecular dehydrohalogenations by favoring ring formation over alkene production. The of the transformation retains the configuration at both carbon centers involved, preserving the relative spatial arrangement of substituents from the precursor. This outcome arises because the SN2 inversion at the halogen-bearing carbon effectively mirrors the prior anti addition in formation, resulting in overall syn for the relative to the original . For instance, yield racemic , while produce meso . Suitable substrates include chlorohydrins and bromohydrins, with primary halides preferred to minimize elimination side products and ensure clean SN2 reactivity; secondary or halides may lead to competing E2 pathways. In unsymmetrical cases, favors attack at the less substituted carbon-halogen bond, directing the epoxide oxygen toward the more substituted position. Industrially, this method was pivotal in the early production of via dehydrochlorination of ethylene chlorohydrin with (Ca(OH)2) or NaOH, a process commercialized in and used until the 1930s when direct oxidation supplanted it, though it remains relevant for synthesis. Epoxides produced via this dehydrohalogenation serve as critical intermediates in the manufacture of pharmaceuticals, such as antiviral drugs and beta-blockers, and polymers like polyurethanes and , where the enables regioselective ring-opening reactions. The method's synthetic importance lies in its compatibility with chiral halohydrins for enantioselective epoxide synthesis, enhancing its utility in asymmetric synthesis.

Isocyanide Synthesis

The , also known as the Hofmann isocyanide synthesis, is a classic method for producing (R-NC) through the treatment of primary amines with and a , involving multiple dehydrohalogenation steps. The overall reaction can be represented as: \mathrm{RNH_2 + CHCl_3 + 3\, KOH \to R-NC + 3\, KCl + 3\, H_2O} This process was first reported by August Wilhelm von Hofmann in 1867, who observed the formation of foul-smelling from and under basic conditions. The reaction's discovery highlighted the unique reactivity of haloforms and laid the foundation for isocyanide chemistry in . The reaction is typically conducted using alcoholic KOH as the base, with heating to , often in a well-ventilated setup due to the characteristic foul odor of the products. It is specific to primary , as secondary and tertiary do not yield isocyanides under these conditions, making it a valuable qualitative test for distinguishing primary in . For example, the test involves heating a small sample of the amine with and ethanolic KOH; a positive result is indicated by the evolution of a pungent, unpleasant smell. Mechanistically, the reaction proceeds via the generation of dichlorocarbene as a key , followed by two successive dehydrohalogenations. The first step is the base-promoted dehydrohalogenation of : \mathrm{CHCl_3 + KOH \to :CCl_2 + KCl + H_2O} The electrophilic dichlorocarbene then adds to the nucleophilic of the primary , forming an intermediate dichloromethylamine: \mathrm{:CCl_2 + RNH_2 \to RNH-CHCl_2} Subsequent dehydrochlorination steps, facilitated by excess , eliminate two equivalents of HCl to yield the : \begin{align*} &\mathrm{RNH-CHCl_2 + KOH \to RN=CHCl + KCl + H_2O} \\ &\mathrm{RN=CHCl + KOH \to R-NC + KCl + H_2O} \end{align*} This sequence underscores the role of dehydrohalogenation in both generation and the transformation of the adduct to the final product.

Dehydrohalogenation in

Coordination Compounds

In coordination compounds, beta-dehydrohalogenation refers to elimination processes in organometallic complexes where a and are removed from adjacent carbons in an alkyl , often facilitated by the metal center. A key variant is beta-hydride elimination, where a from the migrates to the metal, forming a metal-hydride species and an , with the potentially influencing the pathway through competition or assistance. This process is common in d^8 square-planar complexes of and , requiring a vacant coordination site for the migration to occur. The typically begins with reversible of a (e.g., ) to generate a three-coordinate , followed by rate-determining beta-hydride transfer in a syn fashion, yielding a metal-hydride- that can further dissociate. In cases involving beta-halogen substituents, such as in alkyl-metal halides like M-CH2-CH2-X, the elimination can proceed via beta-hydride migration to form M-H + CH2=CH2 + X^-, or compete with beta-X elimination to give M-X + CH2=CH2, depending on ligand electronics and sterics. For square-planar complexes, the process is often irreversible under thermal conditions due to the stability of the resulting . Similar behavior occurs in [Pt(PPh3)2(CH3)(CH2CH(R))]^+ or alkyl complexes, which decompose at 95°C in to form or propene via beta-hydride elimination, with rates influenced by the cis and beta substitution (e.g., k_H/k_D = 3.2). In palladium systems, beta-elimination is exemplified in (beta-haloalkyl)palladium intermediates during , where electron-rich ligands like PPh3 favor beta-X over beta-H pathways at . These eliminations typically require thermal conditions (e.g., 70-100°C) or ligand assistance to modulate rates and prevent premature decomposition; added phosphines (7.5-15 mM) stabilize complexes and control reversibility. In , such as olefin initiators, beta-hydride elimination serves as a chain-transfer step, limiting molecular weight, while in cross-coupling reactions like the Heck or processes, it is essential for product release, with beta-X variants enabling selective C(sp3)-C(sp2) bond formation by avoiding unwanted alkenes. The general equation for beta-hydride elimination in a square-planar complex is: \ce{M-CH2-CH3 ->[thermal] M-H + CH2=CH2} This differs from E2 dehydrohalogenation, which requires a strong and is concerted with anti-periplanar geometry; in metal-mediated versions, the dictates (often syn), and the process is intramolecular without external nucleophiles, enabling reversibility and integration into catalytic cycles.

Ionic and Organometallic Systems

In ionic compounds, dehydrohalogenation reactions often occur reversibly in salts featuring acidic cations hydrogen-bonded to halometallate anions, such as protonated amines paired with tetrahalometallate species. These processes involve the cleavage of N–H (or analogous C–H) bonds in the cation and M–X bonds in the anion, liberating HX gas and yielding a neutral base alongside a neutral metallate fragment. A representative equilibrium is depicted as [R₃NH]⁺ [MX₄]⁻ ⇌ R₃N + HX + MX₃, where R is an and M is a metal like Al or a , with the reaction driven by thermal or mechanochemical stimuli in solid-state or melt conditions. For instance, in chloroaluminate systems like triethylammonium tetrachloroaluminate ([Et₃NH]⁺ [AlCl₄]⁻), the equilibrium shifts toward dehydrohalogenation under heating, releasing HCl and forming Et₃N + AlCl₃, reflecting the tunable acidity of these melts based on AlCl₃ . Hydrogen-bonded halometallates exemplify this behavior, where the second-sphere coordination between the protonated cation (e.g., [B–H]⁺) and the halometallate anion ([X–MXₙ]⁻) facilitates elimination: [B–H]⁺ ··· [X–MXₙ]⁻ ⇌ B + HX + MXₙ. These reactions are reversible through gas-solid of HX, even in non-porous materials, enabling dynamic structural transformations from outer-sphere adducts to inner-sphere coordination complexes involving metals like Cu(II), Zn(II), Co(II), Pt(II), Pd(II), or Hg(II) with halides or . The reversibility stems from the weak hydrogen bonding and the thermodynamic favorability of HX formation, with activation energies lowered by the proximity of the donor and acceptor sites; for example, treatment at moderate temperatures (around 100–200 °C) induces dehydrohalogenation, while exposure to HX vapor promotes the reverse . In organometallic contexts beyond discrete coordination complexes, dehydrohalogenation manifests through thermal elimination processes, notably β-hydride elimination in Grignard reagents (RMgX). Under heat, these species undergo where a β-hydrogen from the R group migrates to the magnesium, forming an and HMgX: e.g., EtMgCl → C₂H₄ + HMgCl. This pathway dominates thermal stability limits, with decomposition onset around 100–150 °C depending on the alkyl chain, contrasting with oxidative or hydrolytic routes. Sigma-bond metathesis can also contribute in mixed systems, exchanging ligands without changes, though elimination prevails in pure Grignard heating. These bulk ionic or melt-like behaviors differ from molecular coordination compounds by emphasizing equilibrium-driven equilibria in extended lattices rather than isolated ligand-metal interactions. Such reversible dehydrohalogenations find applications in precursor synthesis for coordination compounds, where hydrogen-bonded salts serve as starting materials for mechanochemical or thermal conversion to first-sphere complexes, offering solvent-free routes with high . The thermodynamics favor reversibility due to comparable bond strengths (N–H/M–X ~ 100–150 kJ/mol vs. H–X ~ 400 kJ/mol, balanced by from gas evolution), enabling control over reaction direction via temperature or pressure. While primarily synthetic, these processes inform designs for materials by to reversible HX uptake/release in metal-halide frameworks, though direct implementations remain exploratory.

References

  1. [1]
    Illustrated Glossary of Organic Chemistry - Dehydrohalogenation
    Dehydrohalogenation: A reaction in which a hydrogen atom and a halogen atom are removed from adjacent atoms in a molecule, forming (usually) an alkene or an ...
  2. [2]
  3. [3]
    7.6.1. Elimination of Alkyl Halides - Chemistry LibreTexts
    Jun 5, 2019 · The most common elimination reactions are dehydrohalogenation and dehydration. In the mechanism above, X could be Cl, Br, or I for the ...
  4. [4]
    [PDF] Elimination Reactions
    The most common mechanism for dehydrohalogenation is the E2 mechanism. The reaction is concerted—all bonds are broken and formed in a single step. E2 reactions ...<|control11|><|separator|>
  5. [5]
    [PDF] Chapter 6 Ionic Reactions-Nucleophilic Substitution and Elimination ...
    Mechanism of Dehydrohalogenation: the E2 Reaction. The reaction of isopropyl bromide with sodium ethoxide shows second order kinetics. Page 55. Chapter 6. 55.
  6. [6]
    [PDF] Alkyl Halides
    The elimination of H-X is common, and is called a dehydrohalogenation. Often substitution and elimination reactions will occur in competition with each other.
  7. [7]
    Elimination Reactions – Organic Chemistry
    In dehydrohalogenation, it is the hydrogen atom on β-carbon that is eliminated together with halogen, as HX, therefore the reaction is often called as β- ...Missing: definition | Show results with:definition
  8. [8]
    Dehydrohalogenation reactions in second-sphere coordination ...
    This is our attempt to analyze dehydrohalogenation/hydrohalogenation reactions viewed as transformations from the second sphere coordination to first sphere ...Missing: inorganic | Show results with:inorganic
  9. [9]
    Deprotonation, Chloride Abstraction, and Dehydrohalogenation as ...
    Jul 17, 2017 · The process of removal of protons and chloride, dehydrohalogenation, from [(H2L)FeCl2] is investigated systematically, to understand the ...
  10. [10]
    Elimination Reactions of Alkyl Halides - MSU chemistry
    Kinetic studies of these reactions show that they are both second order (first order in R–Br and first order in Nu:(–)), suggesting a bimolecular mechanism for ...
  11. [11]
  12. [12]
    None
    ### Summary of E2 Mechanism, Stereochemistry, Zaitsev's Rule, E1 vs E2 Factors, and Kinetic Isotope Effects (Dehydrohalogenation Focus)
  13. [13]
    NS14. Stereochemistry in Elimination - csbsju
    An E2 elimination can only happen if a hydrogen and a neighbouring leaving group are anti to each other in a chair. These two groups must be trans to each other ...
  14. [14]
  15. [15]
    [PDF] TOPIC 7. ELIMINATION REACTIONS
    Use of bulky bases can lead to formation of “anti-Zaitsev” products. Br t-BuOK ... Formation of “anti-Zaitsev” products via the Hofmann Elimination.
  16. [16]
    [PDF] PRACTICE EXERCISE Elimination Reactions and Alkene Synthesis
    The molecule must rotate around the central cabon-carbon bond to aquire the anticoplanar arrangement required for E2. This is a stereospecific reaction that ...
  17. [17]
    Ch 5: Dehydrohalogenation - University of Calgary
    When heated with strong bases, alkyl halides typically undergo a 1,2-elimination reactions to generate alkenes. · Typical bases are NaOH or KOH or NaOR or KOR ( ...Missing: ionic salts inorganic
  18. [18]
    What methods are used to manufacture alkenes? - TutorChase
    Alkenes are primarily manufactured through the processes of cracking, dehydration of alcohols, and dehydrohalogenation of alkyl halides.<|control11|><|separator|>
  19. [19]
    Alkenes To Alkynes Via Halogenation And Elimination Reactions
    Jun 11, 2013 · -elimination of vicinal dihalides with nanh2 gives alkenyl halides further elimination gives alkynes 3 equiv. In the first step, NaNH2 is the ...
  20. [20]
    9.2 Preparation of Alkynes: Elimination Reactions of Dihalides
    Sep 20, 2023 · Treatment of a 1,2-dihaloalkane (called a vicinal dihalide) with an excess amount of a strong base such as KOH or NaNH2 results in a twofold ...
  21. [21]
    Reagent Friday: Sodium Amide (NaNH2) - Master Organic Chemistry
    Apr 15, 2025 · In conversion of Alkynes from Vicinal Dihalides we use Alcoholic Potash (KOH) as Reagent in first step and then we use Sodamide (NANH2) in ...
  22. [22]
    Heterolysis and the Pyrolysis of Alkyl Halides in the Gas Phase
    Exploring the reaction mechanisms of fast pyrolysis of Xylan model compounds via tandem mass spectrometry and quantum chemical calculations.Missing: dehydrohalogenation | Show results with:dehydrohalogenation
  23. [23]
    Shock tube study and RRKM calculations on thermal decomposition ...
    Thermal decomposition of alkyl halides usually occurs via a four centered cyclic transition state (TS) mechanism [9], [10], [11] and gives the corresponding ...
  24. [24]
    Unimolecular gas-phase pyrolysis of ethyl chloride - RSC Publishing
    The pyrolysis of ethyl chloride yields ethylene and hydrogen chloride, studied at 402–521°C and pressures down to 0.2 mm Hg.
  25. [25]
    [PDF] Pyrolysis of Organic Molecules Relevant to Combustion as ...
    Although alkyl fluorides are resistant to dehydrohalogenation, alkyl chlorides and bromides form alkene and hydrogen halide cleanly over intermediate.
  26. [26]
    Catalytic Dehydrochlorination of 1,2-Dichloroethane into Vinyl ...
    Jan 28, 2019 · (9,10) This problem interrupts the long-term industrial operation and requires decoking treatment every 2 months by burning coke deposit.
  27. [27]
    Influence of EDC Cracking Severity on the Marginal Costs of Vinyl ...
    The selectivity of the EDC (1,2-dichloroethane) cracking process in vinyl chloride monomer (VCM) manufacturing strongly depends on the cracking severity and ...
  28. [28]
    Coke Produced in the Commercial Pyrolysis of Ethylene Dichloride ...
    Influence of EDC Cracking Severity on the Marginal Costs of Vinyl Chloride Production. Industrial & Engineering Chemistry Research 2009, 48 (6) , 2801-2809 ...
  29. [29]
    [PDF] The Mechanisms of Pyrolysis, Oxidation, and Burning of Organic ...
    The pyrolysis of alkyl halides have been reviewed recently by Maccoll [36]. The pyrolysis of alkyl fluorides goes at a high enough temperature so that the ...
  30. [30]
    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.
  31. [31]
    Ch16: HO-C-C-X => epoxide - University of Calgary
    MECHANISM OF HALOHYDRIN TO EPOXIDE ; Step 1: An acid/base reaction. The base deprotonates the alcohol forming an alkoxide intermediate that has enhanced ...<|control11|><|separator|>
  32. [32]
    Synthesis of Epoxides - Organic Chemistry Tutor
    Synthesis of Epoxides · the epoxidation of alkenes with peroxy acids, and · the process of forming epoxides from halohydrins through intramolecular SN2 reactions.
  33. [33]
    Preparation of epoxides: Stereochemistry (video) - Khan Academy
    Sep 5, 2014 · The second way was to first form a halohydrin ... In this video, we'll look at the stereochemistry of epoxide formation for either of these two reactions.
  34. [34]
    Ethylene Oxide - NCBI - NIH
    Production of ethylene oxide began in 1914 by the chlorohydrin process, the main method used until 1937, in which ethylene chlorohydrin is converted to ...Exposure Data · Studies of Cancer in Humans · Summary of Data Reported
  35. [35]
    “Marriage” of Inorganic to Organic Chemistry as Motivation for a ...
    Sep 20, 2024 · Alternatively, an initial slow attack of the dichlorocarbene C atom by the weak nucleophile, but present in huge excess being the reaction ...<|control11|><|separator|>
  36. [36]
    ethyl isocyanide - Organic Syntheses Procedure
    Ethyl isocyanide has been obtained by treating ethylamine and chloroform with potassium hydroxide, by pyrolyzing the complex between ethyl isothiocyanate and ...
  37. [37]
    Pt-Mechanistic Study of the β-Hydrogen Elimination from ... - NIH
    Mechanistic studies of β-hydrogen elimination from transition metal alkyl complexes have shown that the process is fastest when the complex possesses an open ...
  38. [38]
    [PDF] JH PhD Doc - Queen's University Belfast
    HCl ... AlCl3 + [AlCl4]- ⇄ [Al2Cl7]-. Equation 1.2-5. 2 ... acidity by FTIR spectrum of [Et3NH]Cl – AlCl3 χAlCl3 =0.60 with added salts and pyridine as a.
  39. [39]
    Halide-free Grignard reagents for the synthesis of superior MgH 2 ...
    The thermal decomposition mechanism of Grignard reagents is believed to occur along the β-elimination of hydrogen [25,26]; and the energy required to break ...<|separator|>