Fact-checked by Grok 2 weeks ago

Elimination reaction

An elimination reaction is a fundamental type of in which two substituents—typically a and a —are removed from adjacent atoms (usually carbons) in a , resulting in the formation of a multiple bond, most commonly a carbon-carbon double or triple bond. This process, often termed β-elimination or 1,2-elimination, is the reverse of an and serves as a key method for synthesizing unsaturated compounds like alkenes and alkynes from saturated precursors such as alkyl halides or alcohols. The two primary mechanisms for elimination reactions are the E2 (bimolecular) and E1 (unimolecular) pathways. In the E2 mechanism, a strong base abstracts a β- simultaneously as the departs from the α-carbon in a concerted, single-step process; this requires anti-periplanar geometry between the hydrogen and for efficient reaction and is favored with primary or secondary substrates in polar aprotic solvents. The E1 mechanism, in contrast, proceeds in two steps: first, the ionizes to form a intermediate at the α-carbon, followed by of a β-; it is promoted by weak bases, polar protic solvents, and tertiary substrates where stability is high. A third mechanism, E1cB (elimination unimolecular conjugate base), involves initial to form a intermediate before departure and occurs with strong bases and substrates bearing electron-withdrawing groups that stabilize the anion. Common substrates for elimination include haloalkanes (via ) and alcohols (via acid-catalyzed ), with the reaction often requiring heat or a to proceed. These reactions frequently compete with (SN1 or SN2), where the balance shifts based on factors like /nucleophile strength, substrate sterics, and solvent polarity—strong, bulky bases favor elimination over . Product typically follows Zaitsev's rule, yielding the more stable (more substituted) as the major product, though bulky bases can lead to the less substituted Hofmann product. Stereochemistry in E2 reactions is generally anti elimination, producing specific isomers depending on substrate conformation.

Fundamentals of Elimination Reactions

Definition and Scope

An is a type of in which a loses two substituents, typically from adjacent atoms, resulting in the formation of a new π bond. This process often involves the removal of a and a , leading to the creation of multiple bonds such as those found in alkenes or alkynes. Elimination reactions are the reverse of addition reactions and play a crucial role in synthesizing unsaturated compounds from saturated precursors like alkyl halides, alcohols, or amines. The general scope of elimination reactions encompasses both unimolecular (E1) and bimolecular (E2) pathways, with β-elimination—where the substituents are removed from adjacent (β-position) carbons—being the most common variant. These reactions typically produce alkenes from alkyl halides via or alkynes under more forcing conditions, and they can also yield cyclic compounds when intramolecular elimination occurs. In β-elimination, a σ bond between the carbon and (e.g., C–X, where X is ) and a C–H σ bond are cleaved, while a new C=C π bond forms between the adjacent carbons, facilitating the construction of carbon-carbon multiple bonds essential for . A representative example is the of ethyl bromide (CH₃CH₂Br), which yields (CH₂=CH₂) and upon treatment with a strong base such as tert-butoxide ion: \mathrm{CH_3CH_2Br + t-BuO^- \rightarrow CH_2=CH_2 + t-BuOH + Br^-} This illustrates the transformation of a saturated alkyl into an through β-elimination. E1 and E2 represent the primary mechanisms, though their detailed kinetics are distinct.

Comparison to Other Reaction Types

Elimination reactions differ fundamentally from and reactions in , primarily in the nature of the products formed and the changes to molecular unsaturation. In elimination, two substituents—typically a and a —are removed from adjacent atoms (β-elimination), resulting in the formation of an unsaturated product such as an or , thereby increasing the . In contrast, reactions involve the replacement of one substituent (e.g., a ) with another (e.g., a ), preserving the overall saturation level of the molecule, as seen in the conversion of an alkyl halide () to a substituted product (RNu). reactions, on the other hand, occur at existing multiple bonds, incorporating two substituents and reducing unsaturation by forming new single bonds, such as the addition of HX to an to yield an alkyl halide. The predominance of elimination over substitution often arises in competing pathways from a common substrate like an alkyl halide treated with a or , where the reaction can fork toward either an E2 (elimination) or SN2 () outcome:
RX + /Nu⁻ → + H- + X⁻ (elimination) or RNu + X⁻ (). Factors that favor elimination include the use of strong, non-nucleophilic (e.g., tert-butoxide), elevated temperatures (which increase entropy-driven elimination), and bulky leaving groups or substrates that sterically hinder nucleophilic attack. For instance, secondary or alkyl halides with strong at high temperatures preferentially undergo elimination to form rather than products. These conditions exploit the bimolecular nature of both E2 and SN2 but tilt the balance toward β-elimination by promoting over direct displacement.
The recognition of elimination reactions as a distinct class dates to the , when chemists like Aleksandr Zaitsev utilized them for synthesis, establishing empirical rules for product in such processes in 1875.
Reaction TypeProduct CharacteristicsTypical ConditionsRepresentative Example
EliminationUnsaturated (e.g., from β-elimination)Strong , high temperature, bulky groupsCH₃CH₂Br + OH⁻ (heat) → CH₂=CH₂ + H₂O + Br⁻
SubstitutionSaturated (retained carbon skeleton)Weak /, low temperature, CH₃CH₂Br + OH⁻ → CH₃CH₂OH + Br⁻
AdditionDecreased unsaturation (saturated product)/ addition to π-bondCH₂=CH₂ + HBr → CH₃CH₂Br

Primary Mechanisms of β-Elimination

E2 Mechanism

The E2 mechanism represents a concerted, bimolecular process in β-elimination reactions, where a base simultaneously abstracts a β-hydrogen from the substrate as the leaving group departs from the α-carbon, resulting in the formation of a carbon-carbon double bond in a single transition state. This synchronous bond breaking and forming avoids discrete intermediates and requires anti-periplanar alignment of the β-hydrogen and leaving group for effective orbital overlap between the developing π-bond and the breaking σ-bonds. The anti-periplanar geometry minimizes steric interactions and stabilizes the transition state, often observed in staggered conformations of acyclic substrates or axial positions in cyclohexyl systems. Kinetically, E2 reactions obey a second-order rate law, expressed as rate = k [substrate][base], indicating that the reaction rate depends on the concentrations of both the alkyl halide and the base, consistent with the bimolecular collision in the rate-determining step. The energy profile features a single high-energy transition state, where partial bonds form between the base and β-hydrogen, and between the α- and β-carbons for the alkene, while the C-H and C-leaving group bonds weaken concurrently; this contrasts with stepwise mechanisms by lacking a discrete energy well for intermediates. E2 eliminations proceed efficiently with strong, unhindered bases such as (OH⁻) in protic solvents like , and are particularly suited to primary and secondary alkyl halides, where steric demands allow facile base approach to the β-hydrogen. adheres to Zaitsev's rule, favoring the more thermodynamically stable, more substituted as the major product due to and inductive stabilization in the resembling the . A representative example is the E2 reaction of 2-bromobutane with OH⁻, which produces but-2-ene as the major Zaitsev product and but-1-ene as the minor product, along with bromide and water: \ce{CH3-CHBr-CH2-CH3 + ^-OH ->[ethanol][heat] CH3-CH=CH-CH3 + HBr} (with trans-but-2-ene predominating over cis due to lower steric strain). The stereochemistry arises from the anti-periplanar requirement: in the Newman projection looking along the C2-C3 bond, the bromine and β-hydrogen must be trans-diaxial or anti-staggered for elimination, directing the geometry of the resulting alkene. Under conditions with bulky bases like potassium tert-butoxide (t-BuOK), E2 shifts to favor the Hofmann product—the less substituted —because steric bulk hinders access to more substituted β-hydrogens, promoting abstraction from less hindered positions. For instance, treating with t-BuOK yields a higher proportion of but-1-ene relative to compared to unhindered bases.

E1 Mechanism

The E1 mechanism, or unimolecular elimination, proceeds through a process involving the formation of a intermediate. In the first, rate-determining step, the undergoes heterolytic , where the departs to generate a planar and the anion of the . This step is unimolecular, as it depends solely on the concentration. The second step involves the of a β-hydrogen from the by a , resulting in the formation of a carbon-carbon and regeneration of the base. The of the E1 follow a rate law, expressed as rate = k [substrate], indicating that the is independent of the base concentration and is governed by the slow formation. Polar protic , such as or , stabilize the ionic intermediates through , thereby accelerating the rate-determining step and favoring the E1 pathway. This solvent effect arises from the ability of protic to hydrogen-bond with the anion, facilitating its departure. E1 reactions typically require tertiary or secondary alkyl halides as substrates, where stable s can form, and are promoted by weak bases in polar protic media. A common example is the elimination of in , which proceeds as follows: \ce{(CH3)3CBr ->[(slow)] (CH3)3C^+ + Br^-} The tert-butyl intermediate then loses a proton: \ce{(CH3)3C^+ + EtOH -> (CH3)2C=CH2 + EtOH2^+} Overall: \ce{(CH3)3CBr -> (CH3)2C=CH2 + HBr} This yields isobutene as the product. Due to the 's reactivity, rearrangements such as 1,2- shifts can occur if a more stable is accessible, for instance, in secondary systems where a hydride migrates to form a before . In addition to elimination, E1 conditions often compete with solvolysis, where the solvent acts as a nucleophile to form substitution products alongside alkenes, particularly with tertiary substrates in protic media. The product distribution depends on the stability of the carbocation and the nucleophilicity of the medium, but elimination predominates at higher temperatures.

E1cB Mechanism

The E1cB (elimination unimolecular conjugate base) mechanism is a stepwise β-elimination process that proceeds through a carbanion intermediate, distinguishing it within the family of β-elimination reactions. It begins with the reversible deprotonation of an acidic β-hydrogen by a strong base, generating the carbanion conjugate base of the substrate. This intermediate then undergoes a unimolecular expulsion of the leaving group from the adjacent α-carbon, forming the alkene product and regenerating the base. The carbanion's stability is crucial, as it allows accumulation and partitioning between elimination and reverse protonation pathways. The kinetics of E1cB eliminations are second-order overall but reflect the unimolecular nature of the rate-determining step, with the observed rate law given by rate = k [carbanion], where the carbanion concentration depends on the equilibrium constant for deprotonation (K = [carbanion][H⁺]/[substrate]) and thus on base strength and pH. In many cases, the departure of the leaving group is rate-limiting, leading to isotope effects dominated by the C-LG bond breaking, while pre-equilibrium deprotonation can show primary kinetic isotope effects on the β-hydrogen if reprotonation competes significantly. This partitioning highlights the mechanism's sensitivity to conditions that favor carbanion lifetime over rapid reversal. E1cB reactions require strong, non-nucleophilic bases to drive and are favored for substrates bearing poor leaving groups (such as or ) where direct displacement is unlikely, provided the is stabilized by electron-withdrawing groups (e.g., , carbonyl) at or near the β-position. Unlike concerted mechanisms, the intermediate enables both and periplanar geometries for elimination, as the pyramidal or sp²-hybridized anion allows conformational flexibility without strict alignment requirements. A representative example of the E1cB mechanism is the base-catalyzed dehydration of a β-hydroxy carbonyl compound, such as 3-hydroxybutan-2-one, to form an α,β-unsaturated ketone like but-3-en-2-one: \ce{CH3-C(O)-CH2-CH2OH + ^-OH ⇌ CH3-C(O)-CH^- -CH2OH + H2O \xrightarrow{k_2} CH3-C(O)-CH=CH2 + H2O + ^-OH} Here, the enolate carbanion is stabilized by the carbonyl group, facilitating departure of the hydroxide leaving group. Classic examples include the elimination from β-halo nitro compounds, such as 1-bromo-2-nitroethane (BrCH2-CH2-NO2) under basic conditions, where the nitro group stabilizes the adjacent (BrCH2-CH^- -NO2) for clean formation of nitroethene (CH2=CH-NO2). Similarly, β-halo carbonyl compounds, like 4-bromobutan-2-one treated with , undergo E1cB elimination to produce α,β-unsaturated ketones, leveraging carbonyl stabilization of the enolate-like .

Factors Governing Mechanism Selection

Kinetic and Thermodynamic Considerations

The selection of elimination mechanism is primarily governed by kinetic factors that influence reaction rates, including the strength and concentration of the base, the structure of the substrate, and temperature. Strong bases, such as alkoxides or amide ions with pKa values of their conjugate acids exceeding 15-20, promote the bimolecular E2 mechanism by facilitating concerted proton abstraction and leaving group departure, whereas weak bases like water or alcohols (pKa of conjugate acids ≈ -2) favor the unimolecular E1 pathway through carbocation intermediates. High base concentrations accelerate E2 by increasing the likelihood of bimolecular collisions, while low concentrations allow E1 to dominate via unimolecular dissociation. Substrate structure plays a crucial role: primary alkyl halides predominantly undergo E2 due to steric accessibility for base approach, secondary substrates can proceed via either depending on conditions, and tertiary substrates favor E1 owing to stable carbocation formation. Elevated temperatures generally favor elimination over substitution but specifically promote E1 over E2 for secondary and tertiary substrates, as the higher activation energy of E1 (often 25-35 kcal/mol for carbocation formation) is overcome, while E2 activation energies remain lower at 20-30 kcal/mol for concerted processes. Solvent effects further modulate kinetics by stabilizing or destabilizing charged species: polar protic solvents (e.g., , ) solvate anions via hydrogen bonding, reducing base nucleophilicity and favoring E1 through stabilization, whereas polar aprotic solvents (e.g., DMSO, acetone) enhance anion reactivity by lacking such , thereby accelerating E2. Thermodynamic considerations dictate product distribution under conditions where equilibration occurs, such as in E1 mechanisms or reversible eliminations. The Zaitsev product, featuring the more substituted (and thus more stable) alkene due to greater hyperconjugation and inductive stabilization, predominates under thermodynamic control, as reflected in lower free energies (typically 2-5 kcal/mol more stable than less substituted isomers). In contrast, kinetic control in E2 reactions with bulky bases (e.g., tert-butoxide) yields the Hofmann product, the less substituted alkene, due to lower steric hindrance in the transition state for proton abstraction from less hindered positions. Free energy diagrams illustrate this: the Zaitsev pathway has a deeper energy well but may involve a higher transition state barrier under kinetic conditions, while Hofmann paths exhibit shallower wells but accessible early barriers. The interplay of these factors is summarized in the following table of typical conditions favoring each mechanism:
Substrate TypeBase Strength/ConcentrationSolventTemperaturePreferred Mechanism
PrimaryStrong/HighAproticModerateE2
SecondaryWeak/LowProticHighE1
TertiaryAny/LowProticAnyE1
SecondaryStrong/HighAproticLowE2
This matrix highlights preferences based on empirical trends, with pKa of the base's conjugate acid serving as a key selector: bases with > ~15 promote E2 dominance even for secondary substrates.

Stereochemical and Regiochemical Outcomes

In elimination reactions, stereochemical outcomes are profoundly influenced by the involved. The E2 typically proceeds via anti-periplanar , where the β-hydrogen and the are trans to each other in the , favoring the formation of trans () alkenes as the major stereoisomer when both cis and trans products are possible. This arises from optimal orbital overlap in the anti conformation, as demonstrated in studies of acyclic and cyclic substrates. In contrast, the E1 involves a planar , which allows free and leads to a of stereoisomers, often with the more stable trans predominating but without the high selectivity of E2. The E1cB can exhibit syn elimination in certain cases, particularly when the is stabilized, resulting in cis alkenes from substrates where anti is inaccessible. Regiochemical outcomes in β-elimination reactions are governed by Zaitsev's rule, which predicts that the major product is the with the most substituted double bond (thermodynamic control), as observed in the of alcohols or of alkyl halides under equilibrating conditions. This preference stems from the greater stability of more substituted s due to and inductive effects. However, the Hofmann product—the less substituted —becomes favored under kinetic control with bulky bases like tert-butoxide, which abstract the more accessible β-hydrogen due to steric hindrance. labeling experiments confirm this, as in the elimination from where isotopic substitution at the less hindered position increases the Hofmann product yield by slowing abstraction from the Zaitsev site. In cyclic systems, such as cyclohexyl halides, E2 elimination requires both the and β-hydrogen to occupy axial positions for -periplanar alignment, necessitating a chair flip in equatorial-substituted substrates to access the reactive conformer. This geometric constraint enforces , producing alkenes only from the diaxial arrangement and often favoring endocyclic double bonds in appropriately substituted cyclohexanes. Illustrative examples from chiral substrates highlight these principles. Treatment of meso-2,3-dibromobutane with base via E2 yields trans-2-butene due to elimination from the staggered conformation, while the racemic () pair produces cis-2-butene, demonstrating how diastereomeric starting materials dictate stereoisomer formation. Exceptions occur in specialized cases like dihalides, where double E2 elimination with strong bases such as directly forms alkynes, bypassing typical regiochemistry since the product lacks stereoisomers. Similarly, vinylic halides undergo elimination to alkynes under harsh conditions, with no stereochemical outcome at the due to its linearity, though the initial geometry influences reactivity rates.

Variations Beyond β-Elimination

α-Elimination Reactions

α-Elimination reactions involve the removal of two substituents, typically a proton and a , from the same carbon atom (the α-carbon), resulting in the formation of a intermediate. Unlike β-elimination, which requires participation from an adjacent atom and yields alkenes, α-elimination proceeds without involvement of neighboring carbons, directly generating highly reactive species such as dihalocarbenes. This process is particularly useful for introducing functionality in synthetic transformations. The mechanism of α-elimination is often base-catalyzed and follows an E1cB pathway, where initial forms a intermediate, followed by expulsion of the from the same carbon. A classic example is the generation of dichlorocarbene (:CCl₂) from (CHCl₃) under basic conditions: \ce{CHCl3 + OH^- ->{{grok:render&&&type=render_inline_citation&&&citation_id=1&&&citation_type=wikipedia}} H2O + ^-CCl3} \ce{^-CCl3 ->{{grok:render&&&type=render_inline_citation&&&citation_id=2&&&citation_type=wikipedia}} :CCl2 + Cl^-} In the first step, the strong abstracts the α-proton to yield the trichloromethyl anion, which then undergoes rapid α-elimination of to form the dichlorocarbene. This two-step sequence is facilitated by the acidity of the α-hydrogen in polyhalomethanes, with the overall reaction being: \ce{CHCl3 + KOH -> :CCl2 + KCl + H2O} Similar processes apply to other haloforms, such as bromoform for dibromocarbene. Kinetically, α-elimination reactions resemble E2 processes in their concert-like character but occur on a single carbon, often requiring strong bases like aqueous KOH or NaOH. To enhance efficiency and solubility in biphasic systems, phase-transfer catalysis is commonly employed, using quaternary ammonium salts (e.g., benzyltriethylammonium chloride) to transport hydroxide ions into the organic phase, accelerating carbene formation without high temperatures. This method allows controlled generation of carbenes at room temperature, minimizing side reactions. In applications, dihalocarbenes generated via α-elimination add stereospecifically to alkenes, forming 1,1-dihalocyclopropanes that serve as versatile intermediates in , such as for ring expansion or further functionalization. For instance, the addition of :CCl₂ to yields 7,7-dichlorobicyclo[4.1.0]heptane, preserving the alkene's in a syn addition. A variant, the , employs (CH₂I₂) with a zinc-copper couple to generate a methylene carbenoid (equivalent to :CH₂), enabling mild, stereoselective of alkenes without halogenation, particularly useful for allylic alcohols via directed insertion. These methods highlight α-elimination's role in constructing strained rings central to and pharmaceutical synthesis.

Other Specialized Eliminations

Pyrolytic eliminations represent a of thermal decompositions that proceed via concerted, syn-elimination mechanisms, distinct from acid- or -catalyzed β-eliminations. One prominent example is the Chugaev reaction, where alkyl xanthates derived from alcohols undergo to yield s, (), and a . The general transformation is depicted as: \text{ROC(S)SR'} \rightarrow \text{[alkene](/page/Alkene)} + \text{[COS](/page/COS)} + \text{R'SH} This process involves a six-membered cyclic , enforcing strict syn , which favors the formation of alkenes with aligned to the original alcohol's . The reaction is particularly useful for sensitive substrates, as it occurs under neutral conditions at temperatures around 200–300°C. Another variant is the of esters, which also follows a syn-elimination pathway through a similar cyclic involving the acetate carbonyl acting as an internal . This method converts secondary or alcohols to alkenes upon heating to 400–500°C, often in the gas phase, and exhibits with high . Metal-catalyzed eliminations extend these processes to more selective syntheses under milder conditions. Palladium-mediated elimination of propargylic carbonates provides an efficient route to conjugated enynes, proceeding via , , and β-hydride elimination steps to form the (E)-enyne geometry selectively. This reaction operates at with Pd(0) catalysts like Pd(PPh₃)₄, enabling applications in natural product synthesis. , catalyzed by or complexes, can be conceptualized as a reversible process akin to a reverse elimination, where alkenes redistribute fragments through metallacyclobutane intermediates, effectively interconverting olefins without net loss of material. Sigmatropic eliminations encompass pericyclic variants, such as the [2,3]-Wittig rearrangement, in which deprotonated allylic ethers (e.g., allyl benzyl ethers) undergo suprafacial [2,3]-sigmatropic shifts, resulting in homoallylic alcohols that can further eliminate to form dienes under appropriate conditions. This stereospecific process proceeds via a chair-like , preserving allylic and offering control over in complex molecules. Industrially, specialized eliminations are pivotal in large-scale production, exemplified by the catalytic dehydration of ethanol to ethylene using alumina or zeolite catalysts at 300–500°C. This process follows an E1-like mechanism involving carbocation intermediates on acidic sites, yielding over 99% ethylene selectivity and serving as a renewable alternative to petroleum cracking, with global capacity exceeding 1 million tons annually from bioethanol sources.

Applications and Historical Development

Synthetic and Biological Importance

Elimination reactions serve as a cornerstone in for the preparation of alkenes and alkynes, enabling the formation of carbon-carbon multiple bonds from readily available precursors such as alkyl halides and alcohols. These transformations are particularly valuable in constructing complex molecular frameworks, where the and of the elimination dictate the outcome of downstream reactions. A prominent variant is the Julia-Kocienski olefination, which involves the addition of a metallated to a carbonyl compound followed by β-elimination to yield alkenes with high (E)-selectivity under mild conditions. This method has been extensively applied in the of natural products, including terpenoids. Similarly, elimination reactions feature in synthesis, such as in Corey's route to prostaglandin F2α. In biological systems, elimination reactions underpin essential metabolic pathways, exemplified by the enzyme enolase in glycolysis, which catalyzes the dehydration of 2-phosphoglycerate to phosphoenolpyruvate via an E1cB-like mechanism, generating a high-energy phosphate for ATP production. In fatty acid metabolism, β-oxidation incorporates elimination-like dehydrogenation steps, where acyl-CoA dehydrogenase removes hydrogens from the α- and β-carbons to form trans-enoyl-CoA, facilitating the breakdown of fatty acids into acetyl-CoA units for energy generation. Additionally, nucleotide dehydratases, such as the radical-SAM enzyme producing ddhCTP, play roles in restricting viral replication by modulating host metabolism and interfering with viral nucleotide incorporation. These reactions offer advantages in pharmaceutical through precise stereocontrol, as seen in the double elimination for , yielding stereodefined polyenes critical for drugs. Catalytic variants, such as transition-metal-mediated eliminations, align with principles by minimizing waste and enabling efficient, atom-economical processes over stoichiometric methods. Recent advances include photochemical decarboxyolefination for modification and electrocatalytic methods for selective eliminations, enhancing sustainability in as of 2024. Inhibitors targeting viral dehydratases further highlight therapeutic potential, with compounds like brequinar blocking to suppress replication.
Reagent/BaseSubstrate TypeTypical ProductYield Range (%)Reference
t-BuOK in t-BuOHSecondary Internal (Zaitsev)85–95
KOH in EtOHCycloalkyl Cyclo80–90
NaNH₂ in liq. NH₃Vicinal di70–85

Historical Milestones

The study of elimination reactions began in the 19th century with observations of dehydrohalogenation processes leading to alkene formation from alkyl halides. In 1870, Vladimir Markovnikov proposed a rule for the regioselective addition of hydrogen halides to alkenes, which provided foundational insights into the regiochemistry observed in subsequent elimination reactions. In 1875, Alexander Zaitsev formulated his rule, stating that elimination reactions favor the formation of the more substituted (more stable) alkene as the major product, a principle that became central to understanding β-elimination regioselectivity. During the early 20th century, debates arose regarding product distribution in eliminations, particularly the Zaitsev-Hofmann dichotomy, where less substituted alkenes predominate under certain conditions like in quaternary ammonium salt decompositions. In 1927, G. Hanhart and C. K. Ingold rationalized Hofmann's rule through inductive effects influencing base approach. The mechanistic framework advanced significantly in the 1930s and 1940s through the work of Edward D. Hughes and Christopher K. Ingold, who proposed the SN1/E1 mechanisms involving carbocation intermediates in 1935–1940 and the concerted E2 mechanism in 1933, with detailed kinetic studies confirming bimolecular elimination in 1941. Key figures like Ingold further developed the E1cB mechanism in the 1950s, describing eliminations proceeding via a carbanion intermediate stabilized by electron-withdrawing groups, as evidenced in studies of β-halo carbonyl compounds. In the modern era, the 1990s saw the emergence of asymmetric elimination reactions, enabling enantioselective alkene formation using chiral bases or catalysts, as demonstrated in palladium-catalyzed systems for allylic substrates. The 2000s brought computational modeling advances, with density functional theory (DFT) analyses elucidating transition states in E2 and E1cB pathways, revealing stereoelectronic effects on barriers and selectivity.

Key Historical Milestones

  • 1870: Markovnikov's rule established, influencing elimination regiochemistry by analogy to addition patterns.
  • 1875: Zaitsev's rule proposed, predicting preference for more substituted alkenes in dehydrohalogenations.
  • 1920s: Ingold introduces early concepts of concerted elimination mechanisms.
  • 1927: Zaitsev-Hofmann debate resolved mechanistically via inductive effects by Hanhart and Ingold.
  • 1933: Hughes and Ingold suggest E2 and SN2 mechanisms based on kinetic order.
  • 1935: E1 mechanism identified by Hughes through solvolysis studies.
  • 1941: Comprehensive kinetics of E1 and E2 eliminations published by Hughes and Ingold.
  • 1948: Hughes details stereochemistry and product ratios in E2 reactions of secondary alkyl bromides.
  • 1950s: Ingold develops E1cB mechanism for carbanion-mediated eliminations.
  • 1990s: First reports of asymmetric E2 eliminations using chiral auxiliaries for stereocontrol.
  • 2000s: DFT computations model E2 transition states, quantifying base-leaving group interactions.

References

  1. [1]
  2. [2]
    Elimination Reactions – Organic Chemistry
    An elimination reaction removes a small molecule from two adjacent atoms, forming a multiple bond. For alkyl halides, it's called dehydrohalogenation, ...Missing: definition | Show results with:definition
  3. [3]
  4. [4]
    None
    ### Summary of Elimination Reactions from Neuman Chapter 9
  5. [5]
    Illustrated Glossary of Organic Chemistry - Elimination reaction
    An elimination reaction is the mechanistic reverse of an addition reaction. A generalized β-elimination reaction. A β-elimination reaction following the E2 ...
  6. [6]
    Chemical Reactivity - MSU chemistry
    Elimination of vicinal groups, usually called 1,2- or beta-elimination, is the most common type of elimination reaction. Examples of some typical cyclohexyl ...
  7. [7]
    [PDF] Eliminations
    An elimination is when a leaving group and another atom, typically hydrogen, leave a molecule, forming a new π bond, and no new atoms are added.Missing: definition | Show results with:definition
  8. [8]
    [PDF] Klein 6-8 alkene stereo, substn & elim
    A haloalkane, e.g. CH3CH2Br, can in principle undergo either of two polar reactions when it encounters ... t-BuO– + CH3CH2I → CH2=CH2 + t-BuOH + I–. KOH + ...
  9. [9]
    Chemical Reactivity - MSU chemistry
    The reverse is true of elimination reactions, i.e.the number of σ-bonds in the substrate decreases, and new π-bonds are often formed. Substitution reactions, as ...
  10. [10]
    24.4 Common Classes of Organic Reactions
    Elimination reactions are similar to cleavage reactions in inorganic compounds. Much of the approximately 26 million tons of ethylene produced per year in the ...Missing: definition | Show results with:definition
  11. [11]
    Organic Reactions
    The reaction that produces the alkene involves the loss of an HBr molecule to form a C=C double bond. It is therefore an example of an elimination reaction.Missing: definition | Show results with:definition
  12. [12]
    [PDF] ALEKSANDR MIKHAILOVICH ZAITSEV (1841-1910) - IDEALS
    In 1873, he presented the same view of elimination, illustrating it with the same oxidation reactions, at a chemical con- ference in Kazan'. In this ...
  13. [13]
    Elimination by the E2 mechanism - Chemistry LibreTexts
    Jan 22, 2023 · Kinetic studies of these reactions show that they are both second order (first order in R–Br and first order in Nu:(–)), suggesting a ...
  14. [14]
    2.9: The Mechanism of the E2 Reaction - Chemistry LibreTexts
    Jul 21, 2023 · Understanding that E2 reactions require an anti-periplanar geometry can explain cis and trans alkene isomers can form as products. This occurs ...
  15. [15]
    Antiperiplanar Relationships: The E2 Reaction and Cyclohexane ...
    Jul 4, 2025 · In the E2 reaction, the leaving group is always “anti-periplanar” to the hydrogen that is removed on the adjacent carbon (ie the “Beta-Carbon”).
  16. [16]
    Mechanism of the E2 Reaction - Master Organic Chemistry
    Sep 27, 2012 · This paper uses FMO theory to explain the stereoselectivity of the E2 reactions in terms of orbital overlap between the anti-periplanar C-H bond ...
  17. [17]
    11.8 The E2 Reaction and the Deuterium Isotope Effect - OpenStax
    Sep 20, 2023 · Anti periplanar geometry for E2 eliminations has specific stereochemical consequences that provide strong evidence for the proposed mechanism.
  18. [18]
    8.5. Elimination reactions | Organic Chemistry 1: An open textbook
    In E2, elimination shows a second order rate law, and occurs in a single ... The rate at which this mechanism occurs is second order kinetics, and depends on both ...
  19. [19]
    11.7: Elimination Reactions - Zaitsev's Rule - Chemistry LibreTexts
    Nov 6, 2023 · Zaitsev's or Saytzev's (anglicized spelling) rule is an empirical rule used to predict regioselectivity of 1,2-elimination reactions occurring via the E1 or E2 ...
  20. [20]
    Stereoselectivity of E2 Elimination Reactions - Chemistry Steps
    Most often, the E2 elimination occurs in the anti-periplanar geometry since this is the low-energy staggered conformation of the alkyl halide.
  21. [21]
    Bulky Bases in Elimination Reactions - Master Organic Chemistry
    Oct 24, 2012 · When bulky bases like t-butoxide are used in elimination reactions (E2), "non-Zaitsev" (aka "Hofmann") products can result due to steric ...
  22. [22]
    8.1: E2 Reaction - Chemistry LibreTexts
    Dec 15, 2021 · E2 mechanism is the bimolecular elimination mechanism, that the reaction rate depends on the concentration of both substrate and base.
  23. [23]
    Structure and mechanism in organic chemistry - Internet Archive
    Jan 14, 2019 · Structure and mechanism in organic chemistry. by: Ingold, Christopher, Sir, 1893-. Publication date: 1969. Topics: {u'11': u' ...
  24. [24]
    [PDF] Nucleophilic Reactions In the 1930's, Hughes and Ingold ...
    Usually the rate equation for this reaction is written. -d[tButylbromide]/dt = k. 1. [tButylbromide]. Thus the reaction is first order with respect to the ...
  25. [25]
    The Role of Solvent in SN1, SN2, E1 and E2 Reactions
    Polar protic solvents are capable of making hydrogen bonding, i.e., they contain a hydrogen connected to an electronegative atom and thus can make ...
  26. [26]
    The E1 Reaction and Its Mechanism - Master Organic Chemistry
    Sep 19, 2012 · The reaction is proposed to occur in two steps: first, the leaving group leaves, forming a carbocation. Second, base removes a proton, forming the alkene.
  27. [27]
    Mechanism for Nucleophilic Substitution and Elimination Reactions ...
    Mechanism for Nucleophilic Substitution and Elimination Reactions at Tertiary Carbon in Largely Aqueous Solutions: Lifetime of a Simple Tertiary Carbocation.
  28. [28]
    E1 Reactions With Rearrangement - Alkyl and Hyride Shifts
    Nov 9, 2012 · Elimination reactions (E1) that occur with rearrangements – hydride or alkyl shifts. Where there are carbocations (see last post), rearrangement reactions are ...
  29. [29]
    11.10 The E1 and E1cB Reactions - Organic Chemistry | OpenStax
    Sep 20, 2023 · In contrast to the E1 reaction, which involves a carbocation intermediate, the E1cB reaction takes place through a carbanion intermediate. Base ...
  30. [30]
    E1CB Elimination Mechanism - Chemistry Steps
    E1cB (elimination, unimolecular, conjugate base) is a unimolecular elimination mechanism where the intermediate is a carbanion.
  31. [31]
    an elimination reaction with kinetics characteristic of the E1cB ...
    an elimination reaction with kinetics characteristic of the E1cB mechanism ... Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc. 1968 ...
  32. [32]
    E1cB - Elimination (Unimolecular) Conjugate Base
    Feb 11, 2020 · The E1cB - Elimination Unimolecular Conjugate Base mechanism proceeds when deprotonation is followed by an elimination step.
  33. [33]
    Elimination reactions of N-[2-(p-nitrophenyl)ethyl]alkylammonium ...
    Elimination reactions of N-[2-(p-nitrophenyl)ethyl]alkylammonium ions by an E1cB mechanism. Click to copy article linkArticle link copied! James R. Keeffe ...
  34. [34]
    Elimination Reactions of Alkyl Halides - MSU chemistry
    The corresponding designation for the elimination reaction is E2. An energy diagram for the single-step bimolecular E2 mechanism is shown on the right. We ...
  35. [35]
    E2-Elimination - an overview | ScienceDirect Topics
    Factors affecting E2 elimination reaction-. The following factors affect the rate of E2 elimination reaction-. 1. Structure of substrate. 2. Strength and ...
  36. [36]
    [PDF] 6-21 Comparison of E1 and E2 Elimination Mechanisms
    We will review the most important factors that determine the reaction pathway, organized in a sequence that allows you to predict as much as can be ...
  37. [37]
    [PDF] Sn1 Sn2 E1 E2
    SN2/E2. 5. Temperature. Higher temperatures generally favor elimination reactions (E1, E2) because elimination leads to higher entropy (more molecules formed) ...
  38. [38]
    SN2 versus E2 Competition of F– and PH2– Revisited
    Oct 20, 2020 · The E2 reaction pathway for the attack with the F– Lewis base proceeds with an activation barrier that is almost 5 kcal mol–1 lower in energy ...Introduction · Results and Discussion · Conclusions · Methods
  39. [39]
    [PDF] Chapter 8 Alkyl Halides and Elimination Reactions
    Mechanism: This type of mechanism, involving concerted removal of a β-proton by a base and loss of a halide ion, is called an E2 mechanism.
  40. [40]
    Theoretical studies on E2 elimination reactions. Evidence that syn ...
    Theoretical studies on E2 elimination reactions. Evidence that syn elimination is accompanied by inversion of configuration at the carbanionic center.
  41. [41]
    Kinetics and Mechanisms of Dehydration of Secondary Alcohols ...
    Jun 11, 2018 · The E1 elimination mechanism dominates over the corresponding E2 mechanism, with the E2 mechanism being competitive with E1 only for the most ...
  42. [42]
    Stereochemistry of 1,2-elimination reactions at the E2–E1cB ...
    We have synthesized stereospecifically-deuterated β-tosyloxybutanoate esters and thioesters and studied the stereoselectivity of their elimination reactions ...
  43. [43]
    Zaitsev's elimination rule | SpringerLink
    Cite this chapter. Li, J.J. (2014). Zaitsev's elimination rule. In: Name Reactions. Springer, Cham. https://doi.org/10.1007/978-3-319-03979-4_296. Download ...Missing: original paper
  44. [44]
    The Mechanism of the Hofmann Elimination Reaction. Deuterium ...
    Synthesis of Trialkylamines with Extreme Steric Hindrance and Their Decay by a Hofmann-like Elimination Reaction.Missing: original | Show results with:original
  45. [45]
    Studies in the elimination of substituted vinyl halides to acetylenes
    Studies in the elimination of substituted vinyl halides to acetylenes. Click to ... Total Synthesis of Myxovirescins, 1 Strategy and Construction of the ...Missing: stereochemistry | Show results with:stereochemistry<|control11|><|separator|>
  46. [46]
    5: Carbene Reactions - Chemistry LibreTexts
    Aug 1, 2023 · A common example of the alpha-elimination reaction is the deprotonation reaction of chloroform with hydroxide to yield dichlorocarbene.
  47. [47]
    “Marriage” of Inorganic to Organic Chemistry as Motivation for ... - NIH
    Sep 20, 2024 · ... mechanism, an E1cB-type α-elimination, with generation of the active intermediate dichlorocarbene. This then undergoes a direct insertion ...
  48. [48]
    Phase transfer catalysis in dichlorocarbene chemistry: basic ...
    Aug 11, 2012 · The discovery, basic mechanistic concepts, and specific features of generation and reactions of dichlorocarbene under phase transfer catalysis
  49. [49]
    2,2-Dichlorobicyclo[4.1.0]heptane from cyclohexene and ...
    The authors have developed a procedure for the addition of dichlorocarbene to cyclohexane to give dichloronorcarane.
  50. [50]
    8.9: Addition of Carbenes to Alkenes - Cyclopropane Synthesis
    Apr 1, 2024 · One common method of cyclopropane synthesis is the reaction of carbenes with the double bond in alkenes or cycloalkenes.
  51. [51]
    Cyclopropanation of Alkenes - Master Organic Chemistry
    Oct 18, 2023 · This overall process – breakage of C-H and loss of a leaving group on the same carbon – is known as alpha-elimination, as distinguished from ...
  52. [52]
    Simmons–Smith Cyclopropanation: A Multifaceted Synthetic ... - NIH
    Simmons–Smith cyclopropanation is a widely used reaction in organic synthesis for stereospecific conversion of alkenes into cyclopropane.Missing: elimination | Show results with:elimination
  53. [53]
    Studies in Stereochemistry. IV. The Chugaev Reaction in the ...
    Computational study of the mechanism of thermal decomposition of xanthates in the gas phase (the Chugaev reaction). Journal of Physical Organic Chemistry ...
  54. [54]
    The mechanism of the pyrolytic elimination reaction of acetates
    The mechanism of the pyrolytic elimination reaction of acetates | The Journal of Organic Chemistry. Elements of Water Bacteriology.Missing: stereochemistry | Show results with:stereochemistry
  55. [55]
    Preparation of conjugated enynes by the palladium-catalyzed ...
    Propargylic carbonates are cleanly converted to conjugated enynes by palladium-catalyzed elimination reaction under mild and neutral conditions.Missing: mediated | Show results with:mediated
  56. [56]
    Palladium-catalyzed cycloisomerizations of enynes and related ...
    Palladium-catalyzed cycloisomerizations of enynes and related reactions ... Synthesis and Functionalization of Indoles Through Palladium-catalyzed Reactions.<|control11|><|separator|>
  57. [57]
    Transition structure for the [2,3]-Wittig rearrangement and analysis of ...
    Transition structure for the [2,3]-Wittig rearrangement and analysis of stereoselectivities. Click to copy article linkArticle link copied! Yun Dong Wu ...
  58. [58]
    Ethylene Formation from Ethanol Dehydration Using ZSM-5 Catalyst
    Catalysts prepared for ethanol dehydration in a fixed-bed reactor acted as strong active acidic catalysts under reaction conditions at lower temperatures.Introduction · Results and Discussion · Experimental Section · References
  59. [59]
    Substitution, Elimination, and Integration of Methyl Groups in ...
    Aug 16, 2024 · (5) Although less common, the elimination of methyl groups also has been used in the total synthesis of terpenoids. For example, Shenvi and ...
  60. [60]
    Latest Developments of the Julia–Kocienski Olefination Reaction - NIH
    Jun 7, 2024 · The Julia–Kocienski olefination reaction has over past 30 years become one of the key CC connective methods that is used in late-stage natural product ...
  61. [61]
    Synthesis of Prostaglandin F2α by Elias J. Corey (1969) - SynArchive
    Complete schematic view of the 1969 Prostaglandin F2α's synthesis performed by Elias J. Corey.Missing: elimination | Show results with:elimination
  62. [62]
    Biochemical and Biophysical Characterization of the Enolase ... - NIH
    Dec 17, 2018 · Introduction. Enolase is an enzyme that catalyses dehydration of 2-phospho-D-glycerate into phosphoenolpyruvate during glycolysis and its ...
  63. [63]
    Biochemistry, Fatty Acid Oxidation - StatPearls - NCBI Bookshelf - NIH
    Beta-oxidation is a significant source of metabolic energy during interprandial periods and high energy demand states, such as exercise.[1] These metabolic ...Missing: elimination | Show results with:elimination
  64. [64]
    Radical-SAM dependent nucleotide dehydratase (SAND ... - Frontiers
    (A) SAND produces the nucleoside triphosphate analogue ddhCTP in humans. ddhCTP modulates metabolism affecting cell function and restricting viral replication.
  65. [65]
    Stereocontrolled synthesis of vitamin A through a double elimination ...
    Stereocontrolled synthesis of vitamin A through a double elimination reaction. A novel convergent C10 + C10 route | The Journal of Organic Chemistry. ...
  66. [66]
    Catalytic versus stoichiometric reagents as a key concept for Green ...
    Jan 19, 2016 · Catalysts allow us to decrease the activation energy of chemical transformations, enhance the reaction rate, facilitate high selectivity.
  67. [67]
    Dihydroorotate dehydrogenase inhibitors for the treatment of viral ...
    Flavivirus, rhabdovirus and paramyxovirus infections may be treated by administering an inhibitor of the enzyme dihydroorotate dehydrogenase such as ...
  68. [68]
    11.7 Elimination Reactions: Zaitsev's Rule – Organic Chemistry
    According to Zaitsev's rule, formulated in 1875 by the Russian chemist Alexander Zaitsev, base-induced elimination reactions generally (although not always) ...Missing: history | Show results with:history
  69. [69]
    The mechanism and kinetics of elimination reactions - RSC Publishing
    The mechanism and kinetics of elimination reactions. E. D. Hughes and C. K. Ingold, Trans. Faraday Soc., 1941, 37, 657 DOI: 10.1039/TF9413700657. To request ...Missing: 1940s | Show results with:1940s
  70. [70]
    The Rise of the Highes and Ingold Theory from 1930–1942
    E.D. Hughes and C.K. Ingold. A Series of 16 Consecutive Papers on Elimination Reactions and Nucleophilic Substitution. Journal of the Chemical Society: 899–1029 ...
  71. [71]
    Effects on reactivity in E1cB reactions - ACS Publications
    ... Mechanism of TT-LYK Inhibitor in Cobalt Chemical Mechanical Polishing · Substituent-Dependent Photophysical Properties Due to the Thorpe–Ingold Effect on ...Missing: original | Show results with:original
  72. [72]
    Lord Todd – Biographical - NobelPrize.org
    Sir Alexander Robertus Todd was born in Glasgow on October 2, 1907, the elder son of Alexander Todd, a business man of that city, and his wife Jean Lowrie.Missing: eliminations | Show results with:eliminations
  73. [73]
    Asymmetric induction via addition-elimination process
    Publication History. Published. online 1 May 2002 ... Development of Novel Asymmetric Reactions and Their Application to the Synthesis of Natural Products.
  74. [74]
    How Alkyl Halide Structure Affects E2 and SN2 Reaction Barriers
    ... reactions. (4). Already in the 1940s, Hughes and Ingold established that the SN2 reaction is sensitive to steric hindrance. (5-7) About a decade ago, gas ...
  75. [75]
    E2 Reactions - Chemistry LibreTexts
    Jan 22, 2023 · E2, bimolecular elimination, was proposed in the 1920s by British chemist Christopher Kelk Ingold. Unlike E1 reactions, E2 reactions remove ...
  76. [76]
    420. Mechanism of elimination reactions. Part X. Kinetics of olefin ...
    Kinetics of olefin elimination from isopropyl, sec.-butyl, 2-n-amyl, and 3-n-amyl bromides in acidic and alkaline alcoholic media. M. L. Dhar, E. D. Hughes and ...Missing: E2 | Show results with:E2