E1cB-elimination reaction

Last updated
An example of the E1cB reaction mechanism in the degradation of a hemiketal under basic conditions E1cB hemiacetal.svg
An example of the E1cB reaction mechanism in the degradation of a hemiketal under basic conditions

The E1cB elimination reaction is a type of elimination reaction which occurs under basic conditions, where the hydrogen to be removed is relatively acidic, while the leaving group (such as -OH or -OR) is a relatively poor one. Usually a moderate to strong base is present. E1cB is a two-step process, the first step of which may or may not be reversible. First, a base abstracts the relatively acidic proton to generate a stabilized anion. The lone pair of electrons on the anion then moves to the neighboring atom, thus expelling the leaving group and forming a double or triple bond. [1] The name of the mechanism - E1cB - stands for Elimination Unimolecular conjugate Base. Elimination refers to the fact that the mechanism is an elimination reaction and will lose two substituents. Unimolecular refers to the fact that the rate-determining step of this reaction only involves one molecular entity. Finally, conjugate base refers to the formation of the carbanion intermediate, which is the conjugate base of the starting material.

Contents

E1cB should be thought of as being on one end of a continuous spectrum, which includes the E1 mechanism at the opposite end and the E2 mechanism in the middle. The E1 mechanism usually has the opposite characteristics: the leaving group is a good one (like -OTs or -Br), while the hydrogen is not particularly acidic and a strong base is absent. Thus, in the E1 mechanism, the leaving group leaves first to generate a carbocation. Due to the presence of an empty p orbital after departure of the leaving group, the hydrogen on the neighboring carbon becomes much more acidic, allowing it to then be removed by the weak base in the second step. In an E2 reaction, the presence of a strong base and a good leaving group allows proton abstraction by the base and the departure of the leaving group to occur simultaneously, leading to a concerted transition state in a one-step process.

Mechanism

a and b assignments in a molecule with leaving group, LG Assignment of alpha and beta carbons on a molecule.svg
α and β assignments in a molecule with leaving group, LG

There are two main requirements to have a reaction proceed down an E1cB mechanistic pathway. The compound must have an acidic hydrogen on its β-carbon and a relatively poor leaving group on the α- carbon. The first step of an E1cB mechanism is the deprotonation of the β-carbon, resulting in the formation of an anionic transition state, such as a carbanion. The greater the stability of this transition state, the more the mechanism will favor an E1cB mechanism. This transition state can be stabilized through induction or delocalization of the electron lone pair through resonance. In general it can be claimed that an electron withdrawing group on the substrate, a strong base, a poor leaving group and a polar solvent triggers the E1cB mechanism. An example of an E1cB mechanism that has a stable transition state can be seen in the degradation of ethiofencarb - a carbamate insecticide that has a relatively short half-life in Earth's atmosphere. Upon deprotonation of the amine, the resulting amide is relatively stable because it is conjugated with the neighboring carbonyl. In addition to containing an acidic hydrogen on the β-carbon, a relatively poor leaving group is also necessary. A bad leaving group is necessary because a good leaving group will leave before the ionization of the molecule. As a result, the compound will likely proceed through an E2 pathway. Some examples of compounds that contain poor leaving groups and can undergo the E1cB mechanism are alcohols and fluoroalkanes. It has also been suggested that the E1cB mechanism is more common among alkenes eliminating to alkynes than from an alkane to alkene. [2] One possible explanation for this is that the sp2 hybridization creates slightly more acidic protons. Although this mechanism is not limited to carbon-based eliminations. It has been observed with other heteroatoms, such as nitrogen in the elimination of a phenol derivative from ethiofencarb. [3]

Degradation of ethiofencarb illustrating the presence of a stable anion due to resonance between the amide functional group and the carbonyl group. E1cB Mechanism Ethiofencarb2.svg
Degradation of ethiofencarb illustrating the presence of a stable anion due to resonance between the amide functional group and the carbonyl group.

Distinguishing E1cB-elimination reactions from E1- and E2-elimination reactions

All elimination reactions involve the removal of two substituents from a pair of atoms in a compound. Alkene, alkynes, or similar heteroatom variations (such as carbonyl and cyano) will form. The E1cB mechanism is just one of three types of elimination reaction. The other two elimination reactions are E1 and E2 reactions. Although the mechanisms are similar, they vary in the timing of the deprotonation of the α-carbon and the loss of the leaving group. E1 stands for unimolecular elimination, and E2 stands for bimolecular elimination. In an E1 mechanism, the molecule contains a good leaving group that departs before deprotonation of the α-carbon. This results in the formation of a carbocation intermediate. The carbocation is then deprotonated resulting in the formation of a new pi bond. The molecule involved must also have a very good leaving group such as bromine or chlorine, and it should have a relatively less acidic α-carbon.

Example of the preferential elimination of fluorine in an E1cB-elimination reaction. E1cB preferential elimination of fluorine.svg
Example of the preferential elimination of fluorine in an E1cB-elimination reaction.

In an E2-elimination reaction, both the deprotonation of the α-carbon and the loss of the leaving group occur simultaneously in one concerted step. Molecules that undergo E2-elimination mechanisms have more acidic α-carbons than those that undergo E1 mechanisms, but their α-carbons are not as acidic as those of molecules that undergo E1cB mechanisms. The key difference between the E2 vs E1cb pathways is a distinct carbanion intermediate as opposed to one concerted mechanism. Studies have been shown that the pathways differ by using different halogen leaving groups. One example uses chlorine as a better stabilizing halogen for the anion than fluorine, [4] which makes fluorine the leaving group even though chlorine is a much better leaving group. [5] This provides evidence that the carbanion is formed because the products are not possible through the most stable concerted E2 mechanism. The following table summarizes the key differences between the three elimination reactions; however, the best way to identify which mechanism is playing a key role in a particular reaction involves the application of chemical kinetics.

E1E2E1cB
Stepwise reactionConcerted reactionStepwise reaction
Carbocation intermediateSimultaneous removal of proton, formation of double bond, and loss of leaving groupCarbanion intermediate
Strongly acidic mediaNo preferenceStrongly basic media
Good leaving groupsLeaving groupPoor leaving groups
Less acidic B-HAcidic B-HMore acidic B-H

Chemical kinetics of E1cB-elimination mechanisms

When trying to determine whether or not a reaction follows the E1cB mechanism, chemical kinetics are essential. The best way to identify the E1cB mechanism involves the use of rate laws and the kinetic isotope effect. These techniques can also help further differentiate between E1cB, E1, and E2-elimination reactions.

Rate law

When trying to experimentally determine whether or not a reaction follows the E1cB mechanism, chemical kinetics are essential. The best ways to identify the E1cB mechanism involves the use of rate laws and the kinetic isotope effect.

The rate law that governs E1cB mechanisms is relatively simple to determine. Consider the following reaction scheme.

An example of an E1cB-elimination mechanism with a generic leaving group, LG, and ethoxide as the base. E1cB reaction mechanism with leaving group LG.svg
An example of an E1cB-elimination mechanism with a generic leaving group, LG, and ethoxide as the base.

Assuming that there is a steady-state carbanion concentration in the mechanism, the rate law for an E1cB mechanism.

From this equation, it is clear the second order kinetics will be exhibited. [6] E1cB mechanisms kinetics can vary slightly based on the rate of each step. As a result, the E1cB mechanism can be broken down into three categories: [7]

  1. E1cBanion is when the carbanion is stable and/or a strong base is used in excess of the substrate, making deprotonation irreversible, followed by rate-determining loss of the leaving group (k1[base] ≫ k2).
  2. E1cBrev is when the first step is reversible but the formation of product is slower than reforming the starting material, this again results from a slow second step (k−1[conjugate acid] ≫ k2).
  3. E1cBirr is when the first step is slow, but once the anion is formed the product quickly follows (k2 ≫ k−1[conjugate acid]). This leads to an irreversible first step but unlike E1cBanion, deprotonation is rate determining.

Kinetic isotope effect

Deuterium

Deuterium exchange and a deuterium kinetic isotope effect can help distinguish among E1cBrev, E1cBanion, and E1cBirr. If the solvent is protic and contains deuterium in place of hydrogen (e.g., CH3OD), then the exchange of protons into the starting material can be monitored. If the recovered starting material contains deuterium, then the reaction is most likely undergoing an E1cBrev type mechanism. Recall, in this mechanism protonation of the carbanion (either by the conjugate acid or by solvent) is faster than loss of the leaving group. This means after the carbanion is formed, it will quickly remove a proton from the solvent to form the starting material.
If the reactant contains deuterium at the β position, a primary kinetic isotope effect indicates that deprotonation is rate determining. Of the three E1cB mechanisms, this result is only consistent with the E1cBirr mechanism, since the isotope is already removed in E1cBanion and leaving group departure is rate determining in E1cBrev.

Fluorine-19 and carbon-11

Another way that the kinetic isotope effect can help distinguish E1cB mechanisms involves the use of 19F. Fluorine is a relatively poor leaving group, and it is often employed in E1cB mechanisms. Fluorine kinetic isotope effects are also applied in the labeling of Radiopharmaceuticals and other compounds in medical research. This experiment is very useful in determining whether or not the loss of the leaving group is the rate-determining step in the mechanism and can help distinguish between E1cBirr and E2 mechanisms. 11C can also be used to probe the nature of the transition state structure. The use of 11C can be used to study the formation of the carbanion as well as study its lifetime which can not only show that the reaction is a two-step E1cB mechanism (as opposed to the concerted E2 mechanism), but it can also address the lifetime and stability of the transition state structure which can further distinguish between the three different types of E1cB mechanisms. [8]

Aldol reactions

The most well known reaction that undergoes E1cB elimination is the aldol condensation reaction under basic conditions. This involves the deprotonation of a compound containing a carbonyl group that results in the formation of an enolate. The enolate is the very stable conjugate base of the starting material, and is one of the intermediates in the reaction. This enolate then acts as a nucleophile and can attack an electrophilic aldehyde. The Aldol product is then deprotonated forming another enolate followed by the elimination of water in an E1cB dehydration reaction. Aldol reactions are a key reaction in organic chemistry because they provide a means of forming carbon-carbon bonds, allowing for the synthesis of more complex molecules. [9]

An aldol condensation reaction is one of the most common examples of an E1cB mechanism. E1cB aldol condensation.svg
An aldol condensation reaction is one of the most common examples of an E1cB mechanism.

Photo-induced E1cB

A photochemical version of E1cB has been reported by Lukeman et al. [10] In this report, a photochemically induced decarboxylation reaction generates a carbanion intermediate, which subsequently eliminates the leaving group. The reaction is unique from other forms of E1cB since it does not require a base to generate the carbanion. The carbanion formation step is irreversible, and should thus be classified as E1cBirr.

E1cB reaction mechanism through photo-induced decarboxylation. E1cB photo-induced reaction mechanism.svg
E1cB reaction mechanism through photo-induced decarboxylation.

In biology

The E1cB-elimination reaction is an important reaction in biology. For example, the penultimate step of glycolysis involves an E1cB mechanism. This step involves the conversion of 2-phosphoglycerate to phosphoenolpyruvate, facilitated by the enzyme enolase.

See also

Related Research Articles

<span class="mw-page-title-main">Chemical reaction</span> Process that results in the interconversion of chemical species

A chemical reaction is a process that leads to the chemical transformation of one set of chemical substances to another. When chemical reactions occur, the atoms are rearranged and the reaction is accompanied by an energy change as new products are generated. Classically, chemical reactions encompass changes that only involve the positions of electrons in the forming and breaking of chemical bonds between atoms, with no change to the nuclei, and can often be described by a chemical equation. Nuclear chemistry is a sub-discipline of chemistry that involves the chemical reactions of unstable and radioactive elements where both electronic and nuclear changes can occur.

<span class="mw-page-title-main">Elimination reaction</span> Reaction where 2 substituents are removed from a molecule in a 1 or 2 step mechanism

An elimination reaction is a type of organic reaction in which two substituents are removed from a molecule in either a one- or two-step mechanism. The one-step mechanism is known as the E2 reaction, and the two-step mechanism is known as the E1 reaction. The numbers refer not to the number of steps in the mechanism, but rather to the kinetics of the reaction: E2 is bimolecular (second-order) while E1 is unimolecular (first-order). In cases where the molecule is able to stabilize an anion but possesses a poor leaving group, a third type of reaction, E1CB, exists. Finally, the pyrolysis of xanthate and acetate esters proceed through an "internal" elimination mechanism, the Ei mechanism.

The unimolecular nucleophilic substitution (SN1) reaction is a substitution reaction in organic chemistry. The Hughes-Ingold symbol of the mechanism expresses two properties—"SN" stands for "nucleophilic substitution", and the "1" says that the rate-determining step is unimolecular. Thus, the rate equation is often shown as having first-order dependence on the substrate and zero-order dependence on the nucleophile. This relationship holds for situations where the amount of nucleophile is much greater than that of the intermediate. Instead, the rate equation may be more accurately described using steady-state kinetics. The reaction involves a carbocation intermediate and is commonly seen in reactions of secondary or tertiary alkyl halides under strongly basic conditions or, under strongly acidic conditions, with secondary or tertiary alcohols. With primary and secondary alkyl halides, the alternative SN2 reaction occurs. In inorganic chemistry, the SN1 reaction is often known as the dissociative substitution. This dissociation pathway is well-described by the cis effect. A reaction mechanism was first proposed by Christopher Ingold et al. in 1940. This reaction does not depend much on the strength of the nucleophile, unlike the SN2 mechanism. This type of mechanism involves two steps. The first step is the ionization of alkyl halide in the presence of aqueous acetone or ethyl alcohol. This step provides a carbocation as an intermediate.

<span class="mw-page-title-main">Leaving group</span> Atom(s) which detach from the substrate during a chemical reaction

In chemistry, a leaving group is defined by the IUPAC as an atom or group of atoms that detaches from the main or residual part of a substrate during a reaction or elementary step of a reaction. However, in common usage, the term is often limited to a fragment that departs with a pair of electrons in heterolytic bond cleavage. In this usage, a leaving group is a less formal but more commonly used synonym of the term nucleofuge. In this context, leaving groups are generally anions or neutral species, departing from neutral or cationic substrates, respectively, though in rare cases, cations leaving from a dicationic substrate are also known.

S<sub>N</sub>2 reaction Substitution reaction where bonds are broken and formed simultaneously

Bimolecular nucleophilic substitution (SN2) is a type of reaction mechanism that is common in organic chemistry. In the SN2 reaction, a strong nucleophile forms a new bond to an sp3-hybridised carbon atom via a backside attack, all while the leaving group detaches from the reaction center in a concerted fashion.

<span class="mw-page-title-main">Aldol condensation</span> Type of chemical reaction

An aldol condensation is a condensation reaction in organic chemistry in which two carbonyl moieties react to form a β-hydroxyaldehyde or β-hydroxyketone, and this is then followed by dehydration to give a conjugated enone.

In organic chemistry, a carbanion is an anion in which carbon is negatively charged.

In physical organic chemistry, a kinetic isotope effect (KIE) is the change in the reaction rate of a chemical reaction when one of the atoms in the reactants is replaced by one of its isotopes. Formally, it is the ratio of rate constants for the reactions involving the light (kL) and the heavy (kH) isotopically substituted reactants (isotopologues):

The Wolff–Kishner reduction is a reaction used in organic chemistry to convert carbonyl functionalities into methylene groups. In the context of complex molecule synthesis, it is most frequently employed to remove a carbonyl group after it has served its synthetic purpose of activating an intermediate in a preceding step. As such, there is no obvious retron for this reaction. The reaction was reported by Nikolai Kischner in 1911 and Ludwig Wolff in 1912.

<span class="mw-page-title-main">Hammond's postulate</span> Hypothesis in physical organic chemistry

Hammond's postulate, is a hypothesis in physical organic chemistry which describes the geometric structure of the transition state in an organic chemical reaction. First proposed by George Hammond in 1955, the postulate states that:

If two states, as, for example, a transition state and an unstable intermediate, occur consecutively during a reaction process and have nearly the same energy content, their interconversion will involve only a small reorganization of the molecular structures.

The benzilic acid rearrangement is formally the 1,2-rearrangement of 1,2-diketones to form α-hydroxy–carboxylic acids using a base. This reaction receives its name from the reaction of benzil with potassium hydroxide to form benzilic acid. First performed by Justus von Liebig in 1838, it is the first reported example of a rearrangement reaction. It has become a classic reaction in organic synthesis and has been reviewed many times before. It can be viewed as an intramolecular redox reaction, as one carbon center is oxidized while the other is reduced.

<span class="mw-page-title-main">Darzens reaction</span>

The Darzens reaction is the chemical reaction of a ketone or aldehyde with an α-haloester in the presence of a base to form an α,β-epoxy ester, also called a "glycidic ester". This reaction was discovered by the organic chemist Auguste Georges Darzens in 1904.

In chemistry, a reaction intermediate, or intermediate, is a molecular entity arising within the sequence of a stepwise chemical reaction. It is formed as the reaction product of an elementary step, from the reactants and/or preceding intermediates, but is consumed in a later step. It does not appear in the chemical equation for the overall reaction.

The Kornblum–DeLaMare rearrangement is a rearrangement reaction in organic chemistry in which a primary or secondary organic peroxide is converted to the corresponding ketone and alcohol under acid or base catalysis. The reaction is relevant as a tool in organic synthesis and is a key step in the biosynthesis of prostaglandins.

Oxidative decarboxylation is a decarboxylation reaction caused by oxidation. Most are accompanied by α- Ketoglutarate α- Decarboxylation caused by dehydrogenation of hydroxyl carboxylic acids such as carbonyl carboxylic acid, malic acid, isocitric acid, etc.

The Wallach rearrangement, also named Wallach transformation, is a name reaction in the organic chemistry. It is named after Otto Wallach, who discovered this reaction in 1880. In general it is a strong acid-promoted conversion of azoxybenzenes into hydroxyazobenzenes.

In organic chemistry, the Ei mechanism, also known as a thermal syn elimination or a pericyclic syn elimination, is a special type of elimination reaction in which two vicinal (adjacent) substituents on an alkane framework leave simultaneously via a cyclic transition state to form an alkene in a syn elimination. This type of elimination is unique because it is thermally activated and does not require additional reagents, unlike regular eliminations, which require an acid or base, or would in many cases involve charged intermediates. This reaction mechanism is often found in pyrolysis.

More O’Ferrall–Jencks plots are two-dimensional representations of multiple reaction coordinate potential energy surfaces for chemical reactions that involve simultaneous changes in two bonds. As such, they are a useful tool to explain or predict how changes in the reactants or reaction conditions can affect the position and geometry of the transition state of a reaction for which there are possible competing pathways.

The Evelyn effect is defined as the phenomena in which the product ratios in a chemical reaction change as the reaction proceeds. This phenomenon contradicts the fundamental principle in organic chemistry by reactions always go by the lowest energy pathway. The favored product should remain so throughout a reaction run at constant conditions. However, the ratio of alkenes before the synthesis is complete shows that the favored product to is not the favored product. The basic idea here is that the proportions of the various alkene products changes as a function of time with a change in mechanism.

Ether cleavage refers to chemical substitution reactions that lead to the cleavage of ethers. Due to the high chemical stability of ethers, the cleavage of the C-O bond is uncommon in the absence of specialized reagents or under extreme conditions.

References

  1. Grossman, R.B. (2008). The Art of Writing Reasonable Organic Mechanisms . New York: Springer. pp.  53–56. ISBN   978-0-387-95468-4.
  2. Smith, Michael (2007). March's advanced organic chemistry reactions, mechanisms, and structure (6th ed.). Hoboken, N.J.: Wiley-Interscience. pp. 1488–1493. ISBN   978-1-61583-842-4.
  3. Ouertani, Randa; El Atrache, Latifa Latrous; Hamida, Nejib Ben (2013). "Alkaline hydrolysis of ethiofencarb: Kinetic study and mechanism degradation". International Journal of Chemical Kinetics. 45 (2): 118–124. doi:10.1002/kin.20748. ISSN   0538-8066.
  4. Hine, Jack; Burske, Norbert W.; Hine, Mildred; Langford, Paul B. (1957). "The Relative Rates of Formation of Carbanions by Haloforms1". Journal of the American Chemical Society. 79 (6): 1406–1412. doi:10.1021/ja01563a037. ISSN   0002-7863.
  5. Baciocchi, Enrico; Ruzziconi, Renzo; Sebastiani, Giovanni Vittorio (1 August 1982). "Concerted and stepwise mechanisms in the eliminations from 1,2-dihaloacenaphthenes promoted by potassium tert-butoxide and potassium ethoxide in the corresponding alcohols". The Journal of Organic Chemistry. 47 (17): 3237–3241. doi:10.1021/jo00138a007.
  6. McLennan, D. J. (1967). "The carbanion mechanism of olefin-forming elimination". Quarterly Reviews, Chemical Society. 21 (4): 490. doi:10.1039/qr9672100490. ISSN   0009-2681.
  7. Smith, Michael (2007). March's advanced organic chemistry reactions, mechanisms, and structure (6th ed.). Hoboken, N.J.: Wiley-Interscience. pp. 1488–1493. ISBN   978-1-61583-842-4.
  8. Matsson, Olle; MacMillar, Susanna (September 2007). "Isotope effects for fluorine-18 and carbon-11 in the study of reaction mechanisms". Journal of Labelled Compounds and Radiopharmaceuticals. 50 (11–12): 982–988. doi:10.1002/jlcr.1443.
  9. Wade, L.G. (2005). Organic Chemistry. New Jersey: Prentice Hall. pp. 1056–1066. ISBN   0-13-236731-9.
  10. Lukeman, Matthew; Scaiano, Juan C. (2005). "Carbanion-Mediated Photocages: Rapid and Efficient Photorelease with Aqueous Compatibility". Journal of the American Chemical Society. 127 (21): 7698–7699. doi:10.1021/ja0517062. ISSN   0002-7863. PMID   15913358.