Oxymercuration reaction

Last updated

Oxymercuration reaction Oxymercuration-reduction.svg
Oxymercuration reaction

In organic chemistry, the oxymercuration reaction is an electrophilic addition reaction that transforms an alkene (R2C=CR2) into a neutral alcohol. In oxymercuration, the alkene reacts with mercuric acetate (AcO−Hg−OAc) in aqueous solution to yield the addition of an acetoxymercury (−HgOAc) group and a hydroxy (−OH) group across the double bond. Carbocations are not formed in this process and thus rearrangements are not observed. The reaction follows Markovnikov's rule (the hydroxy group will always be added to the more substituted carbon) and it is an anti addition (the two groups will be trans to each other). [2] [3] [4]

Contents

Oxymercuration followed by reductive demercuration is called an oxymercuration–reduction reaction or oxymercuration–demercuration reaction. This reaction, which is almost always done in practice instead of oxymercuration, is treated at the conclusion of the article.

Mechanism

Oxymercuration can be fully described in three steps (the whole process is sometimes called deoxymercuration), which is illustrated in stepwise fashion to the right. In the first step, the nucleophilic double bond attacks the mercury ion, ejecting an acetoxy group. The electron pair on the mercury ion in turn attacks a carbon on the double bond, forming a mercurinium ion in which the mercury atom bears a positive charge. The electrons in the highest occupied molecular orbital of the double bond are donated to mercury's empty 6s orbital and the electrons in mercury's dxz (or dyz) orbital are donated in the lowest unoccupied molecular orbital of the double bond.

In the second step, the nucleophilic water molecule attacks the more substituted carbon, liberating the electrons participating in its bond with mercury. The electrons collapse to the mercury ion and neutralize it. The oxygen in the water molecule now bears a positive charge.

In the third step, a negatively charged acetate ion deprotonates the alkyloxonium ion, forming the waste product HOAc. The two electrons participating in the bond between oxygen and the attacked hydrogen collapse into the oxygen, neutralizing its charge and creating the final alcohol product.

Curved-arrow mechanism for the oxymercuration reaction. Oxymercuration mechanism.svg
Curved-arrow mechanism for the oxymercuration reaction.

Regioselectivity and stereochemistry

Oxymercuration is very regioselective and is a textbook Markovnikov reaction; ruling out extreme cases, the water nucleophile will always preferentially attack the more substituted carbon, depositing the resultant hydroxy group there. This phenomenon is explained by examining the three resonance structures of the mercuronium ion formed at the end of the step one.

By inspection of these structures, it is seen that the positive charge of the mercury atom will sometimes reside on the more substituted carbon (approximately 4% of the time). This forms a temporary tertiary carbocation, which is a very reactive electrophile. The nucleophile will attack the mercuronium ion at this time. Therefore, the nucleophile attacks the more substituted carbon because it retains a more positive character than the lesser substituted carbon.

Stereochemically, oxymercuration is an anti addition. As illustrated by the second step, the nucleophile cannot attack the carbon from the same face as the mercury ion because of steric hindrance. There is simply insufficient room on that face of the molecule to accommodate both a mercury ion and the attacking nucleophile. Therefore, when free rotation is impossible, the hydroxy and acetoxymercuri groups will always be trans to each other.

Shown below is an example of regioselectivity and stereospecificity of the oxymercuration reaction with substituted cyclohexenes. A bulky group like t-butyl locks the ring in a chair conformation and prevents ring flips. With 4-t-butylcyclohexene, oxymercuration yields two products – where addition across the double bond is always anti – with slight preference towards acetoxymercury group trans to the t-butyl group, resulting in slightly more cis product. With 1-methyl-4-t-butylcyclohexene, oxymercuration yields only one product – still anti addition across the double bond – where water only attacks the more substituted carbon. The reason for anti addition across the double bond is to maximize orbital overlap of the lone pair of water and the empty orbital of the mercuronium ion on the opposite side of the acetoxymercury group. Regioselectivity is observed to favor water attacking the more substituted carbon, but water does not add syn across the double bond which implies that the transition state favors water attacking from the opposite side of the acetomercury group. [5]

Oxymercuration stereospecificity Oxymerc stereocontrol.png
Oxymercuration stereospecificity
Oxymercuration regioselectivity Oxymerc regiocontrol.png
Oxymercuration regioselectivity

Oxymercuration–reduction

In practice, the mercury adduct product created by the oxymercuration reaction is almost always treated with sodium borohydride (NaBH4) in aqueous base in a reaction called demercuration. In demercuration, the acetoxymercury group is replaced with a hydrogen in a stereochemically insensitive reaction [6] known as reductive elimination. The combination of oxymercuration followed immediately by demercuration is called an oxymercuration–reduction reaction. [7]

Therefore, the oxymercuration-reduction reaction is the net addition of water across the double bond. Any stereochemistry set up by the oxymercuration step is scrambled by the demercuration step, so that the hydrogen and hydroxy group may be cis or trans from each other. Oxymercuration reduction is a popular laboratory technique to achieve alkene hydration with Markovnikov selectivity while avoiding carbocation intermediates and thus the rearrangement which can lead to complex product mixtures.

Other applications

Oxymercuration is not limited to an alkene reacting with water to add hydroxyl and mercury groups. The carbon–mercury structure can undergo spontaneous replacement of the mercury by hydrogen, rather than persisting until a separate reduction step. In this manner, the effect is for mercury to be a Lewis acid catalyst. For example, using an alkyne instead of an alkene yields an enol, which tautomerizes into a ketone. Using an alcohol instead of water yields an ether (see also Hofmann-Sand reaction). In both cases, Markovnikov's rule is observed.

Other applications of oxymercuration Oxymerc other.png
Other applications of oxymercuration

Using a vinyl ether in the presence of an alcohol allows the replacement of one alkoxy group (RO–) for another by way of an acetal intermediate. An allyl alcohol and a vinyl ether under the conditions of oxymercuration–reaction can give R–CH=CH–CH2–O–CH=CH2, which is suitable for a Claisen rearrangement. [8]

Mercury-catalyzed enol ether exchange Oxymerc vinyl ether rxn.png
Mercury-catalyzed enol ether exchange
Mechanism mercury-catalyzed enol ether exchange Oxymerc vinyl ether mech.png
Mechanism mercury-catalyzed enol ether exchange

See also

Related Research Articles

<span class="mw-page-title-main">Alkene</span> Hydrocarbon compound containing one or more C=C bonds

In organic chemistry, an alkene, or olefin, is a hydrocarbon containing a carbon–carbon double bond. The double bond may be internal or in the terminal position. Terminal alkenes are also known as α-olefins.

<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.

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.

In organic chemistry, Markovnikov's rule or Markownikoff's rule describes the outcome of some addition reactions. The rule was formulated by Russian chemist Vladimir Markovnikov in 1870.

In chemistry, a hydration reaction is a chemical reaction in which a substance combines with water. In organic chemistry, water is added to an unsaturated substrate, which is usually an alkene or an alkyne. This type of reaction is employed industrially to produce ethanol, isopropanol, and butan-2-ol.

A halogen addition reaction is a simple organic reaction where a halogen molecule is added to the carbon–carbon double bond of an alkene functional group.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

In organic chemistry, a nucleophilic addition reaction is an addition reaction where a chemical compound with an electrophilic double or triple bond reacts with a nucleophile, such that the double or triple bond is broken. Nucleophilic additions differ from electrophilic additions in that the former reactions involve the group to which atoms are added accepting electron pairs, whereas the latter reactions involve the group donating electron pairs.

A substitution reaction is a chemical reaction during which one functional group in a chemical compound is replaced by another functional group. Substitution reactions are of prime importance in organic chemistry. Substitution reactions in organic chemistry are classified either as electrophilic or nucleophilic depending upon the reagent involved, whether a reactive intermediate involved in the reaction is a carbocation, a carbanion or a free radical, and whether the substrate is aliphatic or aromatic. Detailed understanding of a reaction type helps to predict the product outcome in a reaction. It also is helpful for optimizing a reaction with regard to variables such as temperature and choice of solvent.

<span class="mw-page-title-main">Electrophilic addition</span> Chemical reaction

In organic chemistry, an electrophilic addition reaction is an addition reaction where a chemical compound containing a double or triple bond has a π bond broken, with the formation of two new σ bonds.

<span class="mw-page-title-main">Hydrohalogenation</span> Electrophilic addition of hydrogen halides to alkenes

A hydrohalogenation reaction is the electrophilic addition of hydrogen halides like hydrogen chloride or hydrogen bromide to alkenes to yield the corresponding haloalkanes.

<span class="mw-page-title-main">Wacker process</span> Chemical reaction

The Wacker process or the Hoechst-Wacker process refers to the oxidation of ethylene to acetaldehyde in the presence of palladium(II) chloride and copper(II) chloride as the catalyst. This chemical reaction was one of the first homogeneous catalysis with organopalladium chemistry applied on an industrial scale.

<span class="mw-page-title-main">Hydrogen iodide</span> Chemical compound

Hydrogen iodide (HI) is a diatomic molecule and hydrogen halide. Aqueous solutions of HI are known as hydroiodic acid or hydriodic acid, a strong acid. Hydrogen iodide and hydroiodic acid are, however, different in that the former is a gas under standard conditions, whereas the other is an aqueous solution of the gas. They are interconvertible. HI is used in organic and inorganic synthesis as one of the primary sources of iodine and as a reducing agent.

In organic chemistry, neighbouring group participation has been defined by the International Union of Pure and Applied Chemistry (IUPAC) as the interaction of a reaction centre with a lone pair of electrons in an atom or the electrons present in a sigma or pi bond contained within the parent molecule but not conjugated with the reaction centre. When NGP is in operation it is normal for the reaction rate to be increased. It is also possible for the stereochemistry of the reaction to be abnormal when compared with a normal reaction. While it is possible for neighbouring groups to influence many reactions in organic chemistry this page is limited to neighbouring group effects seen with carbocations and SN2 reactions.

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.

Morris Selig Kharasch was a pioneering organic chemist best known for his work with free radical additions and polymerizations. He defined the peroxide effect, explaining how an anti-Markovnikov orientation could be achieved via free radical addition. Kharasch was born in the Russian Empire in 1895 and immigrated to the United States at the age of 13. In 1919, he completed his Ph.D. in chemistry at the University of Chicago and spent most of his professional career there.

<span class="mw-page-title-main">Vinyl cation</span> Organic cation

The vinyl cation is a carbocation with the positive charge on an alkene carbon. Its empirical formula is C
2
H+
3
. More generally, a vinylic cation is any disubstituted carbon, where the carbon bearing the positive charge is part of a double bond and is sp hybridized. In the chemical literature, substituted vinylic cations are often referred to as vinyl cations, and understood to refer to the broad class rather than the C
2
H+
3
variant alone. The vinyl cation is one of the main types of reactive intermediates involving a non-tetrahedrally coordinated carbon atom, and is necessary to explain a wide variety of observed reactivity trends. Vinyl cations are observed as reactive intermediates in solvolysis reactions, as well during electrophilic addition to alkynes, for example, through protonation of an alkyne by a strong acid. As expected from its sp hybridization, the vinyl cation prefers a linear geometry. Compounds related to the vinyl cation include allylic carbocations and benzylic carbocations, as well as aryl carbocations.

Desulfonylation reactions are chemical reactions leading to the removal of a sulfonyl group from organic compounds. As the sulfonyl functional group is electron-withdrawing, methods for cleaving the sulfur–carbon bonds of sulfones are typically reductive in nature. Olefination or replacement with hydrogen may be accomplished using reductive desulfonylation methods.

<span class="mw-page-title-main">Carbonyl α-substitution reactions</span>

Alpha-substitution reactions occur at the position next to the carbonyl group, the α-position, and involve the substitution of an α hydrogen atom by an electrophile, E, through either an enol or enolate ion intermediate.

<span class="mw-page-title-main">2-Chlorobutane</span> Chemical compound

2-Chlorobutane is a compound with formula C4H9Cl. It is also called sec-butyl chloride. It is a colorless, volatile liquid at room temperature that is not miscible in water.

References

  1. Organic Syntheses OS 6:766 Link
  2. Loudon, Marc G. (2002). "Addition Reactions of Alkenes". Organic Chemistry (fourth ed.). Oxford University Press. pp. 165–168.
  3. McGraw-Hill Higher Education (2000). Oxymercuration–Demercuration of Alkenes.
  4. Andreas Schleifenbaum (2001). "Oxymercuration". Reaktionen, Reagenzien und Prinzipien. Archived from the original on 29 August 2004.
  5. Pasto, D. J.; Gontarz, J. A. "Studies on the Mechanism of the Oxymercuration of Substituted Cyclohexenes". Journal of the American Chemical Society (1971), 93, pp 6902–6908.
  6. Whitesides, George M.; San Filippo, Joseph Jr. (1970). "Mechanism of reduction of alkylmercuric halides by metal hydrides". J. Am. Chem. Soc. 92 (22): 6611–6624. doi:10.1021/ja00725a039.
  7. Bordwell, Frederick G.; Douglass, Miriam L. "Reduction of Alkylmercuric Hydroxides by Sodium Borohydride". Journal of the American Chemical Society (1966), 88, pp 993–99.
  8. McMurry, J. E.; Andrus A.; Ksander G. M.; Muesser, J. H.; Johnson, M. A. "Stereospecific Total Synthesis of Aphidicolin.". Journal of the American Chemical Society (1979), 101, pp 1330–32.