Enols and Enolates - Alpha Carbon Chemistry | Organic Chemistry 3

The chemistry of the alpha carbon of a carbonyl is studied in this chapter: the alpha position, enols and tautomerization, the preparation of enolates and their reactivity, alpha halogenation, the haloform reaction, the alkylation of enolates, the malonic ester synthesis and acetoacetic ester synthesis.

The Alpha Position

Carbon atoms:

For compounds containing a carbonyl group, Greek letters are used to describe the proximity of each carbon atom to the carbonyl group. The carbonyl group itself does not receive a Greek letter. The first carbon atom after the carbonyl group is the α carbon, the next is the β carbon, and so on.

 

Hydrogen atoms:

Hydrogen atoms are designated with the Greek letter of the carbon to which they are attached ⇒ the hydrogen atoms connected to α carbons are called α protons.

Enols and Tautomerization

Tautomers:

Rapidly interconverting constitutional isomers that differ from each other in the placement of a proton and the position of a double bond. 2 tautomers differ in the arrangement of their atoms and thus are not resonance structures.

In the presence of catalytic acid or base, carbonyls are in equilibrium with their enol forms, which for most carbonyls are very minor (10 kcal.mol-1 less stable).

 

Acid-catalyzed tautomerization:
 

Mechanism:

  1. Protonation of the carbonyl group to form a resonance-stabilized cation.
  2. Deprotonation of the cationic intermediate to form the enol.


     

 

Base-catalyzed tautomerization:
 

Mechanism:

  1. Deprotonation at the α position to form a resonance-stabilized anion.
  2. Protonation of the anionic intermediate to form the enol.

Preparation of Enolates

Enolate formation:

α-hydrogens are acidic ⇒ when treated with a strong base, the α position of a carbonyl is deprotonated to yield a resonance-stabilized intermediate called an enolate.
 

 

Bases for the deprotonation of aldehydes and ketones:

pKa values of the α hydrogens of a carbonyl group: ~ 16-18 (aldehyde) and ~ 19-21 (ketone).

  • KH, LDA, or BuLi (pKa > 21) will irreversibly and completely convert an aldehyde or ketone into an enolate.
  • EtONa (pKa = 16) or KOtBu (pKa = 18) will result in an equilibrium between an alkoxide ion and an enolate.

 

β-dicarbonyl acidity:

The acidity of β-dicarbonyl is due to the resonance stabilization of the corresponding anion
 

Enolate Reactivity

Enolates vs. enols:

  • Enolates are fully negatively charged and are therefore more reactive than enols.
  • Unlike enols, enolates can be isolated and stored for short periods of time.

 

Enolate reactivity:

Enolates are ambident nucleophiles because they have 2 nucleophilic sites: the oxygen atom and the α carbon. Although the oxygen atom of an enolate carries most of the negative charge due to its electronegativity, C-attacks are more common than O-attacks.

 

The 2 general types of reactions of enolates are:

  • nucleophilic substitutions (e.g., halogenation with X2, alkylation with RX)
  • reactions with other carbonyl compounds.

Alpha Halogenation

Alpha halogenation in acidic conditions:
 

Mechanism:

Part 1: Enol formation (acid-catalyzed tautomerization)

  1. Protonation of the carbonyl group to form a resonance-stabilized cation.
  2. Deprotonation of the cationic intermediate to form the enol.

Part 2: Halogenation

  1. Nucleophilic attack of the enol on the bromine to form a resonance-stabilized cation.
  2. Deprotonation of the carbonyl to form the desired product.


     

The enol formation is the rate-determining step of the reaction.

The acidic HBr is the by-product of α bromination, which is capable of catalyzing the nol formation ⇒ the reaction is autocatalytic.

 

Alpha halogenation in basic conditions:
 

Mechanism:

  1. Deprotonation of an α proton to form an enolate.
  2. Nucleophilic attack of the enolate on the bromine to form the desired product.


     

Limitations: The brominated product is more reactive and undergoes further bromination rapidly ⇒ if more than one α proton is present, it is difficult to achieve monobromination.
 

 


Alpha bromination of carboxylic acid:
 

Mechanism:

This process is called the Hell-Volhard-Zelinski reaction.

  1. Reaction between the carboxylic acid and PBr3 to form an acid bromide.
  2. Tautomerization of the acid bromide to form an acid halide enol.


     
  3. Nucleophilic attack of the enol on the bromine and deprotonation of the carbonyl to form the product of an α-halogenation.
  4. Hydrolysis of the acid bromide to regenerate the carboxylic acid. 

The Haloform Reaction

The haloform reaction:
 


Conversion of a methyl ketone into a carboxylic acid.

Mechanism:

  1. Replacement of α protons by bromine atoms, one at a time.
  2. Nucleophilic attack of the hydroxide on the carbonyl to form a tetrahedral intermediate.
  3. Loss of the leaving group, a tribromomethyl carbanion, to re-form the carbonyl.
  4. Deprotonation of the carboxylic acid to form a carboxylate ion and bromoform CHBr3 (shifting the equilibrium in favor of product formation).


     
  5. After the reaction is complete, protonation of the carboxylate ion in the presence of an acid.

The negative charge on CBr3- is stabilized by the electron-withdrawing effects of the 3 bromine atoms, making it a good leaving group. 

Enolate Alkylation

Alkylation via enolate ions:
 

Mechanism:

  1. Deprotonation of an α protons to form an enolate.
  2. Nucleophilic attack of the enolate on the alkyl halide in a SN2 reaction.


     


R-X must be a primary halide (SN2 reaction). With a secondary or tertiary alkyl halide, the enolate acts as a base and the alkyl halide undergoes elimination rather than substitution.

 

Alkylation of unsymmetrical ketones:

With an unsymmetrical ketone, 2 possible enolates can be formed:

  • The kinetic enolate (less substituted enolate). To favor the formation of the kinetic enolate, a sterically hindered base (e.g. LDA) is used to readily deprotonate the less hindered α position. The reaction is performed at low temperature (-78oC) to prevent equilibration of the enolates.
  • The thermodynamic enolate (the more substituted enolate). To favor the formation of the thermodynamic enolate, a nonsterically hindered base (e.g. NaH) is used, and the reaction is performed at room temperature.

 

Malonic Ester Synthesis

1,3-dicarboxylic acid decarboxylation:
 

Mechanism:

  1. Pericyclic reaction to form an enol.
  2. Tautomerization to form a carboxylic acid.


     

 

Malonic ester synthesis:
 


Conversion of an alkyl halide into a carboxylic acid with the introduction of 2 new carbon atoms from diethyl malonate.

Mechanism:

  1. Deprotonation of the relatively acidic diethyl malonate to form a doubly stabilized enolate.
  2. Nucleophilic attack of the enolate on the alkyl halide resulting in alkylation of the enolate.
  3. Acid-catalyzed hydrolysis of the esters to form a 1,3-dicarboxylic acid.
  4. Decarboxylation of the 1,3-dicarboxylic acid at high temperature to form the desired carboxylic acid.

 

 


Dialkylation of the diethyl malonate:

Diethyl malonate can also be dialkylated by a double deprotonation-nucleophilic attack of the enolate sequences. The hydrolysis followed by decarboxylation leads to the formation of a carboxylic acid with 2 alkyl groups at the α position.

 

Acetoacetic Ester Synthesis

Acetoacetic ester synthesis:
 


Conversion of an alkyl halide into a methyl ketone with the introduction of 3 new carbon atoms from ethyl acetoacetate. This process is very similar to the malonic ester synthesis.

Mechanism:

  1. Deprotonation of the relatively acidic ethyl acetoacetate to form a doubly stabilized enolate.
  2. Nucleophilic attack of the enolate on the alkyl halide resulting in alkylation of the enolate.
  3. Acid-catalyzed hydrolysis of the ester to form a β-keto acid.
  4. Decarboxylation of the β-keto acid at high temperature to form the desired methyl ketone.

 

 


Dialkylation of the ethyl acetoacetate:

Ethyl acetoacetate can also be dialkylated by a double deprotonation-nucleophilic attack of the enolate sequences. The hydrolysis followed by decarboxylation leads to the formation of a derivative of acetone with 2 alkyl groups at the α position.

 

Check your knowledge about this Chapter

The alpha position in organic chemistry refers to the carbon atom adjacent to a functional group, such as a carbonyl group (C=O). In the context of carbonyl compounds like aldehydes and ketones, the carbon atom directly bonded to the carbonyl carbon is termed as the α carbon, and the hydrogens attached to this α carbon are known as α hydrogens. The unique chemical environment of the alpha position allows for various important reactions, such as enolization and alkylation, due to the relatively acidic nature of the α hydrogens.

An enol is a species featuring a carbon-carbon double bond (alkene) that is directly connected to a hydroxyl group (-OH), while a ketone contains a carbonyl group (C=O) bonded to two carbon atoms and an aldehyde contains a carbonyl group bonded to at least one hydrogen atom. The 'en' in enol refers to the alkene and the 'ol' refers to the alcohol part of the molecule.

Enols can interconvert with ketones or aldehydes through a process known as tautomerization, where the double bond and hydrogen shift positions, leading to the formation of the more stable carbonyl-containing species.

Tautomerization is a chemical reaction that describes the transfer of a hydrogen atom and a double bond within a molecule, resulting in the conversion of one isomer (tautomer) into another. This process is particularly common between keto and enol tautomers where the hydrogen atom moves from an oxygen (in the case of ketones and aldehydes) to a carbon atom, generating an enol (a compound with a vinyl alcohol moiety).

Tautomerization is important in organic chemistry because it significantly influences the reactivity and stability of compounds, affects biochemical processes such as base pairing in DNA, and plays a crucial role in the mechanisms of certain organic reactions like enolization, which is key for the formation of carbon-carbon bonds in aldol reactions.

The keto-enol tautomeric shift occurs under acidic or basic conditions: 

  • In acidic conditions, the keto form is protonated, and a subsequent rearrangement leads to the formation of the enol.
  • Under basic conditions, a hydroxide ion abstracts a proton from the alpha-carbon, generating an enolate that can then tautomerize to the enol.

Enolates can be prepared using two common methods, namely, the deprotonation of ketones or aldehydes with a strong base, or through the use of a weaker base under equilibrium conditions.

  • Strong bases like LDA (Lithium Diisopropylamide) or NaHMDS (Sodium Hexamethyldisilazide) are typically used for the complete deprotonation at the alpha position.
  • Alternatively, bases such as sodium ethoxide can be used to generate enolates in a reversible manner, allowing for control over the reaction by removing the formed enol or enolate.

The choice of base and reaction conditions can influence the selectivity and yield of the enolate formation. 

Enolates are more reactive than enols due to their increased nucleophilicity. The negative charge on the oxygen atom in enolates increases their ability to participate in nucleophilic attacks on electrophilic centers. Enols, on the other hand, are neutral species and therefore less nucleophilic. They often require acid catalysis to react, whereas enolates can react under both acidic and basic conditions, taking part in a wider array of reactions such as alkylation and acylation.

Alpha halogenation involves the addition of a halogen atom to the α carbon in an aldehyde or ketone. This reaction typically proceeds via an enol intermediate, where the α carbon becomes nucleophilic due to the presence of the electron-withdrawing carbonyl group. The halogenation occurs through nucleophilic attack by a halogen source, such as bromine or chlorine, on the enol intermediate, followed by deprotonation to yield the alpha-halogenated carbonyl compound.

The haloform reaction is a chemical reaction where a methyl ketone is reacted with halogen and a base to produce a haloform (a trihalogenated methane such as chloroform or bromoform) and a carboxylate anion. Compounds that can undergo this reaction include methyl ketones and secondary alcohols that can be oxidized to methyl ketones. The reaction proceeds through multiple halogenation steps at the methyl group adjacent to the carbonyl, followed by the base-induced cleavage of the carbon-carbon bond, leading to the formation of the haloform and a carboxylate salt.

The key steps in enolate alkylation reactions involve the formation of an enolate ion from a carbonyl compound by deprotonation at the alpha position. A strong base such as LDA or NaH is typically used to remove the α hydrogen, forming a resonance-stabilized enolate anion. This nucleophilic enolate can then attack an alkyl halide (typically a primary alkyl halide to avoid side reactions such as elimination), leading to the formation of a new carbon-carbon bond at the alpha position.

Enolate alkylation involves the reaction of an enolate ion with an alkyl halide to form a new carbon-carbon bond. The nature of the alkyl halide greatly impacts the reaction’s outcome due to factors like reactivity and sterics.

  • Primary alkyl halides typically undergo SN2 reactions with enolates, leading to straightforward alkylation.
  • Secondary alkyl halides may also react via an SN2 mechanism, though they can be hindered by steric effects, which can slow down or prevent the reaction.
  • Tertiary alkyl halides are usually too hindered for SN2 reactions and can instead react through an SN1 mechanism or not at all, potentially leading to side reactions, such as eliminations or rearrangements.

The choice of alkyl halide is, therefore, crucial for achieving the desired alkylation product. 

The malonic ester synthesis is a method for preparing carboxylic acids with a substituent on the alpha carbon. It involves the alkylation of malonic ester followed by decarboxylation. This method is particularly useful because it allows the introduction of a wide range of substituents in a controlled fashion, enabling the synthesis of substituted acetic acids that are otherwise difficult to obtain directly through other methods.

Acetoacetic ester synthesis is a powerful method for the preparation of ketones. It begins with the alkylation of acetoacetic ester, an ester derived from a β-keto acid, to form a substituted acetoacetic ester. This substituted ester can then undergo hydrolysis and decarboxylation, effectively removing the two carbon atoms between the carbonyl group of the acetoacetate and the newly introduced alkyl group. The end product is a ketone with a carbon skeleton that has been extended by the length of the alkyl halide used in the alkylation step. This synthesis is particularly useful for creating ketones with complex substituents that may be difficult to introduce by other methods.

In malonic ester and acetoacetic ester syntheses, the decarboxylation step involves a pericyclic reaction where the β-keto acid intermediate undergoes the expulsion of carbon dioxide CO2 to form an enol intermediate. This enol then tautomerizes to produce the corresponding keto form, resulting in malonic acid in malonic ester synthesis and acetoacetic acid in acetoacetic ester synthesis.