Grignard reaction

Last updated

Contents

Classical Grignard reaction
Named after Victor Grignard
Reaction type Coupling reaction
Reaction
Aldehyde or Ketone
+
R-MgX
+
(H3O+)
Secondary/Tertiary Alcohols
Identifiers
Organic Chemistry Portal grignard-reaction
RSC ontology ID RXNO:0000014
A solution of a carbonyl compound is added to a Grignard reagent. (See gallery) Grignard reaction experiment 07.jpg
A solution of a carbonyl compound is added to a Grignard reagent. (See gallery)

The Grignard reaction (French: [ɡʁiɲaʁ] ) is an organometallic chemical reaction in which, according to the classical definition, carbon alkyl, allyl, vinyl, or aryl magnesium halides (Grignard reagent) are added to the carbonyl groups of either an aldehyde or ketone under anhydrous conditions. [1] [2] [3] This reaction is important for the formation of carbon-carbon bonds. [4] [5]

(R2 or R3 could be hydrogen)

Grignard reaction scheme.svg

History and definitions

Grignard reactions and reagents were discovered by and are named after the French chemist François Auguste Victor Grignard (University of Nancy, France), who described them in 1900. [6] He was awarded the 1912 Nobel Prize in Chemistry for this work. [7]

The reaction of an organic halide with magnesium is not a Grignard reaction, but provides a Grignard reagent. [8]

Although Grignard reagents undergo many reactions, the classical Grignard reaction refers only to the reaction of RMgX with ketones and aldehydes, shown in red. X = Cl, Br, I. Grignard with carbonyl.png
Although Grignard reagents undergo many reactions, the classical Grignard reaction refers only to the reaction of RMgX with ketones and aldehydes, shown in red. X = Cl, Br, I.

Classically, the Grignard reaction refers to the reaction between a ketone or aldehyde group with a Grignard reagent to form a primary or tertiary alcohol. [1] However, some chemists understand the definition to mean all reactions of any electrophiles with Grignard reagents. [9] Therefore, there is some dispute about the modern definition of the Grignard reaction. In the Merck Index, published online by the Royal Society of Chemistry, the classical definition is acknowledged, followed by "A more modern interpretation extends the scope of the reaction to include the addition of Grignard reagents to a wide variety of electrophilic substrates." [9] This variety of definitions illustrates that there is some dispute within the chemistry community about the definition of a Grignard reaction.

Shown below are some reactions involving Grignard reagents, but they themselves are not classically understood as Grignard reactions.

Additional reactions which involve Grignard reagents, but are not considered to be Grignard reactions by the classical definition. X = Cl, Br, I. Grignard with others.png
Additional reactions which involve Grignard reagents, but are not considered to be Grignard reactions by the classical definition. X = Cl, Br, I.

Reaction mechanism

Because carbon is more electronegative than magnesium, the carbon attached to magnesium acts as a nucleophile and attacks the electrophilic carbon atom in the polar bond of a carbonyl group. The addition of the Grignard reagent to the carbonyl group typically proceeds through a six-membered ring transition state, as shown below. [10]

The mechanism of the Grignard reaction. Grignard-Reaction Mechanism.png
The mechanism of the Grignard reaction.

Based on the detection of radical coupling side products, an alternative single electron transfer (SET) mechanism that involves the initial formation of a ketyl radical intermediate has also been proposed. [11] A recent computational study suggests that the operative mechanism (polar vs. radical) is substrate-dependent, with the reduction potential of the carbonyl compound serving as a key parameter. [12]

Conditions

If a Grignard reaction is performed in the presence of water, or any labile proton, the labile proton will quench the Grignard reagent as shown in the figure above. Grignard reagents with acidic protons.png
If a Grignard reaction is performed in the presence of water, or any labile proton, the labile proton will quench the Grignard reagent as shown in the figure above.

The Grignard reaction is conducted under anhydrous conditions. [3] Otherwise, the reaction will fail because the Grignard reagent will act as a base rather than a nucleophile and pick up a labile proton rather than attacking the electrophilic site. This will result in no formation of the desired product as the R-group of the Grignard reagent will become protonated while the MgX portion will stabilize the deprotonated species.

To prevent this, Grignard reactions are completed in an inert atmosphere to remove all water from the reaction flask and ensure that the desired product is formed. [13] Additionally, if there are acidic protons in the starting material, as shown in the figure on the right, one can overcome this by protecting the acidic site of the reactant by turning it into an ether or a silyl ether to eliminate the labile proton from the solution prior to the Grignard reaction.

Variants

Other variations of the Grignard reagent have been discovered to improve the chemoselectivity of the Grignard reaction, which include but are not limited to: Turbo-Grignards, organocerium reagents, and organocuprate (Gilman) reagents.

Turbo-Grignards

Turbo-Grignards are Grignard reagents modified with lithium chloride. Compared to conventional Grignard reagents, Turbo-Grignards are more chemoselective; esters, amides, and nitriles do not react with the Turbo-Grignard reagent. [14]

An example reaction of forming a Turbo-Grignard with an ester group. Turbo-Grignard formation.png
An example reaction of forming a Turbo-Grignard with an ester group.

Heterometal-modified Grignard reagents

A conjugated 1,4 addition using a Gilman reagent with an arbitrary R group Cuprate conjugate addition with lewis acid.png
A conjugated 1,4 addition using a Gilman reagent with an arbitrary R group

The behavior of Grignard reagents can be usefully modified in the present of other metals. Copper(I) salts give organocuprates that preferentially effect 1,4 addition. [15] Cerium trichloride allows selective 1,2-additions to the same substrates. Nickel and palladium halides catalyze cross coupling reactions.

See also

Related Research Articles

<span class="mw-page-title-main">Carboxylic acid</span> Organic compound containing a –C(=O)OH group

In organic chemistry, a carboxylic acid is an organic acid that contains a carboxyl group attached to an R-group. The general formula of a carboxylic acid is often written as R−COOH or R−CO2H, sometimes as R−C(O)OH with R referring to the alkyl, alkenyl, aryl, or other group. Carboxylic acids occur widely. Important examples include the amino acids and fatty acids. Deprotonation of a carboxylic acid gives a carboxylate anion.

<span class="mw-page-title-main">Ketone</span> Organic compounds of the form >C=O

In organic chemistry, a ketone is an organic compound with the structure R−C(=O)−R', where R and R' can be a variety of carbon-containing substituents. Ketones contain a carbonyl group −C(=O)−. The simplest ketone is acetone, with the formula (CH3)2CO. Many ketones are of great importance in biology and in industry. Examples include many sugars (ketoses), many steroids, and the solvent acetone.

The Friedel–Crafts reactions are a set of reactions developed by Charles Friedel and James Crafts in 1877 to attach substituents to an aromatic ring. Friedel–Crafts reactions are of two main types: alkylation reactions and acylation reactions. Both proceed by electrophilic aromatic substitution.

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

The aldol reaction is a reaction in organic chemistry that combines two carbonyl compounds to form a new β-hydroxy carbonyl compound. Its simplest form might involve the nucleophilic addition of an enolized ketone to another:

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

<span class="mw-page-title-main">Victor Grignard</span> French chemist (1871–1935)

Francois Auguste Victor Grignard was a French chemist who won the Nobel Prize for his discovery of the eponymously named Grignard reagent and Grignard reaction, both of which are important in the formation of carbon–carbon bonds.

<span class="mw-page-title-main">Enolate</span> Organic anion formed by deprotonating a carbonyl (>C=O) compound

In organic chemistry, enolates are organic anions derived from the deprotonation of carbonyl compounds. Rarely isolated, they are widely used as reagents in the synthesis of organic compounds.

Nucleophilic acyl substitution describes a class of substitution reactions involving nucleophiles and acyl compounds. In this type of reaction, a nucleophile – such as an alcohol, amine, or enolate – displaces the leaving group of an acyl derivative – such as an acid halide, anhydride, or ester. The resulting product is a carbonyl-containing compound in which the nucleophile has taken the place of the leaving group present in the original acyl derivative. Because acyl derivatives react with a wide variety of nucleophiles, and because the product can depend on the particular type of acyl derivative and nucleophile involved, nucleophilic acyl substitution reactions can be used to synthesize a variety of different products.

The Weinreb ketone synthesis or Weinreb–Nahm ketone synthesis is a chemical reaction used in organic chemistry to make carbon–carbon bonds. It was discovered in 1981 by Steven M. Weinreb and Steven Nahm as a method to synthesize ketones. The original reaction involved two subsequent nucleophilic acyl substitutions: the conversion of an acid chloride with N,O-Dimethylhydroxylamine, to form a Weinreb–Nahm amide, and subsequent treatment of this species with an organometallic reagent such as a Grignard reagent or organolithium reagent. Nahm and Weinreb also reported the synthesis of aldehydes by reduction of the amide with an excess of lithium aluminum hydride.

The Blaise ketone synthesis is the chemical reaction of acid chlorides with organozinc compounds to give ketones.

The Sakurai reaction is the chemical reaction of carbon electrophiles with allyltrimethylsilane catalyzed by strong Lewis acids. The reaction achieves results similar to the addition of an allyl Grignard reagent to the carbonyl.

<span class="mw-page-title-main">Grignard reagent</span> Organometallic compounds used in organic synthesis

Grignard reagents or Grignard compounds are chemical compounds with the general formula R−Mg−X, where X is a halogen and R is an organic group, normally an alkyl or aryl. Two typical examples are methylmagnesium chloride Cl−Mg−CH3 and phenylmagnesium bromide (C6H5)−Mg−Br. They are a subclass of the organomagnesium compounds.

In organosilicon chemistry, silyl enol ethers are a class of organic compounds that share the common functional group R3Si−O−CR=CR2, composed of an enolate bonded to a silane through its oxygen end and an ethene group as its carbon end. They are important intermediates in organic synthesis.

<span class="mw-page-title-main">Organozinc chemistry</span>

Organozinc chemistry is the study of the physical properties, synthesis, and reactions of organozinc compounds, which are organometallic compounds that contain carbon (C) to zinc (Zn) chemical bonds.

<span class="mw-page-title-main">Stork enamine alkylation</span> Reaction sequence in organic chemistry

The Stork enamine alkylation involves the addition of an enamine to a Michael acceptor or another electrophilic alkylation reagent to give an alkylated iminium product, which is hydrolyzed by dilute aqueous acid to give the alkylated ketone or aldehyde. Since enamines are generally produced from ketones or aldehydes, this overall process constitutes a selective monoalkylation of a ketone or aldehyde, a process that may be difficult to achieve directly.

Organomanganese chemistry is the chemistry of organometallic compounds containing a carbon to manganese chemical bond. In a 2009 review, Cahiez et al. argued that as manganese is cheap and benign, organomanganese compounds have potential as chemical reagents, although currently they are not widely used as such despite extensive research.

<span class="mw-page-title-main">Carbonyl reduction</span> Organic reduction of any carbonyl group by a reducing agent

In organic chemistry, carbonyl reduction is the conversion of any carbonyl group, usually to an alcohol. It is a common transformation that is practiced in many ways. Ketones, aldehydes, carboxylic acids, esters, amides, and acid halides - some of the most pervasive functional groups, -comprise carbonyl compounds. Carboxylic acids, esters, and acid halides can be reduced to either aldehydes or a step further to primary alcohols, depending on the strength of the reducing agent. Aldehydes and ketones can be reduced respectively to primary and secondary alcohols. In deoxygenation, the alcohol group can be further reduced and removed altogether by replacement with H.

<span class="mw-page-title-main">Reductions with samarium(II) iodide</span>

Reductions with samarium(II) iodide involve the conversion of various classes of organic compounds into reduced products through the action of samarium(II) iodide, a mild one-electron reducing agent.

Reactions of organocopper reagents involve species containing copper-carbon bonds acting as nucleophiles in the presence of organic electrophiles. Organocopper reagents are now commonly used in organic synthesis as mild, selective nucleophiles for substitution and conjugate addition reactions.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

References

  1. 1 2 Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, ISBN   978-0-471-72091-1
  2. Chapter 19: Carboxylic Acids. Organic Chemistry 4e Carey. mhhe.com
  3. 1 2 3 Ouellette, Robert J.; Rawn, J. David (January 1, 2014), Ouellette, Robert J.; Rawn, J. David (eds.), "15 - Alcohols: Reactions and Synthesis", Organic Chemistry, Boston: Elsevier, pp. 491–534, doi:10.1016/b978-0-12-800780-8.00015-2, ISBN   978-0-12-800780-8 , retrieved November 6, 2023
  4. Shirley, D. A. (1954). "The Synthesis of Ketones from Acid Halides and Organometallic Compounds of Magnesium, Zinc, and Cadmium". Org. React. 8: 28–58.
  5. Huryn, D. M. (1991). "Carbanions of Alkali and Alkaline Earth Cations: (ii) Selectivity of Carbonyl Addition Reactions". In Trost, B. M.; Fleming, I. (eds.). Comprehensive Organic Synthesis, Volume 1: Additions to CX π-Bonds, Part 1. Elsevier Science. pp. 49–75. doi:10.1016/B978-0-08-052349-1.00002-0. ISBN   978-0-08-052349-1.
  6. texte, Académie des sciences (France) Auteur du (January 1, 1900). "Comptes rendus hebdomadaires des séances de l'Académie des sciences / publiés... par MM. les secrétaires perpétuels". Gallica. Retrieved April 23, 2023.
  7. Grignard, V. (1900). "Sur quelques nouvelles combinaisons organométaliques du magnésium et leur application à des synthèses d'alcools et d'hydrocabures". Compt. Rend. 130: 1322–25.
  8. IUPAC. Compendium of Chemical Terminology, 2nd ed. (the "Gold Book"). Compiled by A. D. McNaught and A. Wilkinson. Blackwell Scientific Publications, Oxford (1997). ISBN   0-9678550-9-8. doi : 10.1351/goldbook.
  9. 1 2 "Grignard Reaction | The Merck Index Online". www.rsc.org. Retrieved April 23, 2023.
  10. Maruyama, K.; Katagiri, T. (1989). "Mechanism of the Grignard reaction". J. Phys. Org. Chem. 2 (3): 205–213. doi:10.1002/poc.610020303.
  11. Ashby, E. C.; Goel, A. B. (August 1981). "Direct evidence supporting a single electron transfer pathway in the reduction of ketones by primary, secondary, and tertiary Grignard reagents". Journal of the American Chemical Society. 103 (16): 4983–4985. doi:10.1021/ja00406a070. ISSN   0002-7863.
  12. Peltzer, Raphael Mathias; Gauss, Jürgen; Eisenstein, Odile; Cascella, Michele (February 12, 2020). "The Grignard Reaction – Unraveling a Chemical Puzzle". Journal of the American Chemical Society. 142 (6): 2984–2994. doi:10.1021/jacs.9b11829. hdl: 10852/83918 . ISSN   0002-7863. PMID   31951398. S2CID   210709021.
  13. Carey, Francis A. "Grignard reagent". Britannica.
  14. Hermann, Andreas; Seymen, Rana; Brieger, Lukas; Kleinheider, Johannes; Grabe, Bastian; Hiller, Wolf; Strohmann, Carsten (2023). "Comprehensive Study of the Enhanced Reactivity of Turbo‐Grignard Reagents*". Angewandte Chemie International Edition. 62 (25). doi: 10.1002/anie.202302489 . S2CID   257765567.
  15. Woodward, Simon (January 1, 2000). "Decoding the 'black box' reactivity that is organocuprate conjugate addition chemistry". Chemical Society Reviews. 29 (6): 393–401. doi:10.1039/B002690P. ISSN   1460-4744.