User:Elbert Ainsteinium/Michael reaction

From Wikipedia, the free encyclopedia

Examples[edit]

Asymmetric Michael reaction[edit]

*NOTE: THIS IS THE ORIGINAL SECTION WITH NO EDITS. REFORMATTING THIS TO BE UNDER ONE TITLE: EXAMPLES

Researchers have expanded the scope of Michael additions to include elements of chirality via asymmetric versions of the reaction. The most common methods involve chiral phase transfer catalysis, such as quaternary ammonium salts derived from the Cinchona alkaloids; or organocatalysis, which is activated by enamine or iminium with chiral secondary amines, usually derived from proline.

In the reaction between cyclohexanone and β-nitrostyrene sketched below, the base proline is derivatized and works in conjunction with a protic acid such as p-toluenesulfonic acid:

Syn addition is favored with 99% ee. In the transition state believed to be responsible for this selectivity, the enamine (formed between the proline nitrogen and the cycloketone) and β-nitrostyrene are co-facial with the nitro group hydrogen bonded to the protonated amine in the proline side group.

A well-known Michael reaction is the synthesis of warfarin from 4-hydroxycoumarin and benzylideneacetone first reported by Link in 1944:

Several asymmetric versions of this reaction exist using chiral catalysts.

Mukaiyama-Michael addition[edit]

*NOTE: THIS IS THE ORIGINAL SECTION WITH NO EDITS. REFORMATTING THIS TO BE UNDER ONE TITLE: EXAMPLES

In the Mukaiyama–Michael addition the nucleophile is a silyl enol ether and the catalyst usually titanium tetrachloride:


1,6-Michael reaction[edit]

The 1,6-Michael reaction proceeds via nucleophilic attack on the 𝛿 carbon of an α,β-,𝛿-diunsaturated Michael acceptor.[1][2] The 1,6-addition mechanism is similar to the 1,4-addition, with one exception being the nucleophilic attack occurring at the 𝛿 carbon of the Michael acceptor.[2] However, research shows that organocatalysis often favours the 1,4-addition.[1] In many syntheses where 1,6-addition was favoured, the substrate contained certain structural features.[2] Research has shown that catalysts can influence the regioselectivity and enantioselectivity of a 1,6-addition reaction.[2]

For example, the image below shows the addition of ethylmagnesium bromide to ethyl sorbate 1 using a copper catalyst with a reversed josiphos (R,S)-(–)-3 ligand.[2] This reaction produced the 1,6-addition product 2 in 0% yield, the 1,6-addition product 3 in approximately 99% yield, and the 1,4-addition product 4 in less than 2% yield. This particular catalyst and set of reaction conditions led to the mostly regioselective and enantioselective 1,6-Michael addition of ethyl sorbate 1 to product 3.

The Michael addition of ethylmagnesium bromide to ethyl sorbate.
The Michael addition of ethylmagnesium bromide to ethyl sorbate.
















Applications[edit]

Pharmaceuticals[edit]

A Michael reaction is used as a mechanistic step by many covalent inhibitor drugs. Cancer drugs such as ibrutinib, osimertinib, and rociletinib have an acrylamide functional group as a Micheal acceptor. The Micheal acceptor on the drug reacts with a Michael acceptor in the active site of an enzyme. This is a viable cancer treatment because the target enzyme is inhibited following the Michael reaction.[3]

Polymerization reactions[edit]

Mechanism[4][edit]

All polymerization reactions have three basic steps: initiation, propagation, and termination. The initiation step is the Michael addition of the nucleophile to a monomer. The resultant species undergoes a Michael addition with another monomer, with the latter acting as an acceptor. This extends the chain by forming another nucleophilic species to act as a donor for the next addition. This process repeats until the reaction is quenched by chain termination.[5] The original Michael donor can be a neutral donor such as amines, thiols, and alkoxides, or alkyl ligands bound to a metal. [6]

Polymerization mechanism of a Michael addition with a thiol nucleophile

Examples[edit]

Linear step growth polymerizations are some of the earliest applications of the Michael reaction in polymerizations. A wide variety of Michael donors and acceptors have been used to synthesize a diverse range of polymers. Examples of such polymers include poly(amido amine), poly(amino ester), poly(imido sulfide), poly(ester sulfide), poly(aspartamide), poly(imido ether), poly(amino quinone), poly(enone sulfide) and poly(enamine ketone).

For example, linear step growth polymerization produces the redox active poly(amino quinone), which serves as an anti-corrosion coatings on various metal surfaces.[7] Another example includes network polymers, which are used for drug delivery, high performance composites, and coatings. These network polymers are synthesized using a dual chain growth, photo-induced radical and step growth Michael addition system.



Poly(amino quinone)
Poly(amino quinone)
Poly(amido-amine)














Article Draft[edit]

Lead[edit]

y
y


Deprotonation of 1 by a base leads to carbanion 2, stabilized by its electron-withdrawing groups. Structures 2a to 2c are three resonance structures that can be drawn for this species, two of which have enolate ions. This nucleophile reacts with the electrophilic alkene 3 to form 4 in a conjugate addition reaction. Finally, enolate 4 abstracts a proton from protonated base (or solvent) to produce 5.

The reaction is dominated by orbital, rather than electrostatic, considerations. The HOMO of stabilized enolates has a large coefficient on the central carbon atom while the LUMO of many alpha, beta unsaturated carbonyl compounds has a large coefficient on the beta carbon. Thus, both reactants can be considered soft. These polarized frontier orbitals are of similar energy, and react efficiently to form a new carbon–carbon bond.[8]


Like the aldol addition, the Michael reaction may proceed via an enol, silyl enol ether in the Mukaiyama–Michael addition, or more usually, enolate nucleophile. In the latter case, the stabilized carbonyl compound is deprotonated with a strong base (hard enolization) or with a Lewis acid and a weak base (soft enolization). The resulting enolate attacks the activated olefin with 1,4-regioselectivity, forming a carbon–carbon bond. This also transfers the enolate to the electrophile. Since the electrophile is much less acidic than the nucleophile, rapid proton transfer usually transfers the enolate back to the nucleophile if the product is enolizable; however, one may take advantage of the new locus of nucleophilicity if a suitable electrophile is pendant. Depending on the relative acidities of the nucleophile and product, the reaction may be catalytic in base. In most cases, the reaction is irreversible at low temperature. ---citation?---

Article body[edit]

Kohler paragraph:

*The paragraph below has been copied from the Michael reaction Wikipedia page and will be edited here before transferring our final edits to the main page:

As originally defined by Arthur Michael, the reaction is the addition of an enolate of a ketone or aldehyde to an α,β-unsaturated carbonyl compound at the β carbon. A newer definition, proposed by Kohler, is the 1,4-addition of a doubly stabilized carbon nucleophile to an α,β-unsaturated carbonyl compound. Some examples of nucleophiles include beta-ketoesters, malonates, and beta-cyanoesters. The resulting product contains a highly useful 1,5-dioxygenated pattern.


*Edited version of paragraph (fixing for citation about Kohler which doesn’t appear to exist):

Arthur Michael originally defined the Michael reaction as the addition of an enolate of a ketone or aldehyde to an α,β-unsaturated carbonyl compound at the β carbon. The current definition of the Michael reaction has broadened to include nucleophiles other than enolates.[4] Some examples of nucleophiles include doubly stabilized carbon nucleophiles such as beta-ketoesters, malonates, and beta-cyanoesters. The resulting product contains a highly useful 1,5-dioxygenated pattern. Non-carbon nucleophiles such as water, alcohols, amines, and enamines can also react with an α,β-unsaturated carbonyl in a 1,4-addition.[9]

References[edit]

  1. ^ a b Hayashi, Yujiro; Okamura, Daichi; Umemiya, Shigenobu; Uchimaru, Tadafumi (2012-07). "Organocatalytic 1,4-Addition Reaction of α,β-γ,δ-Diunsaturated Aldehydes versus 1,6-Addition Reaction". ChemCatChem. 4 (7): 959–962. doi:10.1002/cctc.201200161. {{cite journal}}: Check date values in: |date= (help)
  2. ^ a b c d e den Hartog, Tim; Harutyunyan, Syuzanna R.; Font, Daniel; Minnaard, Adriaan J.; Feringa, Ben L. (2008-01). "Catalytic Enantioselective 1,6-Conjugate Addition of Grignard Reagents to Linear Dienoates". Angewandte Chemie International Edition. 47 (2): 398–401. doi:10.1002/anie.200703702. {{cite journal}}: Check date values in: |date= (help)
  3. ^ Boike, Lydia; Henning, Nathaniel J.; Nomura, Daniel K. (2022-08-25). "Advances in covalent drug discovery". Nature Reviews Drug Discovery. doi:10.1038/s41573-022-00542-z. ISSN 1474-1776. PMC 9403961. PMID 36008483.{{cite journal}}: CS1 maint: PMC format (link)
  4. ^ a b Mather, Brian D.; Viswanathan, Kalpana; Miller, Kevin M.; Long, Timothy E. (2006-05). "Michael addition reactions in macromolecular design for emerging technologies". Progress in Polymer Science. 31 (5): 487–531. doi:10.1016/j.progpolymsci.2006.03.001. {{cite journal}}: Check date values in: |date= (help)
  5. ^ Huang, Sijia; Sinha, Jasmine; Podgórski, Maciej; Zhang, Xinpeng; Claudino, Mauro; Bowman, Christopher N. (2018-08-14). "Mechanistic Modeling of the Thiol–Michael Addition Polymerization Kinetics: Structural Effects of the Thiol and Vinyl Monomers". Macromolecules. 51 (15): 5979–5988. doi:10.1021/acs.macromol.8b01264. ISSN 0024-9297.
  6. ^ Jung, Hyuk-Joon; Yu, Insun; Nyamayaro, Kudzanai; Mehrkhodavandi, Parisa (2020-06-05). "Indium-Catalyzed Block Copolymerization of Lactide and Methyl Methacrylate by Sequential Addition". ACS Catalysis. 10 (11): 6488–6496. doi:10.1021/acscatal.0c01365. ISSN 2155-5435.
  7. ^ Pham, M.C; Hubert, S; Piro, B; Maurel, F; Le Dao, H; Takenouti, H (2004-02). "Investigations of the redox process of conducting poly(2-methyl-5-amino-1,4-naphthoquinone) (PMANQ) film". Synthetic Metals. 140 (2–3): 183–197. doi:10.1016/S0379-6779(03)00373-4. {{cite journal}}: Check date values in: |date= (help)
  8. ^ Perlmutter, P., ed. (1992-01-01), "Chapter One - Introduction", Tetrahedron Organic Chemistry Series, Conjugate Addition Reactions in Organic Synthesis, vol. 9, Elsevier, pp. 1–61, doi:10.1016/b978-0-08-037067-5.50007-2, retrieved 2022-11-19
  9. ^ Brown, William Henry (2018). Organic chemistry. Brent L. Iverson, Eric V. Anslyn, Christopher S. Foote (Eighth edition ed.). Boston, Mass. ISBN 978-1-337-51640-2. OCLC 1200494733. {{cite book}}: |edition= has extra text (help)CS1 maint: location missing publisher (link)