fbpx
Wikipedia

Baylis–Hillman reaction

In organic chemistry, the Baylis–Hillman, Morita–Baylis–Hillman, or MBH reaction is a carbon-carbon bond-forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst, such as a tertiary amine or phosphine. The product is densely functionalized, joining the alkene at the α-position to a reduced form of the electrophile (e.g. in the case of an aldehyde, an allylic alcohol).[1][2]

(Morita–)Baylis–Hillman reaction
Named after Ken-ichi Morita
Anthony B. Baylis
Melville E. D. Hillman
Reaction type Coupling reaction
Identifiers
Organic Chemistry Portal baylis-hillman-reaction
RSC ontology ID RXNO:0000076
Baylis-Hillman reaction

The reaction is named for Anthony B. Baylis and Melville E. D. Hillman, two of the chemists who developed the reaction at Celanese; and K. Morita, who published earlier work[3] on the same.

The MBH reaction offers several advantages in organic synthesis:

  1. It combines easily prepared starting materials with high atom economy.
  2. It requires only mild conditions and does not require any transition metals.
  3. Asymmetric synthesis is possible from prochiral electrophiles.
  4. The product's dense functionalization enables many further transformations.

Its disadvantage is that the reaction is extremely slow.

Common reagents edit

The most frequently-used catalyst for the reaction is the tertiary amine DABCO (triethylenediamine); other known catalysts include 4-dimethylaminopyridine, DBU (diaza­bicyclo­undecene), and various phosphines.

Reaction mechanism edit

As of 2012, certain questions about MBH reaction's mechanism remain open.

Hill and Isaacs performed the first kinetic experiments in the 1990s, discovering that the reaction rate between acrylonitrile and acetaldehyde was first-order in each reactant and in the DABCO catalyst. α-Deuterated acrylonitrile exhibited no kinetic isotope effect, but the product appeared to catalyze its own generation.[4]

In a model Hoffmann proposed seven years prior,[5] the reaction begins with 1,4-addition of the catalytic amine to the activated alkene. The resulting zwitterionic aza-enolate undergoes aldol addition to the aldehyde. Intramolecular proton shift then generates the final MBH adduct, which eliminates the catalyst.

 

If Hoffmann's model were correct, then the aldol addition would be the rate-limiting step, which accords with the absent kinetic isotope effect.[4] However, Hoffman's mechanism rationalizes neither the product's autocatalysis nor (in the reaction of aryl aldehydes with acrylates) the considerable generation of a dioxanone byproduct.

In more recent work, McQuade et al and Aggarwal et al reevaluated the MBH mechanism, focusing on the proton-transfer and autocatalysis.[6][7] According to McQuade, the reaction between methyl acrylate and p-nitrobenzaldehyde is second-order relative to the aldehyde. Moreover, it showed a significant kinetic isotope effect for the acrylate's α-hydrogen (5.2 in DMSO, but ≥2 in all solvents), which would imply that proton abstraction is the rate-determining step.

To account for this reanalysis, McQuade proposed modifying Hoffman's proposal, such that after the first aldol addition, a second aldol adds to form a hemiacetal alkoxide. Then the rate-determining proton transfer releases the adduct A via a six-membered transition state, which further reacts to produce the MBH product B or a dioxanone byproduct C.

In a further modification, Aggarwal noted that observed that methanol inhibited autocatalysis in the MBH reaction. Thus he proposed that in the reaction's early stages, a mechanism equivalent to McQuade's proposal operates, but after 20% conversion, reaction with an alcoholic solvent can replace the second aldol addition. In support of this contention, Aggarwal and Harvey modeled the two pathways using density functional theory calculations and showed that the computed energy profile matches the experimental kinetic isotope effects and observed rate of reaction.[8] Also they showed that the overall enthalpic barrier of the alcohol-catalyzed pathway is slightly smaller than that of the non-alcohol-catalyzed pathway, rationalizing that as the alcohol (MBH product) concentration increases the alcohol-catalyzed pathway starts to dominate.

 

While McQuade's and Aggarwal's studies received much attention, unequivocal proof of intermediate A's role remained elusive as of 2012. Because A could be formed by separate addition of B to an aldehyde, any isolated A and C could be the result of side reactions, rather than the MBH itself. Nor has a convincing explanation yet been presented for Hill and Isaac's original first-order data.

Aggarwal's modification has its own controversies. If it is correct, then the rate-determining step changes from proton transfer to aldol addition over the course of the reaction[7] — but subsequent computational studies have concluded that proton transfer still has the highest barrier even late into the reaction's process. On the other hand, Coelho and Eberlin et al have obtained electrospray-mass-spectroscopy data that is structural evidence for two different forms of the reaction's proton transfer step.[9]

Scope and limitations edit

 

The MBH reaction is extremely general. In most cases the electrophile is an aldehyde, ketone (but see below), or imine (latterly the aza-Baylis–Hillman reaction); but reports indicate that allyl halides, alkyl halides, and epoxides are also possible.[10][11][12] Using an allene instead of a simple alkene as the precursor gives an intermediate that can react at the γ carbon rather than at the α.[13]

At the same time, it can be challenging to develop suitable reaction conditions. The reaction is slow (times of a fortnight or longer are not uncommon, even with 25-100 mol % catalyst), especially with (as alkene) β-substituted activated olefins, vinyl sulfones, or vinyl sulfoxides; or (as electrophile) hindered aliphatic aldehydes or electron-rich benzaldehydes. Ketones are generally not reactive enough under ordinary conditions to take part in a synthetically useful manner.[14] For example, reaction between sterically hindered t-butyl acrylate and benzaldehyde with catalytic DABCO in the absence of solvent required 4 weeks to give moderate conversion to the final product.

 

In aprotic solvents, the reaction rate is even slower, although recovery is possible with protic additives (e.g. alcohols and carboxylic acids).[15]

At such low rates, the activity of the substrates may induce competing side-reactions: acroleins also oligomerize and allenoates cycloadd. Allyl-halide and alkyl-epoxide electrophiles also often prove unruly.[clarification needed][citation needed] The MBH reaction of an aryl vinyl ketone with an aldehyde is not straightforward (but see § Sila-MBH reaction), since the reactive aryl vinyl ketone readily undergoes Michael addition to another molecule of the aryl vinyl ketone, which then adds to the aldehyde to form a double-MBH adduct.[16]

 

Due to the highly negative volume of activation, sluggish Baylis–Hillman reactions — including ketonic ones — can be realized by conducting the reaction under high pressure (up to 20 kbar).[14]

Variants edit

Sila-MBH reaction edit

 

In the sila-MBH reaction, α-silylated vinyl aryl ketones couple to aldehydes in the presence of catalytic TTMPP, a large triarylphosphine reagent.[17] The zwitterionic enolate produced upon addition of nucleophilic catalyst to the enone adds to an aldehyde carbonyl to generate an alkoxide. This alkoxide undergoes a subsequent 1,3-Brook rearrangement and elimination cascade to afford a siloxy-methylene enone and release the catalyst.

Rauhut-Currier reaction edit

 

The Rauhut-Currier reaction is a vinylogous analogoue of the MBH reaction, in which the electrophile is a Michael acceptor, not an aldehyde or an imine. Intermolecular Rauhut-Currier reactions typically exhibit poor chemoselectivity, because the reaction couples two activated alkenes, but intramolecular Rauhut-Currier reactions have been employed. For example, cyclization of α,β-unsaturated aldehydes can be performed in the presence of proline derivative and acetic acid, affording enantioenriched products.[18]

Tandem strategies edit

As mentioned above, the slow rate of the MBH reaction often enables side-reactions on its activated substrates. In tandem reaction strategies, this is a virtue, for it enables syntheses with high atom economy. For example, in the three-component coupling of aldehydes, amines, and activated alkenes, the aldehyde reacts with the amine to produce an imine prior to forming the aza-MBH adduct, as in the reaction of aryl aldehydes, diphenylphosphinamide, and methyl vinyl ketone, in the presence of TiCl4, triphenylphosphine, and triethylamine:[19]

 

Likewise, activated acetylenes can undergo conjugate addition and remain an activated alkene for the MBH reaction, as in the following enantioselective cyclization reaction in which a phenolate nucleophile adds to a functionalized enyne before aza-MBH ring closure catalyzed by a chiral amine base.[20]

 

Asymmetric synthesis edit

Chiral auxiliaries edit

Oppolzer's sultam can be used as a chiral auxiliary for an asymmetric MBH reaction. When an acrylate substituted with the Oppolzer's sultam reacted with various aldehydes in the presence of DABCO catalyst, optically pure 1,3-dioxan-4-ones were afforded with cleavage of the auxiliary (67-98% yield, >99% ee). The cyclic products could be converted into desired MBH products by use of camphorsulfonic acid and methanol.[21]

 

A related hydrazide auxiliary is the chiral acryloylhydrazide, which reacts diastereoselectively with aldehydes.[22] Both diastereomers could be obtained with different choice of solvents (DMSO vs. mixed THF and H2O), suggesting that the transition structure conformation is solvent-influenced.

Chiral allenes and imines can also be employed for an asymmetric DABCO-catalyzed aza-MBH reaction.[23] Optically active 10-phenylsulfonylisobornyl buta-2,3-dienoate reacts with an aryl imine to afford α-allenylamine in a diastereoselective manner (37-57% yield).

 

Chiral Lewis-basic catalyst edit

Some enantioselective MBH reactions employ chiral tertiary amine catalysts. For example, β-ICD, a cinchona alkaloid derivative, is famous among the quinidine framework-based catalysts, and catalyzed an enantioselective MBH reaction with 1,1,1,3,3,3,-hexafluoroisopropyl acrylate as the activated alkene:[24]

 

The phenolic oxygen of β-ICD was shown to be important in the reaction, implying that β-ICD acts as a Bronsted acid, not just a nucleophile.

Cyclopentenone and various aromatic and aliphatic aldehydes undergo an asymmetric reaction using Fu's planar chiral DMAP catalyst in isopropanol (54-96% yield, 53-98% ee). In this case, magnesium iodide as a Lewis acid cocatalyst was required to accelerate the reaction.[25]

 

P-Chiral phosphines have been investigated.[26]

Simple diamines can also be employed as MBH catalysts. Methyl vinyl ketone and various substituted benzaldehydes were found to undergo asymmetric MBH reaction. The chiral pyrrolidine catalyst was effective for ortho- and para-substituted electron-deficient benzaldehydes (75-99% yield, 8-73% ee).[27]

 

Chiral phosphine MBH catalysts often contain Bronsted acid moieties in their backbones. For example, chiral phosphines containing a Lewis base, a Bronsted acid, and an acid-activated Bronsted base were developed for an asymmetric aza-MBH reaction (86-96% yield, 79-92% ee). The Bronsted acid and base moieties were proposed to be involved in the stabilization of zwitterionic species in a stereoselective manner.[28]

 

BINOL-derived chiral phosphine catalyst is also effective for an asymmetric aza-MBH reaction of N-tosyl imines with activated alkenes such as methyl vinyl ketone and phenyl acrylate.[29]

In addition, a distinct class of chiral phosphine-squaramide molecules could effectively catalyze an intramolecular asymmetric MBH reaction. ω-formylenones reacted to afford enantioenriched cyclic products at ambient temperature (64-98% yield, 88-93% ee).[30]

 

Chiral Lewis acid catalyst edit

Chiral Lewis acid catalysts have been given interests as they could activate the electron-withdrawing group in an enantioselective manner. Chiral cationic oxazaborolidinium catalysts were shown to be effective in the three-component coupling of α,β-acetylenic esters, aldehydes, and trimethylsilyl iodide (50-99% yield, 62-94% ee). Both enantiomeric products could be obtained by using different enantiomers of the catalyst.[31]

 

Complex of metal salt and chiral ligand is a viable strategy, too. La(OTf)3 and camphor-derived chiral ligands could induce enantioselectivity in a DABCO-catalyzed MBH reaction of various aldehydes and acrylates (25-97% yield, 6-95% ee). For these cases, multidentate ligands were usually employed to chelate with the metal, which activates both the zwitterionic enolate and the aldehyde.[32]

 

La(O-iPr)3 and BINOL-derived ligand system, in conjunction with catalytic DABCO, also works for an asymmetric aza-MBH reaction of various N-diphenylphosphinoyl imines and methyl acrylate. Aryl, heteroaryl, and alkenyl imines were all suitable for good yield and enantioselectivity.[33]

Chiral palladium(II) pincer complexes function as Lewis acid in the enantioselective DABCO-catalyzed aza-MBH reaction of acrylonitrile and various tosyl imines to afford functionalized α-methylene-β-aminonitriles (75-98% yield, 76-98% ee). Silver acetate is required to activate the palladium bromide precatalyst in the catalytic cycle.[34]

 

Chiral Bronsted acid cocatalyst edit

A variety of chiral thiourea catalysts are under investigation for asymmetric MBH reactions. Chiral thiourea and bis(thiourea) catalysts can be effective in DABCO-catalyzed MBH and aza-MBH reactions.[35][36] Jacobsen's thiourea catalyst performs an enantioselective aza-MBH reaction, for example (25-49% yield, 87-99% ee).

 

While simple thiourea requires a nucleophilic catalyst in conjunction, bifunctional catalysts such as phosphine-thioureas can be used alone for asymmetric MBH reactions. For example, various acrylates and aromatic aldehydes react in the presence of these catalysts to afford either enantiomeric MBH adducts (32-96% yield, 9-77% ee).[37]

 

MBH reaction can involve proline derivative as a cocatalyst. It was proposed that imidazole nucleophilic catalyst and proline effect the reaction via iminium intermediate.[38] With (S)-proline and DABCO, α-amido sulfones and α,β-unsaturated aldehydes undergo a highly enantioselective aza-MBH reaction (46-87% yield, E/Z 10:1-19:1, 82-99% ee).[39]

 

Applications in organic synthesis edit

The Baylis–Hillman adducts and their derivatives have been extensively utilized for the generation of heterocycles and other cyclic frameworks.[40]

MBH reactions are widely used in organic synthesis. For example, this reaction was used to construct key cyclic intermediates for syntheses of salinosporamide A, diversonol, and anatoxin-a.[41][42][43]

 
 
 

Further reading edit

Many reviews have been written, including:

  • Deevi Basavaiah, Anumolu Jaganmohan Rao, and Tummanapalli Satyanarayana (2003), "Recent Advances in the Baylis−Hillman Reaction and Applications." Chem. Rev., 103 (3), pp. 811–892. doi:10.1021/cr010043d
  • G. Masson, C. Housseman and J. Zhu (2007), "The Enantioselective Morita–Baylis–Hillman Reaction and Its Aza Counterpart." Angewandte Chemie International Edition, 46: 4614–4628. doi:10.1002/anie.200604366
  • Valerie Declerck, Jean Martinez and Frederic Lamaty (2009), "The aza-Baylis−Hillman Reaction" Chem. Rev., 109 (1), pp. 1–48. doi:10.1021/cr068057c
  • Deevi Basavaiah, Bhavanam Sekhara Reddy and Satpal Singh Badsara (2010), "Recent Contributions from the Baylis−Hillman Reaction to Organic Chemistry" Chemical Reviews 110 (9), pp. 5447-5674. doi:10.1021/cr900291g
  • Deevi Basavaiah and Gorre Veeraraghavaiah (2012), "The Baylis–Hillman reaction: a novel concept for creativity in chemistry" Chem. Soc. Rev. doi:10.1039/C1CS15174F

References edit

  1. ^ Baylis, A. B.; Hillman, M. E. D. German Patent 2155113, 1972.
  2. ^ Ciganek, E. Org. React. 1997, 51, 201. doi:10.1002/0471264180.or051.02
  3. ^ K. Morita, Z. Suzuki and H. Hirose, Bull. Chem. Soc. Jpn.,1968, 41, 2815.
  4. ^ a b J. Phys. Org. Chem. 1990, 3, 285.
  5. ^ Angew. Chem. Int. Ed. Engl. 1983, 22, 795.
  6. ^ Organic Letters, 2005, 7, 1, 147-150.
  7. ^ a b Angew. Chem. Int. Ed. 2005, 44, 1706-1708.
  8. ^ J. Am. Chem. Soc. 2007, 129, 15513.
  9. ^ J. Org. Chem., 2009, 74(8), 3031-3037
  10. ^ Tetrahedron Lett. 2001, 42, 85.
  11. ^ Org. Lett. 2010, 12, 2418.
  12. ^ Chem. Commun. 2006, 2977.
  13. ^ J. Am. Chem. Soc. 2009, 131, 4196.
  14. ^ a b Basavaiah, Rao & Satyanarayana 2003.
  15. ^ Fort, Yves; Berthe, Marie Christine; Caubere, Paul (1992). "The 'Baylis - Hillman Reaction' mechanism and applications revisited". Tetrahedron. 48 (31): 6371–6384. doi:10.1016/s0040-4020(01)88227-2.
  16. ^ "Enantioselective Aza-Morita–Baylis–Hillman Reactions of Acrylonitrile Catalyzed by Palladium(II) Pincer Complexes having C2-Symmetric Chiral Bis(imidazoline) Ligands" Hyodo, K.; Nakamura, S.; Shibata, N. Angew. Chem. Int. Ed. 2012, 51, 10337. doi:10.1002/anie.201204891
  17. ^ Trofimov, Alexander; Gevorgyan, Vladimir (2009). "Sila-Morita−Baylis−Hillman Reaction of Arylvinyl Ketones: Overcoming the Dimerization Problem". Organic Letters. 11 (1): 253–255. doi:10.1021/ol8026522.
  18. ^ Marqués-López, Eugenia; Herrera, Raquel P.; Marks, Timo; Jacobs, Wiebke C.; Könning, Daniel; de Figueiredo, Renata M.; Christmann, Mathias (2009). "Crossed Intramolecular Rauhut−Currier-Type Reactions via Dienamine Activation". Organic Letters. 11 (18): 4116–4119. doi:10.1021/ol901614t. hdl:10261/113980.
  19. ^ Shi, Min; Zhao, Gui-Ling (2002). "One-pot aza-Baylis–Hillman reactions of arylaldehydes and diphenylphosphinamide with methyl vinyl ketone in the presence of TiCl4, PPh3, and Et3N". Tetrahedron Letters. 43 (50): 9171–9174. doi:10.1016/S0040-4039(02)02263-3.
  20. ^ Alemán, José; Núñez, Alberto; Marzo, Leyre; Marcos, Vanesa; Alvarado, Cuauhtémoc; Ruano, José Luis García (2010). "Asymmetric Synthesis of 4-Amino-4H-Chromenes by Organocatalytic Oxa-Michael/Aza-Baylis–Hillman Tandem Reactions". Chem. Eur. J. 16 (31): 9453–9456. doi:10.1002/chem.201001293.
  21. ^ J. Am. Chem. Soc. 1997, 119, 4317-4318
  22. ^ Org. Lett. 2000, 2, 6, 729-731
  23. ^ Eur. J. Org. Chem. 2010, 3249-3256
  24. ^ J. Am. Chem. Soc. 1999, 121, 10219-10220
  25. ^ Chem. Commun. 2010, 46, 2644-2646
  26. ^ Xiao, Y.; Sun, Z.; Guo, H.; Kwon, O. (2014). "Chiral Phosphines in Nucleophilic Organocatalysis". Beilstein Journal of Organic Chemistry. 10: 2089–2121. doi:10.3762/bjoc.10.218. PMC 4168899. PMID 25246969.
  27. ^ J. Tetrahedron: Asymmetry, 2010, 1511.
  28. ^ Adv. Synth. Catal. 2009, 351, 331
  29. ^ Chem. Commun. 2003, 1310
  30. ^ Chem. Commun. 2011, 47, 1012
  31. ^ Angew. Chem. Int. Ed. 2009, 48, 4398
  32. ^ J. Org. Chem. 2003, 68, 915-919
  33. ^ J. Am. Chem. Soc. 2010, 132, 11988
  34. ^ Angew. Chem. Int. Ed. 2012, 51, 10337-10341
  35. ^ Adv. Synth. Catal. 2005, 347, 1701-1708
  36. ^ Tetrahedron Lett. 2011, 52, 6234
  37. ^ Tetrahedron 2009, 65, 8185
  38. ^ Chem. Eur, J. 2009, 15, 1734
  39. ^ J. Adv. Synth. Catal. 2011, 353, 1096
  40. ^ Tetrahedron, 2008, 64(20), 4511-4574.
  41. ^ J. Am. Chem. Soc. 2004, 126, 6230-6231.
  42. ^ Angew. Chem. Int. Ed. 2006, 45, 307–309.
  43. ^ Chem. Commun. 2008, 3432.

baylis, hillman, reaction, this, article, relies, excessively, references, primary, sources, please, improve, this, article, adding, secondary, tertiary, sources, find, sources, news, newspapers, books, scholar, jstor, july, 2023, learn, when, remove, this, te. This article relies excessively on references to primary sources Please improve this article by adding secondary or tertiary sources Find sources Baylis Hillman reaction news newspapers books scholar JSTOR July 2023 Learn how and when to remove this template message In organic chemistry the Baylis Hillman Morita Baylis Hillman or MBH reaction is a carbon carbon bond forming reaction between an activated alkene and a carbon electrophile in the presence of a nucleophilic catalyst such as a tertiary amine or phosphine The product is densely functionalized joining the alkene at the a position to a reduced form of the electrophile e g in the case of an aldehyde an allylic alcohol 1 2 Morita Baylis Hillman reactionNamed after Ken ichi Morita Anthony B Baylis Melville E D HillmanReaction type Coupling reactionIdentifiersOrganic Chemistry Portal baylis hillman reactionRSC ontology ID RXNO 0000076 Baylis Hillman reactionThe reaction is named for Anthony B Baylis and Melville E D Hillman two of the chemists who developed the reaction at Celanese and K Morita who published earlier work 3 on the same The MBH reaction offers several advantages in organic synthesis It combines easily prepared starting materials with high atom economy It requires only mild conditions and does not require any transition metals Asymmetric synthesis is possible from prochiral electrophiles The product s dense functionalization enables many further transformations Its disadvantage is that the reaction is extremely slow Contents 1 Common reagents 2 Reaction mechanism 3 Scope and limitations 4 Variants 4 1 Sila MBH reaction 4 2 Rauhut Currier reaction 4 3 Tandem strategies 5 Asymmetric synthesis 5 1 Chiral auxiliaries 5 2 Chiral Lewis basic catalyst 5 3 Chiral Lewis acid catalyst 5 4 Chiral Bronsted acid cocatalyst 6 Applications in organic synthesis 7 Further reading 8 ReferencesCommon reagents editThe most frequently used catalyst for the reaction is the tertiary amine DABCO triethylenediamine other known catalysts include 4 dimethylaminopyridine DBU diaza bicyclo undecene and various phosphines Reaction mechanism editAs of 2012 update certain questions about MBH reaction s mechanism remain open Hill and Isaacs performed the first kinetic experiments in the 1990s discovering that the reaction rate between acrylonitrile and acetaldehyde was first order in each reactant and in the DABCO catalyst a Deuterated acrylonitrile exhibited no kinetic isotope effect but the product appeared to catalyze its own generation 4 In a model Hoffmann proposed seven years prior 5 the reaction begins with 1 4 addition of the catalytic amine to the activated alkene The resulting zwitterionic aza enolate undergoes aldol addition to the aldehyde Intramolecular proton shift then generates the final MBH adduct which eliminates the catalyst nbsp If Hoffmann s model were correct then the aldol addition would be the rate limiting step which accords with the absent kinetic isotope effect 4 However Hoffman s mechanism rationalizes neither the product s autocatalysis nor in the reaction of aryl aldehydes with acrylates the considerable generation of a dioxanone byproduct In more recent work McQuade et al and Aggarwal et al reevaluated the MBH mechanism focusing on the proton transfer and autocatalysis 6 7 According to McQuade the reaction between methyl acrylate and p nitrobenzaldehyde is second order relative to the aldehyde Moreover it showed a significant kinetic isotope effect for the acrylate s a hydrogen 5 2 in DMSO but 2 in all solvents which would imply that proton abstraction is the rate determining step To account for this reanalysis McQuade proposed modifying Hoffman s proposal such that after the first aldol addition a second aldol adds to form a hemiacetal alkoxide Then the rate determining proton transfer releases the adduct A via a six membered transition state which further reacts to produce the MBH product B or a dioxanone byproduct C In a further modification Aggarwal noted that observed that methanol inhibited autocatalysis in the MBH reaction Thus he proposed that in the reaction s early stages a mechanism equivalent to McQuade s proposal operates but after 20 conversion reaction with an alcoholic solvent can replace the second aldol addition In support of this contention Aggarwal and Harvey modeled the two pathways using density functional theory calculations and showed that the computed energy profile matches the experimental kinetic isotope effects and observed rate of reaction 8 Also they showed that the overall enthalpic barrier of the alcohol catalyzed pathway is slightly smaller than that of the non alcohol catalyzed pathway rationalizing that as the alcohol MBH product concentration increases the alcohol catalyzed pathway starts to dominate nbsp While McQuade s and Aggarwal s studies received much attention unequivocal proof of intermediate A s role remained elusive as of 2012 update Because A could be formed by separate addition of B to an aldehyde any isolated A and C could be the result of side reactions rather than the MBH itself Nor has a convincing explanation yet been presented for Hill and Isaac s original first order data Aggarwal s modification has its own controversies If it is correct then the rate determining step changes from proton transfer to aldol addition over the course of the reaction 7 but subsequent computational studies have concluded that proton transfer still has the highest barrier even late into the reaction s process On the other hand Coelho and Eberlin et al have obtained electrospray mass spectroscopy data that is structural evidence for two different forms of the reaction s proton transfer step 9 Scope and limitations edit nbsp The MBH reaction is extremely general In most cases the electrophile is an aldehyde ketone but see below or imine latterly the aza Baylis Hillman reaction but reports indicate that allyl halides alkyl halides and epoxides are also possible 10 11 12 Using an allene instead of a simple alkene as the precursor gives an intermediate that can react at the g carbon rather than at the a 13 At the same time it can be challenging to develop suitable reaction conditions The reaction is slow times of a fortnight or longer are not uncommon even with 25 100 mol catalyst especially with as alkene b substituted activated olefins vinyl sulfones or vinyl sulfoxides or as electrophile hindered aliphatic aldehydes or electron rich benzaldehydes Ketones are generally not reactive enough under ordinary conditions to take part in a synthetically useful manner 14 For example reaction between sterically hindered t butyl acrylate and benzaldehyde with catalytic DABCO in the absence of solvent required 4 weeks to give moderate conversion to the final product nbsp In aprotic solvents the reaction rate is even slower although recovery is possible with protic additives e g alcohols and carboxylic acids 15 At such low rates the activity of the substrates may induce competing side reactions acroleins also oligomerize and allenoates cycloadd Allyl halide and alkyl epoxide electrophiles also often prove unruly clarification needed citation needed The MBH reaction of an aryl vinyl ketone with an aldehyde is not straightforward but see Sila MBH reaction since the reactive aryl vinyl ketone readily undergoes Michael addition to another molecule of the aryl vinyl ketone which then adds to the aldehyde to form a double MBH adduct 16 nbsp Due to the highly negative volume of activation sluggish Baylis Hillman reactions including ketonic ones can be realized by conducting the reaction under high pressure up to 20 kbar 14 Variants editSila MBH reaction edit nbsp In the sila MBH reaction a silylated vinyl aryl ketones couple to aldehydes in the presence of catalytic TTMPP a large triarylphosphine reagent 17 The zwitterionic enolate produced upon addition of nucleophilic catalyst to the enone adds to an aldehyde carbonyl to generate an alkoxide This alkoxide undergoes a subsequent 1 3 Brook rearrangement and elimination cascade to afford a siloxy methylene enone and release the catalyst Rauhut Currier reaction edit nbsp The Rauhut Currier reaction is a vinylogous analogoue of the MBH reaction in which the electrophile is a Michael acceptor not an aldehyde or an imine Intermolecular Rauhut Currier reactions typically exhibit poor chemoselectivity because the reaction couples two activated alkenes but intramolecular Rauhut Currier reactions have been employed For example cyclization of a b unsaturated aldehydes can be performed in the presence of proline derivative and acetic acid affording enantioenriched products 18 Tandem strategies edit As mentioned above the slow rate of the MBH reaction often enables side reactions on its activated substrates In tandem reaction strategies this is a virtue for it enables syntheses with high atom economy For example in the three component coupling of aldehydes amines and activated alkenes the aldehyde reacts with the amine to produce an imine prior to forming the aza MBH adduct as in the reaction of aryl aldehydes diphenylphosphinamide and methyl vinyl ketone in the presence of TiCl4 triphenylphosphine and triethylamine 19 nbsp Likewise activated acetylenes can undergo conjugate addition and remain an activated alkene for the MBH reaction as in the following enantioselective cyclization reaction in which a phenolate nucleophile adds to a functionalized enyne before aza MBH ring closure catalyzed by a chiral amine base 20 nbsp Asymmetric synthesis editChiral auxiliaries edit Oppolzer s sultam can be used as a chiral auxiliary for an asymmetric MBH reaction When an acrylate substituted with the Oppolzer s sultam reacted with various aldehydes in the presence of DABCO catalyst optically pure 1 3 dioxan 4 ones were afforded with cleavage of the auxiliary 67 98 yield gt 99 ee The cyclic products could be converted into desired MBH products by use of camphorsulfonic acid and methanol 21 nbsp A related hydrazide auxiliary is the chiral acryloylhydrazide which reacts diastereoselectively with aldehydes 22 Both diastereomers could be obtained with different choice of solvents DMSO vs mixed THF and H2O suggesting that the transition structure conformation is solvent influenced Chiral allenes and imines can also be employed for an asymmetric DABCO catalyzed aza MBH reaction 23 Optically active 10 phenylsulfonylisobornyl buta 2 3 dienoate reacts with an aryl imine to afford a allenylamine in a diastereoselective manner 37 57 yield nbsp Chiral Lewis basic catalyst edit Some enantioselective MBH reactions employ chiral tertiary amine catalysts For example b ICD a cinchona alkaloid derivative is famous among the quinidine framework based catalysts and catalyzed an enantioselective MBH reaction with 1 1 1 3 3 3 hexafluoroisopropyl acrylate as the activated alkene 24 nbsp The phenolic oxygen of b ICD was shown to be important in the reaction implying that b ICD acts as a Bronsted acid not just a nucleophile Cyclopentenone and various aromatic and aliphatic aldehydes undergo an asymmetric reaction using Fu s planar chiral DMAP catalyst in isopropanol 54 96 yield 53 98 ee In this case magnesium iodide as a Lewis acid cocatalyst was required to accelerate the reaction 25 nbsp P Chiral phosphines have been investigated 26 Simple diamines can also be employed as MBH catalysts Methyl vinyl ketone and various substituted benzaldehydes were found to undergo asymmetric MBH reaction The chiral pyrrolidine catalyst was effective for ortho and para substituted electron deficient benzaldehydes 75 99 yield 8 73 ee 27 nbsp Chiral phosphine MBH catalysts often contain Bronsted acid moieties in their backbones For example chiral phosphines containing a Lewis base a Bronsted acid and an acid activated Bronsted base were developed for an asymmetric aza MBH reaction 86 96 yield 79 92 ee The Bronsted acid and base moieties were proposed to be involved in the stabilization of zwitterionic species in a stereoselective manner 28 nbsp BINOL derived chiral phosphine catalyst is also effective for an asymmetric aza MBH reaction of N tosyl imines with activated alkenes such as methyl vinyl ketone and phenyl acrylate 29 In addition a distinct class of chiral phosphine squaramide molecules could effectively catalyze an intramolecular asymmetric MBH reaction w formylenones reacted to afford enantioenriched cyclic products at ambient temperature 64 98 yield 88 93 ee 30 nbsp Chiral Lewis acid catalyst edit Chiral Lewis acid catalysts have been given interests as they could activate the electron withdrawing group in an enantioselective manner Chiral cationic oxazaborolidinium catalysts were shown to be effective in the three component coupling of a b acetylenic esters aldehydes and trimethylsilyl iodide 50 99 yield 62 94 ee Both enantiomeric products could be obtained by using different enantiomers of the catalyst 31 nbsp Complex of metal salt and chiral ligand is a viable strategy too La OTf 3 and camphor derived chiral ligands could induce enantioselectivity in a DABCO catalyzed MBH reaction of various aldehydes and acrylates 25 97 yield 6 95 ee For these cases multidentate ligands were usually employed to chelate with the metal which activates both the zwitterionic enolate and the aldehyde 32 nbsp La O iPr 3 and BINOL derived ligand system in conjunction with catalytic DABCO also works for an asymmetric aza MBH reaction of various N diphenylphosphinoyl imines and methyl acrylate Aryl heteroaryl and alkenyl imines were all suitable for good yield and enantioselectivity 33 Chiral palladium II pincer complexes function as Lewis acid in the enantioselective DABCO catalyzed aza MBH reaction of acrylonitrile and various tosyl imines to afford functionalized a methylene b aminonitriles 75 98 yield 76 98 ee Silver acetate is required to activate the palladium bromide precatalyst in the catalytic cycle 34 nbsp Chiral Bronsted acid cocatalyst edit A variety of chiral thiourea catalysts are under investigation for asymmetric MBH reactions Chiral thiourea and bis thiourea catalysts can be effective in DABCO catalyzed MBH and aza MBH reactions 35 36 Jacobsen s thiourea catalyst performs an enantioselective aza MBH reaction for example 25 49 yield 87 99 ee nbsp While simple thiourea requires a nucleophilic catalyst in conjunction bifunctional catalysts such as phosphine thioureas can be used alone for asymmetric MBH reactions For example various acrylates and aromatic aldehydes react in the presence of these catalysts to afford either enantiomeric MBH adducts 32 96 yield 9 77 ee 37 nbsp MBH reaction can involve proline derivative as a cocatalyst It was proposed that imidazole nucleophilic catalyst and proline effect the reaction via iminium intermediate 38 With S proline and DABCO a amido sulfones and a b unsaturated aldehydes undergo a highly enantioselective aza MBH reaction 46 87 yield E Z 10 1 19 1 82 99 ee 39 nbsp Applications in organic synthesis editThe Baylis Hillman adducts and their derivatives have been extensively utilized for the generation of heterocycles and other cyclic frameworks 40 MBH reactions are widely used in organic synthesis For example this reaction was used to construct key cyclic intermediates for syntheses of salinosporamide A diversonol and anatoxin a 41 42 43 nbsp nbsp nbsp Further reading editMany reviews have been written including Deevi Basavaiah Anumolu Jaganmohan Rao and Tummanapalli Satyanarayana 2003 Recent Advances in the Baylis Hillman Reaction and Applications Chem Rev 103 3 pp 811 892 doi 10 1021 cr010043d G Masson C Housseman and J Zhu 2007 The Enantioselective Morita Baylis Hillman Reaction and Its Aza Counterpart Angewandte Chemie International Edition 46 4614 4628 doi 10 1002 anie 200604366 Valerie Declerck Jean Martinez and Frederic Lamaty 2009 The aza Baylis Hillman Reaction Chem Rev 109 1 pp 1 48 doi 10 1021 cr068057c Deevi Basavaiah Bhavanam Sekhara Reddy and Satpal Singh Badsara 2010 Recent Contributions from the Baylis Hillman Reaction to Organic Chemistry Chemical Reviews 110 9 pp 5447 5674 doi 10 1021 cr900291g Deevi Basavaiah and Gorre Veeraraghavaiah 2012 The Baylis Hillman reaction a novel concept for creativity in chemistry Chem Soc Rev doi 10 1039 C1CS15174FReferences edit Baylis A B Hillman M E D German Patent 2155113 1972 Ciganek E Org React 1997 51 201 doi 10 1002 0471264180 or051 02 K Morita Z Suzuki and H Hirose Bull Chem Soc Jpn 1968 41 2815 a b J Phys Org Chem 1990 3 285 Angew Chem Int Ed Engl 1983 22 795 Organic Letters 2005 7 1 147 150 a b Angew Chem Int Ed 2005 44 1706 1708 J Am Chem Soc 2007 129 15513 J Org Chem 2009 74 8 3031 3037 Tetrahedron Lett 2001 42 85 Org Lett 2010 12 2418 Chem Commun 2006 2977 J Am Chem Soc 2009 131 4196 a b Basavaiah Rao amp Satyanarayana 2003 Fort Yves Berthe Marie Christine Caubere Paul 1992 The Baylis Hillman Reaction mechanism and applications revisited Tetrahedron 48 31 6371 6384 doi 10 1016 s0040 4020 01 88227 2 Enantioselective Aza Morita Baylis Hillman Reactions of Acrylonitrile Catalyzed by Palladium II Pincer Complexes having C2 Symmetric Chiral Bis imidazoline Ligands Hyodo K Nakamura S Shibata N Angew Chem Int Ed 2012 51 10337 doi 10 1002 anie 201204891 Trofimov Alexander Gevorgyan Vladimir 2009 Sila Morita Baylis Hillman Reaction of Arylvinyl Ketones Overcoming the Dimerization Problem Organic Letters 11 1 253 255 doi 10 1021 ol8026522 Marques Lopez Eugenia Herrera Raquel P Marks Timo Jacobs Wiebke C Konning Daniel de Figueiredo Renata M Christmann Mathias 2009 Crossed Intramolecular Rauhut Currier Type Reactions via Dienamine Activation Organic Letters 11 18 4116 4119 doi 10 1021 ol901614t hdl 10261 113980 Shi Min Zhao Gui Ling 2002 One pot aza Baylis Hillman reactions of arylaldehydes and diphenylphosphinamide with methyl vinyl ketone in the presence of TiCl4 PPh3 and Et3N Tetrahedron Letters 43 50 9171 9174 doi 10 1016 S0040 4039 02 02263 3 Aleman Jose Nunez Alberto Marzo Leyre Marcos Vanesa Alvarado Cuauhtemoc Ruano Jose Luis Garcia 2010 Asymmetric Synthesis of 4 Amino 4H Chromenes by Organocatalytic Oxa Michael Aza Baylis Hillman Tandem Reactions Chem Eur J 16 31 9453 9456 doi 10 1002 chem 201001293 J Am Chem Soc 1997 119 4317 4318 Org Lett 2000 2 6 729 731 Eur J Org Chem 2010 3249 3256 J Am Chem Soc 1999 121 10219 10220 Chem Commun 2010 46 2644 2646 Xiao Y Sun Z Guo H Kwon O 2014 Chiral Phosphines in Nucleophilic Organocatalysis Beilstein Journal of Organic Chemistry 10 2089 2121 doi 10 3762 bjoc 10 218 PMC 4168899 PMID 25246969 J Tetrahedron Asymmetry 2010 1511 Adv Synth Catal 2009 351 331 Chem Commun 2003 1310 Chem Commun 2011 47 1012 Angew Chem Int Ed 2009 48 4398 J Org Chem 2003 68 915 919 J Am Chem Soc 2010 132 11988 Angew Chem Int Ed 2012 51 10337 10341 Adv Synth Catal 2005 347 1701 1708 Tetrahedron Lett 2011 52 6234 Tetrahedron 2009 65 8185 Chem Eur J 2009 15 1734 J Adv Synth Catal 2011 353 1096 Tetrahedron 2008 64 20 4511 4574 J Am Chem Soc 2004 126 6230 6231 Angew Chem Int Ed 2006 45 307 309 Chem Commun 2008 3432 Retrieved from https en wikipedia org w index php title Baylis Hillman reaction amp oldid 1207306389, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.