Acetaldehyde in asymmetric organocatalytic ...

3 downloads 0 Views 1MB Size Report
Paraldehyde,57 vinyl acetate,58 and vinyl ethers59 are being mostly used as the indirect sources of acetaldehyde in asymmetric organic transformations. 4.1.
Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances View Article Online

REVIEW

View Journal | View Issue

Acetaldehyde in asymmetric organocatalytic transformations Cite this: RSC Adv., 2015, 5, 55926

Manjeet Kumar,†a Arvind Kumar,†a Masood Ahmad Rizvi*b and Bhahwal Ali Shah*a The role of acetaldehyde as a nucleophile in various asymmetric C–C bond forming transformations is presented, with consideration given not only to the optimization of reaction parameters relevant to product formation, polymerization, by-products, purification, chirality and instability of the final products, but also to the application of the final products in the synthesis of bioactive molecules. This review is organized according to the use of acetaldehyde in different organocatalytic reactions covering the most Received 31st March 2015 Accepted 17th June 2015

recent reports. In the last section indirect sources of acetaldehyde are discussed from the perspective of difficult handling and its requirement of slow addition as well as being freshly distilled in some of the

DOI: 10.1039/c5ra05695k

transformations, which is one of the most critical issues in the late exploitation of acetaldehyde as a

www.rsc.org/advances

nucleophile in synthetic chemistry.

1. Introduction Asymmetric C–C bond formation is one of the most important and challenging key transformations in organic chemistry,1,2 providing access to a wide variety of building blocks used in chiral drug development.3 The importance of enantioselective C–C bond formation can also be inferred from the large amount of research that has been performed on the reactions like asymmetric aldol,4 Diels–Alder,5 Michael,6 allylic alkylation,7 Mannich,8 and Baylis–Hillman reactions.9 Previously, these transformations were generally achieved by either of two catalytic approaches i.e., metal catalysts employing mainly transition metals or biocatalysts wherein enzymes were used for their efficiency and selectivity.10 However in the last decade a third front, the eld of organocatalysis has grown with a breathtaking pace and has emerged as powerful tool to construct enantiorich compounds.11 This can be easily visualised from the large number of research papers published in last decade.12 Although Hajos et al.13 and Wiechert et al.14 had independently reported the rst highly enantioselective proline catalysed intramolecular aldol reaction in the early 1970's; organocatalysis was not considered as a possible alternative to two main classes of established asymmetric catalysis (metal catalysis and biocatalysis). It was in 2000 when Barbas and co-workers15 employed proline in enantioselective intermolecular cross aldol reaction (Scheme 1), marking the explosive growth of low molecular-weight enamine/iminium-ion

a

Academy of Scientic and Innovative Research, Natural Product Microbes, CSIR-Indian Institute of Integrative Medicine, Canal Road, Jammu-Tawi, 180001, India. E-mail: [email protected]; [email protected]

b

Department of Chemistry, University of Kashmir, Srinagar, India

† These authors contributed equally.

55926 | RSC Adv., 2015, 5, 55926–55937

Scheme 1

L-Proline

catalysed cross-aldol reaction.

organocatalysis. The reaction involves asymmetric enamine catalysis in which carbonyl compounds form reactive enamine intermediate with L-proline catalyst, increasing the feasibility of reacting with electrophiles.

2. Challenges associated with the use of acetaldehyde The recent advances in the design and development of organocatalysts has not only provided the easy and more sustainable alternate to metal complexes in organic synthesis but also offers a greater scope in terms of coupling partners (nucleophiles & electrophiles). Various nucleophiles have made their way into the world of chemical diversity through these synthetic methodologies. Despite all advances the simplest of all nucleophiles, acetaldehyde, was not much used until recently16 due to various

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

Scheme 2

RSC Advances

Possible side product formation in aldol and Mannich

reaction.

reasons such as (a) it's low boiling point (21  C), generally involves sensitive reaction protocols; (b) its ability to act as nucleophile as well as electrophile, give rise to uncontrolled cascade reactions like polyaldolization, double Mannich etc.; (c) its tendency to polymerize at room temperature requires freshly distilled acetaldehyde for product formation in good yields; (d) dehydration of the nal product enables Michael type additions; (e) its high reactivity results in self-aldol condensation prior to cross coupling; (f) undergo Tishchenko-type processes; (g) the small steric difference between the methyl and hydrogen atom suffers with low relative control, resulting in decreased enantioselectivity (Scheme 2).

3.

Acetaldehyde as a nucleophile

3.1. Use of acetaldehyde in aldol reactions Aldol reaction is one of the most studied reactions for C–C bond formation in the history of synthetic chemistry.17 This may be attributed to the versatility offered by this reaction in terms of coupling partners with the ease of use.18 Owing to the difficult handling, acetaldehyde was not directly used as a nucleophile in organocatalysed C–C bond forming reactions until 2002, when Barbas et al.19 reported the rst time direct use of acetaldehyde (1) in self aldol reaction. In the initial investigation, the proline (2) catalysed self aldol reaction in DMSO at room temperature lead to the formation of two products, i.e. (+)-(5S)-hydroxy-(2E)hexenal (3) and 2,4-hexadienal (4), where dehydrated product (3) was the major one. Standardisation of the reaction condition i.e. in THF at 0  C, afforded the desired product in high enantiomeric excess (90%) however the yield was low (10%). Proposed mechanistic origin of stereoselectivity involves the refacial attack of enamine on carbonyl group of acetaldehyde in transition state (TS-1) which further undergoes Mannich type condensation to give the product. Authors also addressed the importance of products by converting them into various important building blocks commonly encountered in the synthesis of macrolide antibiotics such as grahamimycin A and A1, carbomycin B and platenomycin A20(Scheme 3). Aer these initial efforts, other reports followed where Denmark and Bui21 used trialkylsilylenol ether of acetaldehyde in chiral phosphoramide catalysed enantioselective aldol addition with aromatic aldehydes. The reaction of trimethylsilyl enol ether of acetaldehyde (11) and benzaldehyde (12) in the presence of 15 mol% of chiral dimeric phosphoramide catalyst (13), SiCl4 (2 equiv.) and i-Pr2EtN (1 equiv.) gave the desired aldol

This journal is © The Royal Society of Chemistry 2015

Scheme 3

Acetaldehyde in L-proline catalysed self-aldol reaction.

addition in 80% yield and 97.1 : 2.9 enantiomeric ratio. Since the product was unstable thereby reaction was quenched with methanol which resulted in dimethylacetal protected aldol product (14). Furthermore, reaction proceeds through the formation of chlorohydrin intermediate (15) which in the presence of tert-butyl isocyanide had been converted to lactone (16, d.r. 9/1) and b-hydroxy amide (17) in a single step (Scheme 4). Although proline has been used in the self aldol reaction of acetaldehyde19 but reaction afforded the product in poor yields (12%). To overcome the issues related to yields, Hayashi and co-workers22 successfully attempted the use of triuoromethyl substituted diaryl prolinol (18) to give product (19) in good yields and high enantioselectivities. The reaction was screened in different solvents where the best results were obtained in N-methyl pyrrolidone affording the product in 56% yield and 82% enantioselectivity. Furthermore, the product obtained was in situ acetal protected, which had been converted to the different important building blocks (21, 22) in same pot (Scheme 5). In the proposed transition state (TS-2), diarylprolinol forms anti-enamine complex with one equivalent of acetaldehyde and directing the approach towards (S)-phase of other electrophilic acetaldehyde through hydrogen bonding between acidic proton of catalyst and carbonyl oxygen of acetaldehyde.

Scheme 4 Silylenol ether of acetaldehyde in enantioselective aldol

addition.

RSC Adv., 2015, 5, 55926–55937 | 55927

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances

Review

Scheme 5 Asymmetric, catalytic and direct self-aldol reaction of acetaldehyde catalyzed by diaryl prolinol.

3.2. Mannich reaction Chiral b-amino carbonyl compounds are always considered as privileged motif in medicinal chemistry, owing to their high abundance in natural products as well as in drug like molecules. They are largely accessed through Mannich reaction,23 which is arguably the most simple, economical approach to C–C bond formation in nitrogen containing compounds and has ourished over the years due to the ubiquitous potential of this multicomponent reaction to generate diversity by means of substrate combinations. However acetaldehyde was not used as a substrate in Mannich reaction until 2008, when List et al.,24 for the rst time reported the direct use of acetaldehyde in (S)-proline catalysed Mannich reaction. They came up with the observation that all the side reactions (Scheme 2) can be suppressed if higher excess of acetaldehyde (5–10 equivalents) is employed. Thus (S)-proline (20 mol%) catalysed three component reaction of aromatic as well as aliphatic N-Boc-imines (23) with acetaldehyde (10 equiv.) in acetonitrile at 0  C gave the

Scheme 7 Acetaldehyde in trifluoromethyl-substituted diaryl prolinolsilyl ether catalysed Mannich reaction.

b-amino aldehydes (24) in moderate yields (23–58%) and high enantioselectivity ratio upto >99 : 1. The importance of Mannich product was also highlighted by using it as a chiral building block in the synthesis of variety of bioactive molecules e.g., (24) can readily undergo Wittig reaction to give d-aminoa,b-unsaturated ester (25), a synthetic intermediate of substituted piperidines (26) which are highly abundant in bioactive molecules.25 Furthermore, the reductive amination of aldehyde also resulted in the formation of piperidine derivatives (27) which found its application in the synthesis of UK-427, 857 (28), a recently approved CCR5 inhibitor for the HIV treatment.26 To broaden the further application of asymmetric Mannich product, author reduced it to amino alcohol (30), which is a known intermediate in the synthesis (S)-dapoxetine (31), a selective serotonin re-uptake inhibitor.27 Also in situ cyclisation of reduced alcohol lead to the formation of oxazinone (32), an important heterocyclic synthetic intermediate.28 In addition, enantiopure b3-amino acids29 (29) can also be synthesized by oxidation of Mannich product in the presence of sodium hypochlorite (Scheme 6). In the same year Hayashi and co-workers reported the triuoromethyl-substituted diaryl prolinolsilyl ether catalysed

Scheme 8 Acetaldehyde in (S)-proline catalysed double Mannich Scheme 6

Acetaldehyde in (S)-proline catalysed Mannich reaction.

55928 | RSC Adv., 2015, 5, 55926–55937

reaction.

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

RSC Advances

Scheme 9 Chiral amino sulfonamide as an efficient organocatalyst for direct asymmetric Mannich reactions of N-Boc-protected imines.

Mannich reaction of acetaldehyde with N-benzoyl imine (33).30 It was however found that reaction works only in the presence of weak acidic additive like benzoic acids while strong acid like TsOH results in the decomposition of imine. The reaction in the presence of additive as p-nitrobenzoic acid in anhydrous THF affords the product (35) in good yields and excellent enantioselectivities (up to 98%). In this paper authors also presented the mechanistic insights of the reaction through quantum energy calculations where they found that trans N–C(O) bond having (S)-cis-geometry of imine and anti-conformation of enamine are the low energy transition states. According to the energy calculations, attack of trans-enamine at si-face of protonated imine is a highly exothermic process having large negative energy ( 38.82, 31.12 kcal mol 1 in gas phase and THF phase respectively, calculated through B3LYP theory, TS-3). This suggests that formation of enamine from acetaldehyde and diarylprolinolsilyl ether involving several steps (attack of N-atom of catalyst at the carbonyl centre of acetaldehyde, proton transfer, and dehydration) is the rate determining step. Aer which the formation C–C bond formation take place with almost no activation barrier. This was further conrmed by Gan and co-workers31 in the detailed mechanistic studies of reaction through density functional theory calculations (Scheme 7). List et al.,32 also reported the use of acetaldehyde in double Mannich reaction where one equivalent of acetaldehyde was treated with three equivalents of N-Boc-imines (37, 3 equiv.) in the presence of (S)-proline in acetonitrile at 0  C. The reaction was found to give double Mannich product (38) with remarkably high enantioselectivities with both (S) and (R)-proline; however, the products with aliphatic aldehyde were obtained in low yields (Scheme 8, eqn (III)). Methodology was further extended to

Scheme 11 Aldol reaction of acetaldehyde with isatin derivatives.

Acetaldehyde in trifluoromethyl substituted diarylprolinol catalyzed cross aldol reaction.

Scheme 12

cross-Mannich reaction with two different imines wherein, initial mono-addition products were rstly isolated and then subjected to a second reaction with different aromatic N-Boc imines (Scheme 8, eqn (IV)). In 2009, Maruoka and co-workers33 reported the use of acetaldehyde in direct asymmetric Mannich reaction of N-Boc-protected imines (41). The reaction was performed with axially chiral bifunctional amino sulphonamide catalyst (42), where the author proposed that highly acidic triamide group played a crucial role inactivating the electrophile with an added advantage i.e. dibenzylic secondary amine are less nucleophilic than pyrrolidine type catalysts thus reduces the possibility of the side product formation in Mannich reaction. Initially reaction was done in various solvents later on they found the best results in solvent free conditions at 0  C. Reaction was applicable to both aromatic as well as aliphatic imines without any signicant change in either of yields or enantioselectivity. Also, author attempted the reaction with different a-substituted aldehydes which afford anti-Mannich adducts with high enantioselectivities (Scheme 9). 3.3. Cross-aldol reaction

Scheme 10 Acetaldehyde in trifluoromethyl substituted diarylprolinol catalyzed cross Aldol reaction.

This journal is © The Royal Society of Chemistry 2015

The high abundance of asymmetric b-hydroxyl carbonyl moiety in natural products has always excited the research community for development of their new synthetic methods. Over the years cross-aldol reaction has ourished as the simplest and most economical way to introduce this functionality in

RSC Adv., 2015, 5, 55926–55937 | 55929

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances

Review

Scheme 13 4-Hydroxy diarylprolinol is used as a catalyst in the reaction between acetaldehyde and isatin. Scheme 15 Dioctylamino group containing diarylprolinol-based catalysed cross aldol reaction of acetaldehyde.

stereochemical fashion. In 2008, Hayashi et al.34 rst time exploited the nucleophilicity of acetaldehyde in organocatalysed cross-aldol reaction wherein they employed various aromatic as well as aliphatic aldehydes as eletrophiles in the presence of bistriuoromethyl substituted diarylprolinol (18). In an initial attempt they employed L-proline (2) as a catalyst in reaction between 2-chlorbenzaldehyde (44) and acetaldehyde, but instead of giving any cross-aldol product, reaction underwent self-condensation of acetaldehyde and gave crotonaldehyde (45) as a major product (Scheme 10, eqn (V)). Screening different proline derivatives author found that triuoromethylsubstituted diarylprolinol catalyst (18) is suitable. Thus, on employment of (18), the reaction of acetaldehyde with variety of aromatic aldehydes gave cross-aldol product (48) in good yields (52–92 %) and high enantioselectivities (80–99%) (Scheme 10, eqn VI). As the reaction in the presence of diarylprolinol based silyl ether affords the product in poor yields suggesting the role of hydrogen bonding in activating carbonyl group of aromatic aldehyde for reaction (TS-4). Since, the product obtained was

Scheme 14

Primary amine catalysed cross aldol reaction of

acetaldehyde.

55930 | RSC Adv., 2015, 5, 55926–55937

not stable and undergo dehydration when tried to isolate, so the cross-aldol product was further reduced to diol (49) in the presence of sodium borohydride in methanol at 0  C. Isatins have recently gained an increased interest due to its privileged scaffold, a large abundance in biologically potent molecules and greatly its synthetic versatility led to the generation of a large number of structurally diverse organic compounds. Aer the successful implication of acetaldehyde in cross-aldol reaction with various aromatic aldehydes, Nakamura et al.35 and Hayashi et al.36 independently reported the reaction of acetaldehyde with isatin in the presence of N-(heteroarenesulfonyl)prolinamides (52) and 4-hydroxydiarylprolinol (51) organocatalysts respectively. In both cases reactions afforded the product in good yields with high enantioselectivities. The importance of obtained product was also shown in the synthesis of 3-hydroxyindole alkaloids (Scheme 11). The reactions involve isatin aldol product as a key intermediate in the synthesis of ent-convolutamydine E (54) and CPC-1 (55) and a half fragment of madindoline A and B (56). Zhao and coworkers37 also reported a similar reaction of acetaldehyde and isatin where cinchona primary amine (58) and benzoic acid were used as catalysts and co-catalyst respectively. Standardization studies shows that water has a signicant effect on reactivity of the reaction i.e., 3 equiv. of water in THF solvent decreases the reaction time from 48 to 15 h with no change in yield however there is slight increase in enantioselectivity from 87–93%. Proposed mechanism indicates the multiple roles of benzoic acid like it accelerates the enamine

Scheme 16 Organocatalytic high enantioselective synthesis of b-formyl-a-hydroxyphosphonates.

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

RSC Advances

19 Acetaldehyde in N-heterocyclic carbene-catalysed intermolecular Setter reaction.

Scheme

Scheme 17 Acetaldehyde in trifluoromethyl substituted diarylprolinol silyl ether catalysed Michael reaction.

formation between primary amine moiety of catalyst and acetaldehyde, simultaneously, it also protonates the N atom of quinicilidine ring which forms hydrogen bond with isatin carbonyl groups. In the transition state enamine approaches the re-face of isatin ketone group resulting in major S-enantiomer (TS-5, TS-6) (Scheme 12). Other report by Yuan and coworkers,38 where 4-hydroxy diarylprolinol (51) is used as a catalyst in the reaction between acetaldehyde and isatin. Reaction without any co-catalyst affords the products in high yields and good enantioselectivities. To demonstrate the synthetic utility of reaction method, author successfully attempted the synthesis of various biologically active compounds like (R)-convolutamydines B (62) and E (63), ( )-donaxaridine (64) and (R)-chimonamidine (65) (Scheme 13). In the last decade number of chiral primary amine-based organocatalysts had been developed for various carbon–carbon bond forming reactions.39 Cheng et al.,40 had successfully applied primary amines as the organocatalysts in aldol reaction of acetaldehyde with aromatic aldehyde and isatin derivatives. The reaction of L-tert-leucine derivative organocatalyst (66) in conjunction with (H4SiW12O40)0.25 (67) affords the product in excellent yields (34–99%) and high enantioselectivities (up to 92%) with range of aromatic aldehydes. Furthermore primary amines were also employed in aldol reaction of acetaldehyde with isatin derivatives where desired product was obtained in excellent yields and moderate to good enantioselectivites (Scheme 14). Proposed mechanism involves hydrogen bonding between protonated amino group and carbonyl group of

acceptor aldehyde which plays a signicant role in governing the stereochemical outcome. In addition, the bulky tert-butyl group and the substituents on tertiary amine also favour the si-face attack of enamine to the acceptor aldehyde. Hypothesis was further conrmed by the results obtained in case of o- and pchloro substituted benzaldehydes where o-chlorobenzaldeyde gave the product with low enantioselectivity because of steric interference between ortho substituent and tertiary amino group (TS-7) while in case of para no such interaction take place, thus desired product in high enantioselectivity (TS-8). Qiao and co-workers developed a new dioctylamino group containing diarylprolinol-based catalyst (72) for cross aldol reaction of acetaldehyde in aqueous medium.41 The reaction in the presence of catalyst (72, 5 mol%), ionic liquid supported benzoic acid (73, 10 mol%) as co-catalyst, afforded product in good yields upto 97% in brine water. Even the reaction at low catalyst loading of 5 mol% affords high enantioselectivities upto 92% in brine solution with a range of aromatic aldehydes (Scheme 15). Quinine and 4-methoxybenzoic acid mediated cross-aldol reaction of acetaldehyde with a-ketophosphonate afforded a series of b-formyl-a-hydroxyphosphonates in good yields and moderate to good enantioselectivites.42 In an Initial attempt reaction was tried between propanal (74) and a-ketophosphonate (75) which undergo no product formation. The accepted reason was the unfavourable interaction between enamine methyl group and the bulky phosphonate group in the transition state. Thus, the reaction was further tried with acetaldehyde using proline and prolinamide as catalysts. As expected reaction undergo the product formation in signicant quantity (50% and 59% respectively), but aer a prolonged reaction time (7 days). Various catalysts had been screened where quinine derivative (76) were found to be suitable for reaction conditions and affording the product (77) in 75% yield and 93% ee in the presence of 4-methoxybenzoic acid as additive and toluene as solvent (Scheme 16). Both the electron donating and electron withdrawing substituents on the phenyl ring of benzoylphosphonates worked well under the standard reaction conditions. Furthermore the obtained products were screened for their anticancer potential where they were found to suppress the proliferation of human and murine tumor cells, while are mild against immortalized cells (HFF).

3.4. Michael reaction Scheme 18

Michael type addition reaction catalyzed by 4-OT

enzyme.

This journal is © The Royal Society of Chemistry 2015

Acetaldehyde also found its application in one of the vastly studied organo-catalysed reaction i.e. conjugate addition of

RSC Adv., 2015, 5, 55926–55937 | 55931

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances

Review

Acetaldehyde in organocatalysed reaction of nitrodienynes and nitroenynes.

Scheme 21

Scheme 20 Acetaldehyde in regioselective acyloin condensation.

nucleophiles to the b-position of a,b-unsaturated carbonyl compounds, known as Michael reaction.43 List and Hayashi and their co-workers independently reported the use of acetaldehyde in diarylprolinol (79) catalyzed Michael reaction of various substituted nitro styrenes (Scheme 17).44,45 Further List et al., have demonstrated the synthetic potential of obtained Michael product in pharmaceutically important compounds such as (R)-pregabalin (81)46 (S)-baclofen (82),47 (S)-rolipram (83)48 and 3-monosubstituted pyrrolidines (84).49 Inspired by the versatility of proline and its derivatives, Zandvoort and co-workers50 designed new biocatalysts (85) by combining organocatalyst with biocatalyst for Michael-type addition reaction of acetaldehyde with substituted b-nitrostyrenes. Reaction involved the use of 4-oxalocrotonate tautomerase (0.7 mol%) in water, afforded the product in moderate yield and high enantioselectivity up to 89% (Scheme 18). 3.5. Miscellaneous reactions Aer remarkable success in aldol, Mannich, Michael reaction, acetaldehyde was further used up in N-heterocyclic carbenecatalysed intermolecular Stetter reactions i.e.1,4-addition of aldehyde to an a,b-unsaturated compound. Yang et al.,51 used acetaldehyde for rst time as source of acyl anion which further engaged up in the reaction with Michael acceptors. Initially the reaction was screened with thiazolium (89) and triazolium precatalyst (90) with Cs2CO3 which provides the desired product in good yields but in racemic fashion. Further to achieve asymmetric product author employ other NHC catalysts where they found chiral cis-2-aminoindanol-derived triazolium salt (87) having N-penta uorophenyl substituent gave the better results with yield upto 85% and enantioselectivity upto 76%. Same group employed acetaldehyde in regioselective acyloin condensation wherein the reaction of acetaldehyde with electrophilic aromatic aldehydes in the presence of (89, 90) afforded the desired product52 (Scheme 22). According to proposed mechanism, in case of thiazolium catalyst the nucleophilic attack

55932 | RSC Adv., 2015, 5, 55926–55937

of NHC catalyst occurs on aromatic aldehydes rather than acetaldehyde, affords the most resonance-stabilized Breslow intermediate (94). In the next step nucleophilic carbene thus formed reacts with acetaldehyde to give the thermodynamically stable product (92). In contrast, triazolium catalyst (90) involves the rst nucleophilic attack on acetaldehyde 95 rather than aromatic aldehydes because of steric requirements and thus leads to the formation of product (93). In addition, authors had also described the effect of electronic properties of aromatic aldehydes on regioselective outcome. Thiazolium catalyst (89), gave better regioselectivity (higher 3 : 4 ratio) for electron-withdrawing substituents on the aromatic aldehydes while triazolium catalyst (90), the regioselectivity is better (higher 4 : 3 ratio) for electrondonating substituents on the aromatic aldehydes.

3.6. Nucleophilic addition of acetaldehyde to diyne system Importance of acetaldehyde can also be visualized from its application in the synthesis of biologically active compounds like (+)-a-lycorane and chiral b-alkynyl acids. Xue-Ling and co workers53 used acetaldehyde in the organocatalytic reaction of nitrodienynes (96) and nitroenynes (99). The conjugate addition reaction of acetaldehyde with 96 and 99 in the presence of diarylprolinolsilyl ethers (79), affords the desired product 97 and 100 respectively in good yields as well as in high enantioselectivities. Owing to various functionalities present in a single molecule, authors have successfully used their reaction product in enantioselective total synthesis of (+)-a-lycorane (98), which belongs to a alkaloid family, widely known for their different biological activities such as antiviral, antineoplastic, inhibitor of cell growth and cell division in higher plants.54 Finally they have shown that the product can also be converted to chiral b-alkynyl acids (101), which represent an important class of

Scheme 22

Acetaldehyde in organocatalytic asymmetric alkynylation.

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

Scheme

RSC Advances

23

Vinyl

acetate

mediated

cross-aldol

reaction

of

acetaldehyde.

pharmaceutical compounds having wide range of biological activities like PDE IV inhibitors, TNF inhibitors, GPR40 receptor agonists, and GRP receptor antagonists55(Scheme 21). Trost and Quintard56 have used acetaldehyde for organocatalytic asymmetric alkynylation to access propargylic alcohols. The between (1) and (102) reaction in the presence of catalyst (S,S) ProPhenol/POPh3 (103) (1 : 1) at 20  C resulted the product (105) in 78% yield and 86% enantioselectivity. Since the reaction was suffering with self aldol side reaction, therefore to overcome the problem, rate of addition of acetaldehyde in reaction mixture was studied, where slow addition over 30 minutes was found to be the condition of choice which resulted in high product formation and reduces the chances of self aldol reaction (Scheme 22).

4. Enolactetate as indirect source of acetaldehyde in asymmetric transformations To overcome the problems encountered in the direct use of acetaldehyde, a lot of research is ongoing where acetaldehyde has been generated in situ from stable and easy to handle chemicals. Paraldehyde,57 vinyl acetate,58 and vinyl ethers59 are being mostly used as the indirect sources of acetaldehyde in asymmetric organic transformations. 4.1. Vinyl acetate in cross-aldol reaction Since the direct use of acetaldehyde generally suffers with various side uncontrolled reaction. Also, some of the authors suggested that slow addition of acetaldehyde can overcome the problem of high reactivity of acetaldehyde, this can also be achieved by in situ generation of acetaldehyde in reaction mixture. Since in the past vinyl acetate is well known for in situ generation of acetaldehyde during the transesterication of hydroxyl compounds60 but the rst report of vinyl acetate in aldol condensation appears when Gupta and coworkers61 were trying for asymmetric transesterication of tricyclic diketone in presence of Novozym 435 lipase. Instead of getting the expected transacetylated product they got the condensation product of tricyclic ketone and acetaldehyde. Reaction involves rst enzyme-catalysed nucleophilic addition of (1) with acetaldehyde resulting in the formation of mono-adduct which further undergoes second condensation reaction (1) to afford the bisadduct in 97% yield, although in racemic fashion.

This journal is © The Royal Society of Chemistry 2015

Scheme 24 Vinyl acetate mediated asymmetric cross-aldol reaction of acetaldehyde.

In 2011, our group reported the rst attempt on the use of vinyl acetate (109) in asymmetric cross-aldol reaction where lipase was used in tandem with organocatalyst.62 The reaction apparently involves lipase (110) catalyzed in situ generation of active form of acetaldehyde from vinyl acetate followed by organocatalysed aldol reaction with aromatic aldehydes for the preparation of b-hydroxy aldehydes. In the initial studies reaction was studied with triethylamine in tandem with lipase where the product (111) obtained in good yields but is racemic in nature. Aer the initial success, reaction was tested for tolerance of the lipase with organocatalyst towards variously substituted aromatic aldehydes. The reaction of in the presence of L-proline and lipase proceeded well and afforded the corresponding product (113) in moderate yields (42–65%) but in quite low enantioselectivity ranging between 15–20%. To further explore the use of vinyl acetate as an alternative source of acetaldehyde our group have recently employed vinyl acetate in for highly enantioselective cross-aldol reaction63 (Scheme 24). The reaction apparently involves in situ generation of acetaldehyde from vinyl acetate through lipase Novozym 435 followed by the cross aldol reaction in the presence of bistriuoro substituted diphenylprolinol (18). Under the optimized conditions, reaction between acetaldehyde and various aromatic aldehydes afford the cross-aldol product in good yields (60–94%) and high enantioselectivities (up to 99%). Reaction mechanism was proposed by performing density functional energy (DFT) calculations. Four possible transition sates were studied i.e. anti–Re face (TS9), anti–Si face (TS10), syn–Re face (TS11) and syn–Si (TS-12) face. All the transition state involves

RSC Adv., 2015, 5, 55926–55937 | 55933

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances

Trypsin-catalyzed tandem reaction: one-pot synthesis of 3,4-dihydropyrimidin-2(1H)-ones by in situ formed acetaldehyde.

Scheme 25

Scheme 26 Asymmetric cross aldol reaction of acetaldehyde in deep eutectic solvent (DES).

hydrogen bonding between the acidic OH group of the triuoromethyl-substituted catalyst (18) and the carbonyl group of acetaldehyde, which activates the substrate for reaction as well as controlling the direction of approach. DFT calculation predict the most stable transition state as (TS9) where the aryl rings of the aldehyde and catalyst are sufficiently far apart resulted in the formation of stable adduct having minimum energy (20.9 kcal mol 1). Also the synthetic importance of crossaldol product is well addressed in the asymmetric synthesis of a,b-unsaturated d-lactones, a well known structural motif present in a wide variety of bioactive natural scaffolds.64 A simple, mild, one-pot tandem method catalyzed by trypsin was developed65 for the synthesis of 3,4-dihydropyrimidin2(1H)-ones (119) by the Biginelli reaction of urea (117), b-dicarbonyl compounds (118), and in situ-formed acetaldehyde from vinyl acetate (109). Trypsin was found to display dual promiscuous functions to catalyze transesterication and the Biginelli reaction in sequence (Scheme 25). Recently Maria and coworkers66 have reported the asymmetric cross aldol reaction of acetaldehyde in deep eutectic solvent (DES) i.e. choline chloride and glycerol, by using lipase Cal-B and proline derived organocatalysts. Screening of organocatalysts showed that triuoromethyl substituted proline catalyst (18) is the most suitable organocatalyst for this transformation and afford the product (121) in 70% yield and 95% enantiomeric excess. Author have successfully employed the DES solvent with lipase, where both can be reused up to 6 cycles without any loss of activity, while fresh organocatalyst is required every time for good results. 4.2. Paraldehyde in asymmetric Michael reaction Aer the successful employment of vinyl acetate as an indirect source of acetaldehyde, Pericas et al.67 recently showed the use of cyclic trimer paraldehyde (110), as a precursor for acetaldehyde. Since its more stable liquid as compared to acetaldehyde, thereby reduces the challenges associated with handling of acetaldehyde. The reaction involves the acid-catalyzed

55934 | RSC Adv., 2015, 5, 55926–55937

Review

Scheme 27 Paraldehyde mediated asymmetric Michael reaction of acetaldehyde.

depolymerisation of paraldehyde (110), which in situ generates acetaldehyde and further in the presence of amino-catalyst undergoes Michael reaction. Since both catalysts were incompatible to each other so authors tried to exploit the site isolation principle, where polystyrene supported catalysts (123) and (124) were used in the reaction of paraldehyde and trans-b-nitrostyrene. As expected both the catalysts corporate with each other and affords the product (80) in 42% yield and 90% ee. Reaction also shows the improved results when 123 were conned to teabag in CH2Cl2 solvent (Scheme 27).

5.

Conclusions

In summary, this review summarises both the recent developments in the eld of organocatalysed use of acetaldehyde as a substrate in various organic transformations and the application of acetaldehyde derived products in the synthesis of bioactive molecules. Inspite the parallel development of acetaldehyde as an efficient substrate, the discovery of highly effective organocatalysts for nucleophilic addition of acetaldehyde to the various electrophilic centres was undoubtedly a momentous point for recent development of the eld. We hope that this short review will further stimulate collective thinking on design, synthesis of organocatalysts and new approaches for stereocenter forming reactions involving unexplored substrates, to increase the molecular diversity in organic transformations.

Acknowledgements We thank DST, New Delhi for nancial assistance. Also M.K. thanks CSIR, New Delhi for the award of Research Fellowship. IIIM Communication no. 1802.

Notes and references 1 (a) T. Hayashi and M. Kumada, Asymmetric Synthesis, ed. J. D. Morrison, Academic Press, Inc., 1985, vol. 5, p. 147; (b) R. Mahrwald, Enantioselective Organocatalyzed Reactions II: Asymmetric C-C Bond Formation Processes, Springer, New York, 2011. 2 (a) U. Scheffler and R. Mahrwald, Chem.–Eur. J., 2013, 19, 14346; (b) S. Kobayashi, Y. Mori, J. S. Fossey and M. M. Salter, Chem. Rev., 2011, 111, 2626. 3 (a) A. N. Collins, G. N. Sheldrake and J. Crosby, Chirality in Industry II, Wiley, New York, 1997; (b) G. Beck, Synlett, 2002, 6, 837.

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

4 (a) R. Mahrwald, Modern Aldol Reactions, Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim, Germany, 2004, vol. 1 and 2; (b) C. H. Heathcock, in Asymmetric Synthesis, ed. J. D. Morrison, Academic Press, Orlando, FL, 1984, vol. 3, p. 111; (c) C. H. Heathcock, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, Oxford, 1991, vol. 2, p. 133; (d) D. Gryko and D. Walaszek, Stereoselective Organocatalysis: Bond Formation Methodologies and Activation Modes, ed. R. R. Torres, John Wiley & Sons, Inc., Hoboken, New Jersey, ch. 3, 2013; (e) T. D. Machajewski and C.-H. Wong, Angew. Chem., Int. Ed., 2000, 39, 1352; (f) C. Palomo, M. Oiarbide and J. M. Garcia, Chem. Soc. Rev., 2004, 33, 65. 5 (a) K. C. Nicolaou, S. A. Snyder, T. Montagnon and G. Vassilikogiannakis, Angew. Chem., Int. Ed., 2002, 41, 1668; (b) J. A. Norton, Chem. Rev., 1942, 42, 319; (c) R. Huisgen, R. Grashey and J. Sauer, in The Chemistry of the Alkenes, ed. S. Patai, Wiley, London, 1964, p. 739; (d) D. A. Evans and J. S. Johnson, in Comprehensive Asymmetric Catalysis, ed. E. N. Jacobsen, A. Pfaltz and H. Yamamoto, Springer, New York, 1999, vol. 3, p. 1177; (e) W. Oppolzer, in Comprehensive Organic Synthesis, ed. B. M. Trost, Pergamon Press, New York, 1991, vol. 5; (f) H. B. Kagan and O. Riant, Chem. Rev., 1992, 92, 1007. 6 (a) S. C. Jha and N. N. Joshi, ARKIVOC, 2002, 7, 167; (b) M. E. Jung, Stabilized Nucleophiles with Electron Decient Alkenes and Alkynes, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, Elmsford, NY, 1991, vol. 4, ch. 1.1; (c) P. Perlmutter, Conjugate Addition Reactions in Organic Synthesis, Pergamon Press, Elmsford, NY, 1992; (d) Y. Zhang and W. Wang, Catal. Sci. Technol., 2012, 2, 42. 7 (a) T. Graening and H. G. Schmalz, Angew. Chem., Int. Ed., 2003, 42, 2580; (b) B. M. Trost and M. L. Crawley, Chem. Rev., 2003, 103, 2921; (c) B. M. Trost and T. J. Fullerton, J. Am. Chem. Soc., 1973, 95, 292; (d) B. M. Trost and D. L. V. Vranken, Chem. Rev., 1996, 96, 395; (e) B. M. Trost, Org. Process Res. Dev., 2012, 16, 185. 8 (a) M. Ueno and S. Kobayashi, Enantioselective Synthesis of bAmino Acids, ed. E. Juaristi and V. A. Soloshonok, Wiley-VCH, New York, 2nd edn, 2005, p. 139; (b) J. M. Verkade, L. J. Vhemert, P. J. Quaedieg and F. P. Ruties, Chem. Soc. Rev., 2008, 37, 29; (c) A. Cordova, Acc. Chem. Res., 2004, 37, 102. 9 (a) P. Langer, Angew. Chem., Int. Ed., 2000, 39, 3049; (b) D. Basavaiah, A. J. Rao and T. Satyanarayana, Chem. Rev., 2003, 103, 811; (c) Y. Wei and M. Shi, Chem. Rev., 2013, 113, 6659; (d) F. L. Hu and M. Shi, Org. Chem. Front., 2014, 1, 587. 10 (a) S. G. Davies, Organotransition Metal Chemistry Applications to organic Synthesis, Pergamon, Oxford, 1982; (b) E. Meggers, Chem.–Eur. J., 2010, 752; (c) B. G. Davis and V. Boyer, Nat. Prod. Rep., 2001, 18, 618; (d) S. I. Bertelsen and K. A. Jørgensen, Chem. Soc. Rev., 2009, 38, 2178. 11 (a) P. I. Dalko and L. Moisan, Angew. Chem., Int. Ed., 2004, 43, 5138; (b) M. J. Gaunt, C. C. Johansson, A. McNally and N. T. Vo, Drug Discovery Today, 2007, 12, 1; (c) C. Grondal, This journal is © The Royal Society of Chemistry 2015

RSC Advances

12

13 14 15 16 17

18

19 20

21 22 23

24

M. Jeanty and D. Enders, Nat. Chem., 2010, 2, 167; (d) J. Alem´ an and S. Cabrera, Chem. Soc. Rev., 2013, 42, 774; (e) M. C. Holland and R. Gilmour, Angew. Chem., Int. Ed., 2015, 54, 3862; (f) F. G. Finelli, L. S. M. Miranda and R. O. M. A. de Souza, Chem. Commun., 2015, 51, 3708. (a) D. W. C. MacMillan, Nature, 2008, 455, 304; (b) S. V. Pansare and E. K. Paul, Chem.–Eur. J., 2011, 17, 8770; (c) S. Mukherjee, J. W. Yang, S. Hoffmann and B. List, Chem. Rev., 2007, 107, 5471; (d) U. Scheffler and R. Mahrwald, Chem.–Eur. J., 2013, 19, 14346; (e) A. Dondoni and A. Massi, Angew. Chem., Int. Ed., 2008, 47, 4638. Z. G. Hajos and D. R. Parrish, in Ger. Pat. DE 21026, 29 July 1971transJ. Org. Chem., 1974, 39, 1615. U. Eder, G. Sauer and R. Wiechert, in Ger. Pat. DE 2014757, 7 October 1971transAngew. Chem., Int. Ed. Engl., 1971, 10, 496. B. List, R. A. Lerner and C. F. Barbas, J. Am. Chem. Soc., 2000, 122, 2395. (a) B. Alcaide and P. Almendros, Angew. Chem., Int. Ed., 2008, 47, 4632; (b) S. M. Kim, Synlett, 2014, 25, 153. (a) L. G. Wade, Organic Chemistry, Prentice Hall, Upper Saddle River, New Jersey, 6th edn, 2005, p. 1056; (b) M. B. Smith and J. March, Advanced Organic Chemistry, Wiley Interscience, New York, 5th edn, 2001, p. 1218. (a) B. M. Kim, S. F.Williams and S. Masamune, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, Oxford, 1991, vol. 2, p. 239; (b) I. Paterson, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, Oxford, 1991, vol. 2, p. 301; (c) C. Gennari, in Comprehensive Organic Synthesis, ed. B. M. Trost and I. Fleming, Pergamon Press, Oxford, 1991, vol. 2, p. 629; (d) M. Braun, in Stereoselective Synthesis(Houben-Weyl, Methods of Organic Chemistry), ed. G. Helmchen, R. W. Hoffmann, J. Mulzer and E. Schaumann, George Thieme Verlag, Stuttgart, 1995, vol. E21b, p. 1603; (e) A. S. Franklin and I. Paterson, Contemp. Org. Synth., 1994, 1, 317. A. Cordova, W. Notz and C. F. Barbas, J. Org. Chem., 2002, 67, 301. (a) G. E. Keck, A. Palani and S. F. McHardy, J. Org. Chem., 1994, 59, 3113; (b) L. R. Hillis and R. C. Ronald, J. Org. Chem., 1985, 50, 470; (c) C. J. Hayes and C. H. Heathcock, J. Org. Chem., 1997, 62, 2678. S. E. Denmark and T. Bui, J. Org. Chem., 2005, 70, 10190. Y. Hayashi, S. Samanta, T. Itoh and H. Ishikawa, Org. Lett., 2008, 10, 5581. (a) C. Mannich and W. Krosche, Arch. Pharm., 1912, 250, 647; (b) S. Kobayashi and M. Ueno, in Comprehensive Asymmetric Catalysis, Supplement I, ed. E. N. Jacobsen, A. Pfaltz and H. Yamamoto, Springer, Berlin, 2003, ch. 29.5; (c) M. Arend, B. Westermann and N. Risch, Angew. Chem., Int. Ed., 1998, 37, 1044; (d) M. M. B. Marques, Angew. Chem., Int. Ed., 2006, 45, 348; (e) M. Shibasaki and S. Matsunaga, J. Organomet. Chem., 2006, 691, 2089; (f) A. Cordova, Acc. Chem. Res., 2004, 37, 102. J. W. Yang, C. Chandler, M. Stadler, D. Kampen and B. List, Nature, 2008, 452, 453. RSC Adv., 2015, 5, 55926–55937 | 55935

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

RSC Advances

25 (a) F. A. Davis and J. M. Szewczyk, Tetrahedron Lett., 1998, 39, 5951; (b) G. B. Fodor and B. Colasanti, in Alkaloids: Chemical and Biological Perspectives, ed. S. W. Pelletier, Wiley, New York, 1985, pp. 1–90. 26 (a) P. Dorr, M. Westby, S. Dobbs, P. Griffin, B. Irvine, M. Macartney, J. Mori, G. Rickett, C. S. Burchnell, C. Napier, R. Webster, D. Armour, D. Price, B. Stammen, A. Wood and M. Perros, Antimicrob. Agents Chemother., 2005, 49, 4721; (b) D. A. Price, S. Gayton, M. D. Selby, J. Ahman, S. H. Lewandowski, B. L. Stammen and A. Warren, Tetrahedron Lett., 2005, 46, 5005. 27 (a) L. A. Sorbera, J. Castaner and R. M. Castaner, Drugs Future, 2004, 29, 1201; (b) S. A. Siddiqui and K. V. Srinivasan, Tetrahedron: Asymmetry, 2007, 18, 2099. 28 S. G. Davies, A. C. Garner, P. M. Roberts, A. D. Smith, M. J. Sweeta and J. E. Thomson, Org. Biomol. Chem., 2006, 4, 2753. 29 (a) P. V. Ramachandran and T. E. Burghardt, Chem.–Eur. J., 2005, 11, 4387; (b) H. S. Lee, J. S. Park, B. M. Kim and S. H. Gellman, J. Am. Chem. Soc., 2007, 129, 6050. 30 Y. Hayashi, T. Okano, T. Itoh, T. Urushima, H. Ishikawa and T. Uchimaru, Angew. Chem., Int. Ed., 2008, 47, 9053. 31 Q. Chang, J. Zhou and L.-H. Gan, J. Phys. Org. Chem., 2012, 25, 667. 32 C. Chandler, P. Galzerano, A. Michrowska and B. List, Angew. Chem., Int. Ed., 2009, 48, 1978. 33 T. Kano, Y. Yamaguchi and K. Maruoka, Angew. Chem., Int. Ed., 2009, 48, 1838. 34 Y. Hayashi, T. Itoh, S. Aratake and H. Ishikawa, Angew. Chem., Int. Ed., 2008, 47, 2082. 35 N. Hara, S. Nakamura, N. Shibata and T. Toru, Chem.–Eur. J., 2009, 15, 6790. 36 T. Itoh, H. Ishikawa and Y. Hayashi, Org. Lett., 2009, 11, 3854. 37 Q. Guo and J. C.-G. Zhao, Tetrahedron, 2012, 53, 1768. 38 W.-B. Chen, X.-L. Du, L.-F. Cun, X.-M. Zhang and W.-C. Yuan, Tetrahedron, 2010, 66, 1441. 39 (a) S. Pizzarello and A. L. Weber, Science, 2004, 303, 1151; (b) Z. Jiang, Z. Liang, X. Wu and Y. Lu, Chem. Commun., 2006, 2801; (c) T. Kanemitsu, A. Umehara, M. Miyazaki, K. Nagata and T. Itoh, Eur. J. Org. Chem., 2011, 2011, 993. 40 S. Hu, L. Zhang, J. Li, S. Luo and J.-P. Cheng, Eur. J. Org. Chem., 2011, 2011, 3347. 41 Y. Qiao, Q. Chen, S. Lin, B. Ni and A. D. Headley, J. Org. Chem., 2013, 78, 2693. 42 S. Perera, V. K. Naganaboina, L. Wang, B. Zhang, Q. Guo, L. Rout and C.-G. Zhao, Adv. Synth. Catal., 2011, 353, 1729. 43 (a) T. Komnenos, Liebigs Ann. Chem., 1883, 218, 145; (b) A. Michael, J. Prakt. Chem., 1887, 36, 349; (c) B. E. Rossiter and N. M. Swingle, Chem. Rev., 1992, 92, 771; (d) P. Perlmutter, Conjugate Addition Reactions in Organic Synthesis, Pergamon Press, Oxford, 1992. 44 P. Garc´ıa-Garc´ıa, A. Lad´ epˆ eche, R. Halder and B. List, Angew. Chem., Int. Ed., 2008, 47, 4722. 45 Y. Hayashi, T. Itoh, S. Aratake and H. Ishikawa, Angew. Chem., Int. Ed., 2008, 47, 2082.

55936 | RSC Adv., 2015, 5, 55926–55937

Review

46 (a) R. H. Dworkin and P. Kirkpatrick, Nat. Rev. Drug Discovery, 2005, 4, 455; (b) P. J. Siddall, M. J. Cousins, A. Otte, T. Griesing, R. Chambers and T. K. Murphy, Neurology, 2006, 67, 1792. 47 A. Muzyk, S. Rivelli and J. Gagliardi, CNS Drugs, 2012, 26, 69. 48 (a) J. Zhu, E. Mix and B. Winblad, CNS Drug Rev., 2001, 7, 387; (b) A. L. L. Garcia, M. J. S. Carpes, A. C. B. M. de Oca, M. A. G. dos Santos, C. S. C. Santan and C. R. D. Correia, J. Org. Chem., 2005, 70, 1050. 49 (a) J. L. Watts, S. A. Salmon, M. S. Sanchez and R. J. Yancey Jr, Antimicrob. Agents Chemother., 1997, 41, 1190; (b) J. M. Domagala, S. E. Hagen, T. Joannides, J. S. Kiely, E. Laborde, M. C. Schroeder, J. A. Sesnie, M. A. Shapiro, M. J. Suto and S. Vanderroest, J. Med. Chem., 1993, 36, 871; (c) J. Katarzynska, A. Mazur, M. Bilska, E. Adamek, M. Zimecki, S. Jankowski and J. Zabrocki, J. Pept. Sci., 2008, 14, 1283; (d) Y. J. Kim, D. A. Kaiser, T. D. Pollard and Y. Ichikawa, Bioorg. Med. Chem. Lett., 2000, 10, 2417. 50 (a) E. M. Geertsema, Y. Miao, P. G. Tepper, P. de Haan, E. Zandvoort and G. J. Poelarends, Chem.–Eur. J., 2013, 19, 14407; (b) E. Zandvoort, E. M. Geertsema, B.-J. Baas, W. J. Quax and G. J. Poelarends, Angew. Chem., Int. Ed., 2012, 51, 1240. 51 S. M. Kim, M. Y. Jin, M. J. Kim, Y. Cui, Y. S. Kim, L. Zhang, C. E. Song, D. H. Ryu and J. W. Yang, Org. Biomol. Chem., 2011, 9, 2069. 52 M. Y. Jin, S. M. Kim, H. Han, D. H. Ryu and J. W. Yang, Org. Lett., 2011, 13, 880. 53 X.-L. Meng, T. Liu, Z.-W. Sun, J.-C. Wang, F.-Z. Peng and Z.-H. Shao, Org. Lett., 2014, 16, 3044. 54 O. Hoshino, in The Alkaloids, ed. G. A. Cordell, Academic Press, New York, 1998. 55 (a) T. Shimada, H. Ueno, K. Tsutsumi, K. Aoyagi, T. Manabe, S. Y. Sasaki and S. Katoh, Int Appl PCT, WO 2009054479, 2009; (b) J. Houze, J. Liu, Z. Ma, J. C. Medina, M. J. Schmitt, R. Sharma, Y. Sun, Y. Wang and L. Zhu, US Pat. 7,465,804, 2008; (c) S. P. Brown, P. Dranseld, Z. Fu, J. Houze, X. Y. Jiao, T. J. Kohn, V. Pattaropong, M. Vimolratana and M. J. Schmitt, Int Appl PCT, WO 2008130514, 2008; (d) S. B. Bharate, K. V. Nemmani and R. A. Vishwakarma, Expert Opin. Ther. Pat., 2009, 19, 237. 56 A. Quintard and B. M. Trost, Angew. Chem., 2012, 124, 6808. 57 T. V. Vasudevan and M. Sharma, Ind. Eng. Chem. Process Des. Dev., 1983, 22, 161. 58 D. Mukherjee, B. A. Shah, P. Gupta and S. C. Taneja, J. Org. Chem., 2007, 72, 8965. 59 M. B. Boxer and H. Yamamoto, J. Am. Chem. Soc., 2006, 128, 48. 60 (a) U. Alder, D. Breitgoff, P. Klein, K. Laumen and P. Schneirder, Tetrahedron Lett., 1989, 30, 1793; (b) A. M. Brzozowski, U. Derewenda, Z. S. Derewenda, G. G. Dodson, D. M. Lawson, J. P. Turkenburg, F. Bjorkling, B. H. Jensen, S. A. Patkar and L. Thim, Nature, 1991, 351, 91. 61 A. B. Majumder, N. G. Ramesh and M. N. Gupta, Tetrahedron Lett., 2009, 50, 5190.

This journal is © The Royal Society of Chemistry 2015

View Article Online

Published on 17 June 2015. Downloaded by INDIAN INSTITUTE OF TECHNOLOGY BOMBAY on 11/10/2015 19:19:00.

Review

62 M. Kumar, B. A. Shah and S. C. Taneja, Adv. Synth. Catal., 2011, 353, 1207. 63 M. Kumar, A. Kumar, M. Rizvi, M. Mane, K. Vanka, S. C. Taneja and B. A. Shah, Eur. J. Org. Chem., 2014, 24, 5247. 64 (a) M. A. Koch, A. Schuffenhauer, M. Scheck, S. Wetzel, M. Ca-saulta, A. Odermatt, P. Ertl and H. Waldmann, Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 17272; (b) V. Boucard, G. Broustal and J. M. Campagne, Eur. J. Org. Chem., 2007, 225; (c) P. Barbier and F. Schneider, J. Org. Chem., 1988,

This journal is © The Royal Society of Chemistry 2015

RSC Advances

53, 1218; (d) Q. Zhu, L. Qiao, Y. Wu and Y.-L. Wu, J. Org. Chem., 2001, 66, 2692. 65 Z. B. Xie, N. Wang, W. X. Wu, Z. G. Le and X. Q. Yu, J. Biotechnol., 2014, 170, 1. 66 C. R. Muller, I. Meiners and P. D. Maria, RSC Adv., 2014, 4, 46097. 67 C. R. E. X. Fan, S. Sayalero and M. A. Pericas, Chem.–Eur. J., 2013, 19, 10814.

RSC Adv., 2015, 5, 55926–55937 | 55937