Open access peer-reviewed chapter

Total Synthesis of Macrolides

Written By

Chebolu Naga Sesha Sai Pavan Kumar

Submitted: 29 March 2019 Reviewed: 07 June 2019 Published: 22 July 2019

DOI: 10.5772/intechopen.87898

From the Edited Volume

Organic Synthesis - A Nascent Relook

Edited by Belakatte Parameshwarappa Nandeshwarappa

Chapter metrics overview

1,775 Chapter Downloads

View Full Metrics

Abstract

Structurally complex macrolide natural products, isolated from a variety of marine and other sources, continue to provide a valuable source of targets for the synthetic chemist to embark. In this account, we provide the recent progress and pathways in the total synthesis of macrolides and discussed the synthesis of (+)-neopeltolide, aspergillide D, miyakolide and acutiphycin natural products.

Keywords

  • macrolide
  • aldolization
  • macrolactonization
  • ring closing metathesis
  • coupling reactions
  • total synthesis

1. Introduction

Macrolides are a class of antibiotics that consist of a large macrocyclic lactone ring attached to deoxy sugars. These antibiotics are bacteriostatic in nature and act by inhibiting protein synthesis of bacteria. These are obtained mainly from certain actinomycetes genus, such as Streptomyces and related species. The original macrolide complex, erythromycin A, was isolated in 1952 as a natural product of Saccharopolyspora erythraea (formerly Streptomyces erythreus). Other examples include clarithromycin, azithromycin, telithromycin, cethromycin, modithromycin, etc. Macrolides structurally contain three characteristic parts in every molecule, that is, a macrocyclic lactone ring, multiple ketone & hydroxyl group, and two deoxy sugars attached by glycosidic bond. According to the carbon number of lactone ring, macrolides are classified into several types. That is, 12-membered ring, 13-membered ring, 14-membered ring, 15-membered ring, 16-membered rings, etc. (Figure 1). Out of these, most of the antibiotic drugs comprised of 14-membered and 16-membered lactone rings.

Figure 1.

Classification of macrolide antibiotics.

The construction of macrocyclic structures is a recurrent and challenging problem in synthetic organic chemistry. Theoretically, macrocyclic systems can be generated by cyclization of open, long chain precursors or by cleavage of internal bonds in polycyclic systems. In the course of synthesis, numerous problems are encountered to achieve target molecules. Despite the several problems, however, recent interest in the chemistry of macrolide antibiotics and other biologically active macrolactones and macrolactams resulted in the discovery and development of several new synthetic methods for macrolide formation. In this chapter, total synthesis of some of the macrolides is discussed with scrupulous emphasis on the key macrolide ring forming reactions.

Advertisement

2. Synthetic strategy for macrolide synthesis

In the polyoxomacrolide ring, generally we will observe the 1,3-diol systems as a core. There are two synthetic approaches for the edifice of 1,3-diols which are illustrated here. They are asymmetric aldol reaction and the other one is asymmetric epoxide and epoxide ring-opening.

2.1 Aldolization

Asymmetric synthesis of β-hydroxy ketones by aldol reactions of ketones with aldehydes is the general and efficient method for the synthesis of 1,3-diol systems and is of great interest in the field of total synthesis. By using the range of chiral ketones, highly diastereoselective syn and anti aldol products are produced using various boron enolates [1, 2, 3]. Some of the reagents shown below (Figure 2) direct the relative and absolute stereochemistry of C▬C bond formation between various achiral and chiral ketones, thus providing a ubiquitous synthetic tool for macrolide synthesis (Figure 3).

Figure 2.

Reagents for asymmetric Aldol reactions.

Figure 3.

Macrolide antibiotics: evidence for the chemists interested in the stereochemistry of the Aldol reaction.

A highly efficient and extensively used method for diastereoselective aldol reactions is the Evans aldol reaction using boron enolate derived from a chiral imide [4, 5]. Upon treatment of imide 1 with n-Bu2BOTf and i-Pr2NEt in CH2Cl2 followed by addition of aldehyde, aldol reaction proceeds smoothly in stereoselective manner through the chelation transition state to attain 1,2-syn-aldol adduct 2 in high yield and with excellent diastereoselectivity. After the reaction, the chiral auxiliary is cleaved by hydrolysis to acid, then reduction to aldehyde or alcohol, conversion to Weinreb amide, etc. In contrast, addition of a Lewis acid to the boron enolate provides either anti-diol 3 or non-Evans 1,2-syn-aldol 4 with excellent diastereoselectivity [6] (Figure 4).

Figure 4.

Evan’s Aldol strategy.

2.2 Asymmetric epoxidation and dihydroxylation

The asymmetric epoxidation of allylic alcohols introduced by Katsuki and Sharpless in 1980 has tremendous applications in the synthesis of various compounds [7]. The Sharpless asymmetric epoxidation (AE) is the efficacious reagent in the synthetic organic chemistry particularly in the synthesis of variety of natural products. Epoxidation is carried out from allylic alcohols 5 with tert-butyl hydroperoxide in the presence of Ti(OiPr)4. The resulting epoxide stereochemistry is determined by the enantiomer of the chiral tartrate ester (usually diethyl tartrate or diisopropyl tartrate) employed in reaction. When (−)-diester is used, β-epoxide 6 is obtained, while (+)-diester produces α-epoxide 7 (Figure 5).

Figure 5.

Sharpless asymmetric epoxidation strategy.

The Sharpless dihydroxylation [8] is another tool used in the enantioselective preparation of 1,2-diols (9a/9b) from olefins (8). This reaction is performed with osmium catalyst and a stoichiometric oxidant (e.g., K3Fe(CN)6 or NMO). Enantioselectivity is produced by the addition of enantiomerically-enriched chiral ligands [(DHQD)2PHAL also called AD-mix-β, (DHQ)2PHAL also called AD-mix-α or their derivatives] (Figure 6). These reagents are also commercially available as stable and not so expensive.

Figure 6.

Sharpless asymmetric dihydroxylation.

Stereoselective ring-opening of 2,3-epoxy alcohols 10 is extremely valuable for the synthesis of different functionalized compounds [9]. A wide range of nucleophiles such as secondary amine, alcohol, thiol, azide and carboxylic acid predominantly at C-3 position to give 1,2-diol 11 (Figure 7).

Figure 7.

Stereoselective ring-opening of epoxy alcohols.

The logic of macrocyclization in natural product synthesis can be investigated by different strategies; some of them are Prins reaction [10], lactonization [11, 12], ring closing metathesis [13], Wittig reaction [14], Horner Wadsworth Emmons (HWE) reaction [15], Julia-Kocienski reaction [16], metal-mediated cross coupling reaction [17], etc. However, it is true that there is no universal macrocyclization method is reliable in the total synthesis of natural products.

2.3 Macrolactonization

Macrolactonization is the one of the effective and popular methods in the synthesis of macrolactones. The method is based on the lactonization of the corresponding seco-acid. Thus various methods are reported in the literature for the macrolactone synthesis, some of the most commonly used methods are Corey-Nicolaou [18], Shiina [19], Yamaguchi [20], Mitsunobu [21], Keck-Boden [22], and Mukaiyama [23] macrolactonizations (Figure 8).

Figure 8.

Some of the popular methods used for macrolactonization.

Hansen et al. [24] reported the synthesis of (−)-aplyolide A from 12 in which they adopted the Corey-Nicolaou macrolactonisation as the key step with 78% yield (Figure 9).

Figure 9.

Application of Corey-Nicolaou macrolactonisation.

Narasaka et al. [25] used the Mukaiyama method for the effective construction of macrocycle ring from corresponding seco-acid 13 in the synthesis of Prostaglandin F-lactone (Figure 10).

Figure 10.

Mukaiyama method in the synthesis of prostaglandin F-lactone.

Enev et al. [26] in his studies towards the total synthesis of laulimalide, crucial Yamaguchi macrolactonization was employed on the ynoic seco-acid 14 and then reducing the triple bond obtained the desired macrolactone 15 (Figure 11).

Figure 11.

Yamaguchi protocol in the synthesis of laulimalide.

In the synthetic studies towards the synthesis of colletodial, Keck et al. [27] effectively used DCC-DMAP protocol for the macrolactonization of 16 to precursor of colletodial 17 (Figure 12).

Figure 12.

Keck et al. lactonisation for the synthesis of colletodial.

Mitsunobu macrolactonization protocol based on the activation of the seco-acid alcohol 18 to 19 using diisopropyl azodicarboxylate (DIAD) and triphenylphosphine is used in the total synthesis of natural product (+)-amphidinolide K by Williams and Meyer [28] (Figure 13).

Figure 13.

Mitsunobu esterification in the total synthesis of (+)-amphidinolide K.

In the total synthesis of iejimalide by Schweitzer et al. [29], Shiina macrolactonization (2-methyl-6-nitro benzoic anhydride/DMAP) is used as the key step for the construction of macrolactone 21 in moderate yield. Even the yield is somewhat low, other methods failed to construct the lactone while Shiina protocol worked successfully from 20 (Figure 14).

Figure 14.

Shiina macrolactonization towards the synthesis of iejimalide B.

2.4 Ring-closing olefin metathesis

In recent years, ring closing metathesis (RCM) has become one of the most paramount tools in synthetic organic chemistry especially in the field of total synthesis of macrolide natural products [13, 30, 31]. Furthermore, RCM is becoming the most popular way to construct large rings and has the advantage of being compatible with a wide range of functional groups such as ketones, ethers, esters, amides, amines, epoxides, silyl ethers, alcohols, thioesters, etc. In view of this, among the several reagents developed by Grubbs, Shrock, and Chauvin, the catalysts AD represents two generations of ruthenium complexes, while E is the molybdenum Shrock catalyst (Figure 15). A is popularly known as Grubbs first generation catalysts, B and C are Grubbs second generation catalysts and D is Hoveyda-Grubbs catalyst. The choice of the catalysts can be used in the synthetic organic transformations based on the reactivity of the substrate, and other reaction condition parameters. Substitution in the aromatic ring of D has given rise to a new family of third generation catalyst.

Figure 15.

Various catalysts for ring closing metathesis.

Here, some of the applications of ring closing metathesis in the total synthesis of macrolides salicylihalamide A [32], trans-resorcylide [33], (+)-lasiodiplodin [34], oximidine III [35], and Sch 38516 [36] by various metathesis catalysts have been illustrated (Figure 16).

Figure 16.

Some of the applications of ring closing metathesis in total synthesis of macrolides.

2.5 Palladium catalyzed coupling reactions

Palladium-catalyzed coupling reactions have gained more attention in recent years in the field of organic chemistry. In this course, Suzuki reaction using organoboron compounds [37], Heck reaction using alkenes [38], Stille reaction with organostannate [39, 40], Sonogashira reaction with terminal alkyne [41] and Tsuji-Trost reaction with π-allylpalladium intermediate [42, 43], etc. are the most frequently employed reactions in the total synthesis of macrolide natural products. Some of them are depicted here.

Tortosa and co-workers [44] in the synthesis of (+)-superstolide A, Suzuki macrocyclisation approach is used for the construction of 24-membered macrocyclic octane 23 (Figure 17).

Figure 17.

Suzuki macrocyclisation in the synthesis of superstolide A.

The first application of the Heck cyclisation to a macrocyclic substrate was reported by Ziegler and co-workers [45] in 1981 during the synthesis of aglycone of the macrocyclic antibiotic carbomycin B. They achieved the cyclisation to the model substrate 25 in 55% yield, by slow addition to a solution of PdCl2(MeCN)2, Et3N and formic acid in MeCN at ambient temperature (Figure 18).

Figure 18.

Heck cyclisation in the synthesis of carbomycin B.

Stille macrocyclisation as illustrated below used as the key step in the total synthesis of the biselyngbyolide A by Tanabe et al. [46] using Pd2(dba)3 and lithium chloride in DMF (Figure 19).

Figure 19.

Stille macrocyclisation in the total synthesis of biselyngbyolide A.

The utility of Sonogashira macrocyclisation in the first total synthesis of penarolide sulfate A1, an α-glucosidase inhibitor is demonstrated by Mohapatra and co-workers [47]. The macrocyclisation was successfully achieved from compound 28 with catalytic Pd(PPh3)4 and CuI in Et2NH at room temperature (Figure 20).

Figure 20.

Sonogashira macrocyclisation in the synthesis of penarolide sulfate A1.

Towards the total synthesis of antibiotic natural product A26771B, Trost and co-workers [48] effectively constructed the macrolactone 31 (Figure 21) by the use of bidentate phosphine ligand (1,4-bis(diphenylphosphino)butane (DPPB)).

Figure 21.

Tsuji-Trost lactonization for the synthesis of A26771B.

In the end game of total synthesis of macrolides, glycosidation to the aglycon also have more significance. Thus, a wide variety of methods are reported for glycosidation in the literature [49, 50].

Advertisement

3. Total synthesis of selected macrolides

In this section, the total synthesis of selected macrolides is discussed: (+)-neopeltolide (32), aspergillide D (33) and briefly about miyakolide (34) and acutiphycin (35).

3.1 (+)-Neopeltolide

(+)-Neopeltolide is a 14-membered macrolide isolated from north coast of Jamaica by Wright and coworkers from a deep water sponge [51]. It was tested for in vitro antiproliferative activity against several cancer cell lines comprising A549 human lung adenocarcinoma, NCI/ADR-RES ovarian sarcoma and P388 murine leukemia and shows IC50 values 1.2, 5.1 and 0.56 nM. Besides this, neopeltolide also exhibits anti-fungal activity against Candida albicans [52]. The complexity of the structure with six chiral centres, tetrahydropyran ring, and an oxazole-bearing unsaturated side chain and its efficacious biological activity led to several total syntheses, few of them are discussed below.

In 2013, Ghosh et al. [53] in their total synthesis adopted the retrosynthetic pathway as follows. Disconnection of O▬C bond of the oxazole side chain would give acid which can undergo Mitsunobu esterification. Yamaguchi macrolactonization of acid would in turn give the desired macrolactone. The tetrahydropyran ring in acid could be constructed via a hetero Diels-Alder reaction between aldehyde and silyloxy diene ether using Jacobsen’s chromium catalyst.

The synthesis of the macrolactone ring of (+)-neopeltolide began with commercially available 3-methyl glutaric anhydride as shown in the scheme. 3-methyl glutaric anhydride, 36 was desymmetrized using PS-30 ‘Amano’ lipase to obtain acid. The resulting acid was treated with borane-dimethyl sulfide complex to afford alcohol, 37. Alcohol 37 was oxidized to corresponding aldehyde by Swern oxidation and then protected to its acetal, 38. Ester of 38 was then reduced to alcohol and on Swern oxidation obtained aldehyde and the resulting aldehyde was subjected to Brown’s allylation protocol using (+)-Ipc2BOMe and allyl magnesium bromide to attain alcohol, 39. Alcohol 39 was methylated with MeI, and on Lemieux-Johnson oxidation gave aldehyde and on Brown’s allylation protocol afforded alcohol, 40. Acetal protection was deprotected and the aldehyde was converted to α,β-unsaturated ketone, 41 using standard Horner-Wadsworth-Emmons olefination conditions. Secondary alcohol in 41 was then protected with TESOTf to obtain the silyloxy diene, 42 in excellent yield (Figure 22).

Figure 22.

Synthesis of silyloxy diene 42 fragment.

After the completion of requisite silyloxy diene, hetero-Diels Alder reaction of tosyl oxyacetaldehyde, 43 with 42 using chiral chromium catalyst (44) gave tetrahydropyranone, 45 in 83% yield (Figure 23). After protection of ketone group in 45 as ketal and displaced the tosylate to nitrile 46 using NaCN in DMF. Nitrile 46 was hydrolysed to acid and on deprotection of ketone to afford ketone 47. Intramolecular Yamaguchi macrolactonization attained the key macrolactone 48 in 40% yield. Olefin 48 was subjected to hydrogenation with 10% Pd/C to give saturated compound and on reduction with NaBH4/EtOH to attain alcohol 49. Next, the synthesis of unsaturated oxazole side chain 50 is started with known alkyne with LDA and Bu3SnCl to obtain the alkynyl stannate, which on hydrozirconation gave the carbamate in 38% yield. Crucial Stille cross coupling of carbamate with iodooxazole using Pd(MeCN)2Cl2 in DMF gave oxazole which can be easily converted to desired side chain 50. The target neopeltolide compound was furnished by standard Mitsunobu esterification of 49 with acid 50 (Figure 24).

Figure 23.

Hetero-Diels-Alder reaction for the synthesis of tetrahydropyrarone moiety.

Figure 24.

Completion of synthesis of (+)-neopeltolide.

3.2 Paterson strategy

Paterson and coworkers [54] reported the synthesis of neopeltolide as follows. Aldehyde was synthesized starting from known β-keto ester, 51, which on treatment with (S)-BINAP-Ru (II) catalyst under Noyori asymmetric hydrogenation to afford (13S)-alcohol. The alcohol on TBS protection and DIBAL reduction of the ester produced the enantiopure aldehyde 52. Aldehyde 52 was subjected to Brown’s methallylation using 2-methyl propene and (−)-Ipc2BOMe furnished the desired C11 alcohol with 94:6 dr, and the alcohol was methylated into the methyl ether 53 by NaH, MeI. Methyl ether 53 was subjected to ozonolysis to obtain methyl ketone and on Horner Wadsworth Emmons reaction with trimethyl phosphonoacetate to attain ester 54 in E/Z isomers (75,25). Esters 54 were reduced to its alcohol by DIBAL-H and on subsequent oxidation with Dess-Martin periodinane produced aldehyde 55. Next, organo catalytic hydride reduction of enal 55 using MacMillan strategy with imidzolidinone catalyst 56. TFA (20 mol%) and Hantzch ester furnished as 1,4-reduction product 57 with 76:24 of epimers at C9 stereocentre (Figure 25). Further, Jacobsen asymmetric hetero Diels-Alder reaction between 57 and known 2-silyloxy diene 58 produced cis-tetrahydropyranone 60 in 60% yield using chiral tridentate chromium (III) catalyst 59. On PMB deprotection and further oxidation of alcohol to corresponding acid followed by TBS deprotection furnished seco-acid 61. Macrolactonization of 61 under standard Yamaguchi conditions afforded macrolactone 62 in 80% yield. Reduction of macrolactone 62 to the alcohol with NaBH4 in MeOH followed by Mitsunobu esterification with the oxazole side chain 50 achieved (+)-neopeltolide (32) in 52% yield (Figure 26).

Figure 25.

Synthesis of aldehyde 57.

Figure 26.

Total synthesis of (+)-neopeltolide.

3.3 Ulanovskaya strategy

The synthesis of neopeltolide by Ulanovskaya et al. [55] is depicted as follows, Prins desymmetrization of diene 63 followed by benzyl protection and Wacker oxidation of alkene afforded ketone 64. Formation of boron enolate from ketone and on addition of aldehyde 65 gave the anticipated aldol product with >98:2 diastereoselectivity, which was treated with Ph3P=CH2 (Wittig methylenation) followed by cleavage of dioxolone by acidic work-up afforded ketone 66 in 75% yield. Ketone 66 was selectively reduced to syn-alcohol using Et2BOMe and NaBH4 followed by ester hydrolysis gave acid which was subjected to Yamaguchi macrolactonization to furnish desired macrolactone 67. Alkene in 67 was hydrogenated using Pd/C to afford desired alcohol 68 as a major product. Alcohol 68 was subjected to Mitsunobu conditions, followed by hydrolysis with K2CO3 in MeOH to get the inversion product. Subsequent methylation with MeO3BF4 & hydrogenolysis of benzyl ether achieved desired macrolide 49. The final coupling of fragment 49 with oxazole side chain 50 with standard Mitsunobu conditions furnished target (+)-neopeltolide 32 (Figure 27).

Figure 27.

Alternate synthesis of (+)-neopeltolide.

3.4 Aspergillide-D

Bao and coworkers in 2013 isolated 16-membered macrolide, aspergillide D, from the extract of Aspergillus sp. SCSGAF 0076 [56]. Aspergillide D macrolactone contains four chiral centres, α,β-unsaturation, three hydroxyl groups and the first total synthesis was reported by Jena et al. in 2017 as follows [57].

The retrosynthetic analysis of aspergillide D was depicted as shown above, macrolactone could be synthesized from seco acid via intramolecular Shiina esterification. For the total synthesis of Aspergillide D, the acid fragment was synthesized from commercially available D-ribose which was transformed to lactol 69 by using three step sequence, that is, catalytic amount of H2SO4 & acetone to form acetonide which on reduction with NaBH4 and on oxidative cleavage of the diol with NaIO4. The lactol was subjected to Wittig type olefination using PPh3=CH2 and the obtained primary alcohol 70 was oxidized to carboxylic acid 71 by using TEMPO/BAIB conditions (Figure 28). The synthesis of alcohol fragment was started with mono-PMB protection 73 of commercially available 1,8-octane diol 72 and the other alcohol was converted to racemic allyl alcohol 74 by Swern oxidation and subsequent treatment of aldehyde with vinyl magnesium bromide in the presence of CuI. The allylic alcohol 74 was subjected to standard Sharpless kinetic resolution conditions by using (−)-DIPT & Ti(OiPr)4 to obtain enantiomeric epoxy alcohol 75. Upon MOM protection 76 to the secondary alcohol 75 and PMB deprotection produced 77, which on oxidation with Dess-Martin periodinane to afford aldehyde. Aldehyde was converted to olefin 78 by treating PPh3=CH2 in THF. 78 was cleaved to alcohol 79 by reduction with LAH in THF (Figure 29).

Figure 28.

Synthesis of acid fragment 71.

Figure 29.

Synthesis of alcohol fragment 79.

Acid 71 and alcohol 79 fragments were coupled together under Yamaguchi esterification conditions afforded diene ester 80 in 65% yield. Intramolecular RCM was employed on diene ester by using Grubbs’ second generation catalyst in refluxing CH2Cl2 to produce the requisite macrolactone 81. Double bond in 81 was hydrogenated by using PtO2 in MeOH to attain saturated lactone 82. Lactone 82 was reduced with DIBAL-H to afford lactol which on further treatment with Ph3P=CHCO2Et in C6H6 afforded α,β-unsaturated ester 83. The ester was converted to carboxylic acid 84 by LiOH in THF/H2O which on adopting key Shiina’s macrolactonization protocol to provide the desired mactrolactone 85 in 51% yield. On deprotection of acetonide with CuCl2.2H2O gave diol 86 and removal of MOM group, the synthesis of aspergillide D 33 was achieved (Figure 30).

Figure 30.

Completion of synthesis of aspergillide D.

3.5 Miyakolide

Evan’s strategy of bond connections & key reactions in the synthesis of 34 is illustrated [58].

3.6 Acutiphycin

Smith’s strategy [59] & Moslin’s strategy of acutiphycin [60] is shown below (Figure 31).

Figure 31.

Key reactions and strategies in the synthesis of miyakolide and acutiphycin.

Advertisement

4. Conclusions

A number of new macrolide antibiotics with fascinating biological activities have been isolated everyday with the unique and complex structures have been determined with extensive spectroscopic studies. Toward the total synthesis of such macrolide antibiotics, very efficient synthetic strategies and various new methodologies are also developed. Recent advances in macrolide synthesis based on newly developed strategies and methodologies are noteworthy. Further synthetic studies on macrolide antibiotics will make an immense contribution to progress in both organic and medicinal chemistry.

Advertisement

Acknowledgments

The author wishes to thank Vice Chancellor, Dean R & D, VFSTRU for constant support and encouragement. The author wishes to express his gratitude to Prof. V. Anuradha and other staff members of chemistry department, Vignan for the fruitful discussion. IntechOpen OAPF is greatly acknowledged.

References

  1. 1. Cowden CJ, Paterson I. Asymmetric Aldol reactions using boron enolates. Organic Reactions. 1997;51:1-200. DOI: 10.1002/0471264180.or051.01
  2. 2. Kim BM, Williams SF, Masamune S. The Aldol reaction: Group III enolates. In: Trost BM, Fleming I, editors. Comprehensive Organic Synthesis. Oxford, UK: Pergamon Press; 1991. pp. 301-320
  3. 3. Paterson I, Cowden CJ, Wallace DJ. Stereoselective Aldol reactions in the synthesis of polyketide natural products. In: Otera J, editor. Modern Carbonyl Chemistry. Weinheim, Germany: Wiley-VCH; 2000. pp. 249-297
  4. 4. Evans DA, Nelson JV, Vogel E, Taber TR. Stereoselective Aldol condensations via boron enolates. Journal of the American Chemical Society. 1981;103:3099-3111. DOI: 10.1021/ja00401a031
  5. 5. Gage JR, Evans DA. Diastereoselective Aldol condensation using a chiral oxazolidinone auxiliary: (2S*3S*)-3-hydroxy-3-phenyl-2-methylpropanoic acid. Organic Syntheses. 1990;68:83-91. DOI: 10.15227/orgsyn.068.0083
  6. 6. Walker MA, Heathcock CH. Extending the scope of the Evans asymmetric aldol reaction: Preparation of anti and “Non-Evans” Syn aldols. Journal of Organic Chemistry. 1991;56:5747-5750. DOI: 10.1021/jo00020a006
  7. 7. Katsuki T, Sharpless KB. The first practical method for asymmetric epoxidation. Journal of the American Chemical Society. 1980;102:5974-5976. DOI: 10.1021/ja00538a077
  8. 8. Hentges SG, Sharpless KB. Asymmetric induction in the reaction of osmium tetroxide with olefins. Journal of the American Chemical Society. 1980;102:4263-4265. DOI: 10.1021/ja00214a053
  9. 9. Caron M, Sharpless KB. Titanium isopropoxide-mediated nucleophilic openings of 2,3-epoxy alcohols. A mild procedure for regioselective ring-opening. The Journal of Organic Chemistry. 1985;50:1557-1560. DOI: 10.1021/jo00209a047
  10. 10. Crane EA, Scheidt KA. Prins type macrocyclizations as an efficient ring-closing strategy in natural product synthesis. Angewandte Chemie, International Edition. 2010;49:8316-8326. DOI: 10.1002/anie.201002809
  11. 11. Parenty A, Moreau X, Campagne J-M. Macrolactonizations in the total synthesis of natural products. Chemical Reviews. 2006;106:911-939. DOI: 10.1021/cr0301402
  12. 12. Parenty A, Moreau X, Niel G, Campagne J-M. Update 1 of: Macrolactonizations in the total synthesis of natural products. Chemical Reviews. 2013;113:PR1-PR40. DOI: 10.1021/cr300129n
  13. 13. Gradillas A, Perez-Castells J. Macrocyclization by ring-closing metathesis in the total synthesis of natural products: Reaction conditions and limitations. Angewandte Chemie International Edition. 2006;45:6086-6101. DOI: 10.1002/anie.200600641
  14. 14. Nicolaou KC, Harter MW, Gunzner JL, Nadin A. The Wittig and related reactions in natural product synthesis. Liebigs Annalen. 1997;1997(7):1283-1301. DOI: 10.1002/jlac.199719970704
  15. 15. Bisceglia JA, Orelli LR. Recent applications of the Horner-Wadsworth-Emmons reaction to the synthesis of natural products. Current Organic Chemistry. 2012;16:2206-2230. DOI: 10.2174/138527212803520227
  16. 16. Chatterjee B, Bera S, Mondal D. Julia-Kocienski olefination: A key reaction for the synthesis of macrolides. Tetrahedron: Asymmetry. 2014;25:1-55. DOI: 10.1016/j.tetasy.2013.09.027
  17. 17. Ronson TO, Taylor RJK, Fairlamb IJS. Palladium-catalysed macrocyclisations in the total synthesis of natural products. Tetrahedron. 2015;71:989-1009. DOI: 10.1016/j.tet.2014.11.009
  18. 18. Corey EJ, Nicolaou KC. Efficient and mild lactonization method for the synthesis of macrolides. Journal of the American Chemical Society. 1974;96:5614-5616. DOI: 10.1021/ja00824a073
  19. 19. Shiina I, Kubota M, Ibuka R. A novel and efficient macrolactonization of ω-hydroxycarboxylic acids using 2-methyl-6-nitrobenzoic anhydride (MNBA). Tetrahedron Letters. 2002;43:7535-7539. DOI: 10.1016/S0040-4039(02)01819-1
  20. 20. Inanaga J, Hirata K, Saeki H, Katsuki T, Yamaguchi M. A rapid esterification by means of mixed anhydride and its application to large-ring lactonization. Bulletin of the Chemical Society of Japan. 1979;52:1989-1993. DOI: 10.1246/bcsj.52.1989
  21. 21. Kurihara T, Nakajima Y, Mitsunobu O. Synthesis of lactones and cycloalkanes. Cyclization of ω-hydroxy acids and ethyl α-cyano-ω-hydroxycarboxylates. Tetrahedron Letters. 1976;17:2455-2458. DOI: 10.1016/0040-4039(76)90018-6
  22. 22. Boden EP, Keck GE. Proton-transfer steps in Steglich esterification: A very practical new method for macrolactonization. The Journal of Organic Chemistry. 1985;50:2394-2395. DOI: 10.1021/jo00213a044
  23. 23. Mukaiyama T, Usui M, Saigo K. The facile synthesis of lactones. Chemistry Letters. 1976;5:49-50. DOI: 10.1246/cl.1976.49
  24. 24. Hansen TV, Stenstrom Y. First total synthesis of (−)-aplyolide A. Tetrahedron: Asymmetry. 2001;12:1407-1409. DOI: 10.1016/S0957-4166(01)00250-6
  25. 25. Narasaka K, Maruyama K, Mukaiyama T. A useful method for the synthesis of macrocyclic lactone. Chemistry Letters. 1978;7:885-888. DOI: 10.1246/cl.1978.885
  26. 26. Enev VS, Kaehlig H, Mulzer J. Macrocyclization via allyl transfer: Total synthesis of laulimalide. Journal of the American Chemical Society. 2001;123:10764-10765. DOI: 10.1021/ja016752q
  27. 27. Keck GE, Boden EP, Wiley MR. Total synthesis of (+)-colletodial: New methodology for the synthesis of macrolactones. The Journal of Organic Chemistry. 1989;54:896-906. DOI: 10.1021/jo00265a033
  28. 28. Williams DR, Meyer KG. Total synthesis of (+)-amphidinolide K. Journal of the American Chemical Society. 2001;123:765-766. DOI: 10.1021/ja005644l
  29. 29. Schweitzer D, Kane JJ, Strand D, McHenry P, Tenniswood M, Helquist P. Total synthesis of iejimalide B. An application of the Shiina macrolactonization. Organic Letters. 2007;9:4619-4622. DOI: 10.1021/ol702129w
  30. 30. Majumdar KC, Rahaman H, Roy B. Synthesis of macrocyclic compounds by ring closing metathesis. Current Organic Chemistry. 2007;11:1339-1365. DOI: 10.2174/138527207782023166
  31. 31. Grubbs RH, Miller SJ, Fu GC. Ring-closing metathesis and related processes in organic synthesis. Accounts of Chemical Research. 1995;28:446-452. DOI: 10.1021/ar00059a002
  32. 32. Wu Y, Esser L, De Brabander JK. Revision of the absolute configuration of salicylihalamide A through asymmetric total synthesis. Angewandte Chemie, International Edition. 2000;39:4308-4310. DOI: 10.1002/1521-3773(20001201)39:23%3C4308::aid-anie4308%3E3.0.CO;2-4
  33. 33. Garbaccio RM, Stachel SJ, Baeschlin DK, Danishefsky SJ. Concise asymmetric synthesis of radicicol and monocillin I. Journal of the American Chemical Society. 2001;123:10903-10908. DOI: 10.1021/ja011364+
  34. 34. Furstner A, Seidel G, Kindler N. Macrocycles by ring-closing metathesis, XI: Syntheses of (R)-(+)-lasiodiplodin, zeranol and truncated salicylihalamides. Tetrahedron. 1999;55:8215-8230. DOI: 10.1016/S0040-4020(99)00302-6
  35. 35. Wang X, Bowman EJ, Bowman BJ, Porco JA. Total synthesis of the salicylate enamide macrolide oximidine III: Application of relay ring-closing metathesis. Angewandte Chemie International Edition. 2004;43:3601-3605. DOI: 10.1002/anie.200460042
  36. 36. Xu Z, Johannes CW, Salman SS, Hoveyda AH. Enantioselective total synthesis of antifungal agent Sch 38516. Journal of the American Chemical Society. 1996;118:10926-10927. DOI: 10.1021/ja9626603
  37. 37. Miyaura N, Suzuki A. Stereoselective synthesis of arylated (E)-alkenes by the reaction of alk-1-enylboranes with aryl halides in the presence of palladium catalyst. Journal of the Chemical Society, Chemical Communications. 1979;1979(19):866-867. DOI: 10.1039/C39790000866
  38. 38. Heck RF, Nolley JP. Palladium-catalyzed vinylic hydrogen substitution reactions with aryl, benzyl, and styryl halides. The Journal of Organic Chemistry. 1972;37:2320-2322. DOI: 10.1021/jo00979a024
  39. 39. Heravi MM, Mohammadkhani L. Recent applications of Stille reaction in total synthesis of natural products: An update. Journal of Organometallic Chemistry. 2018;869:106-200. DOI: 10.1016/j.jorganchem.2018.05.018
  40. 40. Milstein D, Stille JK. Palladium-catalyzed coupling of tetraorganotin compounds with aryl and benzyl halides. Synthetic utility and mechanism. Journal of the American Chemical Society. 1979;101:4992-4998. DOI: 10.1021/ja00511a032
  41. 41. Sonogashira K, Tohda Y, Hagihara N. A convenient synthesis of acetylenes: Catalytic substitutions of acetylenic hydrogen with bromoalkenes, iodoarenes and bromopyridines. Tetrahedron Letters. 1975;16:4467-4470. DOI: 10.1016/S0040-4039(00)91094-3
  42. 42. Trost BM, Fullerton TJ. New synthetic reactions. Allylic alkylation. Journal of the American Chemical Society. 1973;95:292-294. DOI: 10.1021/ja00782a080
  43. 43. Trost BM. Cyclizations via palladium-catalyzed allylic alkylations [new synthetic methods (79)]. Angewandte Chemie International Edition in English. 1989;28:1173-1192. DOI: 10.1002/anie.198911731
  44. 44. Tortosa M, Yakelis NA, Roush WR. Total synthesis of (+)-superstolide A. Journal of the American Chemical Society. 2008;130:2722-2723. DOI: 10.1021/ja710238h
  45. 45. Ziegler FE, Chakraborty UR, Weisenfeld RB. A palladium-catalyzed carbon-carbon bond formation of conjugated dienones: A macrocyclic dienone lactone model for the carbomycins. Tetrahedron. 1981;37:4035-4040. DOI: 10.1016/S0040-4020(01)93278-8
  46. 46. Tanabe Y, Sato E, Nakajima N, Ohkubo A, Ohno O, Suenaga K. Total synthesis of biselyngbyolide A. Organic Letters. 2014;16:2858-2861. DOI: 10.1021/ol500996n
  47. 47. Mohapatra DK, Bhattasali D, Gurjar MK, Khan MI, Shashidhara KS. First asymmetric total synthesis of penarolide sulfate A1. European Journal of Organic Chemistry. 2008;2008(36):6213-6224. DOI: 10.1002/ejoc.200800680
  48. 48. Trost BM, Brickner SJ. Palladium-assisted macrocyclization approach to cytochalasins: A synthesis of antibiotic A26771B. Journal of the American Chemical Society. 1983;105:568-575. DOI: 10.1021/ja00341a043
  49. 49. Toshima K, Tatsuta K. Recent progress in O-glycosylation methods and its application to natural products synthesis. Chemical Reviews. 1993;93:1503-1531. DOI: 10.1021/cr00020a006
  50. 50. Danishefsky SJ, Bilodeau MT. Glycals in organic synthesis: The evolution of comprehensive strategies for the assembly of oligosaccharides and glycoconjugates of biological consequence. Angewandte Chemie International Edition in English. 1996;35:1380-1419. DOI: 10.1002/anie.199613801
  51. 51. Wright AE, Botelho JC, Guzman E, Harmody D, Linley P, McCarthy PJ, et al. Neopeltolide, a macrolide from a lithistid sponge of the family Neopeltidae. Journal of Natural Products. 2007;70:412-416. DOI: 10.1021/np060597h
  52. 52. Altmann KH, Carreira EM. Unraveling a molecular target of macrolides. Nature Chemical Biology. 2008;4:388-389. DOI: 10.1038/nchembio0708-388
  53. 53. Ghosh AK, Shurrush KA, Dawson ZL. Enantioselective total synthesis of macrolide (+)-neopeltolide. Organic & Biomolecular Chemistry. 2013;11:7768-7777. DOI: 10.1039/C3OB41541D
  54. 54. Paterson I, Miller NA. Total synthesis of the marine macrolide (+)-neopeltolide. Chemical Communications. 2008;2008(39):4708-4710. DOI: 10.1039/B812914B
  55. 55. Ulanovskaya OA, Janjic J, Suzuki M, Sabharwal SS, Schumacker PT, Kron SJ, et al. Synthesis enables identification of the cellular target of leucascandrolide A and neopeltolide. Nature Chemical Biology. 2008;4:418-424. DOI: 10.1038/nchembio.94
  56. 56. Bao J, Xu XY, Zhang XY, Qi SH. A new macrolide from a marine-derived fungus Aspergillus sp. Natural Product Communications. 2013;8:1127-1128. DOI: 10.1177%2F1934578X1300800825
  57. 57. Jena BK, Reddy GS, Mohapatra DK. First asymmetric total synthesis of aspergillide D. Organic & Biomolecular Chemistry. 2017;15:1863-1871. DOI: 10.1039/c6ob02435a
  58. 58. Evans DA, Ripin DHB, Halstead DP, Campos KR. Synthesis and absolute assignment of (+)-miyakolide. Journal of the American Chemical Society. 1999;121:6816-6826. DOI: 10.1021/ja990789h
  59. 59. Smith AB III, Chen SS-Y, Nelson FC, Reichert JM, Salvatore BA. Total syntheses of (+)-acutiphycin and (+)-trans-20,21-didehydroacutiphycin. Journal of the American Chemical Society. 1997;119:10935-10946. DOI: 10.1021/ja972497r
  60. 60. Moslin RM, Jamison TF. Highly convergent total synthesis of (+)-acutiphycin. Journal of the American Chemical Society. 2006;128:15106-15107. DOI: 10.1021/ja0670660

Written By

Chebolu Naga Sesha Sai Pavan Kumar

Submitted: 29 March 2019 Reviewed: 07 June 2019 Published: 22 July 2019