Next Article in Journal
Hypoxia-Inducible Factor 1-Alpha and Glucose Metabolism during Cardiac Remodeling Progression from Hypertrophy to Heart Failure
Next Article in Special Issue
Immunomodulatory Effects of the Pea Defensin Psd1 in the Caco-2/Immune Cells Co-Culture upon Candida albicans Infection
Previous Article in Journal
Solid Lipid Nanoparticles: Applications and Prospects in Cancer Treatment
Previous Article in Special Issue
Influence of Lipidation Pattern of the KR12 Fragment of Peptide LL-37 on Its Antibacterial and Hemolytic Activities
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Epoxide-Based Synthetic Approaches toward Polypropionates and Related Bioactive Natural Products

by
Raúl R. Rodríguez-Berríos
1,
Stephen R. Isbel
2 and
Alejandro Bugarin
2,*
1
Department of Chemistry, University of Puerto Rico, Rio Piedras Campus, P.O. Box 23346, San Juan 00931-3346, Puerto Rico
2
Department of Chemistry & Physics, Florida Gulf Coast University, 10501 FGCU Boulevard South, Fort Myers, FL 33965, USA
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(7), 6195; https://doi.org/10.3390/ijms24076195
Submission received: 28 February 2023 / Revised: 16 March 2023 / Accepted: 23 March 2023 / Published: 24 March 2023
(This article belongs to the Special Issue Antimicrobial and Antiviral Drugs Discovery)

Abstract

:
Polypropionate units are a common structural feature of many of the natural products in polyketides, some of which have shown a broad range of antimicrobial and therapeutic potential. Polypropionates are composed of a carbon skeleton with alternating methyl and hydroxy groups with a specific configuration. Different approaches have been developed for the synthesis of polypropionates and herein we include, for the first time, all of the epoxide-based methodologies that have been reported over the years by several research groups such as Kishi, Katsuki, Marashall, Miyashita, Prieto, Sarabia, Jung, McDonald, etc. Several syntheses of polypropionate fragments and natural products that employed epoxides as key intermediates have been described and summarized in this review. These synthetic approaches involve enatio- and diastereoselective synthesis of epoxides (epoxy-alcohols, epoxy-amides, and epoxy-esters) and their regioselective cleavage with carbon and/or hydride nucleophiles. In addition, we included a description of the isolation and biological activities of the polypropionates and related natural products that have been synthetized using epoxide-based approaches. In conclusion, the epoxide-based methodologies are a non-aldol alternative approach for the construction of polypropionate.

1. Introduction

All living organisms have the capacity to transform and interconvert a variety of molecules using two types of metabolic pathways, primary and secondary metabolism, with this review focusing primarily on the latter. In general, the majority of organisms produce different kinds of secondary metabolites in a limited distribution and play an important role in the well-being of the producer [1]. Polyketides are secondary metabolites derived from the acetate pathway and are an important large family of natural products biosynthesized by animals, bacteria, fungi, and plants. This process occurs through the Claisen condensation of C3-units catalyzed by the enzyme polyketide synthases (PKSs) [2]. Polyketides include groups of fatty acids, aromatic compounds (e.g., anthraquinones, tetracyclines), and polypropionates (Figure 1). Polypropionates are further classified into subgroups including macrolide antibiotics, linear polyketides, and polyethers [3]. In fact, more than 10,000 polyketides have been reported and about 1% of them possess drug activity (typically antibiotic, anticancer, antifungal, antiparasitic, immunomodulatory action, and cholesterol lowering agents), which is approximately five times more bioactive than the average of other natural product families [3,4]. Around 20% of the top-selling, small-molecule drugs are polyketides [5].
The propionate unit consists of an aliphatic chain with alternating methyl and hydroxy groups with a specific configuration at each carbon (Figure 2). For decades, the construction of polypropionate chains has attracted the great interest of synthetic organic chemists due to the challenge represented by the elaboration of their array of contiguous asymmetric centers. As the number of stereocenters increases, the difficulty of synthesizing these compounds also increases. In 1987, the term “stereo-n-ad” was introduced by Hoffman, in which n refers to the number of asymmetric carbons within the propionate chain [6]. For example, three and five chiral centers represent a stereotriad and stereopentad, respectively.
Aldol type reactions represent one of the primarily used methods for the regio-, stereo-, and enantioselective formation of propionates that has been developed [7,8,9]. Furthermore, other methods for the stereoselective assembly of long sequences of stereocenters in polypropionates have been developed. These methods include crotylations [10,11,12], allenylations [13,14,15], selective radical processes [16,17,18,19], sequential substitutions [20,21], epoxide ring openings [22,23,24,25], Diels–Alder [26], and others [27,28,29,30] (Figure 3).
In this review, we describe the structures, classification, and antimicrobial activities of approximately thirty polypropionate natural products that have been synthesized using epoxide-based approaches by different research groups from 1980 to 2020. Some research groups reported their epoxide-based methodologies for a partial and/or total synthesis of polypropionates. Therefore, this review describes and depicts the general epoxide-based approach of each group including key steps to construct the polypropionate units, an example in the synthesis of a polypropionate fragment, and all the polypropionate fragments and/or natural products reported by each group.

1.1. Isolation, Structural Characterization, and Antimicrobial Activities of Polypropionates

1.1.1. Macrolides

Macrolides are a large family of polypropionates in which the majority exhibit antibiotic activity, and their structure is characterized by a range of 12-, 14-, 16-, 18-, or 22-membered macrocyclic lactone rings. For example, methymycin 1 (Figure 4) is a 12-membered macrolide antibiotic biosynthesized by Streptomyces venezuelae [31]. Prelog–Djerassi 2 (Figure 4) is a 6-membered polypropionate lactone that was obtained as a degradation product of 1 and retains the configuration of the four chiral centers corresponding to the C1-C7 polypropionate fragment. Erythromycin A 3a (Figure 4) is a 14-membered macrolide antibiotic produced by Saccharolyspora erythraea and was isolated in 1952 [1]. Erythromycins show activity against Gram-positive bacteria, penicillin resistant Staphilococus strains, and are used to treat infections of Legionella pneumophilia. The mechanism of action consists in the inhibition of the protein biosynthesis in the organism. The structure was elucidated by spectroscopic methods and confirmed by X-ray crystallography [32]. Erythromycin A 3a attaches at the C3 and C5 two sugars, d-desoamine and l-cladinose, respectively, and the aglycone structure. It is worth noting that erythronolide A 3b does not have the corresponding glycoside linkages.
Lankamycin 5a (Figure 4) was collected from a soil sample and isolated from various Streptomyces species: first in 1960 by Gäumann et al. from Streptomyces violaceoniger [33] and later in 1969 by Namiki et al. from the Kuji District of Ibari Prefecture, Japan from S. spinichromogenes var. kujimyceticus [34]. In addition, Kinashi et al. reported that Streptomyces rochei 7434AN4 produces lankamycin 5a and other structurally unrelated polyketide antibiotics [35]. Lankamycin 5a shows moderate inhibitory activities against Mycobacteria [34] and Gram-positive bacteria such as Staphylococcus aureus, Bacillus subtilis, and Micrococcus luteus [36,37]. The relative structure of the neutral 14-membered macrolide antibiotic lankamycin 4a was determined by Keller et al. [38]. The molecule consists of a lactone ring, 11-acetyllankolide, which contains a polypropionate fragment with twelve stereogenic centers. The structure of lankamycin 4a has two sugar moieties, in which d-chalcose is bound at the C5 and the 4-O-acetyl-l-arcanose is bound at C3 [39]. Lankanolide 4b (Figure 4) is the aglycon of 4a. Elaiophylin 5 (Figure 4) is a 16-membered macrodiolide isolated from different cultures of Streptomyces malonosporus and belongs to a group of C2-symmetrical macrodiolides [40]. The elaiophylin 5 structure was determined by extensive chemical degradation and nuclear magnetic resonance (NMR) studies, while X-ray crystallographic analysis revealed the relative and absolute stereochemistry [41]. The structure of 5 consists of a macrolactone, an intramolecular hemiacetal, and a polypropionate fragment with nine asymmetric carbons. Macrolide 5 exhibited potent antimicrobial activity against several strains of Gram-positive bacteria including Bacillus subtilis, Staphylococcus aureus, and Corynebacterium diphtheriae. It also revealed anthelmintic activity against Trichonomonas vaginalis [42,43].
Bafilomycin A1 6 (Figure 4) was isolated in 1983 by Wagner from a broth of Streptomyces griseus [44]. Its relative stereochemistry was determined by Corey and Ponder in 1984 using extensive NMR analysis, and confirmed in 1987 using X-ray crystallography [45,46]. The structure of 6 represents a 16-membered macrolide with a tetraene core, a cyclic hemiacetal, a C14 methoxy, and a C23-isopropyl group within its polypropionate chain. Bafilomycin A1 6 exhibited potent antifungal and antibacterial activity and is also the first known inhibitor of Vacuolar H+-ATPase (VATPase) [47]. Protomycinolide IV (7) (Figure 4) is the aglycone of Mycinamicins [48]. It is a 16-membered macrolide isolated from Micromonospora griseorubida sp. nov. that possesses excellent activity against Gram-positive bacteria [49]. The protomycinolide IV (7) structure is characterized by a short polypropionate chain that contains four chiral carbons of the six in total, a conjugate ketone, and macrocyclic lactone.
The macrocyclic bislactone, lepranthin 8 (Figure 4), was isolated from a species of crustaceous lichen Arthonia impolita (Ehrh.) Borrer by Zopf, approximately 119 years ago (in 1904). The structure was elucidated using NMR and X-ray crystallographic analysis in 1995 by Huneker et al. [50]. Lepranthin 8 consists of a 16-membered homo-macrodiolide and the polypropionate chain contains 12 stereogenic centers in which there are four secondary acetates. Biological activity studies for Lepranthin 8 have not been reported in the literature. Tedanolide 9 (Figure 4) is a very potent cytotoxic macrolide that was isolated from Tedania ignis, a very well-known Caribbean marine sponge [51,52]. The 18-membered macrolactone 9 was elucidated using high resolution fast atom bombardment/mass spectrometry (FAB/MS), infrared (IR), NMR spectroscopy, and X-ray diffraction. Tedanolide 9 exhibited good cytotoxicity against the KB cell line (ED50 = 2.5 × 10−4 ng/mL) and PS cell line (ED50 = 1.6 × 10−5 ng/mL) [51]. Furthermore, it has demonstrated strong antitumoral activities since it has been found to increase the life span of mice implanted with lymphocytic leukemia by 23% [53]. Venturicidins A (10a) and B (10b) (Figure 4) are 20-membered macrolide antibiotics that were isolated from Streptomyees aureofaciens, exhibit a potent antifungal activity [54], and are specific inhibitors of mitochondrial ATPase [55]. The structures of venturicindins 10 were characterized by NMR, MS, and X-ray analysis. Specifically, venturicidin X (10c) (Figure 4) is the aglycone, and the R group of venturicidin A 10a and B 10b are 3-O-carbamyl-2-deoxy-D-rhamnoside and 2-deoxy-D-rhamnoside, respectively. These macrolactones contain a tetrahydropyran ring, ten asymmetric carbons in total, in which seven chiral carbons belong to the polypropionate chain.
Scytophycin C 11 (Figure 4) is a 22-membered macrolide isolated in 1986 by Moore et al. from the terrestrial blue-green alga Scytonema pseudohofmanni [56]. The structure of 11 contains 15 stereogenic centers, a dihydropyran ring bearing two trans-substituted side chains, a polypropionate chain, and a terminal N-methyl-N-vinylformamide moiety. Scytophycin C 11 exhibited potent activity against a variety of human carcinoma cell lines including solid tumors [57].
Lobophorolide 12 (Figure 4) was isolated from the Caribbean brown algae Lobophora variegata by Kubanek et al. in 2003 [58]. They characterized the lobophorolide 12 structure using IR, NMR, and MS analysis, which consists of a 22-membered macrolactone that has attached an aliphatic side chain with a tetrahydropyran ring-termini. Lobophorolide 12 displayed antifungal activity against D. salina and L. thalassiae as well as a high cytotoxicity toward the colon tumor cell line HCT-116 (IC50 = 0.03 μg/mL) [59].
The isolation of mycalolides was reported in 1989 by Fusetani et al. from a sponge of the genus Mycale [60]. Mycalolide A 13 (Figure 4) has a 25-membered macrolide ring that incorporates the tris-oxazole, attached to an aliphatic polyketide side chain that contains eight of the eleven stereogenic centers present in this compound [61]. The relative and absolute configuration were determined through a combination of chemical degradation, extensive NMR analysis, and correlation experiments [62]. Mycalolide A 13 exhibits potent antifungal activity against a wide array of pathogenic fungi and cytotoxicity toward B-16 melanoma cells (IC50 = 0.5–1.0 ng/mL). It was also shown that it selectively inhibits the actomyosin Mg2+-ATPase, acting as an acting-depolymerizing agent [63]. Swinholides A–C (14a-c) (Figure 5) was isolated from the marine sponge Theonella swinhoei [64]. The structure consists of a 44-membered dimeric macrolide, with four tetrahydropyran rings, two conjugated double bonds connected to each ester, and the polypropionate chains had nine chiral centers [65]. Swinholides 14a-c exhibit high cytotoxic activity against a variety of tumor cells and possess antifungal activities [50].
Another group of macrolides named ansa lactams, possess an aromatic ring (naphthalene, naphthoquinone, or benzene) connected to the aliphatic chain in non-adjacent positions. Rifamycin S 15 (Figure 5) is a member of the lactam ansa macrolide antibiotic that was isolated from Norcodia mediterranei in 1951 [66,67]. This macrolactam 15 consists of twenty-four-member rings that contain a naphthoquinone portion that is connected by an aliphatic polypropionate chain with eight consecutive stereogenic centers. Rifamycins have shown activity against Gram-positive and Gram-negative bacteria with the most valuable activity toward Mycobacterium tuberculosis, leprosy, and meningitis. Rifamycin activity works through the inhibition of the DNA-dependent RNA polymerase and at high concentration through the inhibition of the RNA-dependent DNA polymerase of retroviruses [1,68].
The streptovaricin anasamycin antibiotics A–K were first isolated from Streptomyces spectabilis in 1957 by Siminoff [69]. Streptovaricins 16 (Figure 5) showed activity against Gram-positive, Gram-negative bacteria, and Mycobacterium tuberculosis [70]. Ten years later, Rinehart et al. was able to characterize the macrolactams by NMR, chemical derivatization studies, and X-ray analysis [71]. The structures of streptovaricin A 16a, D 16b, and U 16e (Figure 5) consist of a napthoquinoid core and a polypropionate chain with nine contiguous stereogenic centers, five of which have a challenging all-anti relative configuration [72]. Interestingly, the streptovaricin U 16e ansa segment replaced the C10 carbomethoxy for a methyl group and streptovarcin A 16a has two tertiary alcohols at the C6 and C14 positions. Streptovaricin D 16b inhibits the RNA-directed DNA polymerase and was the most active against reverse transcriptase (0.25 μmol/mL for 70% inhibition) in comparison to all of the streptovaricins [73]. Streptovaricin U 16e is a unique, open-chain streptovaricin that inhibits Rauscher leukemia virus RNA-dependent DNA polymerase (40% inhibition at 200 μg/mL) [74]. Protostreptovaricins I–V were isolated after various chromatographic purifications of fractions from the streptovarin complex [75]. The protostreptovaricins are precursors of streptovaricins that are active inhibitors of reverse transcriptase. The ansa chain of 16e and protostreptovaricin I 16c and II 16d (Figure 5) were the same and only differed in the substitution of the napthoquinoid core. In fact, streptovaricin U 16e is most likely the result of the enzymatic hydrolysis of protostreptovaricin I 16c by S. spectabilis.
The ansamycin antibiotic trienomycin A 17 (Figure 5) was isolated and purified by Komiyama et al. in 1985 from the culture broth of Streptomyces sp. No. 83-16. Komiyama reported the physico-chemical characteristics, molecular formula, and biological activities against HeLa S3 (IC50 = 0.1 μg/mL) and PLC hepatoma (IC50 = 0.01 μg/mL) cells in vitro [76]. Then, the structures of trienomycins A 17, B, and C were determined based on their spectroscopical and chemical properties [77,78]. Trienomycins are unique ansamycin lactams with a stereotriad polypropionate segment, a triene, an 1,3,5-trisubstituted benzene, and N-hexahydrobenzoylalanine moieties in the structure. Trienomycins do not exhibit antimicrobial activity against bacteria, fungi, or yeasts. In 1995, Otake et al. isolated trienomycins A 17, B, and C Streptomyces rishiriensis T-23. They also characterized the structure of this mycotrienin antibiotic by NMR studies and reported cytotoxicity against L-5178Y (IC50 = 0.11 μg/mL) cells in vitro [79].

1.1.2. Polyene Macrolides

Polyene macrolides are a larger group of macrolides characterized by the presence of a series of conjugated E double bonds and the macrolactone is in the range of 26–38 atoms. Polyenes have antifungal properties, and the medicinally important ones include the heptaene amphotericin B 20 (Figure 5) and the tetraene nystatin. In 1993, Osada et al. isolated a 32-membered macrolide, RK-397 (18), from a soil sample collected in Japan and produced by Streptomyces sp. 87-397 [80]. They reported the physico-chemical properties, NMR spectra, molecular formula, and biological activities of the oxo pentaene 18 [81]. The polyene 18 (Figure 5) showed cytotoxicity against tumor cells lines K-562 and HL-60, inhibiting human leukemia, and was active against filamentous fungi, yeast, and bacteria. Macrolide structure 18 consists of a conjugated pentaene and a large segment of 1,3-polyols. Roflamycoin 19 (formerly named flavomycoin) (Figure 5) is a 36-membered polyene macrolide that was isolated from Streptomyces roseoflavus and has antifungal activity [82]. Schlegel et al. determined the structure using spectroscopic and chemical degradation methods [83], and then reported the absolute configuration using the 13C acetonide method and synthetic correlation [84]. The structure consists of a conjugated pentaene, a ketone at C17, and a long 1,3-diol chain. Roflamycoin 19 is an ion channel-forming antibiotic that increases the membrane permeability only in the case of sterol-containing membranes. The channels are potential-dependent, have a short lifetime, and high conductance [85]. Amphotericin B 20a (Figure 5) is a polyene antifungal antibiotic that consists of a 38-membered macrolide that was isolated from Streptomyces nodosus from the Orinoco River in Tembladora, Venezuela [86]. Amphotericin 20a is active against yeasts and most fungi including Cryptococcus neoformans, Candida albicans, Sporotrichum, Blastomyces dermatitidis, Histoplasma capsulatum, Coccidioides immitis, and Aspergillus fumigatus. In addition, 20a is used for the treatment of cryptococcosis, histoplasmosis, disseminated candidiasis, coccidioidomycosis, North American blastomycosis, aspergillosis, and sporotrichosis [87]. The structure of 20a was elucidated in 1970 employing chemical manipulations and MS studies [88]. The absolute configuration of 20a was established using X-ray single-crystal analysis [89]. The structure of macrolactone 20a consists of a steretetrad polypropionate portion, 1,3-polyol segments, a cyclic hemiacetal substituted with a carboxylic acid, and a conjugated heptaene. Amphoterolide B 20b (Figure 5) is the aglycone of Amphotericin B 20a.

1.2. Linear Polypropionates

Macrolide formation does not always occur during biosynthesis in some organisms such as marine mollusks, microorganisms, terrestrial plants, and insects, and produce a rare class of polyketide metabolites that can be classified as linear polypropionates [90]. (+)-Discodermolide 21 (Figure 6) is a linear polypropionate that contains a six-membered lactone that was isolated from the Caribbean marine sponge Discodermia dissolute [91]. This natural product has been approved for clinical trials regarding antitumoral activity and has also shown strong antimitotic, antifungal, and immunosuppressant activity [92]. The structure of 21 contains an amide and a polypropionate segment with 13 chiral centers. The absolute configuration was determined by analyzing the NMR data and X-ray results in 1990 [93]. Discodermolide 21 exhibited strong cytotoxicity against murine P388 leukemia and human lung adenocarcinoma A549 cell lines [90]. Zincophorin 22 (Figure 6) was isolated from a strain of Streptomyces griseus and the structure consists of a linear polyketide that contains a tetrahydropyran ring and an methylester termini [94]. This zinc-binding antibiotic (ionophore) possesses a remarkable in vitro activity against Gram-positive bacteria as well as against Clostridium coelchii and influenza WSN/virus [95]. (−)-Serricornin (4,6- dimethyl-7-hydroxy-nonan-3-one) 23 (Figure 6) is a sex pheromone of a female cigarette beetle (Lasiodema serricorw F.) that has been isolated and structurally characterized by chemical and spectroscopic evidence by Chuman et al. in 1979 [96]. Myriaporones (Figure 6) have been isolated from the western Mediterranean Sea marine organism bryozoan Myriapora truncata [97]. The NMR, IR, and FAB/MS data analysis were employed to determine the structures and stereochemistry of each myriaporone (1, 2, 3/4). Interestingly, myriaporones 3/4 (24) (Figure 6) were isolated as inseparable equilibrium mixture of acyclic and cyclic isomers, respectively. In 2004, the research group of Echavarren reported the total synthesis of all myriaporones and the relative and absolute configurations of these cytotoxic compounds [98]. Myriaporones 3/4 (24) were the most active and showed 88% inhibition of L1210 murine leukemia cells at 0.2 μg/mL [97]. Crocacins A–D (25a-d) (Figure 6) are unusual secondary metabolites in which the biosynthesis of the structure consists of a mixture of polypropionate and peptide building blocks [99]. They are potent antifungal and cytotoxic compounds isolated from myxobacteria Chondromyces crocatus and the bacteria Chrondomyces pediculatusi by Reichenbach et al., which named the structures of 25 as (Z)-enamides [100]. The structure of crocacins 25a-d has a common polypropionate steretetrad fragment (C16–C19) and the conjugated (E,E)-dienamide motif (C11–C15). The relative configuration of crocacins 25a-d was confirmed by Jansen et al. in 1999 [101]. Their absolute configuration was later confirmed by its first total synthesis in 2000 by Rizzacasa et al. [102]. Only crocacins A 25a and D 25a have shown high biological inhibition of the electron transport chain in a beef heart mitochondrial respiration assay. They also inhibited the growth of several yeast and fungi in vitro such as Plasmopara viticola and Phytophthora infestans [103]. Dolabriferols 26a-c (Figure 6) were isolated by Gavagnin et al. from the anaspidean mollusk Dolabrifera dolabrifera [104]. The structure of dolabriferol was determined by spectral methods and the relative stereochemistry was determined by X-ray analysis and consisted of two polypropionate subunits linked by an ester. Similar compounds, dolabriferol B 26b and C 26c, were isolated by Rodriguez et al. from a Caribbean mollusk Dolabrifera dolabrifera from Puerto Rico and characterized using spectral data and X-ray analysis [105]. Dolabriferol, in general, did not display any significant activity with the exception of Dolabriferol B 26b and 26c at a concentration of 128 μg/mL, which showed an inhibitory activity (39% and 93%, respectively) against Mycobacterium tuberculosis H37Rv. Sakuda et al., in 1996, isolated aflastatin A 27 (Figure 6) from Streptomyces sp. MRI142, which is an inhibitor of aflatoxin biosynthesis, but does not inhibit the growth of Aspergillus parasiticus [106]. First, they reported the preliminary structure of 27 after employing chemical degradation and spectroscopic analysis, and later confirmed and revised the absolute configuration of all 29 chiral centers [107,108]. The tetramic acid derivative 27 consists of a long and linear polyol and polypropionate chain with a tetrahydropyran ring moiety.

1.2.1. Polyethers

Polyether antibiotics are a large group of secondary metabolites that show a broad spectrum of bioactivities such as antibacterial, antifungal, antiparasitic, antiviral, and tumor cell cytotoxicity [109]. These are characterized by the presence of several tetrahydrofuran, tetrahydropyran, and oxepane all-trans fused rings. Gymnocin-A 28 (Figure 7) is classified as a polycyclic ether and was isolated from the red tide dinoflagellate, Gymnodinium mikimotoi, at Kushimoto Bay, Japan in 2002 [110]. The structure of gymnocin-A 28 was determined by NMR and FAB/MS spectrometry and consists of a linear array of 14 fused ether rings and a 2-methyl-2-butenal side-chain. The absolute configuration was assigned via Mosher ester analysis. Gymnocin-A 28 showed potent cytotoxicity (IC50 = 1.3 μg/mL) against mouse leukemia P388 cells [111]. Muamvatin 29 (Figure 7) is a marine polypropionate that contains a unique 2,4,6-trioxaadamantane ring skeleton that was isolated from the Fijian pulmonated mollusc Siphonaria normalis by Ireland et al. in 1986 [112]. They elucidated the structure by employing NMR analysis, specifically, the relative stereochemistry of the trioxaadamantane ring (tricyclic ketal) constituents was determined based on the results of a two-dimensional NOE experiment. The absolute configuration of (+)-muamvatin 29 was determined by chemical degradation studies [113] and after the total synthesis of 29 reported by Paterson and Perkins in 1993 [114]. Yamada et al. reported the isolation of two new polypropionates, auripyrones A 30a and B 30b (Figure 7), from the sea hare Dolabella auricularia (Aplysiidae) collected in Mie Prefecture, Japan [115]. They characterized 30a and 30b by employing spectroscopic analyses including NOESY. The structures of 30a and 30b consist of polycyclic polypropionates that contain a δ-pyrone ring and a spiroketal moiety. Auripyrones A 30a and B 30b exhibited cytotoxicity against HeLa S3 cells with IC50 = 0.26 and 0.48 μ/mL, respectively. Tautomycin 1 (31) (Figure 7) was isolated from Streptomyces spiroverticillatus from a soil sample collected in Jiangsu Province, China [116]. The name of tautomycin was chosen because it exists in a tautomeric mixture in solution. Compound 31 exhibited strong antifungal activity against Sclerotinia sclerotiorum, and induced morphological change (blebbing) of the human erythroid leukemia cell K562. Compound 31 is also a potent inhibitor of the spread of human myeloid leukemia cell HL60 [116], and the protein phosphatases type 1 and type 2A [117]. The structure of polyether 31 consists of 13 chiral centers and was determined by chemical degradation and conformational calculations [118]. Other characteristic structural features of tautomycin 1 31 are the tetramic acid derivative termini linked to an ester, internal and terminal ketones, and a polypropionate spiroketal moiety. Aplysiatoxin 32 (Figure 7) was first isolated and purified from the marine gastropod mollusc Stylocheilus longicauda and the structure of 32 was elucidated employing UV–Vis, NMR, IR, and MS data analysis and chemical degradation [119]. The absolute configuration of 32 was established using optical, NMR, X-ray crystallographic analysis, and an examination of the degradation products of 32 [120]. The macrolactone 32 has a tetrahydropyran–tetrahydropyran (THP–THP) spiro system and side chain with a bromo phenol moiety. Aplysiatoxin 32 and related metabolites have shown to be potent tumor promoters [121].

1.2.2. Other Related Polypropionates

Bengamide E 33 (Figure 7) was first isolated from an undescribed Jaspidae sponge collected in the Benga lagoon of the Fiji Islands. The structure and stereochemistry including the chirality of the substituted ɛ-caprolactam ring of bengamide E 33 was established by comparing the structure of the analogs bengamides A and B through extensive spectroscopic studies and chemical degradation. Bengamide E 33 was also extracted and characterized from Jaspis cf. coriacea from Fiji [122] and Pachastrissa sp. from Djibouti [123]. Bengamide 33 has shown in vitro cytotoxic activity against the MDA-MB-435 (IC50 = 4–7 nM) human breast carcinoma cell line and also demonstrated an in vitro anthelminthic activity [122]. Multiplolide A 34 (Figure 7) is a ten-membered lactone (decalactone) that was isolated from the broth of Xylaria multiplex BCC 1111 [124]. In 2001, the structure of 34 was elucidated by employing spectroscopic and spectrometric analysis, but the relative configuration of the epoxide moiety could not be assigned. Then, in 2008, the absolute configuration was assigned after the total synthesis of 34 [125]. Multiplolide A 34 exhibited antifungal activity against Candida albicans (IC50 = 7 μg/mL).

2. Polypropionate Epoxide-Based Approaches

2.1. Kishi (Pioneer)

It was 1979 when Kishi reported a stereo- and regioselective method for the synthesis of propionates from allylic alcohols [126,127]. The synthetic approach started from chiral aldehydes such as 35 to induce stereoselectivity, which after 4-steps was converted into chiral allylic alcohol 36 (Figure 8). Adduct 36 was then treated with mCPBA to form epoxide 37 in a 97% yield. This epoxide was then regioselectively opened to form propionate 38 using LiCu(Me)2 in a 95% yield. Grignard reagents such as (CH2=CHMgBr) can also be used to increase the propionate backbone to produce terminal olefins that can be further functionalized (Figure 8). In simple terms, Kishi’s approach to propionates relies on the use of mCPBA as the epoxidating agent [128] and either the Gilman or Grignard reagents for the alkylation step. It is important to note that Kishi was also able to synthesize all four diastereomers by either the pioneer epoxide approach or through hydroboration-oxidation conditions (Figure 8, middle) [126,127]. To validate the scope, Kishi and coworkers worked on the total synthesis of rifamycin S [129,130,131] and the partial synthesis of narasin and salinomycin [132] (Figure 8, bottom).

2.2. Katsuki

A few years later (1989), Katsuki and coworkers reported a different approach toward polypropionates [133]. Katsuki’s approach is based on an epoxide ring-opening using 2-lithio-1,3-dithiane [134]. Figure 9 shows the synthetic sequence, which starts from a dithianative epoxide ring opening of 46 to produce dithioacetal 47. This intermediate was treated with iodomethane to allow an alkylative desulfurization that released an aldehyde, which was then oxidized with NaClO2 to afford carboxylic acid 48. Then, esterification with CH2N2, followed by α-methylation and a reduction with lithium aluminum hydride, produced the expected propionate 49. Further elaboration of intermediate 49 produced the rifamycin S fragment 50. It is worth noting that Katsuki utilized either VO(acac)2 or WO5·HMPA for epoxidations [133]. Furthermore, the same synthetic approach was used during a formal total synthesis of aplysiatoxin 32; this methodology could also be used to synthesize dibromoaplysiatoxin (Figure 9, bottom right).

2.3. Marshall

In 1998, Marshall reported the addition of chiral allenylstannanes such as 53 to chiral aldehydes (e.g., 52) to produce homopropargylic alcohols with two contiguous anti-anti stereocenters (Figure 10, top left) in an 87% yield [135]. The homopropargylic alcohol 54 was then reduced with Red-Al to allylic alcohol 55. Then, using Sharpless asymmetric epoxidation conditions, the expected product 56 was observed with an 87:13 dr. The addition of Me2Cu(CN)Li2 opened the epoxide by regio-and stereoselective addition of a methyl group to afford polypropionate 57 in an 85% yield. Finally, the protection and deprotection sequences delivered zincophorin’s C7–C13 fragment (58). Likewise, this reiterative epoxidation-methyl cuprate sequence to transform allylic alcohols into polypropionates was used to prepare rifamycin-S’s C21–C27 fragment (60), albeit employing other enantiomers during the initial allenylstannane addition (Figure 10, middle) [135]. Due to the fact that by using chiral allenylstannanes, five carbons can be added to the backbone in a single step, so this polypropionate approach is efficient and useful for the synthesis of natural products. For instance, Marshall and coworkers reported the total synthesis of (+)-discodermolide with an excellent stereocontrol observed [136]. They also reported the stereoselective synthesis of bafilomycin A1′s C15–25 fragment, this time using allenylzinc as the key reagent (Figure 10, bottom) [137].

2.4. Miyashita

Miyashita was one of the major contributors to publications related to the synthesis of fragments and natural products containing polypropionate subunits. His methodology relies on the use of trimethylaluminum for the regio- and stereospecific methylation of epoxides. In 1991, his group reported a new iterative method for the construction of polypropionate chains from epoxy acrylates [138]. The method takes advantage of optically active epoxy acrylates that can undergo stereospecific methylation with 10 equiv. of (Me)3Al to afford highly diastereoselective adducts. For example, the (E)-epoxy acylate 61 produced anti adduct 62, while the (Z)-epoxy acylate 63 gave syn adduct 64 (Figure 11, top) in very high yields and diastereoselectivity. It is important to note that having adjacent chiral centers does not affect the yields or diastereoselectivity [138]. Using this stereospecific methylation approach, Miyashita and co-workers synthesized Prelog–Djerassi lactone 2 in six synthetic steps (Figure 11, middle) [139]. They also reported the formal synthesis of protomycinolide IV 7 [139]. In addition, several other fragments and/or natural products have been prepared by Miyashita et al. For instance, the following polypropionate-containing-natural products: (−)-serricornin 23 [140], rifamycins 15 [141], streptovaricin U 16c [142], protostreptovaricin I and II (16c-d) [142], scytophycin C 11 [143,144,145], swinholides A-C 14a-c [146], tedanolide 9 [147], lepranthin 8 [148], and venturicidins 10a-c [149] were all prepared from epoxides and trimethylaluminum for the regio- and stereospecific methylation (Figure 11, bottom). It is worth noting that Miyashita also reported the use of m-chloroperoxybenzoic acid (mCPBA) for highly stereoselective epoxidations of chiral allylic alcohols [150] and organoselenium reagents for a chemoselective reduction of epoxides [151]; both important methodologies for the synthesis of polypropionates.

2.5. Prieto

Prieto’s methodology is a relatively simple substrate-control approach and consists of three reiterative steps: (1) regioselective cleavage of epoxide 71 with an organoaluminium or organomagnesium reagents; (2) cis or trans reduction of the resulting alkyne 72; (3) syn or anti epoxidation of the homoallylic alcohol 73 [152,153] to subsequently produce the 3,4-epoxyalcohol 74. The application of the reiterative three step methodology produced a new propionate unit 75 (Figure 12). The absolute stereochemistry of the polypropionate fragment was introduced during the synthesis of the first epoxide for step one by employing the Sharpless asymmetric epoxidation [154]. The relative configuration of the methyl groups is defined by the cis/trans geometry of the epoxide precursor, which in turn is defined by the geometry of the alkene precursors. The configuration of the secondary hydroxy groups is defined by the configuration of the epoxide precursors. In 2009, Prieto et al. reported the stereoselective synthesis of the very challenging all-anti, the C6–C10 polypropionate chain 83 of streptovaricin U 16e (Figure 12) [155]. For this, the syn-selective epoxidation of cis-homoallylic alcohols 77 and 79 were obtained by either applying the iodocarbonation/methanolysis procedure [40,156] or the VO(acac)2 catalyzed epoxidation reaction [153,157]. Then, the regioselective cleavage of epoxides 76, 79, and 81 with a propynyl aluminum reagent followed by a cis reduction produced homoallylic alcohols 77, 79, and 82 in good yields. Furthermore, the incorporation of the copper-catalyzed Grignard epoxide cleavage reaction [158] for the cleaving of epoxide 76 to complement the alkynyl alane approach [25,25,159,160] was also developed.
To demonstrate the synthetic usefulness of this non-aldol approach for the elaboration of polypropionate-containing natural products, they applied a linear three-reaction sequence to the stereoselective construction of the stereohexads corresponding to the C15–C10 84 [158], C5–C12 85 [161], and C15–C8 86 [161] polypropionate chains of streptovaricin D 16b including the C10 carbomethoxy functionality needed for the elaboration (Figure 13). Prieto also reported the C5–C10 87 [160] and triad C12–C16 88 [161] fragments of elaiophylin 5, in addition to the stereotetrad C22–C27 89 [158] fragment of rifamycin S 15 (Figure 13). In 2012, the linear, convergent, and enantioselective synthesis of the C14–C25 90 fragment of bafilomycin A1 6 was reported in a 16% overall yield. The process consisted of eight steps in its longest linear sequence (Figure 13) using a dithiane substitution reaction to the coupling of the corresponding linear fragments [162]. This group has reported the convergent synthesis of optically active and common precursors of the C1–C5/C15–C11 segment 91a,b of lankanolide 4b [157] (Figure 13) and the convergent precursor for the C3–C9/C12–C18 fragments 92a,b of dolabriferol B 26b (Figure 13) [161]. They also reported a second-generation methodology in 2014 to introduce the hydroxymethyl moiety found at the C16 of tedanolide 9 and C18 of myraporone 3/4 (24) [159]. Specifically, they constructed an optically active stereotetrad 92 that is a common intermediate for the elaboration of C6–C9 93a and C14–C7 93b aliphatic chains of tedanolide 9 and myraporone 3/4 (24), respectively (Figure 13). The most recent paper reported by Prieto et al. describes the enantioselective and stereoselective synthesis of several polypropionate chains, which include the C15–C20 segment 94 of crocacin 25 and the polypropionate chains of C22–C27 95, mycalolide A 13, and C19–C24 95 of the lobophorolide 12 natural products (Figure 13) [161]. In summary, this reiterative, linear, and convergent methodology has demonstrated the versatility to synthesize stereotriads, stereotretrads, and stereohexads with regio-, enantio-, and diastereoselectivity in polypropionate natural products.

2.6. Sarabia

More recently, the group of Sarabia reported the stereoselective synthesis of polypropionate-type frameworks utilizing an epoxy amide-based strategy (Figure 14) [163]. The reiterative methodology of Sarabia consists of four consecutive steps: sulfur ylide reaction, oxirane opening, protection, and reduction of amide. The starting material is a chiral aldehyde 97 that is reacted with a sulfur ylide 98 in a stereoselective fashion induced by the presence of an asymmetric center at the α-position of aldehyde 97 (Figure 14). The resulting trans epoxyamide 99 is then subjected to the epoxide opening reaction mediated by a methyl organocuprate reagent, followed by the protection of the resulting alcohol as a silyl ether and the reduction of the amide to the alcohol. At this point, the oxidation of the alcohol provided a new aldehyde 100, which was ready for the application of a second reaction sequence. This sequence provided the anti-relative configuration between the chiral centers present in the resulting polypropionate chain 101 (Figure 14). A collection of different chiral aldehydes was subjected to reaction with sulfur ylide. The majority showed remarkable stereofacial differentiation, providing a major diastereomer, in contrast to other aldehydes that displayed poor to no stereoselectivity. Despite the difficulties encountered for the preparation of some epoxide amides with respect to the diastereomeric yields, Sarabia was able to prepare various stereo-n-ads such as 108 including the synthesis of the C5–C17 polypropionate chain of streptovaricin U 109 (Figure 14) [164]. In 2010, Sarabia and co-workers reported on the stereoselective preparation of bengamide E 33 and analogs using chiral sulfur ylides [165]. Specifically, they achieved the synthesis of the tetraol polyketide fragment present in bengamide E 33 by using two trans-epoxides synthesized by employing Sharpless asymmetric epoxidation (SAE). This was then followed by conversion into epoxyamides to apply the oxirane-ring-opening process with alkylborates, catalyzed by palladium (0) to yield the corresponding ring opened syn-diol products (Figure 14) [166].

2.7. Jung

Jung et al. developed a “non-aldol aldol” protocol for the synthesis of polypropionates that involves the stereoselective olefination of the aldehyde 110, followed by a hydride reduction to produce the allylic alcohol 111. Sharpless asymmetric epoxidation of 111 yielded epoxy alcohol 112 (Figure 15) [24,167,168]. Silylation of the primary alcohol in 112 and the final Lewis acid catalyzed the rearrangement of the epoxy silyl ether 113, generating the desired syn or anti propionate unit product 114. The epoxy silyl ether epoxide 112 is activated with the Lewis acid TBSOTf, and an intramolecular hydride transfer cleaves the epoxide 113, generating an optically active aldehyde 114 (Figure 15). E-allylic alcohols like 111 generate syn adducts whereas Z-allylic epoxides generate anti adducts. In 1999, Jung et al. examined the Lewis acid promoted rearrangements of several allylic epoxides and their derivatives as a method for the preparation of compounds that can be used in the synthesis of natural products such as Tedanolides [167]. Jung et al. first reported the total synthesis of auripyrone B 30a from the known epoxide 120 in 18 steps and a 17% overall yield. They employed a non-aldol Lewis acid catalyzed rearrangement of an epoxy silyl ether strategy combined with a Paterson aldol process (Figure 15) [168]. In 2010, Jung et al. also reported the total synthesis of auripyrone B 30b in only 20 steps in an 8% yield starting from optically active epoxy silyl ether 115 using a syn non-aldol aldol/anti cuprate opening strategy to produce the C8–C12 polypropionate fragment 119 (Figure 15) [169].

2.8. McDonald

In 2002, McDonald and co-workers reported a new strategy for the construction of alternating 1,3-polyols based on two parts. The first, cross-coupling of six-carbon alkynol 122 (nucleophile) and the six-carbon epoxyalkynol 123 (electrophile) modules by epoxide opening. Second, the resulting internal alkyne 124 was hydrated to form 125 and then reduced with either 1,3-anti-15 or syn-stereoinduction to form the polyol fragment 126 (Figure 16) [170]. They reported the preparation of the C13–C27 polyol fragment 134 of the macrolide polyne RK-397 18. The cross-coupling between terminal alkyne 127 with epoxide 128 produced the dialkyne 129. The hydration-reduction of the resulting internal alkyne in 129 produced the polyol fragment 130, which was then coupled with epoxide 131. The synthesis concluded with regio- and stereoselective synthesis of the polyol fragment 133 (Figure 16) [170]. In another paper, McDonald and co-workers reported the synthesis of the C9–C27 polypropionate fragment 134, which is the degradation product of aflastatin A 27 using the alkyne-epoxide cross-coupling strategy [171]. Particularly, the polypropionate fragments 134 and 135 were obtained by employing the stereoselective vanadium catalyzed and iodocarbonation/methanolysis epoxidations, respectively.

2.9. Lipshutz

Lipshutz and co-workers, in 1984, described a two-step methodology for the formation of all syn-1,3-polyols 146. They employed the vinyl-lithium-based higher order cyano-cuprates to cleave the monosubstituted epoxide 143 and 145, followed by epoxidation of the resulting homoallylic alkenol 144 (Figure 17) [172,173]. An alternative methodology for the synthesis of polypropionate chains was also reported by Lipshutz and Barton in 1988 starting with the regiospecific cleavage of epoxide 143 with the cis- or trans-propenyl high order cuprates (Figure 17) [174]. The cleavage of the disubstituted 3,4-epoxy alcohol 147 yielded the desired polypropionate products 148 and 149 in variable regioselectivities. In contrast, the epoxide cleavage with the vinyl cuprate reagent enhanced the regioselectivity at latter stages, providing 1,3-diol exclusively. In this sequence, the epoxidation of the homoallylic alcohols was achieved in high stereoselectivities using the iodocarbonation/methanolysis sequence, while the VO(acac)2 protocol was also quite efficient. In 1988, Lipshutz and co-workers also reported on the construction of the desired all syn-polyol array 153 and 156 by employing the coupling reaction between dithiane 150 with epoxides 151 and 154 to obtain the products 152 and 155, respectively. Hydrolysis of dithianes followed by a highly syn-selective LiAlH4-LiI reduction method produced all of the syn-polyols 153 and 155 (Figure 17). In 1989, this epoxide and dithiane coupling strategy was applied to the construction in an optically pure form of the C12–C35 fragment 157 of the 36-membered polyene macrolide roflamycoin 19. Unfortunately, the stereochemistry of the eleven chiral centers in this year was unknown and they assumed that 19 had all of its polyhydroxyl groups in a syn-relationship (Figure 17) [175].

2.10. Nicolaou

Nicolao and Uenishi, in 1982, reported an approach that utilizes the allylic alcohol 158 as a starting material, converted in epoxyalcohol 159 by the SAE reaction. This was then converted into the γ,δ-epoxy-α,β-unsaturated ester 160 before a regioselective hydride reduction with DIBAL to afford the allylic and homoallylic alcohol 161 (Figure 18) [176]. The application of the aforementioned sequence generated the 1,3-syn-polyol fragment 162. Years later, Nicolaou and co-workers reported the synthesis of various key building blocks 170 to 173 toward the total synthesis of amphotericin B 20a and amphoterolide B 20b (Figure 18) [177]. They envisaged a convergent synthesis to construct polyol fragments 170 and 172. To achieve this, 163 was converted into epoxide 164 by the SAE reaction, then Swern oxidation followed by the Witting reaction, which produced the conjugated epoxy ester 165. A protection–deprotection strategy provided 167, which was then converted into the epoxide 168. The hydride reduction of 168 provided the syn-1,3-diol unit 169, which is the common precursor for the divergent synthesis of 170 and 172 polyol fragments (Figure 18). For the synthesis of the fragment 172 present, Nicolaou employed an epoxide cleavage reaction with two organometallic reagents; Et2AlC≡CCH2OTBDPS and CH2=CHMgBr catalyzed by copper, and SAE as the key step for its elaboration. The total synthesis of amphotericin B 20a and amphoterolide B 20b was achieved by the efficient coupling of 170, 171, 172, and 173 fragments by employing four aldehyde-phosphonate type condensation reactions and an esterification reaction [178].

2.11. Polypropionate Synthesis by Others

Aside from the synthetic approach toward polypropionates by the research groups described above, there are other research groups that have contributed to this field with one or two publications (also see Table 1). Figure 19 depicts several natural products that these research groups have targeted, with reports varying from the synthesis of fragments to total syntheses. With few exceptions, these other methods are analogs to the synthetic approaches described above, since they follow slightly modified protocols and somewhat different starting materials. However, they do complement the epoxide-approach toolkit. In 1979, Corey reported the use of an epoxide intermediate to construct the C15–C29 chain of rifamycin 15 (Figure 19) [179]. However, it was not until 1982 when he disclosed that a vinyl cuprate was used to open the epoxide [180]. A year earlier (1981), Lipshutz documented the use of high order mixed organocuprates in substitution reactions at unactivated secondary centers [181,182]. In 1982, Lipshutz also documented their reactivity toward epoxides [183], followed by their application in other organic reactions [176,184,185] and the synthesis of fragments of polypropionates [175]. In 1989, Yonemitsu et al. described an alternative synthesis toward two fragments C1–C5/C9–C15 of erythronolide A 3b via regio- and stereoselective reductive ring opening. The key of their methodology was to utilize trisubstituted epoxides (the methyl group pre-installed) and B2H6 as a reducing agent [186]. Schreiber also reported stereochemical studies of epoxides toward the ansa chain of streptovaricin A 16a. He also took advantage of trisubstituted epoxides but used an intramolecular ring opening facilitated by a nearby urethane that was deprotected with Et2AlCl [187]. In 1993, Yadav et al. also utilized trimethylaluminum, an epoxy acrylate intermediate to synthesize fragment C9–C16 of trienoycin A 17 [188]. In addition, Zhai used Me2CuLi to open chiral allylic epoxides [189], Haddad used organocuprates and organoaluminium reagents [190], Sato used NaBH4 to open trisubstituted trimethylsilyl epoxides [191], Mulzer used high order organo cyanocuprates to open two consecutive epoxides [192], and Tan used SmI2 to reduce epoxides and MeMgCl or Me3Al to ring-open the epoxide while installing the methyl group [193]. More recently (2020), Kaliappan employed palladium mediated stereoselective reduction (ring opening) of the trisubstituted epoxide 174 to produce propionate 175 in a 68% yield [194]. It is important to note that this is the first example of palladium chemistry used to construct polypropionates, which led to the synthesis of the C1–C10 fragment of muamvatin 29 (Figure 19, middle).

3. Other Epoxide Based-Syntheses of

Polyols and Polyethers

It is well-known that all of the methodologies described above are paired with the aldol chemistry developed by Evans and others. Nonetheless, Evans also reported the use of epoxides for the synthesis of polyethers such as ferensimycin B 176 [199]. In this case, a neighboring hydroxy group performed an intramolecular epoxide ring opening to form an ether, which in turn could lead to polyethers after a few iterations. Lipshutz also contributed with his high order cuprate protocol for the synthesis of polyols [200] such as roflamycoin 19 [174,176] (Figure 17 and Figure 19). Isobe also utilized Grignard or Gilman reagents to regioselective open epoxides to either form polypropionates or polyethers such as tautomycin 1 31 [197] (Figure 19). DIBAL and REDAL were the reducing agents of choice by Nicolaou for the regioselective reduction of chiral disubstituted epoxides [177] toward polyols such as amphoterecin B 20a and amphotheronolide B 20b [178] (Figure 18 and Figure 19). The total synthesis of multiplolide A 34 was reported by Meshram et al. Treating a disubstituted epoxide with Sc(OTf)3 and water afforded the expected diol in good yield [196]. Aside from the well-known Sharpless dihydroxylation reaction to make polyols, he also used Red-Al to open epoxides in his search for synthetic alternatives toward polyols [201]. Furthermore, Jamison reacted a polyepoxide with water at 60 °C for days to produce a polyether intermediate used to construct the HIJK rings of gymnocin A 28 [198] (Figure 19, bottom) and Table 1.

4. Conclusions

This manuscript describes the isolation and characterization of approximately 30 polypropionate and related natural products that have shown antimicrobial activities against bacteria, fungi, and yeast. Some of the polypropionates also displayed cytotoxicity against different types of cancer lines. The first epoxide-based methodology for polypropionate synthesis was reported by Kishi and collaborators in 1980 and was applied to the first synthesis of the ansa chain of rifamycin S 15. Kishi’s five step methodology was based on Sharpless asymmetric epoxidation followed by lithium dimethyl cuprate ring opening to construct the propionate units. Yamagishi et al. reported the synthesis of two types of Kishi’s intermediates, one for the construction of the ansa chain of rifamycin S 15 and the other for the formal synthesis of the aplysiatoxin 32 C1–C21 fragment, based on the stereoselective vanadium-catalyzed or tungsten-based epoxidations of cis-homoallylic alcohols and the asymmetric epoxidation of allylic alcohols. Marshall et al. were also influenced by the pioneering work of Kishi and employed a Lewis acid mediated allenylmetal addition to chiral aldehydes followed by the Kishi reiterative asymmetric epoxidation-methyl cuprate sequence. The research groups of Miyashita (1992 to 2011) and Prieto (2004 to 2017) applied their epoxide-alkynylaluminum cleavage approaches to the formal and total synthesis of several polypropionate natural products in common. Miyashita’s group employed the regiospecific cleavage of γ,δ-epoxy acrylates with trimethylaluminum as the key steps while Prieto’s reiterative methodology reported the stereoselective vanadium-catalyzed epoxidation of homoallylic alcohols, followed by the epoxide-ring opening with alkynylaluminum reagents or alkenyl Grignards as the crucial steps for the elaboration of propionate motifs. Remarkably, McDonald’s group developed a cross-coupling with elaborated alkynyllithium and epoxyalkyne fragments to construct long polypropionate chains with good regioselectivity. Jaminson was the only group that employed SAE for the construction of trans-syn-fused polycyclic ethers by an epoxide-opening cascade reaction promoted by water. The groups of Nicolaou and Yonemitzu based their epoxide methodology on the regioselective hydride ring-opening trisubstituted 2,3-epoxyalcohols for the construction of the 1,3-polyols and polypropionate fragments, respectively. Lipshutz and Nicolaou based their epoxide-cleavage methodologies to the stereoselective synthesis of 1,3-diols motifs found in polyene macrolides. The most recent report was made in 2020 by Kaliappan et al., which reported the synthesis of the polypropionate fragment of Muamvatin using palladium chemistry to open their epoxide. The most successful organometallic reagents to introduce methyl groups by the cleavage of epoxides were Me3Al and Me2CuLi. Indeed, the Kishi, Marshall, Sarabia, Jung, and Isobe groups employed a methyl cuprate reagent, while Miyashita and Yadav selected the methyl aluminum reagent to construct the polypropionate fragments. Particularly, the Yamagishi, Prieto, Isobe, and Lipshutz groups reported the cleavage of hindered 3,4-epoxy alcohols with dithianes for the convergent synthesis of long polypropionate and polyol chains.
It is important to note that most of the reviewed synthetic approaches were developed/reported a few decades ago, with a clear decline in the number of publications in the past decade. However, it is believed that this review should help to put these methods back into the limelight. Thus, an incremental interest on the topic is expected in the near future, particularly for methods utilizing transition metals or mild reaction conditions.
In brief, this review condenses the known reports of epoxide-based approaches toward polypropionates, which are attractive structural motifs in synthetic organic chemistry and the total synthesis of natural products. As described above, the presence of propionate moieties in natural products has shown good antimicrobial activity. Therefore, it is believed that the known literature and this review will be beneficial in the design of novel approaches and the synthesis of other important natural products in the near future.

Author Contributions

Conceptualization, R.R.R.-B. and A.B.; Data curation, R.R.R.-B.; Formal analysis, R.R.R.-B. and A.B.; Funding acquisition, A.B.; Writing—original draft, R.R.R.-B. and A.B.; Writing—review & editing, R.R.R.-B., S.R.I. and A.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was in part supported by the National Institute of General Medical Sciences of the National Institutes of Health (NIH) under award number 1R15GM141726-01.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data are contained within the article.

Acknowledgments

All authors thank Florida Gulf University (USA) and the University of Puerto Rico, Rio Piedras (Puerto Rico, USA) for partially supporting this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dewick, P.M. Medicinal Natural Products: A Biosynthetic Approach; John Wiley & Sons: Hoboken, NJ, USA, 2009. [Google Scholar]
  2. Birch, A.J. Biosynthesis of Polyketides and Related Compounds. Science 1967, 156, 202–206. [Google Scholar] [CrossRef]
  3. Koskinen, A.M.; Karisalmi, K. Polyketide Stereotetrads in Natural Products. Chem. Soc.Rev. 2005, 34, 677–690. [Google Scholar] [CrossRef] [Green Version]
  4. Rohr, J. A New Role for Polyketides. Angew. Chem. Int. Ed. 2000, 39, 2847–2849. [Google Scholar] [CrossRef]
  5. Cragg, G.M.; Grothaus, P.G.; Newman, D.J. Impact of Natural Products on Developing New Anti-Cancer Agents. Chem. Rev. 2009, 109, 3012–3043. [Google Scholar] [CrossRef] [PubMed]
  6. Hoffmann, R.W. Stereoselective Syntheses of Building Blocks with Three Consecutive Stereogenic Centers: Important Precursors of Polyketide Natural Products [New Synthetic Methods (68)]. Angew. Chem. Int. Ed. 1987, 26, 489–503. [Google Scholar] [CrossRef]
  7. Schetter, B.; Mahrwald, R. Modern Aldol Methods for the Total Synthesis of Polyketides. Angew. Chem. Int. Ed. 2006, 45, 7506–7525. [Google Scholar] [CrossRef]
  8. Li, J.; Menche, D. Direct Methods for Stereoselective Polypropionate Synthesis: A Survey. Synthesis 2009, 2009, 2293–2315. [Google Scholar]
  9. Paterson, I. New Methods and Strategies for the Stereocontrolled Synthesis of Polypropionate-Derived Natural Products. Pure Appl. Chem. 1992, 64, 1821–1830. [Google Scholar] [CrossRef] [Green Version]
  10. Chen, M.; Roush, W.R. Crotylboron-Based Synthesis of the Polypropionate Units of Chaxamycins A/D, Salinisporamycin, and Rifamycin S. J. Org. Chem. 2013, 78, 3–8. [Google Scholar] [CrossRef] [Green Version]
  11. Hoffmann, R.W.; Dresely, S. Stereoselective Synthesis of Alcohols, XXIX. Addition of (A-Methoxycrotyl) Boronates to Aldehydes. Chem. Ber. 1989, 122, 903–909. [Google Scholar] [CrossRef]
  12. Brown, H.; Bhat, K.S.; Randad, R.S. A Highly Diastereoselective Addition of [E]-and [Z]-Crotyldiisopinocampheylboranes to α-Substituted Aldehydes. J. Org. Chem. 1987, 52, 3701–3702. [Google Scholar] [CrossRef]
  13. Yamamoto, Y.; Asao, N. Selective Reactions Using Allylic Metals. Chem. Rev. 1993, 93, 2207–2293. [Google Scholar] [CrossRef]
  14. Denmark, S.E.; Fu, J. Catalytic Enantioselective Addition of Allylic Organometallic Reagents to Aldehydes and Ketones. Chem. Rev. 2003, 103, 2763–2794. [Google Scholar] [CrossRef] [PubMed]
  15. Masse, C.E.; Panek, J.S. Diastereoselective Reactions of Chiral Allyl and Allenyl Silanes with Activated C: X. Pi Bonds. Chem. Rev. 1995, 95, 1293–1316. [Google Scholar] [CrossRef]
  16. Guindon, Y.; Brazeau, J.-F. Diastereoselective Mukaiyama and Free Radical Processes for the Synthesis of Polypropionate Units. Org. Lett. 2004, 6, 2599–2602. [Google Scholar] [CrossRef] [PubMed]
  17. Kiyooka, S.; Shahid, K.A. A Divergent Synthesis of Essentially Enantiopure Syn-and Anti-Propionate Aldol Adducts Based on the Chiral 1, 3, 2-Oxazaborolidin-5-One-Promoted Asymmetric Aldol Reaction Followed by Diastereoselective Radical Reduction. Bull. Chem. Soc. Jpn. 2001, 74, 1485–1495. [Google Scholar] [CrossRef]
  18. Kiyooka, S.; Shahid, K.A. An Efficient Method for the Synthesis of Enantiopure 2, 3-Anti-Propionate Aldols Involving a 3, 5-Syn-or Anti-Diol Subunit through Chiral Borane-Mediated Enantioselective Aldol Reaction Coupled with Radical Reduction. Tetrahedron Lett. 2000, 41, 2633–2637. [Google Scholar] [CrossRef]
  19. Sibi, M.P.; Petrovic, G.; Zimmerman, J. Enantioselective Radical Addition/Trapping Reactions with α,β-Disubstituted Unsaturated Imides. Synthesis of a Nti-Propionate Aldols. J. Am. Chem. Soc. 2005, 127, 2390–2391. [Google Scholar] [CrossRef]
  20. Hanessian, S.; Gai, Y.; Wang, W. Stereocontrolled Functionalization in Acyclic Systems by Exploiting Internal 1, 2-Asymmetric Induction—Generation of Polypropionate and Related Motifs. Tetrahedron lett. 1996, 37, 7473–7476. [Google Scholar] [CrossRef]
  21. Whitehead, A.; McParland, J.P.; Hanson, P.R. Divalent Activation in Temporary Phosphate Tethers: Highly Selective Cuprate Displacement Reactions. Org. Lett. 2006, 8, 5025–5028. [Google Scholar] [CrossRef]
  22. Kumar, N.; Kiuchi, M.; Tallarico, J.A.; Schreiber, S.L. Small-Molecule Diversity Using a Skeletal Transformation Strategy. Org. Lett. 2005, 7, 2535–2538. [Google Scholar] [CrossRef]
  23. Sasaki, M.; Tanino, K.; Miyashita, M. Regioselective Alkyl and Alkynyl Substitution Reactions of Epoxy Alcohols by the Use of Organoaluminum Ate Complexes:  Regiochemical Reversal of Nucleophilic Substitution Reactions. Org. Lett. 2001, 3, 1765–1767. [Google Scholar] [CrossRef]
  24. Jung, M.E.; D’Amico, D.C. Enantiospecific Synthesis of All Four Diastereomers of 2-Methyl-3-[(Trialkylsilyl)Oxy]Alkanals: Facile Preparation of Aldols by Non-Aldol Chemistry. J. Am. Chem. Soc. 1993, 115, 12208–12209. [Google Scholar] [CrossRef]
  25. Tirado, R.; Torres, G.; Torres, W.; Prieto, J.A. Regioselective Cleavage of Cis- and Trans-2-Methyl-3,4-Epoxy Alcohols with Diethylpropynyl Aluminum. Tetrahedron Lett. 2005, 46, 797–801. [Google Scholar] [CrossRef]
  26. Myles, D.C.; Danishefsky, S.J. The Synthesis of Polyoxygenated Natural Products via Fully Synthetic Branched Pyranose Derivatives: Application to the Erythronolide Problem. Pure Appl. Chem. 1989, 61, 1235–1242. [Google Scholar] [CrossRef] [Green Version]
  27. Yadav, J.S.; Rao, K.V.R.; Ravindar, K.; Reddy, B.V.S. Total Synthesis of (+)-Bourgeanic Acid Utilizing Desymmetrization Strategy. Eur. J. Org. Chem. 2011, 2011, 58–61. [Google Scholar] [CrossRef]
  28. Vogel, P.; Turks, M.; Bouchez, L.; Craita, C.; Murcia, M.C.; Fonquerne, F.; Didier, C.; Huang, X.; Flowers, C. Use of Sultines in the Asymmetric Synthesis of Polypropionate Antibiotics. Pure Appl. Chem. 2008, 80, 791–805. [Google Scholar] [CrossRef] [Green Version]
  29. Lautens, M.; Fagnou, K.; Hiebert, S. Transition Metal-Catalyzed Enantioselective Ring-Opening Reactions of Oxabicyclic Alkenes. Acc. Chem. Res. 2003, 36, 48–58. [Google Scholar] [CrossRef] [PubMed]
  30. Shen, X.; Wasmuth, A.S.; Zhao, J.; Zhu, C.; Nelson, S.G. Catalytic Asymmetric Assembly of Stereodefined Propionate Units:  An Enantioselective Total Synthesis of (−)-Pironetin. J. Am. Chem. Soc. 2006, 128, 7438–7439. [Google Scholar] [CrossRef]
  31. Djerassi, C.; Zderic, J.A. The Structure of the Antibiotic Methymycin1. J. Am. Chem. Soc. 1956, 78, 6390–6395. [Google Scholar] [CrossRef]
  32. Barton, D.; Meth-Cohn, O. Comprehensive Natural Products Chemistry; Newnes: Oxford, UK, 1999; ISBN 0-08-091283-4. [Google Scholar]
  33. Gäumann, E.; Hütter, R.; Keller-Schierlein, W.; Neipp, L.; Prelog, V.; Zähner, H. Stoffwechselprodukte von Actinomyceten. 21. Mitteilung. Lankamycin Und Lankacidin. Helv. Chim. Acta 1960, 43, 601–606. [Google Scholar] [CrossRef]
  34. Namiki, S.; Omura, S.; Nakayoshi, H.; Sawada, J. Studies on the Antibiotics from Streptomyces Spinichromogenes Var. Kujimyceticus. I Taxonomic and Fermentation Studies with Streptomyces Spinichromogenes Var. Kujimyceticus. J. Antibiot. 1969, 22, 494–499. [Google Scholar] [CrossRef] [Green Version]
  35. Kinashi, H.; Mori, E.; Hatani, A.; Nimi, O. Isolation and Characterization of Linear Plasmids from Lankacidin-Producing Streptomyces Species. J. Antibiot. 1994, 47, 1447–1455. [Google Scholar] [CrossRef] [PubMed]
  36. Arakawa, K.; Kodama, K.; Tatsuno, S.; Ide, S.; Kinashi, H. Analysis of the Loading and Hydroxylation Steps in Lankamycin Biosynthesis in Streptomyces Rochei. Antimicrob. Agents Chemother. 2006, 50, 1946–1952. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Martin, J.R.; Egan, R.S.; Goldstein, A.W.; Mueller, S.L.; Keller-Schierlein, W.; Mitscher, L.A.; Foltz, R.L. 3 ″-De-O-methyl-2 ″, 3 ″-anhydro-lankamycin, a New Macrolide Antibiotic from Streptomyces Violaceoniger. Helv. Chim. Acta 1976, 59, 1886–1894. [Google Scholar] [CrossRef] [PubMed]
  38. Keller-Schierlein, W.; Roncari, G. Stoffwechselprodukte von Mikroorganismen 46. Mitteilung Die Konstitution Des Lankamycins. Helv. Chim. Acta 1964, 47, 78–103. [Google Scholar] [CrossRef]
  39. Keller-Schierlein, W.; Roncari, G. Stoffwechselprodukte von Actinomyceten. 33. Mitteilung. Hydrolyseprodukte von Lankamycin: Lankavose Und 4-O-Acetyl-arcanose. Helv. Chim. Acta 1962, 45, 138–152. [Google Scholar] [CrossRef]
  40. Arcamone, F.M.; Bertazzoli, C.; Ghione, M.; Scotti, T. Melanosporin and Elaiophylin, New Antibiotics from Streptomyces Melanosporus (Sive Melonsporofaciens) n. Sp. G. Microbiol. 1959, 7, 207–216. [Google Scholar]
  41. Ley, S.V.; Neuhaus, D.; Williams, D.J. A Conformational Study of Elaiophylin by X-Ray Crystallography and Difference 1H NMR Methods; Observation of a Selective Sign Reversal of the Nuclear Overhauser Effect. Tetrahedron Lett. 1982, 23, 1207–1210. [Google Scholar] [CrossRef]
  42. Gerlitz, M.; Hammann, P.; Thiericke, R.; Rohr, J. The Biogenetic Origin of the Carbon Skeleton and the Oxygen Atoms of Elaiophylin, a Symmetric Macrodiolide Antibiotic. J. Org. Chem. 1992, 57, 4030–4033. [Google Scholar] [CrossRef]
  43. Grigoriev, P.A.; Schlegel, R.; Gräfe, U. Cation Selective Ion Channels Formed by Macrodiolide Antibiotic Elaiophylin in Lipid Bilayer Membranes. Bioelectrochemistry 2001, 54, 11–15. [Google Scholar] [CrossRef]
  44. Werner, G.; Hagenmaier, H.; Albert, K.; Kohlshorn, H. The Structure of the Bafilomycins, a New Group of Macrolide Antibiotics. Tetrahedron Lett. 1983, 24, 5193–5196. [Google Scholar] [CrossRef]
  45. Baker, G.H.; Brown, P.J.; Dorgan, R.J.J.; Everett, J.R.; Ley, S.V.; Slawin, A.M.Z.; Williams, D.J. A Conformational Study of Bafilomycin A1 by X-Ray Crystallography and Nmr Techniques. Tetrahedron Lett. 1987, 28, 5565–5568. [Google Scholar] [CrossRef]
  46. Baker, G.H.; Brown, P.J.; Dorgan, R.J.J.; Everett, J.R. The Conformational Analysis of Bafilomycin A1. J. Chem. Soc. Perkin Trans. 2 1989, 8, 1073–1079. [Google Scholar] [CrossRef]
  47. Bowman, E.J.; Siebers, A.; Altendorf, K. Bafilomycins: A Class of Inhibitors of Membrane ATPases from Microorganisms, Animal Cells, and Plant Cells. Proc. Natl. Acad. Sci. USA 1988, 85, 7972–7976. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Kinoshita, K.; Takenaka, S.; Suzuki, H.; Yamamoto, T.; Morohoshi, T.; Hayashi, M. Mycinamicin Biosynthesis: Isolation and Structural Elucidation of Novel Macrolactones and a Seco Acid Produced by a Mutant of Micromonospora Griseorubida. J. Chem. Soc. Chem. Commun. 1992, 10, 957–959. [Google Scholar] [CrossRef]
  49. Kinoshita, K.; Takenaka, S.; Hayashi, M. Isolation of Proposed Intermediates in the Biosynthesis of Mycinamicins. J. Chem. Soc. Chem. Commun. 1988, 14, 943–945. [Google Scholar] [CrossRef]
  50. Polborn, K.; Steglich, W.; Connolly, J.D.; Huneck, S. Structure of the Macrocyclic Bis-Lactone Lepranthin from the Lichen Arthonia Impolita; an X-ray Analysis. Z. Fur Nat. B 1995, 50, 1111–1114. [Google Scholar] [CrossRef]
  51. Schmitz, F.J.; Gunasekera, S.P.; Yalamanchili, G.; Hossain, M.B.; Van der Helm, D. Tedanolide: A Potent Cytotoxic Macrolide from the Caribbean Sponge Tedania Ignis. J. Am. Chem. Soc. 1984, 106, 7251–7252. [Google Scholar] [CrossRef]
  52. McCarthy, M. Tedania Ignis. Available online: https://animaldiversity.org/accounts/Tedania_ignis/ (accessed on 23 December 2022).
  53. Smith, A.B.; Lee, D. Total Synthesis of (+)-Tedanolide. J. Am. Chem. Soc. 2007, 129, 10957–10962. [Google Scholar] [CrossRef]
  54. Brufani, M.; Cerrini, S.; Fedeli, W.; Musu, C.; Cellai, L.; Keller-Schierlein, W. Structures of the Venturicidins A and B. Experientia 1971, 27, 604–606. [Google Scholar] [CrossRef]
  55. Lardy, H.A. Antibiotic Inhibitors of Mitochondrial Energy Transfer. Pharmacol. Ther. 1980, 11, 649–660. [Google Scholar] [CrossRef] [PubMed]
  56. Ishibashi, M.; Moore, R.E.; Patterson, G.M.L.; Xu, C.; Clardy, J. Scytophycins, Cytotoxic and Antimycotic Agents from the Cyanophyte Scytonema Pseudohofmanni. J. Org. Chem. 1986, 51, 5300–5306. [Google Scholar] [CrossRef]
  57. Moore, R.E.; Patterson, G.M.L.; Mynderse, J.S.; Barchi, J.; Norton, T.R.; Furusawa, E.; Furusawa, S. Toxins from Cyanophytes Belonging to the Scytonemataceae. Pure Appl. Chem. 1986, 58, 263–271. [Google Scholar] [CrossRef]
  58. Kubanek, J.; Jensen, P.R.; Keifer, P.A.; Sullards, M.C.; Collins, D.O.; Fenical, W. Seaweed Resistance to Microbial Attack: A Targeted Chemical Defense against Marine Fungi. Proc. Natl. Acad. Sci. USA 2003, 100, 6916–6921. [Google Scholar] [CrossRef] [Green Version]
  59. Blain, J.C.; Mok, Y.-F.; Kubanek, J.; Allingham, J.S. Two Molecules of Lobophorolide Cooperate to Stabilize an Actin Dimer Using Both Their “Ring” and “Tail” Region. Chem. Biol. 2010, 17, 802–807. [Google Scholar] [CrossRef] [Green Version]
  60. Fusetani, N.; Yasumuro, K.; Matsunaga, S.; Hashimoto, K. Mycalolides A–C, Hybrid Macrolides of Ulapualides and Halichondramide, from a Sponge of the Genus Mycale. Tetrahedron Lett. 1989, 30, 2809–2812. [Google Scholar] [CrossRef]
  61. Kernan, M.R.; Molinski, T.F.; Faulkner, D.J. Macrocyclic Antifungal Metabolites from the Spanish Dancer Nudibranch Hexabranchus Sanguineus and Sponges of the Genus Halichondria. J. Org. Chem. 1988, 53, 5014–5020. [Google Scholar] [CrossRef]
  62. Matsunaga, S.; Liu, P.; Celatka, C.A.; Panek, J.S.; Fusetani, N. Relative and Absolute Stereochemistry of Mycalolides, Bioactive Macrolides from the Marine Sponge Mycale Magellanica. J. Am. Chem. Soc. 1999, 121, 5605–5606. [Google Scholar] [CrossRef]
  63. Hori, M.; Saito, S.; Shin, Y.Z.; Ozaki, H.; Fusetani, N.; Karaki, H. Mycalolide-B, a Novel and Specific Inhibitor of Actomyosin ATPase Isolated from Marine Sponge. FEBS Lett. 1993, 322, 151–154. [Google Scholar] [CrossRef] [Green Version]
  64. Carmely, S.; Kashman, Y. Structure of Swinholide-a, a New Macrolide from the Marine Sponge Theonella Swinhoei. Tetrahedron Lett. 1985, 26, 511–514. [Google Scholar] [CrossRef]
  65. Kitagawa, I.; Kobayashi, M.; Katori, T.; Yamashita, M.; Tanaka, J.; Doi, M.; Ishida, T. Absolute Stereostructure of Swinholide A, a Potent Cytotoxic Macrolide from the Okinawan Marine Sponge Theonella Swinhoei. J. Am. Chem. Soc. 1990, 112, 3710–3712. [Google Scholar] [CrossRef]
  66. Frontali, L.; Tecce, G. Rifamycins. In Antibiotics: Volume I Mechanism of Action; Gottlieb, D., Shaw, P.D., Eds.; Springer: Berlin/Heidelberg, Germany, 1967; pp. 415–426. ISBN 978-3-662-38439-8. [Google Scholar]
  67. Sensi, P.; Greco, A.M.; Ballotta, R. Rifomycin. I. Isolation and Properties of Rifomycin B and Rifomycin Complex. Antibiot. Annu. 1959, 7, 262–270. [Google Scholar] [PubMed]
  68. Floss, H.G.; Yu, T.-W. RifamycinMode of Action, Resistance, and Biosynthesis. Chem. Rev. 2005, 105, 621–632. [Google Scholar] [CrossRef]
  69. Siminoff, P.; Smith, R.M.; Sokolski, W.T.; Savage, G. Streptovaricin: I. Discovery and Biologic Activity. Am. Rev. Tuberc. Pulmonary Dis. 1957, 75, 576–583. [Google Scholar]
  70. Milavetz, B.I.; Carter, W.A. Streptovaricins. In Pharmacology & Therapeutics, Part A, Chemotherapy, Toxicology and Metabolic Inhibitors; Pergamon Press: Oxford, UK, 1977; Volume 1, pp. 289–305. [Google Scholar]
  71. Wang, A.H.; Paul, I.C.; Rinehart Jr, K.L.; Antosz, F.J. Chemistry of the Streptovaricins. IX. X-Ray Crystallographic Structure of a Streptovaricin C Derivative. J. Am. Chem. Soc. 1971, 93, 6275–6276. [Google Scholar] [CrossRef] [PubMed]
  72. Miyashita, M.; Yamasaki, T.; Shiratani, T.; Hatakeyama, S.; Miyazawa, M.; Irie, H. The Efficient and Regioselective Synthesis of the Naphthoquinonecore of Streptovaricin U. Chem. Commun. 1997, 18, 1787–1788. [Google Scholar] [CrossRef]
  73. Mizuno, S.; Yamazaki, H.; Nitta, K.; Umezawa, H. Inhibition of RNA Polymerase of Escherichia Coli by an Antimicrobial Substance B44P (Streptovaricin). J. Antibiot. 1968, 21, 66–67. [Google Scholar] [CrossRef] [Green Version]
  74. Knöll, W.M.; Rinehart, K.L., Jr.; Willey, P.F.; LI, L.H. Streptovaricin U, an Acyclic Ansamycin. J. Antibiot. 1980, 33, 249–251. [Google Scholar] [CrossRef] [Green Version]
  75. Deshmukh, P.V.; Kakinuma, K.; Ameel, J.J.; Rinehart, K.L., Jr.; Wiley, P.F.; Li, L.H. Chemistry and Biochemistry of the Streptovaricins. XV. Protostreptovaricins IV. J. Am. Chem. Soc. 1976, 98, 870–872. [Google Scholar] [CrossRef]
  76. Umezawa, I.; Funayama, S.; Okada, K.; Iwasaki, K.; Satoh, J.; Masuda, K.; Komiyama, K. Studies on a Novel Cytocidal Antibiotic, Trienomycin A Taxonomy, Fermentation, Isolation, and Physico-Chemical and Biological Characteristics. J. Antibiot. 1985, 38, 699–705. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Funayama, S.; Okada, K.; Komiyama, K.; Umezawa, I. Structure of Trienomycin A, a Novel Cytocidal Ansamycin Antibiotic. J. Antibiot. 1985, 38, 1107–1109. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  78. Funayama, S.; Okada, K.; Iwasaki, K.; Komiyama, K.; Umezawa, I. Structures of Trienomycins A, B and C, Novel Cytocidal Ansamycin Antibiotics. J. Antibiot. 1985, 38, 1677–1683. [Google Scholar] [CrossRef]
  79. Hiramoto, S.; Ugita, M.; Ando, C.; Sasaki, T.; Furihata, K.; Seto, H.; Otake, N. Studies on Mycotrienin Antibiotics, a Novel Class of Ansamycins V. Isolation and Structure Determination of Novel Mycotrienin Congeners. J. Antibiot. 1985, 38, 1103–1106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Kobinata, K.; Koshino, H.; Kudo, T.; Isono, K.; Osada, H. RK-397, a New Oxo Pentaene Antibiotic. J. Antibiot. 1993, 46, 1616–1618. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  81. Koshino, H.; Kobinata, K.; Isono, K.; Osada, H. Structure of RK-397, a New Oxo Pentaene Antibiotic. J. Antibiot. 1993, 46, 1619–1621. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Schlegel, R.; Thrum, H. A New Polyene Antibiotic, Flavomycoin Structural Investigations. II. J. Antibiot. 1971, 24, 368–374. [Google Scholar] [CrossRef] [Green Version]
  83. Schlegel, R.; Thrum, H.; Zielinski, J.; Borowski, E. The Structure of Roflamycoin, a New Polyene Macrolide Antifungal Antibiotic. J. Antibiot. 1981, 34, 122–123. [Google Scholar] [CrossRef] [Green Version]
  84. Rychnovsky, S.D.; Griesgraber, G.; Schlegel, R. Stereochemical Determination of Roflamycoin: 13C Acetonide Analysis and Synthetic Correlation. J. Am. Chem. Soc. 1995, 117, 197–210. [Google Scholar] [CrossRef]
  85. Grigorjev, P.; Schlegel, R.; Thrum, H.; Ermishkin, L. Roflamycoin—A New Channel-Forming Antibiotic. Biochim. Biophys. Acta (BBA) Biomembr. 1985, 821, 297–304. [Google Scholar] [CrossRef]
  86. Cereghetti, D.M.C.; Erick, M. Amphotericin B: 50 Years of Chemistry and Biochemistry. Synthesis 2006, 2006, 0914–0942. [Google Scholar]
  87. Waugh, C.D. Amphotericin B. In xPharm: The Comprehensive Pharmacology Reference; Enna, S.J., Bylund, D.B., Eds.; Elsevier: New York, NY, USA, 2007; pp. 1–5. ISBN 978-0-08-055232-3. [Google Scholar]
  88. Cope, A.C.; Axen, U.; Burrows, E.P.; Weinlich, J. Amphotericin B. I. Carbon Skeleton, Ring Size, and Partial Structure1. J. Am. Chem. Soc. 1966, 88, 4228–4235. [Google Scholar] [CrossRef]
  89. Ganis, P.; Avitabile, G.; Mechlinski, W.; Schaffner, C.P. Polyene Macrolide Antibiotic Amphotericin B. Crystal Structure of the N-Iodoacetyl Derivative. J. Am. Chem. Soc. 1971, 93, 4560–4564. [Google Scholar] [CrossRef] [PubMed]
  90. Liu, Z.; Liu, H.; Zhang, W. Natural Polypropionates in 1999–2020: An Overview of Chemical and Biological Diversity. Mar. Drugs 2020, 18, 569. [Google Scholar] [CrossRef] [PubMed]
  91. Ruiz, C.; Valderrama, K.; Zea, S.; Castellanos, L. Mariculture and Natural Production of the Antitumoural (+)-Discodermolide by the Caribbean Marine Sponge Discodermia Dissoluta. Mar. Biotechnol. 2013, 15, 571–583. [Google Scholar] [CrossRef] [PubMed]
  92. Smith, A.B.; Freeze, B.S. (+)-Discodermolide: Total Synthesis, Construction of Novel Analogues, and Biological Evaluation. Tetrahedron 2007, 64, 261–298. [Google Scholar] [CrossRef] [Green Version]
  93. Gunasekera, S.P.; Gunasekera, M.; Longley, R.E.; Schulte, G.K. Discodermolide: A New Bioactive Polyhydroxylated Lactone from the Marine Sponge Discodermia Dissoluta. J. Org. Chem. 1990, 55, 4912–4915. [Google Scholar] [CrossRef]
  94. Brooks, H.A.; Gardner, D.; Poyser, J.P.; King, T. The Structure and Absolute Stereochemistry of Zincophorin (Antibiotic M144255): A Monobasic Carboxylic Acid Ionophore Having a Remarkable Specificity for Divalent Cations. J. Antibiot. 1984, 37, 1501–1504. [Google Scholar] [CrossRef] [Green Version]
  95. Danishefsky, S.J.; Selnick, H.G.; Zelle, R.E.; DeNinno, M.P. Total Synthesis of Zincophorin. J. Am. Chem. Soc. 1988, 110, 4368–4378. [Google Scholar] [CrossRef]
  96. Chuman, T.; Kohno, M.; Kato, K.; Noguchi, M. 4.6-Dimethyl-7-Hydroxy-Nonan-3-One, a Sex Pheromone of the Cigarette Beetle (Lasioderma Serricorne F.). Tetrahedron Lett. 1979, 20, 2361–2364. [Google Scholar] [CrossRef]
  97. Cheng, J.-F.; Lee, J.-S.; Sakai, R.; Jares-Erijman, E.A.; Silva, M.V.; Rinehart, K.L. Myriaporones 1−4, Cytotoxic Metabolites from the Mediterranean Bryozoan Myriapora Truncata. J. Nat. Prod. 2007, 70, 332–336. [Google Scholar] [CrossRef] [PubMed]
  98. Pérez, M.; del Pozo, C.; Reyes, F.; Rodríguez, A.; Francesch, A.; Echavarren, A.M.; Cuevas, C. Total Synthesis of Natural Myriaporones. Angew. Chem. Int. Ed. 2004, 43, 1724–1727. [Google Scholar] [CrossRef] [PubMed]
  99. Müller, S.; Rachid, S.; Hoffmann, T.; Surup, F.; Volz, C.; Zaburannyi, N.; Müller, R. Biosynthesis of Crocacin Involves an Unusual Hydrolytic Release Domain Showing Similarity to Condensation Domains. Chem. Biol. 2014, 21, 855–865. [Google Scholar] [CrossRef] [Green Version]
  100. Kunze, B.; Jansen, R.; Höfle, G.; Reichenbach, H. Crocacin, a New Electron Transport Inhibitor from Chondromyces Crocatus (Myxobacteria) Production, Isolation, Physico-Chemical and Biological Properties. J. Antibiot. 1994, 47, 881–886. [Google Scholar] [CrossRef]
  101. Jansen, R.; Washausen, P.; Kunze, B.; Reichenbach, H.; Höfle, G. The Crocacins, Novel Antifungal and Cytotoxic Antibiotics from Chondromyces Crocatus and Chondromyces Pediculatus (Myxobacteria): Isolation and Structure Elucidation. Eur. J. Org. Chem. 1999, 1999, 1085–1089. [Google Scholar] [CrossRef]
  102. Feutrill, J.T.; Lilly, M.J.; Rizzacasa, M.A. Total Synthesis of (+)-Crocacin D. Org. Lett. 2002, 4, 525–527. [Google Scholar] [CrossRef]
  103. Crowley, P.J.; Worthington, P.A.; Swanborough, J.; Sageot, O.-A.; Munns, G.R.; Devillers, I.M.; Godfrey, C.R.A.; Gillen, K.; Aspinall, I.H.; Williams, J. The Crocacins: Novel Natural Products as Leads for Agricultural Fungicides. Chimia 2003, 57, 685. [Google Scholar] [CrossRef]
  104. Ciavatta, M.L.; Gavagnin, M.; Puliti, R.; Cimino, G.; Martinez, E.; Ortea, J.; Mattia, C.A. Dolabriferol: A New Polypropionate from the Skin of the Anaspidean Mollusc Dolabrifera Dolabrifera. Tetrahedron 1996, 52, 12831–12838. [Google Scholar] [CrossRef]
  105. Jiménez-Romero, C.; González, K.; Rodríguez, A.D. Dolabriferols B and C, Non-Contiguous Polypropionate Esters from the Tropical Sea Hare Dolabrifera Dolabrifera. Tetrahedron Lett. 2012, 53, 6641–6645. [Google Scholar] [CrossRef] [Green Version]
  106. Sakuda, S.; Ono, M.; Furihata, K.; Nakayama, J.; Suzuki, A.; Isogai, A. Aflastatin A, a Novel Inhibitor of Aflatoxin Production of Aspergillus Parasiticus, from Streptomyces. J. Am. Chem. Soc. 1996, 118, 7855–7856. [Google Scholar] [CrossRef]
  107. Ikeda, H.; Matsumori, N.; Ono, M.; Suzuki, A.; Isogai, A.; Nagasawa, H.; Sakuda, S. Absolute Configuration of Aflastatin A, a Specific Inhibitor of Aflatoxin Production by Aspergillus Parasiticus. J. Org. Chem. 2000, 65, 438–444. [Google Scholar] [CrossRef] [PubMed]
  108. Sakuda, S.; Matsumori, N.; Furihata, K.; Nagasawa, H. Assignment of the Absolute Configuration of Blasticidin A and Revision of That of Aflastatin A. Tetrahedron Lett. 2007, 48, 2527–2531. [Google Scholar] [CrossRef]
  109. Rutkowski, J.; Brzezinski, B. Structures and Properties of Naturally Occurring Polyether Antibiotics. Biomed Res. Int. 2013, 2013, 162513. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Satake, M.; Shoji, M.; Oshima, Y.; Naoki, H.; Fujita, T.; Yasumoto, T. Gymnocin-A, a Cytotoxic Polyether from the Notorious Red Tide Dinoflagellate, Gymnodinium Mikimotoi. Tetrahedron Lett. 2002, 43, 5829–5832. [Google Scholar] [CrossRef]
  111. Sakai, T.; Matsushita, S.; Arakawa, S.; Mori, K.; Tanimoto, M.; Tokumasu, A.; Yoshida, T.; Mori, Y. Total Synthesis of Gymnocin-A. J. Am. Chem. Soc. 2015, 137, 14513–14516. [Google Scholar] [CrossRef]
  112. Roll, D.M.; Biskupiak, J.E.; Mayne, C.L.; Ireland, C.M. Muamvatin, a Novel Tricyclic Spiro Ketal from the Fijian Mollusk Siphonaria Normalis. J. Am. Chem. Soc. 1986, 108, 6680–6682. [Google Scholar] [CrossRef]
  113. Hoffmann, R.W.; Dahmann, G. The Absolute and Relative Configuration of Muamvatin. Tetrahedron Lett. 1993, 34, 1115–1118. [Google Scholar] [CrossRef]
  114. Paterson, I.; Perkins, M.V. Total Synthesis of the Marine Polypropionate (+)-Muamvatin. A Configurational Model for Siphonariid Metabolites. J. Am. Chem. Soc. 1993, 115, 1608–1610. [Google Scholar] [CrossRef]
  115. Suenaga, K.; Kigoshi, H.; Yamada, K. Auripyrones A and B, Cytotoxic Polypropionates from the Sea Hare Dolabella Auricularia: Isolation and Structures. Tetrahedron Lett. 1996, 37, 5151–5154. [Google Scholar] [CrossRef]
  116. Cheng, X.-C.; Kihara, T.; Kusakabe, H.; Magae, J.; Kobayashi, Y.; Fang, R.-P.; Ni, Z.-F.; Shent, Y.-C.; Keido, K.; Yamaguchi, I. A New Antibiotic, Tautomycin. J. Antibiot. 1987, 40, 907–909. [Google Scholar] [CrossRef] [Green Version]
  117. Suganuma, M.; Okabe, S.; Sueoka, E.; Nishiwaki, R.; Komori, A.; Uda, N.; Isono, K.; Fujiki, H. Tautomycin: An Inhibitor of Protein Phosphatases 1 and 2A but Not a Tumor Promoter on Mouse Skin and in Rat Glandular Stomach. J. Cancer Res. Clin. Oncol. 1995, 121, 621–627. [Google Scholar] [CrossRef] [PubMed]
  118. Cheng, X.-C.; Ubukata, M.; Isono, K. The Structure of Tautomycin, a Dialkylmaleic Anhydride Antibiotic. J. Antibiot. 1990, 43, 809–819. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Kato, Y.; Scheuer, P.J. The Aplysiatoxins. Pure Appl. Chem. 1975, 41, 1–14. [Google Scholar] [CrossRef] [Green Version]
  120. Moore, R.E.; Blackman, A.J.; Cheuk, C.E.; Mynderse, J.S.; Matsumoto, G.K.; Clardy, J.; Woodard, R.W.; Craig, J.C. Absolute Stereochemistries of the Aplysiatoxins and Oscillatoxin A. J. Org. Chem. 1984, 49, 2484–2489. [Google Scholar] [CrossRef]
  121. Fujiki, H.; Sugimura, T. New Classes of Tumor Promoters: Teleocidin, Aplysiatoxin, and Palytoxin. In Advances in Cancer Research; Klein, G., Weinhouse, S., Eds.; Academic Press: Cambridge, MA, USA, 1987; Volume 49, pp. 223–264. ISBN 0065-230X. [Google Scholar]
  122. Thale, Z.; Kinder, F.R.; Bair, K.W.; Bontempo, J.; Czuchta, A.M.; Versace, R.W.; Phillips, P.E.; Sanders, M.L.; Wattanasin, S.; Crews, P. Bengamides Revisited:  New Structures and Antitumor Studies. J. Org. Chem. 2001, 66, 1733–1741. [Google Scholar] [CrossRef] [PubMed]
  123. Fernández, R.; Dherbomez, M.; Letourneux, Y.; Nabil, M.; Verbist, J.F.; Biard, J.F. Antifungal Metabolites from the Marine Sponge Pachastrissa Sp.:  New Bengamide and Bengazole Derivatives. J. Nat. Prod. 1999, 62, 678–680. [Google Scholar] [CrossRef] [PubMed]
  124. Boonphong, S.; Kittakoop, P.; Isaka, M.; Pittayakhajonwut, D.; Tanticharoen, M.; Thebtaranonth, Y. Multiplolides A and B, New Antifungal 10-Membered Lactones from Xylaria Multiplex. J. Nat. Prod. 2001, 64, 965–967. [Google Scholar] [CrossRef]
  125. Ramana, C.V.; Khaladkar, T.P.; Chatterjee, S.; Gurjar, M.K. Total Synthesis and Determination of Relative and Absolute Configuration of Multiplolide A. J. Org. Chem. 2008, 73, 3817–3822. [Google Scholar] [CrossRef] [PubMed]
  126. Johnson, M.R.; Nakata, T.; Kishi, Y. Stereo- and Regioselective Methods for the Synthesis of Three Consecutive Asymmetric Units Found in Many Natural Products. Tetrahedron Lett. 1979, 20, 4343–4346. [Google Scholar] [CrossRef]
  127. Schmid, G.; Fukuyama, T.; Akasaka, K.; Kishi, Y. Synthetic Studies on Polyether Antibiotics. 4. Total Synthesis of Monensin. 1. Stereocontrolled Synthesis of the Left Half of Monensin. J. Am. Chem. Soc. 1979, 101, 259–260. [Google Scholar] [CrossRef]
  128. Hasan, I.; Kishi, Y. Further Studies on Stereospecific Epoxidation of Allylic Alcohols. Tetrahedron Lett. 1980, 21, 4229–4232. [Google Scholar] [CrossRef]
  129. Nagaoka, H.; Rutsch, W.; Schmid, G.; Iio, H.; Johnson, M.R.; Kishi, Y. Total Synthesis of Rifamycins. 1. Stereocontrolled Synthesis of the Aliphatic Building Block. J. Am. Chem. Soc. 1980, 102, 7962–7965. [Google Scholar] [CrossRef]
  130. Iio, H.; Nagaoka, H.; Kishi, Y. Total Synthesis of Rifamycins. 2. Total Synthesis of Racemic Rifamycin S. J. Am. Chem. Soc. 1980, 102, 7965–7967. [Google Scholar] [CrossRef]
  131. Nagaoka, H.; Kishi, Y. Further Synthetic Studies on Rifamycin s. Tetrahedron 1981, 37, 3873–3888. [Google Scholar] [CrossRef]
  132. Tino, J.A.; Lewis, M.D.; Kishi, Y. A New, Efficient Synthesis of the Left Half of Narasin. Heterocyles 1987, 25, 97–104. [Google Scholar]
  133. Katsuki, T.; Hanamoto, T.; Yamaguchi, M. Synthesi of C19–C27 Fragment of Ansa Chain Part of Rifamycin S. Chem. Lett. 1989, 18, 117–118. [Google Scholar] [CrossRef]
  134. Okamura, H.; Kuroda, S.; Ikegami, S.; Ito, Y.; Katsuki, T.; Yamaguchi, M. A Formal Total Synthesis of Aplysiatoxin. Tetrahedron Lett. 1991, 32, 5141–5142. [Google Scholar] [CrossRef]
  135. Marshall, J.A.; Palovich, M.R. Synthesis of Stereopentad Subunits of Zincophorin and Rifamycin-S through Use of Chiral Allenyltin Reagents. J. Org. Chem. 1998, 63, 3701–3705. [Google Scholar] [CrossRef]
  136. Marshall, J.A.; Johns, B.A. Total Synthesis of (+)-Discodermolide. J. Org. Chem. 1998, 63, 7885–7892. [Google Scholar] [CrossRef]
  137. Marshall, J.A.; Adams, N.D. Progress toward the Total Synthesis of Bafilomycin A1: Stereoselective Synthesis of the C15−C25 Subunit by Additions of Nonracemic Allenylzinc Reagents to Aldehydes. Org. Lett. 2000, 2, 2897–2900. [Google Scholar] [CrossRef]
  138. Miyashita, M.; Hoshino, M.; Yoshikoshi, A. Stereospecific Methylation of γ,δ-Epoxy Acrylates by Trimethylaluminum: A Method for the Iterative Construction of Polypropionate Chains. J. Org. Chem. 1991, 56, 6483–6485. [Google Scholar] [CrossRef]
  139. Miyashita, M.; Hoshino, M.; Yoshikoshi, A.; Kawamine, K.; Yoshihara, K.; Irie, H. Synthesis of the Prelog-Djerassi Lactone and Protomycinolide IV Based on the Stereospecific Methylation of γ,δ-Epoxy Acrylates by Trimethylaluminum. Chem. Lett. 1992, 21, 1101–1104. [Google Scholar] [CrossRef]
  140. Miyashita, M.; Toshimitsu, Y.; Shiratani, T.; Irie, H. Enantioselective Synthesis of (−)-Serrieornina, Sex Pheromone of a Female Cigarette Beetle (Lasioderma Serricorne F.). Tetrahedron Asymmetry 1993, 4, 1573–1578. [Google Scholar] [CrossRef]
  141. Miyashita, M.; Yoshihara, K.; Kawamine, K.; Hoshino, M.; Irie, H. Synthetic Studies on Polypropionate Antibiotics Based on the Stereospecific Methylation of γ δ-Epoxy Acrylates by Trimethylaluminum. A Highly Stereoselective Construction of the Eight Contiguous Chiral Centers of Ansa-Chains of Rifamycins. Tetrahedron Lett. 1993, 34, 6285–6288. [Google Scholar] [CrossRef]
  142. Miyashita, M.; Shiratani, T.; Kawamine, K.; Hatakeyama, S.; Irie, H. Tandem Synthesis of Polypropionate Chains–Highly Stereoselective Synthesis of the Ansa Chain of Streptovaricin U and Protostreptovaricins Based on Stereospecific Methylation of γ,δ-Epoxy Acrylates by Trimethylaluminium. Chem. Commun. 1996, 9, 1027–1028. [Google Scholar] [CrossRef]
  143. Grieco, P.A.; Speake, J.D.; Yeo, S.K.; Miyashita, M. Synthesis of the C(19)-C(32) Fragment of Scytophycin C via Stereospecific Methylation of γ,δ-Epoxy Acrylates with Trimethylaluminum. Tetrahedron Lett. 1998, 39, 1125–1128. [Google Scholar] [CrossRef]
  144. Nakamura, R.; Tanino, K.; Miyashita, M. Total Synthesis of Scytophycin C. 1. Stereoselective Syntheses of the C(1)−C(18) Segment and the C(19)−C(31) Segment. Org. Lett. 2003, 5, 3579–3582. [Google Scholar] [CrossRef]
  145. Nakamura, R.; Tanino, K.; Miyashita, M. Total Synthesis of Scytophycin C. 2. Coupling Reaction of the C(1)−C(18) Segment and the C(19)−C(31) Segment, a Key Macrolactonization, and the Crucial Terminal Amidation Reaction. Org. Lett. 2003, 5, 3583–3586. [Google Scholar] [CrossRef]
  146. Hayakawa, H.; Miyashita, M. A Tandem Synthesis of Polypropionate Chains. Highly Stereo- Selective Construction of the C(13)–C(25) Segment Containing Nine Contiguous Chiral Centers of Swinholides A–C Based on the Stereo-Specific Methylation of, -Epoxy Acrylates by Trimethylaluminium. J. Chem. Perkin Trans. 1999, 23, 3399–3401. [Google Scholar] [CrossRef]
  147. Iwata, Y.; Tanino, K.; Miyashita, M. Synthetic Studies of Tedanolide, a Marine Macrolide Displaying Potent Antitumor Activity. Stereoselective Synthesis of the C(13)−C(23) Segment. Org. Lett. 2005, 7, 2341–2344. [Google Scholar] [CrossRef]
  148. Nagumo, S.; Miyashita, M.; Takada, H.; Yasui, E.; Mizukami, M. Synthetic Studies of Lepranthin, a Lichen-Produced Dimeric Macrolide. Stereoselective Synthesis of a Seco-Acid Based on Stereospecific Epoxide-Opening Reactions. Heterocycles 2011, 83, 293. [Google Scholar] [CrossRef]
  149. Suzuki, Y.; Nagumo, S.; Yasui, E.; Mizukami, M.; Miyashita, M. Synthetic Studies of Venturicidins: Stereoselective Synthesis of the C15–C27 Segment Based on Two Types of Stereospecific Epoxide Opening Reactions. Tetrahedron Lett. 2011, 52, 6948–6951. [Google Scholar] [CrossRef]
  150. Maruyama, K.; Ueda, M.; Sasaki, S.; Iwata, Y.; Miyazawa, M.; Mlyashita, M. Highly Stereoselective Epoxidation of a 4-Methyl-5-(Triethylsilyi)Oxyallyl Alcohol System with m.Chloroperoxybenzoic Acid. Tetrahedron Lett. 1998, 39, 4517–4520. [Google Scholar] [CrossRef]
  151. Miyashita, M.; Suzuki, T.; Hoshino, M.; Yoshikoshi, A. The Organoselenium-Mediated Reduction of α,β-Epoxy Ketones, α,β-Epoxy Esters, and Their Congeners to β-Hydroxy Carbonyl Compounds: Novel Methodologies for the Synthesis of Aldols and Their Analogues. Tetrahedron 1997, 53, 12469–12486. [Google Scholar] [CrossRef]
  152. Torres, W.; Torres, G.; Prieto, J.A. Synthesis of Stereotetrads by Regioselective Cleavage of Diastereomeric MEM-Protected 2-Methyl-3,4-Epoxy Alcohols with Diethylpropynyl Aluminum. Synlett 2012, 23, 2534–2538. [Google Scholar] [PubMed] [Green Version]
  153. Torres, G.; Torres, W.; Prieto, J.A. Microwave-Assisted Epoxidation of Hindered Homoallylic Alcohols Using VO(acac)2: Application to Polypropionate Synthesis. Tetrahedron 2004, 60, 10245–10251. [Google Scholar] [CrossRef]
  154. Katsuki, T.; Sharpless, K.B. The First Practical Method for Asymmetric Epoxidation. J. Am. Chem. Soc. 1980, 102, 5974–5976. [Google Scholar] [CrossRef]
  155. Torres, W.; Rodríguez, R.R.; Prieto, J.A. Stereoselective Construction of All-Anti Polypropionate Modules: Synthesis of the C5− C10 Fragment of Streptovaricin U. J. Org. Chem. 2009, 74, 2447–2451. [Google Scholar] [CrossRef] [Green Version]
  156. Tirado, R.; Prieto, J.A. Stereochemistry of the Iodocarbonatation of cis-and trans-3-Methyl-4-Pentene-1, 2-Diols: The Unusual Formation of Several Anti Iodo Carbonates. J. Org. Chem. 1993, 58, 5666–5673. [Google Scholar] [CrossRef]
  157. Rodríguez-Berríos, R.R.; Torres, G.; Prieto, J.A. Stereoselective VO (Acac) 2 Catalyzed Epoxidation of Acyclic Homoallylic Diols. Complementary Preparation of C2-Syn-3, 4-Epoxy Alcohols. Tetrahedron 2011, 67, 830–836. [Google Scholar] [CrossRef] [Green Version]
  158. Rodríguez, D.; Mulero, M.; Prieto, J.A. Highly Regioselective Copper-Catalyzed Cis-and Trans-1-Propenyl Grignard Cleavage of Hindered Epoxides. Application in Propionate Synthesis. J. Org. Chem. 2006, 71, 5826–5829. [Google Scholar] [CrossRef] [PubMed]
  159. Prieto, J.A.; Torres, J.R.; Rodríguez-Berrios, R. Regiocontrolled Ring Opening of Monoprotected 2,3-Epoxy-1,4-Diols by Using Alkynyl Aluminum Reagents: Synthesis of Differentially Monoprotected Alkynyl Triol Derivatives. Synlett 2014, 25, 433–437. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Dávila, W.; Torres, W.; Prieto, J.A. Regioselective Cleavage of 3,4-Epoxy Alcohols with Substituted Alkynylaluminum Reagents: Application to the Stereoselective Synthesis of Polypropionates. Tetrahedron 2007, 63, 8218–8226. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  161. Cruz-Montañez, A.; Morales-Rivera, K.F.; Torres, W.; Valentín, E.M.; Rentas-Torres, J.; Prieto, J.A. Reiterative Epoxide-Based Strategies for the Synthesis of Stereo-n-Ads and Application to Polypropionate Synthesis. A Personal Account. Inorg. Chim. Acta 2017, 468, 28–37. [Google Scholar] [CrossRef] [PubMed]
  162. Valentín, E.M.; Mulero, M.; Prieto, J.A. Concise Epoxide-Based Synthesis of the C14–C25 Bafilomycin A1 Polypropionate Chain. Tetrahedron Lett. 2012, 53, 2199–2201. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  163. Sarabia, F.; Martín-Gálvez, F.; García-Castro, M.; Chammaa, S.; Sánchez-Ruiz, A.; Tejón-Blanco, J.F. Epoxyamide-Based Strategy for the Synthesis of Polypropionate-Type Frameworks. J. Org. Chem. 2008, 73, 8979–8986. [Google Scholar] [CrossRef]
  164. Sarabia, F.; Sánchez-Ruiz, A.; Martín-Ortiz, L.; García-Castro, M.; Chammaa, S. Stereoselective Synthesis of Macrolide-Type Antibiotics from Epoxy Amides. Synthesis of the Polypropionate Chain of Streptovaricin U. Org. Lett. 2007, 9, 5091–5094. [Google Scholar] [CrossRef]
  165. Sarabia, F.; Martin-Galvez, F.; Chammaa, S.; Martin-Ortiz, L.; Sanchez-Ruiz, A. Chiral Sulfur Ylides for the Synthesis of Bengamide E and Analogues. J. Org. Chem. 2010, 75, 5526–5532. [Google Scholar] [CrossRef]
  166. Sarabia, F.; Martin-Galvez, F.; Garcia-Ruiz, C.; Sanchez-Ruiz, A.; Vivar-Garcia, C. Epi-, Epoxy-, and C2-Modified Bengamides: Synthesis and Biological Evaluation. J. Org. Chem. 2013, 78, 5239–5253. [Google Scholar] [CrossRef]
  167. Jung, M.E.; Marquez, R. Rearrangement of Epoxides in Non-Aldol Aldol Process: Allylic vs. Tertiary and Secondary Carbocationic Centers. Tetrahedron Lett. 1999, 40, 3129–3132. [Google Scholar] [CrossRef]
  168. Jung, M.E.; Salehi-Rad, R. Total Synthesis of Auripyrone A Using a Tandem Non-aldol Aldol/Paterson Aldol Process as a Key Step. Angew. Chem. Int. Ed. 2009, 48, 8766–8769. [Google Scholar] [CrossRef]
  169. Jung, M.E.; Chaumontet, M.; Salehi-Rad, R. Total Synthesis of Auripyrone B Using a Non-Aldol Aldol—Cuprate Opening Process. Org. Lett. 2010, 12, 2872–2875. [Google Scholar]
  170. Burova, S.A.; McDonald, F.E. Stereoselective Synthesis of a Family of Alternating Polyols from Six-Carbon Epoxyalkynol Modules. J. Am. Chem. Soc. 2002, 124, 8188–8189. [Google Scholar] [CrossRef] [PubMed]
  171. Robles, O.; McDonald, F.E. Modular Synthesis of the C9–C27 Degradation Product of Aflastatin A via Alkyne−Epoxide Cross-Couplings. Org. Lett. 2008, 10, 1811–1814. [Google Scholar] [CrossRef]
  172. Lipshutz, B.H.; Kozlowski, J.A. A Reiterative Route to Chiral All-Syn-1,3-Polyols. J. Org. Chem. 1984, 49, 1147–1149. [Google Scholar] [CrossRef]
  173. Lipshutz, B.H.; Kozlowski, J.A.; Parker, D.A.; Nguyen, S. More Highly Mixed, Higher Order Cyanocuprates “RT(24bienyl)Cu(CN)Li”. Efficient Reagents Which Promote Selective Ligand Transfer. J. Organomet. Chem. 1985, 285, 431–447. [Google Scholar] [CrossRef]
  174. Lipshutz, B.H.; Barton, J.C. Acyclic Control of Stereochemistry via a Reiterative (E or Z)-1-Propenyllithium-Derived Cuprate Opening of a Chiral Epoxide/Reepoxidation Sequence. J. Org. Chem. 1988, 53, 4495–4499. [Google Scholar] [CrossRef]
  175. Lipshutz, B.H.; Moretti, R.; Crow, R. Progress toward Roflamycoin; Synthesis of the C-12 to C-35 Section in Homochiral Form. Tetrahedron Lett. 1989, 30, 15–18. [Google Scholar] [CrossRef]
  176. Nicolaou, K.C.; Uenishi, J. Stereocontrolled Synthesis of 1, 3, 5… (2 N+ 1) Polyols. J. Chem. Soc. Chem. Commun. 1982, 1985, 1292–1293. [Google Scholar]
  177. Nicolaou, K.C.; Daines, R.A.; Uenishi, J.; Li, W.S.; Papahatjis, D.P.; Chakraborty, T.K. Total Synthesis of Amphoteronolide B and Amphotericin B. 1. Strategy and Stereocontrolled Construction of Key Building Blocks. J. Am. Chem. Soc. 1988, 110, 4672–4685. [Google Scholar] [CrossRef]
  178. Nicolaou, K.C.; Daines, R.A.; Chakraborty, T.K.; Ogawa, Y. Total Synthesis of Amphoteronolide B and Amphotericin B. 2. Total Synthesis of Amphoteronolide B. J. Am. Chem. Soc. 1988, 110, 4685–4696. [Google Scholar] [CrossRef]
  179. Corey, E.J.; Hase, T. Studies on the Total Synthesis of Rifamycin Highly Stereo Selective Synthesis of Intermediates for Construction of the Carbon 15 to Carbon 29 Chain. Tetrahedron Lett. 1979, 20, 335–338. [Google Scholar] [CrossRef]
  180. Corey, E.J.; Pan, B.C.; Hua, D.H.; Deardorff, D.R. Total Synthesis of Aplasmomycin. Stereocontrolled Construction of the C(3)-C(17) Fragment. J. Am. Chem. Soc. 1982, 104, 6816–6818. [Google Scholar] [CrossRef]
  181. Lipshutz, B.H.; Floyd, D.M.; Wilhelm, R.S. Chemistry of Higher Order, Mixed Organocuprates. 1. Substitution Reactions at Unactivated Secondary Centers. J. Am. Chem. Soc. 1981, 103, 7672–7674. [Google Scholar] [CrossRef]
  182. Lipshutz, B.H. Applications of Higher-Order Mixed Organocuprates to Organic Synthesis. Synthesis 1987, 1987, 325–341. [Google Scholar] [CrossRef]
  183. Lipshutz, B.H.; Kozlowski, J.; Wilhelm, R.S. Chemistry of Higher Order Mixed Organocuprates. 2. Reactions of Epoxides. J. Am. Chem. Soc. 1982, 104, 2305–2307. [Google Scholar] [CrossRef]
  184. Lipshutz, B.H.; Wilhelm, R.S.; Kozlowski, J.A.; Parker, D. Substitution Reactions of Secondary Halides and Epoxides with Higher Order, Mixed Organocuprates, R2Cu(CN)Li2: Synthetic, Stereochemical, and Mechanistic Aspects. J. Org. Chem. 1984, 49, 3928–3938. [Google Scholar] [CrossRef]
  185. Lipshutz, B.H.; Wilhelm, R.S.; Kozlowski, J.A. The Chemistry of Higher Order Organocuprates. Tetrahedron 1984, 40, 5005–5038. [Google Scholar] [CrossRef]
  186. Tone, H.; Hikota, M.; Hamada, T.; Nishi, T.; Oikawa, Y.; Yonemitzu, O. Alternative Syntheses of the C9–C15 and C1–C5 Segments of Erythronolide A via Regio- and Stereo-Selective Reductive Ring Opening of 2, 3-Epoxy Alcohols. Chem. Pharm. Bull. 1989, 37, 1155–1159. [Google Scholar] [CrossRef] [Green Version]
  187. Wang, Z.; Schreibet, S.L. Stereochemical Studies of Epoxides Relevant to the Ansa Chain of the Streptovaricins. Tetrahedron Lett. 1990, 31, 31–34. [Google Scholar] [CrossRef]
  188. Yadav, J.S.; Praveen, K.T.K.; Maniyan, P.P. Stereoconvergent Synthesis of C-9 to C-16 Fragment of Trienomycin Based on the Regioselective Opening of γ-δ-Poxy Acrylates with Trimethylaluminium. Tetrahedron Lett. 1993, 34, 2969–2972. [Google Scholar] [CrossRef]
  189. Kong, L.; Zhuang, Z.; Chen, Q.; Deng, H.; Tang, Z.; Jia, X.; Li, Y.; Zhai, H. A Facile Asymmetric Synthesis of (+)-Eldanolide. Tetrahedron Asymmetry 2007, 18, 451–454. [Google Scholar] [CrossRef]
  190. Haddad, N.; Brik, A.; Grishko, M. Studies towards Total Synthesis of Borrelidin, Regioselective Methylation of Bis-Epoxides and Structure Determination. Tetrahedron Lett. 1997, 38, 6079–6082. [Google Scholar] [CrossRef]
  191. Kobayashi, Y.; Uchiyama, H.; Kanbara, H.; Sato, F. Stereoselective Synthesis of Gamma, Delta-Epoxy-Beta-Methyl-Gamma-(Trimethylsilyl)Alkanols. Synthesis of the C(1)–C(7) Segment of 6-Deoxyerythronolide B. J. Am. Chem. Soc. 1985, 107, 5541–5543. [Google Scholar] [CrossRef]
  192. Mulzer, J.; Schöllhorn, B. Enantio- and Regiocontrolled Synthesis of a Central Ionophoric Antibiotic Building Block by Sequential Opening of Two Epoxide Rings with Cuprate Reagents. Angew. Chem. Int. Ed. Engl. 1990, 29, 1476–1478. [Google Scholar] [CrossRef]
  193. Shang, S.; Iwadare, H.; Macks, D.E.; Ambrosini, L.M.; Tan, D.S. A Unified Synthetic Approach to Polyketides Having Both Skeletal and Stereochemical Diversity. Org. Lett. 2007, 9, 1895–1898. [Google Scholar] [CrossRef] [Green Version]
  194. Bandaru, A.; Kaliappan, K.P. Synthesis of the C 1–C 10 Fragment of Muamvatin. Chem. Asian J. 2020, 15, 2208–2211. [Google Scholar] [CrossRef] [PubMed]
  195. Burova, S.A.; McDonald, F.E. Total Synthesis of the Polyene-Polyol Macrolide RK-397, Featuring Cross-Couplings of Alkynylepoxide Modules. J. Am. Chem. Soc. 2004, 126, 2495–2500. [Google Scholar] [CrossRef]
  196. Reddy, B.C.; Bangade, V.M.; Ramesh, P.; Meshram, H.M. Stereoselective Total Synthesis of Multiplolide A and of a Diastereoisomer. Helv. Chim. Acta 2013, 96, 266–274. [Google Scholar] [CrossRef]
  197. Tsuboi, K.; Ichikawa, Y.; Naganawa, A.; Isobe, M.; Ubukata, M.; Isono, K. Synthetic Studies on Tautomycin Synthesis of Segment B. Tetrahedron 1997, 53, 5083–5102. [Google Scholar] [CrossRef]
  198. Van Dyke, A.R.; Jamison, T.F. Functionalized Templates for the Convergent Assembly of Polyethers: Synthesis of the HIJK Rings of Gymnocin A. Angew. Chem. Int. Ed. 2009, 48, 4430–4432. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  199. Evans, D.A.; Polniaszek, R.P.; DeVries, K.M.; Guinn, D.E.; Mathre, D.J. Synthetic Studies in the Lysocellin Family of Polyether Zyxwvu Antibiotics. The Total Synthesis of Ferensimycin B. J. Am. Chem. Soc. 1991, 113, 7613–7630. [Google Scholar] [CrossRef]
  200. Mori, Y.; Takeuchi, A.; Kageyama, H.; Suzuki, M. A Convergent General Synthetic Protocol for Syn-1,3-Polyols. Tetrahedron Lett. 1988, 29, 5423–5426. [Google Scholar] [CrossRef]
  201. Ma, P.; Martin, V.S.; Masamune, S.; Sharpless, K.B.; Viti, S.M. Synthesis of Saccharides and Related Polyhydroxylated Natural Products. 2. Simple Deoxyalditols. J. Org. Chem. 1982, 47, 1378–1380. [Google Scholar] [CrossRef]
Figure 1. Types of polyketides in natural products.
Figure 1. Types of polyketides in natural products.
Ijms 24 06195 g001
Figure 2. General polypropionate structure (i.e., stereooctad).
Figure 2. General polypropionate structure (i.e., stereooctad).
Ijms 24 06195 g002
Figure 3. Methods for the stereoselective synthesis of polypropionates.
Figure 3. Methods for the stereoselective synthesis of polypropionates.
Ijms 24 06195 g003
Figure 4. Structures of reported macrolides 113.
Figure 4. Structures of reported macrolides 113.
Ijms 24 06195 g004
Figure 5. Structures of the reported macrolides 1420.
Figure 5. Structures of the reported macrolides 1420.
Ijms 24 06195 g005
Figure 6. Structures of the reported linear polypropionates 2127.
Figure 6. Structures of the reported linear polypropionates 2127.
Ijms 24 06195 g006
Figure 7. Structures of the reported polypropionates 2834.
Figure 7. Structures of the reported polypropionates 2834.
Ijms 24 06195 g007
Figure 8. Kishi’s allylic epoxidation-based approach toward polypropionates.
Figure 8. Kishi’s allylic epoxidation-based approach toward polypropionates.
Ijms 24 06195 g008
Figure 9. Katsuki’s dithianative epoxide-ring opening approach toward polypropionates.
Figure 9. Katsuki’s dithianative epoxide-ring opening approach toward polypropionates.
Ijms 24 06195 g009
Figure 10. Marshall’s allenylstannane epoxide-based approach toward polypropionates.
Figure 10. Marshall’s allenylstannane epoxide-based approach toward polypropionates.
Ijms 24 06195 g010
Figure 11. Miyashita’s trimethylaluminum epoxide-ring opening toward polypropionates.
Figure 11. Miyashita’s trimethylaluminum epoxide-ring opening toward polypropionates.
Ijms 24 06195 g011
Figure 12. Prieto’s three step epoxide-based reiterative approach for polypropionate construction and the stereoselective synthesis of the C5–C12 fragment of streptovaricin U.
Figure 12. Prieto’s three step epoxide-based reiterative approach for polypropionate construction and the stereoselective synthesis of the C5–C12 fragment of streptovaricin U.
Ijms 24 06195 g012
Figure 13. Polypropioante fragments and natural products reported by Prieto’s group.
Figure 13. Polypropioante fragments and natural products reported by Prieto’s group.
Ijms 24 06195 g013
Figure 14. Sarabia’s methodology for the synthesis of polypropionates.
Figure 14. Sarabia’s methodology for the synthesis of polypropionates.
Ijms 24 06195 g014
Figure 15. Jung methodology for the synthesis of polypropionate.
Figure 15. Jung methodology for the synthesis of polypropionate.
Ijms 24 06195 g015
Figure 16. McDonald epoxide-ring opening toward polypropionates.
Figure 16. McDonald epoxide-ring opening toward polypropionates.
Ijms 24 06195 g016
Figure 17. Lipshutz methodology for the synthesis of polypropionate.
Figure 17. Lipshutz methodology for the synthesis of polypropionate.
Ijms 24 06195 g017
Figure 18. Nicolaou’s methodology for the synthesis of polyols.
Figure 18. Nicolaou’s methodology for the synthesis of polyols.
Ijms 24 06195 g018
Figure 19. Natural products synthesized following the epoxide-based approaches (several authors).
Figure 19. Natural products synthesized following the epoxide-based approaches (several authors).
Ijms 24 06195 g019
Table 1. Summary of the formal and total syntheses of polypropionates and related natural products by research group (chronological order).
Table 1. Summary of the formal and total syntheses of polypropionates and related natural products by research group (chronological order).
GroupYearFragmentNatural ProductRef.
Kishi1980C29–C15Rifamycin S 15[129]
1980Total synthesisRifamycin S 15[130]
1981C29–C15 2nd generationRifamycin S 15[131]
1987Left halfNarasin[132]
Katsuki1989C19–C27 (Kishi’s intermediate)Rifamycin S 15[133]
1991C1–C21 (Kishi’s intermediate)Aplysiatoxin 32[134]
Marshall1998C7–C13Zincophorin 22[135]
C21–C27Rifamycin S 15
1998Total synthesis(+)-Discodermolide 21[136]
2000C15–C25Bafilomycin A1 6[137]
Miyashita1992Total synthesisPrelog-Djerassi lactone 2[139]
Formal synthesisProtomycinolide IV 7
1993Total synthesis(−)-Serricornin 23[140]
1993C27–C20Rifamycins 15[141]
1996C17–C5Streptovaricin U 16e[142]
C17–C5Protostreptovaricin I & II (16c-d)
1998C19–C32Scytophycin C 11[143]
1999C13–C25Swinholides A-C 14a-c[146]
2003C1–C18Scytophycin C 11[144]
2003Total synthesisScytophycin C 11[145]
2005C13–C23Tedanolide 9[147]
2011C1–C10Lepranthin 8[148]
2011C15–C17Venturicidins 10[149]
Prieto2004C5–C10Streptovaricin D 16b[153]
C5–C12Streptovaricin U 16e
2005C15–C10Streptovaricin D 16b[158]
C24–C27Rifamycin S 15
2007C5–C10Elaiophylin 5[160]
2009C6–C10Streptovaricin U 16e[155]
2011C1–C4/C12–C15Lankanolide 4b[157]
2012C14–C25Bafilomycin A1 6[162]
2014C14–C17Tedanolide 9[159]
C6–C9Myraporone ¾ (24)
2017C20–C15Crocacins 24a-d[161]
C3–C9/C12–C18Dolabriferol B 26b
C34–C22 and C22–C27Mycalolide A 13
C19–C24Lobophorolide 12
C5–C12 and C15–C8Streptovaricin D 16b
C12–C16Elaiophylin 5
Sarabia2007C1–C13Streptovaricin U 16e[164]
2010 and 2013Total synthesisBengamide E 33 [165,166]
Jung2010Total synthesisAuripyrone B 30b[169]
McDonald2004C28–C11RK-397 (18)[195]
2008C9–C27Aflastatin A 27[171]
Corey1979C15–C29Rifamycin S 15[179]
Yonemitzu1989C9–C15 and C1–C5Erythronolide A 3b[186]
Lipshutz1989C12–C35Roflamycoin 19[176]
Schreiber1990C5–C15Streptovaricin A 16a[187]
Evans1991Total synthesisFerensimycin B 176[196]
Yadav1993C9–C16Trienomycin 17[191]
Isobe1997C26–C27Tautomycin 1 (31)[197]
Nicolaou 1998Fragments 170, 171, 172,
and 173
Amphoterolide B 20b & Amphotericin B 20a[177]
1998Total synthesisAmphoterolide B 20b & Amphotericin B 20a[178]
Jamison2009HIJK ringsGymnocin A 28[198]
Meshram2013Total synthesisMultiplolide A 34[196]
Kaliappan2020C1–C10Muamvatin 29[194]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Rodríguez-Berríos, R.R.; Isbel, S.R.; Bugarin, A. Epoxide-Based Synthetic Approaches toward Polypropionates and Related Bioactive Natural Products. Int. J. Mol. Sci. 2023, 24, 6195. https://doi.org/10.3390/ijms24076195

AMA Style

Rodríguez-Berríos RR, Isbel SR, Bugarin A. Epoxide-Based Synthetic Approaches toward Polypropionates and Related Bioactive Natural Products. International Journal of Molecular Sciences. 2023; 24(7):6195. https://doi.org/10.3390/ijms24076195

Chicago/Turabian Style

Rodríguez-Berríos, Raúl R., Stephen R. Isbel, and Alejandro Bugarin. 2023. "Epoxide-Based Synthetic Approaches toward Polypropionates and Related Bioactive Natural Products" International Journal of Molecular Sciences 24, no. 7: 6195. https://doi.org/10.3390/ijms24076195

APA Style

Rodríguez-Berríos, R. R., Isbel, S. R., & Bugarin, A. (2023). Epoxide-Based Synthetic Approaches toward Polypropionates and Related Bioactive Natural Products. International Journal of Molecular Sciences, 24(7), 6195. https://doi.org/10.3390/ijms24076195

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop