TY - JOUR AU - Gong, Liu-Zhu AB - Abstract Indole and its structural analogues have been frequently found in numerous alkaloids, pharmaceutical products and related materials. The enantioselective construction of these structures allows efficient total synthesis of optically pure indole alkaloids, and hence has received worldwide interest. In the past decade, asymmetric organocatalysis has been recognized as one of the most powerful strategies to create chiral molecules with high levels of stereoselectivity. In particular, organocatalytic asymmetric cascade reactions often occur with multiple bond-breaking and forming events simultaneously or sequentially, leading to the appearance of various straightforward approaches to access core structures for asymmetric total synthesis. This review will summarize recent applications of asymmetric organocatalysis in the enantioselective synthesis of indole alkaloids. asymmetric catalysis, organocatalysis, indole alkaloids, total synthesis INTRODUCTION Indole alkaloids constitute a large family of important natural and pharmaceutical substances with a diverse range of structural complexity and biological relevance [1]. These alkaloids are often installed with complex structures that are of great challenge to be accessed. Furthermore, the issue to control stereochemistry in the synthesis of a specific target molecule has been considered particularly difficult. As such, the development of new synthetic methods to stereoselectively access core structural skeletons that allow the enantioselective total synthesis of indole alkaloids represents an ambition in chemical synthesis [2,3]. Circumventing the current challenges existing in this highly valuable field basically relies on the development of new catalytic bond-forming strategies to prepare optically pure fragments that enhance efficiency and provide concise approaches to access target molecules [4–95]. Over the past decade, organocatalysis has been included among the most successful concepts in asymmetric catalysis, and it has been used for the enantioselective formation of new chemical bonds with several advantages from an economical and environmental point of view [10–135]. In addition to methodology development, impressive advances have also been made on the organocatalytic asymmetric total synthesis. Previous elegant reviews have summarized the progress and perspective of this field [14–335]. This review will focus on recent representative examples describing asymmetric organocatalytic synthesis of core structural skeletons or intermediates that turn out to be key issues to build up indole alkaloids with excellent efficiency and stereoselectivity (Fig. 1). These examples will be described in different sections grouped on the basis of the core structures, which are accessed by using a conceptually diverse range of organocatalytic asymmetric reactions. Figure 1. Open in new tabDownload slide Representative organocatalysis for the total synthesis of indole alkaloids. (a) A crucial role of organocatalysed reactions applied in the key step of total synthesis. (b) Organocatalysts involved in total synthesis. (c) Representative indole alkaloids. Figure 1. Open in new tabDownload slide Representative organocatalysis for the total synthesis of indole alkaloids. (a) A crucial role of organocatalysed reactions applied in the key step of total synthesis. (b) Organocatalysts involved in total synthesis. (c) Representative indole alkaloids. TOTAL SYNTHESIS OF CYCLOTRYPTAMINE ALKALOIDS Cyclotryptamine alkaloids constitute a large family of natural products that show fascinating biological activities [34,35]. These molecules, such as (+)-folicanthine and (–)-chimonanthine, share an octahydro-3a,3΄a-bispyrrolo[2,3-b]indole subunit characterized by vicinal all-carbon quaternary stereogenic centers. The construction of the bispyrrolidinoindoline core is considered the key issue for total synthesis and thereby has received considerable attention [36]. However, the enantioselective catalytic synthesis of all-carbon quaternary stereogenic centers has been considered a longstanding formidable challenge [37,38]. Thus, the construction of octahydro-3a,3΄a-bispyrrolo[2,3-b]indole skeletons basically relied on asymmetric induction [39,40] and chiral pool-strategy [41–435]. Folicanthine In 2012, Gong and coworkers established an asymmetric substitution of 3-hydroxyoxindole 1 with enamine 2 via TS-1, providing an alkylation product 4 in 82% yield and with 90% ee (Scheme 1) [44]. (E)-Oxime 5 was then obtained via a three-step routine transformations from 4. A HgCl2-promoted Beckman rearrangement selectively delivered amide 6, which was followed by a five-step reaction sequence, including oxidation of indole moiety and alkylation, to give the key diamide intermediate 7. After a known reductive cyclization of 7 [45], the first catalytic enantioselective synthesis of (+)-folicanthine was finished in 12 steps and 3.7% overall yield. Scheme 1. Open in new tabDownload slide First catalytic enantioselective total synthesis of (+)-folicanthine by Gong and coworkers. Scheme 1. Open in new tabDownload slide First catalytic enantioselective total synthesis of (+)-folicanthine by Gong and coworkers. A more concise total synthesis of folicanthine via asymmetric metal/organo relay catalysis was developed by Gong and coworkers [46]. The combination of 1.0 mol% of Rh2(OAc)4 and 5.0 mol% of chiral bifunctional squaramide 9 enabled a three-component coupling reaction of indole, diazooxindole 8 and nitroethylene, affording chiral nitroalkane 10 bearing 3,3΄-bisindole skeleton in a decent yield and with 88% ee (Scheme 2). Oxidation of indole moiety within 10 and a subsequent diastereoselective Michael addition with nitroethylene catalysed by manganese(II) acetate provided Kanai-Matsunaga intermediate 11 [47]. Finally, following Kanai/Matsunaga's [47] and Oguri's procedures [48], the synthesis of (–)-folicanthine was accomplished in seven steps and 14.5% overall yield. Scheme 2. Open in new tabDownload slide Synthesis of (–)-folicanthine via asymmetric Rh(II)/squaramide relay catalysis by Gong and coworkers. Scheme 2. Open in new tabDownload slide Synthesis of (–)-folicanthine via asymmetric Rh(II)/squaramide relay catalysis by Gong and coworkers. Chimonanthine (–)-Chimonanthine, isolated from skin extracts of the Colombian poison-dark frog phyllobates terribilis [49], also has a bispyrrolidinoindoline core. The total synthesis of this molecule has intensively been established. In 2013, Ma and coworkers reported an extremely concise synthesis of chimonanthine [50]. By virtue of chiral anion-phase transfer catalysis pioneered by Toste [51,52], gram-scale asymmetric bromocyclization reaction of protected tryptamine 12 with 13 afforded 15 in 98% yield and 95% ee (Scheme 3). A low-covalent cobalt(I) complex mediated homo-coupling reaction of 15 provided a direct access to octahydro-3a,3΄a-bispyrrolo[2,3-b]indole skeleton 16 [53–565], which could be then converted into (–)-chimonanthine after removal of the N-Boc protecting group and the reduction of methylcarbamate. Scheme 3. Open in new tabDownload slide Chiral anion-phase transfer strategy for the total synthesis of (–)-chimonathine by Ma and Xie. Scheme 3. Open in new tabDownload slide Chiral anion-phase transfer strategy for the total synthesis of (–)-chimonathine by Ma and Xie. Gliocladin C Gliocladin C was isolated from a strain of Gliocladium sp., originally separated from the sea hare exhibited cytotoxic activity against murine P388 lymphocytic leukemia cells [57]. 3-Functionalized indolyloxindole skeleton, commonly existing in a wide selection of alkaloids, including Gliocladin C, has been commonly used as a key intermediate in the total synthesis of natural products [58]. The hexahydropyrrolo indole skeleton, in particular a quaternary stereogenic center at its C3 position, is a signature structural element. In 2011, Overman and coworkers reported a highly efficient Steglich-type rearrangement of bisindole 18, in the presence of 5% Fu's catalyst 19, to give 3,3΄-disubstituted oxindole 20 in 96% yield and 96% ee on a 15-gram scale (Scheme 4) [59]. The carbonyl of oxindole in 20 was easily reduced by NaBH4, and then underwent a highly yielding methylation reaction with trimethyl orthoformate under acidic conditions. Subsequent Soai reduction and Dess-Martin oxidation provided aldehyde intermediate 22, which was able to couple with a trioxopiperazine analogue 23 to furnish compound 24. Finally, (+)-gliocladin C was obtained by the treatment of 24 with Sc(OTf)3. Starting from (+)-di-Boc-gliocladin C, Overman and coworkers accomplished the total synthesis of (+)-gliocladine C, (+)-Leptosin D, (+)-T988C, (+)-Bionectin A and Plectosphaeroic Acid B [60,61]. Scheme 4. Open in new tabDownload slide Steglich-type rearrangement for total synthesis of (+)-gliocladin C by Overman and coworkers. Scheme 4. Open in new tabDownload slide Steglich-type rearrangement for total synthesis of (+)-gliocladin C by Overman and coworkers. In 2013, the Gong group described a primary amine-catalysed highly enantioselective alkylation reaction of 3-hydroxyoxindoles with aldehydes, which allows the asymmetric catalytic total synthesis of (+)-gliocladin C (Scheme 5) [62]. The combined catalysts of primary amine 26 and chiral phosphoric acid 27 provided an efficient approach to access disubstituted oxindoles in 80% yield and 94% ee. Interestingly, the addition of ethanol gave relatively higher enantioselectivity, showing that the alcohol indeed plays an important role in the formation of new carbon–carbon bond. The reduction of the chiral aldehyde 28 to corresponding alcohol and protection with the tert-butyldimethylsilyl (TBS) group gave 29. Oxindole carbonyl turned out to be more reactive after protecting with (Boc)2O and was thereby reduced with NaBH4 in MeOH to give a hemiaminal 30. A subsequent methylation and deprotection reaction sequence provided diol 31, which was then transformed into Overman intermediate [59] by NaIO4 oxidation. From the chiral alkylation adduct 28, the enantioselective total synthesis of (+)-gliocladin C was successfully accomplished in 12 steps and 19% overall yield. Scheme 5. Open in new tabDownload slide Synthesis of (+)-gliocladin C by Gong and coworkers. Scheme 5. Open in new tabDownload slide Synthesis of (+)-gliocladin C by Gong and coworkers. TOTAL SYNTHESIS OF POLYCYCLIC TERPENE INDOLE ALKALOIDS Translating new methodology in total synthesis and building up large collections of natural product families are two fundamental challenges in target molecule synthesis. The MacMillan group recently sought to develop an asymmetric strategy for the collective total synthesis by preparing an advanced core structure based on organocascade catalysis (Fig. 2). The strategy of collective natural product synthesis via chiral imidazolidinone catalysis has been demonstrated through the asymmetric total syntheses of six alkaloid natural products: (–)-strychnine, (+)-aspidospermidine, (+)-vincadifformine, (–)-akuammicine, (–)-kopsinine and (–)-kopsanone [63]. Figure 2. Open in new tabDownload slide Organocascade catalysis developed by MacMillan and coworkers. Figure 2. Open in new tabDownload slide Organocascade catalysis developed by MacMillan and coworkers. The detailed catalytic cycles of this fascinating cascade reaction are shown in Fig. 2 [63]. Under the catalysis of 1-naphthyl-substituted imidazolidinone 32, the tryptamine derivative 33 and propynal underwent an endo-selective Diels-Alder reaction, followed by a methyl selenide-triggered β-elimination, to give a transient intermediate 35. The author proposed a reversible cyclization/ring-opening process of 35. In the second catalytic cycle, both iminium and Brønsted acid catalysis were possible to lower the Lowest Unoccupied Molecular Orbital (LUMO) energy of 38, thus rendered a conjugate addition reaction to generate the common intermediate 39. (–)-Strychnine and (–)-akuammicine Strychnine is the most well-known member of the strychnos alkaloid family. Akuammicine is an alkaloid found in Vinca minor and Aspidosperma. Importantly, strychnine is believed to share a key precursor of these structurally diverse alkaloids. Thus, a common intermediate 39 is exploited in the construction of a diverse collection of target complex molecules. The crucial asymmetric reaction was accomplished by using catalyst 32 and tribromoacetic acid as co-catalyst, giving the tetracyclic spiroindoline 39a in a good yield and with excellent enantioselectivity (82% yield, 97% ee, Scheme 6). Decarbonylation of 39a mediated with Wilkinson's catalyst (Rh(PPH3)3Cl) followed by the introduction of a carbomethoxy group and reduction with DIBAL-H resulted in an isomeric mixture of unsaturated ester 40. An allylation/reduction reaction sequence that occurred with 40 afforded compound 41. As a second key step, a cascade Jeffery-Heck cyclization/lactol formation was initiated. The insertion of palladium into the vinyl iodide and a subsequent carbopalladation would forge the six-membered ring and produce an alkyl palladium intermediate. The following β-hydride elimination step provided an enol that would rapidly engage in lactol cyclization and gave intermediate 42. The removal of the p-methoxybenzyl (PMB) group with trifluoroacetic acid (TFA)/PhSH followed by treatment with malonic acid delivered enantioenriched (–)-strychnine. A three-step reaction sequence starting from 40, including removal of the PMB protecting group, an allylation reaction and a Heck cyclization finally furnished (–)-akuammicine (Scheme 6) [63]. Scheme 6. Open in new tabDownload slide Synthesis of (–)-strychnine and (–)-akuammicine by MacMillan and coworkers. Scheme 6. Open in new tabDownload slide Synthesis of (–)-strychnine and (–)-akuammicine by MacMillan and coworkers. The broad utility of these organocatalytic enantioselective cascade reactions for constructing polycyclic rings is illustrated by total syntheses of several related natural products, such as (+)-aspidospermidine, (+)-vincadifformine, (–)-kopsinine and (–)-kopsanone (Scheme 7) [63]. Scheme 7. Open in new tabDownload slide (a) Application of organocascade catalysis to total synthesis of (+)-vincadifformine. (b) Total synthesis of (–)-kopsinine and (–)-kopsanone. Scheme 7. Open in new tabDownload slide (a) Application of organocascade catalysis to total synthesis of (+)-vincadifformine. (b) Total synthesis of (–)-kopsinine and (–)-kopsanone. Minovincine (–)-Minovincine was first isolated from Vinca minor in 1962 [64]. In 2013, the MacMillan group disclosed an organocatalytic cascade reaction of 2-vinyl indole 33a with but-3-yn-2-one, providing direct access to the key tetracycle skeleton of (–)-minovincine (Scheme 8). The vinyl selenide system was crucial for the cascade enantioselective Diels-Alder cycloaddition, β-elimination and conjugate addition reactions. A palladium-catalysed carbomethoxylation of the dienylogous amide converted the enantioenriched tetracyclic intermediate 52 into a vinylogous carbamate 53 in 92% yield. After a subsequent four-step reaction sequence, (–)-minovincine was accessed in 13% overall yield [65]. Scheme 8. Open in new tabDownload slide Organocascade catalysis for synthesis of (–)-minovincine by MacMillan and coworkers. Scheme 8. Open in new tabDownload slide Organocascade catalysis for synthesis of (–)-minovincine by MacMillan and coworkers. Vincorine (–)-Vincornie, which belongs to the Akuammiline-type alkaloids, has been found to exhibit interesting biological activity and holds an inspiring architecture [66]. In 2013, the MacMillan group reported a concise enantioselective total synthesis of (–)-vincornie based on an organocatalytic Diels-Alder reaction, elimination and conjugate cascade addition (Scheme 9a) [67]. Chiral secondary amine catalyst 55 could give the tetracyclic core structure 56 in 73% yield and 95% ee. Notably, the challenging seven-membered azepanyl ring system was built up by means of a radical cyclization initiated from an acyl telluride precursor 58. This transformation was believed to proceed through carbon-Te bond hemolysis, followed by releasing carbon monoxide to generate an alkyl radical. The challenging seven-membered azepanyl ring was then accomplished through a radical cyclization reaction. Finally, the total synthesis of (–)-vincorine was achieved in nine steps and 9% overall yield from commercially available starting materials. Scheme 9. Open in new tabDownload slide (a) Synthesis of (–)-vincorine by MacMillan and coworkers. (b) Synthesis of (–)-vincorine by Ma and coworkers. Scheme 9. Open in new tabDownload slide (a) Synthesis of (–)-vincorine by MacMillan and coworkers. (b) Synthesis of (–)-vincorine by Ma and coworkers. Alternatively, the Ma group reported the total synthesis of (–)-vincorine via an organocatalysed asymmetric Michael addition reaction (Scheme 9b) [68]. The combination of alkylidene malonate 60, aldehyde 61 and 20 mol% of TMS-protected diphenylprolinol 62 would lead to the Michael adduct 63 in 75% yield as a 5:1 diastereomeric mixture. Subsequently, a sequential process involving oxidation of the aryl selenide moiety of 63 and Et3N-mediated elimination produced olefin 64. E-64 was then treated with strong base LiHMDS and I2 to undergo an oxidative coupling reaction to give 65 in 67% isolated yield [69–71]. (–)-Vincorine was finally furnished in 5% overall yield after additional five-step routine manipulation. Deethylibophyllidine and limaspermidine (+)-Deethylibophyllidine was originally isolated from the bark of Tabernaemontana albif lora and Anacampta disticha in 1980 [72] and (+)-limaspermidine was initially isolated from the trunk bark of a small Venezuelan tree A. rhombeosignatum MARKGRAF in 1979 [73]. Both of these two natural products belong to the family of Apocynaceae alkaloids, and thus commonly have a hydrocarbazole nucleus as core structure. This functionalized hydrocarbazole unit, which contains an all-carbon quaternary stereocenter, poses a particular challenge in organic synthesis. In 2015, Fan and coworkers reported a total synthesis of these two natural products (Scheme 10) [74]. The routes involved a key tandem reaction of aminolysis and aza-Michael addition reaction of spirocyclic para-dienoneimides 66 through organocatalytic enantioselective desymmetrization. The chiral tertiary amine thiourea 67 was identified to be the optimal catalyst in terms of the yield and selectivity (45% yield and 90% ee) in the model reaction of 66 with allyl amine. The syntheses of natural products commenced with a base-promoted intramolecular aza-Michael addition of desymmetrization product 68, affording the versatile tetracyclic precursor 69 in 85% yield. The ozonolysis and palladium-catalysed anti-Markovnikov oxidation of 69 respectively gave the aldehydes 70 and 72, which then underwent an intramolecular aldol reaction to furnish the pentacyclic building block 71 and 73, respectively. Starting from the intermediate 71, an assembly of (+)-deethylibophyllidine was achieved after six additional transformations, while 10 steps were still needed from intermediate 73 to access (+)-limaspermidine. Scheme 10. Open in new tabDownload slide Synthesis of (+)-deethylibophyllidine and (+)-limaspermidine by Fan and coworkers. Scheme 10. Open in new tabDownload slide Synthesis of (+)-deethylibophyllidine and (+)-limaspermidine by Fan and coworkers. Ibophyllidine (+)-Ibophyllidine, originally isolated in 1976 from the plants Taberanthe iboga and T. subsessilis, contains an all-syn-pyrrolidine nucleus, thus holding an uncommon synthetic challenge [75]. In 2012, the Kwon group reported the first enantioselective total synthesis of (+)-ibophyllidine by using the key asymmetric [3 + 2] annulation (Scheme 11) [76]. Based on the concept pioneered by Lu and coworkers [77], catalytic amounts of chiral phosphine 76 was able to attack the β-carbon atom of allenoate 75 to generate a reactive dipole intermediate that could participate in an asymmetric [3 + 2] cycloaddition with imine 74 to afford dihydropyrrole 77 in 93% yield and with 99% ee. A hydrogenation of the pyrroline double bond in 77 catalysed by Raney Ni led to the isolation of an all-syn pyrrolidine derivative 78 in 80% yield. Further transformations, including reduction, Swern oxidation and Wittig reaction, provided the intermediate 80. Silver(I)-catalysed sequential spirocyclization and aza-Morita-Baylis-Hillman reactions provided access to the pentacyclic framework in 81. After a subsequent three-step transformation, (+)-ibophyllidine was obtained in 13% overall yield. Scheme 11. Open in new tabDownload slide Phosphine-catalysed [3 + 2] cycloaddition for total synthesis of (+)-lbophyllidine. Scheme 11. Open in new tabDownload slide Phosphine-catalysed [3 + 2] cycloaddition for total synthesis of (+)-lbophyllidine. TOTAL SYNTHESIS OF TETRAHYDRO-β-CARBOLINE ALKALOIDS The 1,2,3,4-tetrahydro-β-carboline (THBC) ring has been frequently found in a wide range of indole alkaloids and biologically active compounds [78,79]. A catalytic enantioselective Pictet-Spengler reaction of tryptamine and its derivative 83 with aldehyde or ketone 84 is the most straightforward method to access enantiopure 1,2,3,4-tetrahydro-β-carboline skeleton 85 (Scheme 12a) [80,81], which is the core structural element of numerous indole alkaloids. Chiral small molecules, especially chiral Brønsted acids, have been the privileged catalysts for asymmetric Pictet-Spengler reactions [82–87], thus enabling asymmetric catalytic total synthesis of tetrahydro-β-carboline alkaloids, such as (+)-yohimbine by the Maarseceen and Hiemstra group (Scheme 12b) [84] and (–)-arboricine by Jacobsen and coworkers (Scheme 12c) [85]. Recently, the total synthesis of more complex indole alkaloids was also accomplished by using the asymmetric Pictet-Spengler reaction as the key step. Scheme 12. Open in new tabDownload slide (a) Access to 1,2,3,4-tetrahydro-β-carboline via asymmetric Pictet-Spengler reaction. (b) Synthesis of (+)-yohimbine. (c) Synthesis of (–)-arboricine. Scheme 12. Open in new tabDownload slide (a) Access to 1,2,3,4-tetrahydro-β-carboline via asymmetric Pictet-Spengler reaction. (b) Synthesis of (+)-yohimbine. (c) Synthesis of (–)-arboricine. Mitragynine Mitragynine, a potent analgesic reagent [88], was isolated from tropical tree Mitragyna speciosa (Rubiaceae) native to Thailand and Malaysia [89]. In 2012, the Maarseveen and Hiemstra group established a total synthesis of (–)-mitragynine and its congeners, (+)-paynantheine and (+)-speciogynine (Scheme 13) [90]. This synthetic approach commenced with an asymmetric Pictet-Spengler reaction catalysed by quinine-derived thiourea catalyst 90. A reaction sequence involving protection/hydrolysis of dithoacetal gave rise to the α-ketoester 92 in 73% yield. A subsequent intramolecular Tsuji-Trost allylic alkylation reaction of 92 delivered the tetracyclic intermediate as a 4:1 mixture of epimers. The intermediate 93b was treated with ylide in situ generated from methoxymethylene triphenylphosphonium chloride and then with TFA to provide (+)-paynantheine, which was finally converted to (–)-speciogynine via a simple hydrogenation. Following a similar procedure, (–)-mitragynine was also synthesized in 60% yield. Scheme 13. Open in new tabDownload slide Synthesis of (–)-mitragynine, (+)-paynantheine and (+)-speciogynine by Maarseveen and Hiemstra group. Scheme 13. Open in new tabDownload slide Synthesis of (–)-mitragynine, (+)-paynantheine and (+)-speciogynine by Maarseveen and Hiemstra group. Peganumine A Peganumine A, first isolated from the seeds of Peganum harmala L. in 2014 [91], possesses a unique octacyclic architecture and shows significant cytotoxic activity against MCF-7, PC-3, HepG2 cells. Very recently, Zhu and coworkers accomplished the first total synthesis of (+)-Peganumine A by using a catalytic enantioselective Pictet-Spengler reaction as one of the key steps (Scheme 14) [92]. The thioesterification and ozonolysis of commercially available 3,3-dimethyl-pen-4-enoic acid 94 gave 95. Starting from 6-methoxyl-tryptamine 96, a three-step synthetic sequence provided organotin intermediate 98, which underwent a Liebeskind−Srogl cross-coupling reaction with 95 to give 99. The dehydration of 99 followed by a 4-center-3-component Ugi reaction afforded tetracyclic 101. Under the catalysis of chiral thiourea 102, the asymmetric Pictet-Spengler reaction of 101 with 6-methoxyl-tryptamine 96 gave (+)-peganumine in 69% yield and with 92% ee. Scheme 14. Open in new tabDownload slide Synthesis of (+)-peganumine by Zhu and coworkers. Scheme 14. Open in new tabDownload slide Synthesis of (+)-peganumine by Zhu and coworkers. OTHER INDOLE ALKALOIDS Trigonoliimine A Trigonoliimine A with a unique polycyclic skeleton was isolated from the extract of the leaves of Trigonostemon. Lii Y. T. Chang collected in Yunnan province of China [93]. The structure features a tricylic core with a quaternary carbon center [94]. In 2013, Zhu reported an organoctalytic enantioselective Michael addition of α-aryl-isocyanoacetates 103 to vinyl phenyl selenone, which turned out to be the key step for the total synthesis of (+)-trigonoliimine A (Scheme 15a) [95]. A nucleophilic substitution of the phenylselenone 105, obtained from the asymmetric Michael addition, with sodium azide and followed by hydrolysis of nitrile functionality gave intermediate 106, which underwent a reductive alkyla-tion with 2-(1H-indole-3-yl)acetaldehydealdehyde 107 in the presence of NaBH(OAc)3 to afford complex quaternary α-amino ester 108. An efficient Staudinger reduction directly produced the lactam 109 in 78% yield. A subsequent three-step reaction sequence involving reduction of a nitro group and the Bischler-Napieralski reaction furnished (+)-trigonoliimine in 7.5% overall yield. It is noteworthy that employing quinidine-derived similar bifunctional catalyst 111 under identical conditions and following exactly the same synthetic sequence, (–)-trigonoliimine A could be obtained in 6.8% overall yield (Scheme 15b). Scheme 15. Open in new tabDownload slide (a) Synthesis of (+)-trigonoliimine A by Zhu and coworkers. (b) Synthesis of (–)-trigonoliimine A. Scheme 15. Open in new tabDownload slide (a) Synthesis of (+)-trigonoliimine A by Zhu and coworkers. (b) Synthesis of (–)-trigonoliimine A. Anhydronium alkaloids The first isolation of alstonine (from Alstonia constricta [96]) and serpentine (from Rauvolfia serpentine [97]) dates back to about 80 years ago. Biologically, alstonine and serpentine are able to treat psychotic disorders. They are pentacyclic indole alkaloids containing a zwitterionic indolo[2,3-a]quinolizidine motif and three contiguous stereogenic centers. In 2015, Scheidt and coworkers described an enamine-catalysed Michael addition reaction for the enantioselective total syntheses of these natural products (Scheme 16a) [98]. The combination of chiral primary amine 113 (30 mol%) and catechol rendered a highly efficient reaction to deliver piperidine derivative 114 in 78% yield and with 98% ee. Starting from 114, dihydropyran 115 was obtained through several routine manipulations and a key Korte rearrangement. After a highly yielding deprotection reaction, the intermediate 115 underwent a reductive alkylation with indole-3-acetaldehyde to give intermediate 116. Then an oxidative iminium ion cyclization of 116 promoted by mercuric acetate and ethylenediaminetetraacetic acid (EDTA) furnished a heteroyohimbine alkaloid ajmalicine in 41% yield. Dehydrogenation of ajmalicine by using Pd/C as catalyst in aqueous conditions followed by alkoxylation provided serpentine methoxide in 26% yield. Similarly, the synthesis of alstonine hydrogen maleate could be accomplished within four steps from 118 (Scheme 16b). Scheme 16. Open in new tabDownload slide (a) Organocatalytic asymmetric intramolecular Michael addition reaction. (b) Synthesis of ajmalicine and related natural products by Scheidt and coworkers. Scheme 16. Open in new tabDownload slide (a) Organocatalytic asymmetric intramolecular Michael addition reaction. (b) Synthesis of ajmalicine and related natural products by Scheidt and coworkers. Actinophyllic acid (–)-Actinophyllic acid was isolated in 2005 from the leaves of the tree Alstonia actinophylla growing in Cape York Peninsula, Far North Queensland, Australia [99]. (–)-Actinophyllic acid contains the cage-like scaffold, with five contiguous stereogenic centers, one of which is a quaternary carbon, bridged by a tetrahydrofuran lactol. In 2016, the Kwon group reported a catalytic asymmetric total synthesis of (–)-actinophyllic acid by using a [3 + 2] annulation strategy between imine 119 and benzyl allenoate 120 under catalysis of chiral phosphine 121 (Scheme 17) [100]. S-BINOL or R-BINOL was used as an additive to facilitate the proton-transfer steps and to rigidify the transition-state assembly, thereby improving the enantioselectivity without sacrificing the yield (99% yield and 94% ee). The annulation product 122 was converted to 123 by a C-2 selective iodination/deprotection/aza-Michael addition sequence, followed by an intramolecular CuI-catalysed cross-coupling reaction [101] to allow a C(sp2)-C(sp3) bond formation to give tetracyclic product 124. It is noteworthy that the enantiomeric excess of 124 could be 99% by one single recrystallization operation. Further transformations involving a hydrogenolysis over Pd/C to remove the benzyl group, esterification reaction with chloroiodomethane and nucleophilic substitution were performed to furnish the lactone 126. SmI2-mediated intramolecular ketone-lactone pinacol coupling provided diol product 127 in almost quantitative yield. A key radical dehydroxylation along with the removal of the Boc group provided (–)-actinophyllic acid hydrochloride. Scheme 17. Open in new tabDownload slide Synthesis of (–)-actinophyllic acid hydrochloride by Kwon and coworkers. Scheme 17. Open in new tabDownload slide Synthesis of (–)-actinophyllic acid hydrochloride by Kwon and coworkers. Leucomidine B (–)-Leucomidine B was isolated from the barks of Leuconotis griffithii in 2012 by Morita et al. [102]. Recently, the Zhu group reported a formal synthesis of (–)-leucomidine B [103], wherein asymmetric desymmetrization of prochiral synthon represented the key step (Scheme 18) [104]. A chiral imidodiphosphoric acid 130-catalzyed asymmetric desymmetrization reaction of bicyclic bislactone 129 provided monoacid 131 (95% yield and with 84% ee), which was then converted to protected azide intermediate 132 after five-step derivatization. A Staudinger reduction and subsequent aza-Wittig reaction of the deprotected azide from 132 delivered cyclic imine 133. The condensation of methyl 3-(2-nitrophenyl)-2-oxo-propanoate 134 with 133 afforded [3 + 2] annulation product 135 as a 1:1 mixture of diastereomers in 80% yield. Hydrogenative condensation of 135 gave natural product (–)-leucomidine B in 40% yield and its C21 epimer in 40% yield as well. Scheme 18. Open in new tabDownload slide Formal synthesis of (–)-leucomidine B by Zhu and coworkers. Scheme 18. Open in new tabDownload slide Formal synthesis of (–)-leucomidine B by Zhu and coworkers. CONCLUSION This review has briefly illustrated synthetic applications of organocatalytic reactions in the synthesis of relevant indole alkaloids, which have enabled rapid construction of molecular complexity with excellent levels of stereocontrol. Various approaches towards making these natural products have been reported. It is certain that organocatalysis can be therefore considered an additional tool for asymmetric synthesis, besides other kinds of well-established catalytic methods. Hopefully, this review will inspire more efforts devoted to applying organocatalytic transformations to the total synthesis of complex natural products. Despite the elegant progress that has been accomplished, organocatalytic reactions are still underdeveloped, and some of the organocatalytic reactions are far from being satisfactory. Furthermore, the development of efficient methods for the stereoselective construction of privileged heterocyclic systems will undoubtedly lead to concise strategies for the construction of privileged motifs. Therefore, further development of organocatalytic reactions in both breadth and depth is anticipated to improve efficiency, discover new activation modes towards common feedstocks and provide rapid access to target molecules. Acknowledgements We are grateful for financial support from National Natural Science Foundation of China (21232007) and Chinese Academy of Sciences (XDB20020000). REFERENCES 1. Sundberg RJ. The Chemistry of Indoles . New York and London : Academic Press , 1970 . 2. Bandini M , Eichholzer A. Catalytic functionalization of indoles in a new dimension . Angew Chem Int Ed 2009 ; 48 : 9608 – 44 . Google Scholar Crossref Search ADS WorldCat 3. Humphrey GR , Kuethe JT. Practical methodologies for the synthesis of indoles . Chem Rev 2006 ; 106 : 2875 – 911 . Google Scholar Crossref Search ADS PubMed WorldCat 4. Nicolaou KC , Edmonds DJ, Bulger PG. Cascade reactions in total synthesis . Angew Chem Int Ed 2006 ; 45 : 7134 – 86 . Google Scholar Crossref Search ADS WorldCat 5. Nicolaou KC , Sorensen EJ. Classics in Total Synthesis . New York: VCH , 1996 . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC 6. Nicolaou KC , Chen JS. The art of total synthesis through cascade reactions . Chem Soc Rev 2009 ; 38 : 2993 – 3009 . Google Scholar Crossref Search ADS PubMed WorldCat 7. Trost BM. The atom economy: a search for synthetic efficiency . Science 1991 ; 254 : 1471 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 8. Mohr JT , Krout MR, Stoltz BM. Natural products as inspiration for the development of asymmetric catalysis . Nature 2008 ; 455 : 323 – 32 . Google Scholar Crossref Search ADS PubMed WorldCat 9. Shiri M. Indoles in multicomponent processes (MCPs) . Chem Rev 2012 ; 112 : 3508 – 49 . Google Scholar Crossref Search ADS PubMed WorldCat 10. Dalko PI , Moisan L. In the golden age of organocatalysis . Angew Chem Int Ed 2004 ; 43 : 5138 – 75 . Google Scholar Crossref Search ADS WorldCat 11. Melchiorre P , Marigo M, Carlone Aet al. Asymmetric aminocatalysis-gold rush in organic chemistry . Angew Chem Int Ed 2008 ; 47 : 6138 – 71 . Google Scholar Crossref Search ADS WorldCat 12. MacMillan DWC. The advent and development of organocatalysis . Nature 2008 ; 455 : 304 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 13. Enders D , Grondal C, Hüttl MRM. Asymmetric organocatalytic domino reactions . Angew Chem Int Ed 2007 ; 46 : 1570 – 81 . Google Scholar Crossref Search ADS WorldCat 14. Grondal C , Jeanty M, Enders D. Organocatalytic cascade reactions as a new tool in total synthesis . Nat Chem 2010 ; 2 : 167 – 78 . Google Scholar Crossref Search ADS PubMed WorldCat 15. Abbasov ME , Romo D. The ever-expanding role of asymmetric covalent organocatalysis in scalable, natural product synthesis . Nat Prod Rep 2014 ; 31 : 1318 – 27 . Google Scholar Crossref Search ADS PubMed WorldCat 16. Sánchez-Rosello M , Aceña JL, Simón-Fuentes Aet al. A general overview of the organocatalytic intramolecular aza-Michael reaction . Chem Soc Rev 2014 ; 43 : 7430 – 53 . Google Scholar Crossref Search ADS PubMed WorldCat 17. Sun B-F. Total synthesis of natural and pharmaceutical products powered by organocatalytic reactions . Tetrahedron Lett 2015 ; 56 : 2133 – 40 . Google Scholar Crossref Search ADS WorldCat 18. Akiyama T. Stronger Brønsted acids . Chem Rev 2007 ; 107 : 5744 – 58 . Google Scholar Crossref Search ADS PubMed WorldCat 19. Terada M. Chiral phosphoric acids as versatile catalysts for enantioselective transformations . Synthesis 2010 ; 2010 : 1929 – 82 . Google Scholar Crossref Search ADS WorldCat 20. Kampen D , Reisinger CM, List B. Chiral Brønsted acids for asymmetric organocatalysis . Top Curr Chem 2010 ; 291 : 395 – 456 . Google Scholar Crossref Search ADS PubMed WorldCat 21. Yu J , Shi F, Gong L-Z. Brønsted-acid-catalyzed asymmetric multicomponent reactions for the facile synthesis of highly enantioenriched structurally diverse nitrogenous heterocycles . Acc Chem Res 2011 ; 44 : 1156 – 71 . Google Scholar Crossref Search ADS PubMed WorldCat 22. Parmar D , Sugiono E, Raja Set al. Complete field guide to asymmetric BINOL-phosphate derived Brønsted acid and metal catalysis: history and classification by mode of activation; Brønsted acidity, hydrogen bonding, ion pairing, and metal phosphates . Chem Rev 2014 ; 114 : 9047 – 153 . Google Scholar Crossref Search ADS PubMed WorldCat 23. Doyle AG , Jacobsen EN. Small-molecule H-bond donors in asymmetric catalysis . Chem Rev 2007 ; 107 : 5713 – 43 . Google Scholar Crossref Search ADS PubMed WorldCat 24. Enders D , Niemeier O, Henseler A. Organocatalysis by N-heterocyclic carbenes . Chem Rev 2007 ; 107 : 5606 – 55 . Google Scholar Crossref Search ADS PubMed WorldCat 25. Hopkinson MN , Richter C, Schedler Met al. An overview of N-heterocyclic carbenes . Nature 2004 ; 510 : 485 – 96 . Google Scholar Crossref Search ADS WorldCat 26. Flanigan DM , Romanov-Michailidis F, White NAet al. Organocatalytic reactions enabled by N-heterocyclic carbenes . Chem Rev 2015 ; 115 : 9307 – 87 . Google Scholar Crossref Search ADS PubMed WorldCat 27. Erkkilä A , Majander I, Pihko PM. Iminium catalysis . Chem Rev 2007 ; 107 : 5416 – 70 . Google Scholar Crossref Search ADS PubMed WorldCat 28. Mukherjee S , Yang JW, Hoffman Set al. Asymmetric enamine catalysis . Chem Rev 2007 ; 107 : 5471 – 569 . Google Scholar Crossref Search ADS PubMed WorldCat 29. Bertelsen S , Jørgensen KA. Organocatalysis-after the gold rush . Chem Soc Rev 2009 ; 38 : 2178 – 89 . Google Scholar Crossref Search ADS PubMed WorldCat 30. Zhang Z-Z , Schreiner PR. (Thio)urea organocatalysis: what can be learnt from anion recognition? Chem Soc Rev 2009 ; 38 : 1187 – 98 . Crossref Search ADS PubMed 31. Lu X-Y , Zhang C-M, Xu Z-R. Reactions of electron-deficient alkynes and allenes under phosphine catalysis . Acc Chem Res 2001 ; 34 : 535 – 44 . Google Scholar Crossref Search ADS PubMed WorldCat 32. Ye L-W , Zhou J, Tang Y. Phosphine-triggered synthesis of functionalized cyclic compounds . Chem Soc Rev 2008 ; 37 : 1140 – 52 . Google Scholar Crossref Search ADS PubMed WorldCat 33. Denmark SE , Beutner GL. Lewis base catalysis in organic synthesis . Angew Chem Int Ed 2008 ; 47 : 1560 – 638 . Google Scholar Crossref Search ADS WorldCat 34. Cordell GA , Saxton JE. The Alkaloids: Chemistry and Physiology . New York: Academic Press , 1981 ; 20 : 293 – 5 . Google Scholar Google Preview OpenURL Placeholder Text WorldCat COPAC 35. Anthoni U , Christophersen C, Nielsen PH. Alkaloids: Chemical and Biological Perspectives . London: Pergamon , 1999 ; 163 – 236 . Google Scholar Crossref Search ADS Google Preview WorldCat COPAC 36. Steven A , Overman LE. Total synthesis of complex cyclotryptamine alkaloids: stereocontrolled construction of quaternary carbon stereocenters . Angew Chem Int Ed 2007 ; 46 : 5488 – 508 . Google Scholar Crossref Search ADS WorldCat 37. Douglas CJ , Overman LE. Catalytic asymmetric synthesis of all-carbon quaternary stereocenters . Proc Natl Acad Sci USA 2004 ; 101 : 5363 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 38. Quasdorf KW , Overman LE. Catalytic enantioselective synthesis of quaternary carbon stereocentres . Nature 2014 ; 516 : 181 – 91 . Google Scholar Crossref Search ADS PubMed WorldCat 39. Overman LE , Paone DV, Stearns BA. Direct stereo- and enantiocontrolled synthesis of vicinal stereogenic quaternary carbon centers: total syntheses of meso- and (–)-chimonanthine and (+)-calycanthine . J Am Chem Soc 1999 ; 121 : 7702 – 3 . Google Scholar Crossref Search ADS WorldCat 40. Overman LE , Larrow JF, Stearns BAet al. Enantioselective construction of vicinal stereogenic quaternary centers by dialkylation: practical total syntheses of (+)- and meso-chimonanthine . Angew Chem Int Ed 2000 ; 39 : 213 – 5 . Google Scholar Crossref Search ADS WorldCat 41. Movassaghi M , Schmidt MA. Concise total synthesis of (–)-calycanthine, (+)-chimonanthine, and (+)-folicanthine . Angew Chem Int Ed 2007 ; 46 : 3725 – 8 . Google Scholar Crossref Search ADS WorldCat 42. Movassaghi M , Schmidt MA, Ashenhurst JA. Concise total synthesis of (+)-WIN 64821 and (–)-ditryptophenaline . Angew Chem Int Ed 2008 ; 47 : 1485 – 7 . Google Scholar Crossref Search ADS WorldCat 43. Kim J , Movassaghi M. General approach to epipolythiodiketopiperazine alkaloids: total synthesis of (+)-chaetocins A and C and (+)-12,12΄-dideoxychetracin A . J Am Chem Soc 2010 ; 132 : 14376 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 44. Guo C , Song J, Huang J-Zet al. Core-structure-oriented asymmetric organocatalytic substitution of 3-hydroxyoxindoles: application in the enantioselective total synthesis of (+)-folicanthine . Angew Chem Int Ed 2012 ; 51 : 1046 – 50 . Google Scholar Crossref Search ADS WorldCat 45. Fang CL , Horne S, Taylor N et al. Dimerization of a 3-substituted oxindole at C-3 and its application to the synthesis of (.+-.)-folicanthine . J Am Chem Soc 1994 ; 116 : 9480 – 6 . Google Scholar Crossref Search ADS WorldCat 46. Chen D-F , Zhao F, Hu Yet al. C-H functionalization/asymmetric Michael addition cascade enabled by relay catalysis: metal carbenoid used for C-C bond formation . Angew Chem Int Ed 2014 ; 53 : 10763 – 7 . Google Scholar Crossref Search ADS WorldCat 47. Mitsunuma H , Shibasaki M, Kanai Met al. Catalytic asymmetric total synthesis of chimonanthine, folicanthine, and calycanthine through double michael reaction of bisoxindole . Angew Chem Int Ed 2012 ; 51 : 5217 – 21 . Google Scholar Crossref Search ADS WorldCat 48. Wada M , Murata T, Oikawa Het al. Nickel-catalyzed dimerization of pyrrolidinoindoline scaffolds: systematic access to chimonanthines, folicanthines and (+)-WIN 6482 . Org Biomol Chem 2014 ; 12 : 298 – 306 . Google Scholar Crossref Search ADS PubMed WorldCat 49. Duke RK , Allan RD, Johnston GAR et al. Idiospermuline, a trimeric pyrrolidinoindoline alkaloid from the seed of idiospermum australiense . J Nat Prod 1995 ; 58 : 1200 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 50. Xie W , Jiang G, Liu Het al. Highly enantioselective bromocyclization of tryptamines and its application in the synthesis of (–)-chimonanthine . Angew Chem Int Ed 2013 ; 52 : 12924 – 7 . Google Scholar Crossref Search ADS WorldCat 51. Phipps RJ , Hamilton GL, Toste DF. The progression of chiral anions from concepts to applications in asymmetric catalysis . Nat Chem 2012 ; 4 : 603 – 14 . Google Scholar Crossref Search ADS PubMed WorldCat 52. Rauniyar V , Lackner AD, Hamilton GL et al. Asymmetric electrophilic fluorination using an anionic chiral phase-transfer catalyst . Science 2011 ; 334 : 1681 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 53. Movassaghi M , Schmidt MA, Ashehurst JA. Concise total synthesis of (+)-WIN 64821 and (–)-ditryptophenaline . Angew Chem Int Ed 2008 ; 47 : 1485 – 7 . Google Scholar Crossref Search ADS WorldCat 54. Kim J , Ashenhurst JA, Movassaghi M. Total synthesis of (+)-11,11’-dideoxyverticillin A . Science 2009 ; 324 : 238 – 41 . Google Scholar Crossref Search ADS PubMed WorldCat 55. Kim J , Movassaghi M. General approach to epipolythiodiketopiperazine alkaloids: total synthesis of (+)-chaetocins A and C and (+)-12,12-dideoxychetracin A . J Am Chem Soc 2010 ; 132 : 14376 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 56. Movassaghi M , Ahmad OK, Lathrop SP. Directed heterodimerization: stereocontrolled assembly via solvent-caged unsymmetrical diazene fragmentation . J Am Chem Soc 2011 ; 133 : 13002 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 57. Usami Y , Yamaguchi J, Numata A. Gliocladins A-C and glioperazine; cytotoxic dioxo- or trioxopiperazine metabolites from a gliocladium sp. separated from a sea hare . Heterocycles 2004 ; 63 : 1123 – 9 . Google Scholar Crossref Search ADS WorldCat 58. Takahashi C , Numata A, Ito Yet al. Leptosins antitumour metabolites of a fungus isolated from a marine alga . J Chem Soc Perkin Trans 1994 ; 1 : 1859 – 64 . Google Scholar Crossref Search ADS WorldCat 59. DeLorbe JE , Jabri SY, Mennen SMet al. Enantioselective total synthesis of (+)-gliocladine C: convergent construction of cyclotryptamine-fused polyoxopiperazines and a general approach for preparing epidithiodioxopiperazines from trioxopiperazine precursors . J Am Chem Soc 2011 ; 133 : 6549 – 52 . Google Scholar Crossref Search ADS PubMed WorldCat 60. DeLorbe JE , Hornet D, Jove Ret al. General approach for preparing epidithiodioxopiperazines from trioxopiperazine precursors: enantioselective total syntheses of (+)- and (−)-gliocladine C, (+)-leptosin D, (+)-T988C, (+)-bionectin A, and (+)-gliocladin A . J Am Chem Soc 2013 ; 135 : 4117 – 28 . Google Scholar Crossref Search ADS PubMed WorldCat 61. Jabri SY , Overman LE. Enantioselective total synthesis of plectosphaeroic acid B . J Am Chem Soc 2013 ; 135 : 4231 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 62. Song J , Guo C, Adele Aet al. Enantioselective organocatalytic construction of hexahydropyrroloindole by means of α-alkylation of aldehydes leading to the total synthesis of (+)-gliocladin C . Chem Eur J 2013 ; 19 : 3319 – 23 . Google Scholar Crossref Search ADS PubMed WorldCat 63. Jones SB , Simmons B, Mastracchio Aet al. Collective synthesis of natural products by means of organocascade catalysis . Nature 2011 ; 475 : 183 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 64. Cava MP , Tjoa SS, Ahmed QAet al. Alkaloids of Tabernaemontana riedelii and T. rigida . J Org Chem 1968 ; 33 : 1055 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 65. Laforteza BN , Pickworth M, MacMillan DWC. Enantioselective total synthesis of (–)-minovincine in nine chemical steps: an approach to ketone activation in cascade catalysis . Angew Chem Int Ed 2013 ; 52 : 11269 – 72 . Google Scholar Crossref Search ADS WorldCat 66. Ramirez A , Garcia-Rubio S. Current progress in the chemistry and pharmacology of akuammiline alkaloids . Curr Med Chem 2003 ; 10 : 1891 – 915 . Google Scholar Crossref Search ADS PubMed WorldCat 67. Horning BD , MacMillan DWC. Nine-step enantioselective total synthesis of (–)-vincorine . J Am Chem Soc 2013 ; 135 : 6442 – 5 . Google Scholar Crossref Search ADS PubMed WorldCat 68. Zi W-W , Xie W-Q, Ma D-W. Total synthesis of akuammiline alkaloid (–)-vincorine via intramolecular oxidative coupling . J Am Chem Soc 2012 ; 134 : 9126 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 69. Baran PS , Richter JM. Direct coupling of indoles with carbonyl compounds: short, enantioselective, gram-scale synthetic entry into the hapalindole and fischerindole alkaloid families . J Am Chem Soc 2004 ; 126 : 7450 – 1 . Google Scholar Crossref Search ADS PubMed WorldCat 70. Baran PS , Richter JM. Enantioselective total syntheses of welwitindolinone A and fischerindoles I and G . J Am Chem Soc 2005 ; 127 : 15394 – 6 . Google Scholar Crossref Search ADS PubMed WorldCat 71. Baran PS , Richter JM, Lin DW. Short enantioselective total synthesis of stephacidin A . Angew Chem Int Ed 2005 ; 44 : 606 – 9 . Google Scholar Crossref Search ADS WorldCat 72. Medina JD , Di Genova L. Alkaloids of the bark of aspidosperma rhombeosignatum . Planta Med 1979 ; 37 : 165 – 7 . Google Scholar Crossref Search ADS WorldCat 73. Kan C , Husson H-P, Jacquemin Het al. Determination de structures par RMN du 1H A 400 MH: alcaloides de tabernaemontana albiflora . Tetrahedron Lett 1980 ; 21 : 55 – 8 . Google Scholar Crossref Search ADS WorldCat 74. Du J-Y , Zeng C, Han X-Jet al. Asymmetric total synthesis of apocynaceae hydrocarbazole alkaloids (+)-deethylibophyllidine and (+)-limaspermidine . J Am Chem Soc 2015 ; 137 : 4267 – 73 . Google Scholar Crossref Search ADS PubMed WorldCat 75. Khuong-Huu F , Cesario M, Guilhem Jet al. Alcaloides indoliques—CII . Tetrahedron 1976 ; 32 : 2539 – 43 . Google Scholar Crossref Search ADS WorldCat 76. Andrews IP , Kwon O. Enantioselective total synthesis of (+)-ibophyllidine via an asymmetric phosphine-catalyzed 3 + 2 annulation . Chem Sci 2012 ; 3 : 2510 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 77. Xu Z-R , Lu X-Y. Phosphine-catalyzed 3 + 2 cycloaddition reaction of methyl 2,3-butadienoate and N-tosylimines: a novel approach to nitrogen heterocycles . Tetrahedron Lett 1997 ; 38 : 3461 – 4 . Google Scholar Crossref Search ADS WorldCat 78. O’Connor SE , Maresh J. Chemistry and biology of monoterpene indole alkaloid biosynthesis . Nat Prod Rep 2006 ; 23 : 532 – 47 . Google Scholar Crossref Search ADS PubMed WorldCat 79. Ishikura M , Yamada K, Abe T. Simple indole alkaloids and those with a nonrearranged monoterpenoid unit . Nat Prod Rep 2010 ; 27 : 1630 – 80 . Google Scholar Crossref Search ADS PubMed WorldCat 80. Cox ED , Cook JM. The Pictet-Spengler condensation: a new direction for an old reaction . Chem Rev 1995 ; 95 : 1797 – 842 . Google Scholar Crossref Search ADS WorldCat 81. Stöckigt J , Antonchick AP, Wu Fet al. The Pictet-Spengler reaction in nature and in organic chemistry . Angew Chem Int Ed 2011 ; 50 : 8538 – 64 . Google Scholar Crossref Search ADS WorldCat 82. Taylor MS , Jacobsen EN. Highly enantioselective catalytic acyl-Pictet-Spengler reactions . J Am Chem Soc 2004 ; 126 : 10558 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 83. Seayad J , Seayad AM, List B. Catalytic asymmetric Pictet-Spengler reaction . J Am Chem Soc 2006 ; 128 : 1086 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 84. Wanner MJ , Boots RNA, Eradus Bet al. Organocatalytic enantioselective total synthesis of (–)-arboricine . Org Lett 2009 ; 11 : 2579 – 81 . Google Scholar Crossref Search ADS PubMed WorldCat 85. Mergott DJ , Zuend SJ, Jacobsen EN. Catalytic asymmetric total synthesis of (+)-yohimbine . Org Lett 2008 ; 10 : 745 – 8 . Google Scholar Crossref Search ADS PubMed WorldCat 86. Muratore ME , Holloway CA, Pilling AWet al. Enantioselective Brønsted acid-catalyzed N-acyliminium cyclization cascades . J Am Chem Soc 2009 ; 131 : 10796 – 7 . Google Scholar Crossref Search ADS PubMed WorldCat 87. Herle B , Wanner MJ, van Maarseveen JHet al. Total synthesis of (+)-yohimbine via an enantioselective organocatalytic Pictet-Spengler reaction . J Org Chem 2011 ; 76 : 8907 – 12 . Google Scholar Crossref Search ADS PubMed WorldCat 88. Takayama H , Ishikawa H, Kurihara Met al. Studies on the synthesis and opioid agonistic activities of mitragynine-related indole alkaloids: discovery of opioid agonists structurally different from other opioid ligands . J Med Chem 2002 ; 45 : 1949 – 56 . Google Scholar Crossref Search ADS PubMed WorldCat 89. Adkins JE , Boyer EW, McCurdy CR. Mitragyna speciosa: a psychoactive tree from Southeast Asia with opioid activity . Curr Top Med Chem 2011 ; 11 : 1165 – 75 . Google Scholar Crossref Search ADS PubMed WorldCat 90. Kerschgens IP , Claveau E, Wanner MJet al. Total syntheses of mitragynine, paynantheine and speciogynine via an enantioselective thiourea-catalysed Pictet-Spengler reaction . Chem Commun 2012 ; 48 : 12243 – 5 . Google Scholar Crossref Search ADS WorldCat 91. Wang K-B , Di Y-T, Bao Yet al. Peganumine A, a β-carboline dimer with a new octacyclic scaffold from Peganum harmala . Org Lett 2014 ; 16 : 4028 – 31 . Google Scholar Crossref Search ADS PubMed WorldCat 92. Piemontesi C , Wang Q, Zhu J. Enantioselective synthesis of (+)-peganumine A . J Am Chem Soc 2016 ; 138 : 11148 – 51 . Google Scholar Crossref Search ADS PubMed WorldCat 93. Tan C-J , Di Y-T, Wang Y-Het al. Three new indole alkaloids from Trigonostemon lii . Org Lett 2010 ; 12 : 2370 – 3 . Google Scholar Crossref Search ADS PubMed WorldCat 94. Han S , Movassaghi M. Concise total synthesis and stereochemical revision of all (–)-trigonoliimines . J Am Chem Soc 2011 ; 133 : 10768 – 71 . Google Scholar Crossref Search ADS PubMed WorldCat 95. Buyck T , Wang Q, Zhu J. Catalytic enantioselective Michael addition of α-aryl-α-isocyanoacetates to vinyl selenone: synthesis of α,α-disubstituted α-amino acids and (+)- and (–)-trigonoliimine A . Angew Chem Int Ed 2013 ; 52 : 12714 – 8 . Google Scholar Crossref Search ADS WorldCat 96. Sharp TM. The alkaloids of Alstonia barks. Part I. A. constricta, F. Muell . J Chem Soc 1934 ; 287 – 91 . Google Scholar OpenURL Placeholder Text WorldCat 97. Itallie LV , Steenhauer AJ. Rauwolfia serpentina Benth . Arch Pharm 1932 ; 270 : 313 – 22 . Google Scholar Crossref Search ADS WorldCat 98. Younai A , Zeng B-S, Meltzer HYet al. Enantioselective syntheses of heteroyohimbine natural products: a unified approach through cooperative catalysis . Angew Chem Int Ed 2015 ; 54 : 6900 – 4 . Google Scholar Crossref Search ADS WorldCat 99. Carroll AR , Hyde E, Smith Jet al. Actinophyllic acid, a potent indole alkaloid inhibitor of the coupled enzyme assay carboxypeptidase U/hippuricase from the leaves of alstonia actinophylla (apocynaceae) . J Org Chem 2005 ; 70 : 1096 – 9 . Google Scholar Crossref Search ADS PubMed WorldCat 100. Cai L-C , Zhang K, Kwon O. Catalytic asymmetric total synthesis of (−)-actinophyllic acid . J Am Chem Soc 2016 ; 138 : 3298 – 301 . Google Scholar Crossref Search ADS PubMed WorldCat 101. Coste A , Toumi M, Wright Ket al. Copper-catalyzed cyclization of iodo-tryptophans: a straightforward synthesis of pyrroloindoles . Org Lett 2008 ; 10 : 3841 – 4 . Google Scholar Crossref Search ADS PubMed WorldCat 102. Motegi M , Nugroho AE, Hirasawa Yet al. Leucomidines A-C, novel alkaloids from Leuconotis griffithii . Tetrahedron Lett 2012 ; 53 : 1227 – 30 . Google Scholar Crossref Search ADS WorldCat 103. Gualtierotti J-B , Pasche D, Wang Qet al. Phosphoric acid catalyzed desymmetrization of bicyclic bislactones bearing an all-carbon stereogenic center: total syntheses of (−)-rhazinilam and (−)-leucomidine B . Angew Chem Int Ed 2014 ; 53 : 9926 – 30 . Google Scholar Crossref Search ADS WorldCat 104. García-Urdiales E , Alfonso I, Gotor V. Enantioselective enzymatic desymmetrizations in organic synthesis . Chem Rev 2005 ; 105 : 313 – 54 . Google Scholar Crossref Search ADS PubMed WorldCat © The Author(s) 2017. Published by Oxford University Press on behalf of China Science Publishing & Media Ltd. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com This is an Open Access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0/), which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is properly cited. for commercial re-use, please contact journals.permissions@oup.com © The Author(s) 2017. Published by Oxford University Press on behalf of China Science Publishing & Media Ltd. All rights reserved. For permissions, please e-mail: journals.permissions@oup.com TI - Recent progress in organocatalytic asymmetric total syntheses of complex indole alkaloids JF - National Science Review DO - 10.1093/nsr/nwx028 DA - 2017-05-01 UR - https://www.deepdyve.com/lp/oxford-university-press/recent-progress-in-organocatalytic-asymmetric-total-syntheses-of-47JrbkH9ox SP - 381 EP - 396 VL - 4 IS - 3 DP - DeepDyve ER -