Next Article in Journal
Safe and Effective Antioxidant: The Biological Mechanism and Potential Pathways of Ergothioneine in the Skin
Next Article in Special Issue
One-Pot Synthesis of 3,4-Dihydrocoumarins via C-H Oxidation/Conjugate Addition/Cyclization Cascade Reaction
Previous Article in Journal
GSH-Activatable Aggregation-Induced Emission Cationic Lipid for Efficient Gene Delivery
Previous Article in Special Issue
Rh-Catalyzed Cascade C-H Activation/Annulation of N-Hydroxybenzamides and Propargylic Acetates for Modular Access to Isoquinolones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in the Synthesis of 3,n-Fused Tricyclic Indole Skeletons via Palladium-Catalyzed Domino Reactions

1
Department of Chemical Biology, School of Sciences, Henan Agricultural University, Zhengzhou 450002, China
2
Institute of Pesticide, School of Plant Protection, Henan Agricultural University, Zhengzhou 450002, China
*
Authors to whom correspondence should be addressed.
Molecules 2023, 28(4), 1647; https://doi.org/10.3390/molecules28041647
Submission received: 31 December 2022 / Revised: 3 February 2023 / Accepted: 5 February 2023 / Published: 8 February 2023
(This article belongs to the Special Issue Application of Cascade/Tandem Reactions in Green Organic Synthesis)

Abstract

:
3,n-fused (n = 4–7) tricyclic indoles are pervasive motifs, embedded in a variety of biologically active molecules and natural products. Thus, numerous catalytic methods have been developed for the synthesis of these skeletons over the past few decades. In particular, palladium-catalyzed transformations have received much attention in recent years. This review summarizes recent developments in the synthesis of these tricyclic indoles with palladium-catalyzed domino reactions and their applications in the total synthesis of representative natural products.

1. Introduction

Indole is one of the most significant nitrogen heterocycles, owing to its unique biological activity and wide existence in numerous natural products, bioactive molecules, pharmaceuticals, and functional materials [1,2,3,4,5,6]. Among the numerous indole alkaloids, 3,n-fused (n = 4–7) tricyclic indoles have attracted much attention for their typical biological activities and synthetic challenges in recent years [7,8]. The tricyclic indole nucleus is found in a myriad of natural products and pharmaceutical molecules (Scheme 1), such as HKI 0231B [9], dehydrobufotenine [10], ambiguine H [11], tryptorubin A [12], celogentin C [13], chloropeptin I [14], and streptide [15]. Accordingly, a number of strategies have been established for the synthesis of these skeletons, including intramolecular Fischer indole synthesis, Witkop photocyclizations, Diels-Alder reactions, Friedel-Crafts reactions, Pictet-Spengler reactions, 6π-electrocyclizations, and transition-metal-catalyzed domino reactions.
In recent years, numerous advances have been achieved via transition-metal-catalyzed domino reaction. Among them, palladium has been widely used as a versatile catalyst for their typical atom orbital to get a series of tricyclic indole skeletons. The seminal work on the synthesis of tryptophan derivatives was reported by Roberts in 1994 [16]. In this work, the palladium catalyst was employed to complete the bridge between C3 and C4. After that, especially within the last ten years, a variety of protocols for synthesis of these molecules have been developed and refined by researchers. Moreover, the palladium catalyst was also used extensively in cross-coupling reactions, such as Buchwald-Hartwig, Suzuki-Miyaura, Heck, Sonogashira, Negishi, and Stille to synthesize a variety of useful compounds [17,18].
Over the past few years, some reviews on the synthesis of 3,n-fused tricyclic indoles were documented [7,8]. However, to the best of our knowledge, there is no review that summarizes the recent progress for the synthesis of tricyclic indoles via transition metal catalyst. Herein, we want to provide a short review mainly focused on recent advances in the synthesis of tricyclic indoles via the palladium-catalyzed domino reaction up to 2023. Furthermore, the total synthesis of representative natural products obtained by the palladium-catalyzed domino reaction is also described in this paper.

2. Intramolecular Cyclization

The Heck reaction [19] is one of the most powerful methods for generating new C-C bonds, and has been used in numerous syntheses of high value-added chemicals and complex molecules. In 1999, Söderberg and co-workers [20] successfully developed a practical method for synthesizing 3,4-fused tricyclic indole skeletons via two consecutive palladium-catalyzed reactions, using an intramolecular Heck reaction followed by reductive N-heteroannulation. The reaction afforded various tricyclic indole products (3a3d) in 41–78% yields with 5 mol% Pd(OAc)2 as catalyst, 12 mol% 1,3-bis(diphenylphosphino)propane (DPPP) as ligand, 60 psi carbon monoxide as reductant, and polar DMF as solvent (Scheme 2). Notably, compared to the previous Heck reaction, the relative rate of alkene insertion in this reaction is higher, likely due to the presence of the strong electron-withdrawing nitro group. Six years later, the same group [21] reported a similar method for preparation of 3,4-fused tricyclic indoles with 6–8-membered rings. With respect to the scope of substrates, nitrogen- and oxygen-containing rings work well in this reaction as well, except for the carbocyclic ring.
Medium-sized ring fused tricyclic indoles are quite useful structures in numerous natural products such as decursivine and serotobenine. In 2011, Van der Eycken and co-workers [22] reported a novel method for the construction of amide type of 3,4-fused tricyclic indole derivatives through a palladium-catalyzed intramolecular acetylene hydroarylation reaction with excellent regio- and stereoselectivity (Scheme 3). In this report, various nitrogen protecting groups and alkyne substituents, such as methyl (5a), ethyl (5b), tert-butyl (5c), phenyl (5d), and silyl groups were successfully tested and tolerated well, delivering the target products in moderate to good yields, and the structures were confirmed by 1H NMR, 13C NMR spectroscopy, HRMS (EI), and X-ray crystallography. The reaction is rapid, mild, regioselective, stereoselective and proceeds with good yields. It is worth noting that substrates bearing a free NH group (5c) also reacted smoothly in this reaction.
As we know, Larock indole synthesis is one of the most efficient methods for the rapid construction of indole skeletons from halo-anilines and alkynes with palladium catalyst [23,24]. In 2013, Boger and co-workers [25] successfully reported the first example for rapid assembly of 3,(4-6)-tricyclic indoles 7 via intramolecular Larock indole annulation with a powerful Pd2(dba)3/DtPBF catalyst system, and catalyst and other reaction parameters were examined in detail (Scheme 4). This novel transformation features excellent functional-group tolerance (18 examples), good yields (up to 89% yield), and provides an efficient strategy to afford natural product chloropeptin I, chloropeptin II DEF ring system and key isomers directly. The ring size ranges from 6 to 28, which includes peptide chain and conventional carbon chain. More importantly, the TES or TMS group are easily removed from the molecules, leading to the formation of various indoline skeletons. It is worth noting that the established method is a good complement to the Stille or Suzuki cross-coupling reactions for the synthesis of cyclic or macrocyclic ring indole systems [26,27].
In the same year, Jia and co-workers [28] successfully disclosed a new and general method for the synthesis of 3,4-fused tricyclic indole derivatives through intramolecular Larock indole annulation from internal alkyne tethered ortho-iodoaniline derivatives, and these skeletons are often embedded in numerous natural products and bioactive molecules (Scheme 5). After extensive condition studies, the optimal reaction conditions are as follows: Pd(OAc)2 (20 mol%), PPh3 (40 mol%), K2CO3 (2.0 equiv.), and LiCl (1.0 equiv.) in DMF at 100 °C under nitrogen atmosphere. Applying the Pd/PPh3 catalyst system to the indole-synthesized reaction successfully produced 6–18-membered ring fused tricyclic indoles 9 in good to excellent yields (up to 99% yield), and various linkers, including carbon- (9c,9g,9i)), oxygen- (9a,9b,9d,9e,9h), or nitrogen- (9f) are all tolerated well. In addition, the method was highlighted by the total synthesis of the natural product fargesine, which was isolated from the root and stem of Evodia fargesii Dode. Notably, the authors also explored the reaction with the easily accessible ortho-bromoaniline-type substrates. Considering the lower reactivity of the C-Br bond, some representative ligands were tested to complete the ideal conversion, results showed that electron rich Me-phos or DPPP gave the best yields. Later, the method was also extended to the synthesis of 3,5-fused tricyclic indole scaffolds by the same group [29], giving a variety of 3,5-fused macrocyclic with good yields, and the method was also used to afford the tetrahydropyrrolo [4,3,2-de]quinolone ring.
The strategy of diversity oriented synthesis (DOS) is of great importance in organic synthesis, medicine chemistry, agrochemical chemistry, and material sciences, which aims to efficient collections of small molecules with diverse appendages, functional groups, stereochemistry, and skeletons [30]. In 2013, the You group [31] developed an efficient, Pd(0)-catalyzed allylic alkylation protocol for the synthesis of 3,4-fused tricyclic indole derivatives from readily available 3-subsituted indoles at 50 °C (Scheme 6). The phosphine ligand was crucial for this Pd-catalyzed allylic alkylation reaction, results showed that the 2-(2-(diphenylphosphanyl)phenyl)-4,4-dimethyl-4,5-dihydrooxazole (L1) proved to be the best ligand and led to the formation of tricyclic indole product (11a) in 70% yield. Substituents at the nitrogen such as benzyl and allyl could be well-tolerated, and delivered tricyclic indoles in moderate yields. Indole substrates substituted with fluorine (11c) and chlorine (11d) at the 7-position generated the products with 57% and 60% yield, respectively. Moreover, various 3-substituted substrates were used to examine the substrate scope of the Pd-catalyzed allylic dearomatization reaction under slightly modified conditions. In the transformation, screening of various chiral ligands showed that planar chiral ferrocene-based ligand (L2) gave the highest value of ee. Substrates bearing methyl, allyl, benzyl, and various nucleophilic substituents at the 3-position of indole could be applied in the reaction and gave the desired products in good yields with moderate enantioselectivity (12a12h). Notably, when an excess of trifluoroacetic acid (TFA) was added to (12e) at room temperature, tetracyclic product (12i) was achieved with 80% yield and 73% ee via deprotection of the Boc group followed by addition to imine.
As important raw materials, allenes, which possess unique reactivity compared to alkenes, alkynes, and others, owing to the presence of two cumulative orthogonal π-bonds [32,33,34,35,36], have proven to be key synthetic intermediates in construction of various organic molecules. In particular, allenes generally react with aryl halide to produce π-allylpalladium(II) species in the presence of Pd(0) catalyst through a classic Heck insertion process. In 2015, Nemoto and co-workers [37] successfully demonstrated a novel Pd-catalyzed cascade cyclization with allene tethered ortho-iodoanilines 13 as starting materials. This enabled rapidly assembled 3,4-fused tricyclic 3-alkylidene indoline scaffolds through a sequence of oxidative addition, allene insertion, and allylic amination, thus yielding a new class of 3,4-tricyclic 3-alkylidene indoline skeletons in good to excellent yields with broad substrates scope (Scheme 7). Furthermore, the 3,4-tricyclic 3-alkylidene indolines were divergently transformed into three types of 3,4-fused tricyclic indoles in 94% to quantitative yield with rather simple operation, successfully demonstrating the utility of this cascade process. Alternatively, the desired 3,4-fused tricyclic indoles could also be obtained by oxidation of products 14 with DDQ or PCC.
In many macrocyclic natural products, peptide-based macrocycles are pervasive motifs in medicine molecules [38,39,40]. Among them, drugs, such as anticancer agent octreotide, antibiotic vancomycin, and immunosuppressant cyclosporin are typical molecules. Therefore, a concise method for the synthesis of these skeletons is highly attractive and challenging. In 2018, Chen, Liu, Shen, Qi, He, and co-workers [41] described an amide and pyridine directed β-C(sp3)-H arylation reaction to access diverse peptide macrocycles under Pd(II) catalysis (Scheme 8). In this transformation, 3,5-fused tricyclic indole (16a) and 3,7-fused tricyclic indole (16b) were obtained in 87% and 71% yields, respectively.

3. Intermolecular Cyclization

Construction of new C-C bonds by palladium-catalyzed α-arylation of ketones has emerged as an attractive method in recent years [42]. To date, numerous efforts have been devoted to develop this methodology, and various functionalized molecules have been synthesized. However, application of this method for the synthesis of tetracyclic indoles has not been achieved so far. In 2012, Cuny and co-workers [43] reported a straightforward synthesis of tetracyclic indoles via intramolecular α-arylation of ketones using monoligated Pd(0) catalyst (Scheme 9). The choice of catalyst was vital to access the desired tetracyclic indole product in good to excellent yield. In the presence of the Pd-complex the functionalized het(aryl)ketones transformed to the desired products in 50–99% yields (methyl 20a and 20f, gem-dimethyl 20b, phenyl 20c, 3-methoxyphenyl 20g, and pyridyl 20h). Moreover, both the cyclopentenyl and cyclohexenyl substrates were tolerated well, giving 20d, 20e in a 96% and 94% yield, respectively. It is worth mentioning that the Pd catalytic system features good functional-group tolerance, good to excellent yields, low catalyst loading, and short reaction times. Furthermore, double α-arylation of ketones by one-pot procedures were also presented.
In 2018, based on their previous work and others, Lautens and co-workers [44] successfully discovered a novel Pd2(dba)3/P(2-CF3-C6H4)3-catalyzed domino reaction to give 3,4-fused tricyclic indolones in good to high yields and excellent regioselectivities with ortho-iodoacrylamides 21 and internal alkynes 22 via intramolecular carbopalladation/ortho C-H activation/alkyne insertion sequence (Scheme 10). Various coupling partners were investigated and tolerated well. Specifically, the acrylamide substrates containing N-, α-, and aryl-substituents, and several unsymmetrical alkynes were subjected to the reaction conditions and the corresponding reaction products were isolated as single regioisomers (23a23i). Based on previous reports and experiment results, a plausible mechanism via a Pd(0)/Pd(II) catalytic cycle is depicted in the bottom of Scheme 10. Initially, acrylamide substrate (21a) undergoes oxidative addition to give aryl palladium species A. Next, intramolecular carbopalladation of alkene moiety followed by C(sp2)-H activation led to a five-membered palladacycle B. Afterward, coordination and migratory insertion of alkyne (22a) to five-membered palladacycle B furnished the seven-membered palladacycle C. Finally, reductive elimination of the intermediate C releases desired product (23a) and regenerates an active Pd(0) catalyst.
In 2020, Luan and co-workers [45] elegantly developed the first C(sp2)-H amination of alkyne-tethered aryl iodides for the rapid preparation of a series of tricyclic indoles using secondary hydroxylamines as the bifunctional single-nitrogen source under mild reaction conditions (Scheme 11). Initially, though various N-substituted secondary hydroxylamines were systematically evaluated, only Ts-NH-OBz could be promoted for the titled reaction in 8% yield. Next, various benzoyl O-substituents of amino agents, including steric effect and electron effect, have been examined in this transformation. In particular, substrates containing an ester in para-position led to the formation of tricyclic indole (26a) in 97% yield. Notably, the role of PivOH was crucial in obtaining the tricyclic indole products, which may be due to its ability to facilitate the process of C-H activation. For the scope of 3,4-fused tricyclic indoles, various aryl iodides, linkers, and alkyne termini were explored with good to excellent yields (26b26u). Notably, the pyridyl and thienyl group, which might induce strong coordination toward the Pd(II) species, were found to proceed smoothly to give corresponding tricyclic indole products (26s, 26t) in 93% and 92% yields, respectively. More importantly, the reaction works well to afford 3,4-fused tricyclic indoles as well as 3,5-, and 3,6-fused tricyclic indoles (26v26ac) in acceptable yields. Of note, these tricyclic indoles have been observed as the key core for a range of biologically active natural and unnatural products [12,46,47].
On the basis of previous literature reports and detailed mechanistic investigations [48,49,50], the authors proposed a plausible pathway as depicted in Scheme 12. The reaction initiated with generation of aryl palladium species D via oxidative addition of iodobenzene moiety of 24a to Pd(0). Next, intramolecular carbopalladation of alkynes followed by C-H activation led to the formation of five-membered palladacycle E. After being deprotonated with Cs2CO3, the amino agent 25′ coordinates with electrophilic Pd(II) species E and undergoes concerted 1,2-migration insertion of the aryls to afford the six-membered palladacycle H. Finally, reductive elimination of the intermediate H provided the desired tricyclic indole product 26a with the concomitant regeneration of the active Pd(0) catalyst. Alternatively, formation of formal nitrene intermediate I is followed by nitrene insertion to furnish the six-membered palladacycle H. Intermediate H undergoes reductive elimination to yield the desired product 26a and regenerate an active Pd(0) catalyst.
Almost simultaneously, the Zhang group, and Yu and Jiang group independently developed an efficient Pd(0)-catalyzed intermolecular annulation of ortho-alkyne tethered aryl iodides with N,N-di-tert-butyldiaziridinone 28 for the synthesis of 3,4-fused tricyclic indoles (Scheme 13 and Scheme 14).
Zhang and co-workers [51] accessed a range of 3,4-fused tricyclic indoles 29, in moderate to excellent yields with good functional group tolerance, including various ring sizes and various linkers (Scheme 13). Initially, the authors reported a catalyst system without ligand; most of them reacted well and delivered the desired tricyclic indole products in good yields. However, some substrates give lower yields or dramatically shut down the reaction. When an electron-rich and bulky monodentate phosphine ligand P(o-tol)3 was added, it showed a better performance than no ligand, such as (29g,29r29u). Notably, the reaction was highlighted by the synthesis of poly(ADP-ribose) polymerase-1 (PARP-1) Inhibitor rucaparib (AG-014699) in 87% yield, for the treatment of advanced solid tumors [52,53].
In Yu and Jiang group’s work [54], they mainly reported carbon-tethered 3,4-fused tricyclic indoles in moderate to excellent yields with wide functional group tolerance (Scheme 14). However, only one example involved oxygen-tethered 3,4-fused tricyclic indole, 31f, 75% yield. Unfortunately, attempts with alkyl-substituted and unsubstituted alkynes were unsuccessful.
In the same year, Fan, He, and co-workers [55] successfully developed a novel three-component reaction for rapid assembly of 3,4-fused tricyclic indole derivatives from ortho-iodoanilines 32, alkynyl iodides 33, and cyclohexadienimines 34 using Pd(0) as a tandem catalysis (Scheme 15). Notably, the three starting materials are cheap and readily available. From the insight of mechanics, firstly, this tandem process involves a Pd-catalyzed intermolecular coupling/cyclization/Michael addition and a new chemical bond at the position of C3 and C4 was built. Then, a seven-membered C3, C4 bridge was erected through intramolecular cyclization with Pd(0)-catalysis, and the desired tricyclic indoles were obtained via rearomatization with 1 M aqueous hydrochloric acid. A series of ortho-iodoanilines could be tolerated well and delivered desired products (35l35o) in acceptable yields. Moreover, iodoalkynes with different substituents such as also reacted smoothly under the standard reaction conditions yielding the desired products (aryl 35e35i, heterocycle 35j, and alkyl 35k). Notably, HCO2Na plays a critical role in this transformation, which is used to trigger the necessary transformation in the catalytic cycles.
Palladium/norbornene (NBE) chemistry, also known as the Catellani reaction, is widely used in natural product synthesis, complex molecules, functional materials, and other fields, which involves a Pd(0)-catalyzed bifunctionalization of aryl halides with various electrophilic reagents and terminal reagents in the presence of NBE [56]. Seminal works by Catellani, Lautens, Dong, Gu, Liang, Luan, and other groups [57] revealed that a series of polysubstituted aromatics, carbocyclic, and heterocyclic compounds could be obtained by Pd/NBE-catalyzed difunctionalization of aryl halides. In 2021, Zhang, Liang, Li, Quan, and co-workers [58] disclosed an elegant method for the synthesis of 3,4-fused tricyclic indoles via a N–S bond cleavage strategy in one step using palladium and NBE in a co-catalysis (Scheme 16). After extensively screening different amino protecting groups, para-methoxy benzenesulfonyl was used as the optimal protecting group, giving the desired product in moderate yield. Regarding the substrate scope, both six- and seven-membered rings could be formed with carbon and oxygen as a linker (38a38x). Aryl iodides bearing various groups were tolerated well, yielding the desired products (methoxy 38q, fluoro 38r, chloro 38s, trifluoromethyl 38t, and ester 38q). Aryl alkynes bearing different groups in any site of the benzene ring also reacted smoothly under standard conditions to give the expected products 38a38k. Unfortunately, alkyl alkynes and terminal alkynes were not suitable coupling partners for this reaction. Notably, corresponding products can also be obtained by replacing the methyl with ethyl (38w) and propyl (38x). Finally, the proposed mechanism was studied by control experiments and density functional theory (DFT) calculations.
Very recently, Luan and co-workers [59] reported an efficient Pd(0)-catalyzed [2 + 2 + 1] annulation of ortho-alkyne-tethered aryl iodides 39 with tertiary hydroxylamines 40 for the synthesis of 3,4-fused tricyclic indoles 41, which are widely present in various natural products and medicine molecules (Scheme 17). Compared to their previous work [45], tertiary hydroxylamines were used as nitrogen sources, which are more accessible and more stable. To obtain the optimal reaction conditions, various tertiary hydroxylamines were tested to harness the modest synthetic efficiency at this stage. Using N-benzyl-N-methyl-benzyl substituent hydroxylamine (40e) could greatly improve the performance of titled conversion, affording the desired tricyclic indole (41e) in 89% yield through C-N bond cleavage of tertiary hydroxylamine with loss of a larger alkyl group. A series of aryl iodides could be well-tolerated and delivered tricyclic indoles (41h41n) in 63–89% yields. Furthermore, with respect to the alkyne terminus, the aromatic rings bearing various substituents also reacted well, affording the desired products (methoxy 41o, 41s and 41t, fluoro 41p, chloro 41q, and cyano 41r) in good to excellent yields. Particularly, heteroarenes such as pyridine and thiophene reacted with (40e) to afford (41v) and (41u) in 87% and 85% yields, respectively. Notably, this novel method was highlighted by a large-scale reaction and the removal of an alkyl group. After systematic mechanistic studies, the selective C(sp3)-N bond cleavage mechanism is inclined to the SN1 pathway.

4. Application in Total Synthesis of Representative Natural Products

As mentioned above, tricyclic indoles are widely embedded in a number of biologically active natural products. Thus, tremendous efforts have been devoted to the synthesis of some representative natural products using these established methods. Chloropeptin I is an anti-HIV agent obtained from Streptomyces sp. WK-3419 [60,61,62]. In 2003, Hoveyda, Snapper, and co-workers [63] reported the first stereoselective total synthesis of anti-HIV agent chloropeptin I by involving palladium-promoted oxidative addition, C-H cleavage, and reductive elimination (Scheme 18a). Undoubtedly, this example is practical and efficient, although a 1.0 equivalent of Pd(PtBu3)2 was used. Lysergic acid is one of important members of ergot alkaloids [64], which oriented its unique tetracyclic skeleton containing a [cd]-fused indole, and the asymmetric total syntheses was first achieved in 2004 by the Szántay group [65]. Five years later, Fukuyama and co-workers [66,67] also reported a novel method for synthesis of this natural product. In 2011, Ohno, Fujii, and co-workers [68] realized another method for the asymmetric syntheses of lysergic acid (Scheme 18b). Tricyclic indole 44 was obtained in 80% yield via Pd(0)-catalyzed domino cyclization of the allene-tethered indole 43. As the analogs of chloropeptin, complestatin A and B were obtained by the fermentation of Streptomyces sp. MA7234; Singh and co-workers [69] disclosed it as inhibitors of HIV-1 integrase in 2001, and elucidated its structure and partial stereochemistry. In 2011, Boger and co-workers [70] devised an elegant strategy to achieve the synthesis of complestatin A and B via intramolecular Lacork indole cyclization (Scheme 18c).
N-methylwelwitindolinone C isothiocyanate is a member of welwitindolinone alkaloids, which is isolated from the blue-green algae Hapalosiphon wetwitschii, Westiella intricata, Fischerella muscicola, and Fischerella major [71,72]. In 2015, Hatakeyama and co-workers [73] developed a Pd(0)-catalyzed tandem enolate strategy for the total synthesis of N-methylwelwitindolinone C isothiocyanate (Scheme 19a). Notably, a tetracyclic indole skeleton 48 could be obtained from 47 in 100% yields with 2:1 stereoselectivity. In 2017, Jia and co-workers [74] demonstrate a Pd/Me-phos-catalyzed intramolecular Larock indole annulation/Tsuji-Trost allylation cascade reaction to assemble tetracyclic indole product fumigaclavine G in one step with good yields (Scheme 19b). Moreover, festuclavine, pyroclavine, costaclavine, epi-costaclavine, pibocin A, 9-deacetoxyfumigaclavine C, and dihydrosetoclavine, a member of the ergot alkaloids family, were also totally synthesized with this established method. In the same year, Werz and co-workers [75] reported the facile enantioselective total synthesis of (+)-lysergol via Pd(0)-catalyzed anti-carbopalladation/Heck cascade process and further eight steps with an overall yield of 13% (Scheme 19c).
Dehydrobufotenine, a member of the 1,3,4,5-tetrahydropyrrolo [4,3,2-de]quinoline ring system natural products, was originally isolated from the parotid glands of the toad Bufo marinus [76]. In 1996, based on their previous studies about the C–N bond forming reaction, Buchwald and co-workers [77] successfully developed a palladium-catalyzed efficient protocol for the total synthesis of dehydrobufotenine from readily available substrate 53 via intramolecular Buchwald-Hartwig amination/demethylation/methylation reactions. In this manner, dehydrobufotenine was synthesized via a three-step transformation from 53, affording the desired product in 40.5% overall yield (Scheme 20a). As a member of marine natural products, Ambiguine H produced from cyanobacteria have shown excellent biological activities such as antibacterial and anticancer. Considering the importance of the scaffold in pharmaceutical and natural products, numerous efficient methods have been developed to afford it. In 2007, Baran and co-workers [78] developed a practical method for the enantioselective total syntheses of tetracyclic indole product (+)-ambiguine H via a palladium-catalyzed intramolecular Heck annulation followed by further transformation (Scheme 20b). Compared to previous methods, this method featured no need of a protecting group, shorter steps, and single enantiomers. Moreover, this method is also applicable for synthesis of hapalindole U, welwitindolinone A, and fischerindole I from commercially available materials, where fewer steps were involved in these routes. Cycloclavine, a member of clavine-type alkaloids, was first isolated in 1969 from the seeds of the African morning glory [79]. Over the past few years, only two racemic examples have been reported for the synthesis of cycloclavine [80,81]. So, synthesis of cycloclavine molecules using an efficient strategy is necessary. In 2013, Cao and co-workers [82] reported an efficient and concise strategy for the formal synthesis of (±)-cycloclavine with an intramolecular Heck reaction and iron-catalyzed aza-Cope-Mannich cyclization as key steps.

5. Conclusions

3,n-fused tricyclic indoles are widespread in various natural products, drug molecules, and material sciences. Given the importance of these skeletons and complexity of their structures, assembling these skeletons has presented numerous interests and challenges to synthetic chemists. Over the past decades, considerable efforts have been devoted to develop synthetic methods for producing diversified 3,n-fused tricyclic indole derivatives with transition metals as effective catalysts. In particular, Pd(0) complexes stand out as prominent catalysts in the annulation of alkyne-tethered ortho-haloanilines or aryl iodides with various amination agents, and have received much attention. In this review, we mainly summarized the recent progress in the synthesis of 3,n-fused tricyclic indoles with a Pd catalyst, and its applications in the total synthesis of representative natural products. By providing an overview of palladium catalysis to access various 3,n-fused tricyclic indole derivatives, we believe that this review will inspire the further developments of novel strategies to achieve tricyclic indoles by using mild and sustainable conditions, extended the space of these tricyclic indole molecules, and potentially given access to new compounds.

Author Contributions

Conceptualization, L.F., X.Z. and X.L.; writing—original draft preparation, L.F., F.H. and X.Y.; writing—review and editing, L.F. and C.X.; visualization, G.Y.; supervision, X.Y. and C.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Program of National Natural Science Foundation of China (22201066), the Natural Science Foundation of Henan Province (222300420188, 222300420456, 222300420459), the Science and Technology Project of Henan Province (212102311063), and the special fund for topnotch talents in Henan Agricultural University (30500925, 30500740).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

F.L. thanks his family for their unrequited support and encouragement in the past few years. We are grateful to my colleagues, students, and friends for their support and help.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

DtBPF1,1′-Bis(di-tert-butylphosphino)ferrocene
TEStriethylsilyl
TMStrimethylsilyl
DMFN,N-dimethylformamide
DMAN,N-dimethylacetamide
DCMdichloromethane
THFtetrahydrofuran
DMSOdimethyl sulfoxide
TFAtrifluoroacetic acid
dbadibenzylideneacetone
m-CPBAm-chloroperbenzoic acid
Me-phosdicyclohexyl(2′-methyl-[1,1′-biphenyl]-2-yl)phosphane
DPPP1,3-bis(diphenylphosphaneyl)propane
TBABtetrabutylammonium bromide
SN1unimolecular nucleophilic substitution
Boctert-butoxycarbonyl
SCNisothiocyanato
DDQ4,5-dichloro-3,6-dioxocyclohexa-1,4-diene-1,2-dicarbonitrile
OTBStert-butyldimethylsilyloxy
PCCpyridinium chlorochromate
PMBpara-methoxlphenyl
DFTdensity functional theory
o-PBAortho-phenyl benzoic acid
DPPBz1,2-Bis(diphenylphosphino)benzene
NBEnorbornene
NMRnuclear magnetic resonance
HRMShigh resolution mass spectrometry
Xphosdicyclohexyl(2′,4′,6′-triisopropyl-[1,1′-biphenyl]-2-yl)phosphane

References

  1. Cacchi, S.; Fabrizi, G. Synthesis and functionalization of indoles through palladium-catalyzed reactions. Chem. Rev. 2005, 105, 2873–2920. [Google Scholar] [CrossRef] [PubMed]
  2. Humphrey, G.R.; Kuethe, J.T. Practical methodologies for the synthesis of indoles. Chem. Rev. 2006, 106, 2875–2911. [Google Scholar] [CrossRef]
  3. Gribble, G.W. Recentdevelopments in indole ring synthesis—Methodology and applications. J. Chem. Soc. Perkin Trans. 2000, 1, 1045–1075. [Google Scholar] [CrossRef]
  4. Prabagar, B.; Yang, Y.; Shi, Z. Site-selective C-H functionalization to access the arene backbone of indoles and quinolines. Chem. Soc. Rev. 2021, 50, 11249–11269. [Google Scholar] [CrossRef] [PubMed]
  5. Nagaraju, K.; Ma, D. Oxidative coupling strategies for the synthesis of indole alkaloids. Chem. Soc. Rev. 2018, 47, 8018–8029. [Google Scholar] [CrossRef] [PubMed]
  6. Bariwal, J.; Voskressensky, L.G.; Van der Eycken, E.V. Recent advances in spirocyclization of indole derivatives. Chem. Soc. Rev. 2018, 47, 3831–3848. [Google Scholar] [CrossRef] [PubMed]
  7. Nemoto, T.; Harada, S.; Nakajima, M. Synthetic methods for 3,4-fused tricyclic indoles via indole ring formation. Asian J. Org. Chem. 2018, 7, 1730–1742. [Google Scholar] [CrossRef]
  8. Connon, R.; Guiry, P.J. Recent advances in the development of one-pot/multistep syntheses of 3,4-annulated indoles. Tetrahedron Lett. 2020, 61, 151696. [Google Scholar] [CrossRef]
  9. Scopton, A.; Kelly, T.R. Synthesis of HKI 0231B. J. Org. Chem. 2005, 70, 10004–10012. [Google Scholar] [CrossRef]
  10. Ginnon, W.F.; Benigni, J.D.; Suzuki, J.; Daly, J.W. The synthesis of dehydrobufotenine. Tetrahedron Lett. 1967, 8, 1531–1533. [Google Scholar] [CrossRef]
  11. Sahu, S.; Das, B.; Maji, M.S. Stereodivergent total synthesis of hapalindoles, fischerindoles, hapalonamide H, and ambiguine H alkaloids by developing a biomimetic, redox-neutral, cascade Prins-type cyclization. Org. Lett. 2018, 20, 6485–6489. [Google Scholar] [CrossRef] [PubMed]
  12. Wyche, T.P.; Ruzzini, A.C.; Schwab, L.; Currie, C.R.; Clardy, J. Tryptorubin A: A polycyclic peptide from a fungus-derived streptomycete. J. Am. Chem. Soc. 2017, 139, 12899–12902. [Google Scholar] [CrossRef] [PubMed]
  13. Ma, B.; Litvinov, D.N.; He, L.; Banerjee, B.; Castle, S.L. Total synthesis of celogentin C. Angew. Chem. Int. Ed. 2009, 48, 6104–6107. [Google Scholar] [CrossRef]
  14. Garfunkle, J.; Kimball, F.S.; Trzupek, J.D.; Takizawa, S.; Shimamura, H.; Tomishima, M.; Boger, D.L. Total synthesis of chloropeptin II (complestatin) and chloropeptin I. J. Am. Chem. Soc. 2009, 131, 16036–16038. [Google Scholar] [CrossRef]
  15. Isley, N.A.; Endo, Y.; Wu, Z.-C.; Covington, B.C.; Bushin, L.B.; Seyedsayamdost, M.R.; Boger, D.L. Total synthesis and stereochemical assignment of streptide. J. Am. Chem. Soc. 2019, 141, 17361–17369. [Google Scholar] [CrossRef] [PubMed]
  16. Horwell, D.C.; Nichols, P.D.; Ratcliffe, G.S.; Roberts, E. Synthesis of conformationally constrained tryptophan derivatives. J. Org. Chem. 1994, 59, 4418–4423. [Google Scholar] [CrossRef]
  17. Biffis, A.; Centomo, P.; Del Zotto, A.; Zecca, M. Pd metal catalysts for cross-couplings and related reactions in the 21st century: A critical review. Chem. Rev. 2018, 118, 2249–2295. [Google Scholar] [CrossRef]
  18. Shetgaonkar, S.E.; Mamgain, R.; Kikushima, K.; Dohi, T.; Singh, F.V. Palladium-catalyzed organic reactions involving hypervalent iodine reagents. Molecules 2022, 27, 3900. [Google Scholar] [CrossRef]
  19. Heck, R.F. Acylation, methylation, and carboxyalkylation of olefins by Group VIII metal derivative. J. Am. Chem. Soc. 1968, 90, 5518–5526. [Google Scholar] [CrossRef]
  20. Söderberg, B.C.; Rector, S.R.; O’Neil, S.N. Palladium-catalyzed synthesis of fused indoles. Tetrahenron Lett. 1999, 40, 3657–3660. [Google Scholar] [CrossRef]
  21. Söderberg, B.C.; Hubbard, J.W.; Rector, S.R.; O’Neil, S.N. Synthesis offused indoles by sequential palladium-catalyzed Heck reaction and N-heteroannulation. Tetrahedron 2005, 61, 3637–3649. [Google Scholar] [CrossRef]
  22. Peshkov, V.A.; Van Hove, S.; Donets, P.A.; Pereshivko, O.P.; Van Hecke, K.; Van Meervelt, L.; Van der Eycken, E.V. Synthesis of the azocino[cd]indole framework through Pd-catalyzed intramolecular acetylene hydroarylation. Eur. J. Org. Chem. 2011, 2011, 1837–1840. [Google Scholar] [CrossRef]
  23. Larock, R.C.; Yum, E.K. Synthesis of indoles via palladium-catalyzed heteroannulation of internal alkynes. J. Am. Chem. Soc. 1991, 113, 6689–6690. [Google Scholar] [CrossRef]
  24. Larock, R.C.; Yum, E.K.; Refvik, M.D. Synthesis of 2,3-disubstituted indoles via palladium-catalyzed annulation of internal alkynes. J. Org. Chem. 1998, 63, 7652–7662. [Google Scholar] [CrossRef]
  25. Breazzano, S.P.; Poudel, Y.B.; Boger, D.L. A Pd(0)-mediated indole (macro)cyclization reaction. J. Am. Chem. Soc. 2013, 135, 1600–1606. [Google Scholar] [CrossRef] [PubMed]
  26. Mohr, P.J.; Halcomb, R.L. Total synthesis of (+)-phomactin A using a B-alkyl Suzuki macrocyclization. J. Am. Chem. Soc. 2003, 125, 1712–1713. [Google Scholar] [CrossRef]
  27. Nicolaou, K.C.; Bulger, P.G.; Sarlah, D. Palladium-catalyzed cross-coupling reactions in total synthesis. Angew. Chem. Int. Ed. 2005, 44, 4442–4489. [Google Scholar] [CrossRef]
  28. Shan, D.; Gao, Y.; Jia, Y. Intramolecular Larock indole synthesis: Preparation of 3,4-fused tricyclic indoles and total synthesis of fargesine. Angew. Chem. Int. Ed. 2013, 52, 4902–4905. [Google Scholar] [CrossRef] [PubMed]
  29. Gao, Y.; Shan, D.; Jia, Y. Intramolecular Larock indole synthesis for the preparation of tricyclic indoles and its application in the synthesis of tetrahydropyrroloquinoline and fargesine. Tetrahedron 2014, 70, 5136–5141. [Google Scholar] [CrossRef]
  30. Zeng, J.; Sun, G.; Yao, W.; Zhu, Y.; Wang, R.; Cai, L.; Liu, K.; Zhang, Q.; Liu, X.-W.; Wan, Q. 3-Aminodeoxypyranoses in glycosylation: Diversity-oriented synthesis and assembly in oligosaccharides. Angew. Chem. Int. Ed. 2017, 56, 5227–5231. [Google Scholar] [CrossRef]
  31. Xu, Q.-L.; Dai, L.-X.; You, S.-L. Diversity oriented synthesis of indole-based peri-annulated compounds via allylic alkylation reactions. Chem. Sci. 2013, 4, 97–102. [Google Scholar] [CrossRef]
  32. Byun, S.; Farah, A.O.; Wise, H.R.; Katchmar, A.; Cheong, P.H.Y.; Scheidt, K.A. Enantioselective copper-catalyzed borylative amidation of allenes. J. Am. Chem. Soc. 2022, 144, 22850–22857. [Google Scholar] [CrossRef] [PubMed]
  33. Li, M.; Dixon, D.J. Stereoselective spirolactam synthesis via palladium catalyzed arylative allene carbocyclization cascades. Org. Lett. 2010, 12, 3784–3787. [Google Scholar] [CrossRef]
  34. Ma, S.; Zhao, S. Pd(0)-catalyzed insertion−cyclization reaction of 2,3-allenols with aryl or alkenyl halides. Diastereoselective synthesis of highly optically active trans-2,3-disubstituted vinylic oxiranes. J. Am. Chem. Soc. 1999, 121, 7943–7944. [Google Scholar] [CrossRef]
  35. Chen, S.; Gao, Z.; Zhao, H.; Li, B. Palladium-catalyzed regio- and stereoselective arylesterification of ferrocenylallene: A synthetic route to ferrocene-containing disubstituted E-allylic esters. J. Org. Chem. 2014, 79, 1481–1486. [Google Scholar] [CrossRef]
  36. Lee, M.; Nguyen, M.; Brandt, C.; Kaminsky, W.; Lalic, G. Catalytic Hydroalkylation of allenes. Angew. Chem. Int. Ed. 2017, 56, 15703–15707. [Google Scholar] [CrossRef]
  37. Nakano, S.; Inoue, N.; Hamada, Y.; Nemoto, T. Pd-catalyzed cascade cyclization by intramolecular Heck insertion of an allene-allylic amination sequence: Application to the synthesis of 3,4-fused tricyclic indoles. Org. Lett. 2015, 17, 2622–2625. [Google Scholar] [CrossRef]
  38. Walsh, C.T.; O’Brien, R.V.; Khosla, C. Nonproteinogenic amino acid building blocks for nonribosomal peptide and hybrid polyketide scaffolds. Angew. Chem. Int. Ed. 2013, 52, 7098–7124. [Google Scholar] [CrossRef]
  39. Cardote, T.A.F.; Ciulli, A. Cyclic and macrocyclic peptides as chemical tools to recognise protein surfaces and probe protein–protein interactions. ChemMedChem 2016, 11, 787–794. [Google Scholar] [CrossRef]
  40. Marsault, E.; Peterson, M.L. Macrocycles are great cycles: Applications, opportunities, and challenges of synthetic macrocycles in drug discovery. J. Med. Chem. 2011, 54, 1961–2004. [Google Scholar] [CrossRef]
  41. Zhang, X.; Lu, G.; Sun, M.; Mahankali, M.; Ma, Y.; Zhang, M.; Hua, W.; Hu, Y.; Wang, Q.; Chen, J.; et al. A general strategy for synthesis of cyclophane-braced peptide macrocycles via palladium-catalysed intramolecular sp3 C-H arylation. Nat. Chem. 2018, 10, 540–548. [Google Scholar] [CrossRef]
  42. Bellina, F.; Rossi, R. Transition metal-catalyzed direct arylation of substrates with activated sp3-hybridized C-H bonds and some of their synthetic equivalents with aryl halides and pseudohalides. Chem. Rev. 2010, 110, 1082–1146. [Google Scholar] [CrossRef]
  43. Hellal, M.; Singh, S.; Cuny, G.D. Synthesis of tetracyclic indoles via intramolecular α-arylation of ketones. J. Org. Chem. 2012, 77, 4123–4130. [Google Scholar] [CrossRef] [PubMed]
  44. Rodríguez, J.F.; Marchese, A.D.; Lautens, M. Palladium-catalyzed synthesis of dihydrobenzoindolones via C-H bond activation and alkyne insertion. Org. Lett. 2018, 20, 4367–4370. [Google Scholar] [CrossRef] [PubMed]
  45. Fan, L.; Hao, J.; Yu, J.; Ma, X.; Liu, J.; Luan, X. Hydroxylamines as bifunctional single-nitrogen sources for the rapid assembly of diverse tricyclic indole scaffolds. J. Am. Chem. Soc. 2020, 142, 6698–6707. [Google Scholar] [CrossRef]
  46. Qin, H.; Xu, Z.; Cui, Y.; Jia, Y. Total Synthesis of (±)-Decursivine and (±)-Serotobenine: A Witkop photocyclization/elimination/O-Michael addition cascade approach. Angew. Chem. Int. Ed. 2011, 50, 4447–4449. [Google Scholar] [CrossRef]
  47. Smith, A.B.; Kanoh, N.; Ishiyama, H.; Minakawa, N.; Rainier, J.D.; Hartz, R.A.; Cho, Y.S.; Cui, H.; Moser, W.H. Tremorgenic Indole Alkaloids. The total synthesis of (−)-penitrem D. J. Am. Chem. Soc. 2003, 125, 8228–8237. [Google Scholar] [CrossRef] [PubMed]
  48. Ng, K.-H.; Chan, A.S.C.; Yu, W.-Y. Pd-catalyzed intermolecular ortho-C-H amidation of anilides by N-nosyloxycarbamate. J. Am. Chem. Soc. 2010, 132, 12862–12864. [Google Scholar] [CrossRef]
  49. Zheng, H.; Zhu, Y.; Shi, Y. Palladium(0)-catalyzed Heck reaction/C-H activation/amination sequence with diaziridinone: A facile approach to indolines. Angew. Chem. Int. Ed. 2014, 53, 11280–11284. [Google Scholar] [CrossRef]
  50. Sun, X.; Wu, Z.; Qi, W.; Ji, X.; Cheng, C.; Zhang, Y. Synthesis of indolines by palladium-catalyzed intermolecular amination of unactivated C(sp3)-H bonds. Org. Lett. 2019, 21, 6508–6512. [Google Scholar] [CrossRef]
  51. Cheng, C.; Zuo, X.; Tu, D.; Wan, B.; Zhang, Y. Synthesis of 3,4-fused tricyclic indoles through cascade carbopalladation and C–H amination: Development and total synthesis of rucaparib. Org. Lett. 2020, 22, 4985–4989. [Google Scholar] [CrossRef] [PubMed]
  52. Porcelli, L.; Quatrale, A.E.; Mantuano, P.; Leo, M.G.; Silvestris, N.; Rolland, J.F.; Carioggia, E.; Lioce, M.; Paradiso, A.; Azzariti, A. Optimize radiochemotherapy in pancreatic cancer: PARP inhibitors a new therapeutic opportunity. Mol. Oncol. 2013, 7, 308–322. [Google Scholar] [CrossRef] [PubMed]
  53. Plummer, R.; Lorigan, P.; Steven, N.; Scott, L.; Middleton, M.R.; Wilson, R.H.; Mulligan, E.; Curtin, N.; Wang, D.; Dewji, R.; et al. A phase II study of the potent PARP inhibitor, rucaparib (PF-01367338, AG014699), with temozolomide in patients with metastatic melanoma demonstrating evidence of chemopotentiation. Cancer Chemother. Pharmacol. 2013, 71, 1191–1199. [Google Scholar] [CrossRef] [PubMed]
  54. Zhang, L.; Chen, J.; Zhong, T.; Zheng, X.; Zhou, J.; Jiang, X.; Yu, C. Palladium-catalyzed [2 + 2 + 1] annulation of alkyne-tethered aryl iodides with diaziridinone: Synthesis of 3,4-fused tricyclic indoles. J. Org. Chem. 2020, 85, 10823–10834. [Google Scholar] [CrossRef]
  55. Wang, W.; Yang, M.; Han, D.; He, Q.; Fan, R. Tandem palladium catalysis for rapid construction of 3,4-fused tricyclic indoles. Adv. Synth. Catal. 2020, 362, 1281–1285. [Google Scholar] [CrossRef]
  56. Catellani, M.; Frignani, F.; Rangoni, A. A complex catalytic cycle leading to a regioselective synthesis of o,o′-disubstituted vinylarenes. Angew. Chem. Int. Ed. 1997, 36, 119–122. [Google Scholar] [CrossRef]
  57. Wang, J.; Dong, G. Palladium/norbornene cooperative catalysis. Chem. Rev. 2019, 119, 7478–7528. [Google Scholar] [CrossRef]
  58. Zhang, B.-S.; Wang, F.; Gou, X.-Y.; Yang, Y.-H.; Jia, W.-Y.; Liang, Y.-M.; Wang, X.-C.; Li, Y.; Quan, Z.-J. Palladium-catalyzed synthesis of tricyclic indoles via a N–S bond cleavage strategy. Org. Lett. 2021, 23, 7518–7523. [Google Scholar] [CrossRef]
  59. Wu, J.; Li, L.; Liu, M.; Bai, L.; Luan, X. Selective C(sp3)-N bond cleavage of N,N-dialkyl tertiary amines with the loss of a large alkyl group via an SN1 pathway. Angew. Chem. Int. Ed. 2022, 61, e202113820. [Google Scholar]
  60. Gouda, H.; Matsuzaki, K.; Tanaka, H.; Hirono, S.; Ōmura, S.; McCauley, J.A.; Sprengeler, P.A.; Furst, G.T.; Smith, A.B. Stereostructure of (−)-chloropeptin I, a novel inhibitor of gp120−CD4 binding, via high-temperature molecular dynamics, Monte Carlo conformational searching, and NMR spectroscopy. J. Am. Chem. Soc. 1996, 118, 13087–13088. [Google Scholar] [CrossRef]
  61. Matsuzaki, K.; Ikeda, H.; Ogino, T.; Matsumoto, A.; Woodruff, H.B.; Tanaka, H.; Omura, S. Chloropeptin-I and chloropeptin-II, novel inhibitors against GP120-CD4 binding from Streptomyces sp. J. Antibiot. 1994, 47, 1173–1174. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Matsuzaki, K.; Ogino, T.; Sunazuka, T.; Tanaka, H.; Omura, S. Chloropeptins, new anti-HIV antibiotics inhibiting gp120-CD4 binding from Streptomyces sp. II. Structure elucidation of chloropeptin I. J. Antibiot. 1997, 50, 66–69. [Google Scholar] [CrossRef] [PubMed]
  63. Deng, H.; Jung, J.-K.; Liu, T.; Kuntz, K.W.; Snapper, M.L.; Hoveyda, A.H. Total synthesis of anti-HIV agent chloropeptin I. J. Am. Chem. Soc. 2003, 125, 9032–9034. [Google Scholar] [CrossRef] [PubMed]
  64. de Groot, A.N.J.A.; van Dongen, P.W.J.; Vree, T.B.; Hekster, Y.A.; van Roosmalen, J. Ergot alkaloids. Drugs 1998, 56, 523–535. [Google Scholar] [CrossRef]
  65. Moldvai, I.; Temesvári-Major, E.; Incze, M.; Szentirmay, É.; Gács-Baitz, E.; Szántay, C. Enantioefficient synthesis of α-ergocryptine:  First direct synthesis of (+)-lysergic acid. J. Org. Chem. 2004, 69, 5993–6000. [Google Scholar] [CrossRef]
  66. Kurokawa, T.; Isomura, M.; Tokuyama, H.; Fukuyama, T. Synthesis of lysergic acid methyl ester via the double cyclization strategy. Synlett 2009, 2009, 775–778. [Google Scholar]
  67. Inoue, T.; Yokoshima, S.; Fukuyama, T. Synthetic studies toward (+)-lysergic acid: Construction of the tetracyclic ergoline skeleton. Hetetocycles 2009, 79, 373–378. [Google Scholar]
  68. Iwata, A.; Inuki, S.; Oishi, S.; Fujii, N.; Ohno, H. Formal total synthesis of (+)-lysergic acid via zinc(II)-mediated regioselective ring-opening reduction of 2-alkynyl-3-indolyloxirane. J. Org. Chem. 2011, 76, 5506–5512. [Google Scholar] [CrossRef]
  69. Singh, S.B.; Jayasuriya, H.; Salituro, G.M.; Zink, D.L.; Shafiee, A.; Heimbuch, B.; Silverman, K.C.; Lingham, R.B.; Genilloud, O.; Teran, A.; et al. The complestatins as HIV-1 integrase inhibitors. Efficient isolation, structure elucidation, and inhibitory activities of isocomplestatin, chloropeptin I, new complestatins, A and B, and acid-hydrolysis products of chloropeptin I. J. Nat. Prod. 2001, 64, 874–882. [Google Scholar] [CrossRef]
  70. Breazzano, S.P.; Boger, D.L. Synthesis and stereochemical determination of complestatin A and B (neuroprotectin A and B). J. Am. Chem. Soc. 2011, 133, 18495–18502. [Google Scholar] [CrossRef]
  71. Stratmann, K.; Moore, R.E.; Bonjouklian, R.; Deeter, J.B.; Patterson, G.M.L.; Shaffer, S.; Smith, C.D.; Smitka, T.A. Welwitindolinones, unusual alkaloids from the blue-green algae Hapalosiphon welwitschii and Westiella intricata. Relationship to fischerindoles and hapalinodoles. J. Am. Chem. Soc. 1994, 116, 9935–9942. [Google Scholar] [CrossRef]
  72. Jimenez, J.I.; Huber, U.; Moore, R.E.; Patterson, G.M.L. Oxidized welwitindolinones from terrestrial Fischerella spp. J. Nat. Prod. 1999, 62, 569–572. [Google Scholar] [CrossRef]
  73. Komine, K.; Nomura, Y.; Ishihara, J.; Hatakeyama, S. Total synthesis of (−)-N-methylwelwitindolinone C isothiocyanate based on a Pd-catalyzed tandem enolate coupling strategy. Org. Lett. 2015, 17, 3918–3921. [Google Scholar] [CrossRef] [PubMed]
  74. Liu, H.; Zhang, X.; Shan, D.; Pitchakuntla, M.; Ma, Y.; Jia, Y. Total syntheses of festuclavine, pyroclavine, costaclavine, epi-costaclavine, pibocin A, 9-deacetoxyfumigaclavine C, fumigaclavine G, and dihydrosetoclavine. Org. Lett. 2017, 19, 3323–3326. [Google Scholar] [CrossRef] [PubMed]
  75. Milde, B.; Pawliczek, M.; Jones, P.G.; Werz, D.B. Enantioselective total synthesis of (+)-lysergol: A formal anti-carbopalladation/Heck cascade as the key step. Org. Lett. 2017, 19, 1914–1917. [Google Scholar] [CrossRef] [PubMed]
  76. Marki, F.; Robertson, A.V.; Witkop, B. Dehydrobufotenine, a novel type of tricyclic serotonin metabolite from Bufo Marinus. J. Am. Chem. Soc. 1961, 83, 3341–3342. [Google Scholar] [CrossRef]
  77. Peat, A.J.; Buchwald, S.L. Novel Syntheses of Tetrahydropyrroloquinolines: Applications to alkaloid synthesis. J. Am. Chem. Soc. 1996, 118, 1028–1030. [Google Scholar] [CrossRef]
  78. Baran, P.S.; Maimone, T.J.; Richter, J.M. Total synthesis of marine natural products without using protecting groups. Nature 2007, 446, 404–408. [Google Scholar] [CrossRef]
  79. Stauffacher, D.; Niklaus, P.; Tscherter, H.; Weber, H.P.; Hofmann, A. Cycloclavin, ein neues alkaloid aus Ipomoea hildebrandtii vatke—71: Mutterkornalkaloide. Tetrahedron 1969, 25, 5879–5887. [Google Scholar] [CrossRef]
  80. Incze, M.; Dörnyei, G.; Moldvai, I.; Temesvaári-Major, E.; Egyed, O.; Szaántay, C. New routes to clavine-type ergot alkaloids. Part 2: Synthesis of the last, so far not yet synthesized member of the clavine alkaloid family, (±)-cycloclavine. Tetrahedron 2008, 64, 2924–2929. [Google Scholar] [CrossRef]
  81. Petronijevic, F.R.; Wipf, P. Total synthesis of (±)-cycloclavine and (±)-5-epi-cycloclavine. J. Am. Chem. Soc. 2011, 133, 7704–7707. [Google Scholar] [CrossRef] [PubMed]
  82. Wang, W.; Lu, J.-T.; Zhang, H.-L.; Shi, Z.-F.; Wen, J.; Cao, X.-P. Formal synthesis of (±)-cycloclavine. J. Org. Chem. 2014, 79, 122–127. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. Representative examples of biologically active molecules containing tricyclic indole cores.
Scheme 1. Representative examples of biologically active molecules containing tricyclic indole cores.
Molecules 28 01647 sch001
Scheme 2. Synthesis of 3,4-fused tricyclic indoles via two consecutive palladium-catalyzed reactions.
Scheme 2. Synthesis of 3,4-fused tricyclic indoles via two consecutive palladium-catalyzed reactions.
Molecules 28 01647 sch002
Scheme 3. Synthesis of 3,4-fused tricyclic indoles through palladium-catalyzed intramolecular acetylene hydroarylation.
Scheme 3. Synthesis of 3,4-fused tricyclic indoles through palladium-catalyzed intramolecular acetylene hydroarylation.
Molecules 28 01647 sch003
Scheme 4. Synthesis of diversely functionalized 3,4-, 3,5-, and 3,6-fused tricyclic indoles 7 with the alkyne tethered ortho-bromoanilines.
Scheme 4. Synthesis of diversely functionalized 3,4-, 3,5-, and 3,6-fused tricyclic indoles 7 with the alkyne tethered ortho-bromoanilines.
Molecules 28 01647 sch004
Scheme 5. Synthesis of 3,4-fused tricyclic indoles via palladium-catalyzed intramolecular annulation of internal alkyne tethered ortho-iodoanilines and total synthesis of fargesine.
Scheme 5. Synthesis of 3,4-fused tricyclic indoles via palladium-catalyzed intramolecular annulation of internal alkyne tethered ortho-iodoanilines and total synthesis of fargesine.
Molecules 28 01647 sch005
Scheme 6. Diversity oriented synthesis of 3,4-fused tricyclic indole and their derivatives via palladium-catalyzed allylic alkylation reactions.
Scheme 6. Diversity oriented synthesis of 3,4-fused tricyclic indole and their derivatives via palladium-catalyzed allylic alkylation reactions.
Molecules 28 01647 sch006
Scheme 7. Palladium-catalyzed intramolecular annulation of allene tethered ortho-iodoanilines.
Scheme 7. Palladium-catalyzed intramolecular annulation of allene tethered ortho-iodoanilines.
Molecules 28 01647 sch007
Scheme 8. Synthesis of peptide-based macrocycles via palladium-catalyzed intramolecular C(sp3)-H arylation.
Scheme 8. Synthesis of peptide-based macrocycles via palladium-catalyzed intramolecular C(sp3)-H arylation.
Molecules 28 01647 sch008
Scheme 9. Synthesis of tetracyclic indoles via Suzuki reaction and intramolecular α-arylation of ketones.
Scheme 9. Synthesis of tetracyclic indoles via Suzuki reaction and intramolecular α-arylation of ketones.
Molecules 28 01647 sch009
Scheme 10. Synthesis of 3,4-fused tricyclic indolones via palladium-catalyzed cascade annulation between ortho-iodoacrylamides 21 and unsymmetrical internal alkynes 22.
Scheme 10. Synthesis of 3,4-fused tricyclic indolones via palladium-catalyzed cascade annulation between ortho-iodoacrylamides 21 and unsymmetrical internal alkynes 22.
Molecules 28 01647 sch010
Scheme 11. Palladium-catalyzed [2 + 2 + 1] annulation with internal alkyne tethered aryl iodide 24 and secondary hydroxylamines 25. a 120 °C was used.
Scheme 11. Palladium-catalyzed [2 + 2 + 1] annulation with internal alkyne tethered aryl iodide 24 and secondary hydroxylamines 25. a 120 °C was used.
Molecules 28 01647 sch011
Scheme 12. Proposed mechanism of Pd(0)-catalyzed [2 + 2 + 1] annulation.
Scheme 12. Proposed mechanism of Pd(0)-catalyzed [2 + 2 + 1] annulation.
Molecules 28 01647 sch012
Scheme 13. Synthesis of 3,4-fused tricyclic indoles via Pd(0)-catalyzed annulation of alkyne-tethered aryl iodides with diaziridinone.
Scheme 13. Synthesis of 3,4-fused tricyclic indoles via Pd(0)-catalyzed annulation of alkyne-tethered aryl iodides with diaziridinone.
Molecules 28 01647 sch013
Scheme 14. Pd(0)-catalyzed [2 + 2 + 1] annulation of alkyne-tethered aryl iodides with diaziridinone.
Scheme 14. Pd(0)-catalyzed [2 + 2 + 1] annulation of alkyne-tethered aryl iodides with diaziridinone.
Molecules 28 01647 sch014
Scheme 15. Synthesis of 3,4-fused tricyclic indoles via palladium-catalyzed intermolecular [3 + 3 + 1] annulation of ortho-iodoanilines with alkynyl iodides and cyclohexadienimines.
Scheme 15. Synthesis of 3,4-fused tricyclic indoles via palladium-catalyzed intermolecular [3 + 3 + 1] annulation of ortho-iodoanilines with alkynyl iodides and cyclohexadienimines.
Molecules 28 01647 sch015
Scheme 16. Synthesis of 3,4-fused tricyclic indole derivatives via a N–S bond cleavage strategy with palladium and norbornene in a co-catalysis.
Scheme 16. Synthesis of 3,4-fused tricyclic indole derivatives via a N–S bond cleavage strategy with palladium and norbornene in a co-catalysis.
Molecules 28 01647 sch016
Scheme 17. Synthesis of 3,4-fused tricyclic indoles through selective C(sp3)-N bond cleavage of tertiary hydroxylamines with the loss of a larger alkyl group.
Scheme 17. Synthesis of 3,4-fused tricyclic indoles through selective C(sp3)-N bond cleavage of tertiary hydroxylamines with the loss of a larger alkyl group.
Molecules 28 01647 sch017
Scheme 18. Total synthesis of (a) chloropeptin I, (b) (+)-lysergic acid, and (c) complestatin A and B.
Scheme 18. Total synthesis of (a) chloropeptin I, (b) (+)-lysergic acid, and (c) complestatin A and B.
Molecules 28 01647 sch018
Scheme 19. Total synthesis of (a) N-methylwelwitindolinone C isothiocyanate, (b) fumigaclavine G, and (c) (+)-lysergol.
Scheme 19. Total synthesis of (a) N-methylwelwitindolinone C isothiocyanate, (b) fumigaclavine G, and (c) (+)-lysergol.
Molecules 28 01647 sch019
Scheme 20. Synthesis of (a) dehydrobufotenine, (b) (+)-ambiguine H, and (c) (±)-cycloclavine.
Scheme 20. Synthesis of (a) dehydrobufotenine, (b) (+)-ambiguine H, and (c) (±)-cycloclavine.
Molecules 28 01647 sch020
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

Fan, L.; Zhu, X.; Liu, X.; He, F.; Yang, G.; Xu, C.; Yang, X. Recent Advances in the Synthesis of 3,n-Fused Tricyclic Indole Skeletons via Palladium-Catalyzed Domino Reactions. Molecules 2023, 28, 1647. https://doi.org/10.3390/molecules28041647

AMA Style

Fan L, Zhu X, Liu X, He F, Yang G, Xu C, Yang X. Recent Advances in the Synthesis of 3,n-Fused Tricyclic Indole Skeletons via Palladium-Catalyzed Domino Reactions. Molecules. 2023; 28(4):1647. https://doi.org/10.3390/molecules28041647

Chicago/Turabian Style

Fan, Liangxin, Xinxin Zhu, Xingyuan Liu, Fangyu He, Guoyu Yang, Cuilian Xu, and Xifa Yang. 2023. "Recent Advances in the Synthesis of 3,n-Fused Tricyclic Indole Skeletons via Palladium-Catalyzed Domino Reactions" Molecules 28, no. 4: 1647. https://doi.org/10.3390/molecules28041647

Article Metrics

Back to TopTop