Next Article in Journal
Effects of Insecticides and Microbiological Contaminants on Apis mellifera Health
Next Article in Special Issue
Design, Synthesis, and In Vitro Evaluation of Novel Indolyl DiHydropyrazole Derivatives as Potential Anticancer Agents
Previous Article in Journal
Nucleus-Independent Chemical Shift (NICS) as a Criterion for the Design of New Antifungal Benzofuranones
Previous Article in Special Issue
Synthesis and Antiplasmodial Activity of Bisindolylcyclobutenediones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress Concerning the N-Arylation of Indoles

Department of Organic Chemistry, University of Chemistry and Technology, 16628 Prague, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2021, 26(16), 5079; https://doi.org/10.3390/molecules26165079
Submission received: 31 July 2021 / Revised: 17 August 2021 / Accepted: 18 August 2021 / Published: 22 August 2021
(This article belongs to the Special Issue Indole Derivatives: Synthesis and Application II)

Abstract

:
This review summarizes the current state-of-the-art procedures in terms of the preparation of N-arylindoles. After a short introduction, the transition-metal-free procedures available for the N-arylation of indoles are briefly discussed. Then, the nickel-catalyzed and palladium-catalyzed N-arylation of indoles are both discussed. In the next section, copper-catalyzed procedures for the N-arylation of indoles are described. The final section focuses on recent findings in the field of biologically active N-arylindoles.

1. Introduction

Heterocyclic compounds are cyclic compounds that each contain either a heteroatom or a number of heteroatoms. Among the most important groups of heterocyclic compounds are the aromatic heterocyclic compounds. Well-known examples of such compounds include pyridine, pyrrole, thiophene, and furan. Indoles occupy a unique position within the field of heterocyclic compounds. An indole is a condensed compound that is composed of a benzene and pyrrole core. The importance of indoles and their derivatives is illustrated by the examples shown in Figure 1. Tryptophan is an aromatic essential amino acid, and sattazolin [1] and umifenovir [2] are indole derivatives with antiviral effects. N-arylated indoles also belong to the group of important indole derivatives. An example of such a compound is sertindole, which is used as an antipsychotic drug [3]. Due to the significant biological activity of the indole derivatives, much attention has been paid to the modification of the indole unit [4,5,6,7,8].
The significant biological activity exhibited by sertindole has inspired the search for new procedures for the preparation of N-arylindoles. Every indole contains a pyrrole core, so it can be expected that the N-arylation of indoles will proceed in a similar manner to the N-arylation of pyrroles. Among the group of aromatic heterocyclic compounds with one heteroatom, pyrrole is unique in terms of its reactivity. Unlike pyrrole, modifications of heteroatoms in furan and thiophene to form stable derivatives of the starting compound are not feasible. In some cases, phospholes can be P-arylated to form pentasubstituted phospholes [9], although the reactivity of the phosphole is affected by its low aromaticity [10,11,12,13]. By contrast, pyrroles can be very easily arylated at position 1 [14,15] or 2 [16,17]. In addition, the dearomatizations of pyrroles are important [18,19,20,21,22]. Similar reactivity can be expected even in the case of indole arylations. The typical approach for the N-arylation of amines, including nitrogen-containing heteroaromatic compounds, involves the use of transition-metal-catalyzed reactions (Scheme 1). In these cases, Ullman condensation [23,24,25], Buchwald–Hartwig amination [26,27,28], and Chan–Lam coupling [29,30,31] are all frequently used. A common feature of these procedures is the reaction of amines with organohalides or boronic acids in the presence of a transition metal catalyst and a base. A relatively large number of N-arylation reactions of indoles have been reported over the last decade. However, to the best of our knowledge, only a few review articles have addressed the preparation of N-arylindoles [32,33,34]. Thus, we decided to summarize the procedures that have been used for the preparation of N-arylindoles over the past decade. The scope of the present review is limited to works that focused on the N-arylation of indoles or otherwise significantly contributed to this field.

2. Transition-Metal-Free N-Arylation of Indoles

The preparation of N-arylindoles without transition metal catalysts can be easily carried out using benzynes. This approach was first reported in 2018 (Scheme 2) [35]. In this study, the starting indole S2-1 was treated with 2-bromoacetophenone (S2-2) in the presence of tBuOK to produce N-phenylindole S2-3 in a 66% yield. The proposed mechanism accounts for the formation of a benzyne intermediate (S2-6) and ketene (S2-5) from the enolate S2-4. The scope of the reaction is limited to the indole S2-3. However, Chen and Wu published another benzyne-involving route to functionalized indoles, which in contrast, furnishes mainly C3-arylated indole derivatives, along with N-arylated indoles as side-products [36].
The second and seemingly most frequently used approach is based on nucleophilic aromatic substitution. The intramolecular N-arylation of indoles S3-1 was used for the preparation of dibenzo[b,f][1,4]oxazepines S3-3 (Scheme 3) [37]. Based on the general reaction, it appears that the methodology relies on Smiles rearrangement. This assumption has been experimentally confirmed, and the proposed mechanism assumes the formation of diaryl ether S3-4, which is converted into the N-arylated indole S3-3 by means of the Smiles rearrangement. Selected examples (S3-3a–S3-3c) show that the described procedure works in relation to the preparation of pyridine derivatives of S3-3b and in the case of the regioisomer S3-3c of the starting materials S3-2.
A similar concept for the N-arylation of indoles, which is based on intramolecular nucleophilic aromatic substitution, has been described by both Annareddygari [38] and Chen Ma [39] (Scheme 4). The first procedure for the preparation of indolo[1,2-a]quinoxalines S4-3 makes use of the condensation reaction of 2-formylindole S4-1 with substituted anilines S4-2 in N,N-dimethylformamide (DMF) at 120 °C. Similarly, the condensation–intramolecular-based transition-metal-free arylation approach was used by Ngernmeesri for the synthesis of indolo[1,2-a]quinolines [40]. Ma’s procedure starts with amides S4-4, and the N-arylation of the indole is achieved via double intramolecular nucleophilic aromatic substitution. In both cases, the scope of the studied reaction was limited to the preparation of five indole derivatives, S4-6. Intramolecular nucleophilic aromatic substitution has also been used for the preparation of 9H-pyrrolo[1,2-a]indol-9-ones and 10H-indolo[1,2-a]indol-10-ones by means of the base-mediated N-arylation of pyrroles and indoles [41].
Intermolecular nucleophilic aromatic substitution can also be used to prepare N-arylindoles (Scheme 5). This concept for the preparation of N-arylindoles is usually based on the reaction of aryl fluorides with indoles under basic conditions. Diness et al. developed two procedures for the preparation of N-arylindoles. In the first procedure, the reaction of pentafluorobenzene with an indole in the presence of tBuONa afforded N-arylindole S5-2, which was then used for the preparation of pentasubstituted benzene (Scheme 5, Diness’s N-arylation of indoles) [42]. The analogous reaction with a halogenated aromate containing only one fluorine atom proceeded under harsh reaction conditions, as illustrated by the selected synthesis of indole S5-3. However, the overall reaction scope was limited to three examples of N-arylated indoles [43]. Similar reaction conditions were used for the preparation of bis N-arylindoles with structures represented by the indole S5-5 (Scheme 5, Chang’s N-arylation of indoles). The prepared N-arylindoles were polymerized, and their optical and electrochemical properties and thermal stability were determined [44]. The results summarized in Scheme 5 indicate that the use of nucleophilic aromatic substitution for the preparation of N-arylindoles proceeds best in the case of pentafluorobenzenes. This conclusion was confirmed by the work of Zhang [45], who prepared a series of tetrafluorophenylbenzenes through the reaction of indoles with pentafluorobenzenes in the presence of sodium hydroxide under mild reaction conditions (Scheme 5, Zhang’s N-arylation of indoles). Similar reaction conditions (KOH, DMSO, 100 °C, 24 h) were used by Hua to prepare a series of N-arylated indoles (23 compounds, 24–87% isolated yields) [46].
In addition to the above-mentioned procedures for the transition-metal-free arylation of indoles by means of intermolecular nucleophilic aromatic substitution, other procedures have also been used to achieve limited N-arylation of indoles as parts of general procedures for the preparation of N-arylated amines. Such works include the preparation of benzimidazole-pyrrolo[1,2-a]quinoxaline [47], arene-metal π-complexation mediated C–H arylation [48], base-promoted nucleophilic fluoroarene substitution of C–F bonds [49], and the preparation N-arylation of amines via triarylsulfonium triflates [50].
The organocatalytic preparation of N-aryl carbazoles and indoles was reported in 2020 (Scheme 6) [51]. The reaction between the azonaphthalene S6-2 and N-nucleophiles was catalyzed by chiral phosphoric acid derivatives. Although the majority of reported N-arylations have been performed on carbazole (26 examples), the reaction has also been used for the preparation of ten N-arylindoles. It was proposed that the formation of the reaction products S6-3 involves the asymmetric nucleophilic addition of indoles S6-1 to the reaction center, and that the final product of the reaction S6-3 is obtained via rearomatization. Unfortunately, the scope of the reaction was limited to azonaphthalene S6-2, and the scope of the indoles was also limited.
A different approach to the N-arylation of indoles, which involves aza-Michael addition and aromatization, was reported in 2018 (Scheme 7) [52]. The developed methodology is based on the reaction of azoles with cyclohexa-2,4-dienones in the presence of a catalytic amount of scandium triflate in acetonitrile. The procedure can be used for the regioselective preparation of N-aryl pyrazoles, 1,2,4-triazoles, benzimidazoles, and 1,2,3-benzotriazoles in high yields. In the case of substituted indoles, the low selectivity of the arylation reaction was observed. Thus, 3-methylindole S7-1a reacted with the starting dienone to form a mixture of C2- and N-arylated indoles S7-2a and S7-3a. By contrast, 3-ethylindole S7-1b exclusively gave N-arylindole S7-2b in a 75% yield. The observed reactivity was not explained.
The transition-metal-free synthesis of N-arylindoles S8-3 from 1,2-allenic ketones S8-2 and variously substituted indoles S8-1 via a Cs2CO3-promoted tandem benzannulation reaction has recently been demonstrated by Li et al. (Scheme 8) [53]. The substrate scope with respect to both indoles S8-1a and S8-1b and allenic ketones S8-2 is quite broad. Interestingly, the C2-substituted indoles were either unreactive or reacted poorly under the reaction conditions, likely due to steric hindrance. A mechanism that involves a series of Michael additions followed by cyclization and benzannulation may be operative, as proposed by the authors.

3. Transition-Metal-Catalyzed N-Arylation of Indoles

3.1. Nickel-Catalyzed N-Arylation of Indoles

Nickel(0) complexes are highly reactive toward oxidative addition, and therefore, nickel catalysts are used for the cross-coupling reactions of aryl chlorides. Rull [54] was the first to report the Ni(0)-catalyzed arylation of indole S9-1 using aryl chlorides S9-2 (Scheme 9). The N-arylations were carried out using the [(IPr)Ni(styrene)2] complex [IPr=1,3-Bis(2,6-diisopropylphenyl)imidazol-2-ylidene] in the presence of tBuOLi in 1,4-dioxane at 100 °C. As illustrated by selected examples S9-3aS9-3h, the yields of the N-arylindoles were high. No reaction was observed with aryl chlorides bearing trifluoromethyl and benzoyl groups or ortho-substituted chlorobenzenes. Yet, the preparation of indoles with heteroaryl substituents required 10 mol% of the catalyst loading to achieve high yields of the corresponding indoles S9-3g and S9-3h. The need to use 10 mol% of the catalyst was explained by the competitive coordination of heteroaryl halide during the N-arylation reaction.
Different catalytical systems for the N-arylation of indoles were studied by Stradiotto [55], who investigated the reactivity of DPPF-based ligands (Scheme 10) during the Ni(0)-catalyzed N-arylation of indoles (Scheme 10). Heteroaryl chlorides were coupled with alkylamines and indole in the presence of ten structurally diverse 1,1′-bis(bis(alkyl/aryl)phosphino)ferrocene ligands (LX) and Ni(COD)2. The electron-poor ligand LCF3 proved to be extremely potent. It was the only ligand that gave the N-arylated indole S10-3a bearing a naphthyl substituent in a 80% yield. By contrast, the N-arylation of an indole with 4-chlorobenzonitrile proceeded efficiently with several LX variants, including LiPr, LCy, LPh, LCF3, LOMe, and LMe. The amination of 4-chloroquinaldine with the ligand LPh represents the first room-temperature N-arylation of an indole. In addition, ligands with ortho-substituted phenyl (Lo-tol and L1-nap), a sterically demanding dialkylphosphino group (LtBu), and difuranylphosphino variants (Lfur) proved ineffective under the studied conditions. The comparable catalytic performance of LPh, LCF3, and LOMe in the studied nickel-catalyzed N-arylations suggests that the electronic perturbations arising from arylphosphine substitution do not significantly affect the behavior of nickel in this branch of chemistry.
A different procedure for N-arylation is based on the reaction of amines or heterocyclic compounds with phenols (Scheme 11) [56]. Phenols are reluctant to react during cross-coupling reactions, so the starting phenol is converted into a cyanuric acid ester S11-4 via the reaction with cyanuric chloride (TCT) in 1,4-dioxane in the presence of a base. The conditions for the formation of the cyanuric acid ester S11-4 were obtained through extensive optimization. For example, the N-arylation of indole S11-1 with activated phenol S11-4 was performed using the NiCl2(dppf) catalytic system to give the N-arylated indole S11-3 in a 78% yield. Morpholine, carbazole, indoline, and pyrazole are examples of the other tested amines.
The nickel-catalyzed decarbonylation of substituted N-acyl indoles S12-1 represents another approach for the preparation of N-arylated indoles (Scheme 12) [57]. Two conditions were optimized for the Ni(0)-catalyzed decarbonylation of the starting indoles. First, stoichiometric amounts of Ni(cod)2 and PCy3 preferred quantitative decarbonylation at 80 °C (Condition A). Second, catalytic decarbonylation was achieved through the use of the 1,2-ethanediylbis[dicyclohexyl]phosphine (dcype) ligand at 180 °C (Condition B). Condition B proved to be suitable for N-acylindoles bearing an electron withdrawing group, whereas Condition A was suitable for N-acylindoles bearing an acyl group with electron-neutral and electron-rich substituents. Selected examples S12-2a–S12-2f indicate that this method tolerates esters, nitriles, and the ketone functional group. In addition to benzoic acid derivatives, this method has proven suitable for the formation of N-alkenylindoles S12-2e and N-heteroarylindoles S12-2f.
Recently, the use of NiO nanoparticles for the N-arylation of indoles by means of the reaction of boronic acids with indoles in the presence of NiO nanoparticles as a catalyst has been reported [58]. Pure NiO nanoparticles were prepared using the quercetin-mediated (F2-1, Figure 2) hydrothermal method.
In addition to the above-mentioned nickel-based catalytic systems for the N-arylation of indoles, the use of nickel catalysts F3-1 [59], F3-2 [60], F3-3 [61], F3-4 [62], and Ni(COD)2 with a polymeric triazine-phosphite ligand [63] for the reaction of amines or heterocycles with electrophiles has also been reported (Figure 3). A unique procedure for the preparation of N-aryl amines and indoles makes use of the nickel-catalyzed decarbonylative amination of carboxylic acid esters [64]. However, such procedures have mainly been developed for the N-arylation of amines, whereas the N-arylation of indoles has been achieved for only a limited number of indole derivatives.

3.2. Palladium-Catalyzed N-Arylation of Indoles

Similarly to the palladium-catalyzed arylation of pyrroles, the formation of 1, 2, and 3-arylated indoles can also be observed in the palladium-catalyzed arylation of indoles (Scheme 13) [65,66,67]. The regioselectivity of palladium-catalyzed indole arylation can be influenced by the detailed optimization of the reaction conditions. This was demonstrated in 2019 by Zhang et al., who reported the palladium-catalyzed cascade decarboxylative arylation of indoles [68]. The authors optimized the course of the reaction of indolyl carboxylic acids S13-1 with diaryliodonium salts S13-2 and determined that N1/C2 or C2/C3 arylations of indoles can be performed with high regioselectivity. The reaction allows the introduction of variously substituted biphenyls. The proposed key intermediates for both C2/C3 and N1/C2 arylation, S13-5 and S13-6, show that the N-arylation of indoles is the first step in N1/C2 arylation. However, the C3 arylation of the starting material occurs first in the case of C2/C3 arylation of the starting indolyl carboxylic acid S13-1.
A different approach for the preparation of diarylated indoles was reported in 2020 (Scheme 14) [69]. Muller et al. used 5-bromoindole to prepare diarylated indoles via sequential Suzuki reaction and Buchwald–Hartwig amination. The Suzuki reaction was performed first to avoid any side reactions. The subsequent amination reaction uses the same catalytic system as the Suzuki reaction, although potassium carbonate is used instead of CsF. Selected examples S14-2a and S14-2b illustrate how the Suzuki reaction of bromoindole S14-1 with 1-naphthylboronic acid gives worse yields of the indole S14-2b than the Buchwald–Hartwig amination with naphthyl bromide. This approach is also suitable for the preparation of diarylated carbazoles and 10H-phenothiazines.
An alternative approach to N-arylated indoles utilizes reactions the substituted indole S15-1 with substrates with activated C–O and C–N bonds. Selected examples are presented in Scheme 15. In some cases, the palladium-catalyzed reaction of quaternary ammonium salts with indoles in the presence of potassium tert-butoxide can be used to prepare N-arylated indoles, although the reported scope of the reaction is limited to the preparation of four simple N-arylindoles, S15-4. However, the synthesis of N-alkylated indoles by means of this approach is convenient [70]. Additionally, tosylates are sufficiently reactive for the preparation of N-arylindoles. In this way, a large group of quinazolines S15-2 were prepared in high isolated yields, although similar nonaflates do not react under the developed reaction conditions [71]. A similar approach for the synthesis of N-arylindoles S15-3 was reported in 2018 by Kwong, albeit with a significantly narrower scope [72].
A hydrogen-transfer-mediated cross-coupling reaction can also be used for the preparation of N-arylindoles (Scheme 16) [73]. The starting material, as represented by the structure of substituted indolines S16-1, reacts with 2-naphthols and their derivatives in the presence of Pd/C and sodium methoxide. Optimized conditions were applied to prepare 27 indole derivatives S16-2 with a naphthyl group at position 1. The reaction can tolerate, for example, an ester and a hydroxy functional group. However, 1-naphthols do not react under the optimized reaction conditions, as illustrated in Scheme 16. The proposed mechanism assumes the formation of active catalyst S16-5 by means of the reaction of Pd(0) with indoline. The resultant palladium catalyst S16-5 subsequently hydrogenates 2-naphthol S16-6 to form the intermediate S16-7. The final reaction product is formed by the condensation of intermediate S16-7 with indoline S16-1a, followed by palladium-mediated dehydrogenation.
The two-step one-pot procedure for the preparation of N-arylindoles relies on the palladium-catalyzed cyclization of 2-bromophenethylamines S17-1, which is followed by the palladium-catalyzed N-arylation of the resulting indoles S17-2a (Scheme 17) [74]. The optimization of the reaction conditions revealed that the cyclization of amines S17-1 into indoles S17-3 must be carried out at 200 °C. Running the same cyclisation reaction at 140 °C exclusively produces indoline derivatives S17-2b and intermediates. In the next step, the arylation of the resulting indole S17-2a is performed using the same catalytic system at 180 °C.
A similar approach for the preparation of N-arylated indoles was reported by Stradiotto (Scheme 18) [75]. In this case, the starting alkynes S18-1 reacts with ammonia to form indole-based intermediates S18-3. In the next step, the N-arylation of the resulting indole is performed using the same catalytic system in the presence of potassium tert-butoxide. This procedure led to the preparation of 13 indole derivatives. From the selected examples S18-2a–S18-2e, it is evident that the N-arylation of indoles also tolerates ortho-disubstituted aryl halides. The reaction conditions for the N-arylation of indoles developed in relation to the two-step one-pot procedure can also be used for the N-arylation of independently prepared indoles.
Recently, a series of papers have been published on the N-arylation of indoles using palladium catalysts immobilized on mesoporous silica SBA-15 [76,77], carbon nanotubes [78], or magnetic Fe3O4 nanoparticles [79]. Scheme 19 illustrates the immobilization of the palladium catalyst on the surface of SBA-15 through the coordination of palladium to the amidoxime group [76]. The major advantages offered by immobilized catalysts are their recyclability and reusability. This was demonstrated by reusing the immobilized catalyst in five catalytic cycles [76].
Palladium-catalyzed N-arylation of indoles has been used for the total synthesis of (–)-aspergilazine A (Scheme 20) [80]. In this report, the preparation of (–)-aspergilazine A (S20-3) was achieved by means of a cross-coupling reaction between non-brominated S20-1 and brominated indole S20-2 to form (–)-aspergilazine A (S20-3) at a 9% yield. When evaluating this approach to (–)-aspergilazine A, it should be noted that the starting substances contained four N–H bonds that could be arylated. In addition, 80% of the unreacted starting material S20-1 was isolated together with the reaction product S20-3.
Other examples of Buchwald–Hartwig indole N-arylation include the use of the ligands F4-1 [81], F4-2 [82], and tBuXphos (F4-3) [83] (Figure 4). In addition, the palladium-catalyzed amination of triflates [84,85]; cross-coupling reactions of silylated indoles with aryl halides [86,87]; and reactions that are catalyzed by Pd/ZnO nanoparticles [88], palladium immobilized on graphene [89], or silica-starch substrates [90] were used to prepare N-arylindoles. However, these procedures have mainly been developed for the N-arylation of amines, whereas the N-arylation of indoles was given for only a limited number of indole derivatives.

3.3. Copper-Catalyzed N-Arylation of Indoles

Ligand-free cross-coupling reactions can be considered economical and advantageous due to the cost reductions associated with the designed chemical transformations. In addition, indole N-arylations have been the subjects of ligandless experimental setups. This approach for the preparation of N-aryl azoles, including N-arylindoles, was reported by Teo in 2011 (Scheme 21) [91]. The reaction of unsubstituted indoles with aryl iodides in the presence of catalytic amounts of copper oxide and tetrabutylammonium bromide (TBAB) as phase-transfer catalysts in water gave N-arylindoles S21-2 in high isolated yields. Only aryl iodides can be used for the reaction, because the use of the corresponding bromides leads to significantly lower isolated yields of the target compounds, S20-2. The developed reaction conditions can also be used for the N-arylation of imidazole, pyrrole, 7-azaindole, and indazole. Alternatively, 5 mol% of copper(II) sulphate has been used to catalyze indole and azole N-arylation in the presence of sodium hydroxide at 110 °C in dimethyl sulfoxide [92].
Copper nanocatalysts have become a popular tool for the preparation of N-arylindoles. Recent examples of the use of such nanocatalysts for the synthesis of N-arylindoles are listed in Table 1. In most cases, nanocatalysts catalyze the N-arylation of indoles under ligandless conditions. An exception to this concerns the preparation of N-arylindoles in the presence of 1,10-phenanthroline as a ligand, which is catalyzed by electrospun copper oxide nanoparticles [93] and ceria-supported copper [94] (Table 1, entries 1 and 2). Other copper-based nanocatalysts include CuO nanoflakes (CuO NFs@MP) prepared by reducing copper(II) chloride with fruit waste [95] and CuO nanoparticles, which are easily available by reducing copper(II) acetate by means of water extraction from fresh Rosa canina fruits [96] (Table 1, entries 3 and 4). The structures of both nanocatalysts were determined using physicochemical methods, including X-ray diffraction, Fourier transform infrared spectroscopy, field emission scanning electron microscopy, and others. However, the testing of their ability to catalyze the N-arylation of indoles is limited to only a few cases. Substantially better scopes were reported for carbon nanotube-copper oxide (CNT–CuO) [97], copper oxide nanoparticles [98,99], and copper nanoparticles decorated with organically modified montmorillonite (copper-decorated OMMT) [100] (Table 1, entries 5–7).
An alternative route to N-arylindoles is based on the two-step N-arylation of indoles starting from either indolines, S22-1a, or indoline carboxylic acids, S22-1b (Scheme 22) [101]. The reaction of the starting compounds S22-1a and S22-1b with predominantly aryl iodides is catalyzed by a recyclable nano-copper oxide catalyst. Based on the available literature, the authors assume that the preparation of the target compounds S22-2 starts with the aromatization of indolines into indoles, followed by N-arylation. During this study, 27 N-arylated indoles were prepared. Selected examples S22-2aS22-2d show that the scope of the reaction is mainly limited to unsubstituted indoles and aryl iodides with simple substituents. The authors also confirmed that the nano-copper oxide catalyst can be reused in up to four cycles without a significant loss in activity.
Copper catalysts and nitrogen ligands represent a popular combination for cross-coupling reactions. This is also true for the N-arylation of indoles. A first example of the use of N-ligands concerns the preparation of N-arylindoles, which is catalyzed by copper iodide and ligand L1 (Table 2, entry 1) [102]. The authors succeeded in the preparation of 25 N-arylindoles, although a limited number of functional groups were introduced by this procedure. Significantly greater tolerance on the part of the functional groups was achieved by using the hydroxyquinalidine ligand L2 (Table 1, entry 2) [103]. In contrast to previous work [102], the N-arylation can be performed at 90 °C, and the reaction can also be used to prepare N-aryl imidazoles, pyrazoles, pyrroles, benzoimidazoles, and carbazoles. Ligands in the form of natural amino acids allow the N-arylation of indoles under similarly mild reaction conditions [103]. Unfortunately, the scope of the reaction is almost exclusively limited to the preparation of N-arylindoles (Table 2, entries 3 [104] and 4 [105]). The use of metformin as the ligand for N-arylation led to the formation of N-arylindoles, although the reactions were carried out in DMF at 130 °C (Table 2, entry 5) [106]. The scope of this N-arylation is limited to the preparation of N-arylindoles.
The copper-catalyzed N-arylation of indoles in the presence of an L-proline ligand was used to prepare indolo[1,2-a]quinoxalines (Scheme 23) [107]. The authors used the Ugi multicomponent reaction to prepare 2-substituted indoles S23-2, and the subsequent intramolecular cyclization gave the target compounds S23-3. The Ugi reaction is limited to unsubstituted 2-iodaniline and 2-indolyl carboxylic acid. From the examples of S23-3a and S23-3b, it is clear that the scope of the two-step one-pot protocol for the synthesis of indolo[1,2-a]quinoxalines S23-3 is limited to aldehydes and isonitriles without functional groups. It is worth noting that the copper-catalyzed N-arylation gave better yields of N-arylated indoles S23-3 than the palladium-catalyzed cyclization (Pd(OAc)2, BINAP, K2CO3, toluene, reflux, 24 h).
A similar approach for the intramolecular cyclization of Ugi adducts was reported by Liu in 2013 (Scheme 24) [108]. In this approach, the Ugi reaction is carried out among 2-indolylcarbaldehyde S24-1a, 2-iodobenzoic acid S24-1c, amine S24-1b, and isonitriles S24-1d. The Ugi adduct S24-2 is subsequently cyclized under different conditions. The N-arylation products S24-4 are obtained via a copper-catalyzed reaction in the presence of an L-proline ligand. The products of 2,3-cyclization S24-5 are formed during the palladium-catalyzed reaction. Cyclization performed without a transition metal produced isoindolin-1-ones S24-3. The scope of the discovered transformations is limited to a few examples, although a more extensive scope was reported by the same author in 2014 [109].
N,N′-Dimethylethylenediamine (DMEDA) is a popular ligand for use in the copper-catalyzed N-arylation of indoles. Lim used stoichiometric amounts of copper oxide for the arylation of indoles in pyridine (Table 3, entry 1) [110]. Unfortunately, the reaction scope is limited to the preparation of indoles with an ethoxycarbonyl group at position 2 of the indole moiety. An interesting route for the preparation of functionalized N-arylindoles was reported in 2013 (Table 3, entry 2) [111]. In their study, Ham and Kim developed the chemoselective arylation of halogen aryltrifluoroborates to form functionalized organotrifluoroborates. Only two indole derivatives were prepared by this procedure (in 72% and 73% yields, respectively). The presence of a trifluoroborate group in the structure of N-arylindoles can be used in the Suzuki reaction to prepare N-arylindoles with more complex substituents at position 1 [111]. The complementary method for the N-arylation of indoles was performed under similar reaction conditions (Table 3, entries 3 and 4) [112,113].
Dichotomies during the arylations of 3-substituted indoles S25-1a and S25-1b were reported by Schnürch (Scheme 25) [114]. The starting indole, which featured a Boc-protecting group, reacted with arylboronic acids in the presence of a palladium catalyst and one equivalent of copper acetate to give 2-arylindoles S25-3aS25-3c in moderate yields. The formation of N-arylindoles was not observed in this case. The low yields of indoles S25-3a, S25-3b, and S25-3c were explained by the inferior steric requirements for substitution at position 2. Changing the catalytic system to ferric chloride with a DMEDA ligand led to the formation of N-arylated indoles S25-4aS25-4e in high yields. N-arylation can also be performed in the case of unprotected indole S25-1b, although in this case a catalytic amount of copper iodide with a DMEDA ligand and cesium fluoride as a base had to be used.
A series of N-arylation reactions using ligands or reagents obtained from natural sources included the N-arylation of indoles, which was carried out in glycerol (Table 4, entry 1) [115]. Glycerol is considered a cheap and environmentally-friendly alternative to the commonly used solvents DMF and DMSO. A common feature of N-arylation reactions performed in glycerol, DMF, and DMSO is the high reaction temperature. The authors hypothesize that one equivalent of DMSO acts as a ligand to help the oxidative addition of the copper catalyst to the C-halogen bond. The good solubility of the reaction products in organic ethers allowed the recycling of the catalytic system dissolved in glycerol by extraction of the products and the starting compound to the ether. Such recycled catalysts can be used in three catalytic runs. A different ligand concept for the N-arylation of indoles makes use of methyl-α-d-glucopyranoside L1 and glucosamine L2 as ligands for copper-catalyzed N-arylations, as reported by Xuan [116] and Chen [117] (Table 4, entries 2 and 3). Neither reaction is significantly different from the other procedures available for the N-arylation of indoles. In the case of the glucosamine ligand L2, the formation of a radical intermediate T4-1 was proposed, although the formation of this intermediate has not been experimentally verified. The use of other natural substances in this regard includes the use of N-alkyl-glucosamine as a sugar-based surfactant [118], alpha-d-galacturonic acid L3 [119], and l-(-)-quebrachitol L4 [120] as ligands for the copper-catalyzed N-arylation of indoles and nitrogen-containing substances.
In addition to the above-mentioned examples, other ligands containing a nitrogen atom represent popular choices for the copper-catalyzed arylation of indoles. Nitrogen ligands also form part of the system for the N-arylation of indoles in water (Table 5). Schmitt used copper bromide and the ligand L1 for the N-arylation of indoles under mild reaction conditions (Table 5, entry 1) [121]. d-Glucose was used as the reducing agent in this case. The reaction was performed in a commercially available dl-α-tocopherol methoxypolyethylene glycol succinate solution. It was experimentally verified that the use of a surfactant is necessary to achieve a high yield of N-arylindoles. The developed reaction conditions were used for the synthesis of 11 N-arylindoles, and the tolerance of the functional groups in this case was limited to halogens, nitrile, ethers, and the formyl group. A similar surfactant-based strategy for preparing N-arylindoles was reported by Liu in 2013 (Table 5, entry 2) [122]. The reaction conditions for the N-arylation of indoles make use of the bipyridine ligand L2, potassium phosphate as the base, and betaine as the surfactant. In contrast to previous work [121], the arylation reaction must be performed at 90 °C. The authors demonstrated the tolerance of the acetyl group during the preparation of ten indole derivatives under optimized reaction conditions. An extensive series of N-arylated indoles was prepared by means of indole arylation catalyzed by copper iodide and the phenanthroline ligand L3 in 30% aqueous 1,2-dimethoxyethane (DME) (Table 5, entry 3) [123]. The optimum ratio of DME to water was determined using kinetic experiments, which showed that reactions performed in pure water or DME did not achieve quantitative conversion. An alternative procedure for the preparation of N-arylindoles relies on the reaction of boronic acids with indoles in aqueous ammonia and potassium tert-butoxide as a base under ligandless conditions [124]. To complete the list of N-arylations of indoles in water, it is necessary to mention the N-arylation procedures associated with heterocyclic compounds and amines, which can be applied for the preparation of a limited number of N-arylindoles [125,126,127,128,129,130,131].
The copper(I)-oxide-catalyzed N-arylation of indoles, as catalyzed using copper oxide and a pyridine N-oxide ligand L1, was reported in 2016 (Table 6, entry 1) [132]. By varying the ligands during the optimization of the reaction conditions, it was confirmed that the presence of both pyridine N-oxide and phenyl amide moieties in the structure of the ligand L1 is necessary to achieve maximum yields. The optimized reaction conditions are suitable for the preparation of N-aryl pyrazole, pyrrole, and indazole. The reaction scope for indoles is limited to six examples. The use of a pyridine N-oxide ligand was presumably inspired by prior studies that utilized pyridine N-oxide ligands for general N-arylation [133,134,135]. The pyridine-based ligand L2, which contains an iminopyridyl moiety, was used for the N-arylation of indoles in DMSO at 120 °C (Table 6, entry 2) [136]. Another ligand L3 that contains an iminopyridyl moiety in its molecule was reported in 2021 (Table 6, entry 3) [137]. A short communication on the copper-catalyzed N-arylation of indoles was published in 2015 (Table 6, entry 4). Copper(I) chloride with a phosphine ligand L4 in DMSO was found to be the best combination for the N-arylation of indoles. The scope of the reaction is limited to aryl halides with electron-donating or electron-neutral substituents. It should be noted that 4-iodobenzonitrile did not react under the optimized reaction conditions.
Recently, a self-relaying copper(I)-catalyzed sequence was used for the preparation of indolo[1,2-a]quinazolinones S26-3 by means of the reaction of the methyl esters of 3-indolylcarboxylic acid S26-1 with 2-bromobenzamides S26-2 (Scheme 26) [138]. This reaction is sensitive to the structure of the starting indole derivative, as the use of the 3-methylindole led to the formation of the N-arylated indole S26-4 and indoloquinazolinone S26-5. Fifteen indolo[1,2-a]quinazolinone S26-3 derivatives were prepared using this procedure. It is worth noting that despite the harsh reaction conditions, the presence of a chlorine atom was tolerated. It was also verified that the N-arylation of indoles to form the intermediate S26-6 is the first step in the formation of the target compounds S26-3.
The cobalt–copper-cocatalyzed N-arylation and alkenylation of N-heterocycles was reported by Ran (Scheme 27) [139]. The reaction conditions were optimized for the alkenylation of amides, although it was determined that these conditions could also be used for the preparation of N-arylindoles S27-2a and S27-2b in high yields. The authors assume that in the first step, the Co(II) is reduced to Co(I) by means of acetylacetone. The oxidative addition is followed by the substitution of the halide ligand, which is facilitated by the formation of the Cu(I)–N complex S27-6. The product of the reaction is then obtained by reductive elimination. The proposed mechanism was partially confirmed experimentally. A somewhat different approach to the N-arylation of nitrogen-containing heterocycles makes use of the iron-catalyzed bromination of aromatics, followed by copper-catalyzed N-arylation [140]. This approach was used for the synthesis of 1-(4-methoxyphenyl)indole in a 78% yield.
The general copper-catalyzed N-arylation of indoles and other azoles based on C–H bond activation was reported in 2016 (Scheme 28) [141]. The developed methodology is based on N-(quinolin-8-yl)benzamides S28-1, which react with azoles and indoles to give ortho-substituted benzamides S28-2. The methodology was used for the preparation of N-aryl pyrroles, indoles, carbazoles, and pyrazoles. The presence of a quinolin-8-yl group facilitates the formation of the complex S28-3 by means of copper coordination to the quinolin-8-yl moiety. The oxidation of CuII to CuIII forms complex S28-4, which undergoes intramolecular C–H cupration. Eight N-arylated indoles were prepared using this procedure. It was shown that benzamide with a pyrrolyl unit can be hydrolyzed into ortho-substituted benzoic acid. The same research group showed that the methodology developed for C–H functionalization reactions can be extended to picolinamides, rather than benzamides [142].
An alternative approach to the N-arylation of indoles makes use of 2-indolylboronic acids (Scheme 29) [143]. The starting carboxylic acids S29-1 react with aryl bromides or aryl iodides in the presence of catalytic amounts of copper oxide and potassium phosphate as a base. The reported scope of the reaction is limited to substituted phenyl iodides and bromides. Other aromatic and heteroaromatic halides were not tested. In terms of the proposed mechanism, the authors assume the formation of copper carboxylate S29-4 and the subsequent decarboxylation into the 2-indolyl copper reagent S29-5. The final products S29-2 are formed via anion exchange, oxidative addition, and reductive elimination steps.
Arguably the most extensive procedure for the preparation of N-arylindoles is based on the reaction of indoles with triarylbismuth reagents (Scheme 30) [144]. The reaction is catalyzed by copper acetate in the presence of oxygen and one equivalent of base. The triarylbismuth reagents can be prepared using the reaction of bismuth(III) trichloride with Grignard reagents or by the chemical transformation of the ester or formyl group of the corresponding triarylbismuth reagents. The procedure is suitable for the preparation of highly functionalized indoles because the methodology can tolerate a variety of functional groups, including a formyl or ester group, as illustrated by selected examples S30-2bS30-2e. The tryptophan derivative S30-2a was also prepared using this procedure. In addition, a series of heteroatom arylations by organobismuth reagents has been reported [145,146,147,148].
An alternative approach for the preparation of (–)-aspergilazine A that uses copper-catalyzed indole N-arylation was reported in 2017 (Scheme 31) [149]. A direct approach to (–)-aspergilazine A starting from the tryptophan derivative S31-4 proved inefficient because the expected product of the reaction S31-5 was obtained in only a 66% yield. Thus, the preparation of the target compound from indole S31-1 and 2-bromoindole or 2-iodoindole S31-2 was accomplished in nine steps in a total isolated yield of 36%.
Atom-economic arylation of indoles was described in 2015 (Scheme 32) [150]. N1-Unsubstituted indoles react with diaryliodonium salts in the presence of a catalytic amount of copper iodide to give N1- and C3-arylated indoles in average isolated yields. High selectivity was observed for aryl-uracil iodonium triflates S32-3, affording the 3-aryl-1-indoles-uracil conjugates S32-4.
The combination of photocatalysis and copper-catalyzed N-arylation enabled the N-arylation of indoles at room temperature (Scheme 33) [151]. The key step in this conversion is the photoexcitation of a copper complex S33-3. The reaction is also suitable for the preparation of N-aryl benzimidazoles, imidazoles, and carbazoles.

3.4. Miscellaneous Transition-Metal-Catalyzed N–Arylation of Indoles

A straightforward approach to the preparation of indolo[1,2-f]phenanthridine derivatives S34-3 based on the N–arylation of 2-aryl-3-substituted indoles S34-1 using an iridium-catalyzed oxidative [4 + 2] annulation reaction with quinones S34-2 was reported by Guo and Fan (Scheme 34) [152]. A series of 20 N-arylated indoles was prepared, all in satisfactory yields, under optimized reaction conditions. The indoles substituted at positions 2 and 3 smoothly reacted under optimized reaction conditions. However, the reaction is limited for benzoquinone and 2-methylbenzoquinone. Other quinone derivatives, such as 2-chlorobenzoquinone, 2,5-dimethylbenzoquinone, and naphthoquinone, did not furnish the corresponding products. The authors posited a plausible reaction mechanism on the basis of mechanistic studies. Initially, the iridium-catalyzed dual N–H/C–H bond activation of S34-5 is responsible for the formation of five-membered iridacycle S34-6. The subsequent coordination of benzoquinone to the iridium complex S34-6 affords the new iridacycle S34-7, which then undergoes a migratory insertion to deliver a seven-membered intermediate S34-8. Next, the protonolysis of the complex S34-8 with HOAc leads to both intermediate S34-9 and the active Ir(III) catalyst S34-4. Finally, a Zn(OAc)2-promoted intramolecular indolyl N1-attack on the carbonyl group affords the key intermediate S34-10, which undergoes a dehydration reaction to deliver the final product S34-11.
An alternative approach to the N-arylation of indole-2-carboxamides was developed by Kong (Scheme 35) [153]. In this reaction, zinc(II) iodide is used as a catalyst and Ag2CO3 as an oxidant en route to indolo[1,2-a]quinoxalin-6-on S35-2. The scope of the reaction is limited to substrates with a methoxy group or halogens. The presence of the N–CH3 amide moiety in the starting compound is crucial for the success of this reaction, as no reaction occurred in the case of a substrate bearing an N−H amide group or a −CO2(4-MePh) group. It was proposed that this behavior is caused by the steric effect of the N–CH3 group, which can bring the two reaction sites closer, thereby facilitating the intramolecular cyclization reaction. A plausible mechanism can be identified on the basis of controlled experiments. The reaction is initiated by the generation of the indolylzinc(II) intermediate A. The oxidation of A by Ag2CO3 delivers the resonance-stabilized radical intermediate B or C. The intermediate D is obtained via the intramolecular addition of an indole radical onto the N-aryl moiety, which is followed by oxidation and deprotonation to give the desired product S33-2a.

4. N-Arylated Indoles as Biologically Active Substances

In the preceding sections, we have discussed a wide variety of procedures that can be used for the preparation of N-indoles. Such a large number of different procedures have been developed due to the near endless number of applications of N-arylindoles. Among them, N-arylindoles have found applications in the field of materials chemistry. For example, N-arylindoles have been proposed as novel dye-sensitized solar cells [154], electrophosphorescent diodes [155], oxygen sensors [156], and other things [157,158,159,160]. Procedures for the N-arylation of indoles have also been used for the preparation of indoles with biological activity. A number of N-arylindoles have been prepared in structure–activity relationship (SAR) studies in an effort to verify the mechanisms of the biological activity of the substituted indoles; and the determination of the biological activities of N-arylindoles has also been part of a large-scope synthetic study concerning the preparation of N-arylindoles. Examples of such works include inhibitors of human neutrophil elastase (HNE) [161], centromere-associated protein-E (CENP-E) inhibitors [162], CRTh2 antagonists [163], and determinations of cytotoxic activity [164,165]. Ultimately, a large group of N-arylindoles was prepared to study their biological activity. Selected examples are given in Figure 5, Figure 6, Figure 7, Figure 8 and Figure 9.
A series of indole derivatives, F5-1a–F5-1c and F5-2a–F5-2d, was prepared to serve as angiotensin II (Ang II) receptor 1 antagonists (Figure 5) [166]. The introduction of the aryl substituent was accomplished through nucleophilic aromatic substitution by means of the reaction of substituted indoles with 2-fluorobenzonitrile in the presence of potassium carbonate in refluxing DMF. Among the prepared N-arylindoles, the compound F5-2b showed binding affinity to the angiotensin AT1 receptor, with an IC50 value of 1.26 ± 0.08 nM in radioligand binding assays. The same compound also showed the strongest antihypertensive effect, with a maximum reducing response of lower than 40.62 ± 4.08 mmHg MBP at 15 mg/kg 4 h after oral administration in SHRs. The acute toxicity of compound F5-2b was determined by LD50 as 1551.71 mg/kg. Other indole derivatives have been prepared by the same group and studied as angiotensin II receptor 1 antagonists [167,168,169,170]. The examples given in Figure 5 show that a common feature of these compounds is the presence of an ortho-substituted phenyl ring with a carboxylic acid F5-3 [167,168,170] or 1,2,4-oxadiazole F5-4 [169] scaffold at position 1 of the indole unit. Similarly, the introduction of phenyl substituents was accomplished via nucleophilic aromatic substitution in refluxing DMF.
The new quinoline derivatives were prepared by means of a Buchwald–Hartwig reaction, which was catalyzed using a XPhos Pd G2 catalyst and tBuONa as the base (Figure 6) [171]. A number of compounds, including the indole derivative F6-1, were prepared using this procedure. It was determined that the substance F6-1 inhibited 70% of the α-amylase at a concentration of 50 μg/mL, and it also inhibited α-glucosidase at a concentration of 50 μg/mL. Benzimidazole and indole conjugates were prepared via double copper-catalyzed C–N bond formation [172]. The selected compound F6-2 demonstrated the formation of a Nindole–Cphenyl bond by a copper(I)-iodide-catalyzed reaction in DMF at 150 °C. The substance F6-2 showed the best minimum inhibitory concentration (MIC) of 4 μg/mL against Gram-negative (E. coli, P. putida, and S. typhi) and Gram-positive (B. subtilis and S. aureus) bacteria. The antifungal activity against C. albicans and A. niger of F6-2 is represented by an MIC value of 16 μg/mL.
A number of indole derivatives with a 1-methylpyrazol-4-yl substituent at position 1 were investigated with respect to their inhibitory activity in relation to the autotaxin (ATX) enzyme, which is responsible for the hydrolysis of lysophosphatidylcholine (LPC) (Figure 7) [173]. The introduction of the pyrazole skeleton was performed by means of Ullman condensation with copper oxide and 1,10-phenanthroline ligand in pyridine at 110 °C under microwave conditions. A detailed SAR study showed that the indole derivative F7-1a is a potent compound with an IC50 value of 22 ± 6.9 nM in the Bis-pNPP assay and 4 ± 0.5 nM IC50 in the LPC assay. Similar results were observed for the benzoic derivative F7-1b. The crystal structure of the studied compounds indicated that the active compounds bind within the tunnel, which was described as an allosteric binding site. Both B-cell lymphoma (Bcl-2) and myeloid cell leukemia sequence 1 (Mcl-1) are proteins that help with the development and survival of multiple tumor types. Therefore, Fang [174] focused on studying the properties of indole derivatives with a phenyl substituent at position 1 as compounds that exhibit potent inhibitory activity against Bcl-2. The introduction of the aryl substituents was again catalyzed by cuprous oxide, although the reactions were performed in refluxing DMF. Compound F7-2 showed the best activity against Bcl-2, with a Ki value of 0.35 ± 0.03 μM. Docking studies revealed that the active compound F7-2 binds to the active pocket of the Bcl-2 protein through a hydrogen bond with Arg143 and the π–π interaction with Ph109 and Phe101. A number of 2,3-dihydrobenzo[f][1,4]oxazepine derivatives were prepared as antagonists of the prostaglandin E2 receptor subtype 2 (EP2) [175]. The oxazepine derivative with a 4-fluoroindole substituent F7-3 showed a binding affinity of 8 nM for the hEP2 receptor and an IC50 value of 50 nM in the hEP2 cAMP assay. In addition, the compound lacked CYP inhibition at 3 μM and had a 4000-fold higher binding affinity for the hEP2 receptor when compared with hEP1, hEP3, and hEP4. The formation of a N–Caryl bond was accomplished via a CuI-catalyzed reaction in the presence of a N,N′-dimethyl-1,2-cyclohexandiamine ligand and potassium phosphate as the base in 1,4-dioxane at 110 °C.
New substituted indole-3-carbinoles, as potent anti-cancer compounds derived from indole-3-carbinol (I3C), were designed and prepared as new inhibitors of NEDD4-1 ubiquitin ligase activity (Figure 8) [176]. In this case, the N-arylation of indoles was accomplished by means of an l-proline ligand and a copper catalyst. The parent carbinol I3C reached an IC50 value of 284 μM. The introduction of a benzyl group into position 1 decreased the IC50 value to 12.3 μM, whereas the introduction of the 4-tolyl group F8-1 substantially improved the IC50 value to 2.71 μM. Protein thermal shift assays in combination with in-silico binding simulation revealed that the tested derivatives bind to the purified catalytic HECT domain of NEDD4-1. In 2018, Knerr published two papers that evaluated N-arylindoles with trifluoromethoxy and carboxamide groups as selective inhibitors of secreted phospholipase A2 type X [177,178]. The formation of an N–Caryl bond was achieved by means of copper-catalyzed Ullman condensation in the presence of piperidine-2-carboxylic acid as a ligand. The indole F8-2 exhibited the best IC50 value (0.022 μM) against sPLA2-X. The presence of a trifluoromethoxy group is the key to achieving high inhibitory activity. The substitution of CF3O for the methoxy group led to the IC50 value being reduced to 0.62 μM. In another two papers [179,180], the selective activity of N-arylindoles toward P-glycoprotein (P-gp) or σ2 receptor agents was studied. The indole derivative F8-3 was chosen as a good ligand for both P-glycoprotein (P-gp) and the σ2 receptor. The starting compound F8-3 was subject to SAR studies wherein the N-arylation of indoles was performed by a copper-catalyzed reaction in the presence of zinc oxide. On the basis of these studies, the compound F8-4 as a (+)-(S)-enantiomer displayed notable nanomolar σ2 affinity [180].
Another set of indole derivatives with a similar substitution at position 1 of the indole skeleton was reported by Tamoo [181] and Uno (Figure 9) [182]. On the basis of extensive SAR studies, it was found that F9-1 is an inhibitor of cPLA (with an IC50 value of 0.075 μM) [181], and indole F9-2 is a selective inhibitor of both HSP90 (IC50 = 0.069 μM) and SK-Br-3 (IC50 = 0.33 μM) [182]. The introduction of the aryl substituents was accomplished via copper-catalyzed Ullman condensation at reaction temperatures above 100 °C.

5. Conclusions

This review article has summarized the current state-of-the-art procedures for the preparation of N-arylindoles. The syntheses of the target compounds can be accomplished in several ways. The simplest procedures are based on transition-metal-free reactions. The second approach makes use of transition-metal-catalyzed aminations of organohalogenides or organometallic compounds. The metal most commonly used for the N-arylation of indoles is arguably copper, although a number of processes use palladium and nickel catalysts. The N-arylation of indoles can be performed in an inter- or intramolecular manner in high isolated yields. Despite considerable efforts having been made to successfully accomplish N-arylation experiments, two major limitations persist. The first limitation is associated with the low tolerance of the functional groups. Only the N-arylation of indoles, which is accomplished by the effect of triorganobismuth reagents, avoids this trend. The second limitation concerns the reaction temperatures at which the N-arylations take place. In the vast majority of cases, the N-arylations are performed at high temperatures, which commonly exceed 100 °C. In some cases, the arylation reactions are also performed in refluxing DMF. However, N-arylations of indoles at room temperature are rare. Thus, the development of new catalytic systems for the N-arylation of indoles represents a new direction of development in this field. Such efforts will result in higher tolerance of the functional groups, which should streamline the development of new biologically active compounds based on substituted indoles. At the same time, efforts to lower the reaction temperature required for the N-arylation of indoles will greatly benefit the overall economic aspects of the process.

Author Contributions

Conceptualization—T.T.; writing original draft and editing—P.O., J.K., A.P., T.T. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported from the grant of Specific university research grant No A2_FCHT_2021_074 and the APC was funded by University of Chemistry and Technology Prague.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lampis, G.; Deidda, D.; Maullu, C.; Madeddu, M.A.; Pompei, R.; Delle Monachie, F.; Satta, G. Sattabacins and Sattazolins: New Biologically Active Compounds with Antiviral Properties Extracted from a Bacillus sp. J. Antibiot. 1995, 48, 967–972. [Google Scholar] [CrossRef] [Green Version]
  2. Haviernik, J.; Stefanik, M.; Fojtikova, M.; Kali, S.; Tordo, N.; Rudolf, I.; Hubalek, Z.; Eyer, L.; Ruzek, D. Arbidol (Umifenovir): A Broad-Spectrum Antiviral Drug That Inhibits Medically Important Arthropod-Borne Flaviviruses. Viruses 2018, 10, 184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Lewis, R.; Bagnall, A.M.; Leitner, M. Sertindole for schizophrenia. Cochrane Database Syst. Rev. 2005, 3. [Google Scholar] [CrossRef] [PubMed]
  4. Bhattacharjee, P.; Bora, U. Organocatalytic Dimensions to the C–H Functionalization of the Carbocyclic Core in Indoles: A Review Update. Org. Chem. Front. 2021, 8, 2343–2365. [Google Scholar] [CrossRef]
  5. Čubiňák, M.; Edlová, T.; Polák, P.; Tobrman, T. Indolylboronic Acids: Preparation and Applications. Molecules 2019, 24, 3523. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Trubitsõn, D.; Kanger, T. Enantioselective Catalytic Synthesis of N-Alkylated Indoles. Symmetry 2020, 12, 1184. [Google Scholar] [CrossRef]
  7. Urbina, K.; Tresp, D.; Sipps, K.; Szostak, M. Recent Advances in Metal-Catalyzed Functionalization of Indoles. Adv. Synth. Catal. 2021, 363, 2723–2739. [Google Scholar] [CrossRef]
  8. Wen, J.; Shi, Z. From C4 to C7: Innovative Strategies for Site-Selective Functionalization of Indole C-H Bonds. Acc. Chem Res. 2021, 54, 1723–1736. [Google Scholar] [CrossRef]
  9. Polák, P.; Čejka, J.; Tobrman, T. Formal Transition-Metal-Catalyzed Phosphole C-H Activation for the Synthesis of Pentasubstituted Phospholes. Org. Lett. 2020, 22, 2187–2190. [Google Scholar] [CrossRef]
  10. Keglevich, G. 1-(2,4,6-Trialkylphenyl)-1H-Phospholes with a Flattened P-Pyramid: Synthesis and Reactivity. In Phosphorus Heterocycles II; Topics in Heterocyclic Chemistry; Bansal, R., Ed.; Springer: Berlin/Heidelberg, Germany, 2010; Volume 21, pp. 149–173. [Google Scholar]
  11. Quin, L.D. The Continuing Development of the Chemistry of Phospholes. Curr. Org. Chem. 2006, 10, 43–78. [Google Scholar] [CrossRef]
  12. Romero-Nieto, C.; Baumgartner, T. Dithieno[3,2-b:2′,3′-d]phospholes: A Look Back at the First Decade. Synlett 2013, 24, 920–937. [Google Scholar] [CrossRef]
  13. Zagidullin, A.A.; Bezkishko, I.A.; Miluykov, V.A.; Sinyashin, O.G. Phospholes—Development and Recent Advances. Mendeleev Commun. 2013, 23, 117–130. [Google Scholar] [CrossRef]
  14. Ding, Z.; Nie, N.; Chen, T.; Meng, L.; Wang, G.; Chen, Z.; Hu, J. L-Proline N-Oxide Dihydrazides as an Efficient Ligand for Cross-Coupling Reactions of Aryl Iodides and Bromides with Amines and Phenols. Tetrahedron 2021, 79, 131826. [Google Scholar] [CrossRef]
  15. Antilla, J.C.; Baskin, J.M.; Barder, T.E.; Buchwald, S.L. Copper-Diamine-Catalyzed N-Arylation of Pyrroles, Pyrazoles, Indazoles, Imidazoles, and Triazoles. J. Org. Chem. 2004, 69, 5578–5587. [Google Scholar] [CrossRef] [PubMed]
  16. Chen, W.; Li, H.J.; Cheng, Y.F.; Wu, Y.C. Direct C2-Arylation of N-Acyl Pyrroles with Aryl Halides under Palladium Catalysis. Org. Biomol. Chem. 2021, 19, 1555–1564. [Google Scholar] [CrossRef] [PubMed]
  17. Kaloğlu, M.; Düşünceli, S.D.; Özdemir, İ. The First Used Butylene Linked bis(N-Heterocyclic Carbene)-Palladium-PEPPSI Complexes in the Direct Arylation of Furan and Pyrrole. J. Organomet. Chem. 2020, 915, 121236. [Google Scholar] [CrossRef]
  18. Chen, D.; Li, J.; Shan, Y.; Cui, P.; Zhao, Y.; Tian, L.; Qiu, G. Halogen-Radical-Promoted Dearomative Aza-Spirocyclization of Alkynylimines: An Efficient Approach to 3-Halo-Spirocyclohexadienones. Synthesis 2020, 52, 609–618. [Google Scholar] [CrossRef]
  19. Polák, P.; Tobrman, T. Dearomatization Strategy for the Synthesis of Arylated 2H-Pyrroles and 2,3,5-Trisubstituted 1H-Pyrroles. Org. Lett. 2017, 19, 4608–4611. [Google Scholar] [CrossRef]
  20. Zhuo, C.-X.; Cheng, Q.; Liu, W.-B.; Zhao, Q.; You, S.-L. Enantioselective Synthesis of Pyrrole-Based Spiro- and Polycyclic Derivatives by Iridium-Catalyzed Asymmetric Allylic Dearomatization and Controllable Migration Reactions. Angew. Chem. Int. Ed. 2015, 54, 8475–8479. [Google Scholar] [CrossRef]
  21. Zhuo, C.-X.; Liu, W.-B.; Wu, Q.-F.; You, S.-L. Asymmetric dearomatization of pyrroles via Ir-catalyzed allylic substitution reaction: Enantioselective synthesis of spiro-2H-pyrroles. Chem. Sci. 2012, 3, 205–208. [Google Scholar] [CrossRef]
  22. Zhuo, C.-X.; Zhou, Y.; You, S.-L. Highly Regio- and Enantioselective Synthesis of Polysubstituted 2H-Pyrroles via Pd-Catalyzed Intermolecular Asymmetric Allylic Dearomatization of Pyrroles. J. Am. Chem. Soc. 2014, 136, 6590–6593. [Google Scholar] [CrossRef]
  23. Kunz, K.; Scholz, U.; Ganzer, D. Renaissance of Ullmann and Goldberg Reactions—Progress in Copper Catalyzed C–N-, C–O- and C–S-Coupling. Synlett 2003, 2428–2439. [Google Scholar] [CrossRef] [Green Version]
  24. Sambiagio, C.; Marsden, S.P.; Blacker, A.J.; McGowan, P.C. Copper Catalysed Ullmann Type Chemistry: From Mechanistic Aspects to Modern Development. Chem. Soc. Rev. 2014, 43, 3525–3550. [Google Scholar] [CrossRef]
  25. Sperotto, E.; van Klink, G.P.M.; van Koten, G.; de Vries, J.G. The Mechanism of the Modified Ullmann Reaction. Dalton Trans. 2010, 39, 10338–10351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Dorel, R.; Grugel, C.P.; Haydl, A.M. The Buchwald–Hartwig Amination after 25 Years. Angew. Chem. Int. Ed. 2019, 58, 17118–17129. [Google Scholar] [CrossRef] [PubMed]
  27. Forero-Cortés, P.A.; Haydl, A.M. The 25th Anniversary of the Buchwald–Hartwig Amination: Development, Applications, and Outlook. Org. Proc. Res. Develop. 2019, 23, 1478–1483. [Google Scholar] [CrossRef]
  28. Heravi, M.M.; Kheilkordi, Z.; Zadsirjan, V.; Heydari, M.; Malmir, M. Buchwald-Hartwig reaction: An overview. J. Organomet. Chem. 2018, 861, 17–104. [Google Scholar] [CrossRef]
  29. Chen, J.-Q.; Li, J.-H.; Dong, Z.-B. A Review on the Latest Progress of Chan-Lam Coupling Reaction. Adv. Synth. Catal. 2020, 362, 3311–3331. [Google Scholar] [CrossRef]
  30. Munir, I.; Zahoor, A.F.; Rasool, N.; Naqvi, S.A.R.; Zia, K.M.; Ahmad, R. Synthetic Applications and Methodology Development of Chan–Lam Coupling: A Review. Mol. Divers. 2019, 23, 215–259. [Google Scholar] [CrossRef] [PubMed]
  31. West, M.J.; Fyfe, J.W.B.; Vantourout, J.C.; Watson, A.J.B. Mechanistic Development and Recent Applications of the Chan–Lam Amination. Chem. Rev. 2019, 119, 12491–12523. [Google Scholar] [CrossRef]
  32. Joucla, L.; Djakovitch, L. Transition Metal-Catalysed Direct and Site-Selective N1-, C2- or C3-Arylation of the Indole Nucleus: 20 Years of Improvements. Adv. Synth. Catal. 2009, 351, 673–714. [Google Scholar] [CrossRef]
  33. Xu, H. Advances on N-Arylation of Indoles by Cross-Coupling Reactions. Mini-Rev. Org. Chem. 2009, 6, 367–377. [Google Scholar] [CrossRef]
  34. Halder, P.; Roy, T.; Das, P. Recent developments in selective N-arylation of azoles. Chem. Commun. 2021, 57, 5235–5249. [Google Scholar] [CrossRef] [PubMed]
  35. Chang, D.; Gao, F.; Shi, L. Potassium tert-butoxide-mediated generation of arynes from o-bromoacetophenone derivatives. Tetrahedron 2018, 74, 2428–2434. [Google Scholar] [CrossRef]
  36. Chen, J.; Wu, J. Transition-Metal-Free C3 Arylation of Indoles with Aryl Halides. Angew. Chem. Int. Ed. 2017, 56, 3951–3955. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Hu, F.; Liu, H.; Jia, J.; Ma, C. Transition-metal-free synthesis of indole-fused dibenzo[b,f][1,4]oxazepines via Smiles rearrangement. Org. Biomol. Chem. 2016, 14, 11076–11079. [Google Scholar] [CrossRef]
  38. Annareddygari, S.; Kasireddy, V.R.; Reddy, J. Transition-Metal-Free N-Arylation: A General Approach to Aza-Fused Poly-heteroaromatics. J. Heterocycl. Chem. 2019, 56, 3267–3276. [Google Scholar] [CrossRef]
  39. Huang, A.; Liu, F.; Zhan, C.; Liu, Y.; Ma, C. One-Pot Synthesis of Pyrrolo[1,2-a]quinoxalines. Org. Biomol. Chem. 2011, 9, 7351–7357. [Google Scholar] [CrossRef]
  40. Thanetchaiyakup, A.; Rattanarat, H.; Chuanopparat, N.; Ngernmeesri, P. One-Pot Synthesis of Substituted Indolo[1,2-a]quinolines under Transition-Metal-Free Conditions. Tetrahedron Lett. 2018, 59, 1014–1018. [Google Scholar] [CrossRef]
  41. Xu, H.; Sun, L.; Song, C. Base-Mediated N-Arylation for the Synthesis of 9H-Pyrrolo[1,2-a]indol-9-ones and 10H-Indolo[1,2-a]indol-10-ones. Helv. Chim. Acta 2019, 102, e1800195. [Google Scholar] [CrossRef] [Green Version]
  42. Diness, F.; Begtrup, M. Sequential Direct SNAr Reactions of Pentafluorobenzenes with Azole or Indole Derivatives. Org. Lett. 2014, 16, 3130–3133. [Google Scholar] [CrossRef]
  43. Diness, F.; Fairlie, D.P. Catalyst-Free N-Arylation Using Unactivated Fluorobenzenes. Angew. Chem. Int. Ed. 2012, 51, 8012–8016. [Google Scholar] [CrossRef] [PubMed]
  44. Chang, G.; Yang, L.; Liu, S.; Luo, X.; Lin, R.; Zhang, L. Synthesis of indole-based functional polymers with well-defined structures via a catalyst-free C–N coupling reaction. RSC Adv. 2014, 4, 30630–30637. [Google Scholar] [CrossRef]
  45. Liu, C.; Wang, H.; Xing, X.; Xu, Y.; Ma, J.-A.; Zhang, B. Selective C4–F bond cleavage of pentafluorobenzene: Synthesis of N-tetrafluoroarylated heterocyclic compounds. Tetrahedron Lett. 2013, 54, 4649–4652. [Google Scholar] [CrossRef]
  46. Iqbal, M.A.; Mehmood, H.; Lv, J.; Hua, R. Base-Promoted SNAr Reactions of Fluoro- and Chloroarenes as a Route to N-Aryl Indoles and Carbazoles. Molecules 2019, 24, 1145. [Google Scholar] [CrossRef] [Green Version]
  47. Maiti, B.; Sun, C.-M. Novel Approach Towards the Synthesis of Skeletally Diverse Benzimidazole-pyrrolo[1,2-a]quinoxaline by SNAr/Pictet–Spengler Reaction under Focused Microwave Irradiation. New J. Chem. 2011, 35, 1385–1396. [Google Scholar] [CrossRef]
  48. Ricci, P.; Krämer, K.; Cambeiro, X.C.; Larrosa, I. Arene–Metal π-Complexation as a Traceless Reactivity Enhancer for C–H Arylation. J. Am. Chem. Soc. 2013, 135, 13258–13261. [Google Scholar] [CrossRef]
  49. Su, J.; Chen, Q.; Lu, L.; Ma, Y.; Auyoung, G.H.L.; Hua, R. Base-Promoted Nucleophilic Fluoroarenes Substitution of CF Bonds. Tetrahedron 2018, 74, 303–307. [Google Scholar] [CrossRef]
  50. Tian, Z.-Y.; Ming, X.-X.; Teng, H.-B.; Hu, Y.-T.; Zhang, C.-P. Transition-Metal-Free N-Arylation of Amines by Triarylsulfonium Triflates. Chem. Eur. J. 2018, 24, 13744–13748. [Google Scholar] [CrossRef]
  51. Xia, W.; An, Q.-J.; Xiang, S.-H.; Li, S.; Wang, Y.-B.; Tan, B. Chiral Phosphoric Acid Catalyzed Atroposelective C−H Amination of Arenes. Angew. Chem. Int. Ed. 2020, 59, 6775–6779. [Google Scholar] [CrossRef] [PubMed]
  52. Chittimalla, S.K.; Nakka, S.; Koodalingam, M.; Bandi, C. N-Arylation of Heterocycles by a Tandem Aza-Michael Addition Reaction and Aromatization Sequence. Synlett 2018, 29, 57–64. [Google Scholar] [CrossRef] [Green Version]
  53. Li, S.; Wu, X.-X.; Chen, S. Base-Promoted Direct Synthesis of Functionalized N-Arylindoles via the Cascade Reactions of Allenic Ketones with Indoles. Org. Biomol. Chem. 2019, 17, 789–793. [Google Scholar] [CrossRef] [PubMed]
  54. Rull, S.G.; Blandez, J.F.; Fructos, M.R.; Belderrain, T.R.; Nicasio, M.C. C–N Coupling of Indoles and Carbazoles with Aromatic Chlorides Catalyzed by a Single-Component NHC-Nickel(0) Precursor. Adv. Synth. Catal. 2015, 357, 907–911. [Google Scholar] [CrossRef]
  55. Clark, J.S.K.; Voth, C.N.; Ferguson, M.J.; Stradiotto, M. Evaluating 1,1′-Bis(phosphino)ferrocene Ancillary Ligand Variants in the Nickel-Catalyzed C–N Cross-Coupling of (Hetero)aryl Chlorides. Organometallics 2017, 36, 679–686. [Google Scholar] [CrossRef]
  56. Iranpoor, N.; Panahi, F. Direct Nickel-Catalyzed Amination of Phenols via C–O Bond Activation using 2,4,6-Trichloro-1,3,5-triazine (TCT) as Reagent. Adv. Synth. Catal. 2014, 356, 3067–3073. [Google Scholar] [CrossRef]
  57. Morioka, T.; Nakatani, S.; Sakamoto, Y.; Kodama, T.; Ogoshi, S.; Chatani, N.; Tobisu, M. Nickel-Catalyzed Decarbonylation of N-Acylated N-Heteroarenes. Chem. Sci. 2019, 10, 6666–6671. [Google Scholar] [CrossRef] [Green Version]
  58. Krishnaveni, T.; Lakshmi, K.; Kadirvelu, K.; Kaveri, M.V. Exploration of Catalytic Activity of Quercetin Mediated Hydrothermally Synthesized NiO Nanoparticles Towards C–N Coupling of Nitrogen Heterocycles. Catal. Lett. 2020, 150, 1628–1640. [Google Scholar] [CrossRef]
  59. Gatien, A.V.; Lavoie, C.M.; Bennett, R.N.; Ferguson, M.J.; McDonald, R.; Johnson, E.R.; Speed, A.W.H.; Stradiotto, M. Application of Diazaphospholidine/Diazaphospholene-Based Bisphosphines in Room-Temperature Nickel-Catalyzed C(sp2)–N Cross-Couplings of Primary Alkylamines with (Hetero)aryl Chlorides and Bromides. ACS Catal. 2018, 8, 5328–5339. [Google Scholar] [CrossRef]
  60. Park, N.H.; Teverovskiy, G.; Buchwald, S.L. Development of an Air-Stable Nickel Precatalyst for the Amination of Aryl Chlorides, Sulfamates, Mesylates, and Triflates. Org. Lett. 2014, 16, 220–223. [Google Scholar] [CrossRef]
  61. Liu, R.Y.; Dennis, J.M.; Buchwald, S.L. The Quest for the Ideal Base: Rational Design of a Nickel Precatalyst Enables Mild, Homogeneous C–N Cross-Coupling. J. Am. Chem. Soc. 2020, 142, 4500–4507. [Google Scholar] [CrossRef]
  62. Sawatzky, R.S.; Ferguson, M.J.; Stradiotto, M. Thieme Chemistry Journals Awardees—Where Are They Now? Efficient Cross-Coupling of Secondary Amines/Azoles and Activated (Hetero)Aryl Chlorides Using an Air-Stable DPEPhos/Nickel Pre-Catalyst. Synlett 2017, 28, 1586–1591. [Google Scholar]
  63. Panahi, F.; Roozbin, F.; Rahimi, S.; Moayyed, M.; Valaei, A.; Iranpoor, N. A Triazine-Phosphite Polymeric Ligand Bearing Cage-Like P,N-Ligation Sites: An Efficient Ligand in the Nickel-Catalyzed Amination of Aryl Chlorides and Phenols. RSC Adv. 2016, 6, 80670–80678. [Google Scholar] [CrossRef]
  64. Malapit, C.A.; Borrell, M.; Milbauer, M.W.; Brigham, C.E.; Sanford, M.S. Nickel-Catalyzed Decarbonylative Amination of Carboxylic Acid Esters. J. Am. Chem. Soc. 2020, 142, 5918–5923. [Google Scholar] [CrossRef]
  65. Lokhande†, S.K.; Vaidya†, G.N.; Satpute, D.P.; Venkatesh, A.; Kumar, S.; Kumar, D. Structure Ligation Relationship of Amino Acids for the Selective Indole C−H Arylation Reaction: L-Aspartic acid as Sustainable Alternative of Phosphine Ligands. Adv. Synth. Catal. 2020, 362, 2857–2863. [Google Scholar] [CrossRef]
  66. Mohr, Y.; Renom-Carrasco, M.; Demarcy, C.; Quadrelli, E.A.; Camp, C.; Wisser, F.M.; Clot, E.; Thieuleux, C.; Canivet, J. Regiospecificity in Ligand-Free Pd-Catalyzed C–H Arylation of Indoles: LiHMDS as Base and Transient Directing Group. ACS Catal. 2020, 10, 2713–2719. [Google Scholar] [CrossRef]
  67. Yamaguchi, M.; Suzuki, K.; Sato, Y.; Manabe, K. Palladium-Catalyzed Direct C3-Selective Arylation of N-Unsubstituted Indoles with Aryl Chlorides and Triflates. Org. Lett. 2017, 19, 5388–5391. [Google Scholar] [CrossRef] [PubMed]
  68. Ye, Z.; Li, Y.; Xu, K.; Chen, N.; Zhang, F. Cascade π-Extended Decarboxylative Annulation Involving Cyclic Diaryliodonium Salts: Site-Selective Synthesis of Phenanthridines and Benzocarbazoles via a Traceless Directing Group Strategy. Org. Lett. 2019, 21, 9869–9873. [Google Scholar] [CrossRef] [PubMed]
  69. Mayer, L.; Kohlbecher, R.; Müller, T.J.J. Concatenating Suzuki Arylation and Buchwald–Hartwig Amination by a Sequentially Pd-Catalyzed One-Pot Process—Consecutive Three-Component Synthesis of C,N-Diarylated Heterocycles. Chem. Eur. J. 2020, 26, 15130–15134. [Google Scholar] [CrossRef]
  70. Chen, H.; Yang, H.; Li, N.; Xue, X.; He, Z.; Zeng, Q. Palladium-Catalyzed C–N Cross-Coupling of NH-Heteroarenes and Quaternary Ammonium Salts via C–N Bond Cleavage. Org.Process Res. Develop. 2019, 23, 1679–1685. [Google Scholar] [CrossRef]
  71. Ye, X.; Huang, J.; Deng, Z.; Yuan, J.; Peng, Y. Palladium-Catalyzed Cross-Coupling Reactions of 4-Tosyl-oxyquinazolines with Indoles: An Efficient Approach to 4-(1H-Indol-1-yl)quinazolines. Synthesis 2021, 53, 383–390. [Google Scholar]
  72. Choy, P.Y.; Chung, K.H.; Yang, Q.; So, C.M.; Sun, R.W.-Y.; Kwong, F.Y. A General Palladium–Phosphine Complex to Explore Aryl Tosylates in the N-Arylation of Amines: Scope and Limitations. Chem. Asian J. 2018, 13, 2465–2474. [Google Scholar] [CrossRef] [PubMed]
  73. Chen, X.; Yang, Z.; Chen, X.; Liang, W.; Zhu, Z.; Xie, F.; Li, Y. Hydrogen-Transfer-Mediated N-Arylation of Naphthols Using Indolines as Hydrogen Donors. J. Org. Chem. 2020, 85, 508–514. [Google Scholar] [CrossRef] [PubMed]
  74. Monguchi, Y.; Marumoto, T.; Takamatsu, H.; Sawama, Y.; Sajiki, H. Palladium on Carbon-Catalyzed One-Pot N-Arylindole Synthesis: Intramolecular Aromatic Amination, Aromatization, and Intermolecular Aromatic Amination. Adv. Synth. Catal. 2014, 356, 1866–1872. [Google Scholar] [CrossRef]
  75. Crawford, S.M.; Lavery, C.B.; Stradiotto, M. BippyPhos: A Single Ligand With Unprecedented Scope in the Buchwald–Hartwig Amination of (Hetero)aryl Chlorides. Chem. Eur. J. 2013, 19, 16760–16771. [Google Scholar] [CrossRef] [PubMed]
  76. Ghorbani-Vaghei, R.; Hemmati, S.; Hamelian, M.; Veisi, H. An Efficient, Mild and Selective Ullmann-Type N-Arylation of Indoles Catalysed by Pd Immobilized on Amidoxime-Functionalized Mesoporous SBA-15 as Heterogeneous and Recyclable Nanocatalyst. Appl. Organomet. Chem. 2015, 29, 195–199. [Google Scholar] [CrossRef]
  77. Veisi, H.; Poor Heravi, M.R.; Hamelian, M. SBA-15-Functionalized Melamine–Pyridine Group-Supported Palladium(0) as an Efficient Heterogeneous and Recyclable Nanocatalyst for N-Arylation of Indoles through Ullmann-Type Coupling Reactions. Appl. Organomet. Chem. 2015, 29, 334–337. [Google Scholar] [CrossRef]
  78. Veisi, H.; Morakabati, N. Palladium Nanoparticles Supported on Modified Single-Walled Carbon Nanotubes: A Heterogeneous and Reusable Catalyst in the Ullmann-Type N-Arylation of Imidazoles and Indoles. New J. Chem. 2015, 39, 2901–2907. [Google Scholar] [CrossRef]
  79. Ghorbani-Vaghei, R.; Hemmati, S.; Hekmati, M. Pd Immobilized on Modified Magnetic Fe3O4 Nanoparticles: Magnetically Recoverable and Reusable Pd Nanocatalyst for Suzuki-Miyaura Coupling Reactions and Ullmann-Type N-Arylation of Indoles. J. Chem. Sci. 2016, 128, 1157–1162. [Google Scholar] [CrossRef] [Green Version]
  80. Boyd, E.M.; Sperry, J. Total Synthesis of (−)-Aspergilazine A. Org. Lett. 2014, 16, 5056–5059. [Google Scholar] [CrossRef]
  81. Hajipour, A.R.; Dordahan, F.; Rafiee, F. Synthesis of Tertiary Aryl Amines of Various Aryl Halides and Secondary Amines using Ortho-Palladated Complex of Tribenzylamine. Appl. Organomet. Chem. 2013, 27, 704–706. [Google Scholar] [CrossRef]
  82. Monti, A.; Rama, R.J.; Gómez, B.; Maya, C.; Álvarez, E.; Carmona, E.; Nicasio, M.C. N-Substituted Aminobiphenyl Palladacycles Stabilized by Dialkylterphenyl Phosphanes: Preparation and Applications in CN Cross-Coupling Reactions. Inorg. Chim. Acta 2021, 518, 120214. [Google Scholar] [CrossRef]
  83. Wagner, P.; Bollenbach, M.; Doebelin, C.; Bihel, F.; Bourguignon, J.-J.; Salomé, C.; Schmitt, M. t-BuXPhos: A Highly Efficient Ligand for Buchwald–Hartwig Coupling in Water. Green Chem. 2014, 16, 4170–4178. [Google Scholar] [CrossRef]
  84. Izquierdo, J.; Jain, A.D.; Abdulkadir, S.A.; Schiltz, G.E. Palladium-Catalyzed Coupling Reactions on Functionalized 2-Trifluoromethyl-4-chromenone Scaffolds: Synthesis of Highly Functionalized Trifluoromethyl Heterocycles. Synthesis 2019, 51, 1342–1352. [Google Scholar] [CrossRef] [PubMed]
  85. Grimm, J.B.; Lavis, L.D. Synthesis of Rhodamines from Fluoresceins Using Pd-Catalyzed C–N Cross-Coupling. Org. Lett. 2011, 13, 6354–6357. [Google Scholar] [CrossRef]
  86. Shimizu, K.; Minami, Y.; Goto, O.; Ikehira, H.; Hiyama, T. Silicon-based C–N Cross-coupling Reaction. Chem. Lett. 2014, 43, 438–440. [Google Scholar] [CrossRef]
  87. Minami, Y.; Komiyama, T.; Shimizu, K.; Hiyama, T.; Goto, O.; Ikehira, H. Catalytic Carbon–Nitrogen Bond-Forming Cross-Coupling Using N-Trimethylsilylamines. Bull. Chem. Soc. Japan 2015, 88, 1437–1446. [Google Scholar] [CrossRef]
  88. Hosseini-Sarvari, M.; Razmi, Z. Highly Active Recyclable Heterogeneous Pd/ZnO Nanoparticle Catalyst: Sustainable Developments for the C–O and C–N Bond Cross-Coupling Reactions of Aryl Halides under Ligand-Free Conditions. RSC Adv. 2014, 4, 44105–44116. [Google Scholar] [CrossRef]
  89. Fareghi-Alamdari, R.; Haqiqi, M.G.; Zekri, N. Immobilized Pd(0) Nanoparticles on Phosphine-Functionalized Graphene as a Highly Active Catalyst for Heck, Suzuki and N-Arylation Reactions. New J. Chem. 2016, 40, 1287–1296. [Google Scholar] [CrossRef]
  90. Panahi, F.; Daneshgar, F.; Haghighi, F.; Khalafi-Nezhad, A. Immobilized Pd Nanoparticles on Silica-Starch Substrate (PNP-SSS): Efficient Heterogeneous Catalyst in Buchwald–Hartwig C–N Cross-Coupling Reaction. J. Organomet. Chem. 2017, 851, 210–217. [Google Scholar] [CrossRef]
  91. Yong, F.-F.; Teo, Y.-C.; Tay, S.-H.; Tan, B.Y.-H.; Lim, K.-H. A Ligand-Free Copper(I) Oxide Catalyzed Strategy for the N-Arylation of Azoles in Water. Tetrahedron Lett. 2011, 52, 1161–1164. [Google Scholar] [CrossRef]
  92. Liu, L.; Wu, F.; Liu, Y.; Xie, J.; Dai, B.; Zhou, Z. Copper-Catalysed N-Arylation of Pyrrole with Aryl Iodides Under Ligand-Free Conditions. J. Chem. Res. 2019, 38, 180–182. [Google Scholar] [CrossRef]
  93. Khalil, A.; Fihri, A.; Jouiad, M.; Hashaikeh, R. Electrospun Copper Oxide Nanoparticles as an Efficient Heterogeneous Catalyst for N-Arylation of Indole. Tetrahedron Lett. 2014, 55, 5973–5975. [Google Scholar] [CrossRef]
  94. Amadine, O.; Maati, H.; Abdelouhadi, K.; Fihri, A.; El Kazzouli, S.; Len, C.; El Bouari, A.; Solhy, A. Ceria-Supported Copper Nanoparticles: A Highly Efficient and Recyclable Catalyst for N-Arylation of Indole. J. Mol. Cat. A-Chem. 2014, 395, 409–419. [Google Scholar] [CrossRef]
  95. Chaudhary, K.; Subodh; Prakash, K.; Mogha, N.K.; Masram, D.T. Fruit Waste (Pulp) Decorated CuO NFs as Promising Platform for Enhanced Catalytic Response and Its Peroxidase Mimics Evaluation. Arab. J. Chem. 2020, 13, 4869–4881. [Google Scholar] [CrossRef]
  96. Hemmati, S.; Mehrazin, L.; Hekmati, M.; Izadi, M.; Veisi, H. Biosynthesis of CuO Nanoparticles Using Rosa Canina Fruit Extract as a Recyclable and Heterogeneous Nanocatalyst for C-N Ullmann Coupling Reactions. Mater. Chem. Phys. 2018, 214, 527–532. [Google Scholar] [CrossRef]
  97. Lim, J.; Kim, J.D.; Choi, H.C.; Lee, S. CNT-CuO Catalyzed C–N Bond Formation for N-Arylation of 2-Phenylindoles. J. Organomet. Chem. 2019, 902, 120970. [Google Scholar] [CrossRef]
  98. Pai, G.; Chattopadhyay, A.P. Ligand-Free Copper Nanoparticle Promoted N-Arylation of Azoles with Aryl and Heteroaryl Iodides. Tetrahedron Lett. 2014, 55, 941–944. [Google Scholar] [CrossRef]
  99. Suramwar, N.V.; Thakare, S.R.; Karade, N.N.; Khaty, N.T. Green Synthesis of Predominant (111) Facet CuO Nanoparticles: Heterogeneous and Recyclable Catalyst for N-Arylation of Indoles. J. Mol. Catal. A-Chem. 2012, 359, 28–34. [Google Scholar] [CrossRef]
  100. Talukdar, D.; Das, G.; Thakur, S.; Karak, N.; Thakur, A.J. Copper Nanoparticle Decorated Organically Modified Montmorillonite (OMMT): An Efficient Catalyst for the N-Arylation of Indoles and Similar Heterocycles. Catal. Commun. 2015, 59, 238–243. [Google Scholar] [CrossRef]
  101. Reddy, K.H.V.; Satish, G.; Ramesh, K.; Karnakar, K.; Nageswar, Y.V.D. An Efficient Synthesis of N-Substituted Indoles from Indoline/Indoline Carboxylic Acid via Aromatization Followed by C–N Cross-Coupling Reaction by Using Nano Copper Oxide as a Recyclable Catalyst. Tetrahedron Lett. 2012, 53, 3061–3065. [Google Scholar] [CrossRef]
  102. Patil, P.H.; Nallasivam, J.L.; Fernandes, R.A. Unimolecular 4-Hydroxypiperidines: New Ligands for Copper-Catalyzed N-Arylation. Asian J. Org. Chem. 2015, 4, 552–559. [Google Scholar] [CrossRef]
  103. Yang, X.; Xing, H.; Zhang, Y.; Lai, Y.; Zhang, Y.; Jiang, Y.; Ma, D. CuI/8-Hydroxyquinalidine Promoted N-Arylation of Indole and Azoles. Chin. J. Chem. 2012, 30, 875–880. [Google Scholar] [CrossRef]
  104. Elliott, E.-C.; Maggs, J.L.; Park, B.K.; O’Neill, P.M.; Stachulski, A.V. Convenient Syntheses of Halo-dibenz[b,f]azepines and Carbamazepine Analogues via N-Arylindoles. Org. Biomol. Chem. 2013, 11, 8426–8434. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Vaidya, G.N.; Khan, A.; Verma, H.; Kumar, S.; Kumar, D. Structure Ligation Relationship of Amino Acids for the Amination Cross-Coupling Reactions. J. Org. Chem. 2019, 84, 3004–3010. [Google Scholar] [CrossRef] [PubMed]
  106. Chen, H.; Lei, M.; Hu, L. Synthesis of 1-Aryl Indoles via Coupling Reaction of Indoles and Aryl Halides Catalyzed by CuI/metformin. Tetrahedron 2014, 70, 5626–5631. [Google Scholar] [CrossRef]
  107. Balalaie, S.; Bararjanian, M.; Hosseinzadeh, S.; Rominger, F.; Bijanzadeh, H.R.; Wolf, E. Designing a Sequential Ugi/Ullmann Type Reaction for the Synthesis of Indolo[1,2-a]quinoxalinones Catalyzed by CuI/l-Proline. Tetrahedron 2011, 67, 7294–7300. [Google Scholar] [CrossRef]
  108. Zhang, L.; Zhao, F.; Zheng, M.; Zhai, Y.; Liu, H. Rapid and Selective Access to Three Distinct Sets of Indole-Based Heterocycles from a Single Set of Ugi-Adducts under Microwave Heating. Chem. Commun. 2013, 49, 2894–2896. [Google Scholar] [CrossRef]
  109. Zhang, L.; Zheng, M.; Zhao, F.; Zhai, Y.; Liu, H. Rapid Generation of Privileged Substructure-Based Compound Libraries with Structural Diversity and Drug-Likeness. ACS Comb. Sci. 2014, 16, 184–191. [Google Scholar] [CrossRef] [PubMed]
  110. Lee, J.; Choi, J.H.; Shin, S.; Heo, J.-N.; Lim, H.J. N-Arylation of Sterically Hindered NH-Nucleophiles: Copper-Mediated Syntheses of Diverse N-Arylindole-2-carboxylates. Synthesis 2015, 47, 3301–3308. [Google Scholar] [CrossRef]
  111. Lee, J.-H.; Kim, H.; Kim, T.; Song, J.H.; Kim, W.-S.; Ham, J. Functionalization of Organotrifluoroborates via Cu-Catalyzed C–N Coupling Reaction. Bull. KoreanChem. Soc. 2013, 34, 42–48. [Google Scholar] [CrossRef] [Green Version]
  112. Liu, W.; Han, L.-Y.; Liu, R.-L.; Xu, L.-G.; Bi, Y.-L. Copper-Catalyzed N-Arylation of 2-Arylindoles with Aryl Halides. Chin. Chem. Lett. 2014, 25, 1240–1243. [Google Scholar] [CrossRef]
  113. Rodrigues, M.B.; Feitosa, S.C.; Wiethan, C.W.; Rosa, W.C.; da Silveira, C.H.; Pagliari, A.B.; Martins, M.A.P.; Zanatta, N.; Iglesias, B.A.; Bonacorso, H.G. Ullmann-Type Copper-Catalyzed Coupling Amination, Photophysical and DNA/HSA-Binding Properties of New 4-(Trifluoromethyl)quinoline Derivatives. J. Fluor. Chem. 2019, 221, 84–90. [Google Scholar] [CrossRef]
  114. Ghobrial, M.; Mihovilovic, M.D.; Schnurch, M. Exploration of C-H and N-H-Bond Functionalization Towards 1-(1,2-Diarylindol-3-yl)tetrahydroisoquinolines. Beilstein J. Org. Chem. 2014, 10, 2186–2199. [Google Scholar] [CrossRef] [Green Version]
  115. Yadav, D.K.T.; Rajak, S.S.; Bhanage, B.M. N-Arylation of Indoles with Aryl Halides Using Copper/Glycerol as a Mild and Highly Efficient Recyclable Catalytic System. Tetrahedron Lett. 2014, 55, 931–935. [Google Scholar] [CrossRef]
  116. Wei, J.J.; Song, W.B.; Zhu, Y.F.; Wei, B.L.; Xuan, L.J. N,N-Dimethyl-d-glucosamine as an Efficient Ligand for Copper-Catalyzed Ullmann-Type Coupling of N-H Heterocycles with Aryl Halides. Tetrahedron 2018, 74, 19–27. [Google Scholar] [CrossRef]
  117. Chen, Y.; Du, F.; Chen, F.; Zhou, Q.; Chen, G. Methyl-α-d-glucopyranoside as Green Ligand for Selective Copper-Catalyzed N-Arylation. Synthesis 2019, 51, 4590–4600. [Google Scholar] [CrossRef]
  118. Ge, X.; Zhang, S.; Chen, X.; Liu, X.; Qian, C. A Designed Bi-Functional Sugar-Based Surfactant: Micellar Catalysis for C–X Coupling Reaction in Water. Green Chem. 2019, 21, 2771–2776. [Google Scholar] [CrossRef]
  119. Yuan, C.; Zhao, Y.; Zheng, L. α-d-Galacturonic Acid as Natural Ligand for Selective Copper-Catalyzed N-Arylation of N-Containing Heterocycles. Synlett 2019, 30, 2173–2180. [Google Scholar] [CrossRef]
  120. Zhou, Q.; Du, F.; Chen, Y.; Fu, Y.; Sun, W.; Wu, Y.; Chen, G. l-(−)-Quebrachitol as a Ligand for Selective Copper(0)-Catalyzed N-Arylation of Nitrogen-Containing Heterocycles. J. Org. Chem. 2019, 84, 8160–8167. [Google Scholar] [CrossRef]
  121. Bollenbach, M.; Aquino, P.G.V.; de Araújo-Júnior, J.X.; Bourguignon, J.-J.; Bihel, F.; Salomé, C.; Wagner, P.; Schmitt, M. Efficient and Mild Ullmann-Type N-Arylation of Amides, Carbamates, and Azoles in Water. Chem. Eur. J. 2017, 23, 13676–13683. [Google Scholar] [CrossRef] [PubMed]
  122. Liu, S.; Zhou, J. Aqueous Copper-Catalyzed N-Arylation of Indoles: The Surfactant Strategy. New J. Chem. 2013, 37, 2537–2540. [Google Scholar] [CrossRef]
  123. Malavade, V.; Patil, M.; Patil, M. Scope, Kinetics, and Mechanism of “On Water” Cu Catalysis in the C–N Cross-Coupling Reactions of Indole Derivatives. Eur. J. Org. Chem. 2020, 561–569. [Google Scholar] [CrossRef]
  124. Molaei, H.; Ghanbari, M.M. Practical Copper-Catalyzed N-Arylation of Amines with 20% Aqueous Solution of n-Bu4NOH. Chin. Chem. Lett. 2012, 23, 301–304. [Google Scholar] [CrossRef]
  125. Abele, E.; Abele, R. KOH/Adogen 464/Proline System for Highly Effective Cu-Catalyzed “On-Water” N–H Arylation of Heteroaromatic Compounds. Chem. Heterocycl. Com. 2013, 49, 1384–1386. [Google Scholar] [CrossRef]
  126. Engel-Andreasen, J.; Shimpukade, B.; Ulven, T. Selective Copper Catalysed Aromatic N-Arylation in Water. Green Chem. 2013, 15, 336–340. [Google Scholar] [CrossRef] [Green Version]
  127. Mukhopadhyay, C.; Tapaswi, P.K. Highly Efficient and Simple Catalytic System for the N-Arylation of Some Hindered Aza-Heterocycles in Water. Synth. Commun. 2012, 42, 2217–2228. [Google Scholar] [CrossRef]
  128. Salomé, C.; Wagner, P.; Bollenbach, M.; Bihel, F.; Bourguignon, J.-J.; Schmitt, M. Buchwald–Hartwig Reactions in Water Using Surfactants. Tetrahedron 2014, 70, 3413–3421. [Google Scholar] [CrossRef]
  129. Teo, Y.-C.; Yong, F.-F.; Lim, G.S. A Manganese/Copper Bimetallic Catalyst for C–N Coupling Reactions under Mild Conditions in Water. Tetrahedron Lett. 2011, 52, 7171–7174. [Google Scholar] [CrossRef]
  130. Zhou, G.; Chen, W.; Zhang, S.; Liu, X.; Yang, Z.; Ge, X.; Fan, H.-J. A Newly Designed Carbohydrate-Derived Alkylamine Promotes Ullmann Type C–N Coupling Catalyzed by Copper in Water. Synlett 2019, 30, 193–198. [Google Scholar]
  131. Zhou, Q.; Du, F.; Chen, Y.; Fu, Y.; Chen, G. “On Water” Promoted N-Arylation Reactions Using Cu (0)/Myo-inositol Catalytic System. Tetrahedron Lett. 2019, 60, 1938–1941. [Google Scholar] [CrossRef]
  132. Damkaci, F.; Alawaed, A.; Vik, E. N-Picolinamides as Ligands for Ullmann-type CN Coupling Reactions. Tetrahedron Lett. 2016, 57, 2197–2200. [Google Scholar] [CrossRef]
  133. Su, J.; Qiu, Y.; Jiang, S.; Zhang, D. New Ligands for Copper-Catalyzed C–N Coupling Reactions at Gentle Temperature. Chin. J. Chem. 2014, 32, 685–688. [Google Scholar] [CrossRef]
  134. Yang, K.; Qiu, Y.; Li, Z.; Wang, Z.; Jiang, S. Ligands for Copper-Catalyzed C−N Bond Forming Reactions with 1 Mol% CuBr as Catalyst. J. Org. Chem. 2011, 76, 3151–3159. [Google Scholar] [CrossRef]
  135. Wang, Y.; Zhang, Y.; Yang, B.; Zhang, A.; Yao, Q. N-(1-Oxy-2-picolyl)oxalamic Acids as a New Type of O,O-Ligands for the Cu-Catalyzed N-Arylation of Azoles with Aryl Halides in Water or Organic solvent. Org. Biomol. Chem. 2015, 13, 4101–4114. [Google Scholar] [CrossRef]
  136. Taywade, A.; Chavan, S.; Ulhe, A.; Berad, B. Unique CuI-Pyridine Based Ligands Catalytic Systems for N-Arylation of Indoles and Other Heterocycles. Synth. Commun. 2018, 48, 1443–1453. [Google Scholar] [CrossRef]
  137. Echeverry-Gonzalez, C.A.; Ortiz Villamizar, M.C.; Kouznetsov, V.V. The Remarkable Selectivity of the 2-Arylquinoline-Based Acyl Hydrazones Toward Copper Salts: Exploration of Their Catalytic Applications in the Copper Catalysed N-Arylation of Indole Derivatives and C1-Alkynylation of Tetrahydroisoquinolines via the A3 Reaction. New J. Chem. 2021, 45, 243–250. [Google Scholar]
  138. Abe, T.; Takahashi, Y.; Matsubara, Y.; Yamada, K. An Ullmann N-Arylation/2-Amidation Cascade by Self-Relay Copper Catalysis: One-Pot Synthesis of Indolo[1,2-a]quinazolinones. Org. Chem. Front. 2017, 4, 2124–2127. [Google Scholar] [CrossRef]
  139. Ghosh, T.; Maity, P.; Ranu, B.C. Cobalt-Copper Catalyzed C(sp2)—N Cross Coupling of Amides or Nitrogenated Heterocycles with Styrenyl or Aryl Halides. ChemistrySelect 2018, 3, 4406–4412. [Google Scholar] [CrossRef]
  140. Mostafa, M.A.B.; Calder, E.D.D.; Racys, D.T.; Sutherland, A. Intermolecular Aryl C−H Amination through Sequential Iron and Copper Catalysis. Chem. Eur. J. 2017, 23, 1044–1047. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  141. Sadhu, P.; Punniyamurthy, T. Copper(ii)-Mediated Regioselective N-Arylation of Pyrroles, Indoles, Pyrazoles and Carbazole via Dehydrogenative Coupling. Chem. Commun. 2016, 52, 2803–2806. [Google Scholar] [CrossRef] [PubMed]
  142. Pradhan, S.; De, P.B.; Punniyamurthy, T. Copper(II)-Mediated Chelation-Assisted Regioselective N-Naphthylation of Indoles, Pyrazoles and Pyrrole through Dehydrogenative Cross-Coupling. J. Org. Chem. 2017, 82, 4883–4890. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, Y.; Hu, Z.-Y.; Li, X.-C.; Guo, X.-X. Copper-Catalyzed Decarboxylative N-Arylation of Indole-2-carboxylic Acids. Synthesis 2019, 51, 1803–1808. [Google Scholar] [CrossRef]
  144. Petiot, P.; Dansereau, J.; Gagnon, A. Copper-Catalyzed N-Arylation of Azoles and Diazoles Using Highly Functionalized Trivalent Organobismuth Reagents. RSC Adv. 2014, 4, 22255–22259. [Google Scholar] [CrossRef]
  145. Hébert, M.; Petiot, P.; Benoit, E.; Dansereau, J.; Ahmad, T.; Le Roch, A.; Ottenwaelder, X.; Gagnon, A. Synthesis of Highly Functionalized Triarylbismuthines by Functional Group Manipulation and Use in Palladium- and Copper-Catalyzed Arylation Reactions. J. Org. Chem. 2016, 81, 5401–5416. [Google Scholar] [CrossRef]
  146. Jadhav, B.D.; Pardeshi, S.K. A Facile and Practical Copper Diacetate Mediated, Ligand Free C–N Cross Coupling of Trivalent Organobismuth Compounds with Amines and N-heteroarenes. RSC Adv. 2016, 6, 14531–14537. [Google Scholar] [CrossRef]
  147. Le Roch, A.; Hébert, M.; Gagnon, A. Copper-Promoted O-Arylation of the Phenol Side Chain of Tyrosine Using Triarylbismuthines. Eur. J. Org. Chem. 2020, 5363–5367. [Google Scholar] [CrossRef]
  148. Le Roch, A.; Chan, H.-C.; Gagnon, A. Copper-Promoted N-Arylation of the Indole Side Chain of Tryptophan Using Triarylbismuthines. Eur. J. Org. Chem. 2020, 5815–5819. [Google Scholar] [CrossRef]
  149. Alonso, I.; Alvarez, R.; de Lera, Á.R. Indole–Indole Ullmann Cross-Coupling for CAr–N Bond Formation: Total Synthesis of (–)-Aspergilazine A. Eur. J. Org. Chem. 2017, 4948–4954. [Google Scholar] [CrossRef]
  150. Modha, S.G.; Greaney, M.F. Atom-Economical Transformation of Diaryliodonium Salts: Tandem C−H and N−H Arylation of Indoles. J. Am. Chem. Soc. 2015, 137, 1416–1419. [Google Scholar] [CrossRef]
  151. Ziegler, D.T.; Choi, J.; Muñoz-Molina, J.M.; Bissember, A.C.; Peters, J.C.; Fu, G.C. A Versatile Approach to Ullmann C−N Couplings at Room Temperature: New Families of Nucleophiles and Electrophiles for Photoinduced, Copper-Catalyzed Processes. J. Am. Chem. Soc. 2013, 135, 13107–13112. [Google Scholar] [CrossRef] [Green Version]
  152. Guo, S.; Liu, Y.; Zhang, X.; Fan, X. Iridium-Catalyzed Oxidative Annulation of 2-Arylindoles with Benzoquinone Leading to Indolo[1,2-f]phenanthridin-6-ols. Adv. Synth. Catal. 2020, 362, 3011–3020. [Google Scholar] [CrossRef]
  153. Kong, L.; Sun, Y.; Zheng, Z.; Tang, R.; Wang, M.; Li, Y. Chemoselective N–H or C-2 Arylation of Indole-2-carboxamides: Controllable Synthesis of Indolo[1,2-a]quinoxalin-6-ones and 2,3′-Spirobi[indolin]-2′-ones. Org. Lett. 2018, 20, 5251–5255. [Google Scholar] [CrossRef]
  154. Liu, X.; Cao, Z.; Huang, H.; Liu, X.; Tan, Y.; Chen, H.; Pei, Y.; Tan, S. Novel D–D–π-A Organic Dyes Based on Triphenylamine and Indole-Derivatives for High Performance Dye-Sensitized Solar Cells. J. Power Sources 2014, 248, 400–406. [Google Scholar] [CrossRef]
  155. Keruckas, J.; Grazulevicius, J.V.; Volyniuk, D.; Cherpak, V.; Stakhira, P. 3,6-Bis(indol-1-yl)-9-phenylcarbazoles as Electroactive Materials for Electrophosphorescent Diodes. Dye. Pigment. 2014, 100, 66–72. [Google Scholar] [CrossRef]
  156. Hussain, F.; Wang, X.; Wang, S. Impact of Bidentate N,C-Chelate Ligands on the Performance of Phosphorescent Pt(II) Complexes as Oxygen Sensors. J. Organomet. Chem. 2019, 880, 300–311. [Google Scholar] [CrossRef]
  157. Xiang, N.; Gao, Z.; Tian, G.; Chen, Y.; Liang, W.; Huang, J.; Dong, Q.; Wong, W.-Y.; Su, J. Novel Fluorene/Indole-Based Hole Transport Materials with High Thermal Stability for Efficient OLEDs. Dye. Pigment. 2017, 137, 36–42. [Google Scholar] [CrossRef]
  158. Jia, B.; Lian, H.; Chen, Z.; Chen, Y.; Huang, J.; Dong, Q. Novel Carbazole/Indole/Thiazole-Based Host Materials with High Thermal Stability for Efficient Phosphorescent Organic Light-Emitting Diodes. Dye. Pigment. 2017, 147, 552–559. [Google Scholar] [CrossRef]
  159. Selvam, R.; Subramanian, K. Benzimidazole-Indole-Chalcone Connected Methacrylate-Based Side Chain D-π-A Polymer and Its Application in Organic Photovoltaics. J. Polym. Sci. A Polym. Chem. 2017, 55, 997–1007. [Google Scholar] [CrossRef]
  160. Chen, Y.; Xie, J.; Wang, Z.; Cao, J.; Chen, H.; Huang, J.; Zhang, J.; Su, J. Highly Efficient Bipolar Host Material Based-on Indole and Triazine Moiety for Red Phosphorescent Light-Emitting Diodes. Dye. Pigment. 2016, 124, 188–195. [Google Scholar] [CrossRef]
  161. Crocetti, L.; Schepetkin, I.A.; Ciciani, G.; Giovannoni, M.P.; Guerrini, G.; Iacovone, A.; Khlebnikov, A.I.; Kirpotina, L.N.; Quinn, M.T.; Vergelli, C. Synthesis and Pharmacological Evaluation of Indole Derivatives as Deaza Analogues of Potent Human Neutrophil Elastase Inhibitors. Drug Develop. Res. 2016, 77, 285–299. [Google Scholar] [CrossRef] [Green Version]
  162. Hirayama, T.; Okaniwa, M.; Imada, T.; Ohashi, A.; Ohori, M.; Iwai, K.; Mori, K.; Kawamoto, T.; Yokota, A.; Tanaka, T.; et al. Synthetic Studies of Centromere-Associated Protein-E (CENP-E) Inhibitors: 1. Exploration of Fused Bicyclic Core Scaffolds Using Electrostatic Potential Map. Bioorg. Med. Chem. 2013, 21, 5488–5502. [Google Scholar] [CrossRef]
  163. Alonso, J.A.; Andrés, M.; Bravo, M.; Buil, M.A.; Calbet, M.; Castro, J.; Eastwood, P.R.; Eichhorn, P.; Esteve, C.; Gómez, E.; et al. Structure–Activity Relationships (SAR) and Structure–Kinetic Relationships (SKR) of Bicyclic Heteroaromatic Acetic Acids as Potent CRTh2 Antagonists I. Bioorg. Med. Chem. Lett. 2014, 24, 5118–5122. [Google Scholar] [CrossRef] [PubMed]
  164. Bzeih, T.; Lamaa, D.; Frison, G.; Hachem, A.; Jaber, N.; Bignon, J.; Retailleau, P.; Alami, M.; Hamze, A. Csp2–Csp2 and Csp2–N Bond Formation in a One-Pot Reaction between N-Tosylhydrazones and Bromonitrobenzenes: An Unexpected Cyclization to Substituted Indole Derivatives. Org. Lett. 2017, 19, 6700–6703. [Google Scholar] [CrossRef]
  165. Zhang, Q.; Zhong, Y.; Yan, L.-N.; Sun, X.; Gong, T.; Zhang, Z.-R. Synthesis and Preliminary Evaluation of Curcumin Analogues as Cytotoxic Agents. Bioorg. Med. Chem. Lett. 2011, 21, 1010–1014. [Google Scholar] [CrossRef]
  166. Bao, X.; Zhu, W.; Yuan, W.; Zhu, X.; Yan, Y.; Tang, H.; Chen, Z. Design, Synthesis and Evaluation of Novel Potent Angiotensin II Receptor 1 Antagonists. Eur. J. Med. Chem. 2016, 123, 115–127. [Google Scholar] [CrossRef] [PubMed]
  167. Wu, Z.; Anh, N.T.P.; Yan, Y.-J.; Xia, M.-B.; Wang, Y.-H.; Qiu, Y.; Chen, Z.-L. Design, Synthesis and Biological Evaluation of AT1 Receptor Blockers Derived from 6-Substituted Aminocarbonyl Benzimidazoles. Eur. J. Med. Chem. 2019, 181, 111553. [Google Scholar] [CrossRef] [PubMed]
  168. Zhu, W.; Bao, X.; Ren, H.; Da, Y.; Wu, D.; Li, F.; Yan, Y.; Wang, L.; Chen, Z. N-Phenyl Indole Derivatives as AT1 Antagonists with Anti-Hypertension Activities: Design, Synthesis and Biological Evaluation. Eur. J. Med. Chem. 2016, 115, 161–178. [Google Scholar] [CrossRef]
  169. Zhu, W.; Bao, X.; Ren, H.; Liao, P.; Zhu, W.; Yan, Y.; Wang, L.; Chen, Z. Design, Synthesis, and Pharmacological Evaluation of 5-oxo-1,2,4-oxadiazole Derivatives as AT1 Antagonists with Antihypertension Activities. Clin. Exp. Hypertens. 2016, 38, 435–442. [Google Scholar] [CrossRef]
  170. Zhu, W.; Da, Y.; Wu, D.; Zheng, H.; Zhu, L.; Wang, L.; Yan, Y.; Chen, Z. Design, Synthesis and Biological Evaluation of New 5-Nitro Benzimidazole Derivatives as AT1 Antagonists with Anti-Hypertension Activities. Bioorg. Med. Chem. 2014, 22, 2294–2302. [Google Scholar] [CrossRef]
  171. Thiyagamurthy, P.; Teja, C.; Naresh, K.; Gomathi, K.; Nawaz Khan, F.-R. Design, Synthesis and in Silico Evaluation of Benzoxazepino(7,6-b)quinolines as Potential Antidiabetic Agents. Med. Chem. Res. 2020, 29, 1882–1901. [Google Scholar] [CrossRef]
  172. Nandwana, N.K.; Singh, R.P.; Patel, O.P.S.; Dhiman, S.; Saini, H.K.; Jha, P.N.; Kumar, A. Design and Synthesis of Imidazo/Benzimidazo[1,2-c]quinazoline Derivatives and Evaluation of Their Antimicrobial Activity. ACS Omega 2018, 3, 16338–16346. [Google Scholar] [CrossRef] [Green Version]
  173. Miller, L.M.; Keune, W.-J.; Castagna, D.; Young, L.C.; Duffy, E.L.; Potjewyd, F.; Salgado-Polo, F.; Engel García, P.; Semaan, D.; Pritchard, J.M.; et al. Structure–Activity Relationships of Small Molecule Autotaxin Inhibitors with a Discrete Binding Mode. J. Med. Chem. 2017, 60, 722–748. [Google Scholar] [CrossRef] [Green Version]
  174. Xu, G.; Liu, T.; Zhou, Y.; Yang, X.; Fang, H. 1-Phenyl-1H-indole Derivatives as a New Class of Bcl-2/Mcl-1 Dual Inhibitors: Design, Synthesis, and Preliminary Biological Evaluation. Bioorg. Med. Chem. 2017, 25, 5548–5556. [Google Scholar] [CrossRef]
  175. Fox, B.M.; Beck, H.P.; Roveto, P.M.; Kayser, F.; Cheng, Q.; Dou, H.; Williamson, T.; Treanor, J.; Liu, H.; Jin, L.; et al. A Selective Prostaglandin E2 Receptor Subtype 2 (EP2) Antagonist Increases the Macrophage-Mediated Clearance of Amyloid-Beta Plaques. J. Med. Chem. 2015, 58, 5256–5273. [Google Scholar] [CrossRef] [PubMed]
  176. Quirit, J.G.; Lavrenov, S.N.; Poindexter, K.; Xu, J.; Kyauk, C.; Durkin, K.A.; Aronchik, I.; Tomasiak, T.; Solomatin, Y.A.; Preobrazhenskaya, M.N.; et al. Indole-3-carbinol (I3C) Analogues are Potent Small Molecule Inhibitors of NEDD4-1 Ubiquitin Ligase Activity that Disrupt Proliferation of Human Melanoma Cells. Biochem. Pharmacol. 2017, 127, 13–27. [Google Scholar] [CrossRef]
  177. Giordanetto, F.; Knerr, L.; Nordberg, P.; Pettersen, D.; Selmi, N.; Beisel, H.-G.; de la Motte, H.; Månsson, Å.; Dahlström, M.; Broddefalk, J.; et al. Design of Selective sPLA2-X Inhibitor (−)-2-{2-[Carbamoyl-6-(trifluoromethoxy)-1H-indol-1-yl]pyridine-2-yl}propanoic Acid. ACS Med. Chem. Lett. 2018, 9, 600–605. [Google Scholar] [CrossRef] [PubMed]
  178. Knerr, L.; Giordanetto, F.; Nordberg, P.; Pettersen, D.; Selmi, N.; Beisel, H.-G.; de la Motte, H.; Olsson, T.; Perkins, T.D.J.; Herslöf, M.; et al. Discovery of a Series of Indole-2-Carboxamides as Selective Secreted Phospholipase A2 Type X (sPLA2-X) Inhibitors. ACS Med. Chem. Lett. 2018, 9, 594–599. [Google Scholar] [CrossRef] [PubMed]
  179. Abate, C.; Pati, M.L.; Contino, M.; Colabufo, N.A.; Perrone, R.; Niso, M.; Berardi, F. From Mixed Sigma-2 Receptor/P-Glycoprotein Targeting Agents to Selective P-Glycoprotein Modulators: Small Structural Changes Address the Mechanism of Interaction at the Efflux Pump. Eur. J. Med. Chem. 2015, 89, 606–615. [Google Scholar] [CrossRef] [PubMed]
  180. Pati, M.L.; Abate, C.; Contino, M.; Ferorelli, S.; Luisi, R.; Carroccia, L.; Niso, M.; Berardi, F. Deconstruction of 6,7-Dimethoxy-1,2,3,4-tetrahydroisoquinoline Moiety to Separate P-Glycoprotein (P-gp) Activity from σ2 Receptor Affinity in Mixed P-gp/σ2 Receptor Agents. Eur. J. Med. Chem. 2015, 89, 691–700. [Google Scholar] [CrossRef] [PubMed]
  181. Tomoo, T.; Nakatsuka, T.; Katayama, T.; Hayashi, Y.; Fujieda, Y.; Terakawa, M.; Nagahira, K. Design, Synthesis, and Biological Evaluation of 3-(1-Aryl-1H-indol-5-yl)propanoic Acids as New Indole-Based Cytosolic Phospholipase A2α Inhibitors. J. Med. Chem 2014, 57, 7244–7262. [Google Scholar] [CrossRef]
  182. Uno, T.; Kawai, Y.; Yamashita, S.; Oshiumi, H.; Yoshimura, C.; Mizutani, T.; Suzuki, T.; Chong, K.T.; Shigeno, K.; Ohkubo, M.; et al. Discovery of 3-Ethyl-4-(3-isopropyl-4-(4-(1-methyl-1H-pyrazol-4-yl)-1H-imidazol-1-yl)-1H-pyrazolo[3,4-b]pyridin-1-yl)benzamide (TAS-116) as a Potent, Selective, and Orally Available HSP90 Inhibitor. J. Med. Chem. 2019, 62, 531–551. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Selected examples of important indole derivatives.
Figure 1. Selected examples of important indole derivatives.
Molecules 26 05079 g001
Scheme 1. Transition-metal-catalyzed approaches to N-arylamines.
Scheme 1. Transition-metal-catalyzed approaches to N-arylamines.
Molecules 26 05079 sch001
Scheme 2. N-arylation of an indole using a benzyne intermediate.
Scheme 2. N-arylation of an indole using a benzyne intermediate.
Molecules 26 05079 sch002
Scheme 3. Preparation of dibenzo[b,f][1,4]oxazepines S3-3 by means of Smiles rearrangement.
Scheme 3. Preparation of dibenzo[b,f][1,4]oxazepines S3-3 by means of Smiles rearrangement.
Molecules 26 05079 sch003
Scheme 4. Transition-metal-free syntheses of indolo[1,2-a]quinoxalines.
Scheme 4. Transition-metal-free syntheses of indolo[1,2-a]quinoxalines.
Molecules 26 05079 sch004
Scheme 5. Intermolecular nucleophilic aromatic substitution for the synthesis of N-arylindoles.
Scheme 5. Intermolecular nucleophilic aromatic substitution for the synthesis of N-arylindoles.
Molecules 26 05079 sch005
Scheme 6. Atroposelective N-arylation of indoles by arene C–H amination.
Scheme 6. Atroposelective N-arylation of indoles by arene C–H amination.
Molecules 26 05079 sch006
Scheme 7. Synthesis of N-arylindoles via an aza-Michael addition and aromatization sequence.
Scheme 7. Synthesis of N-arylindoles via an aza-Michael addition and aromatization sequence.
Molecules 26 05079 sch007
Scheme 8. Cascade benzannulation for the synthesis N-arylated indoles.
Scheme 8. Cascade benzannulation for the synthesis N-arylated indoles.
Molecules 26 05079 sch008
Scheme 9. Scope of the N-arylation of indoles with aryl chlorides catalyzed by [(IPr)Ni(styrene)2].
Scheme 9. Scope of the N-arylation of indoles with aryl chlorides catalyzed by [(IPr)Ni(styrene)2].
Molecules 26 05079 sch009
Scheme 10. Comparative catalytic screening during the nickel-catalyzed N-arylation of indoles.
Scheme 10. Comparative catalytic screening during the nickel-catalyzed N-arylation of indoles.
Molecules 26 05079 sch010
Scheme 11. Ni-catalyzed N-arylation of phenols via C–O bond activation by means of a TCT reagent.
Scheme 11. Ni-catalyzed N-arylation of phenols via C–O bond activation by means of a TCT reagent.
Molecules 26 05079 sch011
Scheme 12. Ni-catalyzed decarbonylation of N-acylated heteroarenes.
Scheme 12. Ni-catalyzed decarbonylation of N-acylated heteroarenes.
Molecules 26 05079 sch012
Figure 2. The structure of quercetin.
Figure 2. The structure of quercetin.
Molecules 26 05079 g002
Figure 3. Examples of nickel catalysts used for the N-arylation of amines.
Figure 3. Examples of nickel catalysts used for the N-arylation of amines.
Molecules 26 05079 g003
Scheme 13. Palladium-catalyzed site-selective N1/C2 and C3/C2 arylation of indoles.
Scheme 13. Palladium-catalyzed site-selective N1/C2 and C3/C2 arylation of indoles.
Molecules 26 05079 sch013
Scheme 14. Sequential Suzuki arylation and Buchwald–Hartwig amination for the synthesis of N-arylated indoles.
Scheme 14. Sequential Suzuki arylation and Buchwald–Hartwig amination for the synthesis of N-arylated indoles.
Molecules 26 05079 sch014
Scheme 15. Palladium-catalyzed N-arylation of indoles via the reactions with substrates with activated C–O and C–N bonds.
Scheme 15. Palladium-catalyzed N-arylation of indoles via the reactions with substrates with activated C–O and C–N bonds.
Molecules 26 05079 sch015
Scheme 16. Hydrogen-transfer-mediated N-arylation of indolines en route to N-arylated indoles.
Scheme 16. Hydrogen-transfer-mediated N-arylation of indolines en route to N-arylated indoles.
Molecules 26 05079 sch016
Scheme 17. Palladium-catalyzed one-pot synthesis of N-arylated indoles.
Scheme 17. Palladium-catalyzed one-pot synthesis of N-arylated indoles.
Molecules 26 05079 sch017
Scheme 18. Double N-arylation for the preparation of N-arylated indoles.
Scheme 18. Double N-arylation for the preparation of N-arylated indoles.
Molecules 26 05079 sch018
Scheme 19. Amidoxime-functionalized mesoporous SBA-15 as a heterogeneous and recyclable nanocatalyst for indole N-arylation.
Scheme 19. Amidoxime-functionalized mesoporous SBA-15 as a heterogeneous and recyclable nanocatalyst for indole N-arylation.
Molecules 26 05079 sch019
Scheme 20. Use of N-arylation of indoles for the preparation of (–) Aspergilazine A.
Scheme 20. Use of N-arylation of indoles for the preparation of (–) Aspergilazine A.
Molecules 26 05079 sch020
Figure 4. Selected examples of general palladium-catalyzed N-arylations used for the preparation of N-arylindoles.
Figure 4. Selected examples of general palladium-catalyzed N-arylations used for the preparation of N-arylindoles.
Molecules 26 05079 g004
Scheme 21. Ligand-free copper-catalyzed N-arylation of indole.
Scheme 21. Ligand-free copper-catalyzed N-arylation of indole.
Molecules 26 05079 sch021
Scheme 22. Synthesis of N-arylindoles from indolines or indoline carboxylic acids.
Scheme 22. Synthesis of N-arylindoles from indolines or indoline carboxylic acids.
Molecules 26 05079 sch022
Scheme 23. Ugi four-component reaction and subsequent post-modification for the synthesis of indolo[1,2-a]quinoxalines S23-3.
Scheme 23. Ugi four-component reaction and subsequent post-modification for the synthesis of indolo[1,2-a]quinoxalines S23-3.
Molecules 26 05079 sch023
Scheme 24. Diverse synthesis of heterocyclic compounds from the Ugi adducts S24-2.
Scheme 24. Diverse synthesis of heterocyclic compounds from the Ugi adducts S24-2.
Molecules 26 05079 sch024
Scheme 25. Dichotomies in the transition-metal-catalyzed arylation of substituted indoles.
Scheme 25. Dichotomies in the transition-metal-catalyzed arylation of substituted indoles.
Molecules 26 05079 sch025
Scheme 26. Formal double N-arylation synthesis of indolo[1,2-a]quinazolinones S26-3 via self-relay copper catalysis.
Scheme 26. Formal double N-arylation synthesis of indolo[1,2-a]quinazolinones S26-3 via self-relay copper catalysis.
Molecules 26 05079 sch026
Scheme 27. Cobalt–copper-cocatalyzed N-arylation of indoles.
Scheme 27. Cobalt–copper-cocatalyzed N-arylation of indoles.
Molecules 26 05079 sch027
Scheme 28. C–H activation for the N-arylation of indoles.
Scheme 28. C–H activation for the N-arylation of indoles.
Molecules 26 05079 sch028
Scheme 29. Decarboxylative N-arylation of 2-indolylcarboxylic acids.
Scheme 29. Decarboxylative N-arylation of 2-indolylcarboxylic acids.
Molecules 26 05079 sch029
Scheme 30. N-arylation of indoles by means of triarylbismuth reagents.
Scheme 30. N-arylation of indoles by means of triarylbismuth reagents.
Molecules 26 05079 sch030
Scheme 31. Total synthesis of (–)-aspergilazine A via the copper-catalyzed N-arylation of indoles.
Scheme 31. Total synthesis of (–)-aspergilazine A via the copper-catalyzed N-arylation of indoles.
Molecules 26 05079 sch031
Scheme 32. Atom-economical N-arylation of indoles by diaryliodonium salts.
Scheme 32. Atom-economical N-arylation of indoles by diaryliodonium salts.
Molecules 26 05079 sch032
Scheme 33. N-Arylation of indoles under mild conditions.
Scheme 33. N-Arylation of indoles under mild conditions.
Molecules 26 05079 sch033
Scheme 34. Iridium-catalyzed oxidative annulation of 2-arylindoles S34-1.
Scheme 34. Iridium-catalyzed oxidative annulation of 2-arylindoles S34-1.
Molecules 26 05079 sch034
Scheme 35. Novel method for the N–H arylation of indole-2-carboxamides S35-1 developed.
Scheme 35. Novel method for the N–H arylation of indole-2-carboxamides S35-1 developed.
Molecules 26 05079 sch035
Figure 5. Selected examples of bioactive N-arylindoles—part 1.
Figure 5. Selected examples of bioactive N-arylindoles—part 1.
Molecules 26 05079 g005
Figure 6. Selected examples of bioactive N-arylindoles—part 2.
Figure 6. Selected examples of bioactive N-arylindoles—part 2.
Molecules 26 05079 g006
Figure 7. Selected examples of bioactive N-arylindoles—part 3.
Figure 7. Selected examples of bioactive N-arylindoles—part 3.
Molecules 26 05079 g007
Figure 8. Selected examples of bioactive N-arylindoles—part 4.
Figure 8. Selected examples of bioactive N-arylindoles—part 4.
Molecules 26 05079 g008
Figure 9. Selected examples of bioactive N-arylindoles—part 5.
Figure 9. Selected examples of bioactive N-arylindoles—part 5.
Molecules 26 05079 g009
Table 1. Nanocatalyzed N-arylations of indoles.
Table 1. Nanocatalyzed N-arylations of indoles.
Molecules 26 05079 i001
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1CuO nano (5 mol%), 1,10-phen (50 mol%)
KOH (2 equiv), DMSO, reflux, 24 h
4
56–80%
Cl
2Cu@CeO2 (5 mol%), 1,10-phen (50 mol%)
KOH (2 equiv), DMSO, reflux, 24 h
6
82–89%
Cl
3PhI, CuO NFs@MP (15 mg/mmol)
K2CO3 (1.5 equiv), DMF, 100 °C, 6 h
Molecules 26 05079 i002
4CuO NPs (10 mg/1 mmol of indole), TEA (2 equiv),
DMF, reflux, 2 h
Molecules 26 05079 i003
5CNT–CuO (5 mol%), tBuONa (3 equiv), DMSO, 120 °C, 12 h21
42–95%
Cl, FCl, F, COMe, CO2Me, CN, CF3,
3,4-(CH2OCH2), NO2, SMe
6CuO nano (5 mol%), K2CO3 (1 equiv), DMF, reflux, 8 h15
62–98%
Br, CN, NO2, OMeNO2
7Copper-decorated OMMT (5 mol%), K2CO3 (2 equiv),
DMSO, 130 °C, 6 h
8
73–97%
CNOH, NH2,
Table 2. Nitrogen-based ligands for the N-arylation of indoles.
Table 2. Nitrogen-based ligands for the N-arylation of indoles.
Molecules 26 05079 i004
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1CuI (15 mol%)/L1 (15 mol%)
tBuOK (2 equiv), DMSO, 120 °C, 36–48 h
25
50–93%
CH=O, OMeNO2, OMe,
2CuI (2 mol%)/L2 (4 mol%)
K2CO3 (2.5 equiv), DMSO, 90 °C
19
31–98%
Cl, CF3, CN, COCH3, CO2H, F, NH2, NO2
3CuI (5 mol%)/L-proline (10 mol%), K2CO3 (5 equiv)
DMSO, 80–90 °C, 24 h
18
51–97%
Br, Cl, F, MeO,Br, Cl, F, OMe, OH
4CuI (5 mol%), L-methionine (10 mol%), K2CO3 (2 equiv), DMF, 100 °C, 24 h16
72–92%
Br, F, NO2, OBn, OMe,F, CF3, NO2,
5CuI (10 mol%), L3 (20 mol%), Cs2CO3 (2 equiv), DMF, 130 °C, 3–24 h29
62–91%
CHO, NO2, OMeBr, Cl, NO2, OMe
Molecules 26 05079 i005
Table 3. Copper-catalyzed N-arylation of indoles in the presence of a DMEDA ligand.
Table 3. Copper-catalyzed N-arylation of indoles in the presence of a DMEDA ligand.
Molecules 26 05079 i006
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1CuO (2 equiv), K2CO3 (2 equiv), pyridine, reflux, 2–3 d11
15–96%
Cl, F, CO2Et, OBnCF3, F, OMe,
2CuI (10 mol%), DMEDA (20 mol%), Cs2CO3 (1 equiv) DMSO, 120 °C, 4 h Molecules 26 05079 i007
3CuI (10 mol%), DMEDA (20 mol%), K3PO4 (2 equiv), toluene, 110 °C, 24 h13
74–91%
PhCl, CO2Et, F, OMe
4CuI (10 mol%), K3PO4 (3 equiv), DMEDA (20 mol%), toluene, 110 °C, 12 h Molecules 26 05079 i008
Table 4. Copper-catalyzed N-arylations of indoles.
Table 4. Copper-catalyzed N-arylations of indoles.
Molecules 26 05079 i009
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1CuI (10 mol%), DMSO (1 equiv), K2CO3 (2 equiv), glycerol, 120 °C, 24 h12
65–92%
Br, NO2, OMeBr, F, NO2, OMe
2Cu (10 mol%), L1 (20 mol%), Cs2CO3 (3 equiv) DMSO-H2O (1:1), 100 °C, 12–24 h Molecules 26 05079 i010
3Cu2O (10 mol%), L2 (20 mol%), Cs2CO3 (2 equiv), DMSO, 120 °C, 3–24 h15
86–97%
CH3, CHOCl, COMe, CO2Me, COCH2CO2Et OMe, NO2
Molecules 26 05079 i011
Table 5. Copper-catalyzed N-arylation of indoles in water.
Table 5. Copper-catalyzed N-arylation of indoles in water.
Molecules 26 05079 i012
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1CuBr2 (10 mol%), L1 (10 mol%), D-glucose (10 mol%), tBuONa (2 equiv), TPGS-750-M (5 wt%), 50 °C, 24 h11
43–95%
Br, Cl, F, OMOM, CHO, OMe, CH2CN, Cl, NO2 OMeNO2, OMe
2CuI (10 mol%), L2 (20 mol%), K3PO4 (2 equiv), betaine (2 equiv.), water, 90 °C, 10 h10
57–95%
Br, Cl, COMe, F, NO2, OMe
3CuI (5 mol%), L3 (10 mol%), KOH (2 equiv), DME-H2O (3:7), 95 °C, 20 h23
42–95%
Br, CN, COMe, NO2Br, Cl, I, OMe
Molecules 26 05079 i013
Table 6. Pyridine- and phosphorous-based ligands for the N-arylation of indoles.
Table 6. Pyridine- and phosphorous-based ligands for the N-arylation of indoles.
Molecules 26 05079 i014
EntryConditionsNo. of
Examples
indoleFGAr/hetArFG
1Cu2O (10 mol%), L1 (4 mol%), Cs2CO3 (2 equiv), DMSO, 120 °C, 20 h6
72–95%
NO2, OMe
2CuI (10 mol%), L2 (10 mol%), K2CO3 (2 equiv), DMSO, 120 °C, 4–16 h7
84–97%
NO2, OMe
3CuI (10 mol%), L3 (10 mol%), K2CO3 (1.4 equiv), DMSO, 110 °C, 24 h4
69–99
CHOOMe
4CuCl (8 mol%), L4 (16 mol%), NaOH (1 equiv), DMSO, 120 °C, 24 h8
25–98%
Br, Cl, OEt, OMe
Molecules 26 05079 i015
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Oeser, P.; Koudelka, J.; Petrenko, A.; Tobrman, T. Recent Progress Concerning the N-Arylation of Indoles. Molecules 2021, 26, 5079. https://doi.org/10.3390/molecules26165079

AMA Style

Oeser P, Koudelka J, Petrenko A, Tobrman T. Recent Progress Concerning the N-Arylation of Indoles. Molecules. 2021; 26(16):5079. https://doi.org/10.3390/molecules26165079

Chicago/Turabian Style

Oeser, Petr, Jakub Koudelka, Artem Petrenko, and Tomáš Tobrman. 2021. "Recent Progress Concerning the N-Arylation of Indoles" Molecules 26, no. 16: 5079. https://doi.org/10.3390/molecules26165079

Article Metrics

Back to TopTop