Next Article in Journal
Understanding Urea Encapsulation in Different Clay Minerals as a Possible System for Ruminant Nutrition
Next Article in Special Issue
Synthesis and Density Functional Theory Studies of Azirinyl and Oxiranyl Functionalized Isoindigo and (3Z,3’Z)-3,3’-(ethane-1,2-diylidene)bis(indolin-2-one) Derivatives
Previous Article in Journal
Location of Tensile Damage Source of Carbon Fiber Braided Composites Based on Two-Step Method
Previous Article in Special Issue
Spiroindolone Analogues as Potential Hypoglycemic with Dual Inhibitory Activity on α-Amylase and α-Glucosidase
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Indolylboronic Acids: Preparation and Applications

Department of Organic Chemistry, University of Chemistry and Technology Prague, Technická 5, 166 28 Prague 6, Czech Republic
*
Author to whom correspondence should be addressed.
Molecules 2019, 24(19), 3523; https://doi.org/10.3390/molecules24193523
Submission received: 7 September 2019 / Revised: 24 September 2019 / Accepted: 25 September 2019 / Published: 28 September 2019
(This article belongs to the Special Issue Indole Derivatives: Synthesis and Application)

Abstract

:
Indole derivatives are associated with a variety of both biological activities and applications in the field of material chemistry. A number of different strategies for synthesizing substituted indoles by means of the reactions of indolylboronic acids with electrophilic compounds are considered the methods of choice for modifying indoles because indolylboronic acids are easily available, stable, non-toxic and new reactions using indolylboronic acids have been described in the literature. Thus, the aim of this review is to summarize the methods available for the preparation of indolylboronic acids as well as their chemical transformations. The review covers the period 2010–2019.

1. Introduction

Indole is one of the simplest fused heterocyclic compounds known to act as a key scaffold of many naturally occurring molecules. For example, skatole (F1-1), melatonin (F1-2), serotonin (F1-3), and tryptamine (F1-4) are all well-known indole-based compounds (Figure 1). The functionalization of indoles, including the determination of their biological activities, has been the subject of considerable scientific research interest. Indeed, substituted indoles have been reported as affinity probes for the 5-hydroxytryptamine receptor [1], selective tumor-associated carbonic anhydrase inhibitors [2], carboxamide spleen tyrosine kinase inhibitors [3], selective inhibitors with broad hepatitis C virus genotype activity [4], Nav1.7 inhibitors [5], and TNFα inhibitors [6]. Further, the indole skeleton, as a constituent of naturally occurring substances, has been modified to furnish (−)-clavicipitic acid [7], breitfussin B [8], alstoscholarisines A and E [9], (+)-kopsihainanine A [10], and (−)-goniomitine [11].
The synthetic approaches for the preparation of substituted indoles can be divided into two main groups. The first group consists of methodologies involving both transition-metal-catalyzed and transition-metal-free cyclization strategies for ortho-substituted benzenes bearing the nitrogen-containing functional group, which are mostly substituted anilines (Scheme 1, route a). The cyclization pathway for the preparation of indoles is represented by the Baeyer-Emmerling [12], Bartoli [13], Fischer [14], Fukuyama [15], Gassman [16], and Larock [17] methods of indole synthesis. The second approach to indole synthesis is based on the modification of already functionalized indoles by means of traditional electrophilic [18,19] and nucleophilic [20] substitution reactions or transition-metal-catalyzed reactions of indolyl substrates (Scheme 1, route b) [21,22,23,24,25].
Transition metal-catalyzed cross-coupling reactions represent an indispensable tool for organic synthesis when considering the formation of C−C and C−heteroatom bonds. Among the available cross-coupling reactions, the Suzuki-Miyaura reaction [26,27] is the method of choice in terms of designing the construction of a C−C bond under mild reaction conditions. The Suzuki-Miyaura reaction involves a simple reaction between boronic acids S2-2 or boronates/trifluoroborates and electrophilic templates S2-1, as represented by halides, tosylates, triflates, or phosphates, in the presence of a base and transition metal catalyst. The good availability, low toxicity, and high stability of boronic acids all explain the widespread popularity of the Suzuki-Miyaura reaction. The reaction mechanism behind the Suzuki-Miyaura reaction involves a complex process [28] featuring oxidative addition, ligand substitution, transmetalation, and reductive elimination as the main reaction steps (Scheme 2).
Two reaction pathways can be envisioned when designing the synthesis of substituted indoles by means of the Suzuki-Miyaura reaction. The first approach is based on the cross-coupling of aryl- or heteroarylboronic acids with electrophilic indolyl templates. Selected examples show how chloro- [29], bromo- [30], bromoindolyl phosphate [31], and triflates [32] are all capable of coupling with boronic acids under mild reaction conditions. Yet, a wide variety of allylic alcohols [33], heterocyclic halides [34], bromoenol phosphates [35], sulfones [36], and sulfonates [37] have been used as electrophilic templates in reactions with indolylboronic acids under Suzuki-Miyaura reaction conditions involving, for example EvanPhos [38] and tri-tert-butylphosphine ligand [39], nickel [40], and rhodium [41] catalysts, as well as asymmetric transformations [42,43]. The abovementioned examples illustrate the practical application of the Suzuki-Miyaura reaction by means of both electrophilic and nucleophilic indolyl templates for the synthesis of substituted indoles. Aside from the Suzuki-Miyaura reaction, indolylboronic acids constitute an important type of indole intermediate, which has many applications in the synthesis of functionalized indoles. The lack of a comprehensive English written review summarizing the basic reactivities of indolylboronic acids prompted us to focus on this topic.

2. Synthetic Approaches to Indolylboronic Acids

2.1. Introduction

The traditional approach to indolylboronic acids is based on the halide-to-lithium exchange reaction by means of nBuLi. Selected examples reported by Liu [44] show the synthesis of 3-indolylboronic acid via the reaction of 3-bromoinodole S3-1 with nBuLi followed by the addition of triisopropyl borate. The boronic acid S3-2 that was formed in this way was used in the subsequent Suzuki-Miyaura reaction, affording the indole S3-3 with an overall isolated yield of 45% (Scheme 3).
The alternative synthesis of indolylboronic acids by means of the reaction of lithiated indoles with borates takes advantage of ortho-metalation (Scheme 4). When compared to the bromine-lithium exchange reaction, ortho-lithiation reduces the number of steps required to synthesize indolylboronic acids; however, the presence of a directing group is required. Vazquez [45] and Snieckus [46] developed a protocol for indole lithiation by means of nBuLi or LDA. The isolated yields of indolylboronic acid S4-2 and boronate S4-3 were within the yield range of 77−99%. The regioselectivity of the published lithiation in these cases is driven by the presence of a directing group as well as by the highest pKa value of the H2 hydrogen [47].
A similar approach was reported by Garg in relation to the 4-indolylboronic acid pinacol ester S5-2 (Scheme 5) [48]. The carbamate group was used as a traceless directing group, which allows for regioselective lithiation in position 4 of the indole scaffold. Once the Bpin fragment was introduced, the carbamate moiety was removed by means of a Ni-catalyzed reduction by 1,1,3,3-tetramethyl-disiloxane (TMDSO).

2.2. Miyaura Borylation en route to Indolylboronates

Miyaura borylation is a popular method for the synthesis of boronic acids and their derivatives. The typical borylation setup makes use of aryl/hetaryl halides, borylation reagents, and transition metal catalysts. Both iodide and chloride are able to couple with either pinB-Bdan [49] or diisopropylaminoborane [50] under palladium catalysis conditions (Table 1, entries 1 and 2). The isolated yields of indolylboronates ranged from 47−85%. The borylation was also performed in the presence of copper [51], cobalt catalysts [52], and copper-iron cooperative catalysis [53] (Table 1, entries 3−5). While Miyaura borylations of C−Cl, C−Br, and C−I bonds are quite common, the borylation of C−F bonds remains rare. This could be attributed to the relatively strong carbon−fluorine bond, which exhibits high dissociation energy and the shortest bond length in the organic halide series. Despite the unfavorable properties of C−F bonds in relation to transition-metal-catalyzed reactions, some borylations of fluoroindoles have been reported in the literature. Lee [54] and Ruben [55] utilized cobalt and nickel catalysts to carry out their indole borylations (Table 1, entries 6 and 7). Aside from traditional nucleophiles, aryl and hetaryl nitriles also react during borylative cross-coupling (Table 1, entry 8) [56,57]. The reaction is catalyzed by a rhodium catalyst, while bis(neopentyl glycolato)diboron (nepB-Bnep) is used as a borylation reagent. The rhodium precatalyst is activated by the reaction with nepB-Bnep. Then, the addition of activated catalyst S6-1 to nitrile along with E/Z isomerization gives complex S6-3. This complex undergoes β-aryl elimination, thereby affording arylated rhodium species S6-4, which indicates that the synthesis of boronates S6-5 is accomplished (Scheme 6). Alternatively, the borylation of haloindoles can be performed under transition-metal-free conditions. The reported procedures rely on either light-mediated borylation with pinB-Bpin under continuous-flow conditions (Table 1, entry 9) [58] or potassium-methoxide-mediated borylation with silylborane PhMe2Si-Bpin (Table 1, entry 10) [59,60] under mild reaction conditions. Transition-metal-free borylation has also been performed by mixing indolyldiazonium tetrafluoroborate with tetrahydroxydiboron; however, the yield of indolylboronic acid was only 35% [61].
A decarboxylative cross-coupling represents an alternative procedure for the synthesis of indolylboronic acids. The borylation reaction can be performed by starting with either carboxylic acids [62] or thioester [63] under palladium or rhodium catalysis (Table 2, entries 1 and 2). The decarboxylative indole borylation has also been performed under transition-metal-free conditions (Table 2, entries 3 and 4). The reported protocols make use of N-hydroxyphthalimide esters, which are treated with pinB−Bpin under different conditions. Fu et al. [64] rely on the isonicotinate tert-butyl ester when refluxing in trifluorotoluene (TFT), while Glorius et al. [65] make use of pyridine along with a 400-nm light-emitting diode (LED). A common feature of both decarboxylative borylations is the activation of pinB−Bpin by means of pyridine S7-1, thereby affording the ate complex S7-2 (Scheme 7). Next, the ate complex S7-2 undergoes a transformation into radical intermediates via electron transfer, and the formed radicals S7-3 [64] and S7-4 [65] are combined with the activated N-hydroxyphthalimide ester to accomplish the borylation sequence. The isolated yields of indolylboronic acid pinacol esters are moderate in all cases.

2.3. Transition-Metal-Catalyzed Borylation of Indoles by Means of C–H Activation

Since the discovery of the iridium-mediated C−H borylation reactions of aromatic substrates by Iverson and Smith [66], as well as the later improvement of the reaction conditions into iridium-catalyzed C−H borylation with B2pin2 or HBpin by Hartwig [67], the transition-metal-catalyzed borylation reactions have become one of the most versatile tools for the synthesis of borylated aromatic compounds. Heteroarenes, similar to indoles, can also undergo C−H borylation. The site selectivity of the reaction is an issue, as theoretically up to seven different sites can undergo C−H functionalization. Typically, in the case of unsubstituted indoles, the site selectivity is driven by electronic effects, as the reaction proceeds most readily at the C2 position where the C−H bond is the most acidic, and therefore, the most reactive. It is important to note, however, that selectivity can be overridden by steric factors. Several transition metals can be used for the borylation reactions of indoles, including iridium, platinum, nickel, and cobalt, although iridium-based catalytic systems with bipyridine ligands remain the most prevalent. In recent years, several site-selective and non-selective C−H borylation protocols of indoles have been developed, principally focusing on the recyclability of the expensive catalysts and site selectivity. Several boron sources can be used for the transition-metal-catalyzed C−H borylations of arenes (Scheme 8). The most common sources, namely HBpin (S8-1) and B2pin2 (S8-2), were established by Hartwig in relation to the transition-metal-catalyzed C2 borylation of indoles in the early 2000s [68].

2.3.1. C2-Selective Borylation

Recently Hartwig developed a protocol for the iridium-catalyzed borylation of arenes using B2hg2 (S8-4), which could also be successfully applied to the C2 borylation of indoles (S9-1) (Scheme 9) [69]. Tobisu and Chatani developed a one-pot protocol for the Ir/NHC carbene-catalyzed selective borylation of indoles using diisopropylamino borane (S8-7) as the boron source [70]. The formed B−N products S9-3 could not be isolated, although they could be turned into pinacol indol-2-ylboronates S9-4 with good isolated yields (Scheme 9).
Chattopadhyay recently developed L-shaped ligands for the selective borylation of the remote sites (meta and para) of aromatic esters [71] and amides [72], and the protocol could also be applied to indoles. Indoles that lacked the ester- or amide-directing group did not undergo the borylation reaction, most likely due to the preferential complexation of the indole NH group to the catalyst, thus inhibiting the borylation. Interestingly, following the introduction of the directing group onto the indole scaffold, the borylation proceeded without meta- or para-selectivity, but always into the C2 position (S10-1b,2b,3b), which suggests that electronic effects in the case of indoles S10-1a,2a,3a played a crucial role in the site selectivity (Scheme 10). Exchanging the L-shaped ligands for simple tbdpy ligands under these reaction conditions resulted in the formation of complex mixtures of unselectively borylated indoles.
As shown above, the ester and amide functional groups remain stable during the iridium-catalyzed C−H borylation, although the aldehyde group is not stable. Therefore, Ito and Ishiyama developed a protocol whereby heteroaromatic pentafluorophenyl aldimines undergo iridium-catalyzed C−H borylation in moderate to high yields, with one example being the indole substrate S11-1 [73]. The aldimine functional group serves as both a directing group and a protection group, and it can be easily hydrolyzed so as to produce aldehydes (Scheme 11).
In an effort to enhance the prospects of catalyst recovery and recycling, several groups have developed iridium-based heterogeneous catalysts for the C−H borylation of aromatic and heteroaromatic compounds. For instance, Jones et al. prepared and applied the mesoporous silica-supported bipyridine-based iridium catalyst bpy-SBA-15-Ir (S12-1) for the borylation of indole (Scheme 12) [74]. The catalyst did not reach a turnover number (TON) higher than 136, although it could be repeatedly reused following simple filtration with only the gradual loss of activity with subsequent runs. Similarly, Inagaki developed the heterogeneous catalyst Ir-bpy-PMO (S12-2), which in combination with the inexpensive boron source HBpin facilitated the C−H borylation of aromatic and heteroaromatic compounds, while the catalytic activity of Ir-bpy-PMO was comparable to that of homogenous catalytic systems (Scheme 12) [75]. Another example of a heterogeneous catalyst for C−H borylation reactions is based on the covalent triazine framework (CTF) prepared by Van Der Voort et al. They developed a bipyridine-based CTF and then metalated it with the Ir(I) complex so as to prepare an air-stable catalyst (S12-3) for the C−H borylation of various aromatic compounds, including indole [76]. The borylation reactions proceeded in non-polar solvent under mild reaction conditions, and the catalyst could be recycled at least five times with only a small decrease in activity (Scheme 12). Excellent results have been achieved with heterogeneous catalysis when using the bipyridine-based metal-organic framework (mBPV-MOF-Ir) (S12-4) and the 1,10-phenanthroline-based metal-organic framework (mPT-MOF-Ir) catalysts developed by Lin et al. [77]. They used them for the C−H borylation of various aromatic and heteroaromatic substrates with very low catalyst loadings and excellent isolated yields of borylated products. Although prolonged reaction periods and high temperatures were required to achieve full conversion, the products were obtained by removing the catalyst via simple centrifugation and the catalysts could be recycled at least 15 times with no apparent loss of activity (Scheme 12). When using the mBPV-MOF-Ir catalyst, a TON of 9000 was achieved for the borylation of indole and the leached iridium content was found to be 0.3 ppm.
Very recently, the first-of-its-kind mechanochemical iridium-catalyzed C−H borylation of indoles S13-1 was described by Kubota and Ito. The borylation reaction took place in a steel jar in an air atmosphere while the starting compounds were ground by a steel ball for 99 min [78]. The authors report that the careful selection of the grinding apparatus was vital if the reaction was to give results. For substrates with melting points higher than 70 °C, liquid-assisted grinding (LAG) was needed, although a catalytic amount of the organic solvent was sufficient. Several variously substituted 2-indolyl boronates S13-2 were prepared in this way. The isolated yields (in parentheses) differed substantially from the NMR yields of the products due to the reported fast hydrolysis on silica gel (Scheme 13).
In an effort to render transition-metal-catalyzed borylation more economically feasible, several methods based on more abundant metals, such as cobalt, nickel, and platinum, have been developed in recent years. Some of the developed protocols are also applicable for indoles, although in most cases, only a very narrow indole substrate scope was tested. In the identified publications, only indoles with unsubstituted C2 positions were tested; therefore, the borylation proceeded selectively into the C2 position in all cases.
Chirik used a series of pincer-type ligand cobalt complexes for the development of the cobalt-catalyzed borylation of aromatic compounds [79]. The complex S14-3 proved to be effective in terms of the selective borylation of indole S14-1 into the C2 position (Scheme 14). Later, Chirik broadened his application of the cobalt complexes for C−H borylation by means of the preparation and evaluation of a new, air-stable and easily preparable cobalt precatalyst, namely (ArTpy)Co(OAc)2 (S14-4), which was also applicable for the borylation of indoles (Scheme 14) [80]. Recently, Smith and Maleczka have shown that a cobalt-based catalyst supported by NHC ligands (S14-5) can also be used for the borylation of benzylic and heteroaromatic compounds, including the indole S14-1, in moderate yields (Scheme 14) [81].
Nickel-catalyzed C2 borylations of various arenes and heteroarenes have been developed by the Chatani group [82]. In this instance, the catalytic system consisted of the transition metal and an NHC ligand. The reaction proceeded selectively into the C2 position in the case of indoles with nitrogen atoms variously substituted by the alkyl or benzyl groups (S15-1). Similar selectivity was achieved when positions 3 and 5 were substituted by the methyl (S15-2) or halogen and methoxy groups (S15-3), respectively (Scheme 15). The authors managed to employ cheap and air-stable Ni(OAc)2 rather than Ni(cod)2 for the gram-scale preparations, and they achieved excellent results. Similar conditions were used for the borylation of indoles under platinum catalysis conditions, although HBpin proved to be unreactive; therefore, B2pin2 was used as the boron source [83]. The reaction with B2pin2 proceeded selectively at the 2 position, and it offered the indoles S16-2 in moderate to very good yields (Scheme 16). Iwasawa also applied platinum catalysis for the borylation of indoles using the PSiN-pincer platinum complex, albeit with diminished selectivity [84].

2.3.2. C3-Selective Borylation

In 2013, Oestreich reported the new and synthetically useful C3-selective borylation of indoles S17-1 by means of the catalytic generation of borenium ions from HBpin [85]. The protocol required no added base, as H2 was released as the sole by-product and allowed to install the Bpin group onto the indole substrate via a SEAr reaction mechanism. Over the course of the reaction, the B−H bond is split by the ruthenium(II) thiolate complex, thereby affording ruthenium(II) hydride and borenium ions, which undergo an attack by the most nucleophilic site of the indole substrate. The subsequent deprotonation offers C3-borylated indoles in moderate to high yields (Scheme 17).
The regioselectivity of the iridium-catalyzed C−H borylation of indoles can be altered by installing a bulky group onto the indole substrate. The tert-butoxycarbonyl (Boc) group can be used in this way for nitrogen-containing heterocycles such as pyrroles, indoles, azaindoles, and pyrazoles where the N-Boc-protected pyrroles and indole S18-1 are selectively borylated into the 3 position [86]. However, the installation and removal of the directing group extend the time required for the overall procedure. Therefore, it is desirable to use a directing group that can be installed and removed without requiring the isolation of the intermediates. The protocol developed by Smith and Maleczka can be used as an example here, with the Bpin group serving as a traceless directing group for the selective C−H borylation of indoles into the 3 position (Scheme 18) [87]. The indole S18-2 is smoothly converted into the N-borylated indole S18-3 by means of the reaction with HBpin. The addition of Et3N is essential if the reaction is to proceed, as the coordination of the lone nitrogen pair with the boron center in the HBpin renders the B−H bond hydridic enough to react with as weak an acid as S18-2. The addition of the iridium catalyst into the reaction mixture causes the C−H borylation to proceed in the C3 position, while the aqueous workup gives 3-borylated indole S18-4 in a 57% yield (Scheme 18).
Colacot et al. studied the causes of prior inconsistent results and diminished the catalytic activity of the Ir complexes formed in situ using the simple and cheap ligand 1,10-phenanthroline. They determined that catalytically inactive cationic [Ir(COD)(phen)]+Cl- (S19-2) is rapidly formed in non-coordinating solvents, thereby diminishing the reactivity (Scheme 19). The authors managed to isolate the active complex [Ir(Cl)(COD)(phen)] (S19-1) and to utilize it for the selective borylation of the Boc-protected indoles S19-3. They achieved excellent results and then applied the catalyst based on the cheap ligand for the one-pot total synthesis of meridianin G (S19-4), thereby achieving an excellent overall yield [88].

2.3.3. C6-, C7-Selective Borylation

Only one example of the selective C6-borylation of indoles can be found in the literature, and even it is not strictly C6-selective but rather C6-preferential. It was developed by Baran et al. in an attempt to achieve the first total syntheses of verruculogen and fumitremorgin A. The C6 selectivity is achieved by means of the careful substrate selection of only 3-substituted indoles S20-1 and the protection of the indole nitrogen by the robust TIPS group, thereby disabling the C−H activation reaction in the otherwise more reactive positions C2, C3, or C7 [89]. By optimizing the reaction conditions, the authors achieved good isolated yields of S20-3 as well as up to 14:1 C6:C5 selectivity on a limited substrate scope (Scheme 20).
Selective transition-metal-catalyzed C−H functionalization at the 7-position typically requires a substituent to be present at the 2-position so as to block reactivity at that site [90,91,92]. An example from the recent literature was developed by Takai, who exploited the iridium-catalyzed silylation of indole (S21-1) to functionalize the 2-position, followed by iridium-catalyzed borylation in the 7-position in a one-pot manner so as to prepare S21-2 (Scheme 21).
The requirement for an occupied 2-position can be avoided by means of the introduction of a silyl-directing group onto the nitrogen atom, as demonstrated by Hartwig in his one-pot protocol consisting of the Ru-catalyzed silylation of indoles, followed by the Ir-catalyzed borylation of the formed silylindole. The mild hydrolysis of the silyl group then selectively afforded the C7-borylated indoles S22-1 (Scheme 22) [93]. The borylation selectivity most likely arises from the formation of the five-membered metallacycle S22-2 through the C−H bond cleavage at the 7-position rather than of the four-membered metallacycle at the 2-position. Sperry applied an analogous approach for the synthesis of indolequinone natural products where even 5- and 6-methoxy-substituted indoles reacted selectively at the 7-position, thereby broadening the applicability of the developed protocol [94].
Inspired by Hartwig’s protocol, Movassaghi et al. developed a more streamlined C7-selective borylation of 3-alkylindoles. Their design was based on the diboronation/protodeboronation sequence. The authors managed to develop reaction conditions wherein the 3-substituted indoles S23-1 underwent iridium-catalyzed C2/C7 diboronation with an excess of HBpin so as to form S23-2 in situ, and subsequently, selective C2 protodeboronation either with a large excess of trifluoroacetic acid or a catalytic amount of Pd(OAc)2 in AcOH (Scheme 23) [95]. Smith and Maleczka later expanded upon this protocol. They developed a complementary method for selective C2 deborylation using the iridium-catalyzed deborylation of 2,7-diborylated indoles to prepare 7-borylated indoles [96]. They also determined that bismuth(III) acetate can be used for the C2 and C7 protodeboronation of indoles. They prepared a series of diborylated and triborylated indoles that underwent selective C2 and C7 deborylation with Bi(OAc)3 in methanol so as to prepare mono- and bisborylated indoles [97]. Recently, the diboronation/protodeboronation method was used by Burns et al. to achieve the total synthesis of Ts-discorhabdin E from tryptamine [98].

2.3.4. Non-Selective Borylations

Several non-selective methods for the borylation of indoles have emerged over the last decade, which either produce multi-borylated indoles or else are highly substrate-specific and differently substituted indoles undergo borylation at different sites [99]. One such method concerns the ester-directed borylation of sulfur-, oxygen-, or nitrogen-containing five-membered heterocycles developed by Sawamura, whereby indoles S24-1 bear the ester functional group borylate at either the 2- or the 3-position depending on where the ester group resides. If the ester group is present at the nitrogen atom, then the borylation occurs at the 2-position, and it only occurs at the 7-position (S24-2) when the 2- and 3-positions are occupied (Scheme 24) [100].
As mentioned above, the borylations of indoles often proceed at multiple sites, most frequently at the C2 and C7 positions [101]. Protodeboronation at C2 can lead to selective C7 borylation, although that is not the only application of the inherent reactivity regarding indole borylations. Sperry applied a diborylation/oxidation protocol for the short total synthesis of an endogenous plant metabolite 7-hydroxyoxindole-3-acetic acid S25-4, which required only one chromatography separation (Scheme 25) [102]. Utilizing the Me4Phen ligand and an excess of B2pin2 was crucial for achieving the full conversion of the S25-1 into S25-2.
Sperry also developed a reaction condition for the threefold borylation of 3-substituted indoles S26-1 by employing iridium catalysis with the Me4Phen ligand. During the course of the reaction, the C−H borylation first proceeded quickly at the C2 and C7 positions, and then further, at the C5 position due to steric effects [103]. Several triborylated 3-substituted indoles S26-2 were prepared by following the described protocol (Scheme 26).
Sperry then expanded the protocol to achieve the synthesis of 5,7-diborylated-3-methylindole S27-3 from 3-methylindole S27-1 utilizing the triborylation/C2-protodeboronation sequence in a one-pot manner (Scheme 27) [104]. The author showed that the two boryl groups in positions 5 and 7 of the S27-3 can be further functionalized in a selective or non-selective way so as to prepare valuable indoles bearing the aryl group, methoxy group, or halogen atoms in the aforementioned positions. Additionally, this method was applied to achieve the total synthesis of (+)-plakohypaphorine C (S27-5) from Boc-protected L-tryptophan methyl ester (S27-4) (Scheme 27).

2.4. Cyclization Protocols for the Synthesis of Indolylboronic Acid Derivatives

Thus far, the synthesis of indolylboronic acids or esters has been achieved via the manipulation of the indolyl moiety. The opposite approach, which consists of the cyclization of the substituted anilines, can also be used to synthesize 3-indolylboronic acids or boronates (Table 3). The cyclization strategy can be performed under transition-metal-free conditions by means of a stoichiometric amount of BCl3 [105] or a catalytic amount of B(C6F5)3 [106] (Table 3, entries 1 and 2). Palladium [107] and gold [108] catalysts can also be used for the same cyclization, affording 3-indolylboronates in comparable isolated yields. The common feature of the above-mentioned protocols is the ability to provide rapid access to 2-subtituted-3-indolylboronates in good isolated yields and wide substrate scopes. Based on both experimental evidence and theoretical investigation [109], plausible mechanisms have been proposed to explain borylative cyclizations. Generally, simplified versions of the proposed mechanisms rely on the activation of the triple bond by means of Lewis acid complexation, and either transition metals or boron-based reagents can be used as acids (Table 3, entry 5). This complexation facilitates the addition of nitrogen to the triple bond. The synthesis of regioisomeric 2-indolylboronates is also feasible [110], although a procedure starting from 2-iodoanilines and MIDAB-acetylene under typical Miyaura borylation conditions would have to be used.
Save for the ortho-substituted aniline, the ortho-substituted phenylisocyanides undergo smooth cyclization so as to give substituted 2-indolylboronic acid pinacol esters S28-2 (Scheme 28) [111]. This methodology leads to excellent yields of the substituted indoles S28-2, although the presence of alkenyl species bearing an ester group is required. This requirement can be explained by a plausible mechanism. It is expected that the copper catalyst activates the pinB-Bpin, followed by a reaction with the isocyanide S28-1. The formed complex S28-3 then undergoes intramolecular addition so as to give copper enolate S28-4 while the ligand substitution releases the 2-indolylboronates S28-2.

2.5. Lewis-acid-catalyzed Borylation of Indoles

It is clear that the transition-metal-catalyzed borylations of indoles have become the methods of choice when designing the functionalization of indoles via indolylboronic acid intermediates. Yet, transition-metal-free borylations offer a relatively simple and cheap route to indolylboronates under mild reaction conditions. In 2013 [112] and 2014 [113], Wang et al. described a procedure that allowed for the conversion of arylamines into indolylboronic acid pinacol esters by means of the reaction with pinB-Bpin in the presence of tBuONO (Scheme 29). The developed methodology was found to be applicable in relation to diverse aromatic and heteroaromatic substrates substituted with both electron-donating and electron-withdrawing groups. Overall, the isolated yields were within the range of 14−80%, and a single example of indolylboronate S9-2 was isolated in a 22% yield. The authors proposed a single-electron transfer (SET) mechanism for their reaction [112].
A different approach comparable to the transition-metal-catalyzed borylation of indoles was reported by Zhang (Scheme 30) [114]. He observed that a catalytic amount of boron trifluoride diethyl etherate was able to catalyze the conversion of substituted indoles into indolylboronic acid pinacol esters. A wide range of indolylboronates were prepared by heating the starting compounds in noctane/THF to 140 °C. The author was able to synthesize indoles with sensitive bromine function S30-2b, unprotected indole S30-2c, as well as 2,3-disubstituted indole S30-2a. It is expected that the reaction proceeds through an initial attack of BF3 on the indole-affording intermediate S30-3. The intermediate S30-3 then reacts with the pinB−Bpin to generate the dihydroindole S30-4. This intermediate S30-4 is disproportionated to the carbocation S30-5 and pinB−BF3, which deprotonates S30-5 so as to yield the indolylboronates S30-2a-c.
The example presented in Scheme 30 illustrates an application of formal electrophilic aromatic substitution as a tool for the borylation of electron-rich arenes. This concept was further exploited by Zhang [115], Oestreich [116], and Erker [117] (Table 4).
Zhang made use of catecholborane (catBH) along with pentafluoroborane as the Lewis acid in the catalytic setup. The reaction proceeds at room temperature as disproportionation affording a mixture of the indolylboronic acid catechol ester and the corresponding 2,3-dihydroindoles in good isolated yields (Table 4, entry 1). This approach allows for easy access to a variety of borylated indoles. However, the convergent disproportionation reaction is able to increase the yield of borylated indoles up to a 98% yield. In this case, the borylation is performed at 120 °C. Under the above-mentioned conditions, the catalyst enables the conversion of indolines into indoles, thereby allowing for a shift in the equilibrium of disproportionation reactions. A similar approach was reported by Oestreich et al. (Table 4, entry 2) [116]. The same reaction conditions were applied, although the scope of the protocol was extended to electron-rich arenes. Interestingly, the reaction was accelerated when performed in the presence of norbornene or norbornadiene. Substantial progress was achieved by running the borylation reaction in the presence of bisborane T4-3, which allowed the borylation to proceed at room temperature with a good isolated yield of the corresponding boronates (Table 4, entry 3) [117].
In a series of papers, Fontaine showed that the frustrated Lewis pair S31-4 can catalyze transition-metal-free C−H borylation (Scheme 31) [118,119,120,121]. The developed conditions are mostly applicable to five-membered heterocycles, thiophene, pyrrole, or furan. However, the substituted indoles S31-2 and S31-3 were prepared in excellent isolated yields.
A different type of electrophilic borylation was reported by Ingleson [122,123,124,125]. This reaction is applicable for the borylation of a wide range of heteroarenes and arenes, including pyrene and anthracene. The reaction can be performed by means of BCl3 [122] or the less reactive ClBcat [123,124] derivative. It was proposed that during the first step, the borylation reagent S32-3 is activated by the presence of triethylamine and aluminum chloride, thereby affording the borenium cation S32-5 as the main reactive species (Scheme 32). Then, the formed borenium cation reacts with the arenes, and following esterification (e.g., with pinacol), provides the products of borylation. Thus, both pinacol esters [122−124] and MIDA-boronates [122] are available via this protocol.
Recently, another different type of borylation reagent has been reported [126]. The protocol involves the treatment of electron-rich alkenes with four equivalents of boron tribromide in the presence of 2,6-lutidine at room temperature. The reaction is limited to electron-rich alkenes, and only one example of indole borylation is provided (Scheme 33).
The final option for the preparation of indolylboronic acids or boronates is based on the chemical manipulation of indoles bearing the −B(OH)2 or −B(OR)2 function. In this regard, regioselective Friedel−Crafts alkylation [127], boronate ligand interconversion [128], palladium or rhodium C−H activation [129,130], metal-free arylation [131,132], and other approaches [133] have emerged in the literature. Although these procedures allow for access to a novel type of indolylboronic acid or ester, they are not discussed in detail in the present paper because indolylboronic acid esters are already used as starting compounds.

3. Chemical Transformations of Indolylboronic Acid Derivatives

3.1. Formal Substitution of Boron Moieties by Transition-Metal-Catalyzed or Transition-Metal-Free Reactions

Indolylboronic acids and their derivatives are all examples of simple and stable heteroaromatic compounds, which have found many applications in the field of organic synthesis, along with other aromatic and heteroaromatic boronates and boronic acids. Presumably, the simplest modification is represented by the protodeboronation that was reported by Macgregor and Lee (Table 5, entry 1) [134]. They found that 5 mol % of a gold catalyst facilitates the protodeboronation of boronic acids in the presence of dimethyl carbonate and water when using a microwave setup. As the reaction is carried out under non-acidic conditions, acid-sensitive functional groups (e.g., the THP group) are tolerated. The use of THF along with D2O allows for a simple deuterodeboronation. Based on density functional theory calculations, a mechanism explaining the observed reactivity has been proposed. A copper catalyst was used for the direct conversion of boronates, boronic acids, and organotrifluoroborates into the azides (Table 5, entry 2) [135].
The developed conditions saw the azidation performed in excellent isolated yields, while the coupling efficiency did not correlate with the electron-donating ability of the substituents. However, the isolated yield of azidoindole was only 51%. The reaction course could be followed colorimetrically as the dark brown solution turned light green upon completion. Recently, the rapid palladium-catalyzed [11C] C-cyanation of aryl- or heteroarylboronates was reported [136]. The reaction conditions allowed for the rapid conversion of the starting boronates by means of treatment with a [11C] NH4CN/NH3 system in the presence of a PdCl2(PPh3)2 catalyst. The high efficiency of the developed protocol was demonstrated via the cyanation of indolylboronate in a 90% radiochemical yield (rcy) and the synthesis of [11C] C-labeled cetrozole.
The ipso-substitution of B-moiety via a halogen atom represents an alternative approach for the synthesis of halogenated arenes. This transformation can be performed under transition-metal-free conditions by means of iodine and KF in dioxane at 80 °C (Table 6, entry 1) [137]. The isolated yield of iodoindole is high. However, this reaction is less important from a practical point of view, as indoles can be regioselectively lithiated at the 3-position by means of tBuLi [138].
The fluorination of indoles is more import, as fluorinated indoles are of interest due to their potential use in the field of medicine. Thus, Hartwig developed the copper(I)-catalyzed fluorination of pinacol boronic ester via an electrophilic fluorine source in THF at 50 °C (Table 6, entry 2) [139]. Overall, the isolated yields of the fluorinated arenes ranged from 29% to 77%, with a 44% isolated yield of 5-fluoroindole being achieved. Radiofluorination is a more attractive concept in relation to the fluorination of indoles. Two different procedures have thus far been reported (Table 6, entries 3 and 4). Both published procedures rely on Cu-mediated radiofluorination with different fluoride sources. Krasikova [140] achieved a radiochemical conversion (rcc) of 84% within 20 min in DMA at 110 °C under phase-transfer catalysis conditions. The same catalytic system was used by Neumaier, although in that case [18F]F- was eluted with Et4NHCO3 in nbutanol [141]. As demonstrated by the reported examples, the achieved rcc rates were found to be almost quantitative (Table 6, entry 4). The developed reaction conditions were used for the synthesis of a PET probe by means of the radiofluorination of the SnMe3 substrate.
Ermert [142] developed a protocol for the copper(II)-catalyzed fluorination of indolylboronic acid pinacol esters by means of Et4N[18F]F. This protocol was applied for the synthesis of 6-6-[18F]fluoro-L-tryptophan (Scheme 34). The starting bromoindole S34-1 was prepared according to the procedure described in the literature [143]. Then, Miyaura borylation was used for the bromine substitution with the Bpin moiety and the resultant radiolabeled fluorination, along with hydrolysis, afforded the tryptophan derivative S34-3 in an overall isolated yield of 37% and moderate rcc.
A logical extension of the halogenation reactions of indolylboronic acids can be seen in the case of trifluoromethylation, since a trifluoromethyl group is a prominent structural motif present in biologically active compounds used in the fields of pharmaceuticals and agrochemicals. An early report concerning such a procedure was published by Buchwald [144] (Table 7, entry 1). This copper-mediated oxidative cross-coupling uses TMSCF3 (Ruppert’s reagent) as a source of nucleophilic CF3 and oxygen as a stoichiometric reoxidant. The isolated yields of trifluoromethylated indoles were 44% and 61%, which demonstrates good agreement with the overall isolated yields. A similar procedure was published by Gooßen (Table 7, entry 2) [145]. This protocol substitutes the inconveniently volatile and moisture-sensitive Ruppert’s reagent for crystalline potassium (trifluoromethyl)trimethoxyborate. Among the other substrates, the isolated yield of 5-(trifluoromethyl)-1-methylindole was only 49%. To overcome the need for stoichiometric amounts of copper, Liu [146] (Table 7, entry 3) suggested the use of (trifluoromethyl)dibenzothiophenium triflate as an electrophilic CF3 source. The greatest advantage offered by this method is the tolerance of an unprotected NH group, which meant that an unprotected trifluoromethylated indole was obtained in a 61% yield. Alternatively, Shen [147] exploited the use of Togni’s reagent [148] using a similar method and achieved the highest isolated yield of trifluoromethylated indole of all the discussed procedures (Table 7, entry 4). A slightly different approach to trifluoromethylation was described by Molander et al. (Table 7, entry 5) [149]. They performed the radical trifluoromethylation of organotrifluoroborates by using sodium triflinate (Langlois reagent) as the source of the CF3 radicals and five equivalents of tert-butyl hydroperoxide (TBHP) in a mixture of solvents. Additionally, the introduction of the trifluoromethoxy group into an indole moiety via the coupling of the corresponding boronic acids was described by Ritter [150]. His two-step, one-pot procedure includes the reaction of indolylboronic acids with NaOH (1 equiv.) and AgPF6 (2 equiv.) in methanol, followed by treatment with TAS∙OCF3 and the Selectfluor-type reagent (F-TEDA-PF6).
The methylation of indole boronic esters can be achieved by a simple cross-coupling reaction with methyl iodide when compared to trifluoromethylation. The iron-catalyzed methylation approach described by Nakamura [151] (Scheme 35, condition A) presumably involves a radical mechanism whereby the methylation of the starting compound proceeds in the solvent cage, thereby affording 1,5-dimethylindole in a 74% isolated yield. The presence of a bulky bidentate ligand (SciPROP-TB) is required in this case. More traditional procedures make use of the cross-coupling between indolylboronic acid pinacol esters and methyl iodide catalyzed by palladium catalyst A, as reported by Hartwig et al. (Scheme 35, condition B) [152].
The authors provided a simple methylation of indomethacin as an example of the late-stage methylation of a bioactive compound. Alternatively, trimethyl phosphate can be used in combination with iodine salt as a milder source of slow-releasing methyl iodide, as described by Hartwig [153] (Scheme 35, condition C). The isolated yields of methylated indoles were 62–70%. It was proposed that the lithium iodide reacts with the trimethyl phosphate (S35-3) to slowly release methyl iodide (S35-4). This reaction represents a rate-determining step. Subsequently, the methyl iodide reacts with the arylcopper intermediate so as to give the product of the methylation. Any side reactions are suppressed by the slow release of methyl iodide as well as the acid base reaction between ArBpin S35-5 and tBuOLi.
Gomez [154] focused on the direct benzylation of 2-indolylboronic acids by means of benzyl bromides in the presence of a trans-PdBr(N-Succ)(PPh3)2 catalyst. This catalyst turned out to be the most efficient in terms of suppressing any undesired protodeboronation. The optimized conditions were applied on substituted indoles and benzyl bromides, thereby affording trisubstituted indoles in good isolated yields (Scheme 36).
A different approach for the synthesis of diarylmethanes was designed by Watson and Garnsey (Table 8, entry 1) [155]. Their methodology started with benzylamines, which were converted into pyridinium salts by refluxing with 2,4,6-triphenylpyrrilium tetrafluoroborate. Then, the benzylation of the indolylboronic acids was achieved by means of an Ni-catalyzed reaction in 1,4-dioxane at 60 °C. The authors hypothesized that the radical mechanism operated along the lines of the developed protocol. This idea was confirmed by the isolation of the benzylated TEMPO radical. The similar nickel-catalyzed benzylation of indolylboronic acid neopentylglycol ester was reported by Tobisu and Chatani (Table 8, entry 2) [156]. The reaction was conducted in toluene at 120 °C in the presence of the NHC ligand. The synthesis of the alkylated indole was performed on a multigram scale in a 90% isolated yield. Analogously to the benzylation protocols, a representative example of the asymmetric allylation of allyl phosphates with boronic acids was described by Hayashi [157]. The allylation of 5-(nepB)-1-methylindole was achieved by means of a CuCl-catalyzed reaction in the presence of sodium methoxide and the N-heterocyclic carbene. The formed allylated indole was isolated in an 89% yield and exhibited acceptable enantiomeric access of 88%.
The hydroxylation of aromatic or heteroaromatic boronic acids represents an alternative reaction pathway for synthesizing the corresponding phenols. In the case of indolylboronic acids or boronates, two protocols have been described in the literature. The first report relies on a C70 fullerene-catalyzed reaction under photocatalytic conditions (Scheme 37, method A) [158]. However, the isolated yield of 6-hydroxyindoles S37-2 is significantly lower when compared to the tested aromatic boronic acids. A quantitative yield of 5-hydroxyindole was achieved by means of oxone oxidation under phase-transfer conditions (Scheme 37, method B) [159]. In addition to the quantitative yield of S37-3, the described protocol allows for the chemoselective oxidation of boronic acids over pinacol boronates. The chemoselectivity is achieved via the faster trihydroxyboronate transfer from the organic layer to the water layer.
In an effort to expand the diversity-oriented synthesis of substituted indoles, two different carbonylation procedures have been reported (Table 9, entries 1 and 2). In the first report, Skrydstrup [160] made use of a traditional palladium catalyst with a Xantphos ligand and the carbonylation of boronic acids with CO and N,N-diethyl-α-bromo-α,α-difluoroacetamide. It was determined that the reaction is suitable for the carbonylation of unprotected indoles, thereby affording the product in a good isolated yield.
The reaction was also extended to include α-bromo-α,α-difluoroacetates. The presence of two equivalents of 1,4-dinitrobenzene suppressed the reaction, while the two equivalents of TEMPO resulted in the isolation of the corresponding adducts. These experiments indicated a SET mechanism for the carbonylation. Different from Skrydstrup’s protocol, Szostak [161] applied an alternative method for the carbonylation of boronic acids. His protocol is based on the cross-coupling reaction of the activated C−N amide bond. Thus, the cyclic amide T9-3 was found to be the best substrate, affording various ketones along with the indole in good isolated yields (Table 9, entry 2). The methodology of the CO carbonylation of indolylboronic acid pinacol boronic esters was extended to cover methoxycarbonylation (Table 9, entry 3) [162]. Weinreb amides are useful synthons for Weinreb ketone synthesis [163]. Therefore, it is unsurprising that the simple and high-yielding synthesis of Weinreb amides has been pursued (Table 9, entry 4) [164]. Different boronic acids and organotrifluoroborates were mixed with N-methoxy-N-methylcarbamoyl chloride in the presence of PdCl2(PPh3)2 and potassium phosphate in ethanol, thereby affording, for instance, 1H protected and unprotected indole-based amides in good isolated yields.
Among the other aromatic and heteroaromatic boronates, 1-methyl-5-nepBindole smoothly reacted with arylcarbamate in the rhodium-catalyzed reaction so as to yield 5-substituted indole in a 93% isolated yield (Table 10, entry 1). Interestingly, the cross-coupling reaction was achieved by means of a catalytic amount of the Rh catalyst [165]. The Liebeskind−Srogl cross-coupling reaction was used for the synthesis of substituted BODIPY involving Biellmann BODIPY (Table 10, entry 2) [166]. The reaction was catalyzed by Pd2(dba)3 in the presence of a catalytic amount of trifurylphosphine and a stoichiometric amount of CuTC, thereby leading to variously substituted BODIPYs. The reaction can also be performed under microwave heating, and the fluorescent efficiency was tested on prepared substances.
A similar type of cross-coupling reaction was described by Pentelute and Buchwald (Scheme 38) [167]. In their report, they applied an umpolung strategy for the bioconjugation of the selenocysteine in unprotected peptides. The reaction was conducted by mixing boronic acids in the presence of copper (II) sulfate and the phenanthroline ligand. The incorporation of diverse indolylboronic acids S38-2 and S38-3 was achieved in good yields.
Arvidsson et al. [168]. developed a general methodology for the synthesis of substituted β-aryl ethenesulfonyl fluorides by means of oxidative Heck coupling. The optimized ligandless reaction conditions involved copper(II) acetate as the oxidant and THF as the solvent. The reaction was applied to the synthesis of 5-substituted indole and the following reaction with methylamine resulted in β-sultam S39-3 in an 85% isolated yield (Scheme 39).
The conditions necessary for a transition-metal-free reaction between electron-rich aryl/hetarylboronic acids or trifluoroborates and unprotected vicinal diols have also been published (Scheme 40) [169]. Two different reaction conditions were developed in order to gain opposite stereochemical outcomes. Trifluoroacetic anhydride preferentially formed the substituted indole S40-2 in a 76% isolated yield as a single diastereomer upon reaction with boronic acid.
Alternatively, Bu4NHSO4 along with organotrifluoroborate gave the pure diastereomer S40-3 in an 80% yield. Based on the experimental evidence, it is expected that the reaction proceeds according to two different types of alcohol S40-1 activations. In the case of the boronic acids, the formation of a cyclic ester is preferred, leading to the product S40-4, which in turn forms the product S40-2. Meanwhile, Bu4NHSO4 facilitates the formation of carbocation S40-5, resulting in the formation of the opposite diastereomer.
A simple approach to the synthesis of substituted indoles was developed by Aggarwal (Scheme 41) [170]. The protocol starts with easily available pinacol boronic esters, which are combined with vinylmagnesium chlorides or vinyllithiums. The formed ate complexes are transformed into the final product via the use of iodine and sodium methoxide. The synthesis of the substituted indoles S41-1 and S41-3 in acceptable yields is achieved.
In 2018, Niu et al. [171] developed a simple protocol for the transition-metal-free amination of boronic acids. The authors proposed that the initially formed ate complex S42-3 requires hydroxyl group activation so as to accomplish the amination (Scheme 42). Indeed, the activation was achieved with trichloroacetonitrile in tbutanol. Thus, the synthesis of 5-(N-benzylamino)-1H-indole was completed in a 90% isolated yield. Aside from the other aromatic and heteroaromatic substrates, the same methodology was used for the amination of a tripeptide, a steroidal moiety and a protected carbohydrate showing excellent functional group tolerance. In addition to the aforementioned transition-metal-free amination of boronic acids, the copper(II)-acetate-catalyzed oxidative coupling of 1H-pyrazolo[3–b]pyridine-amine and 2-aminobenzimidazoles has been described in the literature [172].

3.2. Cyclization Reactions of Indolylboronic Acid Derivatives

Several cyclizations of diverse substrates have been developed covering both indolylboronic acids and transition-metal catalysis. In 2019, the Darses research group developed a methodology for the asymmetric cyclization of (allyl)propargylether catalyzed by a rhodium catalyst (Scheme 43) [173]. The reaction is performed in methanol at 60 °C, and the reaction outcome depends on the structure of the ligand. When conducted in the presence of 1-Np-MSBod (S43-3), the exclusive formation of 3,4-disubstitutedtetrahydrofurans was observed in 47−60% isolated yields. The indolyl derivative S43-2 was isolated in a 54% yield and 95% ee. The use of the similar bicyclo[2.2.2]octa-2,5-diene ligand led to tetrahydrobenzo[d]ixepines in a similar yield and ee.
As demonstrated by Bower, palladium-catalyzed cyclizations of oxime esters or N-(pentafluorobenzoyloxy)carbamates have been successfully used to synthesize trisubstituted 1-pyrrolines [174] (Scheme 44a) and pyrrolidine [175] (Scheme 44b). Both strategies take advantage of the N-pentafluorobenzoyloxy group, which activates the N−O bond for the oxidative addition of palladium, although it requires two different ligands. The catalyzed alkene 1,2-carboamination reaction requires the tris[3,5-bis(trifluoromethyl)phenyl]phosphine ligand (Scheme 44a), while the aza-Heck cyclization relies on an adamantyl-type ligand (Scheme 44b) to give good isolated yields of the products S44-1−S44-3.
Analogously to Bower’s work, Selander et al. developed the 1,2-aminoarylation of oxime ester-tethered alkenes catalyzed by a nickel catalyst along with the phenanthroline ligand, thereby affording 2,5-disubstituted-1-pyrrolines (Scheme 45) [176]. The cyclization reaction showed a good substrate scope with isolated yields ranging from 38−82%. The indolylboronic acid gave the product S45-2 in a 62% isolated yield. The authors proposed the cleavage of the N−O bond by means of SET, with the subsequent cyclization proceeding via a radical addition to the double bond. This assumption was confirmed by trapping the radical intermediate with TEMPO. The developed conditions were applied to the synthesis of ABI-274, a potent tubulin inhibitor effective in relation to multidrug-resistant cancer cells.
There exists an inverse approach to substituted indoles based on a cyclobutanone ring expansion (Scheme 46) [177]. In this case, the substituted cyclobutanone S46-1 was treated with 5-indolylboronic acid in the presence of a palladium precatalyst and the tricyclohexylphosphine ligand. The corresponding 5-substituted indole was formed in a high isolated yield and an excellent Z:E ratio. The methodology also tolerates other substituted phenyl, styrenyl, and heteroarylboronic acids. Based on the deuterium experiments, the formation of the palladium complex S46-3 by means of 1,2-addition is expected as a key reaction intermediate. The subsequent β-carbon elimination via Csp3−Csp2 bond cleavage and protolysis is responsible for the formation of the final product.

3.3. Multicomponent Reactions of Indolylboronic Acid Derivatives

Multicomponent reactions can significantly facilitate the preparation of complex structures of organic molecules due to lowering the required number of reaction steps. The use of boronic acids is widespread in such reactions, as the creation of the quaternary borate salt known as the “ate” complex enables the transfer of the boron substituent onto, for example, an imine – as is the case in the Petasis-Borono-Mannich multicomponent reaction. The classical metal-free Petasis reaction is the topic of the work published by Manolikakes [178] (Scheme 47). He used sulfonamides as the amine component in a three-component reaction with a glyoxylic and a boronic acid. This way, the medicinally important sulfonamide group was introduced into various α-substituted glycines, although the yield of the indole-based substrate S47-2 was only 48%. The sulfonamide group in the prepared compounds could serve as a protecting group and a possible site of alteration for modulating the biological properties and increasing the metabolic stability of the compound.
A more general methodology for the synthesis of sulfones, sulfonamides [179], and sulfonate esters [180] was developed by Willis et al. (Scheme 48). In their reports, the synthesis of the target compounds is realized via the conversion of aryl/hetarylboronic acid, for example, indolylboronic acid S48-1, into the sulfonates S48-2. The reaction is carried out in the presence of the sulfur dioxide surrogate 1,4-diazabicyclo[2.2.2]octane bis(sulfur dioxide) (DABSO) and a Pd-catalyst. The direct conversion of the sulfinate S48-2 into the sulfones S48-4 is performed under transition-metal-free conditions by means of tert-butyl α-bromoacetate, while the conversion into the sulfonate esters S48-3 is mediated by copper(II) bromide. A similar direct copper-catalyzed sulfonamides synthesis was reported by the same authors [181]. Primary amines and sterically demanding secondary amines are not tolerated in this reaction. The authors also incorporated several pharmaceutically important amines into the structures of the indole-5-sulfonamides so as to display the potential value of this method in relation to drug design and discovery.
In 2012, Nielsen [182] described a three-component coupling reaction involving α-hydroxy aldehydes and boronic acids (Scheme 49) in a mixture of solvents, namely MeOH:hexafluoroisopropanol (HFIP). The products formed in the Petasis reaction varied from 26–95% in terms of the yield. The hydrazido alcohols S49-4 obtained using this protocol were subjected to a cyclization reaction in relation to the corresponding oxadiazolones and oxazolidinones, that is the target compounds of the work due to their possible antibacterial activity. Later on, the same authors [183] reported a four-component Petasis-type reaction that was used for the synthesis of dioxadiazaborocines.
A Petasis-type reaction was utilized by Onomura [184] when designing a versatile procedure for the diasterocontrolled synthesis of cis-2,3-disubstitutedpiperidine S50-3 (Scheme 50). The Lewis-acid-catalyzed reaction of indolyl boronic acid S50-1 with cyclic N,O-acetal S50-2 provided the cyclic carbamate S50-3.
In 2017, Ready [185] reported on the three-component palladium-catalyzed coupling of indoles with boronic esters and allylic acetates. The multicomponent reaction generates ate complexes S51-2 via the reaction of indolylboronic acid pinacol esters S51-1 (Scheme 51). Then, the formed ate complexes S51-2 are rearranged into 2,3-dihydroindoles S51-3. Overall, the isolated yields of the disubstituted indoles S51-4a−S51-4d are high, thereby demonstrating high diastereomeric and enantiomeric ratios.
The other multicomponent reactions include indolylboronic acids are summarized in Figure 2. The palladium-catalyzed multicomponent reaction [186] of indole boronic acids includes the 1,1-diarylation of terminal alkenes by means of boronic acids and aryldiazonium salts. This method is highly enantioselective due to the utilization of a chiral anion phase-transfer catalyst, and the indole-based product F3-1 was obtained in a 29% isolated yield and an excellent enantiomeric ratio.
Another possibility with regard to 1,2-diarylation, as published by Xiao [187], employs nitrogen-centered radicals created from cyclobutanone oxime derivatives. The reaction is carried out in the presence of a copper(I) catalyst and with 4,4′-di-tert-butyl-2,2′-dipyridyl (dtbbpy) serving as the ligand. This multicomponent reaction is able to accommodate the 3-indolyl moiety, affording the product F3-2 in an excellent isolated yield. A related procedure concerning the 1,2-addition of boronic esters and aryl triflate to vinyl magnesium chloride was published by Morken [188]. The stereoselective preparation of trisubstituted alkenes with perfluoroalkyl chains was also described [189]. This four-component protocol makes use of carbon monoxide, terminal alkynes, perfluoralkyliodide, and arylboronic acids. A trisubstituted alkene F3-3 was prepared in a moderate yield, while the overall isolated yields ranged from 43−90%. The double-bond configuration was determined by the through space 13C−19F coupling.

3.4. Addition of Indolylboronic Acids or Boronates to Multiple Bonds

The 1,4-addition of aryl- or hetarylboronic acids can be catalyzed by transition metals. Both published procedures (Table 11, entries 1 and 2) are catalyzed by a rhodium catalyst under identical reaction conditions.
Viaud-Massuard [190] developed the asymmetric addition on 7-azaindoles as an example of an exocyclic benzylidene lactamyl Michael acceptor. The reaction was performed in the presence of the (R,R)-Phbod ligand, which ensured the 94% ee of the indole-based product (Table 11, entry 1). A ligandless condition was described by Howell et al. (Table 11, entry 2) [191]. They studied the Michael addition of boronic acids onto α-methylene-β-lactones under similar reaction conditions [190]. Generally speaking, the reaction provided a mixture of diastereomers, including β-propiolactone substituted by a 1-methyl-5-indolyl substituent. The synthesis of nocardiolactone was also described in this paper. The palladium-catalyzed γ-selective hydroarylation of a β,γ-unsaturated secondary amide was reported by Engle (Table 11, entry 3) [192]. In his report, palladium acetate was used to catalyze a regioselective hydroarylation with arylboronic acids, although the substrate scope was limited to amides bearing the N-8-aminoquinoline (AQ) substituent. Based on mechanistic studies, it was proposed that palladium coordinates to the AQ moiety, and further, that transmetalation represents the rate-controlling step. In a different paper, Scarso et al. studied the course of the conjugated addition of boronic acids to vinylidenebisphosphonate (VBP) [193]. This reaction represents a valuable contribution to the synthesis of gem-bisphosphonates. The reaction takes advantage of a simple copper(II) catalyst and tolerates diverse boronic acids, including 4-indolylboronic acids, which afforded the expected product in an 83% isolated yield (Table 11, entry 4). The same catalytic system is able to facilitate the Friedel–Crafts reaction between VBP and indoles, thereby expanding the scope for the preparation of gem-bisphosphonates. The regio- and stereoselective hydroarylation of propargylic carbamates by means of aryl- heteroarylboronic acids catalyzed by a Ni(cod)2 catalyst was reported by Jarvo (Table 11, entry 5) [194]. The rac-2-(di-tert-butylphosphino)-1,1′-binaphthyl (TrixiePhos) was determined to be the ligand of choice in terms of maintaining the quantitative conversion of the starting alkynes, along with excellent regio- and stereoselectivity. The isolated yields of the trisubstituted alkenes were good, which is illustrated by the provided example (Table 11, entry 5). The methodology was later extended to include the formal synthesis of tamoxifen.
Recently, Hyland described a different kind of (Z)-selective palladium-catalyzed addition of aryl boronic acids to vinylaziridines (Scheme 52) [195]. This reaction makes use of palladium acetate as a catalyst in combination with the phenanthroline ligand and AgSbF6. Although the reaction is said to be (Z)-selective, however the reported Z:E ratio varies between 1.6:1 and 4.5:1. This reactivity is demonstrated by the reaction of indolylboronic acid S52-1, which gave a 38% yield of the substituted indole S52-2 as a mixture of Z:E isomers in a ratio of 1.8:1.
Organocatalytic additions are considered to be more environmentally acceptable than transition-metal-catalyzed reactions. However, practical examples concerning a series of indolylboronic acids can only be found in three reports. In 2012, Schaus described the enantioselective addition of diethyl boronates to o-quinone methides S53-1 (Scheme 53) [196]. The reaction is catalyzed by 10 mol % of the BINOL ligand in toluene at room temperature. Aromatic and heteroaromatic boronates provide a high yield (46−85%) with regard to aromatic phenols, in addition to an excellent enantiomeric ratio of ≥90:10. In the case of indolylboronate S53-2, the isolated yield of the product S53-3 was only 46% and the enantiomeric ratio was 96:4. Moreover, the published protocol was applied to the short synthesis of (S)-4-methoxydalbergione.
A combination of the organocatalytic Michael addition of boronic acids and the diastereoselective intramolecular Passerini reaction was used for the synthesis of chiral γ-lactones S54-4 (Scheme 54) [197,198]. This one-pot sequential synthesis starts with the reaction between the Boc-protected 2-indolylboronic acids S54-1 and 5-hydroxyfuran-2(5H)-one (S54-2) in the presence of a diphenylprolinol catalyst. The organocatalyst activates the starting 5-hydroxyfuran-2(5H)-one, while the subsequent addition of indolylboronic acid S54-1 provides the acyclic intermediate S54-3, which undergoes a diastereoselective Passerini reaction with isonitriles. A series of chiral lactones S54-4aS54-4c substituted with the indole moiety were prepared using this methodology.
In addition to the above-mentioned additions of diverse boronic acids to activated and non-activated double bonds, more traditional 1,2-addition reactions with C-heteroatom double bonds have been described. Thus, the organocatalytic formylation of boronic acids with glyoxylic acid [199], the palladium-imidazolinium carbene-catalyzed arylation of aldehydes with boronic acids [200], the rhodium-catalyzed addition of boronic acids to 2,2-disubstituted malononitriles [201,202], the rhodium-catalyzed addition of boronic acids to α-ketoesters [203], and the addition of boronic acids to nitrosoarenes [204] are all valuable examples.

4. Conclusions

The aforementioned examples of transition-metal-catalyzed cross-coupling reactions, multicomponent reactions, addition reactions, and transition-metal-free alternatives, document tremendous progress in relation to the synthetic applications of indolylboronic acids and aryl boronic acid in general. This progress can be illustrated by over 4000 patent applications describing chemical transformations of indolylboronic acid derivatives. Significant progress has been made in the area of indolylboronic acids synthesis via transition-metal-catalyzed C−H borylations, Miyaura borylation, and ortho-substituted aniline cyclizations. Also transition-metal-catalyzed and transition-metal-free reactions of indolylboronic acid derivatives were intensively studied in the past. It can be expected, that more publications concerning a cooperative catalysis of indolylboronic acids will emerge soon as examples from new area of transition-metal-catalyzed reactions. Moreover, increased attention should focus on multicomponent reactions of indolylboronic acids as a simple and efficient way to complex indole-based molecules. Thus, the discovery of novel transformations and reactions on the part of indole boronates or boronic acids will undoubtedly widen the pool of modified indoles, which may find interesting practical applications in the near future.

Author Contributions

All authors contributed equally to this work.

Funding

This work was funded by specific university research grant number MSMT No 21-SVV/2019 and the APC was funded by University of Chemistry and Technology Prague.

Acknowledgments

Financial support was received from a specific university research grant (MSMT No 21-SVV/2019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Foley, C.A.; Al-Issa, Y.A.; Hiller, K.P.; Mulcahy, S.P. Synthesis and Structure–Activity Relationships of 1-Aryl-β-carbolines as Affinity Probes for the 5-Hydroxytryptamine Receptor. ACS Omega 2019, 4, 9807–9812. [Google Scholar] [CrossRef]
  2. Cornelio, B.; Laronze-Cochard, M.; Miambo, R.; De Grandis, M.; Riccioni, R.; Borisova, B.; Dontchev, D.; Machado, C.; Ceruso, M.; Fontana, A.; et al. 5-Arylisothiazol-3(2H)-one-1,(1)-(di)oxides: A new class of selective tumor-associated carbonic anhydrases (hCA IX and XII) inhibitors. Eur. J. Med. Chem. 2019, 175, 40–48. [Google Scholar] [CrossRef] [PubMed]
  3. Ellis, J.M.; Altman, M.D.; Cash, B.; Haidle, A.M.; Kubiak, R.L.; Maddess, M.L.; Yan, Y.; Northrup, A.B. Carboxamide Spleen Tyrosine Kinase (Syk) Inhibitors: Leveraging Ground State Interactions To Accelerate Optimization. ACS Med. Chem. Lett. 2016, 7, 1151–1155. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Zhang, N.; Turpoff, A.; Zhang, X.; Huang, S.; Liu, Y.; Almstead, N.; Njoroge, F.G.; Gu, Z.; Graci, J.; Jung, S.P.; et al. Discovery of 2-(4-sulfonamidophenyl)-indole 3-carboxamides as potent and selective inhibitors with broad hepatitis C virus genotype activity targeting HCV NS4B. Bioorg. Med. Chem. Lett. 2016, 26, 594–601. [Google Scholar] [CrossRef] [PubMed]
  5. Wu, Y.-J.; Venables, B.; Guernon, J.; Chen, J.; Sit, S.-Y.; Rajamani, R.; Knox, R.J.; Matchett, M.; Pieschl, R.L.; Herrington, J.; et al. Discovery of new indole-based acylsulfonamide Nav1.7 inhibitors. Bioorg. Med. Chem. Lett. 2019, 29, 659–663. [Google Scholar] [CrossRef]
  6. Kanada, R.; Tanabe, M.; Muromoto, R.; Sato, Y.; Kuwahara, T.; Fukuda, H.; Arisawa, M.; Matsuda, T.; Watanabe, M.; Shuto, S. Synthesis of Chiral cis-Cyclopropane Bearing Indole and Chromone as Potential TNFα Inhibitors. J. Org. Chem. 2018, 83, 7672–7682. [Google Scholar] [CrossRef]
  7. Bartoccini, F.; Venturi, S.; Retini, M.; Mari, M.; Piersanti, G. Total Synthesis of (−)-Clavicipitic Acid via γ,γ-Dimethylallyltryptophan (DMAT) and Chemoselective C–H Hydroxylation. J. Org. Chem. 2019, 84, 8027–8034. [Google Scholar] [CrossRef]
  8. Nabi, A.A.; Liyu, J.; Lindsay, A.C.; Sperry, J. C4−H alkoxylation of 6-bromoindole and its application to the synthesis of breitfussin B. Tetrahedron 2018, 74, 1199–1202. [Google Scholar] [CrossRef]
  9. Hu, L.; Li, Q.; Yao, L.; Xu, B.; Wang, X.; Liao, X. Enantioselective and Divergent Syntheses of Alstoscholarisines A, E and Their Enantiomers. Org. Lett. 2018, 20, 6202–6205. [Google Scholar] [CrossRef]
  10. Chen, Z.; Zhou, S.; Jia, Y. Formal Synthesis of (+)-Kopsihainanine A and Synthetic Study toward (+)-Limaspermidine. J. Org. Chem. 2015, 80, 12545–12551. [Google Scholar] [CrossRef]
  11. Zhou, S.; Jia, Y. Total Synthesis of (−)-Goniomitine. Org. Lett. 2014, 16, 3416–3418. [Google Scholar] [CrossRef] [PubMed]
  12. Baeyer, A.; Emmerling, A. Synthese des Indols. Ber. Dtsch. Chem. Ges. 1869, 2, 679–682. [Google Scholar] [CrossRef]
  13. Bartoli, G.; Palmieri, G.; Bosco, M.; Dalpozzo, R. The reaction of vinyl grignard reagents with 2-substituted nitroarenes: A new approach to the synthesis of 7-substituted indoles. Tetrahedron Lett. 1989, 30, 2129–2132. [Google Scholar] [CrossRef]
  14. Fischer, E.; Hess, O. Synthese von Indolderivaten. Ber. Dtsch. Chem. Ges. 1884, 17, 559–568. [Google Scholar] [CrossRef] [Green Version]
  15. Tokuyama, H.; Yamashita, T.; Reding, M.T.; Kaburagi, Y.; Fukuyama, T. Radical Cyclization of 2-Alkenylthioanilides:  A Novel Synthesis of 2,3-Disubstituted Indoles. J. Am. Chem. Soc. 1999, 121, 3791–3792. [Google Scholar] [CrossRef]
  16. Gassman, P.G.; Van Bergen, T.J.; Gruetzmacher, G. Use of halogen-sulfide complexes in the synthesis of indoles, oxindoles, and alkylated aromatic amines. J. Am. Chem. Soc. 1973, 95, 6508–6509. [Google Scholar] [CrossRef]
  17. Larock, R.C.; Yum, E.K. Synthesis of indoles via palladium-catalyzed heteroannulation of internal alkynes. J. Am. Chem. Soc. 1991, 113, 6689–6690. [Google Scholar] [CrossRef]
  18. Tanner, M.E. Mechanistic studies on the indole prenyltransferases. Nat. Prod. Rep. 2015, 32, 88–101. [Google Scholar] [CrossRef]
  19. Bandini, M.; Melloni, A.; Tommasi, S.; Umani-Ronchi, A. A Journey Across Recent Advances in Catalytic and Stereoselective Alkylation of Indoles. Synlett 2005, 1199–1222. [Google Scholar] [CrossRef]
  20. Makosza, M. Vicarious Nucleophilic Substitution of Hydrogen in the Chemistry of Heterocyclic Compounds. Synthesis 1991, 103–111. [Google Scholar] [CrossRef]
  21. Vorobyeva, D.V.; Osipov, S.N. Selective Synthesis of 2- and 7-Substituted Indole Derivatives via Chelation-Assisted Metallocarbenoid C–H Bond Functionalization. Synthesis 2018, 50, 227–240. [Google Scholar]
  22. Petrini, M. Regioselective Direct C-Alkenylation of Indoles. Chem. Eur. J. 2017, 23, 16115–16151. [Google Scholar] [CrossRef] [PubMed]
  23. Bheeter, C.B.; Chen, L.; Soulé, J.-F.; Doucet, H. Regioselectivity in palladium-catalysed direct arylation of 5-membered ring heteroaromatics. Catal. Sci. Technol. 2016, 6, 2005–2049. [Google Scholar] [CrossRef]
  24. Sandtorv, A.H. Transition Metal-Catalyzed C−H Activation of Indoles. Adv. Synth. Catal. 2015, 357, 2403–2435. [Google Scholar] [CrossRef]
  25. Rossi, R.; Bellina, F.; Lessi, M.; Manzini, C. Cross-Coupling of Heteroarenes by C−H Functionalization: Recent Progress towards Direct Arylation and Heteroarylation Reactions Involving Heteroarenes Containing One Heteroatom. Adv. Synth. Catal. 2014, 356, 17–117. [Google Scholar] [CrossRef]
  26. Miyaura, N.; Yamada, K.; Suzuki, A. A new stereospecific cross-coupling by the palladium-catalyzed reaction of 1-alkenylboranes with 1-alkenyl or 1-alkynyl halides. Tetrahedron Lett. 1979, 20, 3437–3440. [Google Scholar] [CrossRef] [Green Version]
  27. Miyaura, N.; Suzuki, A. Stereoselective synthesis of arylated (E)-alkenes by the reaction of alk-1-enylboranes with aryl halides in the presence of palladium catalyst. J. Chem. Soc., Chem. Commun. 1979, 866–867. [Google Scholar] [CrossRef]
  28. Amatore, C.; Le Duc, G.; Jutand, A. Mechanism of Palladium-Catalyzed Suzuki–Miyaura Reactions: Multiple and Antagonistic Roles of Anionic “Bases” and Their Countercations. Chem. Eur. J. 2013, 19, 10082–10093. [Google Scholar] [CrossRef]
  29. Düfert, M.A.; Billingsley, K.L.; Buchwald, S.L. Suzuki-Miyaura Cross-Coupling of Unprotected, Nitrogen-Rich Heterocycles: Substrate Scope and Mechanistic Investigation. J. Am. Chem. Soc. 2013, 135, 12877–12885. [Google Scholar] [CrossRef] [Green Version]
  30. Wilson, K.L.; Murray, J.; Sneddon, H.F.; Jamieson, C.; Watson, A.J.B. Dimethylisosorbide (DMI) as a Bio-Derived Solvent for Pd-Catalyzed Cross-Coupling Reactions. Synlett 2018, 29, 2293–2297. [Google Scholar] [Green Version]
  31. Polák, P.; Tobrman, T. The synthesis of polysubstituted indoles from 3-bromo-2-indolyl phosphates. Org. Biomol. Chem. 2017, 15, 6233–6241. [Google Scholar] [CrossRef] [PubMed]
  32. Rossi, E.; Abbiati, G.; Canevari, V.; Celentano, G.; Magri, E. 2-Trifluoromethanesulfonyloxyindole-1-carboxylic Acid Ethyl Ester: A Practical Intermediate for the Synthesis of 2-Carbosubstituted Indoles. Synthesis 2006, 299–304. [Google Scholar] [CrossRef]
  33. Nazari, S.H.; Bourdeau, J.E.; Talley, M.R.; Valdivia-Berroeta, G.A.; Smith, S.J.; Michaelis, D.J. Nickel-Catalyzed Suzuki Cross Couplings with Unprotected Allylic Alcohols Enabled by Bidentate N-Heterocyclic Carbene (NHC)/Phosphine Ligands. ACS Catal. 2018, 8, 86–89. [Google Scholar] [CrossRef]
  34. Tan, J.; Chen, Y.; Li, H.; Yasuda, N. Suzuki-Miyaura Cross-Coupling Reactions of Unprotected Haloimidazoles. J. Org. Chem. 2014, 79, 8871–8876. [Google Scholar] [CrossRef]
  35. Kotek, V.; Polák, P.; Dvořáková, H.; Tobrman, T. Aluminum Chloride Promoted Cross-Coupling of Trisubstituted Enol Phosphates with Organozinc Reagents En Route to the Stereoselective Synthesis of Tamoxifen and Its Analogues. Eur. J. Org. Chem. 2016, 5037–5044. [Google Scholar] [CrossRef]
  36. Nambo, M.; Keske, E.C.; Rygus, J.P.G.; Yim, J.C.H.; Crudden, C.M. Development of Versatile Sulfone Electrophiles for Suzuki–Miyaura Cross-Coupling Reactions. ACS Catal. 2017, 7, 1108–1112. [Google Scholar] [CrossRef]
  37. Chang, S.; Sun, Y.B.; Zhang, X.R.; Dong, L.L.; Zhu, H.Y.; Lai, H.W.; Wang, D. Pd-catalyzed Desulfitative reaction of Aryltrifluoroborates with sodium Arenesulfinates in water. Appl. Organomet. Chem. 2018, 32, e3970. [Google Scholar] [CrossRef]
  38. Landstrom, E.B.; Handa, S.; Aue, D.H.; Gallou, F.; Lipshutz, B.H. EvanPhos: A ligand for ppm level Pd-catalyzed Suzuki–Miyaura couplings in either organic solvent or water. Green Chem. 2018, 20, 3436–3443. [Google Scholar] [CrossRef]
  39. Zou, Y.; Yue, G.; Xu, J.; Zhou, J. General Suzuki Coupling of Heteroaryl Bromides by Using Tri-tert-butylphosphine as a Supporting Ligand. Eur. J. Org. Chem. 2014, 5901–5905. [Google Scholar] [CrossRef]
  40. Boit, T.B.; Weires, N.A.; Kim, J.; Garg, N.K. Nickel-Catalyzed Suzuki–Miyaura Coupling of Aliphatic Amides. ACS Catal. 2018, 8, 1003–1008. [Google Scholar] [CrossRef]
  41. Schäfer, P.; Palacin, T.; Sidera, M.; Fletcher, S.P. Asymmetric Suzuki-Miyaura coupling of heterocycles via Rhodium-catalysed allylic arylation of racemates. Nat. Commun. 2017, 8, 15762. [Google Scholar] [CrossRef] [PubMed]
  42. Iwamoto, T.; Okuzono, C.; Adak, L.; Jin, M.; Nakamura, M. Iron-catalysed enantioselective Suzuki–Miyaura coupling of racemic alkyl bromides. Chem. Commun. 2019, 55, 1128–1131. [Google Scholar] [CrossRef] [PubMed]
  43. Cobb, K.M.; Rabb-Lynch, J.M.; Hoerrner, M.E.; Manders, A.; Zhou, Q.; Watson, M.P. Stereospecific, Nickel-Catalyzed Suzuki–Miyaura Cross-Coupling of Allylic Pivalates To Deliver Quaternary Stereocenters. Org. Lett. 2017, 19, 4355–4358. [Google Scholar] [CrossRef] [PubMed]
  44. Li, G.; Wang, E.; Chen, H.; Li, H.; Liu, Y.; Wang, P.G. A general and efficient synthesis of substituted fluorenes and heterocycle-fused indenes containing thiophene or indole rings utilizing a Suzuki–Miyaura coupling and acid-catalyzed Friedel–Crafts reactions as key steps. Tetrahedron 2008, 64, 9033–9043. [Google Scholar] [CrossRef]
  45. Vazquez, E.; Davies, I.W.; Payack, J.F. A Non-cryogenic Method for the Preparation of 2-(Indolyl) Borates, Silanes, and Silanols. J. Org. Chem. 2002, 67, 7551–7552. [Google Scholar] [CrossRef] [PubMed]
  46. Schneider, C.; Broda, E.; Snieckus, V. Directed ortho-Metalation–Cross-Coupling Strategies. One-Pot Suzuki Reaction to Biaryl and Heterobiaryl Sulfonamides. Org. Lett. 2011, 13, 3588–3591. [Google Scholar] [CrossRef] [PubMed]
  47. Shen, K.; Fu, Y.; Li, J.-N.; Liu, L.; Guo, Q.-X. What are the pKa values of C–H bonds in aromatic heterocyclic compounds in DMSO? Tetrahedron 2007, 63, 1568–1576. [Google Scholar] [CrossRef]
  48. Mesganaw, T.; Fine Nathel, N.F.; Garg, N.K. Cine Substitution of Arenes Using the Aryl Carbamate as a Removable Directing Group. Org. Lett. 2012, 14, 2918–2921. [Google Scholar] [CrossRef]
  49. Al-Saedy, M.A.E.; Harrity, J.P.A. Synthesis and Stabilities of 3-Borylated Indoles. Synlett 2016, 27, 1674–1676. [Google Scholar]
  50. Guerrand, H.D.S.; Marciasini, L.D.; Jousseaume, M.; Vaultier, M.; Pucheault, M. Borylation of Unactivated Aryl Chlorides under Mild Conditions by Using Diisopropylaminoborane as a Borylating Reagent. Chem. Eur. J. 2014, 20, 5573–5579. [Google Scholar] [CrossRef]
  51. Nitelet, A.; Thevenet, D.; Schiavi, B.; Hardouin, C.; Fournier, J.; Tamion, R.; Pannecoucke, X.; Jubault, P.; Poisson, T. Copper-Photocatalyzed Borylation of Organic Halides under Batch and Continuous-Flow Conditions. Chem. Eur. J. 2019, 25, 3262–3266. [Google Scholar] [CrossRef] [PubMed]
  52. Verma, P.K.; Mandal, S.; Geetharani, K. Efficient Synthesis of Aryl Boronates via Cobalt-Catalyzed Borylation of Aryl Chlorides and Bromides. ACS Catal. 2018, 8, 4049–4054. [Google Scholar] [CrossRef]
  53. Labre, F.; Gimbert, Y.; Bannwarth, P.; Olivero, S.; Duñach, E.; Chavant, P.Y. Application of Cooperative Iron/Copper Catalysis to a Palladium-Free Borylation of Aryl Bromides with Pinacolborane. Org. Lett. 2014, 16, 2366–2369. [Google Scholar] [CrossRef] [PubMed]
  54. Lim, S.; Song, D.; Jeon, S.; Kim, Y.; Kim, H.; Lee, S.; Cho, H.; Lee, B.C.; Kim, S.E.; Kim, K.; et al. Cobalt-Catalyzed C–F Bond Borylation of Aryl Fluorides. Org. Lett. 2018, 20, 7249–7252. [Google Scholar] [CrossRef]
  55. Liu, X.-W.; Echavarren, J.; Zarate, C.; Martin, R. Ni-Catalyzed Borylation of Aryl Fluorides via C–F Cleavage. J. Am. Chem. Soc. 2015, 137, 12470–12473. [Google Scholar] [CrossRef]
  56. Tobisu, M.; Kinuta, H.; Kita, Y.; Rémond, E.; Chatani, N. Rhodium(I)-Catalyzed Borylation of Nitriles through the Cleavage of Carbon–Cyano Bonds. J. Am. Chem. Soc. 2012, 134, 115–118. [Google Scholar] [CrossRef]
  57. Kinuta, H.; Kita, Y.; Rémond, E.; Tobisu, M.; Chatani, N. Novel Synthetic Approach to Arylboronates via Rhodium-Catalyzed Carbon–Cyano Bond Cleavage of Nitriles. Synthesis 2012, 44, 2999–3002. [Google Scholar] [CrossRef]
  58. Chen, K.; Cheung, M.S.; Lin, Z.; Li, P. Metal-free borylation of electron-rich aryl (pseudo)halides under continuous-flow photolytic conditions. Org. Chem. Front. 2016, 3, 875–879. [Google Scholar] [CrossRef]
  59. Yamamoto, E.; Izumi, K.; Horita, Y.; Ito, H. Anomalous Reactivity of Silylborane: Transition-Metal-Free Boryl Substitution of Aryl, Alkenyl, and Alkyl Halides with Silylborane/Alkoxy Base Systems. J. Am. Chem. Soc. 2012, 134, 19997–20000. [Google Scholar] [CrossRef]
  60. Yamamoto, E.; Izumi, K.; Horita, Y.; Ukigai, S.; Ito, H. Formal Nucleophilic Boryl Substitution of Organic Halides with Silylborane/Alkoxy Base System. Top. Catal. 2014, 57, 940–945. [Google Scholar] [CrossRef] [Green Version]
  61. Erb, W.; Hellal, A.; Albini, M.; Rouden, J.; Blanchet, J. An Easy Route to (Hetero)arylboronic Acids. Chem. Eur. J. 2014, 20, 6608–6612. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, C.; Ji, C.-L.; Hong, X.; Szostak, M. Palladium-Catalyzed Decarbonylative Borylation of Carboxylic Acids: Tuning Reaction Selectivity by Computation. Angew. Chem. 2018, 57, 16721–16726. [Google Scholar] [CrossRef] [PubMed]
  63. Ochiai, H.; Uetake, Y.; Niwa, T.; Hosoya, T. Rhodium-Catalyzed Decarbonylative Borylation of Aromatic Thioesters for Facile Diversification of Aromatic Carboxylic Acids. Angew. Chem. Int. Ed. 2017, 56, 2482–2486. [Google Scholar] [CrossRef] [PubMed]
  64. Cheng, W.-M.; Shang, R.; Zhao, B.; Xing, W.-L.; Fu, Y. Isonicotinate Ester Catalyzed Decarboxylative Borylation of (Hetero)Aryl and Alkenyl Carboxylic Acids through N-Hydroxyphthalimide Esters. Org. Lett. 2017, 19, 4291–4294. [Google Scholar] [CrossRef] [PubMed]
  65. Candish, L.; Teders, M.; Glorius, F. Transition-Metal-Free, Visible-Light-Enabled Decarboxylative Borylation of Aryl N-Hydroxyphthalimide Esters. J. Am. Chem. Soc. 2017, 139, 7440–7443. [Google Scholar] [CrossRef]
  66. Iverson, C.N.; Smith, M.R. Stoichiometric and Catalytic B−C Bond Formation from Unactivated Hydrocarbons and Boranes. J. Am. Chem. Soc. 1999, 121, 7696–7697. [Google Scholar] [CrossRef]
  67. Ishiyama, T.; Takagi, J.; Ishida, K.; Miyaura, N.; Anastasi, N.R.; Hartwig, J.F. Mild iridium-catalyzed borylation of arenes. High turnover numbers, room temperature reactions, and isolation of a potential intermediate. J. Am. Chem. Soc. 2002, 124, 390–391. [Google Scholar] [CrossRef]
  68. Ishiyama, T.; Nobuta, Y.; Hartwig, J.F.; Miyaura, N. Room temperature borylation of arenes and heteroarenes using stoichiometric amounts of pinacolborane catalyzed by iridium complexes in an inert solvent. Chem. Commun. 2003, 2924–2925. [Google Scholar] [CrossRef]
  69. Liskey, C.W.; Hartwig, J.F. Borylation of Arenes with Bis(hexylene glycolato)diboron. Synthesis 2013, 45, 1837–1842. [Google Scholar]
  70. Tobisu, M.; Igarashi, T.; Chatani, N. Iridium/N-heterocyclic carbene-catalyzed C–H borylation of arenes by diisopropylaminoborane. Beilstein J. Org. Chem. 2016, 12, 654–661. [Google Scholar] [CrossRef]
  71. Hoque, M.E.; Bisht, R.; Haldar, C.; Chattopadhyay, B. Noncovalent Interactions in Ir-Catalyzed C–H Activation: L-Shaped Ligand for Para-Selective Borylation of Aromatic Esters. J. Am. Chem. Soc. 2017, 139, 7745–7748. [Google Scholar] [CrossRef] [PubMed]
  72. Bisht, R.; Hoque, M.E.; Chattopadhyay, B. Amide Effects in C−H Activation: Noncovalent Interactions with L-Shaped Ligand for meta Borylation of Aromatic Amides. Angew. Chem. Int. Ed. 2018, 57, 15762–15766. [Google Scholar] [CrossRef] [PubMed]
  73. Sasaki, I.; Ikeda, T.; Amou, T.; Taguchi, J.; Ito, H.; Ishiyama, T. Regioselective C–H Borylation of Heteroaromatic Aldimines with Iridium Complexes. Synlett 2016, 27, 1582–1586. [Google Scholar]
  74. Wu, F.; Feng, Y.; Jones, C.W. Recyclable Silica-Supported Iridium Bipyridine Catalyst for Aromatic C–H Borylation. ACS Catal. 2014, 4, 1365–1375. [Google Scholar] [CrossRef]
  75. Maegawa, Y.; Inagaki, S. Iridium–bipyridine periodic mesoporous organosilica catalyzed direct C–H borylation using a pinacolborane. Dalton Trans. 2015, 44, 13007–13016. [Google Scholar] [CrossRef] [PubMed]
  76. Tahir, N.; Muniz-Miranda, F.; Everaert, J.; Tack, P.; Heugebaert, T.; Leus, K.; Vincze, L.; Stevens, C.V.; Van Speybroeck, V.; Van Der Voort, P. Immobilization of Ir(I) complex on covalent triazine frameworks for CH borylation reactions: A combined experimental and computational study. J. Catal. 2019, 371, 135–143. [Google Scholar] [CrossRef]
  77. Manna, K.; Zhang, T.; Greene, F.X.; Lin, W. Bipyridine- and phenanthroline-based metal-organic frameworks for highly efficient and tandem catalytic organic transformations via directed C–H activation. J. Am. Chem. Soc. 2015, 137, 2665–2673. [Google Scholar] [CrossRef] [PubMed]
  78. Pang, Y.; Ishiyama, T.; Kubota, K.; Ito, H. Iridium(I)-Catalyzed C−H Borylation in Air by Using Mechanochemistry. Chem. Eur. J. 2019, 25, 4654–4659. [Google Scholar] [CrossRef]
  79. Obligacion, J.V.; Semproni, S.P.; Chirik, P.J. Cobalt-catalyzed C-H borylation. J. Am. Chem. Soc. 2014, 136, 4133–4136. [Google Scholar] [CrossRef]
  80. Léonard, N.G.; Bezdek, M.J.; Chirik, P.J. Cobalt-Catalyzed C(sp2)–H Borylation with an Air-Stable, Readily Prepared Terpyridine Cobalt(II) Bis(acetate) Precatalyst. Organometallics 2017, 36, 142–150. [Google Scholar] [CrossRef]
  81. Jayasundara, C.R.K.; Sabasovs, D.; Staples, R.J.; Oppenheimer, J.; Smith, M.R.; Maleczka, R.E. Cobalt-Catalyzed C–H Borylation of Alkyl Arenes and Heteroarenes Including the First Selective Borylations of Secondary Benzylic C–H Bonds. Organometallics 2018, 37, 1567–1574. [Google Scholar] [CrossRef]
  82. Furukawa, T.; Tobisu, M.; Chatani, N. Nickel-catalyzed borylation of arenes and indoles via C–H bond cleavage. Chem. Commun. 2015, 51, 6508–6511. [Google Scholar] [CrossRef] [PubMed]
  83. Furukawa, T.; Tobisu, M.; Chatani, N. C–H Functionalization at Sterically Congested Positions by the Platinum-Catalyzed Borylation of Arenes. J. Am. Chem. Soc. 2015, 137, 12211–12214. [Google Scholar] [CrossRef] [PubMed]
  84. Takaya, J.; Ito, S.; Nomoto, H.; Saito, N.; Kirai, N.; Iwasawa, N. Fluorine-controlled C–H borylation of arenes catalyzed by a PSiN-pincer platinum complex. Chem. Commun. 2015, 51, 17662–17665. [Google Scholar] [CrossRef] [PubMed]
  85. Stahl, T.; Müther, K.; Ohki, Y.; Tatsumi, K.; Oestreich, M. Catalytic Generation of Borenium Ions by Cooperative B–H Bond Activation: The Elusive Direct Electrophilic Borylation of Nitrogen Heterocycles with Pinacolborane. J. Am. Chem. Soc. 2013, 135, 10978–10981. [Google Scholar] [CrossRef] [PubMed]
  86. Kallepalli, V.A.; Shi, F.; Paul, S.; Onyeozili, E.N.; Maleczka, R.E., Jr.; Smith, M.R., III. Boc groups as protectors and directors for Ir-catalyzed C-H borylation of heterocycles. J. Org. Chem. 2009, 74, 9199–9201. [Google Scholar] [CrossRef] [PubMed]
  87. Preshlock, S.M.; Plattner, D.L.; Maligres, P.E.; Krska, S.W.; Maleczka, R.E., Jr.; Smith, M.R., III. A Traceless Directing Group for C−H Borylation. Angew. Chem. Int. Ed. 2013, 52, 12915–12919. [Google Scholar] [CrossRef] [PubMed]
  88. Seechurn, C.C.C.J.; Sivakumar, V.; Satoskar, D.; Colacot, T.J. Iridium-Catalyzed C–H Borylation of Heterocycles Using an Overlooked 1,10-Phenanthroline Ligand: Reinventing the Catalytic Activity by Understanding the Solvent-Assisted Neutral to Cationic Switch. Organometallics 2014, 33, 3514–3522. [Google Scholar] [CrossRef]
  89. Feng, Y.; Holte, D.; Zoller, J.; Umemiya, S.; Simke, L.R.; Baran, P.S. Total Synthesis of Verruculogen and Fumitremorgin A Enabled by Ligand-Controlled C–H Borylation. J. Am. Chem. Soc. 2015, 137, 10160–10163. [Google Scholar] [CrossRef] [PubMed]
  90. Meyer, F.-M.; Liras, S.; Guzman-Perez, A.; Perreault, C.; Bian, J.; James, K. Functionalization of Aromatic Amino Acids via Direct C−H Activation: Generation of Versatile Building Blocks for Accessing Novel Peptide Space. Org. Lett. 2010, 12, 3870–3873. [Google Scholar] [CrossRef] [PubMed]
  91. Eastabrook, A.S.; Wang, C.; Davison, E.K.; Sperry, J. A Procedure for Transforming Indoles into Indolequinones. J. Org. Chem. 2015, 80, 1006–1017. [Google Scholar] [CrossRef]
  92. Batool, F.; Parveen, S.; Emwas, A.-H.; Sioud, S.; Gao, X.; Munawar, M.A.; Chotana, G.A. Synthesis of Fluoroalkoxy Substituted Arylboronic Esters by Iridium-Catalyzed Aromatic C–H Borylation. Org. Lett. 2015, 17, 4256–4259. [Google Scholar] [CrossRef] [PubMed]
  93. Robbins, D.W.; Boebel, T.A.; Hartwig, J.F. Iridium-Catalyzed, Silyl-Directed Borylation of Nitrogen-Containing Heterocycles. J. Am. Chem. Soc. 2010, 132, 4068–4069. [Google Scholar] [CrossRef] [PubMed]
  94. Wang, C.; Sperry, J. Iridium-Catalyzed C–H Borylation-Based Synthesis of Natural Indolequinones. J. Org. Chem. 2012, 77, 2584–2587. [Google Scholar] [CrossRef] [PubMed]
  95. Loach, R.P.; Fenton, O.S.; Amaike, K.; Siegel, D.S.; Ozkal, E.; Movassaghi, M. C7-Derivatization of C3-Alkylindoles Including Tryptophans and Tryptamines. J. Org. Chem. 2014, 79, 11254–11263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Kallepalli, V.A.; Gore, K.A.; Shi, F.; Sanchez, L.; Chotana, G.A.; Miller, S.L.; Maleczka, R.E.; Smith, M.R. Harnessing C–H Borylation/Deborylation for Selective Deuteration, Synthesis of Boronate Esters, and Late Stage Functionalization. J. Org. Chem. 2015, 80, 8341–8353. [Google Scholar] [CrossRef] [PubMed]
  97. Shen, F.; Tyagarajan, S.; Perera, D.; Krska, S.W.; Maligres, P.E.; Smith, M.R.; Maleczka, R.E. Bismuth Acetate as a Catalyst for the Sequential Protodeboronation of di-and Triborylated Indoles. Org. Lett. 2016, 18, 1554–1557. [Google Scholar] [CrossRef]
  98. Smith, M.W.; Falk, I.D.; Ikemoto, H.; Burns, N.Z. A convenient C–H functionalization platform for pyrroloiminoquinone alkaloid synthesis. Tetrahedron 2019, 75, 3366–3370. [Google Scholar] [CrossRef]
  99. Wang, G.; Xu, L.; Li, P. Double N,B-Type Bidentate Boryl Ligands Enabling a Highly Active Iridium Catalyst for C–H Borylation. J. Am. Chem. Soc. 2015, 137, 8058–8061. [Google Scholar] [CrossRef]
  100. Kawamorita, S.; Ohmiya, H.; Sawamura, M. Ester-Directed Regioselective Borylation of Heteroarenes Catalyzed by a Silica-Supported Iridium Complex. J. Org. Chem. 2010, 75, 3855–3858. [Google Scholar] [CrossRef]
  101. Demory, E.; Devaraj, K.; Orthaber, A.; Gates, P.J.; Pilarski, L.T. Boryl (Hetero)aryne Precursors as Versatile Arylation Reagents: Synthesis through C−H Activation and Orthogonal Reactivity. Angew. Chem. Int. Ed. 2015, 54, 11765–11769. [Google Scholar] [CrossRef] [PubMed]
  102. Homer, J.A.; Sperry, J. A short synthesis of the endogenous plant metabolite 7-hydroxyoxindole-3-acetic acid (7-OH-OxIAA) using simultaneous C–H borylations. Tetrahedron Lett. 2014, 55, 5798–5800. [Google Scholar] [CrossRef]
  103. Eastabrook, A.S.; Sperry, J. Iridium-Catalyzed Triborylation of 3-Substituted Indoles. Australian J. Chem. 2015, 68, 1810–1814. [Google Scholar] [CrossRef]
  104. Eastabrook, A.S.; Sperry, J. Synthetic Access to 3,5,7-Trisubstituted Indoles Enabled by Iridium-Catalyzed C–H Borylation. Synthesis 2017, 49, 4731–4737. [Google Scholar]
  105. Lv, J.; Zhao, B.; Liu, L.; Han, Y.; Yuan, Y.; Shi, Z. Boron Trichloride-Mediated Synthesis of Indoles via the Aminoboration of Alkynes. Adv. Synth. Catal. 2018, 360, 4054–4059. [Google Scholar] [CrossRef]
  106. Yuan, K.; Wang, S. trans-Aminoboration across Internal Alkynes Catalyzed by B (C6F5)3 for the Synthesis of Borylated Indoles. Org. Lett. 2017, 19, 1462–1465. [Google Scholar] [CrossRef] [PubMed]
  107. Huang, J.; Macdonald, S.J.F.; Harrity, J.P.A. A borylative cyclisation towards indole boronic esters. Chem. Commun. 2010, 46, 8770–8772. [Google Scholar] [CrossRef]
  108. Chong, E.; Blum, S.A. Aminoboration: Addition of B–N σ Bonds across C–C π Bonds. J. Am. Chem. Soc. 2015, 137, 10144–10147. [Google Scholar] [CrossRef]
  109. Jiang, J.; Zhang, Z.; Fu, Y. Theoretical Investigation on the ClBcat-Promoted Synthesis of Heterocyclic Boronic Esters. Asian, J. Org. Chem. 2017, 6, 282–289. [Google Scholar] [CrossRef]
  110. Seath, C.P.; Wilson, K.L.; Campbell, A.; Mowat, J.M.; Watson, A.J.B. Synthesis of 2-BMIDA 6,5-bicyclic heterocycles by Cu(i)/Pd(0)/Cu(ii) cascade catalysis of 2-iodoaniline/phenols. Chem. Commun. 2016, 52, 8703–8706. [Google Scholar] [CrossRef] [Green Version]
  111. Tobisu, M.; Fujihara, H.; Koh, K.; Chatani, N. Synthesis of 2-Boryl- and Silylindoles by Copper-Catalyzed Borylative and Silylative Cyclization of 2-Alkenylaryl Isocyanides. J. Org. Chem. 2010, 75, 4841–4847. [Google Scholar] [CrossRef] [PubMed]
  112. Qiu, D.; Jin, L.; Zheng, Z.; Meng, H.; Mo, F.; Wang, X.; Zhang, Y.; Wang, J. Synthesis of Pinacol Arylboronates from Aromatic Amines: A Metal-Free Transformation. J. Org. Chem. 2013, 78, 1923–1933. [Google Scholar] [CrossRef] [PubMed]
  113. Qiu, D.; Zhang, Y.; Wang, J. Direct synthesis of arylboronic pinacol esters from arylamines. Org. Chem. Front. 2014, 1, 422–425. [Google Scholar] [CrossRef] [Green Version]
  114. Zhong, Q.; Qin, S.; Yin, Y.; Hu, J.; Zhang, H. Boron(III)-Catalyzed C2-Selective C−H Borylation of Heteroarenes. Angew. Chem. Int. Ed. 2018, 57, 14891–14895. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, S.; Han, Y.; He, J.; Zhang, Y. B (C6F5)3-Catalyzed C3-Selective C–H Borylation of Indoles: Synthesis, Intermediates, and Reaction Mechanism. J. Org. Chem. 2018, 83, 1377–1386. [Google Scholar] [CrossRef] [PubMed]
  116. Yin, Q.; Klare, H.F.T.; Oestreich, M. Catalytic Friedel–Crafts C−H Borylation of Electron-Rich Arenes: Dramatic Rate Acceleration by Added Alkenes. Angew. Chem. Int. Ed. 2017, 56, 3712–3717. [Google Scholar] [CrossRef]
  117. Liu, Y.-L.; Kehr, G.; Daniliuc, C.G.; Erker, G. Metal-Free Arene and Heteroarene Borylation Catalyzed by Strongly Electrophilic Bis-boranes. Chem. Eur. J. 2017, 23, 12141–12144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Légaré Lavergne, J.; Jayaraman, A.; Misal Castro, L.C.; Rochette, É.; Fontaine, F.-G. Metal-Free Borylation of Heteroarenes Using Ambiphilic Aminoboranes: On the Importance of Sterics in Frustrated Lewis Pair C–H Bond Activation. J. Am. Chem. Soc. 2017, 139, 14714–14723. [Google Scholar] [CrossRef]
  119. Légaré, M.-A.; Courtemanche, M.-A.; Rochette, É.; Fontaine, F.-G. Metal-free catalytic C-H bond activation and borylation of heteroarenes. Science 2015, 349, 513–516. [Google Scholar] [CrossRef] [Green Version]
  120. Légaré, M.-A.; Rochette, É.; Lavergne, J.L.; Bouchard, N.; Fontaine, F.-G. Bench-stable frustrated Lewis pair chemistry: Fluoroborate salts as precatalysts for the C–H borylation of heteroarenes. Chem. Commun. 2016, 52, 5387–5390. [Google Scholar] [CrossRef]
  121. Jayaraman, A.; Misal Castro, L.C.; Fontaine, F.-G. Practical and Scalable Synthesis of Borylated Heterocycles Using Bench-Stable Precursors of Metal-Free Lewis Pair Catalysts. Org. Process. Res. Dev. 2018, 22, 1489–1499. [Google Scholar] [CrossRef]
  122. Bagutski, V.; Del Grosso, A.; Carrillo, J.A.; Cade, I.A.; Helm, M.D.; Lawson, J.R.; Singleton, P.J.; Solomon, S.A.; Marcelli, T.; Ingleson, M.J. Mechanistic Studies into Amine-Mediated Electrophilic Arene Borylation and Its Application in MIDA Boronate Synthesis. J. Am. Chem. Soc. 2013, 135, 474–487. [Google Scholar] [CrossRef] [PubMed]
  123. Del Grosso, A.; Singleton, P.J.; Muryn, C.A.; Ingleson, M.J. Pinacol Boronates by Direct Arene Borylation with Borenium Cations. Angew. Chem. Int. Ed. 2011, 50, 2102–2106. [Google Scholar] [CrossRef]
  124. Grosso, A.D.; Helm, M.D.; Solomon, S.A.; Caras-Quintero, D.; Ingleson, M.J. Simple inexpensive boron electrophiles for direct arene borylation. Chem. Commun. 2011, 47, 12459–12461. [Google Scholar] [CrossRef] [PubMed]
  125. Solomon, S.A.; Del Grosso, A.; Clark, E.R.; Bagutski, V.; McDouall, J.J.W.; Ingleson, M.J. Reactivity of Lewis Acid Activated Diaza- and Dithiaboroles in Electrophilic Arene Borylation. Organometallics 2012, 31, 1908–1916. [Google Scholar] [CrossRef]
  126. Tanaka, S.; Saito, Y.; Yamamoto, T.; Hattori, T. Electrophilic Borylation of Terminal Alkenes with BBr3/2,6-Disubstituted Pyridines. Org. Lett. 2018, 20, 1828–1831. [Google Scholar] [CrossRef]
  127. Bartolucci, S.; Bartoccini, F.; Righi, M.; Piersanti, G. Direct, Regioselective, and Chemoselective Preparation of Novel Boronated Tryptophans by Friedel–Crafts Alkylation. Org. Lett. 2012, 14, 600–603. [Google Scholar] [CrossRef]
  128. Churches, Q.I.; Hooper, J.F.; Hutton, C.A. A General Method for Interconversion of Boronic Acid Protecting Groups: Trifluoroborates as Common Intermediates. J. Org. Chem. 2015, 80, 5428–5435. [Google Scholar] [CrossRef] [Green Version]
  129. Borah, A.J.; Shi, Z. Palladium-catalyzed regioselective C–H fluoroalkylation of indoles at the C4-position. Chem. Commun. 2017, 53, 3945–3948. [Google Scholar] [CrossRef]
  130. Fei, X.; Li, C.; Yu, X.; Liu, H. Rh(III)-Catalyzed Hydroarylation of Alkyne MIDA Boronates via C–H Activation of Indole Derivatives. J. Org. Chem. 2019, 84, 6840–6850. [Google Scholar] [CrossRef]
  131. Caramenti, P.; Nandi, R.K.; Waser, J. Metal-Free Oxidative Cross Coupling of Indoles with Electron-Rich (Hetero) arenes. Chem. Eur. J. 2018, 24, 10049–10053. [Google Scholar] [CrossRef] [PubMed]
  132. Caramenti, P.; Nicolai, S.; Waser, J. Indole- and Pyrrole-BX: Bench-Stable Hypervalent Iodine Reagents for Heterocycle Umpolung. Chem. Eur. J. 2017, 23, 14702–14706. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  133. Chen, J.; Wu, J. Catalytic vinylogous cross-coupling reactions of rhenium vinylcarbenoids. Chem. Sci. 2018, 9, 2489–2492. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Barker, G.; Webster, S.; Johnson, D.G.; Curley, R.; Andrews, M.; Young, P.C.; Macgregor, S.A.; Lee, A.-L. Gold-Catalyzed Proto- and Deuterodeboronation. J. Org. Chem. 2015, 80, 9807–9816. [Google Scholar] [CrossRef] [PubMed]
  135. Grimes, K.D.; Gupte, A.; Aldrich, C.C. Copper(II)-Catalyzed Conversion of Aryl/Heteroaryl Boronic Acids, Boronates, and Trifluoroborates into the Corresponding Azides: Substrate Scope and Limitations. Synthesis 2010, 1441–1448. [Google Scholar] [CrossRef]
  136. Zhang, Z.; Niwa, T.; Watanabe, Y.; Hosoya, T. Palladium(ii)-mediated rapid 11C-cyanation of (hetero)arylborons. Org. Biomol. Chem. 2018, 16, 7711–7716. [Google Scholar] [CrossRef]
  137. Tramutola, F.; Chiummiento, L.; Funicello, M.; Lupattelli, P. Practical and efficient ipso-iodination of arylboronic acids via KF/I2 system. Tetrahedron Lett. 2015, 56, 1122–1123. [Google Scholar] [CrossRef]
  138. Matsuzono, M.; Fukuda, T.; Iwao, M. Direct C-3 lithiation of 1-(triisopropylsilyl) indole. Tetrahedron Lett. 2001, 42, 7621–7623. [Google Scholar] [CrossRef]
  139. Fier, P.S.; Luo, J.; Hartwig, J.F. Copper-Mediated Fluorination of Arylboronate Esters. Identification of a Copper(III) Fluoride Complex. J. Am. Chem. Soc. 2013, 135, 2552–2559. [Google Scholar] [CrossRef] [Green Version]
  140. Antuganov, D.; Zykov, M.; Timofeev, V.; Timofeeva, K.; Antuganova, Y.; Orlovskaya, V.; Fedorova, O.; Krasikova, R. Copper-Mediated Radiofluorination of Aryl Pinacolboronate Esters: A Straightforward Protocol by Using Pyridinium Sulfonates. Eur. J. Org. Chem. 2019, 918–922. [Google Scholar] [CrossRef]
  141. Zischler, J.; Kolks, N.; Modemann, D.; Neumaier, B.; Zlatopolskiy, B.D. Alcohol-Enhanced Cu-Mediated Radiofluorination. Chem. Eur. J. 2017, 23, 3251–3256. [Google Scholar] [CrossRef] [PubMed]
  142. Schäfer, D.; Weiß, P.; Ermert, J.; Castillo Meleán, J.; Zarrad, F.; Neumaier, B. Preparation of No-Carrier-Added 6-[18F]Fluoro-l-tryptophan via Cu-Mediated Radiofluorination. Eur. J. Org. Chem. 2016, 4621–4628. [Google Scholar]
  143. Konas, D.W.; Seci, D.; Tamimi, S. Synthesis of (l)-4-Fluorotryptophan. Synthetic Commun. 2012, 42, 144–152. [Google Scholar] [CrossRef]
  144. Senecal, T.D.; Parsons, A.T.; Buchwald, S.L. Room temperature aryl trifluoromethylation via copper-mediated oxidative cross-coupling. J. Org. Chem. 2011, 76, 1174–1176. [Google Scholar] [CrossRef] [PubMed]
  145. Khan, B.A.; Buba, A.E.; Gooßen, L.J. Oxidative trifluoromethylation of arylboronates with shelf-stable potassium (trifluoromethyl) trimethoxyborate. Chem. Eur. J. 2012, 18, 1577–1581. [Google Scholar] [CrossRef] [PubMed]
  146. Xu, J.; Luo, D.F.; Xiao, B.; Liu, Z.J.; Gong, T.J.; Fu, Y.; Liu, L. Copper-catalyzed trifluoromethylation of aryl boronic acids using a CF3+ reagent. Chem. Commun. 2011, 47, 4300–4302. [Google Scholar] [CrossRef]
  147. Liu, T.; Shen, Q. Copper-catalyzed trifluoromethylation of aryl and vinyl boronic acids with an electrophilic trifluoromethylating reagent. Org. Lett. 2011, 13, 2342–2345. [Google Scholar] [CrossRef]
  148. Eisenberger, P.; Gischig, S.; Togni, A. Novel 10-I-3 Hypervalent Iodine-Based Compounds for Electrophilic Trifluoromethylation. Chem. Eur. J. 2006, 12, 2579–2586. [Google Scholar] [CrossRef]
  149. Presset, M.; Oehlrich, D.; Rombouts, F.; Molander, G.A. Copper-mediated radical trifluoromethylation of unsaturated potassium organotrifluoroborates. J. Org. Chem. 2013, 78, 12837–12843. [Google Scholar] [CrossRef]
  150. Huang, C.; Liang, T.; Harada, S.; Lee, E.; Ritter, T. Silver-mediated trifluoromethoxylation of aryl stannanes and arylboronic acids. J. Am. Chem. Soc. 2011, 133, 13308–13310. [Google Scholar] [CrossRef]
  151. Nakajima, S.; Takaya, H.; Nakamura, M. Iron-catalyzed methylation of arylboron compounds with iodomethane. Chem. Lett. 2017, 46, 711–714. [Google Scholar] [CrossRef]
  152. Haydl, A.M.; Hartwig, J.F. Palladium-Catalyzed Methylation of Aryl, Heteroaryl, and Vinyl Boronate Esters. Org. Lett. 2019, 21, 1337–1341. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. He, Z.T.; Li, H.; Haydl, A.M.; Whiteker, G.T.; Hartwig, J.F. Trimethylphosphate as a Methylating Agent for Cross Coupling: A Slow-Release Mechanism for the Methylation of Arylboronic Esters. J. Am. Chem. Soc. 2018, 140, 17197–17202. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Kearney, A.M.; Landry-Bayle, A.; Gomez, L. Palladium-catalyzed benzylation of N-Boc indole boronic acids. Tetrahedron Lett. 2010, 51, 2281–2283. [Google Scholar] [CrossRef]
  155. Liao, J.; Guan, W.; Boscoe, B.P.; Tucker, J.W.; Tomlin, J.W.; Garnsey, M.R.; Watson, M.P. Transforming Benzylic Amines into Diarylmethanes: Cross-Couplings of Benzylic Pyridinium Salts via C–N Bond Activation. Org. Lett. 2018, 20, 3030–3033. [Google Scholar] [CrossRef] [PubMed]
  156. Tobisu, M.; Yasutome, A.; Kinuta, H.; Nakamura, K.; Chatani, N. 1,3-Dicyclohexylimidazol-2-ylidene as a Superior Ligand for the Nickel-Catalyzed Cross-Couplings of Aryl and Benzyl Methyl Ethers with Organoboron Reagents. Org. Lett. 2014, 16, 5572–5575. [Google Scholar] [CrossRef]
  157. Takeda, M.; Takatsu, K.; Shintani, R.; Hayashi, T. Synthesis of quaternary carbon stereocenters by copper-catalyzed asymmetric allylic substitution of allyl phosphates with arylboronates. J. Org. Chem. 2014, 79, 2354–2367. [Google Scholar] [CrossRef] [PubMed]
  158. Kumar, I.; Sharma, R.; Kumar, R.; Kumar, R.; Sharma, U. C70 Fullerene-Catalyzed Metal-Free Photocatalytic ipso-Hydroxylation of Aryl Boronic Acids: Synthesis of Phenols. Adv. Synth. Catal. 2018, 360, 2013–2019. [Google Scholar] [CrossRef]
  159. Molloy, J.J.; Clohessy, T.A.; Irving, C.; Anderson, N.A.; Lloyd-Jones, G.C.; Watson, A.J.B. Chemoselective oxidation of aryl organoboron systems enabled by boronic acid-selective phase transfer. Chem. Sci. 2017, 8, 1551–1559. [Google Scholar] [CrossRef] [PubMed]
  160. Andersen, T.L.; Frederiksen, M.W.; Domino, K.; Skrydstrup, T. Direct Access to α,α-Difluoroacylated Arenes by Palladium-Catalyzed Carbonylation of (Hetero) Aryl Boronic Acid Derivatives. Angew. Chem. Int. Ed. 2016, 55, 10396–10400. [Google Scholar] [CrossRef]
  161. Meng, G.; Szostak, M. Palladium-catalyzed Suzuki–Miyaura coupling of amides by carbon–nitrogen cleavage: General strategy for amide N–C bond activation. Org. Biomol. Chem. 2016, 14, 5690–5707. [Google Scholar] [CrossRef] [PubMed]
  162. Yamamoto, Y. The First General and Selective Palladium(II)-Catalyzed Alkoxycarbonylation of Arylboronates: Interplay among Benzoquinone-Ligated Palladium(0) Complex, Organoboron, and Alcohol Solvent. Adv. Synth. Catal. 2010, 352, 478–492. [Google Scholar] [CrossRef]
  163. Nahm, S.; Weinreb, S.M. N-methoxy-N-methylamides as effective acylating agents. Tetrahedron Lett. 1981, 22, 3815–3818. [Google Scholar] [CrossRef]
  164. Krishnamoorthy, R.; Lam, S.Q.; Manley, C.M.; Herr, R.J. Palladium-Catalyzed Preparation of Weinreb Amides from Boronic Acids and N-Methyl-N-methoxycarbamoyl Chloride. J. Org. Chem. 2010, 75, 1251–1258. [Google Scholar] [CrossRef] [PubMed]
  165. Nakamura, K.; Yasui, K.; Tobisu, M.; Chatani, N. Rhodium-catalyzed cross-coupling of aryl carbamates with arylboron reagents. Tetrahedron 2015, 71, 4484–4489. [Google Scholar] [CrossRef] [Green Version]
  166. Betancourt-Mendiola, L.; Valois-Escamilla, I.; Arbeloa, T.; Bañuelos, J.; López Arbeloa, I.; Flores-Rizo, J.O.; Hu, R.; Lager, E.; Gómez-Durán, C.F.A.; Belmonte-Vázquez, J.L.; et al. Scope and Limitations of the Liebeskind–Srogl Cross-Coupling Reactions Involving the Biellmann BODIPY. J. Org. Chem. 2015, 80, 5771–5782. [Google Scholar] [CrossRef] [PubMed]
  167. Cohen, D.T.; Zhang, C.; Pentelute, B.L.; Buchwald, S.L. An Umpolung Approach for the Chemoselective Arylation of Selenocysteine in Unprotected Peptides. J. Am. Chem. Soc. 2015, 137, 9784–9787. [Google Scholar] [CrossRef] [Green Version]
  168. Chinthakindi, P.K.; Govender, K.B.; Kumar, A.S.; Kruger, H.G.; Govender, T.; Naicker, T.; Arvidsson, P.I. A Synthesis of “Dual Warhead” β-Aryl Ethenesulfonyl Fluorides and One-Pot Reaction to β-Sultams. Org. Lett. 2017, 19, 480–483. [Google Scholar] [CrossRef]
  169. Ortega, V.; del Castillo, E.; Csákÿ, A.G. Transition-Metal-Free Stereocomplementary Cross-Coupling of Diols with Boronic Acids as Nucleophiles. Org. Lett. 2017, 19, 6236–6239. [Google Scholar] [CrossRef]
  170. Armstrong, R.J.; Niwetmarin, W.; Aggarwal, V.K. Synthesis of Functionalized Alkenes by a Transition-Metal-Free Zweifel Coupling. Org. Lett. 2017, 19, 2762–2765. [Google Scholar] [CrossRef]
  171. Sun, H.-B.; Gong, L.; Tian, Y.-B.; Wu, J.-G.; Zhang, X.; Liu, J.; Fu, Z.; Niu, D. Metal- and Base-Free Room-Temperature Amination of Organoboronic Acids with N-Alkyl Hydroxylamines. Angew. Chem. Int. Ed. 2018, 57, 9456–9460. [Google Scholar] [CrossRef] [PubMed]
  172. Nageswar Rao, D.; Rasheed, S.; Vishwakarma, R.A.; Das, P. Copper-catalyzed sequential N-arylation of C-amino-NH-azoles. Chem. Commun. 2014, 50, 12911–12914. [Google Scholar] [CrossRef] [PubMed]
  173. Selmani, A.; Serpier, F.; Darses, S. From Tetrahydrofurans to Tetrahydrobenzo[d]oxepines via a Regioselective Control of Alkyne Insertion in Rhodium-Catalyzed Arylative Cyclization. J. Org. Chem. 2019, 84, 4566–4574. [Google Scholar] [CrossRef] [PubMed]
  174. Faulkner, A.; Scott, J.S.; Bower, J.F. An Umpolung Approach to Alkene Carboamination: Palladium Catalyzed 1,2-Amino-Acylation, -Carboxylation, -Arylation, -Vinylation, and -Alkynylation. J. Am. Chem. Soc. 2015, 137, 7224–7230. [Google Scholar] [CrossRef] [PubMed]
  175. Hazelden, I.R.; Carmona, R.C.; Langer, T.; Pringle, P.G.; Bower, J.F. Pyrrolidines and Piperidines by Ligand-Enabled Aza-Heck Cyclizations and Cascades of N-(Pentafluorobenzoyloxy) carbamates. Angew. Chem. Int. Ed. 2018, 57, 5124–5128. [Google Scholar] [CrossRef] [PubMed]
  176. Yang, H.-B.; Pathipati, S.R.; Selander, N. Nickel-Catalyzed 1,2-Aminoarylation of Oxime Ester-Tethered Alkenes with Boronic Acids. ACS Catal. 2017, 7, 8441–8445. [Google Scholar] [CrossRef]
  177. Zhou, Y.; Rao, C.; Song, Q. Z-Selective Synthesis of γ,δ-Unsaturated Ketones via Pd-Catalyzed Ring Opening of 2-Alkylenecyclobutanones with Arylboronic Acids. Org. Lett. 2016, 18, 4000–4003. [Google Scholar] [CrossRef] [PubMed]
  178. Diehl, A.M.; Ouadoudi, O.; Andreadou, E.; Manolikakes, G. Sulfonamides as Amine Component in the Petasis-Borono Mannich Reaction: A Concise Synthesis of α-Aryl-and α-Alkenylglycine Derivatives. Synthesis 2018, 50, 3936–3946. [Google Scholar]
  179. Deeming, A.S.; Russell, C.J.; Willis, M.C. Palladium (II)-Catalyzed Synthesis of Sulfinates from Boronic Acids and DABSO: A Redox-Neutral, Phosphine-Free Transformation. Angew. Chem. Int. Ed. 2016, 55, 747–750. [Google Scholar] [CrossRef]
  180. Vedovato, V.; Talbot, E.P.A.; Willis, M.C. Copper-Catalyzed Synthesis of Activated Sulfonate Esters from Boronic Acids, DABSO, and Pentafluorophenol. Org. Lett. 2018, 20, 5493–5496. [Google Scholar] [CrossRef]
  181. Chen, Y.; Murray, P.R.D.; Davies, A.T.; Willis, M.C. Direct Copper-Catalyzed Three-Component Synthesis of Sulfonamides. J. Am. Chem. Soc. 2018, 140, 8781–8787. [Google Scholar] [CrossRef] [PubMed]
  182. Le Quement, S.T.; Flagstad, T.; Mikkelsen, R.J.T.; Hansen, M.R.; Givskov, M.C.; Nielsen, T.E. Petasis three-component coupling reactions of hydrazides for the synthesis of oxadiazolones and oxazolidinones. Org. Lett. 2012, 14, 640–643. [Google Scholar] [CrossRef] [PubMed]
  183. Flagstad, T.; Petersen, M.T.; Nielsen, T.E. A Four-Component Reaction for the Synthesis of Dioxadiazaborocines. Angew. Chem. Int. Ed. 2015, 54, 8395–8397. [Google Scholar] [CrossRef] [PubMed]
  184. Mizuta, S.; Onomura, O. Diastereoselective addition to N-acyliminium ions with aryl-and alkenyl boronic acids via a Petasis-type reaction. RSC Adv. 2012, 2, 2266–2269. [Google Scholar] [CrossRef]
  185. Panda, S.; Ready, J.M. Palladium Catalyzed Asymmetric Three-Component Coupling of Boronic Esters, Indoles, and Allylic Acetates. J. Am. Chem. Soc. 2017, 139, 6038–6041. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Yamamoto, E.; Hilton, M.J.; Orlandi, M.; Saini, V.; Toste, F.D.; Sigman, M.S. Development and Analysis of a Pd (0)-Catalyzed Enantioselective 1,1-Diarylation of Acrylates Enabled by Chiral Anion Phase Transfer. J. Am. Chem. Soc. 2016, 138, 15877–15880. [Google Scholar] [CrossRef] [PubMed]
  187. Yu, X.Y.; Zhao, Q.Q.; Chen, J.; Chen, J.R.; Xiao, W.J. Copper-Catalyzed Radical Cross-Coupling of Redox-Active Oxime Esters, Styrenes, and Boronic Acids. Angew. Chem. Int. Ed. 2018, 57, 15505–15509. [Google Scholar] [CrossRef]
  188. Lovinger, G.J.; Aparece, M.D.; Morken, J.P. Pd-Catalyzed Conjunctive Cross-Coupling between Grignard-Derived Boron “Ate” Complexes and C(sp2) Halides or Triflates: NaOTf as a Grignard Activator and Halide Scavenger. J. Am. Chem. Soc. 2017, 139, 3153–3160. [Google Scholar] [CrossRef]
  189. Domański, S.; Staszewska-Krajewska, O.; Chaładaj, W. Pd-Catalyzed Carbonylative Carboperfluoroalkylation of Alkynes. Through-Space 13C–19F Coupling as a Probe for Configuration Assignment of Fluoroalkyl-Substituted Olefins. J. Org. Chem. 2017, 82, 7998–8007. [Google Scholar] [CrossRef]
  190. Croix, C.; Prié, G.; Chaulet, C.; Viaud-Massuard, M.-C. Rhodium-Catalyzed 1,4-Addition of Arylboronic Acids to 3-Benzylidene-1H-pyrrolo[2,3-b]pyridin-2(3H)-one Derivatives. J. Org. Chem. 2015, 80, 3264–3269. [Google Scholar] [CrossRef]
  191. Malapit, C.A.; Luvaga, I.K.; Caldwell, D.R.; Schipper, N.K.; Howell, A.R. Rh-Catalyzed Conjugate Addition of Aryl and Alkenyl Boronic Acids to α-Methylene-β-lactones: Stereoselective Synthesis of trans-3,4-Disubstituted β-Lactones. Org. Lett. 2017, 19, 4460–4463. [Google Scholar] [CrossRef] [PubMed]
  192. Matsuura, R.; Jankins, T.C.; Hill, D.E.; Yang, K.S.; Gallego, G.M.; Yang, S.; He, M.; Wang, F.; Marsters, R.P.; McAlpine, I.; et al. Palladium (ii)-catalyzed γ-selective hydroarylation of alkenyl carbonyl compounds with arylboronic acids. Chem. Sci. 2018, 9, 8363–8368. [Google Scholar] [CrossRef] [PubMed]
  193. Chiminazzo, A.; Sperni, L.; Damuzzo, M.; Strukul, G.; Scarso, A. Copper-mediated 1,4-Conjugate Addition of Boronic Acids and Indoles to Vinylidenebisphosphonate leading to gem-Bisphosphonates as Potential Antiresorption Bone Drugs. ChemCatChem 2014, 6, 2712–2718. [Google Scholar] [CrossRef]
  194. Hanna, L.E.; Konev, M.O.; Jarvo, E.R. Nickel-Catalyzed Directed Hydroarylation of Alkynes with Boronic Acids. Eur. J. Org. Chem. 2019, 184–187. [Google Scholar] [CrossRef]
  195. Yin, J.; Mekelburg, T.; Hyland, C. Unusual (Z)-selective palladium (ii)-catalysed addition of aryl boronic acids to vinylaziridines. Org. Biom. Chem. 2014, 12, 9113–9115. [Google Scholar] [CrossRef] [PubMed]
  196. Luan, Y.; Schaus, S.E. Enantioselective Addition of Boronates to o-Quinone Methides Catalyzed by Chiral Biphenols. J. Am. Chem. Soc. 2012, 134, 19965–19968. [Google Scholar] [CrossRef] [PubMed]
  197. Bos, M.; Buttard, F.; Vallée, A.; Riguet, E. Organocatalytic Gram-Scale Synthesis and Alkylation of Heteroaryl and Electron-Rich Aryl α-Substituted γ-Lactones. Synthesis 2019, 51, 3151–3159. [Google Scholar] [CrossRef]
  198. Bos, M.; Riguet, E. Synthesis of Chiral γ-Lactones by One-Pot Sequential Enantioselective Organocatalytic Michael Addition of Boronic Acids and Diastereoselective Intramolecular Passerini Reaction. J. Org. Chem. 2014, 79, 10881–10889. [Google Scholar] [CrossRef] [PubMed]
  199. Huang, H.; Yu, C.; Li, X.; Zhang, Y.; Zhang, Y.; Chen, X.; Mariano, P.S.; Xie, H.; Wang, W. Synthesis of Aldehydes by Organocatalytic Formylation Reactions of Boronic Acids with Glyoxylic Acid. Angew. Chem. Int. Ed. 2017, 56, 8201–8205. [Google Scholar] [CrossRef]
  200. Kuriyama, M.; Ishiyama, N.; Shimazawa, R.; Onomura, O. Palladium-imidazolinium carbene-catalyzed arylation of aldehydes with arylboronic acids in water. Tetrahedron 2010, 66, 6814–6819. [Google Scholar] [CrossRef] [Green Version]
  201. Malapit, C.A.; Caldwell, D.R.; Luvaga, I.K.; Reeves, J.T.; Volchkov, I.; Gonnella, N.C.; Han, Z.S.; Busacca, C.A.; Howell, A.R.; Senanayake, C.H. Rhodium-Catalyzed Addition of Aryl Boronic Acids to 2,2-Disubstituted Malononitriles. Angew. Chem. Int. Ed. 2017, 56, 6999–7002. [Google Scholar] [CrossRef] [PubMed]
  202. Malapit, C.A.; Reeves, J.T.; Busacca, C.A.; Howell, A.R.; Senanayake, C.H. Rhodium-Catalyzed Transnitrilation of Aryl Boronic Acids with Dimethylmalononitrile. Angew. Chem. Int. Ed. 2016, 55, 326–330. [Google Scholar] [CrossRef] [PubMed]
  203. Wang, H.; Zhu, T.-S.; Xu, M.-H. Rhodium-catalyzed enantioselective 1,2-addition of arylboronic acids to heteroaryl α-ketoesters for synthesis of heteroaromatic α-hydroxy esters. Org. Biomol. Chem. 2012, 10, 9158–9164. [Google Scholar] [CrossRef] [PubMed]
  204. Roscales, S.; Csákÿ, A.G. Synthesis of Di (hetero)arylamines from Nitrosoarenes and Boronic Acids: A General, Mild, and Transition-Metal-Free Coupling. Org. Lett. 2018, 20, 1667–1671. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Structures of naturally occurring indole derivatives.
Figure 1. Structures of naturally occurring indole derivatives.
Molecules 24 03523 g001
Scheme 1. Principal approaches for the synthesis of substituted indoles.
Scheme 1. Principal approaches for the synthesis of substituted indoles.
Molecules 24 03523 sch001
Scheme 2. General scheme for the Suzuki-Miyaura reaction.
Scheme 2. General scheme for the Suzuki-Miyaura reaction.
Molecules 24 03523 sch002
Scheme 3. Bromine-lithium exchange reaction for the synthesis of 3-indolylboronic acid S3-2.
Scheme 3. Bromine-lithium exchange reaction for the synthesis of 3-indolylboronic acid S3-2.
Molecules 24 03523 sch003
Scheme 4. The synthesis of indolylboronic acids by means of ortho-metalation.
Scheme 4. The synthesis of indolylboronic acids by means of ortho-metalation.
Molecules 24 03523 sch004
Scheme 5. Carbamate as a traceless directing group for the ortho-lithiation of indoles.
Scheme 5. Carbamate as a traceless directing group for the ortho-lithiation of indoles.
Molecules 24 03523 sch005
Scheme 6. Proposed mechanism for the rhodium-catalyzed borylation of nitriles.
Scheme 6. Proposed mechanism for the rhodium-catalyzed borylation of nitriles.
Molecules 24 03523 sch006
Scheme 7. A mechanistic proposal for transition-metal-free borylation.
Scheme 7. A mechanistic proposal for transition-metal-free borylation.
Molecules 24 03523 sch007
Scheme 8. Common boron sources for the C−H borylation of arenes.
Scheme 8. Common boron sources for the C−H borylation of arenes.
Molecules 24 03523 sch008
Scheme 9. Iridium-catalyzed borylation of indoles using alternative boron sources to HBpin and B2pin2.
Scheme 9. Iridium-catalyzed borylation of indoles using alternative boron sources to HBpin and B2pin2.
Molecules 24 03523 sch009
Scheme 10. Selective C2 borylation utilizing L-shaped ligands.
Scheme 10. Selective C2 borylation utilizing L-shaped ligands.
Molecules 24 03523 sch010
Scheme 11. Aldimine-directed borylation of the protected indolyl aldehyde S11-1.
Scheme 11. Aldimine-directed borylation of the protected indolyl aldehyde S11-1.
Molecules 24 03523 sch011
Scheme 12. Selective C2 borylation of indole using heterogeneous catalysts, including the represented structures of the catalysts.
Scheme 12. Selective C2 borylation of indole using heterogeneous catalysts, including the represented structures of the catalysts.
Molecules 24 03523 sch012
Scheme 13. Mechanochemistry as applied for the iridium-catalyzed borylation of indoles.
Scheme 13. Mechanochemistry as applied for the iridium-catalyzed borylation of indoles.
Molecules 24 03523 sch013
Scheme 14. Cobalt-catalyzed borylations of the indole S14-1.
Scheme 14. Cobalt-catalyzed borylations of the indole S14-1.
Molecules 24 03523 sch014
Scheme 15. Nickel-catalyzed borylation of indoles.
Scheme 15. Nickel-catalyzed borylation of indoles.
Molecules 24 03523 sch015
Scheme 16. Platinum-catalyzed borylation of indoles developed by the Chatani group.
Scheme 16. Platinum-catalyzed borylation of indoles developed by the Chatani group.
Molecules 24 03523 sch016
Scheme 17. Ruthenium-catalyzed C3 borylation with HBpin, including the outlined mechanism.
Scheme 17. Ruthenium-catalyzed C3 borylation with HBpin, including the outlined mechanism.
Molecules 24 03523 sch017
Scheme 18. Stepwise and traceless protecting group protocols that complement C3 borylations of indoles.
Scheme 18. Stepwise and traceless protecting group protocols that complement C3 borylations of indoles.
Molecules 24 03523 sch018
Scheme 19. C3-selective borylation of Boc-protected indoles based on the cheap 1,10-phenanthroline ligand.
Scheme 19. C3-selective borylation of Boc-protected indoles based on the cheap 1,10-phenanthroline ligand.
Molecules 24 03523 sch019
Scheme 20. C6-preferential iridium-catalyzed borylation of 3-substituted indoles.
Scheme 20. C6-preferential iridium-catalyzed borylation of 3-substituted indoles.
Molecules 24 03523 sch020
Scheme 21. Ir-catalyzed silylation/borylation sequence for the selective borylation of the C7 position.
Scheme 21. Ir-catalyzed silylation/borylation sequence for the selective borylation of the C7 position.
Molecules 24 03523 sch021
Scheme 22. One-pot C7-selective borylation of indoles and selected examples.
Scheme 22. One-pot C7-selective borylation of indoles and selected examples.
Molecules 24 03523 sch022
Scheme 23. Selective C7 borylation using the diboronation/protodeboronation sequence developed by Movassaghi.
Scheme 23. Selective C7 borylation using the diboronation/protodeboronation sequence developed by Movassaghi.
Molecules 24 03523 sch023
Scheme 24. Ester-directed iridium-catalyzed borylation of indoles.
Scheme 24. Ester-directed iridium-catalyzed borylation of indoles.
Molecules 24 03523 sch024
Scheme 25. The short total synthesis of 7-hydroxyoxindole-3-acetic acid utilizing a selective diborylation/oxidation sequence.
Scheme 25. The short total synthesis of 7-hydroxyoxindole-3-acetic acid utilizing a selective diborylation/oxidation sequence.
Molecules 24 03523 sch025
Scheme 26. 2,5,7-Selective threefold borylation of 3-substituted indoles.
Scheme 26. 2,5,7-Selective threefold borylation of 3-substituted indoles.
Molecules 24 03523 sch026
Scheme 27. Triborylation/protodeboronation sequence for the synthesis of 5,7-functionalized indoles and the total synthesis of (+)-plakohypaphorine C.
Scheme 27. Triborylation/protodeboronation sequence for the synthesis of 5,7-functionalized indoles and the total synthesis of (+)-plakohypaphorine C.
Molecules 24 03523 sch027
Scheme 28. Borylative cyclization of ortho-substituted phenylisocyanides S28-1 catalyzed by copper(I) acetate.
Scheme 28. Borylative cyclization of ortho-substituted phenylisocyanides S28-1 catalyzed by copper(I) acetate.
Molecules 24 03523 sch028
Scheme 29. Aryldiazonium salts as the precursor of the synthesis of pinacolboronates.
Scheme 29. Aryldiazonium salts as the precursor of the synthesis of pinacolboronates.
Molecules 24 03523 sch029
Scheme 30. Lewis-acid-mediated borylation of indoles by means of pinB-Bpin and BF3∙Et2O.
Scheme 30. Lewis-acid-mediated borylation of indoles by means of pinB-Bpin and BF3∙Et2O.
Molecules 24 03523 sch030
Scheme 31. Frustrated Lewis pair C−H bond activation for the synthesis of indolylboronic acid pinacol esters.
Scheme 31. Frustrated Lewis pair C−H bond activation for the synthesis of indolylboronic acid pinacol esters.
Molecules 24 03523 sch031
Scheme 32. Borylation of arenes and heteroarenes with borenium cations.
Scheme 32. Borylation of arenes and heteroarenes with borenium cations.
Molecules 24 03523 sch032
Scheme 33. The synthesis of indolylboronic acid pinacol ester S33-2 by means of boron tribromide-2,6-lutidine electrophilic borylation.
Scheme 33. The synthesis of indolylboronic acid pinacol ester S33-2 by means of boron tribromide-2,6-lutidine electrophilic borylation.
Molecules 24 03523 sch033
Scheme 34. Radiolabeled synthesis of fluorinated L-tryptophan.
Scheme 34. Radiolabeled synthesis of fluorinated L-tryptophan.
Molecules 24 03523 sch034
Scheme 35. Developed conditions for the methylation of indolylboronic acid pinacol esters.
Scheme 35. Developed conditions for the methylation of indolylboronic acid pinacol esters.
Molecules 24 03523 sch035
Scheme 36. Palladium-catalyzed benzylation of 2-indolylboronic acids S36-1.
Scheme 36. Palladium-catalyzed benzylation of 2-indolylboronic acids S36-1.
Molecules 24 03523 sch036
Scheme 37. The oxidation of indolylboronic acids as a route to the corresponding phenols.
Scheme 37. The oxidation of indolylboronic acids as a route to the corresponding phenols.
Molecules 24 03523 sch037
Scheme 38. An umpolung approach for the chemoselective arylation of unprotected peptides.
Scheme 38. An umpolung approach for the chemoselective arylation of unprotected peptides.
Molecules 24 03523 sch038
Scheme 39. Oxidative boron Heck coupling of indolylboronic acids for the synthesis of sulfonylfluoride and β-sultam.
Scheme 39. Oxidative boron Heck coupling of indolylboronic acids for the synthesis of sulfonylfluoride and β-sultam.
Molecules 24 03523 sch039
Scheme 40. A transition-metal-free diastereoselective reaction between vicinal diols and boronic acid derivatives.
Scheme 40. A transition-metal-free diastereoselective reaction between vicinal diols and boronic acid derivatives.
Molecules 24 03523 sch040
Scheme 41. The Zweifel coupling of 5-indolylboronic pinacol boronic acid esters.
Scheme 41. The Zweifel coupling of 5-indolylboronic pinacol boronic acid esters.
Molecules 24 03523 sch041
Scheme 42. Transition-metal-free amination of indolylboronic acids.
Scheme 42. Transition-metal-free amination of indolylboronic acids.
Molecules 24 03523 sch042
Scheme 43. The formation of 3,4-disubstitutedtetrahydrofurans via the rhodium-catalyzed cyclization of the ether S43-1.
Scheme 43. The formation of 3,4-disubstitutedtetrahydrofurans via the rhodium-catalyzed cyclization of the ether S43-1.
Molecules 24 03523 sch043
Scheme 44. Palladium-catalyzed cyclization via O−N bond activation.
Scheme 44. Palladium-catalyzed cyclization via O−N bond activation.
Molecules 24 03523 sch044
Scheme 45. Nickel-catalyzed cyclization of the activated oximes S45-1 in the presence of indolylboronic acid.
Scheme 45. Nickel-catalyzed cyclization of the activated oximes S45-1 in the presence of indolylboronic acid.
Molecules 24 03523 sch045
Scheme 46. Palladium-catalyzed ring expansion of cyclobutanone en route to Z-disubstituted alkenes.
Scheme 46. Palladium-catalyzed ring expansion of cyclobutanone en route to Z-disubstituted alkenes.
Molecules 24 03523 sch046
Scheme 47. Three-component reaction for the preparation of α-substituted glycines.
Scheme 47. Three-component reaction for the preparation of α-substituted glycines.
Molecules 24 03523 sch047
Scheme 48. The synthesis of sulfones and sulfonate esters by means of DABSO.
Scheme 48. The synthesis of sulfones and sulfonate esters by means of DABSO.
Molecules 24 03523 sch048
Scheme 49. The Petasis reaction of indolylboronic acids.
Scheme 49. The Petasis reaction of indolylboronic acids.
Molecules 24 03523 sch049
Scheme 50. A Petasis-type reaction for the preparation of cis-2-aryl-3-hydroxypiperidines S50-3.
Scheme 50. A Petasis-type reaction for the preparation of cis-2-aryl-3-hydroxypiperidines S50-3.
Molecules 24 03523 sch050
Scheme 51. The procedure for the preparation of C2- and C3-substituted indoles.
Scheme 51. The procedure for the preparation of C2- and C3-substituted indoles.
Molecules 24 03523 sch051
Figure 2. The structures of substances prepared by means of the multicomponent reactions of indolylboronic acids.
Figure 2. The structures of substances prepared by means of the multicomponent reactions of indolylboronic acids.
Molecules 24 03523 g002
Scheme 52. Palladium-catalyzed hydroarylation of vinylaziridines.
Scheme 52. Palladium-catalyzed hydroarylation of vinylaziridines.
Molecules 24 03523 sch052
Scheme 53. Enantioselective addition of indolylboronic acid diethyl esters to o-quinone methide catalyzed by chiral biphenol.
Scheme 53. Enantioselective addition of indolylboronic acid diethyl esters to o-quinone methide catalyzed by chiral biphenol.
Molecules 24 03523 sch053
Scheme 54. Organocatalytic addition of indolylboronic acids and the intramolecular Passerini reaction for the synthesis of cyclic lactones.
Scheme 54. Organocatalytic addition of indolylboronic acids and the intramolecular Passerini reaction for the synthesis of cyclic lactones.
Molecules 24 03523 sch054
Table 1. The preparation of indolylboronates by means of Miyaura borylation.
Table 1. The preparation of indolylboronates by means of Miyaura borylation.
Molecules 24 03523 i001
EntryLGB-reagentConditionsT1-2
1IpinB-Bdan (2 equiv.)Pd2(dba)3 (5 mol %), Cs2CO3, MeOH, reflux Molecules 24 03523 i002
2ClH2B-NiPr2Pd(OAc)2 (2 mol %), XPhos (6 mol %), KI (2 mol %), TEA, EtOAc, 50 °C, 16 h Molecules 24 03523 i003
3BrH-Bpin (2.5 equiv.)CuI (10 mol %), Fe(acac)3 (10 mol %), TMEDA (1.5 equiv.), NaH (1.5 equiv.), -10 °C, 18 h Molecules 24 03523 i004
4I, BrpinB-Bpin[Cu(DPEphos)(DMEGqu)]PF6 (2.5 mol %) DIPEA (1.2 equiv.), pyridine (20 mol %), MeCN/H2O, 12 h blue LED Molecules 24 03523 i005
5BrpinB-Bpin (1.3 equiv.)Co(IMes)2Cl2 (2.5 mol %), KOMe (1.3 equiv.), MTBE, 50 °C, 8 h Molecules 24 03523 i006
6FH-Bpin (2 equiv.)CoCl2 (10 mol %), L1 (15 mol %), Mg (1.5 equiv.), THF, 60 °C, 16 h Molecules 24 03523 i007
7FnepB-Bnep (3 equiv.)Ni(cod)2 (5 mol %), PCy3 (20 mol %), NaOPh (3 equiv.), THF, 110 °C Molecules 24 03523 i008
8CNnepB-Bnep (2 equiv.)[RhCl(cod)]2 (5 mol %), Xantphos (20 mol %), DABCO (1.0 equiv.), toluene, 100 °C, 15 h Molecules 24 03523 i009
9ClpinB-Bpinhν, TMEDA, TBAF, MeCN/H2O, acetone Molecules 24 03523 i010
10BrPhMe2Si-Bpin (1.5 equiv.)KOMe (1.2 equiv.), DME, 30 °C, 1 h Molecules 24 03523 i011
Table 2. Decarboxylative cross-coupling for the synthesis of indolylboronic acids.
Table 2. Decarboxylative cross-coupling for the synthesis of indolylboronic acids.
Molecules 24 03523 i012
EntryRConditionsT2-2
1OHpinB−Bpin (1.5 equiv.), Pd(OAc)2, dppb (10 mol %), TEA, piv2O, dioxane 160 °C, 15 h Molecules 24 03523 i013
2SEtpinB−Bpin (2 equiv.), [Rh(OH)(cod)]2 (5 mol %), P(nBu)3 (50 mol %), KOAc (20 mol %), CPME, 80 °C, 24 h Molecules 24 03523 i014
3ONPhth Molecules 24 03523 i015 Molecules 24 03523 i016
4ONPhthpinB−Bpin (3.5 equiv.), pyridine (1.5 equiv.) Cs2CO3 (0.5 equiv.), EtOAc, 35 °C, 400 nm, 24 h Molecules 24 03523 i017
Table 3. The synthesis of indolylboronic acids or esters via the cyclization of substituted anilines.
Table 3. The synthesis of indolylboronic acids or esters via the cyclization of substituted anilines.
Molecules 24 03523 i018
EntryR1, R2ConditionsT3-2
1H, TsBCl3 (2 equiv.), DCM rt, then pinacol, TEA, rt 1 h36−83%, R3 = Ph, 4-MePh, 4-MeOPh, 3-thienyl, 2-naphthyl, etc.
2R1 = cat, dan
R2 = Me or Bn
B(C6F5)3 (5 mol %)42−60%, R3 = Ph, 4-BrPh, 4-tBuPh, nBu, ferrocenyl
3R1 = cat
R2 = Mbs, Ts
IPrAuTFA (5 mol %), NaTFA (20 mol %), toluene, 80 °C, 20 h then pinacol (3 equiv.), THF, rt, 1 h55−80%, R3 = Ph, 2-thienyl, nBu, cyPr, 2-BrPh, TBDPSOCH2, etc
4R1 = Ts, Ms
R2 = H
pinB-Bpin (2 equiv.), Pd2dba3 (5 mol %), AsPh3 (10 mol %), Cs2CO3, DMA, 60 °C, 30 min.51−79%, R3 = Ph, Bu, BnOCH2, etc
5 Molecules 24 03523 i019
Table 4. Lewis-acid-catalyzed borylation of indoles and electron-rich arenes.
Table 4. Lewis-acid-catalyzed borylation of indoles and electron-rich arenes.
Molecules 24 03523 i020
EntryConditionsT4-2
1H-Bcat (0.55 equiv.)
B(C6F5)3 (5 mol %), C6D6, rt, 2 h
Molecules 24 03523 i021
2H-Bcat (4 equiv.)
B(C6F5)3 (10 mol %), toluene, 120 °C, 72 h
2. pinacol (3 equiv.), TEA, rt, 1 h
Molecules 24 03523 i022
3 Molecules 24 03523 i023 Molecules 24 03523 i024
Table 5. Simple transition-metal-catalyzed modifications of indolylboronic acids.
Table 5. Simple transition-metal-catalyzed modifications of indolylboronic acids.
Molecules 24 03523 i025
EntryT5-1ConditionsR, yield (%)
1 Molecules 24 03523 i026Ph3PAuNTf2 (5 mol %), dimethyl carbonate, μW, H2O (10 equiv.), 90 °C, 1 h−H, 73
2 Molecules 24 03523 i027NaN3 (1.5 equiv.), Cu(OAc)2 (0.1 equiv.), MeOH (0.2 M), 55 °C, 24 h−N3, 51
3 Molecules 24 03523 i028[11C]NH4CN/NH3, PdCl2(PPh3)2, K2CO3, DMF, 100 °C, 5 min11CN, 90 ± 15 (rcy)
Table 6. The halogenation of indolylboronic acids.
Table 6. The halogenation of indolylboronic acids.
Molecules 24 03523 i029
EntryT6-1Halogen source, conditionsT6-2
1 Molecules 24 03523 i030I2 (1 equiv.), KF (3.3 equiv.)
1,4-dioxane (0.1 M), 80 °C, 1 h
Molecules 24 03523 i031
2 Molecules 24 03523 i032(tBuCN)2CuOTf (2 equiv.), [Me3py]PF6 (3 equiv.), AgF (2 equiv.), THF, 50 °C, 18 h Molecules 24 03523 i033
3 Molecules 24 03523 i034[18F]F-, trifluoromethanesulfonate, Cu(OTf)2Py4, DMA, air, 110 °C, 20 min Molecules 24 03523 i035
4 Molecules 24 03523 i036(i) elution of [18F]F- with Et4NHCO3 in nBuOH
(ii) [Cu(OTf)2(py)4], DMA, 110 °C, 20 min, air
Molecules 24 03523 i037
Table 7. Examples of indole trifluoromethylation.
Table 7. Examples of indole trifluoromethylation.
Molecules 24 03523 i038
EntryBX2[CF3], conditionsT7-2
1B(OH)2TMSCF3 (2 equiv.), Cu(OAc)2 (1 equiv.), 1,10-phenanthroline (1.1 equiv.), CsF (2 equiv.), O2, iPrCN, DCE, 4 Å MS, rt Molecules 24 03523 i039
2BpinK[CF3B(OMe)3] (2 equiv.), Cu(OAc)2 (1 equiv.), O2,DMSO, 60 °C Molecules 24 03523 i040
3B(OH)2(trifluoromethyl)dibenzothiophenium triflate, CuOAc (20 mol %), 2,4,6-trimethylpyridine (2 equiv.), DMAC, 0 °C Molecules 24 03523 i041
4B(OH)2Togni’s reagent, CuI (2 mol %), phenanthroline, K2CO3 (2 equiv.), 35 °C, DME Molecules 24 03523 i042
5BF3KNaSO2CF3 (3 equiv.), TBHP (5 equiv.), CuCl (1 equiv.), DCM/MeOH/H2O, rt Molecules 24 03523 i043
Table 8. The benzylation of indolylboronic acids.
Table 8. The benzylation of indolylboronic acids.
Molecules 24 03523 i044
EntryLGBX2ConditionsT8-3
1 Molecules 24 03523 i045B(OH)2PhenNi(OAc)2∙xH2O (5 mol %), K3PO4 (3.4 equiv.), EtOH (5 equiv), dioxane (0.5 M), 60 °C, 5 h Molecules 24 03523 i046
2OMeBnep Molecules 24 03523 i047 Molecules 24 03523 i048
Table 9. Carbonylation of indolylboronic acids.
Table 9. Carbonylation of indolylboronic acids.
Molecules 24 03523 i049
EntryBX2ReagentConditionsT9-2
1BpinBrF2CCONEt2 (2 equiv.), COPd(PPh)4 (5 mol %), Xantphos (10 mol %), K2CO3 (4 equiv.), toluene/H2O (9:1), 80 °C Molecules 24 03523 i050
2B(OH)2 Molecules 24 03523 i051Pd(OAc)2 (3 mol %), PCy3HBF4 (12 mol %), K2CO3, H3BO3, THF, 65 °C, 15 h Molecules 24 03523 i052
3BpinCO (1 atm)Pd(OAc)2/2PPh3 (5 mol%), pbq (1 equiv.), MeOH, rt, 18 h Molecules 24 03523 i053
4B(OH)2ClCO−NMe(OMe)PdCl2(PPh3)2, K3PO4∙H2O, EtOH, 65 °C Molecules 24 03523 i054
Table 10. Palladium- and rhodium-catalyzed arylation of indolylboronic acids.
Table 10. Palladium- and rhodium-catalyzed arylation of indolylboronic acids.
Molecules 24 03523 i055
EntryLGBX2ConditionsT10-3
Molecules 24 03523 i056Bnep[RhCl(cod)] (5 mol %), I(2Ad)∙HCl (20 mol %), NaOEt (2 equiv.), toluene, 130 °C, 20 h Molecules 24 03523 i057
-SMeB(OH)2Pd2(dba)3 (2.5 mol %), TFP (7.5 mol %), CuTC (3 equiv.), THF, 55 °C Molecules 24 03523 i058
Table 11. Transition-metal-catalyzed reactions of indolylboronic acids with alkenes.
Table 11. Transition-metal-catalyzed reactions of indolylboronic acids with alkenes.
Molecules 24 03523 i059
EntryT11-1ConditionsT11-3
1 Molecules 24 03523 i060 Molecules 24 03523 i061 Molecules 24 03523 i062
2 Molecules 24 03523 i063ArB(OH)2 (1.5 equiv.), [Rh(cod)Cl]2 (1 mol %), KOH (1 equiv.), dioxane/H2O (10:1), 60 °C, 1 h Molecules 24 03523 i064
3 Molecules 24 03523 i065ArB(OH)2 (2 equiv.), Pd(OAc)2 (10 mol %), NaF (2 equiv.), TFT, 100 °C, 12 h Molecules 24 03523 i066
4 Molecules 24 03523 i067ArB(OH)2 (1.5 equiv.), Cu(OTf)2 (10 mol %), toluene, 70 °C, 18 h Molecules 24 03523 i068
5 Molecules 24 03523 i069ArB(OH)2 (2.5 equiv.), Ni(cod)2 (10 mol %), TrixiePhos (10 mol %), THF (0.05 M), rt, 24 h Molecules 24 03523 i070

Share and Cite

MDPI and ACS Style

Čubiňák, M.; Edlová, T.; Polák, P.; Tobrman, T. Indolylboronic Acids: Preparation and Applications. Molecules 2019, 24, 3523. https://doi.org/10.3390/molecules24193523

AMA Style

Čubiňák M, Edlová T, Polák P, Tobrman T. Indolylboronic Acids: Preparation and Applications. Molecules. 2019; 24(19):3523. https://doi.org/10.3390/molecules24193523

Chicago/Turabian Style

Čubiňák, Marek, Tereza Edlová, Peter Polák, and Tomáš Tobrman. 2019. "Indolylboronic Acids: Preparation and Applications" Molecules 24, no. 19: 3523. https://doi.org/10.3390/molecules24193523

Article Metrics

Back to TopTop