Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2022 Oct 13.
Published in final edited form as: J Am Chem Soc. 2021 Sep 28;143(40):16502–16511. doi: 10.1021/jacs.1c05902

Nickel-Catalyzed Dearomative Arylboration of Indoles: Regioselective Synthesis of C2- and C3-Borylated Indolines

Grace L Trammel 1, Rositha Kuniyil 2, Phillip F Crook 3, Peng Liu 4, M Kevin Brown 5
PMCID: PMC8781163  NIHMSID: NIHMS1769211  PMID: 34582691

Abstract

Indole dearomatization is an important strategy to access indolines: a motif present in a variety of natural products and biologically active molecules. Herein, a method for transition-metal catalyzed regioselective dearomative arylboration of indoles to generate diverse indolines is presented. The method accomplishes intermolecular dearomatization of simple indoles through a migratory insertion pathway on substrates that lack activating or directing groups on the C2- or C3-positions. Synthetically useful C2- and C3-borylated indolines can be accessed through a simple change in N-protecting group in high regio- and diastereoselectivities (up to >40:1 rr and >40:1 dr) from readily available starting materials. Additionally, the origin of regioselectivity was explored experimentally and computationally to uncover the remarkable interplay between carbonyl orientation of the N-protecting group on indole, electronics of the C2–C3 π-bond, and sterics. The method enabled the first enantioselective synthesis of (−)-azamedicarpin.

Graphical Abstract

graphic file with name nihms-1769211-f0001.jpg

INTRODUCTION

Indolines are prevalent scaffolds that appear in many natural products and biologically significant molecules. For example, the monoterpenoid indole alkaloids include many examples of highly functionalized indolines.1 Additionally, indolines appear in FDA-approved pharmaceuticals, including naturally derived vinblastine and pentopril (Scheme 1A).2 Due to the ubiquitous nature of this motif, various strategies to access indolines have been developed over the last 70 years.3 Of these, indole dearomatization has emerged as an important strategy for preparing indolines;3 this is, in part, due to the wide availability and synthetic accessibility of indoles. Significant effort has led to the development of a multitude of dearomatization strategies, which include cycloadditions, hydrogenation, radical-mediated processes, or reactions of electrophilic indoles.3 Perhaps the most well-established indole dearomatization strategy exploits the inherent nucleophilicity of indole through electrophilic attack at the C3-position. The resulting iminium species is trapped by a nucleophile to generate the indoline (Scheme 1B).3

Scheme 1.

Scheme 1.

Synthesis of Indolines by Dearomatization

More recent efforts have focused on the dearomatization of indoles through the aid of transition-metal catalysis.46 Many of these processes take advantage of the inherent nucleophilicity of indole to direct the regioselectivity of the dearomatization.4 However, an alternative approach that is not reliant on adding electrophiles to C3 and nucleophiles to C2 involves a migratory insertion of the C2–C3 π-bond of indole into a [M]-R bond to access diverse indoline products (Scheme 1C).5 The intramolecular version of this strategy has been advanced in a series of elegant studies from Jia, Lautens, and others.6 While the selectivity of these reactions is not reliant on inherent reactivity, the tether dictates the regiochemical outcome of the reaction.

Within the field of transition-metal catalyzed indole dearomatization, carboboration variants have also been investigated (Scheme 1D).4gj,km,5,6k,p,7 These processes are particularly valuable, as in addition to generating a new C–C bond, the resulting C–B bond can be transformed to various functional groups through stereospecific transformations.8 One approach to carboboration involves 1,2-metalate rearrangements of boronates induced by various electrophiles (often activated by transition metals).4hj,km While these reactions have been developed with several distinct classes of electrophiles, they do rely on the inherent nucleophilicity of indoles as outlined in Scheme 1B. The migratory insertion approach illustrated in Scheme 1C has also been advanced with carboboration. Ma and Xu6p as well as Jia and Lautens6k have developed intramolecular carboboration reactions. Despite the value of intramolecular indole dearomatization, intermolecular variants that would allow for greater synthetic versatility remain quite limited.9 Ito has shown that substrates with extended conjugation, such as indole-carboxylate esters, undergo Cu-catalyzed carboboration.10 Additionally, Engle has demonstrated that a tethered 8-aminoquinoline directing group can facilitate a Pd-catalyzed dearomative intermolecular arylboration of indole.11 While these examples are important advances in indole dearomatization, the requirement of an activating group (e.g., carboxylate) or a tethered directing group to achieve satisfactory reactivity limits the broad utility of these methods and only allows for one regioisomer to be generated.

Recently, our group has developed a Ni-catalyzed 1,2-arylboration reaction of alkenes.12,13 On the basis of the principles learned, we envisioned advancing this area to the challenging problem of dearomatization. In particular, dearomative arylboration of indoles without the aid of activating or directing groups at C2/3 was undertaken, as it would allow for exploration of unprecedented and challenging reactivity.14 Herein, we present the realization of this goal. Most importantly, since the method is not reliant on substituents at C2/3, both regioisomers of product can be accessed to synthesize valuable C2- and C3-borylated indolines (Scheme 1E).

Mechanistically, we hypothesized that [Ni(I)]-Bpin complex II would form through base-assisted transmetalation with B2pin2 (Scheme 2), followed by syn-addition across the C2–C3 π-bond of indole to generate two possible regioisomers (III or IV).12 Reaction with the arylbromide electrophile would form arylboration product V or VI. A significant challenge in the development of this intermolecular dearomatization could be overcoming the energy of aromaticity, ~18.7 kcal/mol,15 at a rate faster than Miyaura borylation. Additionally, controlling the regioselectivity of the migratory insertion event was identified as an opportunity to access both regioisomers of the product (V or VI).

Scheme 2.

Scheme 2.

Proposed Catalytic Cycle

RESULTS AND DISCUSSION

Preliminary investigation of a variety of N-protected indoles revealed that N-Boc derivative 6 reacted to produce arylboration product in moderate yield and high regio- and diastereoselectivity under Ni-catalyzed arylboration conditions (Scheme 3). The regio- and diastereoisomer of the product derived from 6 was established by X-ray crystallography as the syn-C2-borylated product A. Remarkably, a reversal in regioselectivity was observed when N-Piv-indole 7 was subjected to the reaction conditions, forming C3-borylation product B as a single regioisomer, albeit in low yield. This regiodivergence arising from a simple swap between two similar N-protecting groups was both mechanistically intriguing and synthetically desirable, as both C2- and C3-borylated indolines could be accessed from readily available indoles. It should be noted that while change of the N-protecting group to alter the regiochemical course of indole functionalization reactions is known (e.g., N-Ts vs N-H16a,d, N-CO2Me vs N-Bn,16b N-H vs N-TIPS,16c N-H vs N-Boc16e the surprisingly subtle change (i.e., N-Boc vs N-Piv) demonstrated here has, to our knowledge, no precedent.

Scheme 3. Initial Investigationsa.

Scheme 3.

aReactions run on a 0.2 mmol scale. All yields, rr’s, and dr’s were determined by 1H NMR analysis of the crude reaction mixture relative to an internal standard.

Optimization of the reaction of N-Boc-indole revealed that increased quantities of NaOt-Bu and B2pin2 led to higher yields (Table 1, entry 1). While lower equivalents of PhBr was tolerated, the use of 3.0 eq was chosen for scope investigation when less reactive ArBr were evaluated (Table 1, entry 4). The cosolvent mixture of THF and DMA was crucial to obtain the optimal yields (Table 1, entries 2 and 3). In addition, lower equivalents of NaOt-Bu and B2pin2let to lower yields (Table 1, entry 5). Finally, the catalyst loading could be reduced with minimal impact on yield (Table 1, entries 6 and 7).

Table 1.

Optimization of Reaction Conditionsc

graphic file with name nihms-1769211-t0002.jpg
Entry N-R Change from above conditions % yield A:B
1 N-Boc none 84 16:01
2 N-Boc THF only 28 8:01
3 N-Boc DMA only 69 16:01
4 N-Boc 1.5 equiv PhBr 85 20:01
5 N-Boc 1.5 equiv NaOtBu / 2 equiv B2pin2 71 13:01
6 N-Boc 2 mol% Ni(DME)CI2 82 20:01
7 N-Boc 10 mol% Ni(DME)CI2 79 19:01
8 N-Piv none 31 1:>40
9 N-Cy-acyl none 72 1:4
10 N-Boc-piperidine-acyl none 10 1:9
11 N-Boc-piperidine-acyl 30 °C 31 1:9
12 N-Boc-L-proline none 76a 1:>40
13 N-Boc-L-proline 3 equiv. B 2 pin 2 , pre-stir with NaO t Bu 77 b 1:>40
graphic file with name nihms-1769211-t0003.jpg
c

Reactions run on 0.2 mmol scale. All yields, rr’s, and dr’s were determined by 1H NMR analysis of the crude reaction mixture relative to an internal standard.

a

Product formed in 9:1 dr, 89:11 er, 90% es (major diastereomer).

b

Product formed in 9:1 dr, 98:2 er, 99% es (major diastereomer)

With respect to the C3-borylation process, despite extensive investigations, optimization of the reaction of the N-Piv indole was unsuccessful (Table 1, entry 8). It was found that the yield of the reaction could be increased if N-Cy-acyl-indole was used; however, the regioselectivity was diminished (Table 1, entry 9). While, the increased yield is likely the result of a less sterically demanding N-protecting group, sterics clearly play an important role in regiocontrol. At this stage, the hypothesis was advanced that the more electron-withdrawing N-Piv (vs N-Boc) favors C3 borylation, and that perhaps use of a more electron-withdrawing N-protecting group of similar size to N-Cy-acyl may be optimal. Studies along these lines eventually led to the finding that reaction of N-Boc-piperdine-acyl-indole led to superior regiochemical outcome compared to N-Cy-acyl (Table 1, entries 10–11). Further design of the N-protecting group by “moving” the N-Boc substituent closer to the indole, led to use of N-Boc-L-proline, which resulted in good yield and regioselectivity (Table 1, entry 12). Furthermore, this substrate was particularly attractive because it allowed for control of stereochemistry with respect to the existing stereogenic center (9:1 dr). The major diastereomer was confirmed by X-ray crystal structure analysis (see the SI for details). One final aspect of optimization was needed, as it was found that the N-Boc-L-proline was epimerizing under the reaction conditions. Since epimerization likely occurs due to the presence of NaOtBu prior to complexation with B2pin2, the conditions were modified to prestir B2pin2 with NaOtBu before adding the other reagents. With this modified procedure, epimerization was minimized, forming product in 98:2 er (Table 1, entry 13).

After optimization of the reaction conditions for both the C2- and C3-borylation reactions, the scope was evaluated. With respect to the C2 borylation reaction, various substituents were tolerated in the C4–C7 positions of the N-Boc-indoles, producing diverse indoline products as the syn-diastereomers (Scheme 4). 4-Fluoroindole reacted to obtain product 9 in moderate yield, likely due to unfavorable steric interaction with the nearby fluorine in the migratory insertion step. However, 5-fluoroindoline 10 and chloro-substituted indolines 1113 were formed in good yields and regioselectivities. Additionally, electron-rich 5- and 6-methoxyindoles underwent dearomatization in excellent yields (product 15–16), though 6-methoxyindoline 16 formed in diminished regioselectivity (7:1 rr). This observation can perhaps be attributed to destabilization of the forming benzylic alkyl-[Ni] complex in the transition state by electron donation from the methoxy substituent, resulting in higher rates of formation of the other regioisomer. The reverse effect was observed with an amide-substituent in the C6 position leading to formation of a single observable regioisomer of product 17, likely due to a stabilizing effect on the forming electron-rich alkyl-[Ni] bond. Finally, substitution at the C7-position was tolerated with Cl, Me and amide substituents (product 13, 14, 18), or in the form of a cyclized amide (product 19).

Scheme 4. Reaction Scoped.

Scheme 4.

dYields are reported as the average of two isolated experiments, and products were isolated as single diastereomers. Crude yields, dr’s, and rr’s are reported as the average of two experiments in parentheses and were determined through 1H NMR analysis of the crude reaction mixture relative to an internal standard. a4.5 eq B2pin2, 4 eq NaOt-Bu used. b1.5 eq ArBr used. cReaction run at 30 °C.

Excitingly, 2-substituted N-Boc-indoles underwent efficient dearomatization to produce the opposite regioisomer containing a N-substituted quaternary carbon with a C3-benzylic boronic ester (Scheme 4). 2-Arylindoles, including substrates with electron-neutral (20), electron-deficient (21), and electron-rich (22) aryl substituents were effective in the dearomative arylboration. Additionally, a Bdan substituent was tolerated in the C2-position, resulting in a highly functionalized indoline product (23). The products were formed as single diastereomers in high regioselectivity. It is proposed that the reversal of regioselectivity is the result of addition of the [Ni(I)]-Bpin to the C2–C3 π-bond to minimize steric pressure between the Bpin unit and the C2 substituent. Formation of a tertiary alkyl-[Ni] complex is rare12,17 and this species is likely stabilized by both the α-aryl and α-carbamate groups.12c,18 With respect to the limitations, 3-phenyl indole was unreactive and alkyl substituents were not tolerated at the C2 and C3 positions.

Investigation of the aryl bromide scope of the dearomative arylboration reaction revealed that methoxy- (26) and fluoro- (28) substitutions in the para-position as well as para-aniline derivatives (29 and 32) reacted efficiently (Scheme 4). Additionally, acetal and tertiary amine functionalities were well tolerated (33 and 34). The aryl bromides could also be substituted in the meta-position with oxygen and nitrogen functionalities (24-25, 30-31). Finally, heterocycles such as carbazole (35) and N-Me-indole (36) participated in the arylboration reaction, with no observable dearomatization of the N-Me-indole ring.

However, reactions with electron-deficient and sterically hindered aryl bromides resulted in significant Miyaura borylation and minimal dearomatization. We hypothesized that an alternative N-protecting group could be used to tune the electronics of indole in order to increase the rate of the dearomatization reaction relative to Miyaura borylation. N-Dimethylamide-protected indole was identified to enable reaction with challenging electrophiles, including with electron-deficient aryl bromides (products 37 and 38), sterically hindered 2-bromotoluene (product 39), and a heterocyclic aryl bromide (product 40) (Scheme 4). Notably, all yields and regioselectivities were superior to those observed with N-Boc-indole (see the SI for details). For example, reaction 2-fluorobromobenzene did not lead to desired product (<2% by NMR) with N-Boc-indole. A significant increase in regioselectivity was also observed with 6-methoxy N-dimethylamide indole (product 41, 28:1 rr), in contrast to the lower selectivity observed with N-Boc-6-methoxy indole (product 16, 7:1 rr).

With respect to the C3-borylation process, in addition to simple indole 42, electron-donating methoxy-groups were tolerated in the C5 and C6 positions of the indole ring (products 44 and 47), and halides such as fluorine and chlorine were tolerated in the C5 (46) and C6 (45) positions, respectively (Scheme 4). Additionally, aryl bromide electrophiles with diverse functionality reacted smoothly, including electron-donating methoxy-groups (48 and 50), an electron-withdrawing chlorine (49), and N-containing heterocycles (51 and 52). In all cases the diastereoselectivity with respect to the existing proline stereogenic center was 8:1 or greater. In addition, single diastereomers of product could be easily isolated by silica gel column chromatography. Finally, a model that rationalizes the observed stereochemical outcome is shown in Scheme 4.19 It is proposed that minimization of A(1,3) strain for both amides is the primary driver for the illustrated conformation. In this conformation, the bottom face of the indole is partially blocked by the N-Boc unit, while the top face remains accessible.

To demonstrate the utility of the method, the dearomative arylboration reaction of N-Boc-indoles was amenable to glovebox-free, gram-scale reaction setup using standard Schlenk techniques. The indoline products 8 and 53 were obtained with no appreciable difference in yield from that achieved with glovebox setup on 10 × smaller scale (Scheme 5A). The arylboration products could be further elaborated through functionalization of the Bpin moiety to afford diverse indoline products. Matteson homologation produced products 54 and 59. Metal-free cross-coupling with 2-bromoquinoline delivered diarylated indoline 56, and with benzofuran produced highly substituted indoline 58. Boc-deprotection delivered the N-H indoline 55. Additionally, the benzylic Bpin moiety could be oxidized with sodium perborate to deliver benzylic alcohol 57.

Scheme 5. Synthetic Utility of the Arylboration Reaction.

Scheme 5.

aReaction conditions for Scheme 5A: arylboration: indole (5 mmol, 1 equiv), Ni(DME)Cl2 (0.25 mmol, 5 mol %), ArBr (15 mmol, 3 equiv), B2pin2 (12.5 mmol, 2.5 equiv), NaOt-Bu (10 mmol, 2 equiv), THF/DMA (4:1, 0.1 M),10 °C, 24 h. a) n-BuLi, CH2Br2, THF, −78 °C to −20 °C.b) TFA, DCM, 0 to 25 °C. c) 2-bromoquinoline, n-BuLi, Et2O, −78 °C; Troc-Cl, −78 to 25 °C; NaOH, H2O2, THF, 25 °C. d) NaBO3.H2O, THF/H2O (1:1), 25 °C. e) n-BuLi, benzofuran, THF, −78 °C; NBS. f) n-BuLi, CH2Br2, THF, −78 to 25 °C. bReaction conditions for Scheme 5B: arylboration: indole (2 mmol, 1 equiv), Ni(DME)Cl2 (0.1 mmol, 5 mol %), ArBr (6 mmol, 3 equiv), B2pin2 (6 mmol, 3 equiv), NaOt-Bu (4 mmol, 2 equiv), THF/DMA (9:1, 0.1 M), 30 °C, 18 h. g) n-BuLi, CH2Br2, THF, −78 to 25 °C. h) NaBO3.H2O, THF/H2O (1:1), 25 °C. (i) NaH, DMF, 25 °C. j) NaOMe, MeOH/MeCN, 80 °C. k) Pd(OH)2, H2, EtOAc, 25 °C

We envisioned that the C3-borylative dearomatization reaction would be a useful tool to enable the first enantioselective synthesis of (−)-azamedicarpin (63), a pterocarpan natural product that has previously shown modest activity against leukemia cell lines and inhibition of bacterial and fungal biofilm formation (Scheme 5B).20 Arylboration product 61 derived from 6-methoxy indole 60 and a complex arylbromide was isolated as a single diastereomer. Subsequent Matteson homologation, oxidation, and SNAr led to formation the tetracyclic core. Deprotection of the N-Boc-proline and benzyl ether formed (−)-azamedicarpin (63) in 10% overall yield over six steps.

The origin of regiodivergence in the dearomative arylboration reaction was explored through experimental and computational investigations at the ωB97X-D/SDD–6–311++G(d,p)/SMD(THF)//B3LYP-D3(BJ)/SDD–6–31+G(d,p) level of theory. The computed Gibbs free energy profiles of the reaction with N-Boc- and N-Piv-indoles indicate that the migratory insertion into a Ni(I) boryl species is irreversible and thus determines the regioselectivity (see SI Figures S9S12).12b The migratory insertion in the C2- and C3-borylation pathways with N-Boc- and N-Piv-indoles involve different rotamers of the π-indole nickel complex (C1C4) (Scheme 6A, see SI for other higher-energy transition state conformers). In agreement with experiment, the reaction with N-Boc-indole prefers C2-borylation via TSA from conformer C2 in which the carbonyl is syn-periplanar with the C2-position of the indole (Scheme 6A). The C2-borylation transition state TSA is stabilized because the nucleophilic boryl ligand21 prefers addition to the more electrophilic C2 position, and this process generates a stable benzylic alkyl-[Ni] species.12c The lowest energy pathway for C3-borylation of N-Boc-indole proceeds through TSA1 from a different carbonyl conformer C1 with the carbonyl anti-periplanar with the C2-position of the indole, indicating that carbonyl conformation influences regioselectivity. In the migratory insertion with N-Piv-indole, the computed regioselectivity reverses to favor C3-borylation via TSB1G = 16.7 kcal mol−1), in agreement with experimental observation. The lowest energy pathway for C2-borylation proceeds through the syn-periplanar conformer C4, which is significantly higher in energy due to steric strain between the fused benzene ring and the N-Piv tert-butyl group. The C3-borylation proceeds via the anti-periplanar conformer C3 in which the steric repulsion with the fused benzene ring is diminished.

Scheme 6. Mechanism Studiesa.

Scheme 6.

aArylboration conditions for Scheme 6B and C: indole (0.5 mmol, 1 equiv), Ni(DME)Cl2 (0.025 mmol, 5 mol %), PhBr (1.5 mmol, 3 equiv), B2pin2 (1.25 mmol, 2.5 equiv), NaOt-Bu (1.0 mmol, 2 equiv), THF/DMA (4:1, 0.1 M), 10 °C, 24 h. Values given in red and parentheses are relative Gibbs energies in kcal/mol.

To probe the origins of regiodivergence more deeply, two hypotheses were advanced. The first is that the t-butyl group of N-Piv-indole blocks addition of Bpin to C2, therefore resulting in C3-borylation (Scheme 6B). To explore this hypothesis, N-neopentyl acyl-indole 64 which is iso-steric with N-Boc-indole, was subjected to the reaction. C3-borylation product 65 formed preferentially, albeit in diminished regioselectivity, indicating that sterics are important but not the primary driver for regiodivergence. The second hypothesis is that carbonyl orientation could influence the electronics of the C2–C3 π-bond of indole to change regioselectivity, which is supported by the previously discussed transition state geometries. Conformationally locked N-Piv-indole derivative 66 was designed to constrain the carbonyl in the same syn-periplanar orientation as the reactive conformer of N-Boc-indole (Scheme 6C). Upon subjection to the reaction, the regioselectivity was flipped to C2-borylation (product 67), indicating that carbonyl orientation is a major factor in determining regioselectivity.

Intrigued by the remarkable influence that carbonyl orientation has on the regioselectivity, we calculated CHelpG22 atomic charges for C2 and C3 of both conformers of N-Boc- and N-Piv-indoles (Scheme 6D). For N-Boc-indole, both conformations have comparable energies but different polarities of the C2–C3 π-bond. The syn-periplanar conformer I2 has a more polarized C2–C3 π-bond and prefers C2-borylation due to polarity matching with the nucleophilic boryl ligand at the more electrophilic C2 position. Similarly, the syn-periplanar conformer of the N-Piv-indole (I4) has a more polarized C2–C3 π-bond but is much higher in energy due to steric repulsions. Thus, lower energy conformer I3 is the reactive conformer and undergoes C3-borylation in the lowest energy pathway. The less polarized C2–C3 π-bond in I3, and thus less electrophilic C2 position, combined with a steric interaction between the approaching Bpin moiety and the pivalate tert-butyl group in I3 favors C3-borylation. CHelpG atomic charges of the various N-protected indoles investigated in this study were also calculated (Scheme 6D). Indoles undergoing C2-borylation all have a more polar C2–C3 π-bond (Δe > 0.34), indicating electronic control of regioselectivity. Indoles undergoing C3-borylation have a less polar C2–C3 π-bond (Δe ≤ 0.3) and both electronics of the C2–C3 π-bond as well as sterics appear to play a role in the observed regioselectivity. Calculated C2–C3 π-bond polarizations of various C6-substituted N-Boc-indoles are also linearly correlated with experimental regioselectivities, further supporting the importance of C2–C3 π-bond polarization on regioselectivity.23

CONCLUSION

This work presents an intermolecular transition-metal catalyzed dearomatization of simple indoles proceeding through a migratory insertion pathway that does not require extended conjugation or a tethered directing group on the C2- or C3-positions of indole. The process is regiodivergent due to the remarkable interplay between carbonyl orientation of the N-protecting group, C2–C3 π-bond polarization, and sterics. The C2- and C3-borylated indoline products are synthetically useful intermediates that are easily elaborated to a variety of indolines as demonstrated in the first enantioselective synthesis of (−)-azamedicarpin.

Supplementary Material

Supporting Info

ACKNOWLEDGMENTS

We thank Indiana University and the NIH (R35GM131755 and R35GM128779) for financial support. This project was partially funded by the Vice Provost for Research through the Research Equipment Fund and the NSF MRI program, CHE-1726633 and CHE-1920026. DFT calculations were performed at the Center for Research Computing at the University of Pittsburgh, the TACC Frontera Supercomputer, and the Extreme Science and Engineering Discovery Environment (XSEDE). We thank Dr. Maren Pink and Dr. Veronica Carta of the IU Molecular Structure Center for acquisition of X-ray crystal structure data.

Footnotes

Supporting Information

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.1c05902.

Experimental procedures, analytical data for all new compounds (PDF)

Accession Codes

CCDC 20757552075757 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/data_request/cif, or by emailing data_request@ccdc.cam.ac.uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.

Complete contact information is available at: https://pubs.acs.org/10.1021/jacs.1c05902

The authors declare no competing financial interest.

Contributor Information

Grace L. Trammel, Department of Chemistry, Indiana University, Bloomington, Indiana 47405, United States

Rositha Kuniyil, Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States.

Phillip F. Crook, Department of Chemistry, Indiana University, Bloomington, Indiana 47405, United States

Peng Liu, Department of Chemistry, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, United States;.

M. Kevin Brown, Department of Chemistry, Indiana University, Bloomington, Indiana 47405, United States;.

REFERENCES

  • (1).Saya JM; Ruijter E; Orru RVA Total Synthesis of Aspidosperma and Strychnos Alkaloids through Indole Dearomatization. Chem. - Eur. J 2019, 25, 8916–8935. [DOI] [PubMed] [Google Scholar]
  • (2).Vitaku E; Smith DT; Njardarson JT Analysis of the Structural Diversity, Substitution Patterns, and Frequency of Nitrogen Heterocycles among U.S. FDA Approved Pharmaceuticals. J. Med. Chem 2014, 57, 10257–10274. [DOI] [PubMed] [Google Scholar]
  • (3).For Reviews on indole dearomatization, see:; (a) Cerveri A; Bandini M Recent Advances in the Catalytic Functionalization of “Electrophilic” Indoles. Chin. J. Chem 2020, 38 (3), 287–294. [Google Scholar]; (b) Silva TS; Rodrigues MT; Santos H; Zeoly LA; Almeida WP; Barcelos RC; Gomes RC; Fernandes FS; Coelho F Recent advances in indoline synthesis. Tetrahedron 2019, 75, 2063–2097. [Google Scholar]; (c) Chen J-B; Jia Y-X Recent Progress in Transition-Metal-Catalyzed Enantioselective Indole Functionalizations. Org. Biomol. Chem 2017, 15, 3550–3567.28401233 [Google Scholar]; (d) Roche SP; Tendoung J-JY; Träguier B Advances in dearomatization strategies of indoles. Tetrahedron 2015, 71, 3549–3591. [Google Scholar]; (e) Denizot N; Tomakinian T; Beaud R; Kouklovsky C; Vincent G Synthesis of 3-arylated indolines from dearomatization of indoles. Tetrahedron Lett. 2015, 56, 4413–4429. [Google Scholar]
  • (4).For select examples of transition-metal mediated indole dearomatization using the inherent nucleophilicity of indole, see:; (a) Kieffer ME; Chuang KV; Reisman SE Copper-Catalyzed Diastereoselective Arylation of Tryptophan Derivatives: Total Synthesis of (+)-Naseseazines A and B. J. Am. Chem. Soc 2013, 135, 5557–5560. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Brown S; Clarkson S; Grigg R; Thomas WA; Sridharan V; Wilson DM Tetrahedron 2001, 57, 1347–1359. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Trost BM; Quancard J Palladium-Catalyzed Enantioselective C-3 Allyation of 3-Substituted-1H-Indoles Using Trialkylboranes. J. Am. Chem. Soc 2006, 128, 6314–6315. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Bedford RB; Fey N; Haddow MF; Sankey RF Remarkably reactive dihydroindoloindoles via palladium-catalysed dearomatization. Chem. Commun 2011, 47, 3649–3651. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Kirchberg S; Fröhlich R; Studer A Stereoselective Palladium-Catalyzed Carboaminoxylations of Indoles with Arylboronic Acids and TEMPO. Angew. Chem., Int. Ed 2009, 48, 4235–4238. [DOI] [PMC free article] [PubMed] [Google Scholar]; (f) Ramella V; He Z; Daniliuc CG; Studer A Palladium-Catalyzed Dearomatizing Difunctionalization of Indoles and Benzofurans. Eur. J. Org. Chem 2016, 2016, 2268–2273. [DOI] [PMC free article] [PubMed] [Google Scholar]; (g) Namirembe S; Morken JP Reactions of organoboron compounds enabled by catalyst-promoted metallate shifts. Chem. Soc. Rev 2019, 48, 3464–3474. [DOI] [PMC free article] [PubMed] [Google Scholar]; (h) Ishikura M; Terashima M A new pathway into [b]-annelated indole derivatives through trialkyl(1-methyl-2-indolyl)boronates. J. Chem. Soc., Chem. Commun 1991, 1219–1221. [DOI] [PMC free article] [PubMed] [Google Scholar]; (i) Ishikura M; Agata I A concise access to 2-allenylindole derivatives based on the palladium catalyzed cross-coupling reaction of indoylborates. Heterocycles 1996, 43, 1591–1595. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Ishikura M; Kato H A synthetic use of the intramolecular alkyl migration process in indolylborates for intramolecular cyclization: a novel construction of carbazole derivatives. Tetrahedron 2002, 58, 9827–9838. [DOI] [PMC free article] [PubMed] [Google Scholar]; (j) Panda S; Ready JM Tandem Allylation/1,2-Boronate Rearrangement for the Asymmetric Synthesis of Indolines with Adjacent Quaternary Stereocenters. J. Am. Chem. Soc 2018, 140, 13242–13252. [DOI] [PMC free article] [PubMed] [Google Scholar]; (k) Panda S; Ready JM Palladium Catalyzed Asymmetric Three-Component Coupling of Boronic Esters, Indoles, and Allylic Acetates. J. Am. Chem. Soc 2017, 139, 6038–6041. [DOI] [PMC free article] [PubMed] [Google Scholar]; (l) Das S; Daniliuc CG; Studer A Lewis Acid Catalyzed Stereoselective Dearomative Coupling of Indolylboron Ate Complexes with Donor-Acceptor Cyclopropanes and Alkyl Halides. Angew. Chem., Int. Ed 2018, 57, 4053–4057. [DOI] [PMC free article] [PubMed] [Google Scholar]; (m) Silvi M; Schrof R; Noble A; Aggarwal VK Enantiospecific Three-Component Alkylation of Furan and Indole. Chem. - Eur. J 2018, 24, 4279–4282. [DOI] [PMC free article] [PubMed] [Google Scholar]; (n) Simlandy AK; Brown MK Allenylidene Induced 1,2-Metalate Rearrangement of Indole-Boronates: Diastereoselective Access to Highly Substituted Indolines. Angew. Chem., Int. Ed 2021, 60, 12366–12370. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (5).For a review of transition-metal mediated indole dearomatization via migratory insertion pathways, see:; (a) Zeidan N; Lautens M Migratory Insertion Strategies for Dearomatization. Synthesis 2019, 51, 4137–4146. [Google Scholar]
  • (6).For select examples of transition-metal mediated indole dearomatization via migratory insertion pathways, see:; (a) Zhao L; Li Z; Chang L; Xu J; Yao H; Wu X Efficient Construction of Fused Indolines with a 2-Quaternary Center via an Intramolecular Heck Reaction with a Low Catalyst Loading. Org. Lett 2012, 14, 2066–2069. [DOI] [PubMed] [Google Scholar]; (b) Shen C; Liu R-R; Fan R-J; Li Y-L; Xu T-F; Gao J-R; Jia Y-X Enantioselective Arylative Dearomatization of Indoles via Pd-Catalyzed Intramolecular Reductive Heck Reactions. J. Am. Chem. Soc 2015, 137, 4936–4939. [DOI] [PubMed] [Google Scholar]; (c) Petrone DA; Yen A; Zeidan N; Lautens M Dearomative Indole Bisfunctionalization via a Diastereoselective Palladium-Catalyzed Arylcyanation. Org. Lett 2015, 17, 4838–4841. [DOI] [PubMed] [Google Scholar]; (d) Chen S; Wu X-X; Wang J; Hao X-H; Xia Y; Shen Y; Jing H; Liang Y-M Palladium-Catalyzed Intramolecular Dearomatization of Indoles via Decarboxylative Alkynyl Termination. Org. Lett 2016, 18, 4016–4019. [DOI] [PubMed] [Google Scholar]; (e) Petrone DA; Kondo M; Zeidan N; Lautens M Pd(0)-Catalyzed Dearomative Diarylation of Indoles. Chem. - Eur. J 2016, 22, 5684–5691. [DOI] [PubMed] [Google Scholar]; (f) Liu R-R; Xu T-F; Wang Y-G; Xiang B; Gao J-R; Jia Y-X Palladium-catalyzed dearomative arylalkynylation of indoles. Chem. Commun 2016, 52, 13664–13667. [DOI] [PubMed] [Google Scholar]; (g) Qin X; Lee MWY; Zhou JS Nickel-Catalyzed Asymmetric Reductive Heck Cyclization of Aryl Halides to Afford Indolines. Angew. Chem., Int. Ed 2017, 56, 12723–12726. [DOI] [PubMed] [Google Scholar]; (h) Liu R-R; Wang Y-G; Li Y-L; Huang B-B; Liang R-X; Jia Y-X Enantioselective Dearomative Difunctionalization of Indoles by Palladium-Catalyzed Heck/Sonogoshira Sequence. Angew. Chem., Int. Ed 2017, 56, 7475–7478. [DOI] [PubMed] [Google Scholar]; (i) Li X; Zhou B; Yang R-Z; Yang F-M; Liang R-X; Liu R-R; Jia Y-X Palladium-Catalyzed Enantioselective Intramolecular Dearomative Heck Reaction. J. Am. Chem. Soc 2018, 140, 13945–13951. [DOI] [PubMed] [Google Scholar]; (j) Zeidan N; Beisel T; Ross R; Lautens M Palladium-Catalyzed Arylation/Heteroarylation of Indoles: Access to 2,3-Functionalized Indolines. Org. Lett 2018, 20, 7332–7335. [DOI] [PubMed] [Google Scholar]; (k) Shen C; Zeidan N; Wu Q; Breuers CB-J; Liu R-R; Jia Y-X; Lautens M Pd-catalyzed dearomative arylborylation of indoles. Chem. Sci 2019, 10, 3118–3122. [DOI] [PubMed] [Google Scholar]; (l) Weng J-Q; Xing L-L; Hou W-R; Liang R-X; Jia Y-X Palladium-catalyzed dearomative arylphosphorylation of indoles. Org. Chem. Front 2019, 6, 1577–1580. [DOI] [PubMed] [Google Scholar]; (m) Liang R-X; Wang K; Sheng W-J; Jia Y-X Palladium-Catalyzed Dearomative Arylvinylation Reaction of Indoles with NArylsulfonylhydrazones. Organometallics 2019, 38, 3927–3930. [DOI] [PubMed] [Google Scholar]; (n) Wang H; Wu X-F Palladium-Catalyzed Carbonylative Dearomatization of Indoles. Org. Lett 2019, 21, 5264–5268. [DOI] [PubMed] [Google Scholar]; (o) Marchese AD; Lind F; Mahon AE; Yoon Y; Lautens M Forming Benzylic Iodides via a Nickel Catalyzed Diastereoselective Dearomative Carboiodination Reaction of Indoles. Angew. Chem., Int. Ed 2019, 58, 5095–5099. [DOI] [PubMed] [Google Scholar]; (p) Wei F; Wei L; Zhou L; Tung C-H; Ma Y; Xu Z Divergent Synthesis of 3,3-Disubstituted Oxindoles Initiated by Palladium-Catalyzed Intramolecular Arylation of Un-saturated Amides. Asian J. Org. Chem 2016, 5, 971–975. [DOI] [PubMed] [Google Scholar]; (q) Chen S; Cai M; Huang J; Yao H; Lin A Cobalt-Catalyzed Dearomatization of Indoles via Transfer Hydrogenation To Afford Polycyclic Indolines. Org. Lett 2021, 23, 2212–2216. [DOI] [PubMed] [Google Scholar]; (r) Han M-L; Huang W; Liu Y-W; Liu M; Xu H; Xiong H; Dai H-X Pd-Catalyzed Asymmetric Dearomatization of Indoles via Decarbonylative Heck-Type Reaction of Thioesters. Org. Lett 2021, 23, 172–177. [DOI] [PubMed] [Google Scholar]
  • (7).For reviews regarding carboboration, see:; (a) Semba K; Nakao Y Cross-Coupling Reactions by Cooperative Pd/Cu or Ni/Cu Catalysis Based on the Catalytic Generation of Organocopper Nucleophiles. Tetrahedron 2019, 75 (6), 709–719. [Google Scholar]; (b) Xue W; Oestreich M Beyond Carbon: Enantioselective and Enantiospecific Reactions with Catalytically Generated Boryl- and Silylcopper Intermediates. ACS Cent. Sci 2020, 6 (7), 1070–1081.32724842 [Google Scholar]; (c) Whyte A; Torelli A; Mirabi B; Zhang A; Lautens M Copper-Catalyzed Borylative Difunctionalization of Π-Systems. ACS Catal. 2020, 10 (19), 11578–11622. [Google Scholar]
  • (8).Sandford C; Aggarwal VK Stereospecific functionalizations and transformations of secondary and tertiary boronic esters. Chem. Commun 2017, 53, 5481–5494. [DOI] [PubMed] [Google Scholar]
  • (9).Yang P; Xu R-Q; Zheng C; You S-L Pd-Catalyzed Dearomatization of Indole Derivatives via Intermolecular Heck Reactions. Chin. J. Chem 2020, 38, 235–241. [Google Scholar]
  • (10).Hayama H; Kubota K; Iwamoto H; Ito H Copper(I)-catalyzed Diastereoselective Dearomative Carboborylation of Indoles. Chem. Lett 2017, 46, 1800–1802. [Google Scholar]; For related Cu-catalyzed indole dearomatizations, see:; (a) Kubota K; Hayama K; Iwamoto H; Ito H Enantioselective Borylative Dearomatization of Indoles through Copper(I) Catalysis. Angew. Chem., Int. Ed 2015, 54, 8809–8813. [Google Scholar]; (b) Chen L; Shen J-J; Gao Q; Xu S Synthesis of cyclic chiral, -amino boronates by copper-catalyzed asymmetric dearomative borylation of indoles. Chem. Sci 2018, 9, 5855–5859.30079199 [Google Scholar]; (c) Shi Y; Gao Q; Xu S Copper-Catalyzed Asymmetric Dearomative Silylation of Indoles. J. Org. Chem 2018, 83, 14758–14767.30358397 [Google Scholar]; (d) Ito H Copper-Catalyzed Asymmetric Dearomative Borylation: New Pathway to Optically Active Heterocyclic Compounds Pure. Pure Appl. Chem 2018, 90, 703. [Google Scholar]
  • (11).Liu Z; Chen J; Lu H-X; Li X; Gao Y; Coombs JR; Goldfogel MJ; Engle KM Palladium(0)-Catalyzed Directed syn-1,2-Carboboration and -Silylation: Alkene Scope, Applications in Dearomatization, and Stereocontrol by a Chiral Auxiliary. Angew. Chem., Int. Ed 2019, 58, 17068–17073. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (12).(a) Logan KM; Sardini SR; White SD; Brown MK Nickel-Catalyzed Stereoselective Arylboration of Unactivated Alkenes. J. Am. Chem. Soc 2018, 140, 159–162. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Sardini SR; Lambright AL; Trammel GL; Omer HM; Liu P; Brown MK Ni-Catalyzed Arylboration of Unactivated Alkenes: Scope and Mechanistic Studies. J. Am. Chem. Soc 2019, 141, 9391–9400. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Chen L-A; Lear A; Gao P; Brown MK Nickel-Catalyzed Arylboration of Alkenylarenes: Synthesis of Boron-Substituted Quaternary Carbons and Regiodivergent Reactions. Angew. Chem., Int. Ed 2019, 58, 10956–10960. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Lambright AL; Liu Y; Joyner IA; Logan KM; Brown MK Mechanism-Based Design of an Amide-Directed Ni-Catalyzed Arylboration of Cyclopentene Derivatives. Org. Lett 2021, 23, 612–616. [DOI] [PMC free article] [PubMed] [Google Scholar]; (e) Sardini SR; Brown MK Nickel-Catalyzed Arylboration of Cyclopentene. Org. Synth 2020, 97, 355–367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (13).For selected other examples or Ni-catalyzed 1,2-arylbroation, see:; (a) Wang W; Ding C; Pang H; Yin G Org. Lett 2019, 21, 3968–3971. [DOI] [PubMed] [Google Scholar]; (b) Zhang P; Zou C; Zhao Q; Zhang C Nickel-Catalyzed Alkenylboration of Alkenylarenes to Access Homoallylic Boronic Esters. Org. Chem. Front 2021, 8, 2589–2594. [DOI] [PubMed] [Google Scholar]
  • (14). While Ni-catalyzed dearomatization of indoles has been reported (see ref 6g 6o,), these processes are mechanistically unrelated to the method described herein.
  • (15).Pino-Rios R; Solà M The Relative Stability of Indole Isomers Is a Consequence of the Glidewell-Lloyd Rule. J. Phys. Chem. A 2021, 125, 230–234. [DOI] [PubMed] [Google Scholar]
  • (16).For selected examples of N-protecting group-controlled regioselective functionalization of indoles, see:; (a) García-Rubia A; Arrayás RG; Carretero JC Palladium(II)-Catalyzed Regioselective Direct C2 Alkenylationof Indoles and Pyrroles Assisted by the N-(2-Pyridyl)sulfonyl Protecting Group. Angew. Chem., Int. Ed 2009, 48, 6511–6515. [DOI] [PubMed] [Google Scholar]; (b) Cheng B; Huang G; Xu L; Xia Y Mechanism of the N-protecting group dependent annulations of 3-aryloxy alkynyl indoles under gold catalysis: a computational study. Org. Biomol. Chem 2012, 10, 4417–4423. [DOI] [PubMed] [Google Scholar]; (c) Takagi J; Sato K; Hartwig JF; Ishiyama T; Miyaura N Palladium-catalyzed C-H coupling reaction of heteroaromatic compounds with bis(pinacolato)diboron: regioselective synthesis of heteroarylboronates. Tetrahedron Lett. 2002, 43, 5649–5651. [DOI] [PubMed] [Google Scholar]; (d) 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. [DOI] [PubMed] [Google Scholar]; (e) Kallepalli VA; Shi F; Paul S; Onyeozili EN; Maleczka RE Jr; Smith MR III J. Org. Chem 2009, 74, 9199–9201. [DOI] [PubMed] [Google Scholar]
  • (17).For select examples of tertiary alkyl-Ni complexes, see:; (a) Joshi-Pangu A; Wang C-Y; Bisco MR Nickel-Catalyzed Kumada Cross-Coupling Reactions of Tertiary Alkylmagnesium Halides and Aryl Bromides/Triflates. J. Am. Chem. Soc 2011, 133, 8478–8481. [DOI] [PMC free article] [PubMed] [Google Scholar]; (b) Lohre C; Droge T; Wang C; Glorius F Nickel-Catalyzed Cross-Coupling of Aryl Bromides with Tertiary Grignard Reagents Utilizing Donor-Functionalized N-Heterocyclic Carbenes (NHCs). Chem. - Eur. J 2011, 17, 6052–6055. [DOI] [PMC free article] [PubMed] [Google Scholar]; (c) Ando S; Mawatari M; Matsunaga H; Ishizuka T An N-Heterocyclic Carbene-Based Nickel Catalyst for the Kumada-Tamao-Corriu Coupling of Aryl Bromides and Tertiary Alkyl Grignard Reagents. Tetrahedron Lett. 2016, 57, 3287–3290. [DOI] [PMC free article] [PubMed] [Google Scholar]; (d) Primer DN; Molander GA Enabling the Cross-Coupling of Tertiary Organoboron Nucleophiles Through Radical-Mediated Alkyl Transfer. J. Am. Chem. Soc 2017, 139, 9847–9850. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (18).Min GK; Skrydstrup T Regioselective Rh(I)-Catalyzed Sequential Hydrosilylation toward the Assembly of Silicon-Based Peptidomimetic Analogues. J. Org. Chem 2012, 77, 5894–5906. [DOI] [PubMed] [Google Scholar]
  • (19).Benkovics T; Guzei IA; Yoon TP Oxaziridine-Mediated Oxyamination of Indoles: An Approach to 3-Aminoindoles and Enantiomerically Enriched 3-Aminopyrroloindolines. Angew. Chem., Int. Ed 2010, 49, 9153–9157. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • (20).Dejon L; Mohammed H; Du P; Jacob C; Speicher A Synthesis of Chromenoindole derivatives from Robinia pseudoacacia. MedChemComm 2013, 4, 1580–1583. [Google Scholar]; (b) Kakuda S; Ninomiya M; Tanaka K; Koketsu M Synthesis of Pterocarpan Derivatives and their Inhibitory Effects against Microbial Growth and Biofilms. Chemistry Select 2016, 1, 4203–4208. [Google Scholar]; (c) Mendes JA; Salustiano EJ; Pires CS; Oliveira T; Barcellos JCF; Cifuentes JMC; Costa PRR; Rennó MN; Buarque CD 11a-N-tosyl-5-carbapterocarpans: Synthesis, antineoplastic evaluation and in silico prediction of ADMETox properties. Bioorg. Chem 2018, 80, 585–590.30036814 [Google Scholar]
  • (21).(a) Dang L; Lin Z; Marder TB Boryl ligands and their roles in metal-catalysed borylation reactions. Chem. Commun 2009, 3987–3995. [DOI] [PubMed] [Google Scholar]; (b) Wang Z; Bachman S; Dudnik AS; Fu GC Nickel-Catalyzed Enantioconvergent Borylation of Racemic Secondary Benzylic Electrophiles. Angew. Chem., Int. Ed 2018, 57, 14529–14532. [DOI] [PubMed] [Google Scholar]
  • (22).The electrostatic potential-derived CHelpG atomic charges are more suitable to describe the through-space electrostatic interactions. See:; (a) Breneman CM; Wiberg KB Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. J. Comput. Chem 1990, 11, 361–373. [Google Scholar]; (b) Qi X; Kohler DG; Hull KL; Liu P Energy Decomposition Analyses Reveal the Origins of Catalyst and Nucleophile Effects on Regioselectivity in Nucleopalladation of Alkenes. J. Am. Chem. Soc 2019, 141, 118929–11904. [Google Scholar]
  • (23). The computed NPA atomic charges have better agreement with the experimental regioselectivity of C6-substituted indoles than the CHelpG atomic charges. This suggests the NPA charges are more effective to describe the resonance and inductive effects of the C6-substitutents. See the Supporting Information for detailed discussions.

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supporting Info

RESOURCES