banner
News center
We are happy to fulfill your customized request

Ni

Mar 10, 2024

Nature Communications volume 13, Article number: 2938 (2022) Cite this article

5213 Accesses

11 Citations

Metrics details

Hydroarylation of alkynes with unactivated C(sp2)−H bonds via chelated C−H metalation mainly occurs at γ-position to the coordinating atom of directing groups via stable 5-membered metallacycles, while β-C(sp2)−H bond-involved hydroarylation has been a formidable challenge. Herein, we used a phosphine oxide-ligated Ni−Al bimetallic catalyst to enable β-C−H bond-involved hydroarylations of alkynes via a rare 7-membered nickelacycle.

Hydroarylation of alkynes with arenes provides a highly atom- and step-economical route to aromatic alkenes that widely exist in natural products, bioactive compounds, and material molecules1,2,3,4,5,6,7,8. During the past several decades, a large number of transition metal-catalyzed hydroarylation reactions have been reported. According to electronic nature of arenes, these hydroarylations can be divided into three types (Fig. 1a): electron-rich (hetero)arenes (Type I), electron-deficient (hetero)arenes (Type II), and general arenes with unactivated C(sp2)−H bonds (Type III). Type I reactions mainly proceed via a Friedel-Crafts-type pathway, requiring multiple electron-rich substituents to enhance the electron density of (hetero)arenes9,10,11. This structural requirement results in a limited scope of substrates and difficult site selectivity. Especially in case of less electron-rich arenes, a large excess of substrates is often required for reasonable yields. Instead, most of type II reactions proceed via oxidative addition pathway, because the presence of strong electron-negative heteroatoms or electron-withdrawing groups in substrates leads to uneven distribution of electron density, and then electron-deficient C−H bonds will easily undergo oxidative addition with metal. However, unique electron demand renders substrates limited to special heterocycles such as pyridines, polyfluoroarenes, imidazoles, and other analogs12,13,14,15,16,17,18,19,20,21,22,23,24. To activate arenes with unactivated C(sp2)−H bonds for hydroarylation reactions, chelated C−H metalation has been devised by incorporating proper directing groups in substrates (Type III). With the aid of directing groups, unactivated C(sp2)−H bonds can be metallated, and tuning the size of the formed metallacycles would in principle achieve diverse site selectivities. Owing to these advantages, chelated C−H metalation-involved hydroarylation has been widely explored in the past decades by using various metal catalysts such as Ru25,26,27,28,29,30,31,32, Ir33,34,35,36, Rh37,38,39,40,41,42,43,44,45,46,47, Co48,49,50,51,52,53,54, Mn55,56,57, Re58,59, and Fe60 (Fig. 1b, left). However, most examples are limited to C−H bonds at γ-position to the coordinating atoms of directing groups, because the formation of stable 5-membered metallacycles has more favorable entropic effect and ring strain than other larger (6- or 7-membered) or smaller (4-membered) metallacycles61,62,63,64,65. Only two examples are reported for δ-C−H bond-involved hydroarylations via a 6-membered metallacycle (Fig. 1b, middle)66,67. In contrast, there are no reports on other C−H bond-involved hydroarylations, especially very challenging β-C−H bond-involved hydroarylations, because a highly strained 4-membered metallacycle is difficult to form (Fig. 1b, right). Here, we show that a phosphine oxide (PO)-ligated Ni−Al bimetallic catalyst can activate an unactivated β-C−H bond on the phenyl ring of benzimidazole to undergo hydroarylation via a rare 7-membered nickelacycle, bypassing an unstable 4-membered nickelacycle (Fig. 1c). A series of C4-alkenylated 2-phenylbenzimidazoles, including complex bioactive molecules, can be produced in 41−96% yield, providing a distinctive site selectivity beyond traditional selectivity that generally occurs at C2-phenyl ring via a 5-membered metallacycle.

a Three types of hydroarylations of alkynes and (hetero)arenes: electron-rich arenes (type I), electron-deficient arenes (type II), and general arenes with unactivated C(sp2)−H bonds (type III). b Stability of metallacycles and reported various metallacycle-involved hydroarylations: γ-C−H bond-activation-involved hydroarylation (widely explored); δ-C−H bond-activation-involved hydroarylation (rarely explored); β-C−H bond-activation-involved hydroarylation (elusive challenge); c β-C−H bond-involved hydroarylation via 7-membered metallacycles (this work). T.M. = transition metal. Het = heterocycle, D = electron-donating group. W = electron-withdrawing group. DG = directing group. PO = phosphine oxide ligand.

2-Phenyl benzimidazole (1a) bearing both β-C−H bond and γ-C−H bond was selected as a model substrate, because such a structural motif can be found in a large number of bioactive and material molecules68,69. However, due to strong directing ability of imine N atom, transition metal-catalyzed selective C−H activation of 2-phenyl benzimidazoles mainly occurs at γ-C−H bond at ortho-position of C2-phenyl ring38,70,71,72,73,74,75,76. For example, with RhCl(PPh3)3 as a catalyst, the reaction of 1a and oct-4-yne (2a) afforded 3a′ as the sole product in 37% yield with a E:Z ratio of 6.4:1 (Fig. 2, entry 1). Similarly, the use of [Ru(p-cymene)Cl2]2 as a catalyst also generated 3a′ in 32% yield (entry 2). Given important bioactivity of C4-alkenylated benzimidazoles and derivatives77,78, reversing traditional γ-C−H bond activation to β-C−H bond activation in the hydroarylation reaction (product 3a) would be highly desirable. We envisioned to use a ligand-ligated Ni−Al bimetallic catalyst for the investigation:79,80,81 firstly, the coordination of Al-Lewis acid with benzimidazole could direct Ni to activate β-C−H bonds via a 7-membered nickelacycle, bypassing a highly strained 4-membered nickelacycle; secondly, the coordination of Al-Lewis acid would favor inhibiting the activation of γ-C−H bonds on the C2-phenyl ring via a 5-membered nickelacycle. Systematic survey on ligands, Lewis acids, solvents, and temperatures revealed the desired C4-alkenylated product 3a can be indeed obtained. Under the optimal conditions: Ni(cod)2 (10 mol%), PO-3 (10 mol%), AlEt3 (40 mol%) in toluene at 80 oC, 3a was produced in 84% yield (entry 3). Control experiments showed that the presence of nickel, Al-Lewis acid, and phosphine oxide ligand is essential to the reaction efficiency, the removal of any of them would inactivate the reaction (entries 4−6). Notably, under the optimal conditions, the use of Rh or Ru instead of Ni led to neither 3a nor 3a′ (entries 7 and 8), suggesting that the PO−Ni−Al bimetallic system is sensitive to transition metals.

Reaction conditions: 1a (0.2 mmol), 2a (0.4 mmol), AlEt3 (1 mol/L in hexane, 0.08 mmol), toluene (0.4 mL) under N2 for 12 h. Yield was determined by 1H NMR. cod = 1,5-cyclooctadiene. Et = ethyl. Cy3P = triisopropylphosphine. BINAP = 2,2’-bis(diphenylphosphino)-1,1’-binaphthalene. IPr = 1,3-bis(2,6-diisopropylphenyl)-2,3-dihydro-1H-imidazole. SIPr = 1,3-bis (2,6-diisopropylphenyl)imidazolidine. IMes = 1,3-dimesityl-2,3-dihydro-1H-imidazole. iBu = isobutyl. tBu = tert-butyl. Mes = 2,4,6-trimethylphenyl. Dipp = 2,6-diisopropylphenyl.

Ligand examination showed that various phosphines (entries 9−11), N-heterocyclic carbenes (entries 12−14), and commonly-used phosphine oxides (entries 15−18) were all ineffective, suggesting that the optimal PO-3 plays a critical role in the reaction. In addition, the selection of Lewis acid proved important. Although AlMe3 gave a slightly lower yield than that of AlEt3, all other Lewis acids with either bulkier steric hindrance or weaker Lewis acidity led to a big decrease in yield (entries 19−22).

With the optimized conditions in hand, various C2-aryl benzimidazoles were investigated first (Fig. 3). Results showed that various substituents on C6-position, including alkyl group (3b and 3c), alkenyl group (3d), aryl group (3e), and heteroaryl group (3f, 3g, and 3h) did not have a strong influence on the reaction efficiency, providing the corresponding products in 76−96% yield. In addition, the investigation on electronic effect proved that either electron-rich group (3i) or electron-deficient group (3j and 3k) were also well-tolerated, delivering 82−86% yield. Given that C7-substituent would have direct influence on the electronic density of the C4−H bond, various substituents with different electronic property were examined. Methyl group provided 52% yield (3l), but electron-deficient fluoro group (3m) and CF3 group (3n) increased yield to 67% and 78%, respectively. Different from C6- and C7-positions, C5-position locates closely to the reaction site and substituents on this position would have detrimental effect to the reactivity owing to steric hindrance. For example, C5-methyl afforded trace amount of products and only smaller C5-F group still led to the corresponding product in 48% yield (3o). Notably, various N1-substituents are far away from reaction site and can be well compatible with the reaction. For example, methyl group (3p) and tert-butyl group (3q) provided the corresponding products in 75% and 78% yield, respectively. When C2-phenyl group was replaced by alkyl group like methyl group (3r) or benzyl group (3s), the reaction still proceeded smoothly at a little elevated temperature and loadings of Al-Lewis acid. Beyond benzimidazoles, other heterocycles such as benzoxazoles (3t, 3u, 3v, 3w) and triazoles (3x and 3y) also proved to be suitable substrates under modified conditions. To compare the effect of C2-aryl ring, we also prepared a wide range of C2-arylated benzimidazoles and checked their reactivity.

Reaction conditions: 1 (0.4 mmol), 2a (0.8 mmol), AlEt3 (1 mol/L in hexane, 0.16 mmol), toluene (0.8 mL) under N2 for 12 h. Yield of isolated products. *100 °C, AlMe3 (1 mol/L in hexane, 0.32 mmol). †PO-3 was replaced by bulkier PO-6 (see the structure in the SI). ‡30 °C. §mono refers to C5-alkenylated product and di refers to C5 and C8-dialkenylated product. nPr = n-propyl.

Results showed that either electron-rich groups (4a to 4d) or electron-deficient groups (4e to 4i) at para position of C2-phenyl ring can be well-tolerated, providing C4-alkenylated products in 56−85% yield without observing C−H activation on the C2-aryl ring. Similar results appeared for substituents at other positions such as ortho- (4j), meta (4k to 4n), and even multiple sites (4o to 4r). These results showed that the current method can provide high β-site selectivity for all examined substrates.

Next, the scope of alkynes was investigated (Fig. 4). Aryl alkynes were in general ineffective, which may be attributed to their big steric hindrance and strong coordinative ability with nickel, while various alkyl alkynes such as ethyl (5a), n-butyl (5b), n-pentyl (5c), n-hexyl (5d) and i-hexyl (5e) were compatible very well, providing the corresponding products in 85−89% yield. When functional groups such as phenyl group (5f) or hydroxyl group (5g) were incorporated into the alkyl chain, no significant loss of yields were observed. In addition, cyclic alkyne was still compatible with the reaction, providing 51% yield (5h). Notably, besides symmetrical alkynes, non-symmetrical alkynes also displayed good reactivity (5i, 5j, 5k), but the regioselectivity was highly depending on steric hindrance of substituents of alkynes. For example, tert-butyl methyl alkyne (5i) and isopropyl methyl alkyne (5j) gave one single regioisomer product, while n-propyl methyl alkyne (5k) delivered a mixture of regioisomers in a ratio of 2:1.

Reaction conditions: 1a (0.4 mmol), 2 (0.8 mmol), AlEt3 (1 mol/L in hexane, 0.16 mmol), toluene (0.8 mL) under N2 for 12 h. Yield of isolated products. Et = ethyl. nBu = n-butyl. nPent = n-pentyl. nHex = n-hexyl. iPr = isopropyl. TBS = tert-butyldimethylsilyl. tBu = tert-butyl.

To demonstrate the utility of the reaction, a gram-scale reaction of the model substrates was conducted, providing the desired product 3a in 81% yield (Fig. 5a). The alkene motif in the product is a versatile synthetic precursor and it can be transformed into alkyl group through hydrogenation (6, quantitative yield), epoxide (7, 60% yield), and ketone (8, 73% yield) through different extent of oxidation. In addition, the selective C−H alkenylation method can be applied to the late-stage derivation of complex molecules (Fig. 5b). Telmisartan ester, a long-acting antihypertensive drug, can be selectively activated at C4−H of benzimidazole, achieving a new telmisartan derivative 9 in 75% yield. Other bioactive molecules such as tocopherol and estrone-derivatives were also well compatible with the reaction, providing the corresponding products 10 in 52% yield and 11 in 62% yield, respectively. Compared with traditional selectivity that dominantly occurs at ortho-C−H bond of C2-phenyl ring of benzimidazoles66,67,68,69, the current method provides a different molecular elaboration.

a Gram-scale reaction and product transformations. b Late-stage selective β-C−H bond alkenylation of bioactive molecules such as Telmisartan ester, tocopherol derivative and estrone derivative. nPr = n-propyl. m-CPBA = m-chloro perbenzoic acid.

To gain insight into this reaction, additional mechanistic experiments were conducted. Deuterium-labeling experiment disclosed that the alkenyl H completely came from aryl H on the C4-position of benzimidazole (Fig. 6a), suggesting a C4−H bond metalation. The determination of kinetic isotopic effect via either intermolecular competitive experiment or parallel experiments revealed significant isotopic effect (Fig. 6b), suggesting that C4−H bond cleavage may be involved into a rate-determining step. On basis of these results, a plausible mechanism was proposed in Fig. 6c. Al-Lewis acid of phosphine oxide-ligated Ni−Al bimetallic catalyst coordinates to N atom of the imidazole first, and then nickel is directed to selectively activate β-C−H bond to form a 7-membered nickelacycle, bypassing the formation of a more challenging 4-membered nickelacycle and a traditional 5-membered metallacycle. Subsequent alkyne insertion and reductive elimination delivered the desired product 3a and regenerated the bimetallic catalyst.

a Deuterium-labeling experiment, showing that C4−D was completely transferred to the alkene of the product. b Kinetic isotopic effect determination: a significant KIE for C4−H bond suggests that the activation of C4−H bond could be a rate-determining step. c Proposed mechanism.

In summary, we have developed an unactivated β-C(sp2)−H bond-involved hydroarylation of alkynes, providing a series of C4-alkenylated 2-phenyl benzimidazoles including bioactive complex molecules in 36−96% yield, overriding traditional γ-selectivity via a stable 5-membered metallacycle. The phosphine oxide-ligated Ni−Al bimetallic catalyst effectively directed Ni to generate β-selectivity via a 7-membered metallacycle, bypassing a highly strained 4-membered nickelacycle. The ligand-ligated bimetallic catalyst provides an efficient tool for site selective C−H bond activation and would find wide applications in other types of reactions in future.

In an argon-filled glove-box, to an oven-dried sealed tube were added Ni(cod)2 (11.0 mg, 0.04 mmol), PO-3 (17.2 mg, 0.04 mmol), toluene (0.8 mL), 1 (0.40 mmol), AlEt3 (1 mol/L in hexane, 160 μL, 0.16 mmol), and 2 (0.80 mmol) in sequence. The tube was then sealed, removed out of the glove-box, and heated at 80 °C with heating mantle as the heat source for 12 h. Then the mixture was cooled to room temperature and concentrated in vacuo. The crude product was purified by flash column chromatography using ethyl acetate/hexane as eluent.

The authors declare that the data supporting the findings of this study are available within the article and its Supplementary Information file. For the experimental procedures and data of NMR see Supplementary Methods in Supplementary Information file. The X-ray crystallographic coordinates for structures reported in this study have been deposited at the Cambridge Crystallographic Data Centre (CCDC), under deposition number CCDC 2101580. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/structures/.

Nevado, C. & Echavarren, A. M. Transition metal-catalyzed hydroarylation of alkynes. Synthesis 2, 167–182 (2005).

Google Scholar

Wang, X., Zhou, L. & Lu, W. Hydroarylation of alkynes via aryl C–H bond cleavage. Curr. Org. Chem. 14, 289–307 (2010).

Article CAS Google Scholar

de Haro, T. & Nevado, C. In Comprehensive Organic Synthesis (ed. Knochel, P.) 2nd edn, 1621 (Elsevier, 2014).

Yamamoto, Y. Synthesis of heterocycles via transition-metal catalyzed hydroarylation of alkynes. Chem. Soc. Rev. 43, 1575–1600 (2014).

Article CAS PubMed Google Scholar

Manikandan, R. & Jeganmohan, M. Recent advances in the ruthenium-catalyzed hydroarylation of alkynes with aromatics: synthesis of trisubstituted alkenes. Org. Biomol. Chem. 13, 10420–10436 (2015).

Article CAS PubMed Google Scholar

Boyarskiy, V. P., Ryabukhin, D. S., Bokach, N. A. & Vasilyev, N. A. Alkenylation of arenes and heteroarenes with alkynes. Chem. Rev. 116, 5894–5986 (2016).

Article CAS PubMed Google Scholar

Biffis, A., Tubaro, C. & Baron, M. Advances in transition-metal-catalysed alkyne hydroarylations. Chem. Rec. 16, 1742–1760 (2016).

Article CAS PubMed Google Scholar

Santhoshkumar, R. & Cheng, C.-H. Hydroarylations by cobalt-catalyzed C–H activation. J. Org. Chem. 14, 2266–2288 (2018).

CAS Google Scholar

Kitamura, T. Transition-metal catalyzed hydroarylation reactions of alkynes through direct functionalization of C–H bonds: a convenient tool for organic synthesis. Eur. J. Org. Chem. 2009, 1111−1125 (2009).

Jia, C. et al. Efficient activation of aromatic C–H bonds for addition to C–C multiple bonds. Science 287, 1992–1995 (2000). 2000.

Article CAS PubMed ADS Google Scholar

Tsukada, N., Mitsuboshi, T., Setoguchi, H. & Inoue, Y. Stereoselective cis-addition of aromatic C-H bonds to alkynes catalyzed by dinuclear palladium complexes. J. Am. Chem. Soc. 125, 12102–12103 (2003).

Article CAS PubMed Google Scholar

Nakao, Y. Hydroarylation of alkynes catalyzed by nickel. Chem. Rec. 11, 242–251 (2011).

Article CAS PubMed Google Scholar

Gao, K. & Yoshikai, N. Low-valent cobalt catalysis: new opportunities for C−H functionalization. Acc. Chem. Res. 47, 1208–1219 (2014).

Article CAS PubMed Google Scholar

Wang, Y.-X. & Ye, M. Recent advances in Ni−Al bimetallic catalysis for unreactive bond transformation. Sci. China Chem. 61, 1004–1013 (2018).

Article CAS Google Scholar

Nakao, Y., Kanyiva, K. S., Oda, S. & Hiyama, T. Hydroheteroarylation of alkynes under mild nickel catalysis. J. Am. Chem. Soc. 128, 8146–8147 (2006).

Article CAS PubMed Google Scholar

Kanyiva, K. S., Nakao, Y. & Hiyama, T. Nickel-catalyzed addition of pyridine-N-oxides across alkynes. Angew. Chem. Int. Ed. 46, 8872–8874 (2007).

Article CAS Google Scholar

Nakao, Y., Kashihara, N., Kanyiva, K. S. & Hiyama, T. Nickel-catalyzed alkenylation and alkylation of fluoroarenes via activation of C-H bond over C-F bond. J. Am. Chem. Soc. 130, 16170–16171 (2008).

Article CAS PubMed Google Scholar

Nakao, Y., Kanyiva, K. S. & Hiyama, T. A strategy for C-H activation of pyridines: direct C-2 selective alkenylation of pyridines by nickel/lewis acid catalysis. J. Am. Chem. Soc. 130, 2448–2449 (2008).

Article CAS PubMed Google Scholar

Nakao, Y., Idei, H., Kanyiva, K. S. & Hiyama, T. Direct alkenylation and alkylation of pyridone derivatives by Ni/AlMe3 catalysis. J. Am. Chem. Soc. 131, 15996–15997 (2009).

Article CAS PubMed Google Scholar

Liu, S., Sawicki, J. & Driver, T. G. Ni-catalyzed alkenylation of triazolopyridines: synthesis of 2,6-disubstituted pyridines. Org. Lett. 14, 3744–3747 (2012).

Article CAS PubMed Google Scholar

Yu, M.-S., Lee, W.-C., Chen, C.-H., Tsai, F.-Y. & Ong, T.-G. Controlled regiodivergent C−H bond activation of imidazo[1,5‑a]pyridine via synergistic cooperation between aluminum and nickel. Org. Lett. 16, 4826–4829 (2014).

Article CAS PubMed Google Scholar

Li, Z. & Duan, W.-L. Palladium-catalyzed C-H alkenylation of arenes with alkynes: stereoselective synthesis of vinyl chlorides via a 1,4-chlorine migration. Angew. Chem. Int. Ed. 57, 16041–16045 (2018).

Article CAS Google Scholar

Yin, G. et al. Ligand-controlled Ni(0)−Al(III) bimetal-catalyzed C3−H alkenylation of 2-pyridones by reversing conventional selectivity. ACS Catal. 11, 4606–4612 (2021).

Article CAS Google Scholar

Zhang, T. et al. A directive Ni catalyst overrides conventional site selectivity in pyridine C–H alkenylation. Nat. Chem. 13, 1207–1213 (2021).

Article CAS PubMed Google Scholar

Kakiuchi, F., Yamamoto, Y., Chatani, N. & Murai, S. Catalytic addition of aromatic C−H bonds to acetylenes. Chem. Lett. 24, 681−682 (1995).

Cheng, K., Yao, B., Zhao, J. & Zhang, Y. RuCl3-catalyzed alkenylation of aromatic C−H bonds with terminal alkynes. Org. Lett. 10, 5309–5312 (2008).

Article CAS PubMed Google Scholar

Hashimoto, Y., Hirano, K., Satoh, T., Kakiuchi, F. & Miura, M. Ruthenium(II)-catalyzed regio- and stereoselective hydroarylation of alkynes via directed C−H functionalization. Org. Lett. 14, 2058–2061 (2012).

Article CAS PubMed Google Scholar

Reddy, M. C. & Jeganmohan, M. Ruthenium-catalyzed highly regio- and stereoselective hydroarylation of aryl carbamates with alkynes via C−H bond activation. Chem. Commun. 49, 481–483 (2013).

Article CAS Google Scholar

Villuendas, P. & Urriolabeitia, E. P. Ru-catalyzed regioselective C−H-hydroarylation of alkynes with benzylthioethers using sulfur as directing group. Org. Lett. 17, 3178–3181 (2015).

Article CAS PubMed Google Scholar

Zhang, W. et al. Highly stereoselective ruthenium(II)-catalyzed direct C2-syn-alkenylation of indoles with alkynes. Org. Lett. 17, 1349–1352 (2015).

Article CAS PubMed Google Scholar

Zhang, J., Shrestha, R., Hartwig, J. F. & Zhao, P. A decarboxylative approach for regioselective hydroarylation of alkynes. Nat. Chem. 8, 1144–1151 (2016).

Article CAS PubMed Google Scholar

Biafora, A., Khan, B. A., Bahri, J., Hewer, J. M. & Goossen, L. J. Doubly regioselective C−H hydroarylation of unsymmetrical alkynes using carboxylates as deciduous directing groups. Org. Lett. 19, 1232–1235 (2017).

Article CAS PubMed Google Scholar

Satoh, T., Nishinaka, Y., Miura, M. & Nomura, M. Iridium-catalyzed regioselective reaction of 1-naphthols with alkynes at the peri-position. Chem. Lett. 1999, 615−616 (1999).

Tsuchikama, K., Kasagawa, M., Hashimoto, Y.-K., Endo, K. & Shibata, T. Cationic iridium–BINAP complex-catalyzed addition of aryl ketones to alkynes and alkenes via directed C–H bond cleavage. J. Organomet. Chem. 693, 3939–3942 (2008).

Article CAS Google Scholar

Takebayashi, S. & Shibata, T. [Ir(cod)2]BARF-catalyzed C−H bond alkenylation and alkylation of ferrocenes. Organometallics 31, 4114–4117 (2012).

Article CAS Google Scholar

Nagamoto, M., Fukuda, J.-i, Hatano, M., Yorimitsu, H. & Nishimura, T. Hydroxoiridium-catalyzed hydroarylation of alkynes and bicycloalkenes with N‑sulfonylbenzamides. Org. Lett. 19, 5952–5955 (2017).

Article CAS PubMed Google Scholar

Aulwurm, U. R., Melchinger, J. U. & Kisch, H. Transition metal complexes of diazenes. 35. Synthesis of l-(arylamino)indoles by rhodium-catalyzed addition of alkynes to l,2-diaryldiazenes. Organometallics 14, 3385–3395 (1995).

Article CAS Google Scholar

Lim, Y.-G., Lee, K.-H., Koo, B. T. & Kang, J.-B. Rhodium(I)-catalyzed ortho-alkenylation of 2-phenylpyridines with alkynes. Tetrahedron Lett. 42, 7609–7612 (2001).

Article CAS Google Scholar

Lim, S.-G., Lee, J. H., Moon, C. W., Hong, J. B. & Jun, C.-H. Rh(I)-catalyzed direct ortho-alkenylation of aromatic ketimines with alkynes and its application to the synthesis of isoquinoline derivatives. Org. Lett. 5, 2759–2761 (2003).

Article CAS PubMed Google Scholar

Shibata, Y., Otake, Y., Hirano, M. & Tanaka, K. Amide-directed alkenylation of sp2 C−H bonds catalyzed by a cationic Rh(I)/BIPHEP complex under mild conditions: dramatic rate acceleration by a 1-pyrrolidinecarbonyl group. Org. Lett. 11, 689–692 (2009).

Article CAS PubMed Google Scholar

Schipper, D. J., Hutchinson, M. & Fagnou, K. Rhodium(III)-catalyzed intermolecular hydroarylation of alkynes. J. Am. Chem. Soc. 132, 6910–6911 (2010).

Article CAS PubMed Google Scholar

Zhao, P., Niu, R., Wang, F., Han, K. & Li, X. Rhodium(III)- and ruthenium(II)-catalyzed olefination of isoquinolones. Org. Lett. 14, 4166–4169 (2012).

Article CAS PubMed Google Scholar

Min, M., Kim, D. & Hong, S. AgSbF6-controlled diastereodivergence in alkyne hydroarylation: facile access to Z- and E-alkenyl arenes. Chem. Commun. 50, 8028–8031 (2014).

Article CAS Google Scholar

Nobushige, K., Hirano, K., Satoh, T. & Miura, M. Rhodium(III)-catalyzed ortho-alkenylation through C−H bond cleavage directed by sulfoxide groups. Org. Lett. 16, 1188–1191 (2014).

Article CAS PubMed Google Scholar

Martínez, Á. M. RhI/RhIII catalyst-controlled divergent aryl/heteroaryl C–H bond functionalization of picolinamides with alkynes. Chem. Sci. 6, 5802–5814 (2015).

Article PubMed PubMed Central CAS Google Scholar

Shibata, K., Natsui, S. & Chatani, N. Rhodium-catalyzed alkenylation of C−H bonds in aromatic amides with alkynes. Org. Lett. 19, 2234–2237 (2017).

Article CAS PubMed Google Scholar

Xu, H.-J. et al. Rh(III)-catalyzed meta-C−H alkenylation with alkynes. J. Am. Chem. Soc. 141, 76–79 (2019).

Article CAS PubMed Google Scholar

Halbritter, G., Knoch, F., Wolski, A. & Kisch, H. Functionalization of aromatic azo compounds by the cobalt-catalyzed, regioselective double addition of tolane: 2,6-distilbenylazobenzenes and 2,3-dihydrocinnolines. Angew. Chem. Int. Ed. 33, 1603–1605 (1994).

Article Google Scholar

Gao, K., Lee, P.-S., Fujita, T. & Yoshikai, N. Cobalt-catalyzed hydroarylation of alkynes through chelation-assisted C−H bond activation. J. Am. Chem. Soc. 132, 12249–12251 (2010).

Article CAS PubMed Google Scholar

Lee, P.-S., Fujita, T. & Yoshikai, N. Cobalt-catalyzed, room-temperature addition of aromatic imines to alkynes via directed C-H bond activation. J. Am. Chem. Soc. 133, 17283–17295 (2011).

Article CAS PubMed Google Scholar

Ding, Z. & Yoshikai, N. Mild and efficient C2-alkenylation of indoles with alkynes catalyzed by a cobalt complex. Angew. Chem. Int. Ed. 51, 4698–4701 (2012).

Article CAS Google Scholar

Fallon, B. J. et al. C−H activation/functionalization catalyzed by simple, well-defined low-valent cobalt complexes. J. Am. Chem. Soc. 137, 2448–2451 (2015).

Article CAS PubMed Google Scholar

Wang, S. et al. Cobalt(III)-catalyzed alkenylation of arenes and 6-arylpurines with terminal alkynes: efficient access to functional dyes. Chem. Commun. 52, 2709–2712 (2016).

Article CAS Google Scholar

Zhou, X. et al. Cp*CoIII-catalyzed branch-selective hydroarylation of alkynes via C−H activation: efficient access to α‑gem-vinylindoles. ACS Catal. 7, 7296–7304 (2017).

Article CAS Google Scholar

Zhou, B., Chen, H. & Wang, C. Mn-catalyzed aromatic C−H alkenylation with terminal alkynes. J. Am. Chem. Soc. 135, 1264–1267 (2013).

Article CAS PubMed Google Scholar

Shi, L., Zhong, X., She, H., Lei, Z. & Li, F. Manganese catalyzed C−H functionalization of indoles with alkynes to synthesize bis/trisubstituted indolylalkenes and carbazoles: the acid is the key to control selectivity. Chem. Commun. 51, 7136–7139 (2015).

Article CAS Google Scholar

Wang, H., Pesciaioli, F., Oliveira, J. C. A., Warratz, S. & Ackermann, L. Synergistic manganese(I) C−H activation catalysis in continuous flow: chemoselective hydroarylation. Angew. Chem. Int. Ed. 56, 15063–15067 (2017).

Article CAS Google Scholar

Kuninobu, Y., Kikuchi, K., Tokunaga, Y., Nishina, Y. & Takai, K. Hydroarylation of acetylenes, acrylates, and isocyanates with heteroaromatic compounds under rhenium catalysis. Tetrahedron 64, 5974–5981 (2008).

Article CAS Google Scholar

Chang, Y.-C., Prakash, S. & Cheng, C.-H. ReI-catalyzed highly regio- and stereoselective C–H addition to terminal and internal alkynes. Org. Chem. Front. 6, 432–436 (2019).

Article CAS Google Scholar

Wong, M. Y., Yamakawa, T. & Yoshikai, N. Iron-catalyzed directed C2-alkylation and alkenylation of indole with vinylarenes and alkynes. Org. Lett. 17, 442–445 (2015).

Article CAS PubMed Google Scholar

Apolinar, O. et al. Sulfonamide directivity enables Ni-catalyzed 1,2-diarylation of diverse alkenyl amines. ACS Catal. 10, 14234–14239 (2020).

Article CAS Google Scholar

Kang, T. et al. Nickel-catalyzed 1,2-carboamination of alkenyl alcohols. J. Am. Chem. Soc. 143, 13962–13970 (2021).

Article CAS PubMed Google Scholar

Patzke, B. & Stanger, A. Synthesis, characterization, and reactions of the new seven-membered nickelacycle (2,2′-bipyridine)-6,7-dihydro-5H-dibenzo[c,e]nickelepine. Organometallics 15, 2633–2639 (1996).

Article CAS Google Scholar

Zhang, W.-B., Yang, X.-T., Ma, J.-B., Su, Z.-M. & Shi, S.-L. Regio- and enantioselective C–H cyclization of pyridines with alkenes enabled by a nickel/N-heterocyclic carbene catalysis. J. Am. Chem. Soc. 141, 5628–5634 (2019).

Article CAS PubMed Google Scholar

Cai, Y., Ye, X., Liu, S. & Shi, S.-L. Nickel/NHC-catalyzed asymmetric C−H alkylation of fluoroarenes with alkenes: synthesis of enantioenriched fluorotetralins. Angew. Chem. Int. Ed. 58, 13433–13437 (2019).

Article CAS Google Scholar

Wang, D. et al. Rhodium-catalysed direct hydroarylation of alkenes and alkynes with phosphines through phosphorous-assisted C−H activation. Nat. Comm. 10, 3539 (2019).

Article ADS CAS Google Scholar

Zhang, Z., Cordier, M., Dixneuf, P. H. & Soulé, J.-F. Late-stage diversification of biarylphosphines through rhodium(I)-catalyzed C−H bond alkenylation with internal alkynes. Org. Lett. 22, 5936–5940 (2020).

Article CAS PubMed Google Scholar

Xing, Y. et al. Chemically stable anion exchange membranes based on C2-protected imidazolium cations for vanadium flow battery. J. Membr. Sci. 618, 118696 (2021).

Article CAS Google Scholar

L’abbate, F. P. et al. Hemozoin inhibiting 2-phenylbenzimidazoles active against malaria parasites. Eur. J. Med. Chem. 159, 243–254 (2018).

Article PubMed PubMed Central CAS Google Scholar

Bellina, F. & Rossi, R. Regioselective functionalization of the imidazole ring via transition metal-catalyzed C-N and C-C bond forming reactions. Adv. Synth. Catal. 352, 1223–1276 (2010).

Article CAS Google Scholar

Tang, Z., Mai, S., Zhou, Y. & Song, Q. Divergent synthesis of α-aryl ketones/esters via rhodium-catalyzed selective deesterification and decarbonylation of diazo compounds. Org. Chem. Front. 5, 2583–2587 (2018).

Article CAS Google Scholar

Dhole, S. & Sun, C.-M. Direct access to dihydrobenzoimidazo[2,1-a]isoquinolines through ruthenium-catalyzed formal [4+2] annulation. Adv. Synth. Catal. 361, 535–541 (2019).

Article CAS Google Scholar

Gui, J. et al. C−H methylation of heteroarenes inspired by radical SAM methyl transferase. J. Am. Chem. Soc. 136, 4853–4856 (2014).

Article CAS PubMed PubMed Central Google Scholar

Gao, G.-L., Yang, C. & Xia, W. Selective C–H trifluoromethylation of benzimidazoles through photoredox catalysis. Chem. Commun. 53, 1041–1044 (2017).

Article CAS Google Scholar

Chatani, N., Fukuyama, T., Kakiuchi, F. & Murai, S. Ru3(CO)12-catalyzed coupling of heteroaromatic C-H/CO/olefins. Regioselective acylation of the imidazole ring. J. Am. Chem. Soc. 118, 493–494 (1996).

Article CAS Google Scholar

Fukuyama, T., Chatani, N., Tatsumi, J., Kakiuchi, F. & Murai, S. Ru3(CO)12-catalyzed site-selective carbonylation reactions at a C–H bond in Aza-heterocycles. J. Am. Chem. Soc. 120, 11522–11523 (1998).

Article CAS Google Scholar

Morningstar, M. L. et al. Synthesis, biological activity, and crystal structure of potent nonnucleoside inhibitors of HIV-1 reverse transcriptase that retain activity against mutant forms of the enzyme. J. Med. Chem. 50, 4003–4015 (2007).

Article CAS PubMed PubMed Central Google Scholar

White, A. W. et al. Resistance-modifying agents. 9. Synthesis and biological properties of benzimidazole inhibitors of the DNA repair enzyme poly(ADP-ribose) polymerase. J. Med. Chem. 43, 4084–4097 (2000).

Article CAS PubMed Google Scholar

Donets, P. A. & Cramer, N. Diaminophosphine oxide ligand enabled asymmetric nickel-catalyzed hydrocarbamoylations of alkenes. J. Am. Chem. Soc. 135, 11772–11775 (2013).

Article CAS PubMed Google Scholar

Liu, Q.-S. et al. Ni–Al bimetallic catalyzed enantioselective cycloaddition of cyclopropyl carboxamide with alkyne. J. Am. Chem. Soc. 139, 18150–18153 (2017).

Article CAS PubMed Google Scholar

Li, J.-F., Luan, Y.-X. & Ye, M. Bimetallic anchoring catalysis for C–H and C–C activation. Sci. China Chem. 64, 1923−1937 (2021).

Download references

This work was supported by the National Natural Science Foundation of China (21871145, M.Y.; 22188101, M.Y.), the Natural Science Foundation of Tianjin of China (19JCZDJC37900, M.Y.), the Haihe Laboratory of Sustainable Chemical Transformations and “Frontiers Science Center for New Organic Matter”, Nankai University (63181206, M.Y.).

State Key Laboratory and Institute of Elemento-Organic Chemistry, College of Chemistry, Nankai University, Tianjin, 300071, China

Shao-Long Qi, Yu-Peng Liu, Yi Li, Yu-Xin Luan & Mengchun Ye

Haihe Laboratory of Sustainable Chemical Transformations, Tianjin, 300192, China

Mengchun Ye

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

S.-L.Q. discovered and developed the reactions. Y.-P.L., Y.L., and Y.-X.L. performed part of synthetic experiments. M.Y. conceived, designed the investigations, and wrote the manuscript. S.-L.Q. wrote the Supplementary Information.

Correspondence to Mengchun Ye.

The authors declare no competing interests.

Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Qi, SL., Liu, YP., Li, Y. et al. Ni-catalyzed hydroarylation of alkynes with unactivated β-C(sp2)−H bonds. Nat Commun 13, 2938 (2022). https://doi.org/10.1038/s41467-022-30367-8

Download citation

Received: 25 November 2021

Accepted: 28 April 2022

Published: 26 May 2022

DOI: https://doi.org/10.1038/s41467-022-30367-8

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Nature Communications (2022)

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.