Download PDF
Review  |  Open Access  |  19 Sep 2023

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Views: 835 |  Downloads: 172 |  Cited:   0
Chem Synth 2023;3:39.
10.20517/cs.2023.20 |  © The Author(s) 2023.
Author Information
Article Notes
Cite This Article

Abstract

Nickel-catalyzed reductive cross-coupling (RCC) reactions using two carbon electrophiles as coupling partners provide one of the most reliable and straightforward protocols for facile construction of valuable C-C bonds in the realm of organic chemistry. In recent years, significant progress has been made in the methodological developments and mechanistic studies of these reactions. This review summarizes four widely accepted mechanisms for RCC reactions that have been proposed by experiments or density functional theory calculations. The major difference between these four types of mechanisms lies in the oxidation state of the active catalyst, the change in the valence of nickel during the catalytic cycle, the involvement of carbon radicals, and the form in which the radicals are present. Herein, we focus on covering representative advancements in experimental and theoretical mechanistic studies, aiming to offer vital mechanistic insights into key intermediates, reaction rates, the activation modes of electrophiles, rate- or selectivity-determining steps, and the origin of the cross-selectivity.

Keywords

Mechanisms, nickel catalysis, reductive cross-coupling, C-C bond formation, carbon radicals

INTRODUCTION

Nickel-catalyzed RCC reaction, also known as cross-electrophile coupling (XEC), offers a powerful strategy for C−C bond formation, which has attracted considerable research interest[1-8]. This protocol selectively joins two commercially available electrophiles under reducing conditions with a wide substrate scope of both coupling partners [Scheme 1A and B]. Typically, bidentate and tridentate nitrogen ligands, such as bipyridine (bpy), bioxazoline (biOx), biimidazole (biIm), bisoxazoline (box), terpyridine (terpy), and pyridine-biscarboxamidine (PyBCam), and bidentate phosphine ligands, such as Xantphos, BINAP, and so on, were usually used in this system [Scheme 1C]. In general, when nitrogen ligands are used, the reaction tends to proceed via a radical pathway, while a closed-shell reaction generally occurs with phosphine ligands. There are various mild reducing conditions available, including metal reductants such as Mn and Zn[9,10] and organic reductants such as B2pin2 and hydrazine[11,12]. In addition, photoredox catalysis[13-17] and electrochemistry[18-23] can also be utilized [Scheme 1D]. This strategy enables the construction of a range of C(sp2)−C(sp2)[24-26], C(sp2)−C(sp3)[27-31], and C(sp3)−C(sp3)[32-34] bonds with high levels of cross-selectivity and stereoselectivity. For example, Kim et al. discovered the first nickel-catalyzed RCC of aryl chlorides with primary alkyl chlorides, utilizing a small amount of iodide or bromide in conjunction with the pyridine-2,6-bis(N-cyanocarboxamidine) (PyBCamCN) ligand and Zn reductants[28].

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Scheme 1. Nickel-catalyzed reductive cross-coupling reactions. (A) Experimental studies; (B) The scope of substrates; (C) Some common ligands; (D) Reducing conditions; (E) Several mechanistic questions.

With the continuous development of RCC reactions, some mechanistic problems have sparked intense research interest of chemists [Scheme 1E]: (1) The mechanism of RCC deviates from traditional cross-coupling (XC) reactions that typically proceed through oxidative addition (OA)/transmetalation (TM)/reductive elimination (RE) process[35,36]; (2) Nickel intermediates likely possess various oxidation states ranging from Ni0 to NiIV; (3) Reaction appears to undergo both two-electron and single-electron redox processes; (4) The radical forms are involved in cage-escaped radicals or solvent caged radicals; (5) The origin of cross-selectivity over homo-selectivity involved in reactions. Until now, considerable achievements have been made in both experimental and computational mechanistic studies to answer these mechanistic problems[37-44]. In this review, we focus primarily on the mechanistic aspects of RCC reactions, with an emphasis on experimental studies, such as stoichiometric, competitive, radical, kinetics, kinetic isotope effect (KIE) experiments, and so on. However, we also take into account theoretical research where appropriate. The aim of this review is to enhance the comprehension of the mechanism and selectivity of these reactions, thereby aiding chemists in designing novel catalytic systems. In the following sections, we will provide an integrated discussion of these mechanisms, including their fundamental steps, catalytic systems, detailed mechanistic studies, and the origin of cross-selectivity.

THE INVESTIGATION OF MECHANISM I

Scheme 2 depicts the scenarios of Mechanism I, which consists of five elementary steps: OA, radical addition, RE, radical formation, and reduction. This mechanism essentially proceeds through a radical chain process and involves a key Ni0/NiII/NiIII catalytic cycle. It is worth emphasizing that radicals are generated at one nickel center, and the resulting cage-escaped radicals subsequently combine with another nickel species prior to the OA step. Additionally, when a photoredox catalyst is employed, the formation of radicals may exhibit variability and typically be independent of the catalytic cycle of nickel. A detailed discussion on this will be presented later. Finally, it should be noted that in the product release process, no reducing agents are involved. Alternatively, there is another possibility that radicals may directly combine with Ni0 to form NiI[45,46] intermediates. Xu et al. has computationally demonstrated that even though the addition of radicals to Ni0 is energetically more favorable than the OA of Ni0 to substrates aryl bromides, OA still occurs due to the low concentration of radicals and the abundance of aryl bromides[47]. Hence, it will not be further elaborated here.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Scheme 2. The catalytic cycle of Mechanism I.

Mechanism I predominantly occurs in the Ni0-catalyzed reductive couplings between C(sp2) and C(sp3) electrophiles in the presence of pyridine-type nitrogen ligands and Zn or Mn reductants. The substrates for this mechanism mainly consist of unactivated simple halogenated hydrocarbons. As shown in Figure 1A, Biswas et al. conducted a series of experimental mechanistic investigations on the Ni0/L1 (L1 = 4,4'-di-tert-butyl-2,2'-bipyridine (dtbpy))-catalyzed RCC of iodobenzene 1a with iodooctadecane 1b, using Mn as a reductant[48]. Compared to the previously published reactions[49,50], several small modifications were made to these reaction conditions.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 1. Experimental mechanistic investigations of Mechanism I. (A) Model reaction used in experimental mechanistic investigations; (B) Competition reaction studies between 1a and 1b; (C) Stoichiometric studies of IM1; (D) Radical clock experiments; (E) Radical lifetime studies. DMA: N,N-Dimethylformamide; TMSCl: Trimethylchlorosilane; ND: not detected; Conversion with respect to the amount of L1Ni0(cod).

According to Mechanism I, the reaction begins with the initial OA to give NiII intermediate. To determine whether aryl or alkyl NiII intermediate was formed initially, competition reaction studies between 1a and 2a were performed [Figure 1B]. The results showed that iodobenzene reacts with Ni0 species 4.7 times faster than iodooctadecane, which preferentially leads to NiII aryl intermediate. Wotal et al. also isolated NiII acyl intermediates and discovered that they can react with a series of carbon electrophiles[51]. In addition, the similar aryl nickel complex [(dtbpy)NiII(o-tolyl)I] was synthesized by Sheng et al.[52]. Subsequently, stoichiometric experiment studies between isolated and characterized NiII aryl intermediates IM1 and 1b were conducted to further verify that IM1 acts as the initial key intermediate of the catalytic cycle [Figure 1C]. Significantly, cross-product 4c can be formed through this process without the Mn reductant, which not only confirms that NiII aryl species is the active intermediate of the reaction but also proves the release of the product does not require reducing agents. Moreover, radical clock experiments implied that the alkyl radical would be formed in the system [Figure 1D]. Furthermore, the findings from radical lifetime studies reveal a positive linear correlation between the 7c/8c ratio and the catalyst concentration [Ni], suggesting the involvement of a radical chain process [Figure 1E]. Additionally, they carried out organometallic experiments and excluded the possibility of an organozinc reagent acting as an intermediate. Based on all the experimental mechanistic investigations, it has been confirmed that this type of transformation [Figure 1A], which is restricted to the coupling of C(sp2) electrophiles with C(sp3) electrophiles, predominantly proceeds through a radical chain pathway.

More importantly, Biswas et al. also elucidated the origin of the cross-selectivity over homo-selectivity[48] [Figure 2]. They found that iodobenzene PhI reacts with Ni0 species faster than iodooctadecane C8H17I to afford NiII aryl intermediate, while L1NiII species may react with C8H17I rather than PhI, resulting in alkyl radicals. This sequential activation mode accounts for the experimentally observed cross-selectivity. Typically, due to the lower stability of aryl radicals compared to alkyl radicals, C8H17I is more prone to generate radicals by reacting with Ni0 species. In contrast, PhI is more likely to undergo OA with Ni0 species, which can be attributed to the favorable π-metal interactions between the substrate and the metal[53]. Several theoretical calculations have confirmed that C(sp2) electrophiles undergo OA more rapidly than C(sp3) electrophiles. For instance, Ren et al. demonstrated that the activation energy barrier of aryl halides is 4.4 kcal/mol lower than alkyl halides[54]. Similarly, Kumar et al. obtained identical results[55].

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 2. The origin of the cross-selectivity.

It is gratifying to note that Ren et al. performed detailed density functional theory (DFT) calculations on NiBr2/4,4'-di-methyl-2,2'-bipyridine-catalyzed RCC of aryl bromide and alkyl bromide[54]. They found that radical addition is the rate-limiting step with the energy barrier of 10.4 kcal/mol, suggesting that the radical chain mechanism is a feasible process. In addition, Wang et al. disclosed this mechanism still operates when aryl iodides and tertiary alkyl halides are used[56]. These two research groups also independently calculated an alternative Mechanism II (discussed below), and both mechanisms are energetically feasible, making it difficult to distinguish between them based solely on calculations. Recently, Ji et al. carried out experimental mechanistic investigations on the Ni(1,2-Dimethoxyethane (DME))Br2/box-catalyzed enantioselective coupling of acid chlorides with α-bromobenzoates, indicating that this system also follows a radical chain mechanism[57]. It is worth noting that some of the recently developed Ni/photoredox[58-62] and Ni/electrochemistry[63,64] dual catalytic RCCs also involve a radical chain process. Despite the changes in reduction conditions, the essence of the Ni-involved catalytic cycle remains unchanged, primarily influencing the formation pathway of the radicals.

Although Weix proposed a possible mechanism and suggested that radicals may be generated with the assistance of NiIX species, no conclusive evidence has been provided to support this claim. Building on previous studies[65-69], there are four different pathways that can be taken, including single electron transfer, either outer-sphere or inner-sphere, two-electron OA, and concerted halogen-atom abstraction, as shown in Figure 3.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 3. Several pathways for radical formation.

Thankfully, Lin et al. used electrochemical methods and DFT calculations to investigate the activation modes of radical formation in the (bpy)Ni-catalyzed system[69]. They ruled out electron transfer and two-electron OA and revealed radical formation occurs through a halogen atom abstraction process via transition state TS1, with the energy barrier of 7.4 kcal/mol [Figure 4]. Note that, unlike the study of Weix, the radical is assisted by NiI aryl intermediates IM4 rather than NiIX in this process. Additionally, Diccianni et al. further confirmed that alkyl radicals are generated in a similar manner in the (Xantphos)Ni-catalyzed system with the aid of kinetic studies and DFT calculations[65].

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 4. Concerted halogen atom abstraction process. Selected bond distances are given in Å.

Of particular note, photoredox catalysts can also aid in the generation of radicals. In a study by Wang et al., they reported a Ni/photoredox-catalyzed enantioconvergent RCC between α-bromophosphates and aryl iodides[60] [Figure 5A]. To determine whether the radicals were generated by the photoredox catalyst or the nickel, they conducted comparison experiments using substrates 3a and 5b under conditions with photoredox and Ni catalysts, respectively [Figure 5B]. The results showed that the photoredox catalyst was responsible for generating the desired product 10c, suggesting that the radicals are formed by the photoredox catalyst in this system. The process of free radical formation is illustrated in Figure 5C. Initially, the photoredox catalyst 4CzIPN is excited by light, and the resulting excited 4CzIPN* subsequently undergoes reductive quenching with Hantzsch ester (HEH), leading to 4CzIPN−• with a strong reducing capacity [4CzIPN/4CzIPN−• = -1.21 vs. saturated calomel electrode (SCE)]. Eventually, the single-electron transfer process occurs between 4CzIPN−• and 5b to release radicals and regenerate photoredox catalyst 4CzIPN. Moreover, they also confirmed that the radical addition step is the enantioselectivity-determining step through DFT calculations. Similar results were also reported in the work by Guo et al. for the Ni/photoredox-catalyzed enantioselective three-component carboarylation of alkenes with tertiary and secondary alkyltrifluoroborates and aryl bromides[70]. It is important to note that the photoredox catalysts assisted generation of radicals occurs independently of the Ni catalyst. Consequently, this catalytic cycle involves a Ni0/NiII/NiIII/NiI/Ni0 sequence, in which the reduction occurs directly from NiI to Ni0, rather than the NiI→NiII→Ni0 pathway proposed by Weix, where NiI reacts with alkyl halides to generate radicals and NiII, followed by the reduction of NiII to Ni0.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 5. Photocatalyst-assisted radical formation. (A) Model reaction of Ni/photoredox-catalyzed enantioconvergent reductive cross-couplings; (B) Competition experiments between 3a and 5b; (C) Photoredox catalyst-mediated radical formation process. SCE: saturated calomel electrode.

THE INVESTIGATION OF MECHANISM II

As shown in Scheme 3, unlike the radical chain process, Mechanism II proceeds through successive OA, reduction, OA, RE, and reduction, and it features a key Ni0/NiII/NiI/NiIII process. Notably, the second OA can occur via a concert two-electron (NiI→NiIII, black line) or stepwise single-electron process (NiI→NiII→NiIII, blue line). When C(sp2) electrophiles are used, the reaction generally undergoes a two-electron OA process. On the other hand, the reaction can generate radicals through a single electron process when the C(sp3) electrophiles are present. The main differences between Mechanism II and Mechanism I can be summarized as follows: (1) Mechanism II involves twice OA and reduction and requires a reduction step before the release of the product, while Mechanism I only experiences once OA and reduction in the whole catalytic cycle; (2) Mechanism I necessarily involves the radicals, and the formation of radicals occurs prior to OA, whereas Mechanism II may or may not involve radicals. Even if radicals are involved, they are of the solvent-caged type and undergo a radical rebound process during the reaction while being cage-escaped radicals in Mechanism I.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Scheme 3. The catalytic cycle of Mechanism II.

Mechanism II is mainly observed in reductive coupling systems involving C(sp2)-C(sp2) electrophiles or C(sp2)-C(sp3) electrophiles, which are catalyzed by nickel/pyridine-type ligands in the presence of metal reductants. Generally, when C(sp3) electrophiles are involved, the system may generate radical intermediates, making it difficult to distinguish from Mechanism I merely from a computational point of view. In such cases, corresponding experimental mechanistic studies are needed to differentiate between them. As early as 2014, Jiang et al. investigated the mechanisms on (dtbpy)Ni/Zn-catalyzed RCC reaction of aryl halides from a theoretical calculation perspective[71]. They confirmed the feasibility of the Ni0/NiII/NiI/NiIII/NiI cycle and identified the second OA is the rate-determining step.

Later, Liu et al. reported a similar reaction of aryl halides with vinyl bromides [Figure 6A] and further explored the mechanism of this reaction by a combination of experimental and DFT calculations, offering a more in-depth understanding[72]. Firstly, they conducted a control experiment using TEMPO (2,2,6,6-Tetramethylpiperidin-1-oxyl) to demonstrate that the reaction does not involve a radical chain process and that the participation of vinyl radicals is less likely. Next, a possible mechanism was determined through calculations, as shown in Figure 6B. The results revealed that the Ni0 catalyst initially undergoes an OA process with vinyl bromides through the concerted two-electron, three-member transition state TS2, resulting in the formation of the vinyl NiII intermediate IM7. This intermediate is subsequently reduced by Zn to form the vinyl NiI intermediate IM8. The second OA involving bromobenzene then occurs via the transition state TS3, generating the NiIII intermediate IM10. This process has an activation energy of 14.2 kcal/mol and constitutes the rate-determining step of the reaction.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 6. Computational mechanistic studies. (A) Model reaction used for DFT calculations; (B) Energy profile for oxidative addition process. Gibbs free energies are given in kcal/mol. Computations at the B3LYP-D3(BJ)/SDD for Ni, 6-311G(d,p) for Br, 6-31G(d,p) level. Selected bond distances are given in Å.

Thereafter, Long et al. reported a similar mechanism for a reductive coupling of 2-Haloanilines[73]. They also discovered that the energy required for single-electron halide abstraction is less favorable than that of the two-electron OA process by 15.4 kcal/mol. This further implies that the formation of aryl radicals in this system is quite challenging. Recently, Long et al. developed a (bpy)Ni/Mn-catalyzed reductive coupling of 2-halophenol derivatives[74]. They successfully synthesized the NiII aryl intermediate through the reaction of Ni0 and aryl halides. Moreover, this intermediate was found to efficiently catalyze the reaction, leading to the desired coupling products. This observation suggests that the OA of Ni0 to aryl halides takes place in the catalytic cycle, and the resulting NiII aryl intermediate might be a key active species. All of the aforementioned findings indicate that the Ni0/NiII/NiI/NiIII process is possible for Ni/(pyridine-type ligands)-catalyzed reductive couplings involving C(sp2) and C(sp2) electrophiles using Zn or Mn reductants.

According to Mechanism II, when involving C(sp3) electrophiles, a stepwise single-electron OA process can occur. Ren et al. revealed that the stepwise single-electron OA process between the NiI aryl intermediate IM11 and CyBr 7b is completed through the successive halogen atom abstraction transition state TS4 and the radical addition transition state TS5[54] [Figure 7]. This process forms the NiIII intermediate IM12, requiring relatively low activation energy barriers of 5.0 and 10.4 kcal/mol, respectively. A similar result was obtained for nickel-catalyzed reductive XC of activated primary amines with aryl halides, as reported by Yue et al.[75].

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 7. Stepwise single-electron oxidative addition process. Computations at the B3LYP/6-31g* level. Selected bond distances are given in Å.

THE INVESTIGATION OF MECHANISM III

In general, a fundamental question in these catalytic systems is identifying the nature of the active catalyst at the outset of the reaction. Mechanism III is distinct from the previously discussed two mechanisms, as it is characterized by the absence of Ni0 throughout the entire process, with NiI species serving as the active catalyst. As illustrated in Scheme 4, starting from the active catalyst LNiIX, the reaction can proceed through two distinct OA pathways. The first possibility involves a bimolecular OA to form LNiIIX(R1), while the second involves a single-molecule process that yields LNiIII(X)2(R1). These two intermediates are then reduced to LNiI(R1), followed by halogen atom abstraction to form R2 radical and LNiIIX(R1), which, upon radical rebound, produce LNiIII(X)(R1)(R2). Finally, RE occurs, leading to the formation of the product and regeneration of the active catalyst. Generally speaking, in the reductive coupling system involving C(sp2) and C(sp3) electrophiles, catalyzed by Ni catalyst and pyridine-type nitrogen ligands, it has been observed that when the substrates change to alkyl halides featuring more complex structures, such as alkenes and heteroatoms, the reaction mechanism tends to shift towards Mechanism III.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Scheme 4. The catalytic cycle of Mechanism III.

This mechanism was corroborated through various experimental mechanistic investigations, including kinetic, spectroscopic, and organometallic studies, conducted by Lin et al.[76]. They used the model reaction depicted in Figure 8A, where aryl bromide 5a and alkyl bromide 8b can be effectively cross-coupled to generate product 12c in the presence of Zn and NiBr2·DME/1,10-phenanthroline. Firstly, they carried out substrate probe experiments to demonstrate the presence of radical intermediates in the system. Subsequently, the kinetic studies were performed, and the reaction order was obtained, as illustrated in Figure 8B. The zero-order dependence on substrates 5a and 8b, together with the first-order dependence on [Ni], and the observation that the reaction rate increases along with increasing agitation rate and Zn loading. Integrating these results suggests that the reduction of Ni by Zn is the rate-determining step of the reaction. After that, they identified NiII intermediates IM13 and IM14 as the catalyst resting state through EPR, 1H NMR, and UV-visible spectroscopy analysis [Figure 8C]. In contrast to the previous system where Ni0 served as the reducing species, the resting state IM13 can only be reduced by Zn to NiI species IM15, which was confirmed by comparing its cyclic voltammetry (CV) and EPR data with those of isolated and X-ray characterized (phen*)NiIBr IM16 (phen* = 2, 9-di-sec-butyl-phenanthroline) [Figure 8D]. Therefore, NiIBr IM15 serves as the starting point for the reaction. Several reports have also demonstrated that the NiIX complex serves as the initial active catalyst in nickel-catalyzed reductive coupling reactions[77-80].

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 8. Experimental mechanistic investigations of Mechanism III. (A) Model reaction used for mechanistic investigations; (B) Kinetic studies of model reaction; (C) Spectroscopic studies; (D) Reduction of NiII by Zn; (E) Competition experiments between 5a and 8b; (F) Stoichiometric experiments; (G) Control experiments. a and b: refer to conversion and yields, respectively; DMA: N, N-dimethylacetamide, DME: 1,2-Dimethoxyethane.

Subsequently, competition experiments were conducted to explore the electrophile activation of substrates [Figure 8E]. The reaction rates of 5a and 8b with IM15 were 1.2 and 0.42 mM/min, respectively, indicating that the OA of NiIBr with PhBr is faster than its reaction with haloalkanes to generate radicals. They speculate that the OA process takes precedence over radical formation. These findings contradict the results proposed by the radical chain mechanism, which suggests that NiIBr preferentially activates alkyl halides to generate radicals, occurring prior to OA. It is important to note that Ting et al. further demonstrated that aryl halides undergo OA with NiI halide complexes through quantitative experiments, in which they obtained structurally characterized bipyridine-ligated NiIII aryl complexes[81]. In addition, Till et al. discovered that (dtbpy)NiIBr can be generated through pulse radiolysis, and it is capable of undergoing OA with aryl halides[82]. Then, Breitenfeld et al. performed stoichiometric experiments to demonstrate that 6ao-TolBr reacts with IM15 via bimolecular OA, leading to IM13 and IM17[83,84] [Figure 8F]. EPR spectroscopy reveals that IM17 can also only be reduced by Zn to form NiI aryl species. Similar structures of (IPr)NiI alkyl and aryl complexes have been characterized by Laskowski et al., and they confirmed that NiI alkyl species can activate alkyl halides via a radical process[85]. In order to identify how NiII aryl intermediate interacts with substrate 8b, an array of control experiments was carried out [Figure 8G] . Stoichiometric experiments have shown that the efficiency of the reaction is low in the absence of Zn, with only 30% of 8b being converted (entry 1). However, with the addition of Zn, the conversion of 8b increases to 100% (entry 2), suggesting that Zn plays a crucial role in promoting the reaction. This finding contrasts with the results reported by Weix, where the reaction could efficiently yield the desired product even without Zn. Moreover, the formation of 14c infers the existence of the radical intermediate that undergoes cyclization and dimerization, while the cross-coupled product 15c is obtained through a radical addition and reduction elimination process. They also found that when the substituent Ar is o-Tol, both 15c and 14c can be obtained (entry 2). When the substituent is Mes, only dimer 14c can be formed (entry 3). This suggests that the larger steric hindrance prevents radical addition to Ni center. Taken together, these results are consistent with Mechanism III.

By comparing the catalytic systems of Weix and Diao, significant differences were found. (1) The types of electrophilic substrates used are different: Weix employed simple linear alkyl halides such as iodobenzene, while Diao used structurally complex alkyl halides substituted with heteroatoms and olefins; (2) The ligands used are different: although both ligands belong to the pyridine-type nitrogen ligands, they have different substitution patterns, one being a bipyridine and the other being a phenanthroline; (3) The reducing agents used are different: Weix used Mn, while Diao used Zn. These differences in catalytic systems are important factors that lead to changes in the reaction mechanism. More importantly, Ju et al. discovered that the use of biOx ligand excludes the reduction step of (biOx)NiII[86]. This can be attributed to the lack of ligand redox activity, resulting in more negative reduction potentials of (biOx)NiII complexes, rendering them unable to be reduced by Zn and Mn. This further highlights the significant impact of ligands on the mechanism[87].

Similar to Weix, Diao found that the cross-selectivity also originates from the different activation sequences of the two electrophiles. Unlike the Weix’s system, which utilizes Ni0 and NiII species, Diao employs NiIBr and NiIPh to activate different electrophiles, respectively. Specifically, the OA is mainly influenced by steric effects, while the formation of radicals via halogen-atom abstraction is related to electronic effects. Therefore, sterically assessable NiIBr preferentially undergoes two-electron OA with PhBr 5a to give C(sp2) NiII species, while electron-rich but sterically hindered NiIPh predominantly activates alkyl bromides 8b via halogen-atom abstraction to forms C(sp3) radicals [Figure 9], thus resulting in cross-selectivity.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 9. The origin of the cross-selectivity.

Meanwhile, by combining experimental and DFT calculations, Shu et al. investigated the mechanism of dipyridine-ligated nickel-catalyzed reductive dicarbofunctionalization of propene with tert-butyl iodide and iodobenzene with the use of Zn reductants[88]. Their findings confirmed the feasibility of the pathway involving NiI species. More recently, Zhu et al. developed an RCC reaction of α-oxy halides enabled by Mn reductants, photocatalysis, electrocatalysis, or mechanochemistry in the presence of nickel and phenanthroline ligands[89]. Surprisingly, through detailed experimental and theoretical studies, they found the mechanisms of all four catalytic systems are consistent with Mechanism III. Besides, they noted that the NiIII intermediate, obtained through OA, may trigger comproportionation with NiI species to afford the NiII intermediate. Intriguingly, Day et al. disclosed that polypyridine-ligated NiII halide complexes can undergo the comproportionation with Ni0 to form NiI species[90]. These electron-transfer events were corroborated by electrochemical techniques and detailed quantum mechanical calculations. It is worth noting that Tang et al. conducted comprehensive mechanistic studies on the OA of Ni(I) to aryl iodides using electroanalytical and statistical modeling techniques[91]. And predicted OA rates can be utilized not only to interpret observed reactivities but also to rationalize the mechanism. This indicates that an expanding array of techniques can assist us in predicting reaction models, understanding reaction mechanisms, and guiding the design of novel catalytic systems.

THE INVESTIGATION OF MECHANISM IV

In contrast to the aforementioned initial three mechanisms, Mechanism IV fundamentally encompasses an SN2 process and the Ni0/NiII catalytic cycle. As depicted in Scheme 5, commencing with Ni0, the reaction proceeds sequentially through OA, TM, an SN2 reaction, a second TM, and ultimately RE. Notably, this particular mechanism employs organic Grignard reagents as reductants, as opposed to their conventional use as coupling reagents, thereby deviating significantly from traditional XC reactions. Mechanism IV is chiefly observed in the intramolecular reductive coupling reaction system, which is facilitated by Ni0 catalysts in the presence of phosphine ligands and organic Grignard reagents. The primary substrates for this mechanism encompass a range of active pyrrole and amine derivatives. It is important to note that Mechanism IV demonstrates fundamental differences in terms of reaction conditions and substrates when contrasted with the previously discussed three mechanisms.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Scheme 5. The catalytic cycle of Mechanism IV.

Chen et al. conducted a comprehensive mechanistic investigation on Ni0/BINAP-catalyzed stereospecific intramolecular RCC reactions of benzylic ethers through a combination of experimental and computational approaches, thereby confirming the feasibility of this particular mechanism[92]. In this discussion, we concentrate on the key transition states for the product formation, as illustrated in Figure 10A. The reaction proceeds sequentially through the OA transition state TS6, TM transition state TS7 and intramolecular SN2-like attack transition state TS8 to achieve C−O bond cleavage, alkyl transfer, and C−Cl bond cleavage, ultimately leading to the product 16c. The energy barriers associated with these steps are 21.9, 12.0, and 9.3 kcal/mol, respectively. This process takes place under a closed-shell system without involving radicals and exhibits stereospecificity. Among these steps, the OA of the C−O bond occurs with stereoinversion of the benzylic stereogenic center, which is facilitated by the assistance of Grignard reagents, constituting the rate- and selectivity-determining step for product formation. Concurrently, they performed a 13C KIE study of the benzylic ether of 2-naphthyltetrahydropyran 9b, which revealed a significant (> 1.01) KIE numerical distribution for benzylic C11 compared to other atoms [Figure 10B]. This further substantiated that the activation of the C−O bond is the rate-determining step.

Mechanisms of nickel-catalyzed reductive cross-coupling reactions

Figure 10. Experimental and computational mechanistic studies. (A) DFT calculations for key transition states; (B) 13C KIE experiments and KIE numerical distribution of 9b. Computations at the B3LYP-D3(BJ)-SMD/def2-TZVPP//B3LYP-D3(BJ)/def2-SVP level. Selected bond distances are given in Å. The hydrogen atom was omitted for simplification.

Moreover, the activation of both the C−O and C−Cl bonds collectively control the stereospecificity of overall reactions.

They also discovered that the intramolecular reductive coupling reaction of halogenated sulfonylamine derivatives, catalyzed by air-stable ((R)-BINAP)NiCl2 and MeMgI organic Grignard reagents, still adheres to the Ni0/NiII catalytic cycle[93,94]. Furthermore, Xu et al. conducted a detailed theoretical study on the Ni0/XantPhos-catalyzed intramolecular reductive coupling of tetrahydropyrans, indicating that Mechanism IV is applicable to this system[95]. Similarly, nickel-catalyzed intramolecular reductive coupling of difluoromethyl moiety and benzylic ether, reported by Lucas et al., also follows a comparable reaction mechanism[96]. It is noteworthy that Sanford et al. found that when utilizing non-cyclic 1,3-diol derivatives featuring two C(sp3)−I bonds for intramolecular reductive coupling, the activation of C(sp3)−I bonds proceeds via a radical pathway rather than the SN2 mechanism[97]. This finding highlights the notion that the intrinsic characteristics of the substrates can indeed exert a substantial impact on the reaction mechanism.

CONCLUSION AND OUTLOOK

In summary, nickel-catalyzed reductive coupling reactions exhibit considerable diversity in their mechanisms, distinct from those catalyzed by palladium and platinum. This can be ascribed to the unique characteristics of nickel, including high paring energy, low electronegativity and redox potential, and multiple oxidation states (0, I, II, III, IV)[35]. These enable nickel catalysts to preferentially undergo both two- and one-electron redox processes, leading to comparatively diverse mechanistic scenarios. Additionally, these mechanisms are closely related to substrates, ligands, and reducing conditions, making it challenging to discern a unifying pattern. From the four potential reaction mechanisms summarized, we can identify some basic trends: (1) C(sp2) electrophiles tend to undergo two-electron OA, while C(sp3) electrophiles prefer a single-electron pathway initiated by halogen atom transfer; (2) When nitrogen ligands are employed and C(sp3) electrophiles are involved, the reaction is inclined to proceed via a radical pathway, whereas a closed-shell reaction generally occurs with phosphine ligands; (3) Reaction systems involving photoredox catalysts or electrocatalysis typically undergo a single-electron transfer process. However, determining the conceivable reaction mechanism requires considering various reaction conditions, such as the choice of ligands and substrates, and whether the reducing system involves conventional metal and organic reducing agents or emerging photoredox and electrocatalysis. It is particularly important to ascertain whether a NiII catalyst precursor is ultimately reduced to Ni0 or NiI, as this determines the active catalyst in the reaction.

As for C(sp2)−C(sp3) XECs, the origin of the cross-selectivity over homo-selectivity can be ascribed to the different activation sequences of two electrophiles. Notably, different systems use different Ni species to activate electrophiles. For instance, in the Ni/dtbpy-catalyzed system for the XEC of iodobenzene with iodooctadecane, with Mn as the reducing agent, Ni0 and NiII are always used to activate the electrophiles. In contrast, NiIBr and NiIPh species can activate electrophiles when using more complex alkyl bromide containing olefins and heteroatoms substrate, Zn reductants, and 1,10-phenanthroline ligands. Generally speaking, due to the favorable π-metal interactions and the instability of aryl radicals, C(sp2) electrophiles tend to proceed through OA, while C(sp3) electrophiles are more likely to generate radicals. Additionally, OA and radical formation processes are also influenced by steric hindrance and electronic effects, respectively.

Despite significant progress made by both experimental and theoretical studies in characterizing the structures of key intermediates, providing reaction rates, identifying the activation modes of electrophiles, determining rate- or selectivity-determining steps, and identifying the origin of cross-selectivity, several limitations and challenges in mechanistic studies remain to be addressed: (1) There are limited studies on the mechanism of RCC reactions, especially those involving photoredox and electrochemistry catalysis; (2) The electron transfer process involved in the reduction process is still unclear; (3) So far, there is no data-driven and artificial intelligence linkages to aid further mechanistic exploration and reaction prediction. Therefore, the mechanistic investigation of nickel-catalyzed reductive couplings is far from complete, and concerted endeavors of experimental and computational studies are highly demanded to help chemists design more powerful and novel catalytic systems.

DECLARATIONS

Acknowledgments

We sincerely thank all the leading chemists and co-workers who have contributed to the methodological developments and mechanistic studies of nickel-catalyzed reduction cross-couplings.

Authors’ contributions

Guided this work, gave valuable suggestions and discussion for the review, and revised the paper: Zhang SQ, Hong X

Wrote the paper: Wu H

Availability of data and materials

Not applicable.

Financial support and sponsorship

This work was supported by the National Natural Science Foundation of China (22122109 and 22271253, Hong X; 22103070, Zhang SQ); National Key R&D Program of China (2022YFA1504301, Hong X); Zhejiang Provincial Natural Science Foundation of China under Grant No. LDQ23B020002 (Hong X); the Starry Night Science Fund of Zhejiang University Shanghai Institute for Advanced Study (SN-ZJU-SIAS-006, Hong X); Beijing National Laboratory for Molecular Sciences (BNLMS202102, Hong X); CAS Youth Interdisciplinary Team (JCTD-2021-11, Hong X); Fundamental Research Funds for the Central Universities (226-2022-00140, 226-2022-00224 and 226-2023-00115, Hong X); the Center of Chemistry for Frontier Technologies and Key Laboratory of Precise Synthesis of Functional Molecules of Zhejiang Province (PSFM 2021-01, Hong X); the State Key Laboratory of Clean Energy Utilization (ZJUCEU2020007, Hong X); the State Key Laboratory of Physical Chemistry of Solid Surfaces (202210, Hong X); the Leading Innovation Team grant from Department of Science and Technology of Zhejiang Province (2022R01005, Hong X) are gratefully acknowledged. Calculations were performed on the high‐performance computing system at the Department of Chemistry, Zhejiang University.

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2023.

REFERENCES

1. Lucas EL, Jarvo ER. Stereospecific and stereoconvergent cross-couplings between alkyl electrophiles. Nat Rev Chem 2017;1:0065.

2. Poremba KE, Dibrell SE, Reisman SE. Nickel-catalyzed enantioselective reductive cross-coupling reactions. ACS Catal 2020;10:8237-46.

3. Wang X, Dai Y, Gong H. Nickel-catalyzed reductive couplings. Top Curr Chem 2016;374:43.

4. Liu J, Ye Y, Sessler JL, Gong H. Cross-electrophile couplings of activated and sterically hindered halides and alcohol derivatives. Acc Chem Res 2020;53:1833-45.

5. Ni S, Li CX, Mao Y, et al. Ni-catalyzed deaminative cross-electrophile coupling of Katritzky salts with halides via C─N bond activation. Sci Adv 2019;5:eaaw9516.

6. Xue W, Jia X, Wang X, Tao X, Yin Z, Gong H. Nickel-catalyzed formation of quaternary carbon centers using tertiary alkyl electrophiles. Chem Soc Rev 2021;50:4162-84.

7. Pang X, Su PF, Shu XZ. Reductive cross-coupling of unreactive electrophiles. Acc Chem Res 2022;55:2491-509.

8. Yi L, Ji T, Chen K, Chen X, Rueping M. Nickel-catalyzed reductive cross-couplings: new opportunities for carbon-carbon bond formations through photochemistry and electrochemistry. CCS Chem 2022;4:9-30.

9. Durandetti M, Gosmini C, Périchon J. Ni-catalyzed activation of α-chloroesters: a simple method for the synthesis of α-arylesters and β-hydroxyesters. Tetrahedron 2007;63:1146-53.

10. Ye Y, Qi X, Xu B, et al. Nickel-catalyzed cross-electrophile allylation of vinyl bromides and the modification of anti-tumour natural medicine β-elemene. Chem Sci 2022;13:6959-66.

11. Charboneau DJ, Hazari N, Huang H, Uehling MR, Zultanski SL. Homogeneous organic electron donors in nickel-catalyzed reductive transformations. J Org Chem 2022;87:7589-609.

12. Charboneau DJ, Huang H, Barth EL, et al. Tunable and practical homogeneous organic reductants for cross-electrophile coupling. J Am Chem Soc 2021;143:21024-36.

13. Zhou J, Wang D, Xu W, Hu Z, Xu T. Enantioselective C(sp3)-C(sp3) reductive cross-electrophile coupling of unactivated alkyl halides with α-Chloroboronates via dual nickel/photoredox catalysis. J Am Chem Soc 2023;145:2081-7.

14. Zhang P, Le CC, MacMillan DW. Silyl radical activation of alkyl halides in metallaphotoredox catalysis: a unique pathway for cross-electrophile coupling. J Am Chem Soc 2016;138:8084-7.

15. Dongbang S, Doyle AG. Ni/photoredox-catalyzed C(sp3)-C(sp3) coupling between aziridines and acetals as alcohol-derived alkyl radical precursors. J Am Chem Soc 2022;144:20067-77.

16. Bacauanu V, Cardinal S, Yamauchi M, et al. Metallaphotoredox difluoromethylation of aryl bromides. Angew Chem Int Ed Engl 2018;57:12543-8.

17. Yang T, Wei Y, Koh MJ. Photoinduced nickel-catalyzed deaminative cross-electrophile coupling for C(sp2)–C(sp3) and C(sp3)–C(sp3) bond formation. ACS Catal 2021;11:6519-25.

18. Liu Y, Tao X, Mao Y, et al. Electrochemical C-N bond activation for deaminative reductive coupling of Katritzky salts. Nat Commun 2021;12:6745.

19. Liu D, Liu ZR, Wang ZH, et al. Paired electrolysis-enabled nickel-catalyzed enantioselective reductive cross-coupling between α-chloroesters and aryl bromides. Nat Commun 2022;13:7318.

20. Zackasee JLS, Zubaydi SA, Truesdell BL, Sevov CS. Synergistic catalyst-mediator pairings for electroreductive cross-electrophile coupling reactions. ACS Catal 2022;12:1161-6.

21. Perkins RJ, Pedro DJ, Hansen EC. Electrochemical Nickel Catalysis for Sp2-Sp3 cross-electrophile coupling reactions of unactivated alkyl halides. Org Lett 2017;19:3755-8.

22. Hu X, Cheng-Sánchez I, Cuesta-Galisteo S, Nevado C. Nickel-catalyzed enantioselective electrochemical reductive cross-coupling of aryl aziridines with alkenyl bromides. J Am Chem Soc 2023;145:6270-9.

23. Zhang B, Gao Y, Hioki Y, et al. Ni-electrocatalytic Csp3-Csp3 doubly decarboxylative coupling. Nature 2022;606:313-8.

24. Nohira I, Chatani N. Nickel-catalyzed cross-electrophile coupling between C(sp2)-F and C(sp2)-Cl bonds by the reaction of ortho-fluoro-aromatic amides with aryl chlorides. ACS Catal 2021;11:4644-9.

25. Ni S, Zhang W, Mei H, Han J, Pan Y. Ni-catalyzed reductive cross-coupling of amides with aryl iodide electrophiles via C-N bond activation. Org Lett 2017;19:2536-9.

26. Zhu Z, Gong Y, Tong W, Xue W, Gong H. Ni-catalyzed cross-electrophile coupling of aryl triflates with thiocarbonates via C-O/C-O bond cleavage. Org Lett 2021;23:2158-63.

27. Zhao C, Jia X, Wang X, Gong H. Ni-catalyzed reductive coupling of alkyl acids with unactivated tertiary alkyl and glycosyl halides. J Am Chem Soc 2014;136:17645-51.

28. Kim S, Goldfogel MJ, Gilbert MM, Weix DJ. Nickel-catalyzed cross-electrophile coupling of aryl chlorides with primary alkyl chlorides. J Am Chem Soc 2020;142:9902-7.

29. Guo P, Wang K, Jin WJ, et al. Dynamic kinetic cross-electrophile arylation of benzyl alcohols by nickel catalysis. J Am Chem Soc 2021;143:513-23.

30. Zhu Z, Lin L, Xiao J, Shi Z. Nickel-catalyzed stereo- and enantioselective cross-coupling of gem-difluoroalkenes with carbon electrophiles by C-F bond activation. Angew Chem Int Ed Engl 2022;61:e202113209.

31. Chi BK, Widness JK, Gilbert MM, Salgueiro DC, Garcia KJ, Weix DJ. In-situ bromination enables formal cross-electrophile coupling of alcohols with aryl and alkenyl halides. ACS Catal 2022;12:580-6.

32. Lin T, Gu Y, Qian P, Guan H, Walsh PJ, Mao J. Nickel-catalyzed reductive coupling of homoenolates and their higher homologues with unactivated alkyl bromides. Nat Commun 2020;11:5638.

33. Kang K, Weix DJ. Nickel-catalyzed C(sp3)-C(sp3) cross-electrophile coupling of in situ generated nhp esters with unactivated alkyl bromides. Org Lett 2022;24:2853-7.

34. Cui N, Lin T, Wang YE, et al. Nickel-catalyzed reductive coupling of γ-metalated ketones with unactivated alkyl bromides. Org Lett 2022;24:3987-92.

35. Diccianni JB, Diao T. Mechanisms of nickel-catalyzed cross-coupling reactions. Trends Chem 2019;1:830-44.

36. Hu X. Nickel-catalyzed cross coupling of non-activated alkyl halides: a mechanistic perspective. Chem Sci 2011;2:1867-86.

37. Everson DA, Weix DJ. Cross-electrophile coupling: principles of reactivity and selectivity. J Org Chem 2014;79:4793-8.

38. Weix DJ. Methods and mechanisms for cross-electrophile coupling of Csp2 halides with alkyl electrophiles. Acc Chem Res 2015;48:1767-75.

39. Diccianni J, Lin Q, Diao T. Mechanisms of nickel-catalyzed coupling reactions and applications in alkene functionalization. Acc Chem Res 2020;53:906-19.

40. Yang T, Jiang Y, Luo Y, Lim JJH, Lan Y, Koh MJ. Chemoselective union of olefins, organohalides, and redox-active esters enables regioselective alkene dialkylation. J Am Chem Soc 2020;142:21410-9.

41. Duan A, Xiao F, Lan Y, Niu L. Mechanistic views and computational studies on transition-metal-catalyzed reductive coupling reactions. Chem Soc Rev 2022;51:9986-10015.

42. Wang Y, Ren Q. DFT study of the mechanisms of transition-metal-catalyzed reductive coupling reactions. Curr Org Chem 2020;24:1367-83.

43. Gu J, Wang X, Xue W, Gong H. Nickel-catalyzed reductive coupling of alkyl halides with other electrophiles: concept and mechanistic considerations. Org Chem Front 2015;2:1411-21.

44. Zhang SQ, Hong X. Mechanism and selectivity control in Ni- and Pd-catalyzed cross-couplings involving carbon-oxygen bond activation. Acc Chem Res 2021;54:2158-71.

45. Gutierrez O, Tellis JC, Primer DN, Molander GA, Kozlowski MC. Nickel-catalyzed cross-coupling of photoredox-generated radicals: uncovering a general manifold for stereoconvergence in nickel-catalyzed cross-couplings. J Am Chem Soc 2015;137:4896-9.

46. Kolahdouzan K, Khalaf R, Grandner JM, Chen Y, Terrett JA, Huestis MP. Dual photoredox/nickel-catalyzed conversion of aryl halides to aryl aminooxetanes: computational evidence for a substrate-dependent switch in mechanism. ACS Catal 2020;10:405-11.

47. Xu J, Yang Y, Zhao X, Liu C, Zhang D. DFT mechanistic study of IrIII/NiII-metallaphotoredox-catalyzed difluoromethylation of aryl bromides. Inorg Chem 2021;60:8682-91.

48. Biswas S, Weix DJ. Mechanism and selectivity in nickel-catalyzed cross-electrophile coupling of aryl halides with alkyl halides. J Am Chem Soc 2013;135:16192-7.

49. Everson DA, Shrestha R, Weix DJ. Nickel-catalyzed reductive cross-coupling of aryl halides with alkyl halides. J Am Chem Soc 2010;132:920-1.

50. Everson DA, Jones BA, Weix DJ. Replacing conventional carbon nucleophiles with electrophiles: nickel-catalyzed reductive alkylation of aryl bromides and chlorides. J Am Chem Soc 2012;134:6146-59.

51. Wotal AC, Ribson RD, Weix DJ. Stoichiometric reactions of acylnickelII complexes with electrophiles and the catalytic synthesis of ketones. Organometallics 2014;33:5874-81.

52. Sheng J, Ni HQ, Zhang HR, Zhang KF, Wang YN, Wang XS. Nickel-catalyzed reductive cross-coupling of aryl halides with monofluoroalkyl halides for late-stage monofluoroalkylation. Angew Chem Int Ed Engl 2018;57:7634-9.

53. Brauer DJ, Krueger C. Bonding of aromatic hydrocarbons to nickel (0). Structure of bis (tricyclohexylphosphine)(1,2-.eta.2-anthracene) nickel (0)-toluene. Inorg Chem 1977;16:884-91.

54. Ren Q, Jiang F, Gong H. DFT study of the single electron transfer mechanisms in Ni-Catalyzed reductive cross-coupling of aryl bromide and alkyl bromide. J Organomet Chem 2014;770:130-5.

55. Kumar GS, Peshkov A, Brzozowska A, Nikolaienko P, Zhu C, Rueping M. Nickel-catalyzed chain-walking cross-electrophile coupling of alkyl and aryl halides and olefin hydroarylation enabled by electrochemical reduction. Angew Chem Int Ed Engl 2020;59:6513-9.

56. Wang X, Ma G, Peng Y, et al. Ni-catalyzed reductive coupling of electron-rich aryl iodides with tertiary alkyl halides. J Am Chem Soc 2018;140:14490-7.

57. Ji H, Lin D, Tai L, et al. Nickel-catalyzed enantioselective coupling of acid chlorides with α-bromobenzoates: an asymmetric acyloin synthesis. J Am Chem Soc 2022;144:23019-29.

58. Guan H, Zhang Q, Walsh PJ, Mao J. Nickel/photoredox-catalyzed asymmetric reductive cross-coupling of racemic α-chloro esters with aryl iodides. Angew Chem Int Ed Engl 2020;59:5172-7.

59. Lau SH, Borden MA, Steiman TJ, Wang LS, Parasram M, Doyle AG. Ni/photoredox-catalyzed enantioselective cross-electrophile coupling of styrene oxides with aryl iodides. J Am Chem Soc 2021;143:15873-81.

60. Wang H, Zheng P, Wu X, Li Y, Xu T. Modular and facile access to chiral α-aryl phosphates via dual nickel- and photoredox-catalyzed reductive cross-coupling. J Am Chem Soc 2022;144:3989-97.

61. Yuan M, Song Z, Badir SO, Molander GA, Gutierrez O. On the nature of C(sp3)- C(sp2) bond formation in nickel-catalyzed tertiary radical cross-couplings: a case study of Ni/photoredox catalytic cross-coupling of alkyl radicals and aryl halides. J Am Chem Soc 2020;142:7225-34.

62. Paul A, Smith MD, Vannucci AK. Photoredox-assisted reductive cross-coupling: mechanistic insight into catalytic aryl-alkyl cross-couplings. J Org Chem 2017;82:1996-2003.

63. Hamby TB, LaLama MJ, Sevov CS. Controlling Ni redox states by dynamic ligand exchange for electroreductive Csp3-Csp2 coupling. Science 2022;376:410-6.

64. Li Z, Sun W, Wang X, Li L, Zhang Y, Li C. Electrochemically enabled, nickel-catalyzed dehydroxylative cross-coupling of alcohols with aryl halides. J Am Chem Soc 2021;143:3536-43.

65. Diccianni JB, Katigbak J, Hu C, Diao T. Mechanistic characterization of (Xantphos)Ni(I)-mediated alkyl bromide activation: oxidative addition, electron transfer, or halogen-atom abstraction. J Am Chem Soc 2019;141:1788-96.

66. Funes-Ardoiz I, Nelson DJ, Maseras F. halide abstraction competes with oxidative addition in the reactions of aryl halides with [Ni(PMen Ph(3-n))4]. Chemistry 2017;23:16728-33.

67. Kehoe R, Mahadevan M, Manzoor A, et al. Reactions of the Ni(0) Compound Ni(PPh 3)4 with unactivated alkyl halides: oxidative addition reactions involving radical processes and nickel(I) intermediates. Organometallics 2018;37:2450-67.

68. Yoo C, Ajitha MJ, Jung Y, Lee Y. Mechanistic study on C-C bond formation of a nickel(I) monocarbonyl species with alkyl iodides: experimental and computational investigations. Organometallics 2015;34:4305-11.

69. Lin Q, Fu Y, Liu P, Diao T. Monovalent nickel-mediated radical formation: a concerted halogen-atom dissociation pathway determined by electroanalytical studies. J Am Chem Soc 2021;143:14196-206.

70. Guo L, Yuan M, Zhang Y, et al. General method for enantioselective three-component carboarylation of alkenes enabled by visible-light dual photoredox/nickel catalysis. J Am Chem Soc 2020;142:20390-9.

71. Jiang F, Ren Q. Theoretical investigation of the mechanisms of the biphenyl formation in Ni-catalyzed reductive cross-coupling system. J Organomet Chem 2014;757:72-8.

72. Liu J, Ren Q, Zhang X, Gong H. Preparation of vinyl arenes by nickel-catalyzed reductive coupling of aryl halides with vinyl bromides. Angew Chem Int Ed Engl 2016;55:15544-8.

73. Long C, Ni S, Su M, Wang X, Tan W. Highly chemoselective access to 2,2′-diaminobiaryls via Ni-catalyzed protecting-group-free coupling of 2-haloanilines. ACS Catal 2020;10:13641-9.

74. Long CY, Chen H, Ma C, et al. Highly chemoselective Ni-catalyzed protecting-group-free 2,2'-biphenol synthesis and mechanistic insights. Org Lett 2022;24:4155-9.

75. Yue H, Zhu C, Shen L, et al. Nickel-catalyzed C-N bond activation: activated primary amines as alkylating reagents in reductive cross-coupling. Chem Sci 2019;10:4430-5.

76. Lin Q, Diao T. Mechanism of Ni-catalyzed reductive 1,2-dicarbofunctionalization of alkenes. J Am Chem Soc 2019;141:17937-48.

77. Cornella J, Gómez-Bengoa E, Martin R. Combined experimental and theoretical study on the reductive cleavage of inert C-O bonds with silanes: ruling out a classical Ni(0)/Ni(II) catalytic couple and evidence for Ni(I) intermediates. J Am Chem Soc 2013;135:1997-2009.

78. Chen H, Yue H, Zhu C, Rueping M. Reactivity in nickel-catalyzed multi-component sequential reductive cross-coupling reactions. Angew Chem Int Ed Engl 2022;61:e202204144.

79. Wang X, Ma Y, Ye J, Liu Z, Cheng R. Theoretical investigation of mechanism on nickel-catalyzed electrochemical cross-coupling of aryl bromides and arylamines. Catal Sci Technol 2023;13:1032-40.

80. Tang T, Jones E, Wild T, Hazra A, Minteer SD, Sigman MS. Investigating oxidative addition mechanisms of allylic electrophiles with low-valent Ni/Co catalysts using electroanalytical and data science techniques. J Am Chem Soc 2022;144:20056-66.

81. Ting SI, Williams WL, Doyle AG. Oxidative addition of aryl halides to a Ni(I)-bipyridine complex. J Am Chem Soc 2022;144:5575-82.

82. Till NA, Oh S, MacMillan DWC, Bird MJ. The application of pulse radiolysis to the study of Ni(I) intermediates in ni-catalyzed cross-coupling reactions. J Am Chem Soc 2021;143:9332-7.

83. Breitenfeld J, Wodrich MD, Hu X. Bimetallic oxidative addition in nickel-catalyzed alkyl-aryl kumada coupling reactions. Organometallics 2014;33:5708-15.

84. Breitenfeld J, Ruiz J, Wodrich MD, Hu X. Bimetallic oxidative addition involving radical intermediates in nickel-catalyzed alkyl-alkyl kumada coupling reactions. J Am Chem Soc 2013;135:12004-12.

85. Laskowski CA, Bungum DJ, Baldwin SM, Del Ciello SA, Iluc VM, Hillhouse GL. Synthesis and reactivity of two-coordinate Ni(I) alkyl and aryl complexes. J Am Chem Soc 2013;135:18272-5.

86. Ju L, Lin Q, LiBretto NJ, et al. Reactivity of (bi-Oxazoline)organonickel complexes and revision of a catalytic mechanism. J Am Chem Soc 2021;143:14458-63.

87. Thane TA, Jarvo ER. Ligand-based control of nickel catalysts: switching chemoselectivity from one-electron to two-electron pathways in competing reactions of 4-halotetrahydropyrans. Org Lett 2022;24:5003-8.

88. Shu W, García-Domínguez A, Quirós MT, Mondal R, Cárdenas DJ, Nevado C. Ni-catalyzed reductive dicarbofunctionalization of nonactivated alkenes: scope and mechanistic insights. J Am Chem Soc 2019;141:13812-21.

89. Zhu C, Lee SC, Chen H, Yue H, Rueping M. Reductive cross-coupling of α-oxy halides enabled by thermal catalysis, photocatalysis, electrocatalysis, or mechanochemistry. Angew Chem Int Ed Engl 2022;61:e202204212.

90. Day CS, Rentería-gómez Á, Ton SJ, Gogoi AR, Gutierrez O, Martin R. Elucidating electron-transfer events in polypyridine nickel complexes for reductive coupling reactions. Nat Catal 2023;6:244-53.

91. Tang T, Hazra A, Min DS, et al. Interrogating the mechanistic features of Ni(I)-mediated aryl iodide oxidative addition using electroanalytical and statistical modeling techniques. J Am Chem Soc 2023;145:8689-99.

92. Chen PP, Lucas EL, Greene MA, et al. A unified explanation for chemoselectivity and stereospecificity of Ni-catalyzed kumada and cross-electrophile coupling reactions of benzylic ethers: a combined computational and experimental study. J Am Chem Soc 2019;141:5835-55.

93. Hewitt KA, Xie P, Thane TA, et al. Nickel-catalyzed domino cross-electrophile coupling dicarbofunctionalization reaction to afford vinylcyclopropanes. ACS Catal 2021;11:14369-80.

94. Lucas EL, Hewitt KA, Chen PP, Castro AJ, Hong X, Jarvo ER. Engaging sulfonamides: intramolecular cross-electrophile coupling reaction of sulfonamides with alkyl chlorides. J Org Chem 2020;85:1775-93.

95. Xu Z, Yang Y, Jiang J, Fu Y. Theoretical investigation on Ni-catalyzed C(sp3)-F activation and ring contraction of tetrahydropyrans: exploration of an SN2 pathway. Organometallics 2018;37:1114-22.

96. Lucas EL, Mcginnis TM, Castro AJ, Jarvo ER. Nickel-catalyzed cross-electrophile coupling of the difluoromethyl group for fluorinated cyclopropane synthesis. Synlett 2021;32:1525-30.

97. Sanford AB, Thane TA, McGinnis TM, Chen PP, Hong X, Jarvo ER. Nickel-catalyzed alkyl-alkyl cross-electrophile coupling reaction of 1,3-dimesylates for the synthesis of alkylcyclopropanes. J Am Chem Soc 2020;142:5017-23.

Cite This Article

Export citation file: BibTeX | RIS

OAE Style

Wu H, Zhang SQ, Hong X. Mechanisms of nickel-catalyzed reductive cross-coupling reactions. Chem Synth 2023;3:39. http://dx.doi.org/10.20517/cs.2023.20

AMA Style

Wu H, Zhang SQ, Hong X. Mechanisms of nickel-catalyzed reductive cross-coupling reactions. Chemical Synthesis. 2023; 3(4): 39. http://dx.doi.org/10.20517/cs.2023.20

Chicago/Turabian Style

Wu, Hongli, Shuo-Qing Zhang, Xin Hong. 2023. "Mechanisms of nickel-catalyzed reductive cross-coupling reactions" Chemical Synthesis. 3, no.4: 39. http://dx.doi.org/10.20517/cs.2023.20

ACS Style

Wu, H.; Zhang S.Q.; Hong X. Mechanisms of nickel-catalyzed reductive cross-coupling reactions. Chem. Synth. 2023, 3, 39. http://dx.doi.org/10.20517/cs.2023.20

About This Article

© The Author(s) 2023. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, 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.

Data & Comments

Data

Views
835
Downloads
172
Citations
0
Comments
0
2

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Cite This Article 14 clicks
Like This Article 2 likes
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Chemical Synthesis
ISSN 2769-5247 (Online)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/