Keywords

1 Introduction

Nitriles are common and ubiquitous organic compounds. They act as different functional molecules such as pharmaceuticals, pesticides, organic materials, and polymers; they are also important building blocks in organic synthesis. Though cyano group can be readily converted to carbonyl and aminoalkyl groups by a broad range of methods, it has rarely been regarded as a leaving group in organic syntheses, except for acyl and alkoxycarbonyl cyanides and some reactions involving electron transfer and/or addition/elimination pathways to result in the release of cyanide. C–CN bonds generally tolerate the various reaction conditions of many organic transformations, owing partly to their high bond dissociation energies (>100 kcal/mol). Nevertheless, low-valence transition metal complexes have shown that C–CN bonds of nitriles can be cleaved. Nitriles coordinate to a metal center either in a η 1- or in a η 2-manner. High-valence Lewis acidic metal complexes favor η 1-coordination of nitrogen atom in nitriles, whereas low-valence metal complexes often show η 2-coordination which can be strengthened through π-back donation (Scheme 1) [1]. In many cases, C–CN bond activation is initiated by η 2-coordination. Two main pathways have been revealed for the cleavage step: oxidative addition and formation of silylisonitrile complexes. Oxidative addition was already reported in 1971, when a group in DuPont described how benzonitrile adds to nickel(0) species at room temperature (Scheme 2) [2]. This elemental reaction of benzonitrile and other nitriles has been studied extensively [227] and is discussed in more detail in Chap. 1. The DuPont team has studied the catalytic isomerization of 2-methyl-3-butenenitrile (2M3BN) to 3- and/or 4-pentenenitrile (3PN, 4PN) via oxidative addition of the C–CN bond of 2M3BN to form a π-allylnickel intermediate (Scheme 3) [28]. This reaction is a part of DuPont's adiponitrile (ADN) process and represents a very early example of catalytic C–CN bond activation. The catalytic isomerization is still a topic of recent research by several groups [2937] but will not be discussed further in this chapter. The DuPont team [38] and more recently Jones [14] have also revealed the effect of Lewis acid additives on the oxidative addition of C–CN bonds.

Scheme 1
scheme 1

C–CN activation by transition metal complexes

Scheme 2
scheme 2

Oxidative addition of benzonitrile to Ni(PEt3)4

Scheme 3
scheme 3

Isomerization of 2M3BN via C–CN activation catalyzed by Ni/ZnCl2 in the DuPont's ADN process

C–CN bond activation through the formation of silylisonitrile metal complexes has been disclosed for rhodium complexes by Brookhart and coworkers for the first time (Scheme 4) [39, 40]. Silyliron [41], silylene–iron [41], and silylene–ruthenium [42] complexes have also been demonstrated to activate C–CN bonds. Detailed mechanistic studies have been performed on the C–CN bond activation of this type to show the intermediacy of η 2-iminoacyls, for which C–CN cleavage takes place leading to the formation of silylisonitriles.

Scheme 4
scheme 4

C–CN activation by silylrhodium(I) complexes

In spite of the rich chemistry of stoichiometric C–CN bond activation by various transition metal complexes via the different pathways described above, their application in catalytic transformations of nitriles directed to organic synthesis has rarely emerged until the last 10 years. This review features the progress of catalytic reactions via C–CN activation (for a previous review on this topic see [43]). Particular focus of this review is on C–CN activation by metal complexes to give possibly organometallic intermediates bearing organic and/or cyano groups bound to a metal center, and thus, conventional synthetic transformations involving C–CN activation by electron transfer and/or addition/elimination pathways are not covered even when metal catalysts are involved.

2 Coupling Reactions via C–CN Bond Activation

Metal-catalyzed cross-coupling reaction is undoubtedly one of the most important methodologies in modern organic synthesis. This transformation typically employs aryl and heteroaryl halides and a wide range of nucleophiles to construct typically substituted arenes. Recent studies have also shown the use of aryl sulfonates and esters as aryl pseudohalides. These aryl electrophiles are known to undergo oxidative addition to palladium(0) and nickel(0) species to initiate a catalytic cycle for cross-coupling. Thus, the application of C–CN bond activation to cross-coupling catalysis, particularly through oxidative addition, can be envisaged, making aryl cyanides an alternative for aryl electrophiles in cross-coupling reactions.

2.1 Hydrodecyanation Reactions

Oxidative addition of C–CN bonds to nickel(0) can be followed by transmetalation with various main-group organometallic reagents, and subsequent reductive elimination can result in the functionalization of C–CN bonds of nitriles (Scheme 5). As the simplest case, C–CN bonds can be transformed to C–H bonds via transmetalation with metal hydrides. Indeed, nickel-catalyzed hydrodecyanation of various aromatic and aliphatic nitriles proceeds with tetramethyldisiloxane as a hydride donor (Scheme 6) [44]. While a wide range of nitriles can be decyanated by this protocol, a relatively high amount of catalyst is required in this process, presumably because of the formation of catalytically inactive (PCy3)2Ni(CN)2. The use of AlMe3 as a Lewis acid is effective in some cases to promote the C–CN bond activation. Under these reaction conditions, the relative reactivity order of different aryl electrophiles is estimated: Ar–SMe>Ar–CN>Ar–OAr>Ar–OMe.

Scheme 5
scheme 5

A general catalytic cycle for nickel-catalyzed coupling reactions via C–CN activation

Scheme 6
scheme 6

Hydrodecyanation of nitriles with hydrosilanes catalyzed by nickel

Alternatively, iron- or rhodium-catalyzed hydrodecyanation reaction is proposed to be initiated by the C–CN activation of aromatic and aliphatic nitriles with silylmetal species to give iminoacylmetal intermediates and then silylisonitrile metal complexes, which produce decyanated products and silyl cyanides upon the reaction with hydrosilanes to reproduce the catalytically active silylmetal intermediates (Scheme 7). Irradiation is required to generate coordinatively unsaturated iron complexes (Scheme 8) [45, 46]. A similar reaction mechanism also operates with rhodium catalysis (Scheme 9) [47, 48]. Rhodium-catalyzed reaction tolerates a wide range of nitriles, including tertiary alkyl cyanides, using triisopropylsilane as a reducing agent. These catalytic hydrodecyanation reactions can be nicely combined with conventional transformations of nitriles, demonstrating that cyano group can serve as a “removable directing group” (Scheme 10).

Scheme 7
scheme 7

A plausible catalytic cycle for hydrodecyanation of nitriles with hydrosilanes

Scheme 8
scheme 8

Hydrodecyanation of nitriles with hydrosilanes catalyzed by iron

Scheme 9
scheme 9

Hydrodecyanation of nitriles with hydrosilanes catalyzed by rhodium

Scheme 10
scheme 10

Synthetic applications utilizing rhodium-catalyzed hydrodecyanation

2.2 Cross-Coupling Reactions

The transmetalation mentioned above for metal hydrides can naturally be extended to the use of main-group organometallic reagents to perform cross-coupling reactions using aryl cyanides instead of aryl halides. This reaction was first demonstrated with modified aryl, alkenyl, and alkyl Grignard reagents in the presence of Ni/PMe3 catalyst (Scheme 11) [49, 50]. Alkynylzinc reagents also cross-couple with aryl cyanides to give various disubstituted acetylenes (Scheme 11) [51]. More recently, milder nucleophiles, such as arylboron reagents, have been introduced to undergo the cross-coupling reaction (Scheme 11) [52].

Scheme 11
scheme 11

Cross-coupling reaction of aryl cyanides with carbon nucleophiles catalyzed by nickel

Arylrhodium species generated upon the cleavage of an Ar–CN bond by a silylrhodium intermediate, derived from the reaction of rhodium(I) with disilanes (see below), can be trapped by aryl halides in an intramolecular manner to give dibenzofurans and carbazoles (Scheme 12) [53]. The arylrhodium species can also react intermolecularly with vinylsilanes to give 2-arylethenylsilanes as a Heck-type product (Scheme 12) [54].

Scheme 12
scheme 12

C–C bond-forming coupling reactions of aryl cyanides catalyzed by rhodium

In addition to carbon nucleophiles, nitrogen-based nucleophiles can be used for the nickel-catalyzed cross-coupling with aryl cyanides in the presence of CsF, the role of which is yet to be clarified (Scheme 13) [55]. Silylphosphines are reported to serve as a phosphorus-based nucleophile to give a variety of aryl(diphenyl)phosphines by nickel-catalyzed cross-coupling reaction with aryl cyanides (Scheme 13) [56]. The use of a stoichiometric amount of strong bases such as KOt-Bu and NaOMe is crucial, and nucleophilic MPPh3 (M = K, Na) may be generated in situ through the reaction of the bases with the silylphosphine reagents.

Scheme 13
scheme 13

Cross-coupling reaction of aryl cyanides with nitrogen- and phosphorus-based nucleophiles catalyzed by nickel

2.3 Silylation and Borylation Reactions

C–CN bond activation by silylrhodium species can be applied to a catalytic cycle for silylative decyanation of nitriles, when rhodium intermediates active in C–CN activation are generated from disilanes instead of hydrosilanes (Scheme 14). The reaction proceeds with a range of nitriles, including aryl, alkenyl, and alkyl cyanides, to give the corresponding tetraorganosilanes (Scheme 15) [53, 57], although the yields of tetra-alkylsilanes are modest.

Scheme 14
scheme 14

A plausible mechanism for rhodium-catalyzed decyanative silylation of nitriles

Scheme 15
scheme 15

Decyanative silylation of nitriles catalyzed by rhodium

More recently, decyanative borylation of nitriles is found to proceed with diboranes to give organoboron compounds through C–CN activation (Scheme 16) [58, 59]. A borylrhodium(I) species is expected to be responsible for the C–CN activation in a manner similar to that by a silylrhodium(I) species based on theoretical calculations (Scheme 17) [60]. The intermediates subsequently react with diborane reagents via oxidative addition to give rhodium(III) intermediates, which undergo reductive elimination to give cyanoboranes and borylation products and to regenerate the catalytically active borylrhodium(I) species.

Scheme 16
scheme 16

Decyanative borylation of nitriles catalyzed by rhodium

Scheme 17
scheme 17

A plausible mechanism for rhodium-catalyzed decyanative borylation of nitriles

2.4 Cycloaddition Reactions

C–CN activation via oxidative addition can be followed by the activation of another C–C bond to develop cycloaddition reactions. The reaction of o-arylcarboxybenzonitrile with alkynes proceeds in this manner to give coumarins, aryl cyanides, and an alkyne-arylcyanation product in the presence of catalytic amounts of nickel and aluminum-based Lewis acid (Scheme 18) [61]. Likewise, o-cyanophenylbenzamides undergo the transformation to give quinolones (Scheme 18 [62]. A catalytic cycle involving a five-membered nickelacycle intermediate, generated possibly by the oxidative addition of Ar–CN bonds, and the subsequent C–C bond activation [63] is proposed (Scheme 19).

Scheme 18
scheme 18

Cycloaddition reactions of aryl cyanides with alkynes via C–CN and acyl–Ar activation

Scheme 19
scheme 19

A plausible mechanism for nickel-catalyzed cycloaddition via C–CN and acyl–Ar activation

3 Cyanation Reactions via C–CN Bond Activation

C–CN activation by metal complexes often generates metal cyanides, which can serve as cyanating agents to give nitriles as a product. Because many of the cyanation reactions have conventionally been performed by using a stoichiometric amount of generally highly toxic metal cyanides and/or hydrogen cyanides, cyanation reactions via metal cyanides generated catalytically in situ through C–CN activation can be less toxic. Thus, they can be practically useful alternative protocols to introduce a cyano functionality into organic molecules using commonly available less toxic nitriles. Moreover, if both organic and cyano groups of nitriles can be introduced at the same time through C–CN activation, nitriles having higher complexity can be readily accessed in a single operation. These cyanation reactions via C–CN activation are described in this section.

3.1 Cyanation of Aryl Halides and Arenes

Copper has been demonstrated to mediate cyanation of aryl bromides and iodides through the activation of C–CN bonds (Scheme 20). Phenylacetonitrile [64], malononitrile [65], and even acetonitrile as a reaction solvent [66] have been reported to serve as cyanating agents. 2-Phenylpyridines [67] and indoles [68] are directly cyanated by copper-mediated cyanation reaction using phenylacetonitrile, which is supposedly oxidized first at its benzylic position to give benzoyl cyanide, which further reacts with copper complexes to generate a cyanocopper species responsible for the cyanation event. Nevertheless, detailed mechanisms of these cyanation reactions remain elusive.

Scheme 20
scheme 20

Cyanation of aryl halides and arenes with nitriles catalyzed or mediated by copper

Palladium complexes have also been reported to catalyze the cyanation of aryl halides via C–CN activation of nitriles, such as phenylacetonitrile [69] and ethyl cyanoacetate [70] (Scheme 21). An excess amount of acetonitrile can also be activated to serve as a cyanating agent in the presence of palladium catalyst [71]. The mechanism of C–CN activation in these palladium-catalyzed cyanation reactions is also yet to be understood.

Scheme 21
scheme 21

Cyanation of aryl halides with nitriles catalyzed by palladium

3.2 Carbocyanation of Unsaturated Bonds

The oxidative addition of C–CN bonds of nitriles can be followed by migratory insertion of unsaturated compounds into C–metal bonds and subsequent reductive elimination to develop addition reactions of organic and cyano groups of nitriles across unsaturated compounds through C–CN cleavage, namely by carbocyanation reaction (Scheme 22) [72, 73]. Initially, these synthetically novel transformations were attempted by using benzoyl cyanide and alkynes in the presence of a palladium catalyst (Scheme 23) [74, 75]. The reaction proceeds to give cis-adducts but with a mechanistic scenario different from that shown in Scheme 22. Thus, arylacetylenes are first acylated by the nitrile substrate, and the thus generated HCN adds across the aroyl(aryl)acetylenes to give formal trans-aroylcyanation products, which finally isomerize under the reaction conditions to give cis-adducts.

Scheme 22
scheme 22

A possible catalytic cycle for carbocyanation of unsaturated bonds

Scheme 23
scheme 23

Benzoylcyanation of arylacetylenes catalyzed by palladium

Given a number of examples of stoichiometric studies on the oxidative addition of C–CN bonds to nickel(0), nickel catalysts were envisaged to catalyze the carbocyanation reaction. Indeed, aryl cyanides add across alkynes in the presence of Ni/PMe3 catalyst to give various tetra-substituted olefins (Scheme 24) [76, 77]. Allyl cyanides are also viable nitrile substrates to undergo alkyne-carbocyanation reaction by using less electron-donating P(4-CF3–C6H4)3 as a ligand [78]. Nevertheless, the scope of nitriles is limited and the catalyst loadings of the original protocols are quite high for nickel-catalyzed carbocyanation. DFT calculations of the arylcyanation of alkynes have revealed that the oxidative addition of Ar–CN bonds to nickel(0) via η 2-arene nickel intermediates is the rate-determining step [79]. Therefore, it has been envisioned that the promotion of this step could overcome the limitations.

Scheme 24
scheme 24

Aryl- and allylcyanation of alkynes catalyzed by nickel

As mentioned above, the presence of Lewis acidic additives, such as triorganoaluminums and -aluminums, are known to facilitate the oxidative addition of C–CN bonds to nickel(0) species through the coordination of the nitrogen atom of cyano group to the Lewis acids [14, 38]. The arylcyanation of alkynes is indeed significantly accelerated by using aluminum-based Lewis acid cocatalysts (Scheme 25) [80, 81]. By cooperative nickel/Lewis acid catalysis, the scope of aryl cyanides has been improved to include those having labile bromo and chloro groups as well as sterically demanding substrates. The reaction conditions of allylcyanation can also be made milder by using AlMe2Cl as a Lewis acid cocatalyst, allowing highly stereoselective preparation of tri-substituted alkenes bearing a bulkier substituent at the cyano-substituted carbon [82]. Similar regioselectivity is also observed in the arylcyanation of alkynes and can be ascribed to aryl- or allylnickelation proceeding preferentially at sterically less hindered carbons of coordinated alkynes [79]. The protocol employing α-siloxyallyl cyanides affords tri-substituted ethenes bearing a formyl functionality upon the hydrolysis of silyl enol ether products, and can thus be used for the synthesis of functionalized multi-substituted olefins such as plaunotol, a diterpene known for antibacterial activity against Helicobacter pylori (Scheme 26).

Scheme 25
scheme 25

Aryl- and allylcyanation of alkynes catalyzed cooperatively by nickel/aluminum

Scheme 26
scheme 26

Total synthesis of plaunotol through allylcyanation of alkynes

The scope of nitriles for the carbocyanation reaction of alkynes can be expanded by cooperative nickel/Lewis acid catalysis. Alkenyl [80, 81] and alkynyl cyanides [83, 84] also participate in the addition reaction to give highly conjugated nitrile products by nickel/BPh3 catalysis (Scheme 27). The use of aluminum-based Lewis acids causes isomerization of the double bond of alkenylcyanation products, whereas alkynylcyanation is sluggish with the aluminum reagents.

Scheme 27
scheme 27

Alkenyl- and alkynylcyanation of alkynes catalyzed by nickel/BPh3

Some nitriles also add across 1,2-dienes (Scheme 28). Alkynylcyanation takes place predominantly across the internal double bond of 1,2-dienes to give selectively cyanoalkyl-substituted enynes [83, 84]. Cyanoformates also add across 1,2-dienes in a similar manner in the presence of nickel catalyst alone to give cyanoalkyl-substituted acrylates [85, 86], whereas carbocyanation of 1,2-dienes with other nitriles remains unexplored. The 1,2-diene–carbocyanation can be initiated by oxidative addition followed by the coordination of 1,2-dienes at the terminal double bond, and subsequent migratory insertion into the C–Ni bond (Scheme 29). The subsequently formed allylnickel intermediates can be isomerized to give π-allylnickel species, which is likely responsible for the carbocyanation of the internal olefins.

Scheme 28
scheme 28

Carbocyanation of 1,2-dienes using alkynyl cyanides and cyanoformates

Scheme 29
scheme 29

A plausible catalytic cycle for carbocyanation of 1,2-dienes

Alkynes can also be functionalized both stereoselectively and regioselectively by cooperative catalysis (Scheme 30) [87]. Cyanoesterification of silyl-substituted alkynes proceeds to give β-cyanoesters with silyl group at α-position as a single product. The regiochemistry in contrast to other carbocyanation reactions may be derived from interaction of carbonyl with the silyl group prior to the carbonickelation event. A catalytic cycle through the oxidative addition of C–CN bonds is supported by the stoichiometric reaction of cyanoformamide with nickel(0) complex and BPh3 to give an oxidative adduct and its further reaction with an alkyne, as well as by its use as a catalyst, both giving the corresponding cyanocarbamoylation product (Scheme 31). Palladium complexes, on the other hand, have been reported to catalyze intramolecular cyanoesterification of alkynes to give lactone products (Scheme 32) [88].

Scheme 30
scheme 30

Cyanoesterification of silyl-substituted alkynes catalyzed by nickel/BPh3

Scheme 31
scheme 31

Oxidative addition of cyanoformamides to nickel(0) assisted by BPh3 and its intermediacy in cyanocarbamoylation of alkynes

Scheme 32
scheme 32

Intramolecular cyanoesterification of alkynes catalyzed by palladium

Alkanenitriles including acetonitriles and propionitriles undergo a reaction across alkynes to give cis-methylcyanation and ethylcyanation products, respectively (Scheme 33) [80, 89, 90]. Although the methylcyanation proceeds with excellent regioselectivity, formal trans-adducts derived from the isomerization of double bonds are contaminated. A trace amount of hydrocyanation product is observed in the ethylcyanation reaction, possibly through the β-hydride elimination from ethylnickel species, which can be generated upon the oxidative addition of Et–CN bond to nickel(0). Byproducts derived from this unwanted pathway can be observed in much higher amounts with alkyl cyanides having higher alkyl chains.

Scheme 33
scheme 33

Alkylcyanation of alkynes catalyzed by nickel/aluminum

When alkanenitriles having a heteroatom functionality at the γ-position are used for the alkylcyanation, hydrocyanation byproducts can be suppressed (Scheme 34) [91]. With these particular nitriles, oxidative adducts can possibly form metallacycle intermediates through intramolecular coordination of heteroatom functionalities to retard the unwanted β-hydride elimination because of the absence of vacant coordination sites. Even secondary alkyl cyanides participate in the addition reaction in good yields when they have an amino group at their γ-position. Oxygen- and sulfur-based functional groups can also serve as directing groups to give the corresponding functionalized alkylcyanation products.

Scheme 34
scheme 34

Heteroatom-directed alkylcyanation of alkynes catalyzed by nickel/AlMe3

While carbocyanation reactions across alkynes show a broad scope of nitriles as described above, intermolecular reactions across simple alkenes are generally sluggish, probably because of the reluctant C(sp3)–CN bond-forming reductive elimination, which competes with β-hydride elimination. For example, the reaction of aryl cyanides with vinylsilanes gives a Heck-type product in modest yield, possibly through migratory insertion of the alkenes into the Ar–Ni bond followed by β-hydride elimination (Scheme 35). Bicyclic alkenes, typically norbornene and norbornadiene, on the other hand, successfully undergo the nickel-catalyzed carbocyanation to give exo-cis adducts exclusively (Scheme 36) [81, 84, 92]. Palladium can catalyze the cyanoesterification of norbornene and norbornadiene [9395]. The observed exo-selectivity can be ascribed to the higher electron density of the exo-face [96]. These functionalized norbornenes can be monomers for ring-opening metathesis polymerization to give functional polymer materials [97101]. Successful C(sp3)–CN bond-forming reductive elimination can be possible because of unfavorable β-hydride elimination to give rise to anti-Bredt olefin products.

Scheme 35
scheme 35

Attempted arylcyanation of vinylsilanes to result in a Heck-type coupling

Scheme 36
scheme 36

Carbocyanation of norbornadiene

Intramolecular arylcyanation of simple alkenes proceeds smoothly by cooperative nickel/Lewis acid catalysis (Schemes 37, 38, and 39) [102104], whereas palladium catalysts have been shown to be useful for intramolecular cyanocarbamoylation (Scheme 40) [105109]. Proper chiral phosphorus ligands have been identified for each substrate structure to allow for the access to nitriles having a quaternary stereocenter with high enantiomeric excess. Some of these optically active nitrile products serve as synthetic precursors for biologically active natural products (Schemes 38 and 39) [110]. The transformation can be an alternative to asymmetric intramolecular Heck reactions, which have been applied extensively for natural product syntheses [111]. The intramolecular carbocyanation of alkenes can be advantageous in terms of the retention of cyano functionality in a product, while (pseudo)halogen functionalities in starting materials are lost in the Heck cyclization.

Scheme 37
scheme 37

Enantioselective intramolecular arylcyanation of alkenes catalyzed by nickel/BPh3

Scheme 38
scheme 38

Total synthesis of (–)-esermethole through enantioselecive intramolecular arylcyanation of alkenes

Scheme 39
scheme 39

Formal total synthesis of (–)-eptazocine through enantioselecive intramolecular arylcyanation of alkenes

Scheme 40
scheme 40

Enantioselective intramolecular cyanocarbamoylation of alkenes catalyzed by palladium

The reaction intermediates of the intramolecular arylcyanation are fully characterized by NMR and X-ray analyses (Scheme 41). The oxidative addition of an Ar–CN bond to nickel(0) assisted by AlMe2Cl takes place directly from the Lewis acid adduct of η 2-nitrile complexes at room temperature [103]. The oxidative adduct undergoes cyclization and reductive elimination to give another η 2-nitrile complex derived from the intramolecular C–C bond forming event upon heating at 60°C. Treatment of the η 2-nitrile complex with the starting aryl cyanide gives the initially observed η 2-nitrile complex through the exchange of nitrile ligand. The overall scheme shows that the rate-determining step of the intramolecular arylcyanation is either the exchange of bound phosphorus by the tethered alkene or the migratory insertion step rather than oxidative addition of Ar–CN bond.

Scheme 41
scheme 41

η 2-Coordination and oxidative addition of aryl cyanides to nickel(0) assisted by AlMe2Cl and its intermediacy in intramolecular arylcyanation of alkenes

4 Summary, Conclusions, Outlook

Synthetic reactions involving C–CN activation by metal catalysis are reviewed in this chapter. All the transformations presented herein provide the synthetic community with novel reaction modes of nitriles, and thus with new ideas and strategies for the syntheses of target molecules. In addition, the developments in novel metal catalysis for the activation of unreactive C–CN bonds should be of interest to the organometallic community. These catalyst designs will further stimulate the development of more active catalysts as well as novel catalysts for the activation of other unreactive bonds. At the moment, each reaction has a different maturity. Hydrodecyanation reaction covers a wide range of nitrile substrates including the most challenging alkanenitriles using nickel and rhodium catalysis, whereas cross-coupling reactions using nitriles as electrophiles are limited to aryl cyanides. Given the recent extensive studies on the cross-coupling of alkyl electrophiles, the use of alkanenitriles as alkyl electrophiles for cross-coupling should be an interesting and important direction for this particular transformation. Carbocyanation reaction also covers a broad range of nitriles for the addition across alkynes, whereas alkene-carbocyanation is totally underexplored except for intramolecular reactions. Because C(sp3)–CN bond-forming reductive elimination seems unfavorable compared with competitive β-hydride elimination of alkylnickel species, different metal systems and/or different mechanistic scenarios have to be envisioned to develop the potentially highly valuable C–C bond forming reaction.