Qiaodong Wen,
Ping Lu* and
Yanguang Wang*
Department of Chemistry, Zhejiang University, Hangzhou, 310027, P. R. China. E-mail: pinglu@zju.edu.cn; orgwyg@zju.edu.cn; Fax: +86-571-87952543; Tel: +86-571-87952543
First published on 23rd September 2014
C–C bond activation and cleavage have become more and more attractive in recent years for the great challenge of breaking C–C bonds and the great advantage of constructing new chemical bonds. Among them, the activation of inert C–CN bonds by transition metals has been widely investigated and a number of improvements have been achieved by many groups. This review mainly summarizes recent advances in this field, based on different kinds of reactions derived from C–CN bond activation, including cyanofunctionalization, cross coupling reactions and cyanation. Moreover, mechanistic studies involving transition-metal-catalyzed C–CN bond activation are also mentioned.
C–CN bond is relatively stronger and stable in comparison with C–X bond referring to have higher bond dissociation energy (BDE). BDEs of C–CN bonds in CH3–CN, CH2CH–CN, HCC–CN, and Ph–CN are 121.1, 132.1, 152.1, and 132.7 kcal mol−1,5 respectively. In comparison with these data, BDEs of C–X bonds in Ph–F, Ph–Cl, Ph–Br, and Ph–I are 127.4, 97.3, 82.7, and 66.9 kcal mol−1, respectively.6 Thus, the selective cleavage of C–CN bonds by using transition metal catalyst is a challengeable target and of great importance in organic synthesis. Reactions involving the cleavage of C–CN bonds can be simply categorized into three types according to the utility degree of nitriles. In cyanofunctionalizations, both R and CN units are presented in products (route a, Scheme 1). When “CN” is used as the leaving group, R–CN becomes an alternative substrate of R–X/R–OTf for cross coupling reactions which have been widely applied in making C–Si, C–P, C–B, C–H, as well as C–C bonds (route b, Scheme 1). With the disclosure of the safe cyanation reagents in the recent years, R–CN has also been developed for the operational benign cyanation reagents, such as benzyl nitrile,7 and acetonitrile.8 In these reactions, “CN” unit is successfully transferred from starting materials to target molecules (route c, Scheme 1).
In their mechanistic studies, the trans-Pd(CN)(CO2Et)(PPh3)2 (A) was isolated and proved to be the key intermediate by the addition of Pd(PPh3)4 to ethyl cyanoformate (7a) (Scheme 6). Single crystal analysis of the Me analogue of trans-Pd(CN)(CO2Et)(PPh3)2 (A) indicated that A possessed a square planar structure with two triphenylphosphines in trans positions.10a 1,2-Migratory insertion of trans-A to the double bond of norbornene (alkoxypalladation) followed by the reductive elimination gave the cyanoesterification product (8a) in syn-addition manner. Two groups (CN and COOEt) occupied the exo-positions instead of the endo-positions. That an alternative 1,2-migratory insertion of trans-A to norbornene (cyanopalladation generating B′) was ruled out because of the high affinity of cyano to palladium center.10b
Meanwhile, Nakao and Hiyama reported a nickel-catalyzed C–CN bond cleavage of cyanoformates (7) and their sequential cyanoesterifications of cumulative 1,2-dienes (9), affording various 2-(1-cyanoalk-1-yl)acrylates (10), as well as their regioisomers (11) (Scheme 7).11a By reversing the molar ratio of 7a to 9a to 1.0:1.2, 11a was formed predominantly.11b Except for cyanoformates (7), benzoyl cyanide (2b) was also suitable for this reaction under modified conditions and provided 12 with a little lower yield.
In their proposed mechanism, Ni catalyst inserted to C(sp2)–CN bond (Scheme 8). 1,2-Migratory insertion of C–Ni–CN (A) on 1,2-diene occurred on the less steric hindered terminal double bond with the ester group migrated to the central carbon of 1,2-diene. In this way, C was regioselectively formed through the possible complex B. The formed Ni(II) complex (C) could be stabilized by a π-allylnickel intermediate (D). By reductive elimination in two different types, two different products (10a and E-11a) were obtained. Z-11a was proposed to be obtained via equilibrium between intermediates E and F and followed by reductive elimination from intermediate F. The regioisomer E-11a was thought to be the thermodynamic product, whereas 10a was a kinetic product. This proposal was proved by the condition controlled experiments when the reaction temperature was increased to 100 °C and the amount of 10a was significantly decreased.
In 2011, Douglas demonstrated an intramolecular Pd-catalyzed cyanoesterification, accessing to functionalized lactones (13) in moderate to excellent yields (Scheme 9).12 The decarboxylation and decarbonylation byproducts were successfully suppressed by performing the reactions in microwave reactor at high temperature. The authors also hypothesized that the 1,2-migratory insertion was the product-determining step, rather than the step of C(sp2)–CN bond activation, because of the higher reactivity of electron-rich alkynes.
Aromatic substrate (16 and 18) provided indole derivatives 17 and 19, respectively. When cyanoamidation occurred on triple bonds, the stereoselectivity was largely depended upon the R group which was attached on the triple bonds. When R was phenyl or CH2OTBS group, yields were moderate and the products were in exclusively Z-form. When R was altered to butyl, a mixture of Z and E in a ratio of 69:31 was isolated in 97% yield. When cyanoamidation occurred on double bonds, a quaternary carbon center was created and 3,3′-disubstituted oxindole 19 was obtained almost quantitatively, which could be used as the key intermediate in the preparation of physostigmine.
All these reactions occurred in a same manner as the previous cyanofunctionalizations. Pd(0) was oxidatively added to amido cyanides (14, 16, and 18) to form amidopalladium cyanide intermediates (A) (Scheme 11). Thus, a sequential amidopalladation occurred on triple or double bonds to generate intermediates B. Finally, amido and cyano groups were installed into both ends of triple or double bonds by reductive elimination. When two reaction sites (amido cyanide and unsaturated bonds) existed in one molecule, an intramolecular reaction happened to afford lactams successfully.
Later on, Takemoto practiced the enantioselective synthesis of 3,3′-disubstituted oxindole (S-19) by finely tuning the catalyst, chiral ligand, additive, and solvent (Scheme 12).14 By screening these factors, a combination of Pd(dba)2, an optically active phosphoramidite, and a suitable Lewis base in decalin was optimized to be the best conditions to reach the effective enantioselective synthesis of 3,3′-disubstituted oxindole (S-19). It was also noticed that this combination was effective to a broad range of substrates and might be applied in the synthesis of more complicated molecules.
They proposed a possible mechanism (Scheme 14).15b Ni(0) catalyst inserted to Ar–CN bond of arenecarbonitrile (20) through oxidative addition, forming a Ni(II) complex (A). Then the Ni(II) complex (A) regioselectively added to the internal alkyne by 1,2-migratory insertion depending on the steric hindrance. A subsequently reductive elimination provided a cyanoarylation product (22) in syn-addition manner and regenerated the Ni(0) catalyst for the next catalytic cycle. An alternative 1,2-migratory insertion, arylnickelation instead of cyanonickelation, could not be ruled out although the repulsion between ligand and bulkier R2 group was apparent.
Besides aryl nitriles, allyl cyanides (23) could be used as the substrate for the oxidative addition of Ni(0), and provided the cyanoallylation products (24) when the resulting nickel species reacted with alkynes (Scheme 15).16 Thus, reaction between allyl cyanides (23) and oct-4-yne (21a) in the presence of 10 mol% Ni(cod)2 and 20 mol% P(4-CF3–C6H4)3 provided 6-substituted (2Z,5E)-2,3-dipropylhexa-2,5-dienenitriles (24) in regio- and stereoselective way. Partial isomerism of the double bond was observed. Synthetic utility of 2,5-dienenitriles was demonstrated by selective reductions. For instance, the resulting acrylonitriles could easily be reduced into acrylaldehydes and allyl alcohols by selecting suitable reductants without contamination of the reductive products derived from alkene unit.
Addition of α-siloxyallyl cyanide (25), which derived from the reaction between acrolein and TMSCN, onto oct-4-yne (21a) under the aforementioned reaction conditions afforded (Z)-5-formyl-2,3-dipropylpent-2-enenitrile (26) in 81% yield (Scheme 16). The silyl enol ether (29) was isolated and indicated to be the key intermediate by using t-butyldimethylsilyl (TBDMS) instead of TMS. To these cyanoalkylation reactions, terminal alkynes worked well and provided 2-substituted (Z)-5-formyl-pent-2-enenitriles (30) in high regioselectivity and high stereoselective. Furthermore, the formyl group could perform the aldehyde reactions and provided α-functionalized acrolein (27) and hepta-2,6-dienenitrile (28) effectively.
A possible mechanism was proposed by authors (Scheme 17). Ni catalyst firstly added to C(sp3)–CN bond of allyl cyanides (23, 25) via oxidative addition, giving a π-allylnickel intermediate (A). 1,2-Migratory addition of π-allylnickel intermediate alkynes, and followed by reductive elimination, acrylonitriles (24, 29) were obtained. Here the products having the bulkier group at a cyano-substituted carbon were much favored in consideration of the steric hindrance, especially to cases of terminal alkynes in which 2-substituted (Z)-5-formyl-pent-2-enenitriles (30) were afforded in excellent regioselectivity.
A breakthrough work was reported in 2007. Nakao and Hiyama demonstrated a Lewis-acid assisted Ni-catalyzed cyanoarylation of alkynes with nitriles, which efficiently improved the reactivities of some electron-rich aryl cyanides (20a), comparing to the previous works (Scheme 18).17a Moreover, the amount of the catalyst and the ligand was significantly decreased. By surveying the substituted group on aryl cyanides, it was indicated that selective activation of Ar–CN bond over Ar–Br/Cl was remarkable. Bromo or chloro (22b) was survived after reaction. Moreover, sterically hindered 2,6-dimethyl benzonitrile (20c) participated the reaction and the Ar-CN bond was activated if the amount of the catalyst, ligand, and Lewis acid was magnified 5 times, the reaction time was prolonged to 134 h, and the reaction temperature was raised to 100 °C. The corresponding product (22c) was obtained in 78% yield. Another interesting case was the cyanoarylation of 1-methyl-1H-indole-3-carbonitrile (31) over 21a. Activation of Ar–CN occurred prior to the activation of Ar–H at 2-position of indole in the presence of Lewis acid.18
In the aid of Lewis acid, cyanoalkenylation was also successfully achieved by fine tuning ligand and Lewis acid. Thus, a series of conjugated dienenitriles were obtained in good to excellent yields (Scheme 19). Alkene configuration of trans-cinnamonitrile (33) was remained after reaction and the resulting dienenitrile (34) could be transferred into pyridine (35) by a sequence of reduction, 6π-electrocyclization, and aromatization. 2-Menzylidenemalononitrile (36) participated the reaction and provided dicyano-substituted 1,3-diene (37) in 81% yield. In this case, a selective activation of Ar–CN bond which appeared trans to phenyl group was demonstrated. Further, other cyanofunctionalizations, such as aforementioned cyanoesterification17b and cyanoamidation17c were all optimized in the aid of Lewis acid.
Finally, C(sp3)–CN bond in acetonitrile was activated. Reaction of acetonitrile (38) and oct-4-yne gave trisubstituted acrylonitrile (39a) in 71% yield (Scheme 20). That methyl group came from acetonitrile albeit AlMe3 was confirmed by using CH3CN-d3. Other nitriles, such as 2-(trimethylsilyl)-acetonitrile and propiononitrile, provided the corresponding addition products, but in relatively lower yields. In case of propiononitrile (40), it provided the syn-adduct (39b) in 24% yield only. One of the by-products was (E)-2-propylhex-2-enenitrile (41) (Scheme 21). Oxidative addition of Ni(0) to propiononitrile leading to the formation of ethylnickel cyanide (A), which underwent β-elimination to generate highly active species, HNiCN (B). In the presence of oct-4-yne, hydrocyanation occurred and provided the by-product 41.
The dramatic effects of Lewis acids could be contributed to the coordination between R–CN and Lewis acid which accelerated the rate of oxidative addition of Ni(0) to R–CN bond (Scheme 22). Regioselectivity was ascribed to the repulsion between the bulkier R2 and R when the unsymmetrical alkyne and nickel species coordinated. Ultimately, the bulkier R2 group and cyano group (CN) bonded to the same carbon of alkynes. Moreover, these reactions occurred in exclusive syn-addition manner, the trans-adducts came from the isomerism of syn-adducts in the presence of Lewis acid or ligand, or under heat.
Short after, Nakao and Hiyama reported cyanoalkylation of alkynes with arylacetonitriles (42), also catalyzed by nickel and Lewis acid (Scheme 23).19 Various arylacetonitriles could undertake this reaction in good yields, including some heteroaryl derivatives. As arylacetonitriles, β-elimination resulted from the normal alkanenitriles, such as aforementioned byproduct (41), could be effectively diminished. To the symmetrical alkynes, addition occurred in an exclusive syn-fashion. However, to the unsymmetrical alkynes, poor regioselectivities were observed except for the cases where the steric hindrance could be differentiated, such as methyl versus tert-butyl, methyl versus TMS. The bulkier group always linked to carbon with CN attached in the final.
To make a clear outcome of the mechanism, the authors selected optically active (S)-α-phenylpropionitrile (44) as the substrate and detected all the byproducts. It was worth noting that the recovered starting materials showed 80% ee and the product did not undertake racemization under their conditions. These facts indicated that the absolute configuration of the chiral carbon retained after these sequential steps, including the oxidative addition, 1,2-migratory insertion, and reductive elimination. The partial loss of the optical purity could be ascribed to the equilibrium between the β-elimination and its reversed 1,2-migratory insertion as illustrated in Scheme 24.
In 2010, Nakao and Hiyama reported heteroatom-directed cyanoalkylation of alkynes (Scheme 25).20 The heteroatoms, like nitrogen atom, played a significant role in suppressing the aforementioned unexpected β-hydride elimination via the intramolecular chelating of nitrogen with nickel center. The amino effect at the γ-position of butyronitrile (46a) was impressive. When the amino was on the β-position of propiononitrile, only trace amount of the syn-adduct was detected under the same reaction conditions. It was also noticeable that when the amino group was on δ-position of pentanenitrile, with one more methylene unit between pyrrolidine and cyano (46b), cyanoalkylation occurred, but on the γ-position of the pyrrolidyl group. This selectivity was also observed in case where 5-pyrrolidyl hexanenitrile (46c) was used. In this case, a longer reaction time was required and the hydrocyanation by-product (41) was isolated in 9%.
The above experiments indicated that a 5-membered azanickelacycle (C) was the key intermediate (Scheme 26). With one more CH2, a 6-membered azanickelacycle (F) (or even 7-membered in case with two more CH2) could be converted to 5-membered azanickelacycle (C) by a sequence of β-hydrogen elimination and hydronickelation. This proposed mechanism properly explained the fact that the amino-substituted butyronitrile, valeronitrile and hexanenitrile could undertake cyanoalkylation smoothly. Similar to the above mentioned cases, the syn-adducts formed having a bulkier group and CN bonded to the same sp2-carbon when unsymmetrical alkynes were used.
Other heteroatom (O, S) in the appropriate position of alkanenitrile (48, 50) could also function as the directing heteroatom via its coordination to nickel center and afford the corresponding cyanoalkylation products (49, 51) (Scheme 27).
Owning to homo- and/or cross-cyclotrimerization in nickel catalyst system, it was uneasy for alkynyl cyanide (52) to undertake the catalytic insertion of Ni(0) to C(sp)–CN bond. Nevertheless, Nakao and Hiyama reported nickel/BPh3-catalyzed cyanoalkynylation of alkynes, accessing to conjugated enyne structures (53) (Scheme 28).21 With the aid of Lewis acid, the trimerization products (54) were effectively inhibited.
After investigating the substrate scope, they found that bulky silyl substituented alkynyl cyanides (55) might suppress the homo- and/or cross-cyclotrimerization of the nitriles, resulting in a lower requirement of catalyst amounts and approached excellent yields of 56 (Scheme 29). Diynyl cyanides were also suitable for this reaction. 1,2-Dienes (57) reacted with alkynyl cyanides (55) under the same conditions, giving conjugated enynes (58) in moderate to good yields. The regioselectivity was largely depending upon the substituted group attached on 1,2-diene.
The oxidative insertion of C(sp)–CN bond with Ni(0) catalyst assisted by BPh3 was thought to be the initial step in the proposed mechanism to form key intermediate A (Scheme 30). Followed by the coordination with alkynes or 1,2-dienes, respectively, complexes B and D were formed. In case of the alkyne as substrate, the alkynyl group migrated from the nickel center to the less hindered carbon of the coordinated alkyne to generate C. In case of 1,2-diene as substrate, the alkynyl migrated from nickel center to the central carbon of 1,2-diene to generate the allylnickel species E. Finally, the reductive elimination of C or E took place and provided the addition products, 56 and 58 respectively, with both high regioselectivity and high stereoselectivity.
In 2011, Kurahashi and Matsubara reported an intermolecular cycloaddition of o-arylcarboxy benzonitriles (59) and alkynes (21) through Ni-catalyzed cleavage of C–CN and C–CO bond, affording coumarins (60) in good yields (Scheme 31).22 It was worth noting that aryl cyanide (20d) was isolated, as well as the compound (22d) formed from its carbocyanation with alkyne. Similarly, o-arylamide-substituted benzonitrile (61) could also undertake this reaction with oct-4-yne, which gave the quinolinone (62) in 83% yield.23
In 2013, Nakao and Hiyama chose polyfluorobenzonitriles (20e, 20f) to conduct cyanoarenylation with alkynes, affording a series of corresponding polyfluoroarylcyanation products (22e, 22f) by applying Ni(cod)2/DPEphos/BPh3 catalytic system in the solvent of cyclopentylmethylether (CPME) (Scheme 32).24 Under their reaction conditions, C–CN bond activation was prior to C–H and C–F bond activation, so the resulting cyanoarenylation products could be obtained smoothly without the interference with the byproducts derived from C–H and C–F bond activations. Of note, the existence of fluorine atoms on the arenes promoted the C–CN bond activation, driving polyfluorobenzonitriles more reactive than non-fluorine substituted aryl nitrile which was demonstrated by competition experiment between pentafluorobenzonitrile and benzonitrile. Moreover, the dramatic effect of BPh3 and DPEphos was clearly illustrated by the single crystal structure of the nickel insertion intermediate (cis-A) which was responsible for the subsequent 1,2-migratory insertion of ArF–Ni–CN on alkyne. Combination of Ni(cod)2, DPEphos, and BPh3 in CPME was also used for cyanoarenylation of terminal alkynes and alkenes. As an interesting example, selective activation of C–H bond albeit C–CN bond was achieved by using PCyp3 as ligand. Thus, compound (63) was obtained from 22f in 90% yield with high chemoselectivity.
Meanwhile, Nakao and Hiyama also reported the intramolecular cyanoarenylation of various specially designed alkenes with Ni(cod)2/AlMe2Cl catalyst. A number of substrates tolerated the reaction conditions (Scheme 34).26 Amino and silyl tethers provided the corresponding cyclized products 67 and 68 in 86% and 92% yields, respectively, by choosing proper ligand. Insertion of olefin into Ar–CN bond was found to be in a 5-exo-trig fashion. The new constructed ring could be enlarged to six- or seven-membered ring, as indicated by the representative compounds 69 and 70. The cyanoarenylation reaction proceeded in a syn-addition manner which was clearly shown by the diastereoselective preparation of compounds 72a and 72b (Scheme 35).
Not only the reaction could be conducted in high diastereoselectivity, but also in high enantioselectivity by using a chiral ligand, such as (R,R)-i-Pr-Foxap (Scheme 36). This asymmetric synthesis had been successfully applied in the preparation of (−)-esermethole.
Mechanism of this intramolecular cyanoarenylation was proposed based on the isolation of key intermediates 75 and 76 (Scheme 37). Structures 75 and 76 were established by their single crystal analysis. Reaction of stoichiometric 64, Ni(cod)2, P(n-Bu)3, and AlMe2Cl at room temperature immediately provided η2-nitrile complex 75. Existence of Lewis acid facilitated the oxidative addition of Ni(0) to Ar–CN bond and afforded complex 76 after 6 h. 1,2-Migratory insertion onto intramolecular olefin provided η2-nitrile complex 77 via intermediate A. Finally, cyclized product was obtained via ligand exchange.
Above all, a number of cyanofunctionalizations on unsaturated alkynes, alkenes, 1,2-dienes were disclosed, including cyanoeserification, cyanoamidation, cyanoacylation, cyanoarenylation, cyanoalkenylation, cyanoalkylation, and cyanoalkynylation based on the carbon type bonded to CN group in the substrates. All of these reactions were of great synthetic value because they simultaneously installed two groups into one molecule with 100% atom efficiency. More noteworthy, reactions proceeded in syn-addition fashion with high stereoselectivity.
The decyanative silylation was well extended to allyl cyanides. Thus, allyl cyanide (23a) furnished allyl silane (79) in 59% yield (Scheme 39). A relative lower reaction temperature was applied for this transformation due to the relative facile cleavage of C–CN bond and the formation of the stable π-allyl rhodium complex.28 However, when benzyl cyanide (42a) was used as the substrate, a higher reaction temperature was applied for complete conversion of benzyl cyanide. In this case, enamine formation (81) was isolated in a certain amount. The enamine byproduct (84) became dominant when 2,2-diphenylacetonitrile (82) was used.
It was also noteworthy that even the unactivated alkyl cyanide could be used for the decyanative silylation when triisopropyl phosphite was added (Scheme 40). Thus, 4-phenylbutanenitrile (85) provided corresponding silylated product (86) in 48% yield with 34% recovery of the starting material. In this case, the byproduct resulting from the β-hydrogen elimination of the possible alkyl metal intermediate was not observed.
In their proposed catalytic cycle (Scheme 41), silylrhodium species (A) was generated via σ-bond metathesis between Rh–X and Si–Si bonds. Alternatively, A could be formed via an oxidative/elimination process. Migration of silyl of Rh–Si (A) to the nitrogen atom of nitrile generated η2-iminoacyl complex (B) as the key intermediate in this catalytic cycle.27b,c Subsequent migration of aryl/alkyl to rhodium center provided aryl/alkyl rhodium species (C) coordinated with silyl isocyanide. In the presence of disilane (75), σ-bond metathesis between C and disilane (75) gave rise to the silylated product (R–Si) and regenerated Rh–Si (A) for completing the catalytic cycle. Meanwhile, silyl isocyanide was excluded and isomerized into thermodynamically stable silyl cyanide in the final. When benzyl cyanide (42a) was used, two possible ways leading to the formation of enamine (81) was proposed. σ-Bond metathesis between B and disilane (75) produced D and regenerated A. Intramolecular migration of silyl from imino carbon to nitrogen via aza-Brook-type rearrangement formed aminocarbene (E). Through this path, enamine was finally obtained via the carbene insertion to adjacent C–H bond. Alternatively, β-hydrogen elimination of B, and followed by hydrorhodation across CN bond of N-silyl ketenimine (F) generated G. As a result of σ-bond metathesis between G and disilane (75), enamine (81) was afforded in a certain amount.
Further investigation of decyanative silylation revealed that 1,5-rhodium migration step might be involved after the Rh-catalyzed C–CN bond cleavage if the substrates (87) were rationally designed (Scheme 42).29 Thus a mixture of decyanative silyated isomers (88 and 89) was obtained after the silylation occurred on aryl rhodium species (A and B) before and after 1,5-migration. Product ratio was highly dependent on the acidity of Ar–H which was putatively to be activated. With aryl possessing the electron donating group (R2 = OMe), the direct decyanative silylation product was dominant. In a contrast, with aryl bearing the electron withdrawing group (3,4,5-trifluorophenyl), the 1,5-migrated product was isolated as a major product. 1,5-Mgiration was indicated to be irreversible by conducting a comparative experiment and the rate of 1,5-migration resulted in the product distribution. Migration distance between the metal center and the Ar–H was suitable to form the six-membered metallacyclic intermediate (or transition state), no 1,4-, 1,6- or 1,7-migration was observed. Similar 1,5-migration was observed in case when diboron was used instead of disilane in the decyanative borylation.
In order to clarify the mechanism, a controlled experiment was conducted (Scheme 44). By using hexamethyldisilane or tetraphenyldiphosphine instead of Me3Si–PPh2 under the same reaction conditions, the former reaction did not occur, while the later one furnished corresponding phosphinated product in 45% yield. Thus, the possible Si-assisted C–CN bond cleavage in the presence of nickel catalyst was excluded.
A plausible mechanism was proposed (Scheme 45). Ni(II) was firstly reduced to Ni(0) (A) in the presence of base for the subsequent oxidative addition to Ar–CN bond via the complexes (B and C) between aryl nitrile and nickel. As a result, aryl nickel cyanide (D) was formed. Ligand exchange through two possible routes as proposed by the authors provided intermediate (E). Subsequent reductive elimination of E produced the phosphinated product and regenerated Ni(0) (A) for next catalytic cycle.
The mechanistic pathway supposed that the diboron (94) played an important role in the cleavage of C–CN bond (Scheme 47). σ-Bond metathesis between Rh–Cl bond and B–B bond generated Rh–B bond and B–Cl bond which was confirmed by 11B NMR analysis. This step was the same to the previous metal-silicon bond formation. Addition of Rh–B species (A) across nitrile triple bond resulted in the formation of adduct (B) which isomerized to C for the sterically favored syn-β-aryl elimination. As a result, R–Rh (D) was formed and borylated via σ-bond metathesis with diboron. Although β-carbon elimination was proposed in Chatani's report,31a a mechanism including the sequential 2,1-insertion and deinsertion could not be excluded.31c,d
In 2009, Tobisu and Chatani reported a protocol for Rh-catalyzed reductive cleavage of C–CN bond with hydrosilane as the reducing agent, converting varying nitriles to the corresponding alkanes or arenes (Scheme 48).32 They selected [RhCl(cod)]2 as the rhodium catalyst, P(OBu)3 as the ligand and triisopropylsilane as the reducing agent, affording decyanated products with high efficiencies. A variety of functional groups tolerated the reaction conditions, such as dimethylamino, methoxy, ester, and boronic ester. A deuterated product (98) was also gained when the corresponding deuteriosilane was used in the reaction. Besides of aryl nitriles could be decyanated, benzilic nitriles as well as normal alkyl cyanides could also be decyanated in excellent yields without the contamination from β-hydrogen elimination.
In 2013, Maiti demonstrated nickel catalyzed decyanation of nitriles selecting tetramethyldisiloxane (TMDSO) as the hydride source (Scheme 49).33 In the catalytic system of Ni(acac)2, PCy3, and AlMe3, aryl and aliphatic cyanides reacted with TMDSO and decyanated to arenes and aliphatic hydrocarbons, respectively. A variety of functional groups on aromatic ring suffered this reaction condition, such as methoxy, ester, and carbonyl. Cleavage of Ar–SMe bond occurred prior to the cleavage of Ar–CN bond. Heterocyclic cyanides also preceded smoothly under the reaction conditions. The deuterated product was isolated when Ph2SiD2 was used as the reductant. The trans-[Ni(PCy3)2(CN)2] was isolated from the reaction, but failed to catalyze the decyanation, indicating that it was the cause of catalyst deactivation.
Short after, Maiti reported nickel-catalyzed decyanation of aryl and aliphatic cyanides using hydrogen gas as the hydride source (Scheme 50).34 The C–CN bond was efficiently activated in assistance of AlMe3, which was also responsible to consume the in situ produced HCN gas. It was worthy to note that aliphatic cyanides (103), possessing β-hydrogens, were also applicable for this decyanation without the formation of the byproduct arising from β-hydrogen elimination. A variety of functional groups tolerated the reaction conditions, such as MeO-, HO-, F-, ester, amide, ketone, as well as the tethered olefin.
In 2013, Enthaler reported decyanation of aryl and alkyl cyanides catalyzed by their pre-prepared nickel complex and employed tert-butylmagnesium chloride as the hydride source (Scheme 51).35 Most substituent groups could tolerant the reaction conditions except for thioether and several halogens (Br and F). In these cases, the cleavage of C–S and C-X bonds was observed. Hydrogen atom in the product came from Grignard reagent which was further confirmed by using C2D5MgBr. Finally, they proposed a possible mechanism, in which C–CN bond was supposed to be cleaved through a single-electron transfer (SET) step and a radical procedure (Scheme 52).
Scheme 52 Proposed mechanism for Ni-catalyzed decyanation of aryl nitriles with tert-butylmagnesium chloride. |
Later in 2003, Miller and Dankwardt reported the nickel catalyzed cross-coupling of aryl nitriles with the modified alkyl and alkenyl Grignard reagents, also providing corresponding styrene and alkyl arene derivatives in good yields (Scheme 54).37 By the addition of either t-BuOLi or PhSLi to the Grignard reagents, the nucleophilic addition of Grignard to aryl nitriles could be effectively diminished. Undoubtedly, this procedure contained a nickel catalyzed activation of the C–CN bond.
In 2009, Shi reported a cross coupling of aryl nitriles with aryl/alkenyl boronic esters through Ni-catalyzed C–CN bond cleavage (Scheme 55).38 With the aid of t-BuOK as a base and CuF2 as an additive, aryl/alkenyl boronic esters coupled with aryl nitriles and provided biaryls in moderate to good yields. A variety of aryl nitriles were tested, including the aryls with electron-withdrawing and electron-donating groups.
Considering the relative reactivity of C–Cl, C–CN, and C–OMe bond activations in cross-coupling reactions, the authors successfully prepared polysubstituted benzene (111) with different aryl groups by orthogonal design (Scheme 56).
In 2012, Wang developed Ni-catalyzed cross coupling of aryl nitriles and arylmanganese reagents through the cleavage of C–CN bond (Scheme 57).39 The arylmanganese reagents, which had been commonly used in many cross-coupling reactions, showed high efficiency in these cross-coupling reactions. A number of aryl nitriles and arylmanganese chlorides could be used for coupling. In comparison with the Grignard reagents or boronic esters used above, this reaction had great advantages, such as mild conditions, and no requirement of base and additive.
During the process of the rhodium catalyzed decyanative silylation of aryl cyanides as previously mentioned in Scheme 41,28 Ar–Rh was formed as a key intermediate in the catalytic cycle. By designation of the structure of Ar–Rh with a tethered electrophile, such as aryl chloride, C–C bond could be formed intramolecularly. Thus, substituted dibenzofurans (113) and carbazole derivatives (114) were well synthesized by the Rh-catalyzed Si-assisted C–CN bond cleavage and a sequential coupling with intramolecular Ar–Cl (Scheme 58). Besides of these, the benzylic C–CN bond could be arylated with intramolecular Ar–Cl. Thus, fluorene (116) was obtained in 60% yield, accompanying with the formation of the decyanated silylation product (117) in 15% yield. This methodology used the aryl chloride instead of organometallic reagent to realize the cross coupling between C–CN and C–Cl bonds. From this point of view, it was of synthetic value because aryl chlorides were more stable and more widely existed.
Considering the selective C–Br and C–CN bond activation by using suitable transition metal catalyst, 1,4-divinyl phenylene (122), with different substituted groups at ends could be approached (Scheme 60).
In the proposed mechanism (Scheme 61), Rh catalyst transferred to an active silylrhodium species (A) by exclusion of TMSCl, and then Si-assisted C–CN bond cleavage occurred and arylrhodium species (C) was subsequently generated via η2-iminoacyl complex (B). During this process, silyl isocyanide was formed. Finally, a Mizoroki–Heck type reaction between intermediate (C) and vinylsilane occurred to form the desire product via addition and elimination, accompanied by the generation of rhodium hydride species (E). Subsequent reaction between rhodium hydride species (E) and disilane regenerated silylrhodium species (A) for the next catalytic cycle. In some cases, silylated product was obtained from the reaction between arylrhodium species (C) and disilane which was described in detail in the section of the coupling with heteroatoms.
Thus, a number of σ-bonds, including C–Si, C–P, C–B, C–H, and C–C bonds, could be feasibly constructed via transition metal catalyzed C–CN bond cleavage and the subsequent cross coupling. Of particular note, advantageous of R–CN were obvious. First, they could be treated as a succedaneum for R–X/R–OTs in cross coupling. Second, based on the selectivity and the reactivity of C–CN, C–X, and C–O bonds, these protocols enabled orthogonal cross coupling possible for having higher conjugated system. Third, the decyanation reaction took the advantage of “CN” properties, such as α-C–H acidity, ortho-directing effect, and electron-withdrawing nature, and could be utilized in the synthesis of challengeable compounds.
In their proposed mechanism (Scheme 63), light triggered the exclusion of CO from Cp(CO)2FeMe and left the space for oxidative addition to Si–H bond in Et3SiH. After reductive elimination, a key intermediate (C), with 16 valence electrons, was generated. Coordinating this species with acetonitrile in η2-fashion (D), followed by silyl-assisted iron migration insertion, iron isocyanide complex (E) was formed. Exclusion of silyl isocyanide (Et3SiNC) from E regenerated A for completing the catalytic cycle. As a result, Et3SiCN was formed after isomerism, while Et3Si–SiEt3was formed from the reaction between key intermediate (C) and Et3SiH via addition/elimination sequence.
Scheme 63 Proposed mechanism for the Fe-catalyzed Si-assisted photocyanation of silane with acetonitrile. |
The general Pd-catalyzed cyanation of aryl halides contained three steps, oxidative addition, ligand exchange, and reductive elimination (Scheme 65). The key obstacle for this general one was the poison of the palladium catalyst which resulted from overloading of cyanide anion to palladium center. In the cyanation with benzyl cyanide, after the oxidative addition of Pd(0) to aryl halides, B was formed and subsequently coordinated with benzyl cyanide to generate key intermediate (C). Thus, C–CN bond was activated and cleaved by the assistance of halide. Finally, reductive elimination afforded aryl nitriles and regenerated Pd(0) for completing the catalytic cycle.
Subsequently, Shen developed palladium catalyzed cyanation of aryl halides, introducing ethyl cyanoacetate as the cyanating reagent (Scheme 66).42 The combination of Pd(OAc)2, TMEDA, DPPE, Na2CO3 and KI could make the best transformation from aryl halides to aryl nitriles. A broad substrate scope was presented, including the acetyl, hydroxy, amino, ester, and amide. 5-Bromo-1-mthylindole proceeded the reaction affording the desired 5-cyano-1-mthylindole in 52% yield without the byproduct resulting from the C–H activation of 3-position of indole. The major defect for this transformation was the competitive aromatic nucleophilic substitution of enolate anion on the highly electron-deficient aryl halides. It happened spontaneously in basic environment without the participation of palladium. The authors claimed the complex of the ethyl cyanoacetate enolate and the palladium was the key intermediate in this reaction. It could form the palladium-cyano species and facilitate the cleavage of C–CN bond.
In 2012, Zhou reported Cu-catalyzed cyanation of aryl iodides using malononitrile as the cyanide source (Scheme 67).43 After screening the reaction conditions, Cu(OAc)2, 1,10-phenanthroline, t-BuONa were all the best choices, along with KF as a beneficial additive. The intermediate, a complex of [Cu(phen)(CN)2] (126), was formed by proceeding the reaction in the absence of aryl iodides. It was indicated that [Cu(phen)(CN)2] could react with the aryl iodides to form the aryl nitriles in good yields with or without KF. Thus, a copper-catalyzed C–CN bond cleavage of malononitrile assisted by KF might be involved in the reaction pathway, followed by the copper-catalyzed cyanation of aryl iodides with intermediate complex.
Later, Shen subsequently demonstrated a copper-catalyzed cyanation of 2-phenylpyridines and the 1-pyridinylindoles with acetonitrile (Scheme 68).8c It was hypothesized by the authors that hexamethyldisilane was employed as a Lewis acid to weaken the C–CN bond for its facile cleavage. Thus, a SN2-type nucleophilic substitution occurred on the methyl of acetonitrile and resulted in the formation of copper cyanide species for the subsequent cyanation.
After further investigation, a different pathway was illustrated (Scheme 70) in comparison with previously Pd-catalyzed pathway (Scheme 65). An aerobic C–H oxidation of benzyl cyanide in the catalyst system was supposed to produce benzaldehyde cyanohydrin which was responsible for slowly releasing cyano anions for the subsequent Cu-catalyzed cyanation of arenes. It was pointed that matching two Cu-catalyzed reactions rates, those were Cu-catalyzed C–H aerobic oxidation of benzyl cyanide and heteroatom-directed ortho-C–H functionalization, was the key issue for approaching a successful transformation.
Afterwards, benefitted from the advantageous of benzyl cyanide as cyanation reagent and inexpensive copper salts as catalysts, cyanations of aryl halides,7c aryl boronic acids,7d and indoles7e,f were realized with nice substrate diversity and high yields (Scheme 71). Of particular note, by using the standard reaction conditions for cyanation of aryl halides, cyanation of 124 provided 125a in 50% yield without the observation of 125b.
Scheme 71 Selected examples for cyanation of aryl halides, aryl boronic acids, and indoles with benzyl cyanide. |
In 2012, a copper-catalyzed oxidative cyanation of aryl halides with acetonitrile as the cyanation reagent was reported by Li (Scheme 72).8a They employed Ag2O/air as the oxidant and triphenylphosphine oxide as ligand, converting various aryl iodides and bromides to aryl nitriles in moderate to good yields. Cu(I) species was thought to be responsible for the cleavage of Ar–X to form Ar–Cu species, followed by coordination with acetonitrile. In assistance of Ag2O/O2, acetonitrile was oxidized and decomposed to produce formic acid and “CN” for the following cyanation. Thus, the cleavage of C–CN of acetonitrile belonged to the Cu-catalyzed oxidative cleavage (Scheme 73).
In 2013, Zhu also introduced acetonitrile as the cyanide source to the copper-mediated C2-cyanation of indoles and pyrroles without the participation of phosphine ligand (Scheme 74).8b With the pyrimidyl or pyridinyl on the N atom as the directing group, cyanation occurred selectively at C2-position. AgOAc and the O2 atmosphere were used as oxidants in this reaction, also involving copper-mediated C–CN cleavage of acetonitrile via α-C–H oxidation.
Throughout the cyanation reactions by using organic nitriles as “CN” surrogates, both cleavage and formation of C–CN bond were involved. Although organic sources had significantly advantageous over traditional used metal cyanides, such as non-toxic and soluble in common organic solvents, only few organic nitriles had been hitherto examined and developed. Further investigation in this area for easy synthesis of aryl nitriles is still highly demanded.
Although above mentioned cyanofunctionalizations, cross-coupling reactions, and cyanations have well presented the utilities of nitriles in organic synthesis in recent years, the huge potentials of using the abundant, stable, and low-cost R–CN in feasibly making various C–C, C–H, as well as C-heteroatom bonds in organic synthesis still need to be explored and developed.
This journal is © The Royal Society of Chemistry 2014 |