Kingsley
Cavell
School of Chemistry, Cardiff University, Main Building, Park Place, Cardiff, UK CF10 3AT. E-mail: cavellkj@cardiff.ac.uk
First published on 14th October 2008
N-Heterocyclic carbenes (NHCs) are commonly regarded as strong donor ligands that are valuable in coordination chemistry and catalysis. Many reports describing this aspect of their chemistry have been published. The alternative view of NHCs as reaction intermediates has been little explored, and yet excellent examples exist and will be reviewed in this perspective. Group 10 hydrocarbyl complexes of NHCs [(R)(NHC)ML2 where R = H, alkyl, aryl, acyl; M = Ni, Pd, Pt] undergo a facile reductive elimination to generate M(0) and R-substituted-azolium salt as products. On the other hand, 2-Hazolium salts will oxidatively add to M(0) complexes to afford reactive NHC–M–H compounds, suitable as catalysts for selected reactions. Combining the oxidative addition and reductive elimination steps into a single cycle, in the presence of an alkene, provides of a novel and potentially exciting, atom efficient catalytic C–C bond forming process for the substitution, and annulation of heterocyclic rings.
![]() Kingsley Cavell | Kingsley Cavell was appointed Professor of Inorganic Chemistry, Cardiff University in 2000, where he is currently the Head of School and Group leader of the Metals in Synthesis Research Group. He received his PhD from La Trobe University (Melbourne, Australia) and then undertook post-doctoral studies at the University of Manchester (with Professor Hank Skinner and Dr Geoff Pilcher) and at the University of Amsterdam (with Professor Dr Kees Vrieze). This was followed by a period with the CSIRO (Div. of Materials Sci. and Tech., Melbourne) (1980–1986), and the University of Tasmania (1986–2000). He has been Guest Professor at the Institut für Chemie und Petrochemie der RWTH Aachen, Germany (1989/90 and 1994; with Professor Dr W. Keim), and in the Department of Chemistry, Huazhong University of Science and Technology, Wuhan, PR China (1997; with Professor Li Guangxing). He was awarded the 2001 RSC Industrially-Sponsored Award in Homogeneous Catalysis, sponsored by Synetix. His research is focussed on organometallic chemistry, homogeneous catalysis and on elucidation of reaction mechanisms relating to catalytic processes; a key component of the research has been directed at the study of heterocyclic carbenes as ligands and as reaction intermediates. |
In the current literature NHCs have been viewed almost exclusively as strong donor ligands for the synthesis of metal complexes. Therefore, not surprisingly, NHCs have been widely applied in catalysis, and with some success, particularly in selected reactions such as olefin metathesis.3 Such powerful ligands should in principle provide stable catalyst systems, and in some cases examples of carbene complexes have been used at very high temperatures (120–180 °C or higher) offering some promise in this respect.
Early investigations into the application of NHCs in catalysis were initiated in the 1970s.4 However, it was only after the isolation of the first free NHCs,5 and the influential catalytic studies by Herrmann that the field really emerged.6 Nevertheless, NHCs have not been universally applicable as ligands in catalysis, and it soon became apparent that despite strong metal–NHC bonding interactions, fundamental “problems” exist in the chemistry of these ligands; NHCs are susceptible to loss by reductive elimination, leading to decomposition of the catalyst, in some instances before effective catalysis can take place.7,8 Unlike their role as spectator ligands in catalysis, an alternative function of NHCs, as intermediates in chemical reactions, which is more in keeping with the role of “traditional” carbenes in synthesis and Schrock type carbenes in catalysis, is virtually unexplored. In the following discussion the view that NHCs act only as simple donor ligands towards transition metals will be clearly dispelled, and a facile coupling reaction between NHCs and olefinic moieties is presented. The application of NHCs as reaction intermediates, or “substrate”, which finish up as products, will be described.
Fundamental aspects of the chemistry of NHCs and related imidazolium salts are reported, and subsequently how this reaction behaviour has been utilised to construct a novel atom efficient C–C coupling reaction via C–H activation. This perspective is not intended as a comprehensive review and hence there has been a selective reporting of the literature most relevant to the topic under discussion.
![]() | (1) |
![]() | ||
Scheme 1 |
A comprehensive experimental and theoretical study was able to demonstrate that the observed reaction was a concerted reductive elimination process; kinetic studies were consistent with reductive elimination, and DFT calculations supported a reductive elimination mechanism.7 Calculations on a “real” system indicated an “associative” pathway with a small activation barrier (14.1 kcal mol−1) and overall endothermic reaction (−9.2 kcal mol−1) (Scheme 2). In the transition structure (TS) the change in orbital populations is in accord with reductive elimination; there is mixing of Ccarbene p(π), the CMe and the Pd d orbitals. This mixing of orbitals allows the Ccarbene–CMe bond to form (Fig. 1). The plane of the NHC ligand is at a substantial angle to the coordination plane (commonly 60–80°) and hence the Ccarbene p(π), which is perpendicular to the NHC plane is correctly oriented to interact with the Me group (Fig. 1). The d orbital with the appropriate orientation for the three-centre interaction is the dxy orbital. In the encounter complex the population of the dx2−y2 orbital increases along with the population of the 5s orbital; the Ccarbene–CMe bond is fully occupied with no d involvement and the N p(π) → Ccarbene p(π) is purely p in character, as in the starting complex. The p(π) occupancy of the Ccarbene has increased due to greater delocalization of the imidazolium cation.
![]() | ||
Fig. 1 |
![]() | ||
Scheme 2 |
The competitive interplay between decomposition on one hand, and effective catalysis on the other is readily seen from an investigation of a stoichiometric Heck reaction between n-butyl acrylate and the complex (tmiy)2Pd(Ar)I (Ar = 4-nitrobenzene), at sub-ambient temperatures. The reaction beautifully illustrates this competition between reductive elimination and Heck coupling (Scheme 3).12 Under these conditions, both Heck chemistry and reductive elimination are occurring, and the relative degree of each is temperature dependent; at lower temperatures reductive elimination predominates, however, as the temperature is raised and as catalytic conditions are applied the Heck reaction proceeds almost exclusively. In the former case the coordinated NHC is behaving as a reaction intermediate, whereas in the latter it is a “non-participative” spectator ligand. Interestingly, the decomposition products isolated from the reaction (products in boxes) provide a clear steer on the mechanism of the Heck reaction. Furthermore, from Scheme 3 it is evident that as well as carbene–aryl reductive coupling carbene–hydride reductive elimination also occurs, to give 2-hydridoimidazolium salt, 7.
![]() | ||
Scheme 3 |
In subsequent studies it was demonstrated that reductive elimination is also extremely facile for carbene–Pd–acyl complexes, yielding 2-acylimidazolium salts and Pd(0).14 The product distribution in such cases was shown to depend on the structure of the complexes from which reductive coupling took place (Scheme 4); a cis-(NHC)/(Me) arrangement (pathway A) gave 2-methylimidazolium as the major product, some 2-acylimidazolium was also observed; a probable cis-Me/CO structure (pathway B) gave 2-acylimidazolium salt as the major product, and 2-methylimidazolium as the minor product. Significantly, in each of these examples the NHC is again acting as substrate and ends up in the products.
![]() | ||
Scheme 4 |
An investigation of Ni-NHC based olefin dimerization catalysis has provided a further notable example of the impact of reductive coupling on the coordinated NHC, and on the catalytic behaviour of the metal complex.8Catalyst systems of the type [NiL2X2 + AlEtnX3−n] (where L = PR3 and X = halide) are well known and generally afford highly active catalysts for olefin dimerization and isomerization. In the study of interest here, Ni(NHC)2I2 complexes, 8 (Scheme 5), were synthesised, and treated with AlEt2Cl as cocatalyst, in the presence of 1-butene, in toluene at 20 °C. Analysis of the products from the reaction showed only reductive elimination products, I, II, III (Scheme 5). No butene dimers were observed. Decreasing the temperature to −15 °C, and using the complex with 1,3-diisopropylimidazol-2-ylidene as the NHC ligand generated very small amounts of butene dimers. These results unambiguously demonstrate that Ni–NHC complexes are capable of olefin dimerization, however, decomposition of the catalystvia reductive elimination prevents completion of the desired catalytic cycle. Moreover, the specific decomposition products observed here provide unequivocal evidence that this dimerization reaction follows the Cossee–Arlman mechanism of activation, insertion and elimination, commonly invoked for olefin chain-growth processes.
![]() | ||
Scheme 5 |
When the matching dimerization process, using the same NHC–Ni complexes as catalyst precursors, was conducted in ionic liquid (IL) solvent [1-butyl-3-methylimidazolium chloride–AlCl3–N-methylpyrrole (0.45 : 0.55 : 0.1 ratio)] rather than toluene, the catalyst was found to be highly active with no evidence of decomposition.8 Furthermore, product distributions for each of the catalyst systems studied was surprising similar, indicating a common active species may have been formed in each case. To explain such an interpretation it was suggested that reductive elimination of the NHC–Ni did indeed occur, as outlined in Scheme 5; however, the ionic liquid solvent could now oxidatively add to the Ni(0) thus formed to yield a new Ni–NHC complex, stabilized by the IL solvent, and able to effectively catalyse the dimerization process (Scheme 6). Such a proposal would of course require evidence that oxidative addition of imidazolium salt is indeed feasible (vide infra).
![]() | ||
Scheme 6 |
![]() | ||
Scheme 7 |
A DFT computational study (b), on the influence of the spectator ligand bite angle and the twist angle of the NHC on reductive elimination from PdL2 complexes (L = phosphine), demonstrated that bite angle has a significant impact.16 Increasing the bite angle (across a range 80–130°; 12) lowered Eact and increased the exothermicity of the reaction. Complexes with a large bite angle >110° showed a thermodynamic preference for products, complexes with a bite angle <110° prefer the reverse reaction—oxidative addition. Interestingly, steric bulk on the phosphines appears to be just as important as bite angle in influencing the reaction.16,23 Rotation of the NHC with respect to the coordination plane of the complex 13 had little influence on Eact but has significant effect on the thermodynamics of the decomposition reaction. The instability of the complexes where the carbene twists towards coplanarity with the coordination plane can be attributed to steric strain rather than electronic influences.
A further computational study (c), using DFT methods, investigated the influence of the NHC N-substituents on both the Eact and the overall thermodynamics of the reductive elimination reaction, 14.17 Electronic factors have a major influence on the activation barrier. The NHC pπ orbital plays a key role in the reductive elimination process and N-substituents that remove π-density act to promote elimination, whereas increased electron donation from the N-substituents stabilise against reductive elimination.17
An experimental study, (d), found that chelating spectator ligands impart a degree of stability to the complexes of type 15.18 When two monodentate phosphine ligands are used decomposition is rapid at 20 °C, however, if the monodentate ligands are replaced by dppp no decomposition is detected after 24 h.7 Comparing different chelating phosphine ligands, it was found that the rate of decomposition could ostensibly be linked to the chelate ring size; at 65 °C, with dppp decomposition was complete after 6 hrs, with dppe only a small amount of decomposition had occurred in this time.18
Having recognised the problem presented by reductive elimination it is possible to consider methods of limiting the impact of the reaction. A number of solutions have been proposed and tried with mixed success. For example the use of chelating ligands incorporating an NHC component has worked relatively well but is limited to catalyst systems in which a high degree of coordinative unsaturation is not essential.15,20,24
Another approach is the application of imidazolium based ILs as solvents, which has been shown to have clear benefits (vide supra), by providing highly active and very stable catalyst systems.8
![]() | ||
Scheme 8 |
Recent DFT computational studies explored different aspects of the oxidative addition reaction.32,33 In the first study, oxidative addition of azolium salts to a model Wilkinson's catalyst was investigated.32 Ironically, the calculations show that the reverse reaction (reductive elimination to give the original imidazolium salt and the Rh(I) complex) is thermodynamically preferred, suggesting that reductive elimination could occur in NHC–Rh(I)–H complexes. The second computational study looked at the oxidative addition of 2-substituted azolium salts to group 10 metals (eqn (6)) to give complexes 19.33 It appeared that all three metals would have little trouble undergoing oxidative addition, particularly when the 2-substituent was H. Interestingly, it was predicted that “under the right conditions” even the C2–Me “blocked” IL [R = Me, eqn (6)], would also undergo oxidative addition. Furthermore, oxidative addition of oxazolium salts [X = O, eqn (6)] to a group 10 metal centre is highly energetically favourable, providing a valuable synthetic route to new NHC complexes.
![]() | (6) |
![]() | ||
Scheme 9 |
A range of olefins and of azolium salts have been utilised as substrates in this reaction (Table 1), providing a variety of 2-substituted heterocyclic compounds.34,35Catalytic performance is modest and high catalyst loadings are required. Of significance, neutral azoles could also be used in the reaction [eqn (8)], further extending its potential.35 However, in the case of the neutral azole it was necessary to activate the 2-H by adding a strong Lewis acid to the system.
![]() | (8) |
![]() |
A comprehensive DFT investigation of the mechanism has been carried out, and provides a valuable indication why catalyst performance is only modest.35 The reaction mechanism appears quite complex, and subtle changes to the catalyst lead to changes in the rate-determining step. For small, difficult to dissociate (very basic ligands), Scheme 9, L= PMe3, the main activation barrier is ligand dissociation prior to alkene coordination. For bulkier, more readily dissociated ligands (L = PPh3, PBut3) the main energy barrier is the reductive elimination step, although the coordination of alkene can also be high in energy in these systems. In all cases, the oxidative addition step was facile and unlikely to be rate-determining. Concurrent, and extensive, experimental studies assessing ligand (L) sterics and electronic properties, and L : Ni ratios were in very good agreement with the computational studies.35
In subsequent studies the intramolecular 2-substitution, or annulation of a range of azolium salts was investigated.36 The reaction converts N-alkenyl and N-substituted alkenyl substituents into five- and six-membered fused ring azolium species, efficiently and in high yield (Scheme 10). Catalyst loadings of 2–5% gave the best results. Lower catalyst loadings lead to reduced conversion; at lower catalyst loadings, i.e. lower Ni : substrate ratios the equilibrium favours coordination of a second molecule of alkenylimidazolium to the metal centre, shutting down the catalyst. Highly branched rings and larger ring products were also generated by this reaction. A common feature of the fused ring products is the creation of a chiral centre on ring closure, a potentially important feature.
![]() | ||
Scheme 10 |
Structures such as 23 and 27 are oils at room temperature and therefore have potential as IL solvents. Therefore, an asymmetric version of this reaction could provide a catalytic route to chiral ILs. For this reason (S)-QUINAP and (R)-BINAP were tested as ligands, L, in the reaction, eqn (10). However, no annulated products were observed, only alkene isomerisation products were obtained. It appears that the chelating spectator ligands block a coordination site and shut down the final reductive elimination step in the reaction (Scheme 9); the Ni(II) then acts as a standard isomerisation catalyst. Chiral monodentate ligands, which we are currently investigating, will be required for this reaction.
![]() | (10) |
![]() | ||
Scheme 11 |
The oxidative addition of imidazolium salts to generate “normal” C2-bound carbene complexes had been shown to be a facile process and ensuing studies showed that 2-blocked imidazolium salts will also oxidatively add to Pt(0)40 and Ni(0)41 to form abnormal, C4/5 bound carbene–M–H complexes, Scheme 12. The C4/5-NHC–M–H (M = Ni) thus formed could be expected to also undergo insertion of olefins and subsequent reductive elimination to generate C4/5 substituted imidazolium salts, in much the same way that has been demonstrated for normal C2 bound NHCs (vide supra).
![]() | ||
Scheme 12 |
In preliminary studies styrene and also dimethyl fumarate were reacted with complexes 33/34 to observe whether insertion and possibly reductive elimination occurred.40,41
Unexpectedly, the reaction that took place was an apparent equilibrium process in which reductive elimination of the hydride and abnormal carbene predominated forming C4/5 hydridoimidazolium salt, 38, and a Pt(0) diolefin complex, 37, eqn (15). There was no evidence of olefin insertion, nor was the normal C2 bound NHC eliminated. Therefore, initial studies indicate that both oxidative addition of 2-blocked imidazolium salt, and reductive elimination of the abnormal NHC does occur, whereas, migratory insertion did not.
![]() | (15) |
These results are entirely consistent with the behaviour of catalysts for 2-substitution of azolium salts (eqn (7), Scheme 9), discussed above. The reaction of azolium salt and olefinic species in the presence of the most active catalyst systems developed to date, results in the formation of 2-substituted azolium salt only (Table 1); further reaction to give C4/5 substituted azolium has not been observed.
![]() | (16) |
![]() | ||
Scheme 13 |
Despite similarities between the Rh(I) catalysed reaction and our Ni(0)/Pd(0) catalysed reaction, thorough mechanistic studies by the Bergman group (Scheme 14),46 in combination with our investigations on the mechanism of the Ni/Pd catalysed process (Scheme 9),32,33,35 demonstrate the distinct mechanistic differences between the two processes. The former mechanism proposes a 1,2-H shift followed by C–H cleavage and migration of the H to the Rh centre, i.e. a unique C–H activation and intramolecular H-transfer process. The mechanism for the M(0) catalysed reaction involves a primary oxidative addition step to generate an NHC–M–H intermediate into which the olefinic moiety inserts to give NHC–M–alkyl species, which undergoes reductive elimination to generate the product.
![]() | ||
Scheme 14 |
Bergman extended the reaction to include; catalytic arylation of heterocyclic compounds (again highlighting the likely intermediacy of an NHC–Rh complex);47Rh(I) catalysed alkylation of quinolines and pyridines;48 and enantioselective annulation of an indole, using a catalyst system {[RhCl(COE)2]2 plus added phosphoramidite as the chiral ligand}, ultimately generating biologically active dihydropyrroloindoles.49 In important, related, studies Bergman and co-workers applied the Rh-catalysed C–H activation and functionalisation of heterocyclic species to the total synthesis of vasicoline50 and to the synthesis of kinase inhibitors.51 In the former reaction 3,4-dihydroquinazolines were both inter- and intra-molecularly coupled with olefinic species (Scheme 15). The studies of Bergman and co-workers beautifully illustrate several key applications for this type of unique chemistry—involving azole substitution and annulation.
![]() | ||
Scheme 15 |
This journal is © The Royal Society of Chemistry 2008 |