Anda di halaman 1dari 9

View Online / Journal Homepage / Table of Contents for this issue

This article is published as part of the Dalton Transactions themed issue entitled:

Computational Chemistry of Molecular Inorganic Systems


Guest Editor: Professor Stuart MacGregor Heriot-Watt University, Edinburgh, U.K. Published in issue 42, 2011 of Dalton Transactions

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Image reproduced with the permission of Hirofumi Sato

Articles in the issue include: COMMUNICATION: Comparison of different rutheniumalkylidene bonds in the activation step with N-heterocyclic carbene Ru-catalysts for olefins metathesis Albert Poater, Francesco Ragone, Andrea Correa and Luigi Cavallo Dalton Trans., 2011, DOI: 10.1039/C1DT10959F HOT ARTICLES: Matrix infrared spectroscopic and density functional theoretical investigations on thorium and uranium atom reactions with dimethyl ether Yu Gong and Lester Andrews Dalton Trans., 2011, DOI: 10.1039/C1DT10725A Reductive coupling of carbon monoxide by U(III) complexesa computational study Georgina Aitken, Nilay Hazari, Alistair S. P. Frey, F. Geoffrey N. Cloke, O. Summerscales and Jennifer C. Green Dalton Trans., 2011, DOI: 10.1039/C1DT10692A Prediction of high-valent iron K-edge absorption spectra by time-dependent Density Functional Theory P. Chandrasekaran, S. Chantal E. Stieber, Terrence J. Collins, Lawrence Que, Jr., Frank Neese and Serena DeBeer Dalton Trans., 2011, DOI: 10.1039/C1DT11331C Visit the Dalton Transactions website for more cutting-edge inorganic and organometallic research www.rsc.org/dalton

Dalton Transactions
Cite this: Dalton Trans., 2011, 40, 11184 www.rsc.org/dalton

Dynamic Article Links

PAPER

Organometallic reactivity: the role of metalligand bond energies from a computational perspective
Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Natalie Fey,* Benjamin M. Ridgway, Jes us Jover, Claire L. McMullin and Jeremy N. Harvey*
Received 13th May 2011, Accepted 29th July 2011 DOI: 10.1039/c1dt10909j The association and dissociation of ligands plays a vital role in determining the reactivity of organometallic catalysts. Computational studies with density functional theory often fail to reproduce experimental metalligand bond energies, but recently functionals which better capture dispersion effects have been developed. Here we explore their application and discuss future challenges for computational studies of organometallic catalysis.

Introduction
Many of the important mechanistic steps in homogeneous organometallic catalytic cycles involve the addition or loss of ligands (which might also be substrates or solvent molecules). In an inuential review on thermochemistry of transition metal compounds, Simoes and Beauchamp referred to metalhydrogen and metalcarbon bond strengths as holding the keys to catalysis.1 However, a strong case could be made that the strengths of metal bonds to other ligandsphosphines foremost amongst themare equally crucial. Hence, in order to provide a good understanding of catalysis, computational methods must have the ability to predict bond energies with reasonable accuracy, i.e. usually better than 3 kcal mol-1 when compared to experimental data. This level of accuracy typically has not been achieved by computational studies of organometallic catalysis, preventing quantitative insight into catalytic mechanisms; this also casts doubt on the ability of computation to predict mechanisms correctly and continues to hamper its routine use for the design and development of new catalysts. Nevertheless, computational studies of mechanism and the related elucidation of ligand effects have been successful in the past where prediction errors have been systematic, thus allowing the evaluation of small modications to the catalyst structure and/or substrates considered. Several factors contribute to make accurate computational prediction challenging: 1) Size limitations. Many of the transition metal complexes relevant to organometallic catalysis are large, as well as conformationally and congurationally variable. Until recently, calculations had to rely on suitable, simplied models of

School of Chemistry, University of Bristol, Cantocks Close, Bristol, BS8 1TS, UK. E-mail: Natalie.Fey@Bristol.ac.uk, Jeremy.Harvey@ Bristol.ac.uk; Fax: (44) 0117 9251295 Electronic supplementary information (ESI) available: Computational details, including implementation of DFT-D corrections for LKB, calculated bond energies for complexes with different approaches, calculated Pd-P bond lengths. See DOI: 10.1039/c1dt10909j

the active catalyst to limit the complexity and length of calculations with electronic structure methods. While computational resources have improved, it is worth noting that most calculations routinely neglect other species present in the reaction mixture, such as solvents and counterions, that conformers and isomers continue to pose considerable challenges, and that higher levels of theory such as CCSD(T), used e.g. to provide accurate benchmark data where experimental data is not available, continue to be inaccessible to any but the smallest catalyst models. 2) Mechanistic alternatives. Both experimental and computational studies of homogeneous catalysis often nd evidence of more than one possible reaction pathway, where a given combination of substrate, ligand and reaction conditions determines which pathway is accessed. Even with small model complexes, the full exploration of a mechanistic manifold requires considerable computational effort. 3) Methodological limitations. Computational studies of transition metal complexes predominantly use density functional theory (DFT),2 which represents a suitable compromise between accessible system size and accuracy, given currently available processing power. Calculations can be improved further by including solvation effects and corrections for zero-point energy, enthalpy and entropy (see e.g. reference 3), but DFT also has some wellknown shortcomings,2b,4 which are more difcult to address, such as the neglect of dispersion effects,5 self-interaction errors6 and medium- to long-range correlation errors.7 These factors also affect the prediction of metalligand bond energies: 1) Model ligands often do not fully capture the steric and electronic properties of those used synthetically, which may affect structures, bond energies and calculated isomer preferences. In addition, larger ligands are often both more exible and more sensitive to conformational effects, adding conformational noise to calculated energies.8 2) Several different mechanisms and intermediates may need to be considered to identify the correct pathway and thus predict bond energies in agreement with experiment;3a,9 in addition, energetic differences between pathways may be subtle. 3) Corrections are often computationally costly, especially those requiring frequency calculations, and may This journal is The Royal Society of Chemistry 2011

11184 | Dalton Trans., 2011, 40, 1118411191

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

not be feasible for large molecules and multiple pathways. In addition, some corrections are sensitive to complex size, whereas others are less so, requiring careful consideration of models and simplications. Recently, dispersion-corrected density functionals45 have been applied to the computational study of transition metal complexes relevant to organometallic catalysis with promising results, and the remainder of this article will seek to highlight key developments in this area, as well as discussing how further improvements might be achieved, with a view to bringing accurate computational prediction and catalyst design into reach. These functionals have of course also been applied to problems in other areas of chemistry,10 but we do not focus on these applications here.

Due to the tting process needed, improvement of density functional theory is often not systematic12 and functionals may be better suited to problems which are similar to the data used for tting. Most new formulations are evaluated rst in benchmark studies, which provide a useful overview of strengths and weaknesses of different functionals and tting philosophies. Recently, data relevant to transition metal compounds and organometallic catalysis has been included in such benchmark studies and references 18b and 21 provide some examples. Structures The importance of dispersion contributions was demonstrated impressively in structural studies by Maseras and Eisenstein, where standard DFT on model complexes failed to reproduce the crystallographically observed geometries of [OsCl2 H2 (PiPr3 )2 ]22 and [Ir(H)2 (PtBu2 Ph)2 ]+ ,23 whereas integrated molecular orbital/molecular mechanics (IMOMM) calculations,24 where the ligand bulk is described by a molecular mechanics approach which explicitly captures dispersion interactions, were more successful. This led Maseras to propose the use of quantum mechanical/molecular mechanical (QM/MM) hybrid methods as superior to pure DFT calculations at the time.25 More recently, Grimme and Djukic investigated the aggregation of cationic [Rh(CNPh)4 ]+ complexes with a range of dispersion-corrected density functionals, focusing especially on dimers of this complex.26 They found that dimers as observed experimentally were not predicted by standard functionals, and that both dispersion and solvation effects had to be captured to achieve at least qualitative predictions of binding, balancing the charge repulsion against attractive interactions. In this study, only one functional (TPSS-D3 with the COSMO solvation model) predicted a conformer preference in agreement with the crystal structure of this dimer, whereas other conformations are favoured if dispersion is not included. Further analysis of the potential energy curves for dimer dissociation led them to conclude that in this case the interaction is dominated by ligandligand attraction and not by metallophilicity. On the other hand, several recent studies noted that geometries optimised with dispersion corrected functionals are not significantly improved compared to other functionals,21f,27 but that the contribution to bond energies is considerable.27 This will be discussed further below. It is perhaps also worth noting in this context that the recently reported successful crystal structure prediction relied on capturing dispersion effects.28 Ruthenium-catalysed alkene metathesis The contribution of these weak interactions to organometallic catalysts was considered in greater detail for computational studies of ruthenium-catalysed alkene metathesis (Scheme 1),29 where mechanistic studies of the catalytic cycle using the real catalyst, both at Bristol30 and in other groups,31 found calculated phosphine dissociation energies from the precatalyst [(L)(PR3 )Ru(Cl)2 (CHPh)] (L = PR3 or NHC) to show the opposite trend from that observed experimentallyinitiation of the carbene precatalysts is predicted to be faster, yet experimentally observed to be slower, than that of the bisphosphine complex. Detailed analysis, including IMOMM calculations on a range of complexes Dalton Trans., 2011, 40, 1118411191 | 11185

Dispersion Weak hydrogen bonds and van der Waals interactions between different parts of organometallic complexes can be important for the reproduction of observed structures and bond energies, but these forces are not captured well by standard density functional theory. This arises because all density functional approaches in use today are attempts to approximate the exact functional11 and are thus tted, either to reproduce an external data set (from experimental measurements or calculations at a higher level of theory) or the theoretically expected properties of the exact functional.2,12 Shortrange exchange and correlation energies, which contribute to chemical bonds, are captured well by functionals in use today, but medium-range, non-covalent interactions such as those leading to van der Waals attraction, weak hydrogen bonding and pp stacking, are described less well. A number of strategies have been described which seek to address this problem with DFT while retaining the computational efciency of this level of theory. Here, we focus mainly on approaches which have found application in organometallic chemistry and will not aim to review the theoretical aspects of dispersion corrections rigorously. More complete and technical reviews can be found e.g. in reference 13. The main strategies can be grouped as follows: A) Corrected : Dispersion can be described well by a simple potential function of the form C6 R-6 , where C6 are the dispersion coefcients and R are interatomic distances and this pragmatic way5a has been implemented and parameterised extensively by Grimme and co-workers.5,14 We note that it was also proposed separately and in some cases much earlier by many other groups.15 In Grimmes implementation, the use of such corrections is usually denoted by adding the extension -D to the standard abbreviations used for functionals, i.e. B3LYPD adds these empirical dispersion corrections to the standard B3LYP functional. B) Fitted : It is possible to empirically t density functionals so they capture dispersion interactions; this has been implemented e.g. by Truhlar and Zhao, who included a set of noncovalent interaction energies in the tting of the Minnesota functionals (of which M06 and M06-L are recommended for transition metal complexes4 ), summarised in references 4 and 16. C) Intrinsic: Finally, it is also possible to develop functionals which intrinsically describe dispersion interactions, and such functionals have been reported e.g. by Becke (XDM,17 DF0718 ) Perdew (random phase approximation, RPA)19 and Furche (resolution of identity approximation to RPA, RI-RPA).20 This journal is The Royal Society of Chemistry 2011

Table 1 Evaluation of different density functionals against experimental data for ruthenium-catalysed metathesis Set 1a , 21e (kcal mol-1 ) 12.7 10.8 18.6 40.3 39.4 Set 2b , 33 (kcal mol-1 ) 13.2 10.7 19.2 39.8 37.6 43.1 42.8 36.9 Set 3c , 33 (kcal mol-1 ) -1.1 -1.4 4.8 25.4 23.3 28.8 28.5 27.0 Set 4d , 3b (kcal mol-1 ) 14.9 (14.1)e 21.2 (17.9)e -15.2 -19.6 -21.6 (-24.2)e -17.9 (-20.0)e -17.5 2.0

Functional BP86 B3LYP PBE M06 M06-L BP86-D B3LYP-D B97-D Expt.
a

40.2

Dissociation of PCy3 from (H2 IMes)(PCy3 )Cl2 Ru CHPh, single point calculations on M06-L geometries. b Activation enthalpies for dissociation of PCy3 from (H2 IMes)(PCy3 )Cl2 Ru CHPh in gas phase, single point calculations on B3LYP geometries c As b , but in toluene. d Enthalpy of binding of PPh3 to [Ru(CO)Cl(PPh3 )2 (CH CHPh) in CH2 Cl2 , single point calculations on BP86 geometries. e Numbers in parentheses indicate full geometry optimisation with functional shown.

Scheme 1 Alkene metathesis.

with different ligands, allowed one of us to suggest a number of reasons for this discrepancy.30 While the likely impact of conformational effects, solvation and basis set size was found to be small in this work, introduction of dispersion via the IMOMM hybrid approach gave much better agreement with the experimental observations, indicating the importance of this effect. These observations led to the adoption of the phosphine dissociation energy for the ruthenium precatalyst for metathesis as a test case for the development of new density functionals.21e,32 In most studies, calculations at higher levels of theory were used to provide benchmark data,32 but recent studies by Truhlar,21e 3b Goddard34 and Hillier and Percy35 also consider Jensen,33 Buhl, experimental results for related complexes and achieve reasonably good agreement. Table 1 lists examples from the work of Truhlar,21e 3b which cover the broadest range of functionJensen33 and Buhl, als. For this reaction, functionals with an empirical dispersion correction45 as well as the Minnesota functionals4,16 are found to give much better agreement with both experimental data and higher levels of theory for this crucial catalyst activation step than other functionals.3b,21e Complex bond-dissociation enthalpies The work by Jensen and co-workers33 mentioned in the preceding section goes beyond consideration of the ruthenium metathesis pre-catalyst and investigates the reproduction of a range of experimental bond dissociation enthalpies measured in solution for organometallic complexes by different DFT approaches. Their work conrms that commonly used density functionals without dispersion corrections underestimate metalphosphine bond strengths, whereas dispersion-corrected approaches generally give much better agreement between the calculated and experimental data. Interestingly, when considering the dissociation of different
11186 | Dalton Trans., 2011, 40, 1118411191

phosphines (PR3 ) from [Ni(CN)2 (PR3 )3 ] they found standard DFT approaches were more sensitive to changes in the coordinated phosphine ligands from PEt3 to PEtPh2 , predicting relative enthalpy differences on changing the ligand in better agreement with experimental observations than many of the dispersion-corrected functionals, even though absolute enthalpies are underestimated. We have recently described our work to develop a ligand knowledge base for P-donor ligands (LKB-P),36 which seeks to capture the properties of ligands in a range of different coordination environments and use the resulting steric and electronic ligand parameters both in maps of ligand space and to derive statistical models for the interpretation and prediction of ligand effects on the properties and catalytic activity of organometallic complexes.37 In this context, we have explored using thermodynamic data reported in the literature as test cases for the derivation of linear regression models. Our intention was to reproduce the experimental data for a range of different ligands with a suitable computational approach, and then to expand the calculated data set with a more extensive range of ligands than experimentally accessible. This data would then be used to train and test statistical models based on LKB parameters to explore and illustrate the prediction of ligand effects. One of the experimental datasets we considered arose from solution calorimetric studies by Nolan and co-workers for the substitution reaction: [RuCl2 (p-cymene)]2(soln) + 2 PR3(soln) 2 [RuCl2 (pcymene)(PR3 )](soln) , measured in dichloromethane at 30 C.38 This work reported results for 23 ligands; DFT calculations39 with the standard BP86 functional and a medium-sized basis set (bond energies denoted here as BE1, see supporting information for computational details) showed a systematic underestimation of these enthalpies of substitution, as well as considerable scatter of results (Fig. 1, see supporting information for results table, Table S1). The latter could be traced to conformational effects for some of the larger ligands, which mesh with the p-cymene group, making it computationally challenging to locate a low energy conformer; we have removed the most exible ligands, P(OMe)3

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Fig. 1 Bond energies (kcal mol-1 ) for [RuCl2 (p-cymene)]2(soln) + 2 PR3(soln) 2 [RuCl2 (p-cymene)(PR3 )](soln) (see text and supporting information for details). Linear regression equations tted for BE1 (BP86/BS1) (R2 = 0.824), BE2 (BP86/BS2 with SP solvation) (R2 = 0.907) and BE2 (with solvation and dispersion) (R2 = 0.584). The solid black line indicates where quantitative prediction of the experimental data by calculation would be observed.

This journal is The Royal Society of Chemistry 2011

and P(OPh)3 from further consideration, but, despite our best efforts, some of the scatter remains. The data is described reasonably well by a linear relationship, R2 = 0.824 (BE1, light blue circles in Fig. 1). We also investigated calculations with a simplied complex, where the p-cymene ligand is replaced by benzene, but again we observed some scatter when comparing to the experimental data which we tentatively attribute to conformer effects. Extension to a larger basis set and inclusion of solvation effects, both from single point calculations (bond energies calculated with the larger basis set are denoted as BE2, solvation effects indicated by (solv), see supporting information for computational details) improved the correlation with experimental data slightly (R2 = 0.907, BE2 (solv), green triangles in Fig. 1), but calculations continued to underestimate the metalligand bond energies. Introduction of a dispersion energy correction according to Grimme5b (see supporting information, indicated by (disp)) for these complexes improved the quantitative agreement with experimental data for some of the ligands (Table S1), but in many cases led to an overestimation of bond energies, as well as increased scatter, so overall the data is described less well by a linear relationship (R2 = 0.584, BE2 (solv + disp), blue dots in Fig. 1). In this preliminary study, we have used the geometries optimised with our basic approach throughout, thereby not accounting for the effects of solvation and dispersion on calculated complex structures. In addition, solvation effects have been modelled with a continuum dielectric eld, rather than explicit solvent molecules, which may not fully capture interactions between solutes and solvent; the biggest change is likely to occur on coordination of the ligands to the metal centre, reducing dispersive ligand-solvent interactions considerably. In addition, we have not included vibrational corrections to the potential energy, assuming that these would be small and not substantially affected by changing the ligands. Conformational noise and insufcient representation of dispersive interactions with the solvent may well account for some of the observed scatter. Nevertheless, the magnitude of ligand binding energies looks in better agreement when including dispersion corrections (BP86-D). Ruthenium-catalysed alcohol dehydrogenation Alcohol dehydrogenation, catalysed by ruthenium complexes of the general form [RuH2 (X)(PPh3 )3 ], where X = N2 , PPh3 , has recently been investigated with dispersion-corrected functionals as (B97-D)9a well as the tted M06 functional by the groups of Buhl 40 and Bolm (B3LYP, B3LYP-D, M06). In line with observations for alkene metathesis catalytic cycles as discussed above, dispersion needs to be captured for the accurate description of associative and dissociative steps in the catalytic cycle. Only when such effects are taken into account, can competing reaction pathways be evaluated reliably as ligand/substrate/solvent bond energies play a crucial role. Experimental data for comparison is limited in this case, but both groups report that their results are close to experimentally observed reaction rates and that several pathways may be energetically competitive. Interestingly, Bolm et al. found B3LYP to give reasonably good results for the overall reaction as well,40 which they attribute to a partial cancellation of dispersion contributions for different steps; they also note the similarity between free energy diagrams for a model ligand PMe3 calculated This journal is The Royal Society of Chemistry 2011

with B3LYP, where dispersion effect would be more limited, and the B3LYP-D free energy prole for the full PPh3 ligated system, lending some justication to the past success of carefully chosen model systems in computational studies of catalysis. Oxidative addition to palladium(0) The oxidative addition of an aryl halide to a palladium centre occurs in a range of catalytic cycles, most importantly perhaps in the family of cross-coupling reactions.41 This reaction can occur via a number of different pathways, 1) ligand loss, followed by aryl halide coordination and oxidative addition (dissociative), 2) associative displacement, where ligand loss and aryl halide coordination are concerted, and precede the oxidative addition step, or 3) involving a bisphosphine ligated metal centre throughout (Scheme 2).3a Computational studies with standard DFT approaches predicted the dissociative route as most likely,41e,42 but failed to achieve quantitative agreement with available experimental data. For standard DFT, the associative displacement pathway gives rise to a very high barrier42d,43 and has thus been ruled out from consideration, and the bisphosphine pathway is less likely to occur for complexes with sterically hindered ligands such as PtBu3 used experimentally.

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Scheme 2 Possible reaction pathways for oxidative addition of ArX to [PdL2 ].

We recently reported a B3LYP-D study of this reaction step for the addition of phenyl halides to [Pd(PtBu3 )2 ],3a in which we demonstrated that near-quantitative agreement with experimental data44 can be achieved on inclusion of both dispersion and solvation effects when free energies are calculated. While the bisligated pathway 3) can be ruled out for a ligand of this size, the balance between the dissociative and associative pathways is more subtle and only inclusion of dispersion effects reproduced halide effects observed experimentally. One noteworthy observation was that for the large ligand considered in this work, compared to the B3LYP electronic energies in vacuum, the entropic corrections yielding free energies, and the B3LYP-D dispersion correction tended to be of similar magnitude but opposite sign for different species. The effect of these two corrections nearly cancelled out when considering the relative energy of the crucial ligand loss Dalton Trans., 2011, 40, 1118411191 | 11187

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Fig. 2 Energy proles for palladium-catalysed oxidative addition of PhBr to [Pd(PtBu3 )2 ], showing potential energy and free energy calculated with B3LYP and B3LYP-D (kcal mol-1 , numbers on diagram are for best method, DGvdW (B3LYP-D).3a

species and the associative displacement transition state (Fig. 2). Of course, where accurate results are needed, e.g. as in our case where we were focusing on rather subtle substituent effects and seeking close agreement with accurate experimental data, the net effect of these corrections was of vital importance. While the observation that vacuum B3LYP potential energies are in reasonable agreement with the free energy results obtained after dispersion and solvation corrections is striking, this can be expected, as entropic effects tend to disfavour more highly associated species, and dispersion corrections tend to favour such species. The sterically hindered PtBu3 ligands are able to stabilise the formally 14-electron Pd(0) complex [PdL2 ],45 but higher coordination numbers have been observed for other ligands, and [Pd(PPh3 )4 ] is a useful pre-catalyst which has been characterised crystallographically (CSD Refcode TTPPDB).46 Kinetic studies of this system,47 as well as alkylphosphine substituted pre-catalysts,48 suggest that ligand dissociation to form the [PdL3 ] complex is easy, so that [PdL4 ] is absent in solution, with [PdL3 ] and [PdL2 ] present and likely to be in rapid equilibrium. Standard DFT calculations (B3LYP and BP86, see supporting information for computational details) of ligand dissociation energies for [PdLn ] complexes where n = 24 and L = PPh3 give rise to the relative potential energies summarised in Table 2, showing the [PdL3 ] complex to be most stable. However, ligand dissociation will be driven by free energies rather than potential energies. Hampered by the size of complexes, we have not calculated the entropic contributions to ligand loss here, but one can use the rough rule whereby each dissociation process will introduce a favourable T DS contribution of ca. 10 kcal mol-1 near room temperature. This rule suggests that PdL2 + 2 L is lower in free energy than either PdL4 or PdL3 + L according to B3LYP and BP86. This low predicted stability of the triply and quadruply ligated species is further demonstrated by the optimised geometries, which show signicantly longer PdP bonds than the lower-coordinate complexes (see supporting information, Table S2). In contrast, optimisations with the dispersion-tted M06 and M06-L functionals show the complexes with lower ligand coordination numbers to be signicantly less favourable than [PdL4 ], in better agreement with experimental data. Arguably the preference for [PdL4 ] is now too strong, but these calculations do not take account of solvation, basis set and entropic effects,
11188 | Dalton Trans., 2011, 40, 1118411191

Table 2 Stability of [PdLn ] complexes, where L = PPh3 and of [Pd(binap)n ] complexes, DFT potential energies, kcal mol-1 Complex PdL4 PdL3 + L PdL2 + 2 L Pd(binap)2 Pd(binap) + binap B3LYP 0.0 -6.8 -0.5 0.0 8.4 BP86 0.0 -4.6 5.6 0.0 17.4 M06 0.0 29.6 57.9 0.0 47.0 M06-L 0.0 25.5 57.2 0.0 52.9

where dispersive ligand-solvent interaction especially may affect the calculated predictions, as discussed above. Similarly, the experimentally stable [Pd(binap)2 ] complex is clearly favoured only if a dispersion corrected density functional is used. Ligand size effects Our ligand knowledge base for P-donor ligands (LKB-P)36 as briey described above contains a range of steric and electronic parameters for 348 ligands, derived from DFT calculations in different coordination environments. While our work is predominantly aimed at the interpretation and prediction of ligand effects in transition metal complexes, ligand properties can also be examined in detail for individual complexes. The database includes bond energies for dissociation of these ligands from different fragments, and here this provides us with an opportunity to explore the effect of ligand size on bond energies and dispersion contributions in a large and varied dataset, representative of this important class of monodentate ligands. We have calculated dispersion energy contributions according to Grimmes approach (DFT-D revision 2)5b for two of the largest transition metal complexes used in our LKB-P work,36 [PdCl3 L]- and [(PH3 )3 PtL], as well as the uncomplexed ligands and metal fragments after ligand dissociation. Since parameters for dispersion corrections involving platinum have not been explicitly tted for this version of DFT-D,5b we will focus here on the palladium complex, while the corresponding platinum results may be found in the supporting information (Fig. S1, S2, Table S3).
The related complex bis(2,2-bis(di-p-tolylphosphino)-1,1-binaphthyl) palladium(0) has been characterised crystallographically as the benzene solvate (CSD Refcode QIPBEW) and has been used as a starting geometry here.51

This journal is The Royal Society of Chemistry 2011

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Fig. 3 Ligand bond energy (kcal mol-1 ) in [PdCl3 L]- calculated at BP86 (BE) and BP86-D (BEvdW ) level for ligands in LKB-P. Regression coefcient of linear relationship R2 = 0.419, excluding data highlighted with bold circles.

As expected, adding dispersion energies to the bond dissociation energies for the [PdCl3 L]- complex calculated with our standard approach, i.e. comparing BP86 and BP86-D data, shows the two datasets to be linearly correlated, albeit with considerable scatter (R2 = 0.358, Fig. 3). Analysing the data by ligand type suggests that ligands involving P-halide bonds (green circles) consistently have lower dispersion corrected bond energies than other ligands, as well as contributing to noise in the data set. We thus excluded these ligands from further analysis, as well as some additional outliers with unusual ligand structures that appeared on the periphery of the data set (identied in Fig. S3). This gives a slightly improved correlation between the two datasets, R2 = 0.419 for 306 ligands as shown in Fig. 3. The slope of the tted trendline is of interest here, suggesting that dispersion contributions have a bigger impact where the BP86 binding energies are calculated to be small, than they do where binding energies are quite large. From our LKB analysis, we know that larger ligands are often predicted by the BP86 functional to bind more weakly to the metal fragments, due to unfavourable steric interactions with cis-coordinated ligands. On inclusion of dispersion effects, these repulsive interactions are to some extent countered by attractive intramolecular forces, thus increasing the binding energy more noticeably than for complexes where steric bulk is less important. For a related discussion, see reference 49. We have also explored the relationship between the increase in metalligand binding energies due to dispersion corrections and a steric parameter, He8 _steric.36a This steric descriptor is calculated as the energy of interaction between a ring of 8 helium atoms positioned where cis-coordinated ligands might be found in a transition metal complex, and the free ligand (Scheme 3).36 As shown in Fig. 4, there is a correlation between ligand size and the impact of including dispersion contributions on metalligand bond energies, but it is perhaps not best described by a linear relationship (R2 (linear) = 0.657, R2 (ln) = 0.737). The curvature seems to arise mostly from small ligands, where presumably dispersion effects are smaller. However, there is no obvious cut-off value for He8 _steric and the dataset retains considerable scatter even without these values. We have investigated some of the most This journal is The Royal Society of Chemistry 2011

Scheme 3

LKB-P He8 _steric descriptor.36

pronounced outliers (see supporting information, Fig. S4 for a plot identifying these), and they are often conformationally exible, suggesting perhaps that dispersion corrections would also alter conformational preferences, something not taken into account by calculating the dispersion corrections for BP86 optimised geometries. Preliminary calculations (summarised in the supporting information) suggest that dispersion can indeed change the energetic order of conformers for both ligands and complexes, which can amplify conformational noise considerably. We also note that these outliers include ligands popular in organometallic catalysis, e.g. PtBu3 and PCy3 , suggesting perhaps that mechanistic studies involving these ligands should carefully consider the energetic and conformational effects of dispersion. In an LKB context at least, the effect of dispersion on binding energies can be captured by combining BP86 results with steric descriptors, which should permit the analysis and prediction of experimental data without re-calculating descriptors with improved functionals. Interpretation of models derived from this LKB might need to consider the potential exaggeration of steric effects. We have discussed the presence of conformational noise in our databases extensively, especially for bidentate ligands,50 and the effect of dispersion, and indeed solvation as discussed above, on conformational preferences will need to be investigated further.

Conclusions
A large number of computational studies have used density functional theory to address mechanistic issues in organometallic catalysis. Almost all reaction mechanisms involve a ligand Dalton Trans., 2011, 40, 1118411191 | 11189

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

Fig. 4 Relationship between difference in bond energy for [PdCl3 L]- (kcal mol-1 ) between BP86 and BP86-D, and the LKB-P He8 _steric descriptor (kcal mol-1 ).36

association or dissociation step, or a related process, so the ability of theory to predict such bond energies accurately is important. For small ligands (CO, PH3 ) and small metal complexes, standard DFT produces reasonable results in many cases, but with increasing computer power, more and more calculations address complexes with large ligands, and over the past few years this has led to the realization that signicant errors can occur in calculations due to the neglect of dispersion. In this contribution, we have reviewed some recent studies using various improved DFT methods that account better for dispersion interactions, and presented some new results. Methods such as DFT-D lead to much stronger metalligand bonds in the case of bulky metal fragments and ligands, and this can lead to signicantly improved agreement with experiment, for bond energies, but also for structures, and for predicted reaction mechanisms. Encouragingly, our overview of calculated energies for a fairly large number of phosphines binding to a ruthenium and a palladium fragment suggests that the error in standard DFT functionals is fairly systematic, as shown by roughly linear correlations between DFT, DFT-D and (where available) experimental bond energies. However, these calculations show signicant scatter in the effect of dispersion corrections, suggesting that this relationship may not be simple and may indeed vary depending on ligand type, size and conformational preferences. Also, much more work will be needed before it can be concluded that approaches such as DFT-D or the Minnesota functionals reliably predict energies and free energies for ligand binding. Approximate cancellation between dispersion effects and entropic effects may have served to partly mask the signicant errors in both of these terms in previous work, and careful consideration of entropic effects in solution will be needed. In addition, the impact of dispersion effects on conformational preferences and structures will need to be explored in detail. Finally, the interplay between dispersion interactions and solvation free energy will need to be
11190 | Dalton Trans., 2011, 40, 1118411191

addressed carefully. Ligand binding leads to signicant changes in the metal fragment due to ligand dispersion energy, but also to signicant changes in the metal fragmentsolvent, ligandsolvent, and solventsolvent dispersion free energies. It is not yet clear that continuum solvent models capture these effects accurately enough to yield reliable metalligand binding free energies.

Acknowledgements
We would like to thank our colleagues, Prof. Guy Orpen and Prof. Guy Lloyd-Jones, for many helpful discussions and gratefully acknowledge funding from EPSRC (BMR, CLM, NF, incl. ARF (NF, EP/E059376/1)), CCDC (CLM), and AstraZeneca (JJM).

Notes and references


and J. L. Beauchamp, Chem. Rev., 1990, 90, 629688. 1 J. A. M. Simoes 2 (a) C. J. Cramer and D. G. Truhlar, Phys. Chem. Chem. Phys., 2009, 11, 1075710816; (b) J. N. Harvey, Annu. Rep. Prog. Chem., Sect. C, 2006, 102, 203226. 3 (a) C. L. McMullin, J. Jover, J. N. Harvey and N. Fey, Dalton Trans., Inorg. Chem., 2009, 2010, 39, 1083310836; (b) N. Sieffert and M. Buhl, 48, 46224624. 4 Y. Zhao and D. G. Truhlar, Acc. Chem. Res., 2008, 41, 157167, and references cited. 5 (a) S. Grimme, J. Comput. Chem., 2004, 25, 14631473; (b) S. Grimme, J. Comput. Chem., 2006, 27, 17871799; (c) S. Grimme, J. Antony, S. Ehrlich and H. Krieg, J. Chem. Phys., 2010, 132, 154104. 6 M. Lundberg and P. E. M. Siegbahn, J. Chem. Phys., 2005, 122, 224103. 7 S. Grimme, Angew. Chem., Int. Ed., 2006, 45, 44604464. 8 M. Besora, A. A. C. Braga, G. Ujaque, F. Maseras and A. Lledos, Theor. Chem. Acc., 2010, 128, 639646. J. Am. Chem. Soc., 2010, 132, 80568070; 9 (a) N. Sieffert and M. Buhl, (b) N. Schneider, M. Finger, C. Haferkemper, S. Bellemin-Laponnaz, P. Hofmann and L. H. Gade, Chem.Eur. J., 2009, 15, 1151511529. 10 (a) T. Ansbacher, H. K. Srivastava, J. M. L. Martin and A. Shurki, J. Comput. Chem., 2010, 31, 7583; (b) S. Grimme and J. P. Djukic, Inorg. Chem., 2010, 49, 29112919; (c) J. N. Harvey, Faraday Discuss., 2010, 145, 487505; (d) B. G. Janesko, J. Chem. Theory Comput., 2010, 6, 18251833; (e) R. Lonsdale, J. N. Harvey and A. J. Mulholland, J.

This journal is The Royal Society of Chemistry 2011

11 12 13

14 15

16 17 18 19 20 21

22 23 24 25 26 27 28

Phys. Chem. Lett., 2010, 1, 32323237; (f) M. Mura, A. Gulans, T. Thonhauser and L. Kantorovich, Phys. Chem. Chem. Phys., 2010, 12, 47594767; (g) P. E. M. Siegbahn, M. R. A. Blomberg and S. L. Chen, J. Chem. Theory Comput., 2010, 6, 20402044; (h) P. Spies, R. Frohlich, G. Kehr, G. Erker and S. Grimme, Chem.Eur. J., 2008, 14, 333343; (i) S. Osuna, M. Swart and M. Sola, Phys. Chem. Chem. Phys., 2011, 13, 35853603. P. Hohenberg and W. Kohn, Phys. Rev., 1964, 136, B864. J. P. Perdew, A. Ruzsinszky, J. M. Tao, V. N. Staroverov, G. E. Scuseria and G. I. Csonka, J. Chem. Phys., 2005, 123, 062201. (a) E. R. Johnson, I. D. Mackie and G. A. DiLabio, J. Phys. Org. Chem., V 2009, 22, 11271135; (b) L. A. Burns, A. azquez- Mayagoitia, B. G. Sumpter and C. D. Sherrill, J. Chem. Phys., 2011, 134, 084107. T. Schwabe and S. Grimme, Acc. Chem. Res., 2008, 41, 569579. (a) R. Ahlrichs, R. Penco and G. Scoles, Chem. Phys., 1977, 19, 119130; (b) A. D. Becke and E. R. Johnson, J. Chem. Phys., 2005, 123, 154101; (c) P. Jure cka, J. Cern y, P. Hobza and D. R. Salahub, J. Comput. Chem., 2007, 28, 555569. Y. Zhao and D. G. Truhlar, Chem. Phys. Lett., 2011, 502, 113. (a) E. R. Johnson and A. D. Becke, J. Chem. Phys., 2006, 124, 174104; (b) E. R. Johnson and A. D. Becke, Chem. Phys. Lett., 2006, 432, 600603. (a) A. D. Becke and E. R. Johnson, J. Chem. Phys., 2007, 127, 124108; (b) E. R. Johnson and A. D. Becke, Can. J. Chem., 2009, 87, 13691373. A. Ruzsinszky, J. P. Perdew and G. I. Csonka, J. Chem. Phys., 2011, 134, 114110. H. Eshuis, J. Yarkony and F. Furche, J. Chem. Phys., 2010, 132, 234114. (a) C. Flener-Lovitt, D. E. Woon, T. H. Dunning and G. S. Girolami, J. Phys. Chem. A, 2010, 114, 18431851; (b) F. Furche and J. P. Perdew, J. Chem. Phys., 2006, 124, 044103; (c) C. A. Jimenez-Hoyos, B. G. Janesko and G. E. Scuseria, J. Phys. Chem. A, 2009, 113, 1174211749; (d) S. T. Mutter and J. A. Platts, Chem.Eur. J., 2010, 16, 53915399; (e) Y. Zhao and D. G. Truhlar, J. Chem. Theory Comput., 2009, 5, 324333, and references cited; (f) M. Buhl, C. Reimann, D. A. Pantazis, T. Bredow and F. Neese, J. Chem. Theory Comput., 2008, 4, 14491459. F. Maseras and O. Eisenstein, New J. Chem., 1998, 22, 59. G. Ujaque, A. C. Cooper, F. Maseras, O. Eisenstein and I. G. Caulton, J. Am. Chem. Soc., 1998, 120, 361365. F. Maseras and K. Morokuma, J. Comput. Chem., 1995, 16, 11701179. F. Maseras, Chem. Commun., 2000, 18211827. S. Grimme and J.-P. Djukic, Inorg. Chem., 2011, 50, 26192628. (a) H. Jacobsen, Chem. Phys., 2008, 345, 95102; (b) J. Goodman, V. V. Grushin, R. B. Larichev, S. A. Macgregor, W. J. Marshall and D. C. Roe, J. Am. Chem. Soc., 2010, 132, 1201312026. (a) B. Civalleri, C. M. Zicovich-Wilson, L. Valenzano and P. Ugliengo, CrystEngComm, 2008, 10, 16931693; (b) G. M. Day, T. G. Cooper, A. J. Cruz-Cabeza, K. E. Hejczyk, H. L. Ammon, S. X. M. Boerrigter, J. S. Tan, R. G. Della Valle, E. Venuti, J. Jose, S. R. Gadre, G. R. Desiraju, T. S. Thakur, B. P. van Eijck, J. C. Facelli, V. E. Bazterra, M. B. Ferraro, D. W. M. Hofmann, M. A. Neumann, F. J. J. Leusen, J. Kendrick, S. L. Price, A. J. Misquitta, P. G. Karamertzanis, G. W. A. Welch, H. A. Scheraga, Y. A. Arnautova, M. U. Schmidt, J. van de Streek, A. K. Wolf and B. Schweizer, Acta Crystallogr., Sect. B: Struct. Sci., 2009, 65, 107125, Part 2; (c) M. A. Neumann, F. J. J. Leusen and J. Kendrick, Angew. Chem., Int. Ed., 2008, 47, 24272430; (d) A. Asmadi, M. A. Neumann, J. Kendrick, P. Girard, M. A. Perrin and F. J. J. Leusen, J. Phys. Chem. B, 2009, 113, 1630316313.

29 (a) R. H. Grubbs, Angew. Chem., Int. Ed., 2006, 45, 37603765; (b) T. M. Trnka and R. H. Grubbs, Acc. Chem. Res., 2001, 34, 1829. 30 A. C. Tsipis, A. G. Orpen and J. N. Harvey, Dalton Trans., 2005, 2849 2858. 31 (a) C. Adlhart and P. Chen, Angew. Chem., Int. Ed., 2002, 41, 4484 4485; (b) C. Adlhart and P. Chen, J. Am. Chem. Soc., 2004, 126, 3496 3510; (c) L. Cavallo, J. Am. Chem. Soc., 2002, 124, 89658973. 32 M. Piacenza, I. Hyla-Kryspin and S. Grimme, J. Comput. Chem., 2007, 28, 22752285. 33 Y. Minenkov, G. Occhipinti and V. R. Jensen, J. Phys. Chem. A, 2009, 113, 1183311844. 34 D. Benitez, E. Tkatchouk and W. A. Goddard, Organometallics, 2009, 28, 26432645. 35 I. W. Ashworth, I. H. Hillier, D. J. Nelson, J. M. Percy and M. A. Vincent, Chem. Commun., 2011, 47, 54285430. 36 (a) N. Fey, A. C. Tsipis, S. E. Harris, J. N. Harvey, A. G. Orpen and R. A. Mansson, Chem.Eur. J., 2006, 12, 291302; (b) J. Jover, N. Fey, J. N. Harvey, G. C. Lloyd-Jones, A. G. Orpen, G. J. J. Owen-Smith, P. Murray, D. R. J. Hose, R. Osborne and M. Purdie, Organometallics, 2010, 29, 62456258. 37 N. Fey, A. G. Orpen and J. N. Harvey, Coord. Chem. Rev., 2009, 253, 704722. 38 (a) S. Serron, S. P. Nolan, Y. A. Abramov, L. Brammer and J. L. Petersen, Organometallics, 1998, 17, 104110; (b) S. A. Serron and S. P. Nolan, Organometallics, 1995, 14, 46114616. 39 J. Jover, N. Fey, and J. N. Harvey, unpublished results. 40 A. J. Johansson, E. Zuidema and C. Bolm, Chem.Eur. J., 2010, 16, 1348713499. 41 (a) J. F. Hartwig, Nature, 2008, 455, 314322; (b) S. L. Buchwald, Acc. Chem. Res., 2008, 41, 14391564, and references cited; (c) C. A. Fleckenstein and H. Plenio, Chem. Soc. Rev., 2010, 39, 694711; (d) C. Torborg and M. Beller, Adv. Synth. Catal., 2009, 351, 30273043; (e) L. Xue and Z. Lin, Chem. Soc. Rev., 2010, 39, 16921705. 42 (a) M. Ahlquist, P. Fristrup, D. Tanner and P.-O. Norrby, Organometallics, 2006, 25, 20662073; (b) J. C. Green, B. J. Herbert and R. Lonsdale, J. Organomet. Chem., 2005, 690, 60546067; (c) K. C. Lam, T. B. Marder and Z. Y. Lin, Organometallics, 2007, 26, 758760; (d) Z. Li, Y. Fu, Q.-X. Guo and L. Liu, Organometallics, 2008, 27, 40434049; (e) C. L. McMullin, B. Ruhle, M. Besora, A. G. Orpen, J. N. Harvey and N. Fey, J. Mol. Catal. A: Chem., 2010, 324, 4855. 43 J. Jover, N. Fey, M. Purdie, G. C. Lloyd-Jones and J. N. Harvey, J. Mol. Catal. A: Chem., 2010, 324, 3947. 44 F. Barrios-Landeros, B. P. Carrow and J. F. Hartwig, J. Am. Chem. Soc., 2009, 131, 81418154. 45 (a) C. Dai and G. C. Fu, J. Am. Chem. Soc., 2001, 123, 27192724; (b) J. F. Hartwig, M. Kawatsura, S. I. Hauck, K. H. Shaughnessy and L. M. Alcazar-Roman, J. Org. Chem., 1999, 64, 55755580. 46 V. G. Andrianov, I. S. Akhrem, N. M. Chistovalova and Y. T. Struchkov, Zh. Strukt. Khim. (Russ. J. Struct. Chem.), 1976, 17, 135. 47 C. Amatore and F. Puger, Organometallics, 1990, 9, 22762282. 48 E. A. Mitchell and M. C. Baird, Organometallics, 2007, 26, 52305238. 49 S. Grimme, R. Huenerbein and S. Ehrlich, ChemPhysChem, 2011, 12, 12581261. 50 N. Fey, J. N. Harvey, G. C. Lloyd-Jones, P. Murray, A. G. Orpen, R. Osborne and M. Purdie, Organometallics, 2008, 27, 13721383. 51 L. M. Alcazar-Roman, J. F. Hartwig, A. L. Rheingold, L. M. LiableSands and I. A. Guzei, J. Am. Chem. Soc., 2000, 122, 46184630.

Downloaded by University of Bristol on 25 June 2012 Published on 19 August 2011 on http://pubs.rsc.org | doi:10.1039/C1DT10909J

This journal is The Royal Society of Chemistry 2011

Dalton Trans., 2011, 40, 1118411191 | 11191

Anda mungkin juga menyukai