Catalytic Role of Nickel in the Decarbonylative Addition of

Oct 30, 2013 - Density functional theory calculations have been used to investigate the catalytic role of nickel(0) in the decarbonylative addition of...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/Organometallics

Catalytic Role of Nickel in the Decarbonylative Addition of Phthalimides to Alkynes Albert Poater,*,† Sai Vikrama Chaitanya Vummaleti,‡ and Luigi Cavallo‡,† †

Institut de Quı ́mica Computacional i Catàlisi and Departament de Quı ́mica, Universitat de Girona, E-17071 Girona, Catalonia, Spain KAUST Catalyst Center, 4700 King Abdullah University of Science and Technology, Thuwal 23955-6900, Kingdom of Saudi Arabia



S Supporting Information *

ABSTRACT: Density functional theory calculations have been used to investigate the catalytic role of nickel(0) in the decarbonylative addition of phthalimides to alkynes. According to Kurahashi et al. the plausible reaction mechanism involves a nucleophilic attack of nickel at an imide group, giving a sixmembered metallacycle, followed by a decarbonylation and insertion of an alkyne leading to a seven-membered metallacycle. Finally a reductive elimination process produces the desired product and regenerates the nickel(0) catalyst. In this paper, we present a full description of the complete reaction pathway along with possible alternative pathways, which are predicted to display higher upper barriers. Our computational results substantially confirm the proposed mechanism, offering a detailed geometrical and energetical understanding of all the elementary steps.



intermediate.8 This reaction has been quite well explored with PtCl2 as catalyst,9 since late-transition-metal species have proved their efficiency with respect to the activation of alkynes through a nucleophilic attack,10 a process pioneered by Murai et al.11 The formed C−M−N intermediate can undergo carboamination with an alkyne, as depicted in the second step in Scheme 1.12 The carboamination reaction thus forms C−C and C−N bonds simultaneously. In this context, Kurahashi et al. recently expanded the portfolio of this reaction to the nickel-catalyzed decarbonylative addition of a family of phthalimide, such as 1 of Scheme 2, to 4octyne to form isoquinolone 2 with high yields.7 The quinolone skeleton is widely found in various natural products, being a medicinal drug that exhibits a broad range of biological properties. Indeed, cycloenones serve both as key intermediates in the synthesis of a wide array of significant bioactive

INTRODUCTION Transition-metal-promoted activation and synthetic utilization of C−H, C−C, and C−N bonds is an increasingly important synthetic strategy for the atom-economical construction of C− C and C−X (X = N, O, S, P, halides) bonds.1 The C−H bond activation in solution by metal insertion has been relatively well studied in comparison to the C−C single-bond activation, often by oxidative addition to a transition metal,2 due to its low reactivity.3 Along the same line, examples of C−N bond activation by metal complexes are much less common due to the lack of reactivity. Thus, addition of a C−N bond, i.e. carboamination, to alkenes and alkynes, such as that shown in Scheme 1, is still considered a challenging subject.4 On the Scheme 1. Representation of the Metal-Catalyzed Decarbonylative Addition of Phthalimides to Alkynes

Scheme 2. Mechanistic Proposal of Ni-Catalyzed Decarbonylative Addition of Phthalimides to Alkynes

other hand, CN as a nitrile is one of the most useful organic compounds due to its versatility5 as a building block in organic synthesis, becoming key in many synthetic targets including natural products, pharmaceuticals, and materials.6 Bearing in mind that studies on the insertion of alkynes into phthalimides through decarbonylation are even more scarce,7 recent literature on C−N bond activation suggests that the direct oxidative addition of a C−N bond to a low-valent transition-metal catalyst generates the active C−M−N © 2013 American Chemical Society

Received: July 15, 2013 Published: October 30, 2013 6330

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

compounds such as prostaglandins13 and as interesting natural products in their own right as jamone or pentenomycins.14 A proposed plausible mechanism is outlined in Scheme 2. The Ni(0) center corresponds to a Ni complex in which the cyclooctadiene (cod) ligands of the original Ni(cod)2 complex have been replaced by phosphines. The first step is the nucleophilic attack of a Ni(0) center to the amide 1,15 giving first a six-membered metallacycle. Subsequent decarbonylation produces the active C−M−N species, which is followed by alkyne insertion16 into the Ni−C bond leading to the sevenmembered metallacyclic intermediate. A similar mechanism has been previously reported in nickel-catalyzed decarbonylative cross-coupling reactions of anhydrides involving an oxidative addition step.17 Finally, a reductive elimination step produces the desired product 2, regenerating the starting Ni(0) center.6,18,19 We believe that understanding the details of the metal insertion into the C−N bonds and knowledge of the whole reaction mechanism could help the future development of this important reaction. Therefore, in this paper we investigate in detail the reaction pathway proposed in Scheme 2 by Kurahashi7 using density functional theory (DFT) calculations. For this system, bearing in mind previous calculations on Nicatalyzed reactions,20 here we analyze in detail the whole reaction pathway, from the initial activation of the Ni(cod)2 complex to the final product 2. Particular attention will be given to the decarbonylation step, since it requires the difficult activation of a C−C bond with aromatic character.21



Finally, to evaluate the strength of bonds, a Mayer bond order (MBO) analysis has been envisaged. We have calculated MBOs through the expression

BAB = 2

∑ ∑ [(PαS)μν (PαS)νμ + (P β S)μν (P β S)νμ ] (2)

μ∈A ν∈B α

β

where S is the atomic orbital overlap matrix and P and P are the density matrices for the α and β electrons, respectively.31 The Mayer definition can be seen as an extension of the Wiberg index.32 This leads to the classical integer values for homonuclear diatomics when minimal or small basis sets are used. Noninteger values are found for larger basis sets and in more complicated molecules, and these reflect the polarized character of the bonds as well as delocalization and multicenter effects. Mayer bond orders are a valuable tool in the analysis of the bonding in main-group and transition-metal33 systems. It is worth mentioning here that, for all the species presented in the paper, we have computed the closed-shell singlet and triplet electronic states. Since the triplet state is always much higher in energy, in the following sections, we confine our discussion only to the singlet ground state.



RESULTS AND DISCUSSION We started with a geometry optimization of substrate 1, product 2, and its possible isomer 2′ (see Figure 1). The

THEORETICAL METHODS

All of the density functional theory (DFT) calculations were performed using the Gaussian09 package.22 The BP86 GGA functional of Becke and Perdew23 was used with the standard split-valence basis set with a polarization function of Ahlrichs and co-workers for H, C, N, and P atoms (SVP keyword in Gaussian),24 while the quasirelativistic small-core Stuttgart effective core potential (ECP) together with the SDD basis set was used for nickel (SDD keyword in Gaussian09), for geometry optimizations.25 The reported energies have been obtained via single-point energy calculations with the M06 functional26 with the triple-ζ basis set of Ahlrichs for main-group atoms (TZVP keyword in Gaussian09),27 using the computational scheme described above for geometries (BP86 with SVP and SDD basis sets). Solvent effects of toluene were included with the default Gaussian PCM implementation.28 The M06 energy calculations were carried out with the scf=tight and integral (grid=ultrafinegrid) keywords. Zero-point energies and thermal corrections calculated at the BP86 level were added to the M06 functional in solvent energies to approximate free energies in solvent. As a reference system we used the separated reactants, instead of using the prereactant complex approach. In some cases we discuss the change in the local aromaticity of a given ring. As a structure-based measure, we used the harmonic oscillator model of aromaticity (HOMA) index, defined by Kruszewski and Krygowski as29

HOMA = 1 −

α n

Figure 1. Substrate 1, product 2 of the nickel-catalyzed decarbonylative addition of phthalimide to alkyne, and its isomer 2′.

optimized geometry of 2, the only compound with known crystal data, is in perfect agreement with the experimental structure (rmsd = 0.025 Å on distances and 0.7° on angles).34,35 From an energy point of view, product 2 is 27.7 kcal mol−1 lower in energy relative to substrate 1, suggesting that the reaction is thermodynamically favored.6 As for the relative stability of the product, isomer 2′ is 5.7 kcal mol−1 less stable than isomer 2, which supports the fact that 2′ is not formed experimentally. The possible explanation for the higher stability of 2 is the following: in substrate 1, the C−N bond distances (C of both carbonyls and N of amide) are 1.427 and 1.435 Å and the corresponding Mayer bond orders are 1.016 and 0.994, respectively. The longer and weaker C−N bond corresponds to that closer to the N atom of the pyridine-type ring of 1, suggesting that this bond can be more easily activated, thus leading to isomer 2 rather than isomer 2′. For this reason, we did not investigate in detail activation of the C−N bond leading to isomer 2′. The potential free energy surface of the first half (corresponding to decarbonylation of the substrate 1 up to intermediate 7) and of the second half (alkyne addition from intermediate 7 to the product 2) of the studied reaction pathway and the structural details are given in Figures 2 and 3, respectively. Going into detail, the reaction starts from the neutral Ni(cod)2 species 3. In the presence of excess of trimethylphosphine, a cod ligand of 3 can be replaced with two PMe3 molecules, giving the Ni species 4. Displacement of cod is favored by 7.5 kcal mol−1. In principle, also the second cod ligand could be replaced by two more PMe3 molecules but, for

n

∑ (R opt − R i)2 i=1

(1)

where n is the number of bonds considered and α is an empirical constant (for CC bonds α = 257.7 and for CN bonds α = 93.2) fixed to give HOMA = 0 for a model nonaromatic system and HOMA = 1 for a system with all bonds equal to an optimal value Ropt, which is 1.388 Å for CC bonds (1.334 Å for CN bonds), assumed to be achieved for fully aromatic systems.27a Ri stands for a running bond length. HOMA calculations have proved to be useful in similar studies.30 6331

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

Figure 2. Stationary points along the first half of the reaction pathway. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å (from species 5 all methyl groups of phosphines are omitted for the sake of clarity).

Figure 3. Stationary points along the second half of the reaction pathway. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å.

0.898 and 0.918 for species 6 and 7, respectively, in comparison to 0.933 for species 1). The overall barrier for the decarbonylation step, through transition state 6-7, amounts to 7.9 kcal mol−1. Moving to the second half of the reaction pathway (see Figure 3), the key intermediate 7 loses a phosphine ligand to form 8, with a release of only 0.7 kcal mol−1. This stabilization might be much larger with sterically demanding phosphines, such as PPh3 and PCy3. Furthemore, it is worth mentioning that coordination of the CO molecule in 7 and 8 is highly labile, since the corresponding Ni−CO dissociation energies are only 2.7 and 8.1 kcal mol−1, respectively. These relatively small energy differences between the decoordination of PMe3 and CO ligands from species 7 or 8 prompted us to explore all alternative coordination of PMe3 and CO to the nickel (see

the sake of simplicity, we considered replacement of the cod ligand of 4 by the substrate 1, leading to the Ni-coordinated intermediate 5, in which metal coordinates to the C−C double bond, joining the fused rings of 1. The substrate-bound intermediate 5 is only 1.4 kcal mol−1 below 4. Then, we explored a structural rearrangement of 5 leading to species 6, with the Ni atom inserted into the imide ring by breaking the weaker C−N bond (vide supra). This step proceeds through transition state 5-6 and requires overcoming a barrier of 19.1 kcal mol−1. Then, from intermediate 6 decarbonylation occurs, leading to the desired C−M−N key intermediate 7. This step is assisted by an interaction between the Ni atom and the closest aromatic C atom of the pyridine ring of 1. This interaction is promoted by an increase in the aromaticity character of the pyridine ring when moving from 6 to 7 (HOMA indices are 6332

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

Figures 4−9 in the next section for more details). The coordination of an alkyne substrate (4-octyne) to the nickel atom in 8 is endoergonic and leads to intermediate 9, which lies 15.0 kcal mol−1 above 8. The next step corresponds to insertion of the alkyne into the Ni−C(pyridine) bond through transition state 9-10, with a barrier of 46.2 kcal mol−1 above 5, leading to intermediate 10. The penultimate step corresponds to the formation of the C(alkyne)−N bond by reductive elimination from 10, through transition state 10-11. This is a rather low energy step, with a barrier of only 14.1 kcal mol−1, and leads to intermediate 11, with the product 8 coordinated to the Ni center and an energy gain of 34.2 kcal mol−1. As a final step, species 11 would release the product 2, regenerating the catalyst with recoordination of a free PMe3 ligand. Having completed the description of the whole catalytic cyle, we note that there are two rather high in energy transition states, namely 6-7 in Figure 2 and 9-10 in Figure 3. The former corresponds to the decarbonylation step and has an overall barrier of 26.9 kcal mol−1 from the most stable intermediate 5, while the latter corresponds to insertion of the alkyne into the Ni−C(pyridine) bond and has an overall energy barrier of 46.2 kcal mol−1 from the most stable intermediate 5. This is clearly a high barrier, which is however consistent with the remarkably high temperature of 110 °C needed experimentally to achieve a yield of at least 80% in 7 h.6 Nevertheless, to investigate the possible existence of alternative pathways that could present lower energy barriers, we performed additional calculations by considering first the possible initial insertion of the alkyne into the Ni−N bond, rather than into the Ni−C bond, and then we explored the effect of changing the number of ligands, specifically CO or PMe3, coordinated to the Ni center (vide infra). Focusing on the competition between alkyne addition to the Ni−C or Ni−N bond of 9, we located transition state 9-10′, corresponding to the insertion through the Ni−N bond. This alternative transition state depicted in Figure 4 costs 8.6 kcal

Figure 5. Stationary points along the first half of the reaction pathway without any phosphine bonded to nickel). Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å.

the first half of the reaction pathway with no phosphine bonded to nickel, species 12. From a thermodynamics point of view, our results clearly show that species 12−14 are much higher in energy with respect to 5−7. This destabilization might be attributed to the vacancies in the metal first coordination sphere. For instance, a comparison between the stabilities of species 14 and 7 suggests that the presence of two phosphines stabilizes the latter complex by 15.2 kcal mol−1. A similarly high energy profile is calculated when only one PMe3 molecule is bonded to nickel (see Figure 6). The higher

Figure 4. Computed structures for alkyne insertion into the Ni−N bond of 9. Free energies in solution in kcal mol−1 (toluene as the solvent) relative to species 3 are given. Selected distances are given in Å.

Figure 6. Stationary points along the first half of the reaction pathway with one phosphine bonded to the nickel. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å.

mol−1 more than transition state 9-10. Furthermore, isomer 10′ is 19.9 kcal mol−1 higher in energy with respect to 10, which excludes initial alkyne reactivity with the Ni−N bond of 9. To determine the effect of changing the number of ligands of the Ni center, we focused on the steps starting from 5 and 8 as key intermediates (see Figures 5-10). Bearing in mind the possible binding of up to three phosphines instead of one, Clot et al. evidenced that only a single phosphine is required when using a bulky phosphine such as PtBu3,36 while attention should be paid when smaller phosphines such as PMe3 are considered. According to our calculations, removing both phosphine ligands from 5 results in free energy profiles higher than those reported in Figure 2. Going into details, Figure 5 presents

barrier corresponding to the decarbonylation step increases by 11.9 kcal mol−1 in comparison to that of Figure 2. Although experimentally a very high temperature is used, which clearly favors dissociation of ligands, this difference should be enough to conclude that the monocoordinated phosphine pathway should not be operative. Next we move to the second half of the reaction, which is rate determining through transition state 9-10, associated with alkyne insertion into the Ni−C(pyridine) bond. Also in this case we explored different pathways corresponding to different 6333

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

ligand coordinations at the Ni center. We start by describing the situation of no phosphine or CO coordinated to the metal (see Figure 7). However, as in the case of the decarbonylation

Figure 7. Stationary points along the second half of the reaction pathway with neither a phosphine nor a CO bonded to nickel from species 17. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å.

Figure 8. Stationary points along the second half of the reaction pathway with a CO bonded to nickel from species 13. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequencies characterizing each transition state are given in brackets. Selected distances are given in Å.

steps of Figure 5, a very high energy profile is calculated for alkyne addition to the Ni−C(pyridine) bond in the presence of a naked Ni center. This profile is so high in energy that it can be excluded from occurring even under the drastic conditions used experimentally. Then, we considered the case when CO is still coordinated to the metal, while no phosphine is bound (see Figure 8). The upper barrier of this alternative pathway, to reach transition state 22-23, is placed 11.6 kcal mol−1 higher in energy relative to transition state 9-10 of Figure 3, while the overall barrier from the alkyne coordinates species 21 is 25.3 kcal mol−1, versus an overall barrier of 30.8 kcal mol−1 relative to the alkyne free intermediate 8 in Figure 3. As discussed in the case of the first half of the reaction (see Figure 6), this difference should be enough to conclude that alkyne insertion along the pathway with no PMe3 coordinated to the Ni center should not be operative. Next we investigated the situation where only one PMe3, and no CO, is coordinated to the nickel (see Figure 9). In this case, the rate-determining transition state 25-26 is only 3.9 kcal mol−1 higher in energy than transition state 9-10 of Figure 3, with an overall energy barrier of only 20 kcal mol−1 from the alkyne-coordinated intermediate 25. Furthermore, intermediate 25 can be easily reached by CO dissociation from 9, which costs only 0.7 kcal mol−1. The small free energy difference we calculated between transition states 9-10 and 25-26 indicates that CO dissociation could indeed promote alkyne addition in the presence of a single PMe3 molecule, making this another very viable pathway under the drastic conditions used experimentally. Finally, we studied the effect of having two phosphines and no CO bonded to nickel (see Figure 10). The computed energy profile presents only the two intermediates 28 and 29, which are 8.7 and 7.1 kcal mol−1 higher in energy in comparison to

Figure 9. Stationary points along the second half of the reaction pathway with a phosphine bonded to nickel from species 25. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequency characterizing the transition state structures is given in brackets. Selected distances are given in Å.

the corresponding intermediates 10 and 11, respectively, in Figure 3. All attempts to obtain intermediate 9 failed because of the steric hindrance caused by the presence of two phosphine ligands, which are in fact prone to decoordinate from nickel. Taken together, these observations suggest that alkyne addition in presence of two phosphines is unlikely. Overall, the above calculations indicate that under the experimental reaction conditions, i.e. high temperature (110 °C), excess phosphine, and long duration of the reaction, the rate-determining second half of the reaction, corresponding to alkyne insertion into the Ni−C(pyridine) bond, could occur along two pathways, which differ in the presence or the absence of the CO ligand from the Ni center (see Figures 3 and 9, respectively). 6334

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

for access to remarkable computational resources. A.P. thanks the Spanish MICINN for a Ramón y Cajal contract (RYC2009-05226), European Commission for a Career Integration Grant (CIG09-GA-2011-293900), and Generalitat de Catalunya (2012BE100824).



Figure 10. Stationary points along the second half of the reaction pathway with two phosphines bonded to nickel from species 23. Free energies in solution are given in kcal mol−1 (toluene as the solvent) relative to species 3. The imaginary frequency characterizing the transition state structures is given in brackets. Selected distances are given in Å.



CONCLUSIONS In summary, we have reported the first theoretical study describing the complete reaction pathway for the nickel(0)catalyzed decarbonylative addition of phthalimides to alkynes via C−N bond activation. We discussed the reaction in two parts. In the first half of the reaction, in the presence of excess PMe3, one of the two cod ligands of the starting Ni(cod)2 complex is replaced by two PMe3 ligands. Dissociation of the second cod ligand results in the bonding of the phthalamide substrate 1 through the aromatic rings. This intermediate initiates the insertion of the nickel into one of the amide bonds of the substrate to give a six-membered metallacycle. In the second half of the reaction, the Ni metallacycle species is then capable of promoting decarbonylation of the substrate, to form a crucial C−M−N intermediate. This step is followed by dissociation of one PMe3 molecule to decrease the steric hindrance around the Ni center, favoring coordination of the alkyne substrate. The penultimate step in the reaction pathway is insertion of alkyne into the Ni−C(pyridine) bond, followed by a reductive elimination step that liberates the observed product. Insertion of the alkyne into the Ni−C bond resulted as the rate-limiting step, with a barrier of 46.2 kcal mol−1 relative to the most stable intermediate. We also investigated possible alternative pathways by changing the number of ligands around the Ni center. These calculations indicated that CO dissociation prior to alkyne insertion could represent a viable competing reaction pathway under the drastic conditions used experimentally.



ASSOCIATED CONTENT



AUTHOR INFORMATION

REFERENCES

(1) (a) Science of Synthesis; Trost, B. M., Ed.; Thieme: Stuttgart, Germany, 2001; Vol. 1. (b) Tsuji, J. Transition Metal Reagents and Catalysts; Wiley: New York, 2000. (c) Comprehensive Organometallic Chemistry II; Abel, E. W., Stone, F. G. A., Wilkinson, G., Eds.; Pergamon: Oxford, U.K., 1995; Vol. 12. (d) Comprehensive Organic Synthesis; Trost, B. M., Fleming, I., Eds.; Pergamon: Oxford, U.K., 1991; Vol. 4. (2) (a) Crabtree, R. H. Chem. Rev. 1985, 85, 245−269. (b) Rybtchinski, B.; Milstein, D. Angew. Chem., Int. Ed. 1999, 38, 870−883. (c) Murakami, M.; Ito, Y. In Activation of Unreactive Bonds and Organic Synthesis; Murai, S., Ed.; Springer: Berlin, 1999; p 97. (3) (a) Kaneda, K.; Azuma, H.; Wayaku, M.; Teranishi, S. Chem. Lett. 1974, 215−216. (b) Suggs, J. W.; Jun, C.-H. J. Chem. Soc., Chem. Commun. 1985, 92−93. (c) Edelbach, E. L.; Lachicotte, R. J.; Jones, W. D. J. Am. Chem. Soc. 1998, 120, 2843−2853. (d) Liou, S.-Y.; van der Boom, M. E.; Milstein, D. Chem. Commun. 1998, 687−688. (e) Jun, C.-H.; Lee, H. J. Am. Chem. Soc. 1999, 121, 880−881. (f) Jun, C.-H.; Lee, D.-Y.; Kim, Y.-H.; Lee, H. Organometallics 2001, 20, 2928−2931. (g) Miller, J. A. Tetrahedron Lett. 2001, 42, 6991−6993. (h) Müller, C.; Lachicotte, R. J.; Jones, W. D. Organometallics 2002, 21, 1975− 1981. (4) (a) Luo, F.-T.; Wang, R.-T. Tetrahedron Lett. 1992, 33, 6835− 6838. (b) Arcadi, A. Synlett 1997, 941−943. (c) Bouyssi, D.; Marcello, C.; Balme, G. Synlett 1997, 944−946. (d) Cacchi, S.; Fabrizi, G.; Marinelli, F.; Moro, L.; Pace, P. Synlett 1997, 1363−1366. (e) Arcadi, A.; Cacchi, S.; Marinelli, F. Tetrahedron Lett. 1992, 33, 3915−3918. (f) Iritani, K.; Matsubara, S.; Utimoto, K. Tetrahedron Lett. 1988, 29, 1799−1802. (g) Cacchi, S. J. Organomet. Chem. 1999, 576, 42−64. (h) Cacchi, S.; Fabrizi, G.; Pace, P. J. Org. Chem. 1998, 63, 1001− 1011. (5) Nakai, K.; Kurahashi, T.; Matsubara, S. Org. Lett. 2013, 15, 856− 859. (6) (a) Fleming, F. F. Nat. Prod. Rep. 1999, 16, 597−606. (b) Greenham, N. C.; Moratti, S. C.; Bradley, D. D. C.; Friend, R. H.; Holmes, A. B. Nature 1993, 365, 628−630. (7) (a) Kajita, Y.; Matsubara, S.; Kurahashi, T. J. Am. Chem. Soc. 2008, 130, 6058−6059. (b) Kajita, Y.; Kurahashi, T.; Matsubara, S. J. Am. Chem. Soc. 2008, 130, 17226−17227. (8) (a) Murahashi, S.-I.; Imada, Y.; Nishimura, K. Tetrahedron 1991, 50, 453−464. (b) Calet, S.; Urso, F.; Alper, H. J. Am. Chem. Soc. 1989, 111, 931−934. (c) Roberto, D.; Alper, H. J. Am. Chem. Soc. 1989, 111, 7539−7543. (9) (a) Fürstner, A.; Davies, P. W. J. Am. Chem. Soc. 2005, 127, 15024−15025. (b) Shimada, T.; Nakamura, I.; Yamamoto, Y. J. Am. Chem. Soc. 2004, 126, 10546−10547. (10) (a) Aubert, C.; Buisine, O.; Malacria, M. Chem. Rev. 2002, 102, 813−834. (b) Méndez, M.; Echavarren, A. M. Eur. J. Org. Chem. 2002, 15−28. (c) Lloyd-Jones, G. C. Org. Biomol. Chem. 2003, 1, 215−236. (d) Méndez, M.; Mamane, V.; Fürstner, A. Chemtracts 2003, 16, 397− 425. (11) Chatani, N.; Furukawa, N.; Sakurai, H.; Murai, S. Organometallics 1996, 15, 901−903. (12) (a) Chan, Y. W.; Renner, M. W.; Balch, A. L. Organometallics 1983, 2, 1888−1889. (b) Bonanno, J. B.; Henry, T. P.; Neithamer, D. R.; Wolczanski, P. T.; Lobkovsky, E. B. J. Am. Chem. Soc. 1996, 118, 5132−5133. (c) Tayebani, M.; Gambarotta, S.; Yap, G. Organometallics 1998, 17, 3639−3641. (d) Lin, B. L.; Clough, C. R.; Hillhouse, G. L. J. Am. Chem. Soc. 2002, 124, 2890−2891. (e) Fan, L.; Yang, L.; Guo, C.; Foxman, B. M.; Ozerov, O. V. Organometallics 2004, 23, 4778−4787. (f) Ozerov, O. V.; Guo, C.; Papkov, V. A.; Foxman, B. M. J. Am. Chem. Soc. 2004, 126, 4792−4793. (g) Weng, W.; Guo, C.; Moura, C.; Yang,

S Supporting Information *

Text, figures, and tables giving complete computational methods, Cartesian coordinates, absolute energies, and drawings of all the stationary points located. This material is available free of charge via the Internet at http://pubs.acs.org. Corresponding Author

*[email protected]. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS A.P. and L.C. thank the HPC team of Enea for using the ENEA-GRID and the HPC facilities CRESCO in Portici (Italy) 6335

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336

Organometallics

Article

L.; Foxman, B. M.; Ozerov, O. V. Organometallics 2005, 24, 3487− 3499. (h) Takano, K.; Inagaki, A.; Akita, M. Chem. Lett. 2006, 35, 434−435. (i) Burling, S.; Mahon, M. F.; Powell, R. E.; Whittlesey, M. K.; Williams, J. M. J. Am. Chem. Soc. 2006, 128, 13702−13703. (j) Ruck, R. T.; Zuckerman, R. L.; Krska, S. W.; Bergman, R. G. Angew. Chem., Int. Ed. 2004, 43, 5372−5374. (k) Liu, Z.; Larock, R. C. J. Am. Chem. Soc. 2005, 127, 13112−13113. (l) Yoshida, H.; Shirakawa, E.; Honda, Y.; Hiyama, T. Angew. Chem., Int. Ed. 2002, 41, 3247−3249. (m) Lei, Y.; Wrobleski, A. D.; Golden, J. E.; Powell, D. R.; Aubé, J. J. Am. Chem. Soc. 2005, 127, 4552−4553. (n) Ito, M.; Sakaguchi, A.; Kobayashi, C.; Ikariya, T. J. Am. Chem. Soc. 2007, 129, 290−291. (o) Ueno, S.; Chatani, N.; Kakiuchi, F. J. Am. Chem. Soc. 2007, 129, 6098−6099. (13) (a) Noyori, R.; Suzuki, M. Science 1993, 259, 44−45. (b) Straus, D. S.; Glass, C. K. Med. Res. Rev. 2001, 21, 185−210. (c) Corey, E. J. Angew. Chem., Int. Ed. Engl. 1991, 30, 455−465. (14) (a) Dobbs, D. A.; Vanhessche, K. P. M.; Brazi, E.; Rautenstrauch, V.; Lenoir, J.-Y.; Genêt, J.-P.; Wiles, J.; Bergens, S. H. Angew. Chem., Int. Ed. 2000, 39, 1992−1995. (b) Fràter, G.; Bajgrowicz, J. A.; Kraft, P. Tetrahedron 1998, 54, 7633−7703. (c) Seepersaud, M.; Al-Abed, Y. Tetrahedron Lett. 2000, 41, 4291− 4293. (15) (a) Trost, B. M.; Chen, F. Tetrahedron Lett. 1971, 12, 2603− 2607. (b) Sano, K.; Yamamoto, T.; Yamamoto, A. Chem. Lett. 1984, 941−944. (c) Sano, K.; Yamamoto, T.; Yamamoto, A. Bull. Chem. Soc. Jpn. 1984, 57, 2741−2747. (16) (a) Lochow, C. F.; Miller, R. G. J. Am. Chem. Soc. 1976, 98, 1281−1283. (b) Kondo, T.; Tsuji, Y.; Watanabe, Y. Tetrahedron Lett. 1987, 28, 6229−6230. (c) Tsuda, T.; Kiyoi, T.; Saegusa, T. J. Org. Chem. 1990, 55, 2554−2558. (d) Tanaka, K.; Fu, G. C. J. Am. Chem. Soc. 2001, 123, 11492−11493. (e) Jun, C.-H.; Jo, E.-A.; Park, J.-W. Eur. J. Org. Chem. 2007, 1869−1881. (f) Murakami, M.; Itahashi, T.; Ito, Y. J. Am. Chem. Soc. 2002, 124, 13976−13977. (g) Nozaki, K.; Sato, N.; Takaya, H. J. Org. Chem. 1994, 59, 2679−2681. (h) Nishihara, Y.; Inoue, Y.; Itazaki, M.; Takagi, K. Org. Lett. 2005, 7, 2639−2641. (i) Nakao, Y.; Hirata, Y.; Hiyama, T. J. Am. Chem. Soc. 2006, 128, 7420−7421. (j) Kokubo, K.; Matsumasa, K.; Miura, M.; Nomura, M. J. Org. Chem. 1996, 61, 6941−6946. (k) Hua, R.; Shimada, S.; Tanaka, M. J. Am. Chem. Soc. 1998, 120, 12365−12366. (l) Kashiwabara, T.; Kataoka, K.; Hua, R.; Shimada, S.; Tanaka, M. Org. Lett. 2005, 7, 2241−2244. (m) Hua, R.; Onozawa, S.; Tanaka, M. Chem. Eur. J. 2005, 11, 3621−3630. (n) Shirakawa, E.; Yamasaki, K.; Yoshida, H.; Hiyama, T. J. Am. Chem. Soc. 1999, 121, 10221−10222. (o) Hua, R.; Onozawa, S.; Tanaka, M. Organometallics 2000, 19, 3269−3271. (p) Shirakawa, E.; Yamamoto, Y.; Nakao, Y.; Tsuchimoto, T.; Hiyama, T. Chem. Commun. 2001, 1926−1927. (q) Nakao, Y.; Satoh, J.; Shirakawa, E.; Hiyama, T. Angew. Chem., Int. Ed. 2006, 45, 2271−2274. (r) Hua, R.; Takeda, H.; Onozawa, S.; Abe, Y.; Tanaka, M. J. Am. Chem. Soc. 2001, 123, 2899−2900. (s) Sugoh, K.; Kuniyasu, H.; Sugae, T.; Ohtaka, A.; Takai, Y.; Tanake, A.; Machino, C.; Kambe, N.; Kurosawa, H. J. Am. Chem. Soc. 2001, 123, 5108−5109. (t) Kuniyasu, H.; Kambe, N. Chem. Lett. 2006, 35, 1320−1325. (u) Hirai, T.; Kuniyasu, H.; Kato, T.; Kurata, Y.; Kambe, N. Org. Lett. 2003, 5, 3871−3873. (17) (a) Johnson, J. B.; Bercot, E. A.; Rowley, J. M.; Coates, G. W.; Rovis, T. J. Am. Chem. Soc. 2007, 129, 2718−2725. (b) O’Brien, E. M.; Bercot, E. A.; Rovis, T. J. Am. Chem. Soc. 2003, 125, 10498−10499. (18) Kadnikov, D. V.; Larock, R. C. J. Org. Chem. 2004, 69, 6772− 6780. (19) (a) Ruchelman, A. L.; Houghton, P. J.; Zhou, N.; Liu, A.; Liu, L. F.; LaVoie, E. J. J. Med. Chem. 2005, 48, 792−804. (b) Le, T. N.; Gang, S. G.; Cho, W.-J. J. Org. Chem. 2004, 69, 2768−2772. (20) (a) Sakaki, S.; Kitaura, K.; Morokuma, K.; Ohkubo, K. Inorg. Chem. 1983, 22, 104−108. (b) Ohnishi, Y.; Nakao, Y.; Sato, H.; Shigeyoshi Sakaki, S. J. Phys. Chem. A 2007, 111, 7915−7924. (c) Ohnishi, Y.; Nakao, Y.; Sato, H.; Nakao, Y.; Hiyama, T.; Sakaki, S. Organometallics 2009, 28, 2583−2594. (21) (a) Dyker, G. Angew. Chem., Int. Ed. 1999, 38, 1698. (b) Kakiuchi, F.; Chatani, N. Adv. Synth. Catal. 2003, 345, 1077−

1101. (c) Dick, A. R.; Sanford, M. S. Tetrahedron 2006, 62, 2439− 2463. (d) Godula, K.; Sames, D. Science 2006, 312, 67−72. (22) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö .; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09; Gaussian, Inc.: Wallingford CT, 2009. (23) (a) Perdew, J. P. Phys. Rev. B 1986, 33, 8822−8824. (b) Perdew, J. P. Phys. Rev. B 1986, 34, 7406−7406. (c) Becke, A. D. Phys. Rev. A 1988, 38, 3098−3100. (24) (a) Schaefer, A.; Horn, H.; Ahlrichs, R. J. Chem. Phys. 1992, 97, 2571−2577. (b) Schaefer, A.; Huber, C.; Ahlrichs, R. J. Chem. Phys. 1994, 100, 5829−5835. (25) (a) Haeusermann, U.; Dolg, M.; Stoll, H.; Preuss, H. Mol. Phys. 1993, 78, 1211−1224. (b) Kuechle, W.; Dolg, M.; Stoll, H.; Preuss, H. J. Chem. Phys. 1994, 100, 7535−7542. (c) Leininger, T.; Nicklass, A.; Stoll, H.; Dolg, M.; Schwerdtfeger, P. J. Chem. Phys. 1996, 105, 1052− 1059. (26) Zhao, Y.; Truhlar, D. Theor. Chem. Acc. 2008, 120, 215−241. (27) Weigend, F.; Ahlrichs, R. Phys. Chem. Chem. Phys. 2005, 7, 3297−3305. (28) (a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995− 2001. (b) Tomasi, J.; Persico, M. Chem. Rev. 1994, 94, 2027−2094. (29) (a) Kruszewski, J.; Krygowski, T. M. Tetrahedron Lett. 1972, 3839−3842. (b) Krygowski, T. M. J. Chem. Inf. Comput. Sci. 1993, 33, 70−78. (30) (a) Poater, A.; Cavallo, L. Inorg. Chem. 2009, 48, 2340−2342. (b) Poater, A.; Ribas, X.; Llobet, A.; Cavallo, L.; Solà, M. J. Am. Chem. Soc. 2008, 130, 17710−17717. (31) (a) Mayer, I. Chem. Phys. Lett. 1983, 97, 270−277. (b) Mayer, I. Int. J. Quantum Chem. 1984, 26, 151−154. (32) Wiberg, K. B. Tetrahedron 1968, 24, 1083−1096. (33) (a) Poater, A.; Francesco, R.; Correa, A.; Cavallo, L. Dalton Trans. 2011, 40, 11066−11069. (b) Ramón, R. S.; Gaillard, S.; Poater, A.; Cavallo, L.; Slawin, A. M. Z.; Nolan, S. P. Chem. Eur. J. 2011, 17, 1238−1246. (c) Poater, A.; Cavallo, L. J. Mol. Catal. A 2010, 324, 75− 79. (34) Standard deviations for distances and angles: Sn−1 = ((∑iN= 1(CV − EV)2)/(N − 1))1/2, where CV is the calculated value, EV is the experimental value (X-ray data), and N is the number of distances or angles taken into account (distances and angles used are given in Table S2 of the Supporting Information). (35) (a) Mola, J.; Rodrı ́guez, M.; Romero, I.; Llobet, A.; Parella, T.; Poater, A.; Solà, M.; Benet-Buchholz, J. Inorg. Chem. 2006, 45, 10520− 10529. (b) Samojlowicz, C.; Bieniek, M.; Pazio, A.; Makal, A.; Wozniak, K.; Poater, A.; Cavallo, L.; Wojcik, J.; Zdanowski, K.; Grela, K. Chem. Eur. J. 2011, 17, 12981−12993. (c) Bianchettin, G.; Bonaccini, C.; Oliva, R.; Tramontano, A.; Cividini, A.; Casato, M.; Merlini, G.; Silini, E.; Mondelli, M. U. J. Virol. 2007, 81, 4564−4571. (d) Oliva, R.; Cavallo, L. J. Phys. Chem. B 2009, 113, 15670−15678. (36) Chaumontet, M.; Piccardi, R.; Audic, N.; Hitce, J.; Peglion, J.-L.; Clot, E.; Baudoin, O. J. Am. Chem. Soc. 2008, 130, 15157−15166.

6336

dx.doi.org/10.1021/om400693v | Organometallics 2013, 32, 6330−6336