DFT Study of the Mechanisms of Iron-Catalyzed ... - ACS Publications

Nov 22, 2016 - catalytic cycle is the reductive elimination step, where the energy ..... Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Broth...
0 downloads 0 Views 2MB Size
Article pubs.acs.org/Organometallics

DFT Study of the Mechanisms of Iron-Catalyzed Regioselective Synthesis of α‑Aryl Carboxylic Acids from Styrene Derivatives and CO2 Qinghua Ren,* Ningning Wu, Ying Cai, and Jianhui Fang Department of Chemistry, Innovative Drug Research Center, Shanghai University, 99 Shangda Road, Shanghai 200444, People’s Republic of China S Supporting Information *

ABSTRACT: The mechanisms of highly regioselective iron(II)catalyzed synthesis of α-carboxylic acids from alkene derivatives and CO2 have been investigated using density functional theory (DFT) calculations at the B3LYP-D3 level. The results show that the overall catalytic cycle includes β-hydride elimination, hydrometalation, oxidative addition of EtMgBr, reductive elimination, and carboxylation using CO2. However, the first and second steps could be replaced by a favored concerted one-step mechanism without forming the iron hydride complex. The rate-limiting step for the whole catalytic cycle is the reductive elimination step, where the energy barrier ΔE is 37.3 kcal/mol in the gas phase and the Gibbs free energy in solvent THF ΔGsol is 30.3 kcal/mol, computed using the SMD method. The mechanisms to obtain the byproduct βcarboxylic acids are also studied.

1. INTRODUCTION Iron-catalyzed reactions have become increasingly important in the construction of complex molecular frameworks in modern organic synthesis. In comparison to the traditionally used transition-metal catalysts, such as those with Pd, Ni, Rh, Au, etc., iron catalysts have many advantages because of their toxicologically benign nature, low cost, natural abundance, and easy handling.1 Up to now, iron catalysts have been applied to various important reactions, such as substitutions, additions, eliminations, oxidations, reductions, cyclizations, iosmerizations, and rearrangements.2 Carbon dioxide is an extremely attractive C1 feedstock in organic synthesis due to its relatively nontoxic, abundant, and inexpensive advantages.3 Using CO2 as a carboxylative agent in transition-metal-catalyzed carboxylation is an important route to produce carboxylic acids and derivatives.4 For example, Hoberg et al. reported the nickel-catalyzed synthesis of carboxylic acids from alkenes and carbon dioxide.5 Rovis and co-workers developed the nickel-catalyzed reductive carboxylation of electron-deficient and electron-neutral styrene derivatives under an atmospheric pressure of CO2.6 Recently, many successful experimental works on the Ni-catalyzed reactions of carbon dioxide have been reported.7 Other transition-metal-catalyzed carboxylation using CO2 such as Pd-catalyzed,8 Au-catalyzed,9 and Cu-catalyzed10 reactions have also been broadly developed. The iron-catalyzed synthesis of carboxylic acids from alkenes and carbon dioxide was previously reported by Hoberg et al.11 and currently has matured into a class of effective iron-catalyzed hydrocarboxylation of aryl alkenes using CO2 as the C1 feedstock.12 Recently, Greenhalgh and Thomas have reported © XXXX American Chemical Society

that simple iron salts are effective catalysts for the hydrocarboxylation of aryl alkenes using CO2 as the C1 feedstock.13 Styrene and its derivatives can be catalyzed by FeCl2 to give αaryl carboxylic acids using EtMgBr as the hydride source. On the other hand, in the area of mechanistic investigations, Darensbourg and Yeung gave a concise review of computational studies of carbon dioxide−epoxide copolymerization reactions.14 Sperger et al. reviewed the computational studies of synthetically relevant homogeneous organometallic catalysis involving Ni, Pd, Ir, and Rh.15 Yuan and Lin had studied the mechanisms of nickel-catalyzed reductive carboxylation of styrenes with CO2 using DFT at B3LYP level.16 Xu and coworkers investigated the mechanisms of the formation of acrylates from ethylene and CO2 on Ni complexes using DFT calculations.17 Ostapowicz et al. studied Rh-catalyzed hydrocarboxylation of olefins with CO2 and H2 using DFT calculations.18 However, mechanistic investigations of iron-catalyzed reactions have been less reported.19 Wallentin and co-workers gave a proposed mechanistic pathway of iron-mediated crosscoupling reactions of organohalides and Grignard reagents.20 Bedford chased the active species in Fe-catalyzed cross-coupling reactions.21 Thomas and co-workers gave a mechanistic proposal for their experimental results which included the hydrometalation and β-hydride elimination steps.13,22 However, until now, Greenhalgh and Thomas had not theoretically investigated the mechanisms of the iron-catalyzed reactions for the regioselective synthesis of α-aryl carboxylic acids. Received: August 27, 2016

A

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics Mechanistic insight into the catalytic reactions not only is academically important but also will help to maximize the efficiency of processes or even to develop entirely new transformations.20 It is valuable to carry out more comprehensive computational studies to explore the mechanisms of those kinds of reactions. The present study is the first one to explore the detailed mechanisms of highly regioselective iron(II)-catalyzed synthesis of α-carboxylic acids from alkene derivatives and carbon dioxide using DFT calculations. We hope to clarify the questions of which catalytic mechanism is feasible and which step is the ratedetermining step.

Scheme 2. Outline of the Mechanisms of the Iron-Catalyzed Reaction for the Main Product P1 from Styrene (R1), Ethylmagnesium Bromide (R2), and Carbon Dioxide

2. COMPUTATIONAL DETAILS All calculations were carried out using density functional theory (DFT) with the B3LYP-D3 method where the Gaussian keyword “EmpiricalDispersion = GD3” was used.23 Sameera et al. used the B3LYP-D3 method to study the mechanisms of iron-catalyzed asymmetric Mukaiyama aldol reaction.24 Heggen and Thiel also used the B3LYP+D functional to investigate the mechanism of ironcatalyzed cross-coupling reactions.25 For our calculations, the 6-31g (d,p) basis set was used for C, H, N, and O,26 the 6-311g (d,p) basis set was used for Br, and the SDD quasirelativistic pseudopotential and associated basis set was used for Fe.27 The gas-phase geometries of all intermediates and transition states were fully optimized without any symmetry restriction, following the vibrational frequencies analysis to ensure that the local minima have 0 imaginary frequencies and the transition state has exactly 1. The IRC calculations were used to confirm the nature of the transition states. The SMD model on the gas-phase optimized geometries was used as the solvent effect calculated model with THF as the solvent.28 All DFT calculations were implemented using the Gaussian 09 program.29

3. RESULTS AND DISCUSSION 3.1. Model Reaction. We select one typical reaction in the experimental work of Greenhalgh and Thomas13 as an example, which is shown in Scheme 1. Styrene (R1), ethylmagnesium

reacts with EtMgBr (R2) to form the active catalyst 1 (L-Fe+Et), where the iron(II) ion connects with one Et group and has a +1 charge (considering the computational costs, another Cl− ion has not been included in the calculations). The starting step is the same as that in the proposal of Greenhalgh and Thomas.13 The active catalyst 1 was proposed by Greenhalgh and Thomas.13 For the catalytic cycle starting from the active catalyst 1, we have investigated the low-spin state of the singlet FeII catalyst and the high-spin state of the quintet FeII catalyst. Our calculations show that the low-spin state of the singlet FeII catalyst is favored over the high-spin state of the quintet FeII catalyst; thus, here we only discuss the results of the low-spin state of the singlet FeII catalytic processes. The results for the quintet FeII catalyst are given in Table S3 and Figure S1 in the Supporting Information. The whole catalytic cycle includes several basic steps. 3.2.1. β-Hydride Elimination. The approach of styrene (R1) toward the active iron(II) catalyst 1 (L-Fe+-Et) leads to the formation of complex 2. From 2, the β-H atom of the ethyl part moves from the C atom to the iron atom to form the intermediate 3, which passes through the transition state TS1. 3.2.2. Hydrometalation. The iron hydride complex 3 can undergo hydrometalation to form the intermediate 4 after passing through the transition sate TS2. The proposed mechanism of Greenhalgh and Thomas includes these two steps.13 In the optimized geometry of TS1 which is shown in Figure 1, we can see that the β-H atom of ethyl moves from the carbon atom to the iron atom. The C−H and Fe−H distances are 1.99 and 1.46 Å in TS1, in comparison to 1.09 Å for the C−

Scheme 1. Model Reaction of Styrene (R1), Ethylmagnesium Bromide (R2), and Carbon Dioxide

bromide (R2), and carbon dioxide (CO2) are chosen as the substrates, and FeCl2 is used as the catalyst precursor. The solvent is THF and N,N′-dimethylethanediamine (DMEDA) is used as the ligand (L) in place of N,N,N′,N′-tetramethylethylenediamine (TMEDA) in order to reduce computational costs. 3.2. Full Catalytic Cycle. In the last few decades, mechanistic investigations in transition-metal-catalyzed crosscoupling reactions have been widely reported.30 Yuan and Lin investigated the detailed mechanisms for the carboxylative cyclization of propargylamine using CO2 catalyzed by NHCgold(I) complexes with DFT calculations.31 On the basis of previous studies and the proposed mechanism of Greenhalgh and Thomas,13 we give a mechanistic proposal to form the main product P1 which is outlined in Scheme 2. The precatalyst FeCl2 initially combines with the ligand L (DMEDA) and B

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 1. Fully optimized structures of 1−3 and the transition states TS1−TS3. Color scheme: C, cyan; N, blue; H, white; Fe, mauve. Distances are given in Å.

H bond in 2 and 1.50 Å for the Fe−H bond in 3, respectively. Similarly, it can be seen that the H atom moves from the iron atom to the carbon atom of the styrene part in the optimized geometry of TS2 shown in Figure 1. The Fe−H and C−H distances are 1.47 and 1.83 Å in TS2. In Bleith and Gade’s work on the mechanism of the iron(II)-catalyzed hydroslylation of ketones, we can see similar processes which include the steps of the σ-bond metathesis of an alkoxide complex with the silane, subsequent coordination of the ketone to the iron hydride complex, and insertion of the ketone into the Fe−H bond to reestablish the alkoxide complex.32 However, our calculations show that the β-H atom of the ethyl part in 2 can move directly from the C atom of the ethyl part to the C atom of the styrene part around the iron center without forming the iron hydride complex 3, which passes through the concerted transition state TS3 to form 4. From Figure 1, we can see that the H−C distance in the ethyl part in the optimized geometry of TS3 is 1.67 Å and the H−C distance in the styrene part in the optimized geometry of TS3 is 1.63 Å. A concerted mechanism was not proposed by Greenhalgh and Thomas.13 3.2.3. Oxidative Addition of EtMgBr. The molecular structure of the intermediate 4 is like the complex of ethylene and the intermediate 5, where their optimized geometries are shown in Figure 2. We can see that the distance of Fe and C of ethylene in 4 is long at 2.38 Å. Then the complex 4 can release ethylene to form the intermediate 5. From 5, EtMgBr is approached to form the addition product intermediate 6 after passing through the transition state TS4. It can be seen that the C−Mg, C−Fe, and Mg−Fe bonds are 2.14, 2.53, and 2.54 Å in

Figure 2. Fully optimized structures of 4−8 and the transition states TS4−TS6. Color scheme: C, cyan; N, blue; H, white; Fe, mauve; Mg, orange; Br, violet. Distances are given in Å.

TS4, which are shown in Figure 2. The C−Fe (2.07 Å) and Mg−Fe bonds (2.49 Å) are formed in the intermediate 6 shown in Figure 2. 3.2.4. Reductive Elimination. The reductive elimination of the phenyl part and MgBr part in intermediate 6 gives the intermediate 7 after passing though the transition state TS5, where their optimized geometries are shown in Figure 2. We can see that the Fe−C (of the alkyl-substituted styrene part), Fe−Mg, and Mg−C distances are 1.99, 2.49, and 2.38 Å in the intermediate 6, respectively. The Mg−C distance becomes shorter at 2.18 Å in the intermediate 7 in comparison to that of 2.38 Å in the intermediate 6, which means that the alkylsubstituted styrene part connects with the MgBr part in 7. 3.2.5. Carboxylation using CO2. The complex 7 reacts with carbon dioxide to form the intermediate 8 after passing though the transition state TS6 and regenerates the original active catalyst 1. Finally, the intermediate 8 reacts with H+ ion to form the final product P1. The optimized geometries of TS6 and 8 are also shown in Figure 2. A mechanistic study by Hazari and co-workers on the Pd-catalyzed reaction for the catalytic hydroboration of CO2 reported similar processes.8a 3.3. Energy Profile. The electronic energy profiles for the overall catalytic cycle are depicted in Figure 3a. We can see that C

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

mol, respectively, in comparison to 21.0 and 8.5 kcal/mol in the gas phase. Their values become larger when considering the solvent effect. The ΔGsol value of TS3 is 18.3 kcal/mol, in comparison to 18.1 kcal/mol in the gas phase. It almost remains unchanged. From Figure 3b, it can be seen that ΔGsol value of TS3 is much smaller than the ΔGsol value of TS1, which means that the concerted path from 2 to 4 without formation of the iron hydride complex 3 is still favored when the solvent effect is considered. We can also see that the rate-limiting step in solvent THF is still the reductive elimination step when passing through the transition state TS5, where the value of ΔGsol in solvent THF is 30.3 kcal/mol. The solvent effect does not change the conclusion. The mechanism calculated in the gas phase can be effectively used for the model reaction. All calculated energy values of ΔE, ΔG, and ΔGsol for the processes to obtain the main product P1 can be seen in Table S1 in the Supporting Information. 3.5. Mechanism To Obtain the Byproduct. In the experimental works of Greenhalgh and Thomas,13 the byproduct is β-carboxylic acid P2, which is shown in Scheme 1. We have also studied the catalytic mechanism to form P2 byproduct, where the full catalytic cycle is outlined in Scheme 3. The whole catalytic cycle also includes several basic steps Scheme 3. Outline of the Mechanism of the Iron-Catalyzed Reaction for the Byproduct P2 from Styrene (R1), Ethylmagnesium Bromide (R2), and Carbon Dioxide Figure 3. (a) Energy profiles for the overall catalytic cycle to obtain the main product. (b) Gibbs free energies in the gas phase, ΔG (solid line), and in solvent THF, ΔGsol (dashed line), for the overall catalytic cycle to obtain the main product P1 (energies in kcal/mol).

the catalytic cycle has two paths from 2 to 4. One is proposed by Greenhalgh and Thomas,13 which passes through the transition state TS1 and TS2, where the energy barriers of TS1 and TS2 are 23.8 and 15.0 kcal/mol, respectively. Another path is to pass directly through the concerted transition state TS3, where the energy barrier of TS3 is 18.8 kcal/mol, which is smaller than that of TS1 (23.8 kcal/mol). This means that the concerted catalytic path without formation of the iron hydride complex 3 is favored. From Figure 3a, we can see the energy barrier from 4 to form the addition product intermediate 6 after passing through the transition state TS4, which needs 18.8 kcal/mol. For the overall catalytic cycle, the energy barrier of the reductive elimination step to get the intermediate 7 after passing though the transition state TS5 is the highest, at 37.3 kcal/mol. This means that the reductive elimination step is the rate-determining step in the whole catalytic cycle. The reductive elimination is often identified as the rate-determining step in iron-catalyzed crosscoupling reactions.19a,33,2a 3.4. Solvent Effect. All of the above calculations were performed in the gas phase. In order to clarify whether solvent can have a significant effect on our calculated system, we calculate the Gibbs free energy in solvent THF for the whole catalytic cycle. All of the solution-phase free energies in this article correspond to the reference state of 1 mol/L and 298 K. The Gibbs free energies in the gas phase, ΔG, and in solvent THF, ΔGsol, for the overall catalytic cycle are shown in Figure 3b. The former is depicted as a solid line, and the latter is drawn as a dashed line. It can be seen that the Gibbs free energies of TS1 and TS2 in THF solvent are 26.3 and 11.6 kcal/

which are the same as the mechanism to give the main product P1: (I) β-hydride elimination, (II) hydrometalation, (III) oxidative addition of EtMgBr, (IV) reductive elimination, and (V) carboxylation using CO2. However, the main difference between them is the hydrometalation step. The H atom moves to the α-C of styrene to produce the main product P1 in the main cycle shown in Scheme 2; however, the H atom moves to the β-C of styrene to product the byproduct P2 in the cycle shown in Scheme 3. Other steps are similar. D

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

TS3a (21.5 kcal/mol) is still smaller than ΔGsol of TS1 (26.3 kcal/mol), which means that the concerted path from 2 to 4a without formation of the iron hydride complex 3 is still favored when the solvent effect is considered. Furthermore, the ΔGsol of TS3a (21.5 kcal/mol) is also larger than that of the TS3 (18.3 kcal/mol), which means it is difficult to form 4a (and finally to produce the byproduct P2) from 2 in comparison to formation of 4 (and finally to produce the main product P1) from 2 when the solvent effect is considered. The results are in agreement with the experimental results reported by Greenhalgh and Thomas.13 From Figure 4b, it also can be seen that the value of ΔGsol in solvent THF for the transition state TS6a is 25.6 kcal/mol, which is smaller than that of the transition state TS5a (35.5 kcal/mol). Therefore, the rate-limiting step in solvent THF for the catalytic cycle to produce the byproduct P2 is still the reductive elimination step when passing through the transition state TS5a, which is the same as the mechanism of forming the main product P1. We can also see that the value of ΔGsol for the transition state TS5a (35.5 kcal/mol) is still larger than that of TS5 (30.3 kcal/mol), which means it is difficult to produce the byproduct P2 in comparison to production of the main product P1 when the solvent effect is considered. The conclusion is consistent with the experiments reported by Greenhalgh and Thomas.13 All calculated energy values of ΔE, ΔG, and ΔGsol for the processes to obtain the byproduct P2 can be seen in Table S2 in the Supporting Information.

The energy profiles for the mechanisms to produce the byproduct P2 are shown in Figure 4a. It can be seen that one

4. CONCLUSION We have studied the detailed mechanisms of highly regioselective iron-catalyzed synthesis of carboxylic acids from styrene and CO2 using DFT calculations. Our results show that the catalytic cycle includes β-hydride elimination, hydrometalation, oxidative addition of EtMgBr, reductive elimination, and carboxylation using CO2. However, the first and second steps can be replaced by a concerted one-step mechanism without formation of the iron hydride complex. The energy barrier of 18.8 kcal/mol for the concerted mechanism is smaller than that of 23.8 kcal/mol for the β-hydride elimination step. The rate-limiting step for the whole catalytic cycle to produce P1 is the reductive elimination step with an energy barrier of 37.3 kcal/mol in the gas phase, and the Gibbs free energy in solvent THF ΔGsol is 30.3 kcal/mol. The solvent effect does not change the conclusion. The calculated result for the mechanism to produce the byproduct P2 shows that the energy barrier for the rate-limiting step is much higher than that to produce the main product P1. The rate-limiting step for the whole catalytic cycle to produce P2 byproduct is still the reductive elimination step when the solvent effect is considered.

Figure 4. (a) Energy profiles for the overall catalytic cycle to obtain the byproduct P2. (b) Gibbs free energies in the gas phase, ΔG (solid line), and in solvent THF, ΔGsol (dashed line), for the overall catalytic cycle to obtain the byproduct P2 (energies in kcal/mol).

path from 2 to 4 is to pass through the transition states TS1 and TS2a, where the energy barriers of TS1 and TS2a are 23.8 and 13.6 kcal/mol, respectively. The favored path from 2 to 4 is still to pass directly through the concerted transition state TS3a, where the energy barrier of TS3a is 22.4 kcal, which is smaller than that of TS1 (23.8 kcal/mol). At the same time, the energy barrier of TS3a (22.4 kcal/mol) is larger than that of TS3 (18.8 kcal/mol). This means that it is difficult to form 4a (and finally to produce the byproduct P2) from 2 in comparison to the formation of 4 (and finally to produce the main product P1) from 2. The results are in agreement with the experimental results reported by Greenhalgh and Thomas.13 From Figure 4a, we also can see that the energy barrier of TS5a is 38.1 kcal/mol, which is larger than that of TS5 (37.3 kcal/mol) of the rate-determining step for the main cycle. Finally, the energy barrier of TS6a is 43.2 kcal/mol, which means that the last step becomes the rate-determining step in the whole catalytic cycle to give the byproduct P2. Therefore, the energy barrier for the rate-determining step to product P2 (TS6a = 43.2 kcal/mol) is much higher than that to produce P1 (TS5 = 37.3 kcal/mol). This can explain the transferring preference to produce P1 but not P2. The conclusion is consistent with the experiments reported by Greenhalgh and Thomas.13 The Gibbs free energies in the gas phase, ΔG, and in solvent THF, ΔGsol, for the overall catalytic cycle to produce P2 are shown in Figure 4b. The former is depicted as a solid line, and the latter is drawn as a dashed line. It can be seen that ΔGsol of



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.6b00681. Calculated energy values for all optimized structures, energy profiles for the quintet FeII catalytic process, and optimized structures (PDF) Detailed molecular coordinates for all optimized structures (XYZ) E

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics



Guiet, L.; Kaslin, A.; Murphy, E.; Thomas, S. P. RSC Adv. 2013, 3, 25698−25701. (d) Greenhalgh, M. D.; Jones, A. S.; Thomas, S. P. ChemCatChem 2015, 7, 190−222. (13) Greenhalgh, M. D.; Thomas, S. P. J. Am. Chem. Soc. 2012, 134, 11900−11903. (14) Darensbourg, D. J.; Yeung, A. D. Polym. Chem. 2014, 5, 3949− 3962. (15) Sperger, T.; Sanhueza, I. A.; Kalvet, I.; Schoenebeck, F. Chem. Rev. 2015, 115, 9532−9586. (16) Yuan, R.; Lin, Z. Organometallics 2014, 33, 7147−7156. (17) Guo, W.; Michel, C.; Schwiedernoch, R.; Wischert, R.; Xu, X.; Sautet, P. Organometallics 2014, 33, 6369−6380. (18) Ostapowicz, T. G.; Holscher, M.; Leitner, W. Eur. J. Inorg. Chem. 2012, 2012, 5632−5641. (19) (a) Kleimark, J.; Hedstrom, A.; Larsson, P. F.; Johansson, C.; Norrby, P. ChemCatChem 2009, 1, 152−161. (b) Bekhradnia, A.; Norrby, P. O. Dalton Trans. 2015, 44, 3959−3962. (c) Bogdanovic, B.; Schwickardi, M. Angew. Chem., Int. Ed. 2000, 39, 4610. (d) Ren, Q.; Guan, S.; Shen, X.; Fang, J. Organometallics 2014, 33, 1423−1430. (e) Ren, Q.; Shen, X.; Wan, J.; Fang, J. Organometallics 2015, 34, 1129−1136. (f) Adhikari, D. RSC Adv. 2015, 5, 95379−95384. (20) Cassani, C.; Bergonzini, G.; Wallentin, C. J. ACS Catal. 2016, 6, 1640−1648. (21) Bedford, R. B. Acc. Chem. Res. 2015, 48, 1485−1493. (22) Greenhalgh, M. D.; Kolodziej, A.; Sinclair, F.; Thomas, S. P. Organometallics 2014, 33, 5811−5819. (23) (a) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098−3100. (b) Becke, A. D. J. Chem. Phys. 1993, 98, 5648−5652. (c) Lee, C. T.; Yang, W. T.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (d) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. J. Phys. Chem. 1994, 98, 11623− 11627. (e) Hertwig, R. H.; Koch, W. Chem. Phys. Lett. 1997, 268, 345−351. (f) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys. 2010, 132, 154104. (24) Sameera, W. M.; Hatanaka, M.; Kitanosono, T.; Kobayashi, S.; Morokuma, K. J. Am. Chem. Soc. 2015, 137, 11085−11094. (25) Heggen, B.; Thiel, W. J. Organomet. Chem. 2016, 804, 42−47. (26) (a) Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. J. Chem. Phys. 1980, 72, 650−654. (b) Lin, X.; Phillips, D. L. J. Org. Chem. 2008, 73, 3680−3688. (27) (a) Fuentealba, P.; Preuss, H.; Stoll, H.; Vonszentpaly, L. Chem. Phys. Lett. 1982, 89, 418−422. (b) Dunning, T. H.; Hay, P. J. Modern Theoretical Chemistry; Plenum: New York, 1977; Vol. 3, pp 1−28. (28) Marenich, A. V.; Cramer, C. J.; Truhlar, D. G. J. Phys. Chem. B 2009, 113, 6378−6396. (29) 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, Revision D.01; Gaussian, Inc., Wallingford, CT, 2009. (30) (a) Villa, M.; Jacobi von Wangelin, A. J. Angew. Chem., Int. Ed. 2015, 54, 11906−11908. (b) Bauer, G.; Wodrich, M. D.; Scopelliti, R.; Hu, X. Organometallics 2015, 34, 289−298. (c) Zheng, C.; Zhuo, C.; You, S. J. Am. Chem. Soc. 2014, 136, 16251−16259. (d) Yamamoto, Y. Organometallics 2013, 32, 5201−5211. (31) Yuan, R.; Lin, Z. ACS Catal. 2015, 5, 2866−2872. (32) Bleith, T.; Gade, L. H. J. Am. Chem. Soc. 2016, 138, 4972−4983.

AUTHOR INFORMATION

Corresponding Author

*Q.R.: tel, +86 21 66132404; fax, +86 21 66134594; e-mail, [email protected]. ORCID

Qinghua Ren: 0000-0001-8565-9653 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the Shanghai Higher Education Connotation Construction “085” Project “Materials Genome Engineering” Funding (B.58-B111-12-101, B.58-B111-12-103) and the high-performance computing platform of Shanghai University.



REFERENCES

(1) (a) Mako, T. L.; Byers, J. A. Inorg. Chem. Front. 2016, 3, 766− 790. (b) Kuzmina, O. M.; Steib, A. K.; Moyeux, A.; Cahiez, G.; Knochel, P. Synthesis 2015, 47, 1696−1705. (c) Merel, D. S.; Tran Do, M. L.; Gaillard, S.; Dupau, P.; Renaud, J. Coord. Chem. Rev. 2015, 288, 50−68. (d) Bauer, E. B. Curr. Org. Chem. 2008, 12, 1341−1369. (2) (a) Bauer, I.; Knolker, H. J. Chem. Rev. 2015, 115, 3170−3387. (b) Bolm, C.; Legros, J.; Paih, J. L.; Zani, L. Chem. Rev. 2004, 104, 6217−6254. (3) (a) Jessop, P. G.; Joo, F.; Tai, C. C. Coord. Chem. Rev. 2004, 248, 2425−2442. (b) Jacquet, O.; Frogneux, X.; Gomes, C. D. N.; Cantat, T. Chem. Sci. 2013, 4, 2127−2131. (c) Ostapowicz, T. G.; Schmitz, M.; Krystof, M.; Klankermayer, J.; Leitner, W. Angew. Chem., Int. Ed. 2013, 52, 12119−12123. (d) Sakakura, T.; Choi, J.; Yasuda, H. Chem. Rev. 2007, 107, 2365−2387. (4) (a) Cai, X.; Xie, B. Synthesis 2013, 45, 3305−3324. (b) Shao, P.; Wang, S.; Chen, C.; Xi, C. Chem. Commun. 2015, 51, 6640−6642. (c) Takimoto, M.; Hou, Z. Chem. - Eur. J. 2013, 19, 11439−11445. (d) Yu, D.; Teong, S. P.; Zhang, Y. Coord. Chem. Rev. 2015, 293−294, 279−291. (e) Yeung, C. S.; Dong, V. M. Top. Catal. 2014, 57, 1342− 1350. (5) (a) Hoberg, H.; Peres, Y.; Krüger, C.; Tsay, Y. Angew. Chem., Int. Ed. Engl. 1987, 26, 771−773. (b) Hoberg, H.; Peres, Y.; Milchereit, A.; Gross, S. J. Organomet. Chem. 1988, 345, C17−C19. (6) Williams, C. M.; Johnson, J. B.; Rovis, T. J. Am. Chem. Soc. 2008, 130, 14936−14937. (7) (a) Wang, X.; Nakajima, M.; Martin, R. J. Am. Chem. Soc. 2015, 137, 8924−8927. (b) Li, S.; Yuan, W.; Ma, S. Angew. Chem. 2011, 123, 2626−2630. (c) Liu, Y.; Cornella, J.; Martin, R. J. Am. Chem. Soc. 2014, 136, 11212−11215. (d) González-Sebastian, L.; Flores-Alamo, M.; García, J. J. Organometallics 2015, 34, 763−769. (e) Moragas, T.; Gaydou, M.; Martin, R. Angew. Chem., Int. Ed. 2016, 55, 5053−5057. (f) León, T.; Correa, A.; Martin, R. J. Am. Chem. Soc. 2013, 135, 1221−1224. (8) (a) Suh, H. W.; Guard, L. M.; Hazari, N. Chem. Sci. 2014, 5, 3859−3872. (b) Sun, J.; Bao, M.; Feng, X.; Yu, X.; Yamamoto, Y.; Almansour, A. I.; Arumugam, N.; Kumar, R. S. Tetrahedron Lett. 2015, 56, 6747−6750. (c) Hoberg, H.; Minato, M. J. Organomet. Chem. 1991, 406, C25−C28. (9) Dupuy, S.; Gasperini, D.; Nolan, S. P. ACS Catal. 2015, 5, 6918− 6921. (10) Ohmiya, H.; Tanabe, M.; Sawamura, M. Org. Lett. 2011, 13, 1086−1088. (11) (a) Hoberg, H.; Jenni, K.; Krüger, C.; Raabe, E. Angew. Chem., Int. Ed. Engl. 1986, 25, 810−811. (b) Hoberg, H.; Jenni, K.; Angermund, K.; Krüger, C. Angew. Chem., Int. Ed. Engl. 1987, 26, 153−155. (12) (a) Jin, G.; Werncke, C. G.; Escudie, Y.; Sabo-Etienne, S.; Bontemps, S. J. Am. Chem. Soc. 2015, 137, 9563−9566. (b) Huehls, C. B.; Lin, A.; Yang, J. Org. Lett. 2014, 16, 3620−3623. (c) Frank, D. J.; F

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics (33) (a) Ren, Q.; Guan, S.; Jiang, F.; Fang, J. J. Phys. Chem. A 2013, 117, 756−764. (b) Ren, Q.; Shen, X. Acta Phys. Chim. Sin. 2015, 31, 852−858.

G

DOI: 10.1021/acs.organomet.6b00681 Organometallics XXXX, XXX, XXX−XXX