A DFT Study on Palladium and Nickel-Catalyzed ... - ACS Publications

Aug 29, 2017 - Department of Applied Chemistry, Zhejiang Gongshang University, ... and Chemical Engineering, South China University of Technology, ...
0 downloads 0 Views 5MB Size
Article pubs.acs.org/Organometallics

A DFT Study on Palladium and Nickel-Catalyzed Regioselective and Stereoselective Hydrosilylation of 1,3-Disubstituted Allenes Hujun Xie,*,† Jian Kuang,† Lihong Wang,† Yang Li,† Lvtao Huang,† Ting Fan,*,‡ Qunfang Lei,§ and Wenjun Fang§ †

Department of Applied Chemistry, Zhejiang Gongshang University, Hangzhou 310018, People’s Republic of China School of Chemistry and Chemical Engineering, South China University of Technology, Guangzhou 510641, People’s Republic of China § Department of Chemistry, Zhejiang University, Hangzhou 310027, People’s Republic of China ‡

S Supporting Information *

ABSTRACT: Theoretical calculations have been carried out to illustrate the mechanisms and origins of regio/stereoselective hydrosilylation of 1,3-disubstituted allenes catalyzed by Pd and Ni catalysts. Investigations indicated that the mechanisms are divided into three important steps: coordination of silane and allene to the metal center, concerted oxidative addition, and silyl/hydride migration as well as C−H/C−Si bond reductive elimination. We found that concerted oxidative addition is the rate-limiting and regioselective step for the allene-hydrosilylation reaction. Moreover, the utilization of different palladium and nickel catalysts containing N-heterocyclic carbene (NHC) ligands can lead to switchable regioselectivity in allene hydrosilylations. The reaction with Pd−NHC catalysts favors allylsilane products, whereas the reaction with Ni−NHC catalysts affords alkenylsilane products. Intrinsic electronic effects of Pd or Ni catalysts and steric repulsions between catalysts and substrates (silyl and allene) can significantly affect the regioselectivity. These computations are in accordance with observations from experiments and can help to design new palladium and nickel catalysts for the regio/stereoselectivities of allene-hydrosilylation reactions.



INTRODUCTION Organosilicon compounds are useful building blocks in organic synthesis.1 They are widely used to generate C−C bonds in organic synthesis, with example reactions including the Hiyama− Denmark reaction,2 cycloaddition reactions, and Sakurai allylation reactions.3 The hydrosilylation of unsaturated compounds is a good way to obtain organosilicon species.4 Alkenylsilanes can be produced via alkyne hydrosilylation, and allylsilanes can be accessed via 1,3-diene hydrosilylation. For hydrosilylation reactions, it is difficult to control the regio and stereoselectivities, which often leads to regio and/or stereoisomeric mixtures.5 Significant developments have been made to produce high regio and stereoselective organosilicon products for the hydrosilylation of alkynes and dienes.6−14 Allenes seem to be a potential substrate to afford organosilicon compounds with more structural diversity and stereochemical complexity compared with alkynes and dienes. However, allene has two orthogonal and almost comparable π bonds, which make it very difficult to get high regio and stereoselectivities.15−16 Recently, Montgomery and co-workers found that palladium and nickel can catalyze hydrosilylation of 1,3-disubstituted allenes with high stereoselectivity and regioselectivity as shown in eqs (1) and (2).17 Regardless of the use of symmetrical or unsymmetrical allenes, a Pd−L1 catalyst leads to E-allylsilanes, while a © 2017 American Chemical Society

Received: June 29, 2017 Published: August 29, 2017 3371

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics Ni−L2 catalyst leads to Z-alkenylsilanes. For the hydrosilylation of unsymmetrical allenes (eq (2)), E-allylsilanes and Z-alkenylsilanes have two forms depending on whether −SiR3 or −H are added to the bulky carbon or not. To differentiate these two forms, they are labeled as E-allyl (−SiR3 adding to bulky carbon), E-allyl′ (−SiR3 adding to less bulky carbon), Z-alkenyl (−H adding to bulky carbon) and Z-alkenyl′ (−H adding to less bulky carbon). Experimentally, it was found that an Ni−L2 catalyst could lead to excellent selectivity for one of the isomers, specifically the Z-alkenyl isomer and not the Z-alkenyl′ isomer, when allene substituents are sufficiently different in size. However, a Pd−L1 catalyst leads to a mixture of E-allyl and E-allyl′. To know the difference between Pd and Ni catalysts, we theoretically studied the detailed mechanisms of 1,3-disubstituted allene-hydrosilylation reactions catalyzed by palladium and nickel catalysts. In comparison with nickel and palladium, we hope to get some clues of how to differentiate the two π bonds of allenes.

polarization functions of Pd(ζf) = 1.472, Ni(ζf) = 3.130, and Si(ζd) = 0.262 were added.23 Frequency analyses for all stationary points were carried out to confirm local minima or transition states (TS). Intrinsic reaction coordinate (IRC) calculations were also done to check a TS connecting relevant minima.24 All computations were implemented via the Gaussian09 software package.25 Solvent effects are also taken into account. We employed single-point energy calculations through the conductor-like polarizable continuum model (CPCM) with the UAHF radii.26 According to the reaction conditions, we used tetrahydrofuran as solvent. Entropic contributions are very important for reactions in which the number of reaction molecules is changed, such as 1-to-2 reactions. The corrections were performed using freevolume theory.27 For 1-to-1 or 2-to-2 reactions, corrections were unnecessary. While for 2-to-1 (or 1-to-2) reactions, it requires to correct −2.6 (or 2.6) kcal/mol at the temperature of 298.15 K. Previous calculations also confirm the validity of this correction method.28 For the reaction mechanism discussed below, we use the corrected relative Gibbs free energies to investigate the mechanisms.



COMPUTATIONAL DETAILS All species were optimized by DFT computations with the hybrid Becke3LYP (B3LYP) functional.18 Previous work has confirmed the rationality of this selected functional to study palladium- and nickel-catalyzed organic reactions.19,20 We also performed singlepoint calculations by the B3LYP-D3 method21 for the 1,3dimethylallene hydrosilylation catalyzed by the N-heterocyclic carbene complex (IPr) of palladium (Figure 1) and nickel (Figure 3) for the generation of allylsilane and alkenylsilane based on optimized structures in the gas phase. The calculation results (Figures S1,S2 in Supporting Information) showed that different theoretical methods (B3LYP and B3LYP-D3) give similar trends, which is also consistent with experimental observations.17 For the basis sets, we selected the 6-31G(d,p) basis set for C, H, O, and N atoms and the double-ζ valence basis set (LanL2DZ)22 for Si, Pd, and Ni atoms. Furthermore,



RESULTS AND DISCUSSION On the basis of experimental and theoretical investigations,16,17 a mechanism is proposed for the hydrosilylation of allenes using 1,3-dimethylallene as a model described in Scheme 1. There are three major steps: silane and allene coordination, concerted oxidative addition, and reductive elimination. The generation of allylsilanes or alkenylsilanes depends on whether −H or −SiR3 migrates to the middle carbon of allene. Symmetrical 1,3-Disubstituted Allene Hydrosilylation by Pd and Ni Catalysts. In eq (1), it shows that 1,3-dihexylallene reacts with HSiMe2Bn in the presence of Pd−L1.

Figure 1. Calculated reaction profiles (kcal/mol) of 1,3-dimethylallene-hydrosilylation reaction catalyzed by Pd−L1(IPr) catalyst to produce allylsilane (path a) and alkenylsilane (path b). 3372

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

related to H atom migration from silane to the allene internalcarbon atom. In path b, the concerted step takes place via TS3−7a to produce another η3 allyl complex 7a with a barrier of 28.8 kcal/mol. The imaginary frequency of TS3−7a is 149i cm−1, which is related to the silyl group transfer from silane to the allene internal-carbon atom. The 4.3 kcal/mol energy difference between paths a and b accounts for experimental regioselectivity (allylsilane/alkenylsilane = 98:2).17 From 5a, it undergoes reductive elimination to form E-allylsilane, which is the main stereoproduct. From 7a, it can form Z-alkenylsilane. We also consider isomerization of 5a or 7a to 5a′ or 7a′ (Scheme 2),

Scheme 1. Two Possible Catalytic Cycles for the 1,3-Dimethylallene Hydrosilylation

Scheme 2. Molecular Structures of 5a′ and 7a′ The ratio of allylsilane to alkenylsilane is over 98:2. The ratio of E-allylsilane to Z-allylsilane is over 98:2. However, when the reaction is catalyzed by Ni−L2, the selectivity is reversed. The ratio of allylsilane to alkenylsilane is below 98:2. The ratio of Z-alkenylsilane to E-alkenylsilane is over 98:2. To reduce computational cost, 1,3-dimethylallene was used to model 1,3-dihexylallene. We first considered the hydrosilylation of 1,3-dimethylallenes catalyzed by a Pd−L1 catalyst. From 1a, complex 2a can be produced through HSiBnMe2 coordination to a palladium atom, and this step releases 3.4 kcal/mol of energy (Figures 1,2). From 2a, the product for the oxidative addition (OA) of the Si−H bond cannot be located because the OA product will lead to stable complex 2a after optimization.29,30 Complex 2a is a σ-Si−H complex as indicated by the elongated Si−H bond of 1.618 Å compared to the single bond (1.48 Å) based on the covalent radius, which clearly shows a 2e···3c interaction (Pd···H−Si).31,32 Subsequently, 3a and 4a can be produced by the coordination of 1,3-dimethylallene to a Pd atom, and these two processes need energies of 9.2 and 14.5 kcal/mol, respectively. From 4a, path a is involved in the Si−H oxidative addition concerted with hydride migration to the allene central carbon, which can finally give allylsilane. From 3a, path b is related to Si−H oxidative addition concerted with silyl migration to the allene central carbon, which can finally afford alkenylsilane. Path a is energetically more favored than path b. The concerted step in path a takes place via TS4−5a to produce η1 allyl complex 5a.33 The overall activation barrier is 24.5 kcal/mol from 2a to TS4−5a. The imaginary frequency of TS4−5a is 673i cm−1, which is

which further undergoes reductive elimination to form Z-allylsilane or E-alkenylsilane, but the barrier of isomerization is very high, over 20 kcal/mol (Figure S3). On the basis of the calculations, oxidative addition is the ratecontrolling and regioselective step for the allene-hydrosilylation reaction catalyzed by a palladium catalyst. In the experiment, it was found that when the catalyst Pd−L1 was changed to be Pd−L2, where L2 is relatively sterically larger than L1, no products of the allene hydrosilylation were obtained. Figure S4 describes the calculated reaction profiles of allene hydrosilylation to afford allylsilane (path a) and alkenylsilane (path b). The mechanism of allene hydrosilylation catalyzed by a Pd−L2 catalyst for the generation of allylsilane and alkenylsilane is similar to that catalyzed by a Pd−L1 catalyst. Oxidative addition is still the rate-limiting step. The activation barrier is 38.0 kcal/mol for path a and 40.6 kcal/mol for path b, which is kinetically unfavorable based on the experimental conditions of room temperature.17 The origin is due to the steric interactions between the bulky ligand with allene and silyl groups in TS3−4b and TS3−6b (Figure S5). The symmetrical 1,3-disubstituted allene hydrosilylation catalyzed by Ni catalysts are considered. The calculated reaction

Figure 2. Optimized structures (Å) for selected compounds related to allene-hydrosilylation reaction catalyzed by Pd−L1(IPr) catalyst. 3373

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 3. Calculated reaction profiles (kcal/mol) of 1,3-dimethylallene hydrosilylation catalyzed by Ni−L1(IPr) catalyst to give allylsilane (path a) and alkenylsilane (path b).

Figure 4. Calculated reaction profiles (kcal/mol) of 1,3-dimethylallene hydrosilylation catalyzed by Ni−L2(IPr*OMe) catalyst to give allylsilane (path a) and alkenylsilane (path b).

from the Ni center. We think the reason for this is ascribed to the steric repulsions caused by the ligand, silane, and allene. The barrier of path b (13.3 kcal/mol for TS3−6c) is lower by 1.7 kcal/mol than path a (15.0 kcal/mol for TS3−4c), agreeing with the experimental selectivity (alkenylsilane/allylsilane = 60:40).17 Our calculations also indicated that path a is more unfavorable than path b both thermodynamically and kinetically. In addition, 6c undergoes C−H reductive elimination to give Z-alkenylsilanes and 4c undergoes C−Si reductive elimination via E-allylsilane, which agrees with experimental observations that

profiles of Ni-catalyzed hydrosilylation of allene are depicted in Figure 3 (small NHC ligand L1) and 4 (bulky NHC ligand L2), and important geometric structures are displayed in Figure 5. We initially studied an Ni−L1-catalyzed allene-hydrosilylation reaction. Figure 3 describes the calculated reaction profiles of the allene-hydrosilylation reaction to generate allylsilane (path a) and alkenylsilane (path b). It has been shown that the hydrosilylation-reaction mechanisms are similar to the palladium-catalyzed ones. In calculations, we have tried our best to find the isomer of 3c; however, the HSiBnMe2 escaped 3374

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 5. Optimized structures (Å) for selected compounds related to allene hydrosilylation catalyzed by different Ni-NHC catalysts.

both Z-alkenylsilane and E-allylsilane are the major stereoselective products.17 We then study Ni-catalyzed allene-hydrosilylation reaction with bulky ligand L2. Figure 4 described the calculated reaction profiles of the allene-hydrosilylation reaction to generate allylsilane (path a) and alkenylsilane (path b). The important geometric parameters are presented in Figure 5. From computations, it shows that the allene-hydrosilylation-reaction mechanism is similar to when using the small nickel catalyst (Ni−L1). The present calculations suggest that the barrier of path a (26.8 kcal/mol for TS3−4d) is greater than that of path b (23.9 kcal/mol for TS3−6d) by 2.9 kcal/mol, which is in accordance with the experiments that the main product is alkenylsilane for the Ni−L2 catalyst (allylsilane/alkenylsilane = 2:98).17 Furthermore, from 6d, it can generate Z-alkenylsilane, which is consistent with the experiments.17 From the results above, the rate-controlling step for allene hydrosilylation catalyzed by Pd and Ni catalysts is related to the oxidative addition of hydrosilanes. The activation barriers of path a for 1a (Pd−L1), 1b (Pd−L2), 1c (Ni−L1), and 1d (Ni−L2) are 24.5, 38.0, 15.0, and 26.8 kcal/mol, respectively. It requires the activation barriers of path b for 1a, 1b, 1c, and 1d, which are 28.8, 40.6, 13.3, and 23.9 kcal/mol, respectively. On the basis of the activation barriers, it shows that the barriers for Ni-catalyzed pathways are lower than the respective Pd-catalyzed pathways.

Furthermore, Ni and Pd complexes go through different pathways, showing different regioselectivity. Take catalysts 1a and 1c as examples. NBO analysis34 (Table 1) involving the Table 1. NBO Charges for Hydrosilane and Allene Moieties of All Compounds Involved in 2a → 5a/7a and 2c → 4c/6c moiety

2a

5a

7a

2c

4c

6c

hydrosilane allene M−L1

0.02 0.00 −0.02

0.20 −0.48 0.28

0.36 −0.70 0.34

−0.08 0.00 0.08

0.15 −0.46 0.31

0.34 −0.70 0.36

oxidative-addition processes (2a → 5a/7a, 2c → 4c/6c) suggests that electron transfer takes place from transition-metal hydrosilane complexes to allene. For the hydrosilylation reaction catalyzed by the Pd−L1(IPr) catalyst, the charge in the allene group changes from 0.00 in 2a to −0.48 in 5a and −0.70 in 7a. For the hydrosilylation reaction catalyzed by the Ni−L1(IPr) catalyst, the charge in the allene group changes from 0.00 in 2c to −0.46 in 4c and −0.70 in 6c. We then checked the frontier molecular orbitals located on M−H and M−Si of complexes 2a and 2c (Figure 6). HOMO−4 is located on Pd−H and HOMO−5 is on Pd−Si for complex 2a. However, for complex 2c, HOMO−4 is located on Ni−Si and Ni−H. In addition, the Ni complex is more active than Pd complex because the HOMO−4 orbital of 3375

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Scheme 3. Pd- and Ni-Catalyzed Hydrosilylation of an Unsymmetric Allene

Ni with bulky ligand L2 is 98:2 (Scheme 3). On the basis of the computational and experimental results discussed above, Pd carbene catalysts facilitate the generation of allylsilane, while Ni carbene catalysts favor the generation of alkenylsilane. Herein we only consider the regioselectivity for the allene-hydrosilylation reaction. 1-Cyclohexyl-3-methyl-allene is used in the calculation to reduce computational cost. As depicted in Figures 7,8, the barriers for the generation of allylsilanes are energetically more favored than alkenylsilanes. The computations also demonstrated that the catalytic mechanism for the generation of allylsilanes is similar to that palladium-catalyzed symmetrical allene hydrosilylation. The rate-controlling step in path a′ is related to the oxidative-addition step involving the Si−H bond concerted with hydride migaration with a barrier (TS2−4e′) of 27.2 kcal/mol, which finally leads to E-allyl′ (5e′). The rate-controlling step (path a) is related to the oxidative-addition step involving the Si−H bond concerted with hydride migration, and it overcomes a barrier (TS3−4e) of 25.9 kcal/mol, which finally leads to E-allyl (5e). As 4e and 4e′ are difficult to convert to each other based on our studies on 7a isomerization (Figure S3), regioselectivity is controlled by oxidative addition. The calculated barrier of path a′ (27.2 kcal/mol for TS2−4e′) is higher by 1.3 kcal/mol than that of path a (25.9 kcal/mol TS3−4e), agreeing with the experimentally determined regioselectivity (E-allyl/E-allyl′ = 60:40).17 Path a′

Figure 6. Frontier molecular orbitals located on M−H and M−Si for complexes 2a and 2c. (a) HOMO−4 for complex 2a. (b) HOMO−5 for complex 2a. (c) HOMO−4 for complex 2c.

2c is higher in energy than that of 2a, and the energy of LUMO of the allene is 0.89 eV. The energy difference of paths a and b is −4.3 kcal/mol for 1a (Pd−L1) and −2.6 kcal/mol for 1b (Pd−L2). The energy difference changes to 1.7 kcal/mol for 1c (Ni−L1) and 2.3 kcal/mol for 1d (Ni−L2). This shows that bulky NHC ligands promote path b, -SiR3 migration. To sum up, the regioselectivity is caused by electronic effects (HOMO−4 orbital) and ligand steric effects. Unsymmetrical 1,3-Disubstituted Allene Hydrosilylation by Pd and Ni Catalysts. For unymmetrical 1,3-disubstituted allene hydrosilylation, it was found in experiments that the regioselectivity of the allene-hydrosilylation reaction catalyzed by the Pd−L1 catalyst is 60:40, and the regioselectivity catalyzed by

Figure 7. Calculated reaction profiles (kcal/mol) of allene hydrosilylation catalyzed by Pd−L1(IPr) catalyst to generate E-allylsilanes (path a) and E-allylsilanes′ (path a′). 3376

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 8. Calculated reaction profiles (kcal/mol) of allene hydrosilylation catalyzed by Pd−L1(IPr) catalyst to generate Z-alkenylsilanes (path b) and Z-alkenylsilanes′ (path b′).



CONCLUSIONS We study the hydrosilylation of 1,3-disubstituted allene catalyzed by different palladium and nickel catalysts with the aid of theoretical calculations to illustrate the mechanisms and origins of how regio and stereoselectivities yield either allylsilanes or alkenylsilanes. The catalytic cycle is divided into three important steps. The first step is that silane and allene are coordinated to a palladium or nickel atom. The second step is concerted oxidative addition of the Si−H bond with silyl or hydride migration. The final step is reductive elimination. The computations show that oxidative addition is the rate-limiting and regioselective step. The regioselectivity is controlled by the concerted oxidativeaddition step, governed by transition metals as well as the ligands used. The utilization of nickel and palladium catalysts in the allene-hydrosilylation reactions leads to regioselectivity reversals. Allylsilane is the main product via palladium catalyst, and alkenylsilane is the main product via nickel catalyst. The regioselectivity is caused by the electronic effects (HOMO−4 orbital) and ligand steric effects. For the unsymmetrical allenes, the alterations of the bulky NHC ligand makes catalysts differentiate these two almost similar π bonds. Pd catalysts are less active than the corresponding Ni ones. In addition, catalysts with bulky NHC ligands are less active than corresponding ones with less bulky NHC ligands. Thus, the unsymmetrical allene hydrosilylations catalyzed by Pd with bulky NHC ligands such as L2 is proposed to take place with high regioselectivity at high reaction temperatures based on our calculations. The calculations indicated that the high selectivity of hydrosilylation of allenes can be obtained via selecting suitable catalysts, which includes different transition metals and ligands. We hope that our calculations can help understand the mechanism of the allene hydrosilylation and help design efficient

has a relatively high barrier, which originated from the steric repulsions between the bulky ligand and the Cy group of allene (Figure 9). For the unsymmetrical 1,3-disubstituted allene-hydrosilylation reaction catalyzed by a bulky Ni−L2(IPr*OMe) catalyst, Figure 10 describes the calculated reaction profiles to afford alkenylsilanes (path b for Z-alkenyl, path b′ for Z-alkenyl′), and Figure 11 displays the calculated reaction profiles to give allylsilanes. The optimized geometric parameters of important compounds are shown in Figure 12. The computations indicated that the barriers for the generation of alkenylsilanes are energetically more favored than allylsilanes. As 6f and 6f′ are difficult to convert to each other based on our studies on 7a isomerization shown in Figure S3, regioselectivity is controlled by oxidative addition. The reaction barrier of path b′ (24.4 kcal/mol for TS3−6f′) is higher than that of path b by 4.4 kcal/mol (20.0 kcal/mol for TS3−6f), which agrees with experiments regioselectivity (Z-alkenyl/Z-alkenyl′ = 98:2).17 The high regioselectivity of hydrosilylation of unsymmetrical allenes catalyzed by Ni−L2 is due to the bulkiness of L2 ligands. If Pd−L2 was applied to catalyze hydrosilylation of unsymmetrical allenes, the regioselectivity should be also high. We then investigate hydrosilylations of unsymmetrical allenes catalyzed by bulky Pd−L2 catalysts to give allylsilane and alkenylsilane. The rate-controlling step is involved in oxidative addition of the Si−H bond concerted with silyl/hydride insertion. The computations (Figure 13) showed the barriers for the generation of allylsilanes are kinetically more favorable than alkenylsilanes. In addition, the reaction barrier of path a′ is higher by 2.7 kcal/mol than that of path a (36.5 kcal/mol versus 33.8 kcal/mol), which means that the regioselectivity of the allene-hydrosilylation reaction should be high, as high as 90:10. However, the activation barriers are significantly higher at room temperature.17 3377

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 9. Important geometric parameters (Å) of compounds related to allene hydrosilylation catalyzed by Pd−L1(IPr) catalyst.

Figure 10. Calculated reaction profiles (kcal/mol) of allene hydrosilylation catalyzed by an Ni−L2(IPr*OMe) catalyst to generate Z-alkenylsilanes (path b) and Z-alkenylsilanes′ (path b′). 3378

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 11. Calculated reaction profiles (kcal/mol) of allene hydrosilylation catalyzed by Ni−L2(IPr*OMe) catalyst to generate E-allylsilanes (path a) and E-allylsilanes′(path a′).

Figure 12. Important geometric parameters (Å) of compounds related to allene hydrosilylation catalyzed by Ni−L2(IPr*OMe) catalyst. 3379

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

Figure 13. Key transition states (kcal/mol) calculated for oxidative-addition step involving Si−H bond, concerted with hydride/silyl insertion step catalyzed by Pd−L2(IPr*OMe) catalyst. Soc. 2011, 133, 20712−20715. (g) Chaulagain, M. R.; Mahandru, G. M.; Montgomery, J. Tetrahedron 2006, 62, 7560−7566. (5) (a) De Bo, G.; Berthon-Gelloz, G.; Tinant, B.; Markó, I. E. Organometallics 2006, 25, 1881−1890. (b) Denmark, S. E.; Wang, Z. Org. Lett. 2001, 3, 1073−1076. (c) Lappert, M. F.; Nile, T. A.; Takahashi, S. J. Organomet. Chem. 1974, 72, 425−439. (d) Denmark, S. E.; Pan, W. Org. Lett. 2003, 5, 1119−1122. (e) Ojima, I.; Kumagai, M. J. Organomet. Chem. 1978, 157, 359−379. (f) Onozawa, S.; Sakakura, T.; Tanaka, M. Tetrahedron Lett. 1994, 35, 8177−8180. (g) Tsuji, J.; Hara, M.; Ohno, K. Tetrahedron 1974, 30, 2143−2146. (h) Radkowski, K.; Sundararaju, B.; Fürstner, A. Angew. Chem., Int. Ed. 2013, 52, 355−360. (i) Sundararaju, B.; Fürstner, A. Angew. Chem., Int. Ed. 2013, 52, 14050− 14052. (j) Rummelt, S. M.; Fürstner, A. Angew. Chem., Int. Ed. 2014, 53, 3626−3630. (6) (a) Glaser, P. B.; Tilley, T. D. J. Am. Chem. Soc. 2003, 125, 13640− 13641. (b) Beddie, C.; Hall, M. B. J. Am. Chem. Soc. 2004, 126, 13564− 13565. (c) Brunner, H. Angew. Chem., Int. Ed. 2004, 43, 2749−2750. (7) Sommer, L. H.; Pietrusza, E. W.; Whitmore, F. C. J. Am. Chem. Soc. 1947, 69, 188−189. (8) Adamski, A.; Kubicki, M.; Pawluc, P.; Grabarkiewicz, T.; Patroniak, V. Catal. Commun. 2013, 42, 79−83. (9) Zhou, H.; Moberg, C. Org. Lett. 2013, 15, 1444−1447. (10) (a) Trost, B. M.; Ball, Z. T. J. Am. Chem. Soc. 2001, 123, 12726− 12727. (b) Trost, B. M.; Ball, Z. T. J. Am. Chem. Soc. 2005, 127, 17644− 17655. (11) Wu, J. Y.; Stanzl, B. N.; Ritter, T. J. Am. Chem. Soc. 2010, 132, 13214−13216. (12) (a) Yang, Y. F.; Chung, L. W.; Zhang, X. H.; Houk, K. N.; Wu, Y. D. J. Org. Chem. 2014, 79, 8856−8864. (b) Ding, S. T.; Song, L. J.; Chung, L. W.; Zhang, X. H.; Sun, J. W.; Wu, Y. D. J. Am. Chem. Soc. 2013, 135, 13835−13842. (c) Song, L. J.; Ding, S. T.; Wang, Y.; Zhang, X. H.; Wu, Y. D.; Sun, J. W. J. Org. Chem. 2016, 81, 6157−6164. (13) Sudo, T.; Asao, N.; Gevorgyan, V.; Yamamoto, Y. J. Org. Chem. 1999, 64, 2494−2499. (14) Miller, Z. D.; Li, W.; Belderrain, T. R.; Montgomery, J. J. Am. Chem. Soc. 2013, 135, 15282−15285. (15) (a) Burks, H. E.; Liu, S.; Morken, J. P. J. Am. Chem. Soc. 2007, 129, 8766−8773. (b) Pelz, N. F.; Morken, J. P. Org. Lett. 2006, 8, 4557− 4559. (c) Pelz, N. F.; Woodward, A. R.; Burks, H. E.; Sieber, J. D.; Morken, J. P. J. Am. Chem. Soc. 2004, 126, 16328−16329. (d) Yang, F. Y.; Cheng, C. H. J. Am. Chem. Soc. 2001, 123, 761−762. (e) Trost, B. M.; Frederiksen, M. U.; Papillon, J. P. N.; Harrington, P. E.; Shin, S.; Shireman, B. T. J. Am. Chem. Soc. 2005, 127, 3666−3667. (f) Denmark, S. E.; Fujimori, S. J. Am. Chem. Soc. 2005, 127, 8971−8973. (g) O’Neil, G. W.; Phillips, A. J. Tetrahedron Lett. 2004, 45, 4253−4256. (h) Heffron, T. P.; Simpson, G. L.; Merino, E.; Jamison, T. F. J. Org. Chem. 2010, 75, 2681−2701. (16) Xie, H. J.; Zhao, L. J.; Yang, L.; Lei, Q. F.; Fang, W. J.; Xiong, C. H. J. Org. Chem. 2014, 79, 4517−4527. (17) Miller, Z. D.; Dorel, R.; Montgomery, J. Angew. Chem., Int. Ed. 2015, 54, 9088−9091. (18) (a) Becke, A. D. J. Chem. Phys. 1993, 98, 5648−5652. (b) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (c) Becke, A. D. J. Chem. Phys. 1993, 98, 1372−1377. (d) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098−3100. (19) (a) Xie, H. J.; Zhang, H.; Lin, Z. Y. Organometallics 2013, 32, 2336−2343. (b) Xie, H. J.; Lin, F. R.; Lei, Q. F.; Fang, W. J. Organometallics 2013, 32, 6957−6968. (c) Xie, H. J.; Zhang, H.; Lin, Z.

catalysts to control the regio and stereoselectivities of hydrosilylation reactions.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.7b00495. Calculated energy profiles for the 1,3-dimethylallene hydrosilylation (PDF) Cartesian coordinates for the geometries of different compounds (XYZ)



AUTHOR INFORMATION

Corresponding Authors

*H.X.: E-mail: [email protected] *T.F.: E-mail: [email protected]; Fax: +86-57128008900 ORCID

Hujun Xie: 0000-0002-0035-2634 Wenjun Fang: 0000-0002-5610-1623 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Natural Science Foundation of China (21203166 and 21473157), the Natural Science Foundation of Zhejiang Province (LY16B030001), the Center for Collaborative Innovation in Food Safety and Nutrition (2017SICR112), the Zhejiang Provincial Top Key Discipline of Food Science and Biotechnology (JYTsp2014111), and College Students’ Science and Technology Innovation Activities in Zhejiang Province (2016R408026).



REFERENCES

(1) (a) Masse, C. E.; Panek, J. S. Chem. Rev. 1995, 95, 1293−1316. (b) Fleming, I.; Barbero, A.; Walter, D. Chem. Rev. 1997, 97, 2063− 2192. (2) (a) Nakao, Y.; Hiyama, T. Chem. Soc. Rev. 2011, 40, 4893−4901. (b) Chang, W.-T. T.; Smith, R. C.; Regens, C. S.; Bailey, A. D.; Werner, N. S.; Denmark, S. E. Org. React. 2011, 75, 213−745. (3) (a) Angle, S. R.; El-Said, N. A. J. Am. Chem. Soc. 2002, 124, 3608− 3613. (b) Judd, W. R.; Ban, S.; Aube, J. J. Am. Chem. Soc. 2006, 128, 13736−13741. (c) Hosomi, A.; Endo, M.; Sakurai, H. Chem. Lett. 1976, 5, 941−942. (4) (a) Lewis, L. N.; Stein, J.; Gao, Y.; Colborn, R. E.; Hutchins, G. Platinum Metals Rev. 1997, 41, 66−75. (b) Stein, J.; Lewis, L. N.; Gao, Y.; Scott, R. A. J. Am. Chem. Soc. 1999, 121, 3693−3703. (c) Marciniec, B.; Maciejewski, H.; Pietraszuk, C.; Pawluc, P. In Hydrosilylation: A Comprehensive Review on Recent Advances, 1st ed.; Marciniec, B., Ed.; Advances in Silicon Science, Vol. 1, Springer: Berlin, Germany, 2009. (d) Clarson, S. J. Book review. Silicon 2009, 1, 57. (e) Rooke, D. A.; Ferreira, E. M. Angew. Chem., Int. Ed. 2012, 51, 3225−3230. (f) Kawasaki, Y.; Ishikawa, Y.; Igawa, K.; Tomooka, K. J. Am. Chem. 3380

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381

Article

Organometallics

(33) (a) Obora, Y.; Tsuji, Y.; Kawamura, T. J. Am. Chem. Soc. 1993, 115, 10414−10415. (b) Obora, Y.; Tsuji, Y.; Kawamura, T. J. Am. Chem. Soc. 1995, 117, 9814−9859. (c) Tsuji, Y.; Funato, M.; Ozawa, M.; Ogiyama, H.; Kajita, S.; Kawamura, T. J. Org. Chem. 1996, 61, 5779− 5787. (d) Sakaki, S.; Satoh, H.; Shono, H.; Ujino, Y. Organometallics 1996, 15, 1713−1720. (e) Biswas, B.; Sugimoto, M.; Sakaki, S. Organometallics 1999, 18, 4015−4026. (34) (a) Reed, A. E.; Weinstock, R. B.; Weinhold, F. J. Chem. Phys. 1985, 83, 735−746. (b) Reed, A. E.; Weinhold, F. J. Chem. Phys. 1985, 83, 1736−1740. (c) Reed, A. E.; Curtiss, L. A.; Weinhold, F. Chem. Rev. 1988, 88, 899−926.

Y. New J. Chem. 2013, 37, 2856−2861. (d) Xie, H. J.; Lin, F. R.; Yang, L.; Chen, X. S.; Ye, X. C.; Tian, X.; Lei, Q. F.; Fang, W. J. J. Organomet. Chem. 2013, 745−746, 417−422. (e) Xie, H. J.; Lin, F. R.; Lei, Q. F.; Fang, W. J. J. Phys. Org. Chem. 2013, 26, 933−938. (f) Yang, L.; Ren, G. R.; Ye, X. C.; Que, X. Y.; Lei, Q. F.; Fang, W. J.; Xie, H. J. J. Phys. Org. Chem. 2014, 27, 237−244. (g) Xie, H. J.; Fan, T.; Lei, Q. F.; Fang, W. J. Sci. China: Chem. 2016, 59, 1432−1447. (20) (a) Tang, S. Y.; Guo, Q. X.; Fu, Y. Chem. - Eur. J. 2011, 17, 13866− 13876. (b) Perez-Rodriguez, M.; Braga, A. A. C.; de Lera, A. R.; Maseras, F.; Alvarez, R.; Espinet, P. Organometallics 2010, 29, 4983−4991. (c) Surawatanawong, P.; Hall, M. Organometallics 2008, 27, 6222−6232. (d) Lam, K. C.; Marder, T. B.; Lin, Z. Y. Organometallics 2010, 29, 1849−1857. (e) Xue, L. Q.; Lin, Z. Y. Chem. Soc. Rev. 2010, 39, 1692− 1705. (f) Yu, H. Z.; Fu, Y.; Guo, Q. X.; Lin, Z. Y. Organometallics 2009, 28, 4507−4512. (g) Zheng, W. X.; Ariafard, A.; Lin, Z. Y. Organometallics 2008, 27, 246−253. (h) Lam, K. C.; Marder, T. B.; Lin, Z. Y. Organometallics 2007, 26, 758−760. (i) Ariafard, A.; Lin, Z. Y. J. Am. Chem. Soc. 2006, 128, 13010−13016. (21) Grimme, S. WIREs Comput. Mol. Sci. 2011, 1, 211−228. (22) (a) Check, C. E.; Faust, T. O.; Bailey, J. M.; Wright, B. J.; Gilbert, T. M.; Sunderlin, L. S. J. Phys. Chem. A 2001, 105, 8111−8116. (b) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 299−310. (23) (a) Ehlers, A. W.; Bohme, M.; Dapprich, S.; Gobbi, A.; Hollwarth, A.; Jonas, V.; Kohler, K. F.; Stegmann, R.; Veldkamp, A.; Frenking, G. Chem. Phys. Lett. 1993, 208, 111−114. (b) Huzinaga, S. Gaussian Basis Sets for Molecular Calculations; Elsevier: Amsterdam: The Netherlands, 1984. (24) (a) Fukui, K. J. Phys. Chem. 1970, 74, 4161−4163. (b) Fukui, K. Acc. Chem. Res. 1981, 14, 363−368. (25) 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, O.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, revision A.01; Gaussian, Inc.: Wallingford, CT, 2009. (26) (a) Barone, V.; Cossi, M. J. Phys. Chem. A 1998, 102, 1995−2001. (b) Cossi, M.; Rega, N.; Scalmani, G.; Barone, V. J. Comput. Chem. 2003, 24, 669−681. (27) Benson, S. W. The Foundations of Chemical Kinetics; Krieger Pub Co: Malabar, FL, 1982. (28) (a) Okuno, Y. Chem. - Eur. J. 1997, 3, 212−218. (b) Ardura, D.; López, R.; Sordo, T. L. J. Phys. Chem. B 2005, 109, 23618−23623. (c) Schoenebeck, F.; Houk, K. N. J. Am. Chem. Soc. 2010, 132, 2496− 2497. (d) Liu, Q.; Lan, Y.; Liu, J.; Li, G.; Wu, Y. D.; Lei, A. J. Am. Chem. Soc. 2009, 131, 10201−10210. (e) Wang, M. Y.; Fan, T.; Lin, Z. Y. Organometallics 2012, 31, 560−569. (f) Wang, M. Y.; Fan, T.; Lin, Z. Y. Polyhedron 2012, 32, 35−40. (g) Liu, B. W.; Gao, M.; Dang, L.; Zhao, H. T.; Marder, T. B.; Lin, Z. Y. Organometallics 2012, 31, 3410−3425. (29) Iglesias, M.; Sanz Miguel, P. J.; Polo, V.; Fernandez-Alvarez, F. J.; Perez-Torrente, J. J.; Oro, L. A. Chem. - Eur. J. 2013, 19, 17559−17566. (30) (a) Choi, S. H.; Feng, J. W.; Lin, Z. Y. Organometallics 2000, 19, 2051−2054. (b) Corey, J. Y.; Braddock-Wilking, J. Chem. Rev. 1999, 99, 175−292. (c) Speier, J. L. Adv. Organomet. Chem. 1979, 17, 407−447. (d) Parshall, G. W. Homogeneous Catalysis: the Application and Chemistry of Catalysis by Soluble Transition Metal Complexes; Wiley: New York, 1992; p 39. (31) Pyykkö, P.; Atsumi, M. Chem. - Eur. J. 2009, 15, 12770−12779. (32) Popp, B. V.; Stahl, S. S. J. Am. Chem. Soc. 2007, 129, 4410−4422. 3381

DOI: 10.1021/acs.organomet.7b00495 Organometallics 2017, 36, 3371−3381