Catalytic Preparation of Cyclic Carbonates from CO2 and Epoxides by

Jan 23, 2018 - Extensive DFT calculations have been performed to elucidate the plausible mechanisms and effects of the cocatalyst and meso-substitutio...
3 downloads 7 Views 2MB Size
Article Cite This: Organometallics XXXX, XXX, XXX−XXX

Catalytic Preparation of Cyclic Carbonates from CO2 and Epoxides by Metal−Porphyrin and −Corrole Complexes: Insight into Effects of Cocatalyst and meso-Substitution Ping Li†,‡ and Zexing Cao*,†,‡ †

State Key Laboratory of Physical Chemistry of Solid Surfaces and ‡Fujian Provincial Key Laboratory of Theoretical and Computational Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China S Supporting Information *

ABSTRACT: Extensive DFT calculations have been performed to elucidate the plausible mechanisms and effects of the cocatalyst and meso-substitution of the macro ring on the coupling reaction of CO2 with epoxide catalyzed by metal prophyrin and corrole complexes. Here the catalytic generation of cyclic carbonate from CO2 and epoxide by the Al−, Mg−, and Zn−porphyrin catalysts is predicted to follow a similar multistep mechanism, in which the ring-opening and -closure steps experience relatively high free energy barriers, while the CO2 insertion is facile. The computational results lend support to the remarkable catalytic role of a quaternary ammonium salt cocatalyst in the order PPNCl > PPNBr > PPNI. Introduction of the aryl and chlorinated aryl substituents on the porphyrin ligand reduces the electrophilicity of an aluminum center to some extent but remarkably enhances binding interactions among catalyst, cocatalyst, and reactant, which may facilitate the initial reaction. The ring-closure step is less influenced by the meso-substitution. The corrole-based complexes have similar catalytic behaviors, and they are also promising catalysts for the cycloaddition of CO2 to epoxides. centers.34−38 Among the selected main-group metal porphyrins prepared by Jing and co-workers, the metal dependence of catalytic activity is AlIII > MgII > SnIV > SnII, and the metalloporphyrins containing magnesium(II) and aluminum(III) can trigger the coupling of CO2 and epoxides into the cyclic carbonates with high yield at room temperature.38 More recently, Qin et al. successfully designed and prepared an aluminum porphyrin complex with a quaternary ammonium salt cocatalyst, which shows high activity and selectivity for the cycloaddition of CO2 and epoxides to cyclic carbonates,39 and they found that the cocatalyst PPNCl containing Cl− enhances the catalytic activity more remarkably than PPNBr and PPNI, containing Br− and I−, respectively. Such cocatalyst effect on the catalytic activity was ascribed to the nucleophilicity decrease of halide ions, going from Cl− to Br− and I−. On the contrary, for the cycloaddition of CO2 and epoxides to the cyclic carbonates catalyzed by alkali halides with the hydroxyl substances as the cocatalyst, KI and NaI show the highest activity.40,41 Subsequent DFT calculations indicate that the binary synergistic action of K+ and I− ions results in a stepwise mechanism for the coupling reaction and reduces the barriers remarkably, compared to the one-step cycloaddition.41 Similar reaction mechanisms, including the epoxy ring-opening, the CO2 electrophilic attack, and the intramolecular cyclization,

1. INTRODUCTION Carbon dioxide is one of the main greenhouse gases, but it is also widely accepted to be an attractive and abundant C1 source in organic synthesis. There is long-standing interest in the catalytic conversion of CO2 into useful organic compounds,1−6 both academically and industrially, and various catalysts and chemical processes have been developed. In CO2 catalysis, the preparation of cyclic carbonates, which serve as the building blocks for polycarbonates associated with polar aprotic and electrolyte solvents as well as important intermediates in organic synthesis, has received considerable attention,7,8 where the coupling of CO2 and epoxides to produce the cyclic carbonate is among the most effective routes for the CO2 fixation.9−11 In past decades, a variety of catalysts, such as alkali metal salts,12,13 metal oxides,14,15 zeolites,16,17 smectites,18,19 quaternary onium salts,20,21 transition-metal complexes,22,23 functional organic polymer,24,25 and ionic liquids,26−33 among others, have been developed for the selective coupling of CO2 and epoxides. However, some of the reported catalytic systems suffer from relatively low reactivity and turnover frequencies (TOFs), high catalyst loadings, harsh reaction conditions, and water and/or air sensitivity. Alternatively, the metalloporphyrin systems have been developed and widely utilized for the capture and conversion of CO2, owing to their unique structural properties and abilities to accommodate various metal ions as the catalytic © XXXX American Chemical Society

Received: November 15, 2017

A

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 1. Predicted relative free energy profile, electronic energies (in parentheses) (a), and corresponding optimized structures (bond lengths in Å) of the reactant, the transition state, and the product involved in the one-step mechanism (b) for the cycloaddition reaction CO2 with an epoxide catalyzed by the Al−porphyrin catalyst.

were proposed in the fixation of CO2 and epoxides to generate cyclic carbonates catalyzed by the quaternary ammonium salt42 and the LiBr salt,43 both experimentally and theoretically. For the formation of cyclic carbonates from CO2 and epoxides catalyzed by metalloporphyrins, various density functional methodologies44−46 have been used to investigate plausible catalytic mechanisms for the magnesium porphyrin system. In the catalytic cycle, the ring-opening and -closure steps are predicted to experience relatively high barriers,34,35 where the cooperative action of the halide ion as the nucleophilic group and the metal center to bind the epoxide play an important catalytic role. Despite these important contributions, especially the considerable progress achieved experimentally, further theoretical efforts are required to understand the mechanistic details for the CO2 fixation by various metalloporphyrins as well as the catalytic role of halide ions Cl−, Br−, and I−. Herein, extensive density functional calculations on the coupling of CO2 and epoxides to the cyclic carbonates catalyzed by metalloporphyrins and metallocorroles containing aluminum, zinc, and magnesium have been performed, and detailed mechanisms and activity dependence on the metal centers and halide ions as well as the meso-substitution effect have been discussed based on the calculations.

improve the estimation of relative energies, the single-point energy calculations were performed at the M06/6-311+G(2d,2p) level, based on the optimized geometries by M06/6-31G(d). For selected large computational models, the B3LYP functional45,54 was used in DFT calculations. In consideration of the overestimation of the entropic contribution to the Gibbs free energy from the gas-phase calculations for the reaction step with different numbers of reactant and product molecules in the condensed phase, a correction of −2.6 (or 2.6) kcal/mol (T = 298.15K) was applied to calibrate the relative free energies for the 2:1 (or 1:2) transformation according to the free volume theory.55

3. RESULTS AND DISCUSSION 3.1. Catalytic Coupling of CO2 with Epoxide by the Al−Porphyrin Complex. Recent experimental studies reveal that an aluminum porphyrin complex has high activity and selectivity for the cyclic carbonate synthesis in the presence of a quaternary ammonium salt cocatalyst.39 In order to clarify the mechanistic details and the role of cocatalyst, the direct catalytic cycloaddition without the cocatalyst was investigated at first. Figure 1 shows the one-step concerted mechanism for the cycloaddition of CO2 to an epoxide, and the predicted free energy barrier is 58.2 kcal/mol. Clearly, this reaction cannot occur under mild condition without the aid of cocatalyst, as observed experimentally.39 Since the concerted process in the absence of cocatalyst experiences a substantially high free energy barrier, a stepwise mechanism for the cycloaddition of CO2 to an epoxide catalyzed by the Al−porphyrin complex with a quaternary ammonium salt cocatalyst was investigated. In order to reduce the computational cost, only the naked halide ion without the ammonium cation, directly involved in the coupling reaction, was considered in all calculations. On the basis of previous studies,34,35 here a three-step mechanism, including the ringopening, CO2-insertion, and ring-closure steps, was explored. Predicted relative energies in the catalytic cycloaddition in the presence of cocatalyst ligand X (X = Cl−, Br−, and I−) are

2. COMPUTATIONAL DETAILS The geometries of all species involved in the catalytic coupling of CO2 and epoxides into the cyclic carbonates are optimized by the M06 functional47 incorporated in Gaussian 09 program.48 In calculation, the double-ζ valence basis set of LanL2DZ49,50 was used for the Cl and aluminum atoms, where a polarization function (ζd = 0.64)51 was augmented for Cl. For all other atoms, the basis set of 6-31G(d)52 was used. The optimized structures were assessed by the vibrational analysis at the same level of theory. The stable structures of reactants, intermediates, and products are identified by all positive frequencies, while the transition states are characterized by only one imaginary frequency. The minimum energy path (MEP) is constructed by the intrinsic reaction coordinate (IRC) method53 to verify if the transition state correlates two desired minima on the potential energy surface. To B

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 2. Predicted relative free energies and electronic energies (in parentheses) for the catalytic generation of the cyclic carbonate from CO2 and an epoxide catalyzed by the Al−porphyrin catalyst in the presence of cocatalyst ligand X−.

Figure 3. Predicted relative free energies and electronic energies (in parentheses) for the ring-opening and -closure steps in the catalytic coupling reaction of CO2 and an epoxide catalyzed by the Al−porphyrin catalyst in the presence of cocatalyst PPNX.

shown in Figure 2. Selected optimized structures of species involved in the coupling reaction are collected in Figure S1. As Figure 2 shows, the CO2 insertion is relatively facile, and a similar phenomenon was found in previous experimental and theoretical studies.34,35 In the coupling reaction of CO2 and PO, the ring-opening and -closure steps play a crucial role. In the ring-opening step, the cocatalysts Cl− and Br− exhibit higher activity than that of I−, and the ring closure is the rate-

determining step for all cocatalyst ligands. The predicted relative energies in Figures 1 and 2 show that the cocatalyst can remarkably enhance the catalytic activity of the Al−porphyrin compound, as observed experimentally.39 Considering that the influence of the sterically hindered countercation of cocatalyst on key reaction steps in the coupling reaction, we reinvestigated the ring-opening and -closure steps by using the complete quaternary ammonium salt cocatalyst PPNX (PPN = bisC

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 4. Predicted relative free energy profile and electronic energies (in parentheses) for the cycloaddition reaction catalyzed by the Mg− porphyrin catalyst in the presence of the cocatalyst ligand Cl−, in which the zero-point energy correction was considered.

Figure 5. Predicted relative free energy profile and electronic energies (in parentheses) for the cycloaddition reaction catalyzed by the Zn−porphyrin catalyst in the presence of the cocatalyst ligand Cl−, in which the zero-point energy correction was considered.

(triphenylphosphoranylidene)ammonium, X = Cl−, Br−, and I−) in the computational model, and the predicted relative energies are depicted in Figure 3. The predicted relative free energies in Figure 3 show that the ring opening is the rate-determining step, and the corresponding free energy barriers are 13.3, 14.8, and 15.6 kcal/mol for these cocatalysts PPNCl, PPNBr, and PPNI, respectively. Here the free energy spans for PPNCl, PPNBr, and PPNI are 13.3, 15.4, and 19.1 kcal/mol, respectively, suggesting that the catalytic activities of these cocatalysts decrease in the order PPNCl > PPNBr > PPNI, as observed in experiment.39 The relative activity of cocatalysts can be ascribed to their

nucleophilicity in the order Cl− > Br− > I−. The initial complexes of the reactant and the catalyst exhibit the optimized distances of 3.45, 3.52, and 3.87 Å, respectively, between the halide ion X− (X− = Cl−, Br−, and I−) and the CH2 of epoxide, the same as their nucleophilicity order. We note that the presence of the sterically hindered PPN cation makes the carbon atom in methylene as the only accessible site for the nucleophilic attack of X− (refer to Figure S2). The catalytic cycloadditions of CO2 to the epoxide, yielding the catalystbound cyclic carbonate, have overall Gibbs free energies of the reaction ΔG of −11.5, −10.8, and −10.0 kcal/mol (298.15 K) D

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 6. Predicted relative free energies and electronic energies (in parentheses) for the ring-opening and -closure steps in the catalytic coupling reaction of CO2 and an epoxide catalyzed by the Mg−porphyrin catalyst in the presence of cocatalyst PPNX.

Figure 7. Predicted relative free energies and electronic energies (in parentheses) for the ring-opening and -closure steps in the catalytic coupling reaction of CO2 and an epoxide catalyzed by the Zn−porphyrin catalyst in the presence of cocatalyst PPNX.

porphyrin complexes were investigated here. The predicted relative energy profiles for the catalytic cycloaddition of CO2 and PO by Mg− and Zn−porphyrins as well as the cocatalyst Cl− are depicted in Figures 4 and 5, respectively. The optimized structures of initial reactive complexes, transition states, intermediates, and products are collected in Figures S3 and S4. As Figures 4 and 5 show, both coupling reactions follow similar mechanisms with the cycloaddition catalyzed by the Al− porphyrin catalyst, including the ring-opening, CO2-insertion, and ring-closure steps. Similarly, the insertion of CO2 is quite

relative to their reactant complexes for the cocatalysts PPNCl, PPNBr, and PPNI, respectively. 3.2. Cycloaddition Catalyzed by Mg− and Zn− Porphyrin Complexes. Except for the aluminum(III) porphyrin complex, the zinc(II) and magnesium(II) porphyrins also showed very high catalytic activity toward the coupling of CO2 and PO.34,35,56,57 Obviously, the metal center in the Mg− and Zn−porphyrin complexes do not require an axial counterion, unlike the Al(III) center. For comparison, the cycloaddition of CO2 and PO catalyzed by the Mg− and Zn− E

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

Figure 8. Predicted relative free energy profile and electronic energies (in parentheses) for the cycloaddition reaction catalyzed by the Al−corrole complex in the presence of the cocatalyst ligand Cl−, in which the zero-point energy correction was considered.

and Zn−porphyrins, the ring-opening and -closure steps have been investigated by using the complete quaternary ammonium salt cocatalyst PPNX (PPN = bis(triphenylphosphoranylidene)ammonium, X = Cl−, Br−, and I−) in the computational model, and the predicted relative energies are shown in Figures 6 and 7. As Figures 6 and 7 show, here the ring-opening step is predicted to be the rate-determining step in the catalytic conversion of CO2 to cyclic carbonates for both catalysts, as we found for the Al−porphyrin catalyst. Accordingly, the presence of the sterically hindered countercation of cocatalyst generally has remarkable impact on the ring-opening process, and the ring closure is less influenced. The predicted free energy spans for PPNCl, PPNBr, and PPNI cocatalysts are 12.4, 13.8, and 18.8 kcal/mol for the Mg−porphyrin and 15.6, 16.3, and 22.0 kcal/mol for the Zn−porphyrin, respectively, showing that PPNCl generally has the highest performance as the cocatalyst among those explored for the Al−, Mg−, and Zn−porphyrin catalysts. The selected optimized structures of initial reactive complexes, transition states, intermediates, and products are collected in Figures S5 and S6. 3.3. Coupling Reaction of CO2 with Epoxide Catalyzed by the Al−Corrole Complex. The corrole derivatives, as the ring-contracted analogues of porphyrins, are also aromatic macrocyclic systems, and they can serve as the trianionic ligands to accommodate high oxidation-state metals. Previous experimental studies show that the metal−corrole complexes can catalyze the CO2 fixation with epoxides in combination with a cocatalyst.58,59 Here, an Al−corrole complex was constructed, and its catalytic behaviors for the coupling reaction of CO2 and epoxide have been investigated. The predicted relative energy profiles are displayed in Figure 8. Similarly, the catalytic addition of CO2 with epoxide includes the epoxide ring-opening, CO2-insertion, and ring-closure steps, and corresponding free energy barriers are 1.0, 19.7, and 18.2 kcal/mol, respectively, where both the CO2 insertion and the ring closure experience relatively high free energy barriers. Accordingly, the metal corrole complexes exhibit similar activity

facile, and the ring closure is the rate-determining step with the free energy barriers of 17.6 and 19.5 kcal/mol for the Mg− and Zn−porphyrins, respectively. In the ring opening, the metal ion in metalloporphyrins as the Lewis acidic center can activate the C−O bond of the epoxide and make the carbon atom in the methylene more accessible for the nucleophilic attack of halide ion through the dative bonding interactions between the metal center and the epoxy oxygen. The cooperation of the catalyst and cocatalyst may facilitate the cleavage of epoxy O−C bond and form a more stable the ring-opening intermediate with the energy release of 9.8 and 8.8 kcal/mol for Mg-cp2 and Zn-cp2, respectively. In particular, the predicted free energy barriers are 8.6 and 5.0 kcal/mol for the ring-opening step catalyzed by Mg− and Zn−porphyrin catalysts, respectively. In the molecular complexes (M-cp3, M = Al, Mg, and Zn) of CO2 and the ring-opening intermediate (M-cp2), we note that Mg-cp3 and Zn-cp3 have stronger noncovalent interactions than that of Al-cp3, resulting in shorter distances between the metal-bound oxygen and the carbon of CO2 in Mg-cp3 and Zncp3 as compared with that of Al-cp3. Taking the cocatalyst Cl− as an example, the distances between the Mg- and Zn-bound oxygen and the carbon of CO2 are 2.482 and 2.460 Å, respectively, shorter than 2.565 Å in Al-cp3. Such relatively strong noncovalent interactions make the insertion of CO2 in Mg-cp3 and Zn-cp3 more facile than that in Al-cp3, as shown in Figures 2, 4, and 5. During the CO2 insertion, the oxygen ligand exchange from the epoxy O coordination to the oxygen of CO2 is involved in this step, along with the C−O bond coupling. The intermediate M-cp4 (M = Al, Mg, and Zn) undergoes an intramolecular ring closure to yield the product of cyclic carbonate, and this process is the rate-determining step for the coupling of CO2 and PO catalyzed by metalloporphyrins, and the predicted free energy barriers are 18, 17.6, and 19.5 kcal/ mol for Al−, Mg−, and Zn−porphyrin catalysts, respectively. In order to gain insight into effects of the sterically hindered countercation of cocatalyst and different halide ions on key reaction steps in the coupling reaction catalyzed by the Mg− F

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics

optimized structures of species involved in those reaction routes are shown in Figure S10. Clearly, the meso-substitutions of the porphyrin and corrole rings cannot enhance their catalytic activity dynamically for the ring-opening step, which makes the ring opening of PO become difficult slightly. Why was such substitution of the porphyrin ring extensively implemented in development of the porphyrinbased catalysts? We note that the initial reaction system includes the catalyst, the cocatalyst, and the reactant, and the formation of reactive configuration (Cp1) should be important. Our calculations reveal that the meso-substitutions may enhance their binding interactions and thus trigger the initial reaction process. In particular, for the chlorinated aryl-substituted, arylsubstituted, and unsubstituted Al−porphyrin catalysts, the predicted binding energies of their corresponding initial reactive complexes (Cp1) are 56.9, 47.8, and 35.1 kcal/mol, and such binding enhancement can be ascribed to the more strong noncovalent interactions by the introduction of substituents.

for CO2 coupling to the epoxide, in comparison with the metalloporphyrin complexes. Optimized structures for the intermediates and transition states involved in the coupling reaction are displayed in Figure S7. We note that the use of a complete quaternary ammonium salt cocatalyst PPNX in calculation results in quite different relative activity of PPNCl, PPNBr, and PPNI here, in comparison with the metalloporphyrin catalysts investigated, and PPNI is predicted to be most effective among them (see Figures S8 and S9). The presence of the sterically hindered countercation of cocatalyst increases the free energy barrier for the ring-opening step to some extent, while it reduces the free energy barrier for the ring closure remarkably in the coupling reaction catalyzed by the Al−corrole complex. 3.4. meso-Substitution Influence of the Macrocyclic Complex on Catalysis. The meso-substitution of the porphyrin ring can modify the activity of porphyrin-based catalysts,35,37 and furthermore the electronic environment of the substituted groups containing electron-withdrawing and -donating groups also affects the catalytic performance remarkably. We note that here these substituents are generally far from the catalytic site, and the electronic properties of the metal center might be less influenced directly by the mesosubstitution. Selected Mulliken charges of the Al center and the cocatalyst ligand of Cl− in the initial complex (Cp1) and the ring-closure precursor (Cp4), and free energy barriers (kcal/mol) for the ring-opening (TS1) and ring-closure (TS3) steps are compiled into Table 1. The Al center and the Cl− ligand of cocatalyst for

4. CONCLUSION The cycloaddition of CO2 to epoxide catalyzed by the metal porphyrin and corrole complexes have been investigated by extensive DFT calculations. The presented results show that the cyclic carbonate synthesis through the coupling reaction of CO2 and epoxide catalyzed by substituted and unsubstituted porphyrin- and corrole-based complexes follows a similar multistep mechanism, including the ring-opening, CO2insertion, and ring-closure steps, where the ring-opening and -closure steps play a crucial role in catalysis by the metalloporphyrin complexes. The presence of cocatalyst can remarkably reduce the free energy barrier, and the predicted free energy spans for the cocatalysts of PPNCl, PPNBr, and PPNI are 13.3, 15.4, and 19.1 kcal/mol for the Al−porphyrin catalyst, respectively, suggesting that PPNCl is the most effective cocatalyst here, as observed experimentally. The cooperative effect arising from the Lewis acidic metal center of the catalyst and the nucleophilic X− ligand of cocatalyst facilitate the initial ring opening of epoxide. The mesosubstitution of the porphyrin and corrole rings with the aryl and chlorinated aryl groups generally reduces the electrophilicity of metal center, resulting in higher barriers for the ringopening step. However, such substitution enhances the binding interactions among the catalyst, cocatalyst, and reactant through noncovalent interactions, which may drive the initial reaction. Corrole and its derivatives, as the trianionic ligands, can be used to develop corrole-based catalysts with high oxidation-state metals for the catalytic conversion of CO2 into high-value products, and PPNI is predicted to be the most effective cocatalyst among these PPNXs considered here, differing from the catalytic CO2 fixation with epoxides by the metal porphyrin complexes.

Table 1. Predicted Mulliken Charges of the Al Center and the X− Ligand in the Initial Complex (Cp1) and the RingClosure Precursor (Cp4), and Free Energy Barriers (kcal/ mol) for the Ring-Opening (TS1) and Ring-Closure (TS3) Steps catalyst unsubstituted aryl substituted unsubstituted aryl substituted chlorinated aryl substituted

qAl/qX (Cp1)

ΔG⧧ (TS1)

Al−Corrole Complex 1.704/−0.841 1.0 1.692/−0.794 1.7 Al−Porphyrin Complex 1.454/−0.864 2.2 1.431/−0.757 12.1 1.420/−0.748 8.9

qAl/qX (Cp4)

ΔG⧧ (TS3)

1.677/−0.190 1.678/−0.186

18.2 22.6

1.452/−0.199 1.433/−0.192 1.416/−0.187

17.9 17.8 18.0

the most of unsubstituted catalyst are more electrophilic (higher positive charges) and more nucleophilic (higher negative charges), respectively, than those for the substituted catalytic system, which result in quite low free energy barriers for the ring-opening step, as shown in Table 1. For example, the charge populations at the Al center and the Cl− ligand are +1.454 and −0.864, respectively, for the unsubstituted Al− porphurin complex (Cp1), and the predicted free energy barrier is 2.2 kcal/mol for the ring-opening step, much lower than 12.1 kcal/mol for the aryl-substituted system with the charge populations of +1.431 at Al and −0.757 at the Cl− ligand. Table 1 shows the free energy barriers are more sensitive to the charge population at the Cl− ligand, compared to the Al center. On the contrary, these charge populations at Al and X are less changed for the ring-closure precursor (Cp4), and the predicted free energy barriers vary from 17.8−22.6 kcal/mol for the substituted and unsubstituted systems. All the



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.organomet.7b00830. Optimized structures of all reactants, intermediates, transition states, and products by using different computational models, NBO charges, the predicted relative energies for selected steps in the CO2 fixation G

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics



(26) Han, L.; Park, S.-W.; Park, D.-W. Energy Environ. Sci. 2009, 2, 1286−1292. (27) Zhang, Y.; Chan, J. Y. G. Energy Environ. Sci. 2010, 3, 408−417. (28) Peng, J.; Deng, Y. New J. Chem. 2001, 25, 639−641. (29) Kim, H. J. Catal. 2003, 220, 44−46. (30) Kawanami, H.; Sasaki, A.; Matsui, K.; Ikushima, Y. Chem. Commun. 2003, 896−897. (31) Sun, J.; Fujita, S.-i.; Zhao, F.; Arai, M. Green Chem. 2004, 6, 613−616. (32) Cho, H. C.; Lee, H. S.; Chun, J.; Lee, S. M.; Kim, H. J.; Son, S. U. Chem. Commun. 2011, 47, 917−919. (33) Zhao, Y.; Yao, C.; Chen, G.; Yuan, Q. Green Chem. 2013, 15, 446−452. (34) Wang, Q.; Guo, C. H.; Jia, J.; Wu, H. S. J. Mol. Model. 2015, 21, 179. (35) Ema, T.; Miyazaki, Y.; Shimonishi, J.; Maeda, C.; Hasegawa, J. Y. J. Am. Chem. Soc. 2014, 136, 15270−15279. (36) Ohkawara, T.; Suzuki, K.; Nakano, K.; Mori, S.; Nozaki, K. J. Am. Chem. Soc. 2014, 136, 10728−10735. (37) Jin, L.; Jing, H.; Chang, T.; Bu, X.; Wang, L.; Liu, Z. J. Mol. Catal. A: Chem. 2007, 261, 262−266. (38) Bai, D.; Duan, S.; Hai, L.; Jing, H. ChemCatChem 2012, 4, 1752−1758. (39) Qin, Y.; Guo, H.; Sheng, X.; Wang, X.; Wang, F. Green Chem. 2015, 17, 2853−2858. (40) Ma, J.; Liu, J.; Zhang, Z.; Han, B. Green Chem. 2012, 14, 2410− 2420. (41) Huang, J.-W.; Shi, M. J. Org. Chem. 2003, 68, 6705−6709. (42) Wang, J.-Q.; Dong, K.; Cheng, W.-G.; Sun, J.; Zhang, S.-J. Catal. Sci. Technol. 2012, 2, 1480. (43) Ren, Y.; Guo, C. H.; Jia, J. F.; Wu, H. S. J. Phys. Chem. A 2011, 115, 2258−2267. (44) Becke, A. D. Phys. Rev. A: At., Mol., Opt. Phys. 1988, 38, 3098− 3100. (45) Lee, C.; Yang, W.; Parr, R. G. Phys. Rev. B: Condens. Matter Mater. Phys. 1988, 37, 785−789. (46) Perdew, J. P.; Wang, Y. Phys. Rev. B: Condens. Matter Mater. Phys. 1992, 45, 13244−13249. (47) Zhao, Y.; Truhlar, D. G. Theor. Chem. Acc. 2008, 120, 215−241. (48) 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.02; Gaussian, Inc.: Wallingford, CT, 2009. (49) Hay, P. J.; Wadt, W. R. J. Chem. Phys. 1985, 82, 270−283. (50) Wadt, W. R.; Hay, P. J. J. Chem. Phys. 1985, 82, 284−298. (51) Höllwarth, A.; Böhme, M.; Dapprich, S.; Ehlers, A. W.; Gobbi, A.; Jonas, V.; Köhler, K. F.; Stegmann, R.; Veldkamp, A.; Frenking, G. Chem. Phys. Lett. 1993, 208, 237−240. (52) Hariharan, P. C.; Pople, J. A. Theoret. Chim. Acta (Berl.) 1973, 28, 213−222. (53) Fukui, K. Acc. Chem. Res. 1981, 14, 363−368. (54) Devlin, F. J.; Finley, J. W.; Stephens, P. J.; Frisch, M. J. J. Phys. Chem. 1995, 99, 16883−16902. (55) Benson, S. W. The Foundations of Chemical Kinetics; R.E. Krieger: Malabar, FL, 1982. (56) Maeda, C.; Taniguchi, T.; Ogawa, K.; Ema, T. Angew. Chem., Int. Ed. 2015, 54, 134−138.

with an epoxide by the Al−corrole catalyst in the presence of cocatalyst PPNX (PDF)

AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Phone: +86-592-2186081. ORCID

Zexing Cao: 0000-0003-0803-7732 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work was supported by the National Science Foundation of China (21373164 and 21673185).



REFERENCES

(1) Sakakura, T.; Choi, J.-C.; Yasuda, H. Chem. Rev. 2007, 107, 2365−2387. (2) Babu, R.; Kim, S.-H.; Kathalikkattil, A. C.; Kuruppathparambil, R. R.; Kim, D. W.; Cho, S. J.; Park, D.-W. Appl. Catal., A 2017, 544, 126− 136. (3) Martín, C.; Fiorani, G.; Kleij, A. W. ACS Catal. 2015, 5, 1353− 1370. (4) Miceli, C.; Rintjema, J.; Martin, E.; Escudero-Adán, E. C.; Zonta, C.; Licini, G.; Kleij, A. W. ACS Catal. 2017, 7, 2367−2373. (5) Xu, F.; Cheng, W.; Yao, X.; Sun, J.; Sun, W.; Zhang, S. Catal. Lett. 2017, 147, 1654−1664. (6) Wu, L.-X. Int. J. Electrochem. Sci. 2017, 8963−8972. (7) Clements, J. H. Ind. Eng. Chem. Res. 2003, 42, 663−674. (8) Schäffner, B.; Schäffner, F.; Verevkin, S. P.; Börner, A. Chem. Rev. 2010, 110, 4554−4581. (9) Darensbourg, D. J.; Holtcamp, M. W. Coord. Chem. Rev. 1996, 153, 155−174. (10) Darensbourg, D. J.; Mackiewicz, R. M.; Phelps, A. L.; Billodeaux, D. R. Acc. Chem. Res. 2004, 37, 836−844. (11) Mikkelsen, M.; Jørgensen, M.; Krebs, F. C. Energy Environ. Sci. 2010, 3, 43−81. (12) Guo, L.; Wang, C.; Luo, X.; Cui, G.; Li, H. Chem. Commun. 2010, 46, 5960−5962. (13) Kihara, N.; Hara, N.; Endo, T. J. Org. Chem. 1993, 58, 6198− 6202. (14) Yamaguchi, K.; Ebitani, K.; Yoshida, T.; Yoshida, H.; Kaneda, K. J. Am. Chem. Soc. 1999, 121, 4526−4527. (15) Yano, T.; Matsui, H.; Koike, T.; Ishiguro, H.; Fujihara, H.; Yoshihara, M.; Maeshima, T. Chem. Commun. 1997, 1129−1130. (16) Tu, M.; Davis, R. J. J. Catal. 2001, 199, 85−91. (17) Doskocil, E. J. Microporous Mesoporous Mater. 2004, 76, 177− 183. (18) Fujita, S.-i.; Bhanage, B. M.; Ikushima, Y.; Shirai, M.; Torii, K.; Arai, M. Catal. Lett. 2002, 79, 95−98. (19) Bhanage, B. M.; Fujita, S.-i.; Ikushima, Y.; Torii, K.; Arai, M. Green Chem. 2003, 5, 71−75. (20) Caló, V.; Nacci, A.; Monopoli, A.; Fanizzi, A. Org. Lett. 2002, 4, 2561−2563. (21) Buckley, B. R.; Patel, A. P.; Wijayantha, K. G. Chem. Commun. 2011, 47, 11888−11890. (22) Buchard, A.; Kember, M. R.; Sandeman, K. G.; Williams, C. K. Chem. Commun. 2011, 47, 212−214. (23) Decortes, A.; Martinez Belmonte, M.; Benet-Buchholz, J.; Kleij, A. W. Chem. Commun. 2010, 46, 4580−4582. (24) Ijpeij, E. G.; Coussens, B.; Zuideveld, M. A.; van Doremaele, G. H.; Mountford, P.; Lutz, M.; Spek, A. L. Chem. Commun. 2010, 46, 3339−3341. (25) Zhou, H.; Zhang, W.-Z.; Liu, C.-H.; Qu, J.-P.; Lu, X.-B. J. Org. Chem. 2008, 73, 8039−8044. H

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX

Article

Organometallics (57) Maeda, C.; Shimonishi, J.; Miyazaki, R.; Hasegawa, J.; Ema, T. Chem. - Eur. J. 2016, 22, 6556−6563. (58) Nakano, K.; Kobayashi, K.; Ohkawara, T.; Imoto, H.; Nozaki, K. J. Am. Chem. Soc. 2013, 135, 8456−8459. (59) Tiffner, M.; Gonglach, S.; Haas, M.; Schöfberger, W.; Waser, M. Chem. - Asian J. 2017, 12, 1048−1051.

I

DOI: 10.1021/acs.organomet.7b00830 Organometallics XXXX, XXX, XXX−XXX