Вы находитесь на странице: 1из 4

Communication

Cite This: J. Am. Chem. Soc. 2017, 139, 17747−17750 pubs.acs.org/JACS

Molecular Iridium Complexes in Metal−Organic Frameworks


Catalyze CO2 Hydrogenation via Concerted Proton and Hydride
Transfer
Bing An,†,§ Lingzhen Zeng,†,§ Mei Jia,†,§ Zhe Li,† Zekai Lin,‡ Yang Song,‡ Yang Zhou,† Jun Cheng,*,†
Cheng Wang,*,† and Wenbin Lin*,†,‡

Collaborative Innovation Center of Chemistry for Energy Materials, State Key Laboratory of Physical Chemistry of Solid Surfaces,
Department of Chemistry, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, P. R. China

Downloaded via LAWRENCE BERKELEY NATL LABORATORY on July 14, 2020 at 21:22:46 (UTC).

Department of Chemistry, University of Chicago, 929 East 57th Street, Chicago, Illinois 60637, United States
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

*
S Supporting Information

Scheme 1. Schematic Showing a Soxhlet-Type Reflux-


ABSTRACT: Molecular iridium catalysts immobilized in Condensing System To Place MOF Catalysts at a Dynamic
metal−organic frameworks (MOFs) were positioned in Triphasic Interface of CO2/H2−H2O-Catalyst for CO2
the condensing chamber of a Soxhlet extractor for efficient Hydrogenation to Formate
CO2 hydrogenation. Droplets of hot water seeped through
the MOF catalyst to create dynamic gas/liquid interfaces
which maximize the contact of CO2, H2, H2O, and the
catalyst to achieve a high turnover frequency of 410 h−1
under atmospheric pressure and at 85 °C. H/D kinetic
isotope effect measurements and density functional theory
calculations revealed concerted proton−hydride transfer in
the rate-determining step of CO2 hydrogenation, which
was difficult to unravel in homogeneous reactions due to
base-catalyzed H/D exchange.

H ydrogenation of CO2 to formic acid (HCO2H) is of great


interest, not only as an attractive path to convert CO2 to
platform chemicals, but also as a key step in hydrogen storage
using HCO2H as a chemical carrier.1 However, low solubility of
hydrogen in water has limited the catalytic turnover frequency immobilized in MOFs exhibit high activity to hydrogenate CO2
(TOF).2 Herein, we report the design of metal−organic to formate in the reflux-condensing setup.
framework (MOF) catalysts for efficient CO2 hydrogenation in Importantly, performing the reactions with D2 and/or D2O in
a Soxhlet-type reflux-condensing system. Solid MOF catalysts place of H2 and/or H2O under neutral pH revealed ultralarge H/
were packed in a plate that was located below the condenser. D kinetic isotope effects (KIEs) which were previously masked
Droplets of hot condensed water passed through the catalyst bed by base-catalyzed H/D exchange in homogeneous systems. The
intermittently, creating a dynamic gas/liquid/solid interface to ultralarge KIEs for CO2 hydrogenation suggest concerted
proton−hydride transfer via quantum tunneling in the rate-
maximize the contact among CO2, H2, H2O, and the catalyst to
determining step (RDS), which was further supported by density
produce formic acid (Scheme 1). The generated formic acid in functional theory (DFT) calculations.
the aqueous solution was then siphoned to the refluxing flask and H2bpydcOH was prepared from H2bpydc in four steps in a
neutralized by sodium bicarbonate to formate, pushing the 22% overall yield (see Supporting Information [SI] for details,
chemical equilibrium of H2(g) + CO2(g) ⇄ HCO2H(aq) to the Figures S1−S11). The hydroxyl group in bpydcOH serves as an
right. electron-donating group and a hydrogen-bonding site, which has
Key to this reaction design is the construction of solid MOF been reported to facilitate CO2 hydrogenation in homogeneous
catalysts3 based on molecular functionalities.4 A number of Ru,5 Ir complexes.11 Less sterically demanding 4,4′-biphenyl-
Rh,6 Ir,7 Co,8 Re,9 and Fe10 complexes have been reported to be dicarboxylic acid (H2bpdc) was mixed with H2bpydcOH or
highly active in CO 2 hydrogenation, among which Ir- H2bpydc in MOF synthesis to preserve open channels in
polypyridine-based complexes form a versatile category.11 We metalated MOFs. The reaction between the ligand mixture and
incorporated Ir complexes into UiO-type MOFs using 2,2′- ZrCl4 in dimethylformamide (DMF) at 120 °C yielded
bipyridine-5,5′-dicarboxylate ligands with or without −OH
substitution on the 6-position (bpydcOH or bpydc) to provide Received: October 13, 2017
isolated anchoring sites for Ir. After activation, the Ir complexes Published: November 28, 2017

© 2017 American Chemical Society 17747 DOI: 10.1021/jacs.7b10922


J. Am. Chem. Soc. 2017, 139, 17747−17750
Journal of the American Chemical Society Communication

Table 1. Hydrogenation of CO2 to Formate Using Different


Catalysts
entry catalyst Ir (wt%) TONb TOFb (h−1)
1a
mbpyOH-IrCl3-UiO 2.6 543 ± 35 36 ± 2
2a mbpy-IrCl3-UiO 3.8 23 ± 4 1.5 ± 0.2
3a mbpyOH-UiO no formate
4a mbpy-UiO no formate
5a mbpyOH-[IrIII]-UiO 2.4 6149 ± 50 410 ± 3
6a mbpy-[IrIII]-UiO 3.8 417 ± 22 28 ± 2
7a Ir-NP 100 0 0
8c mbpyOH-[IrIII]-UiO 2.4 0 0
9d bpydcOH-IrCl3(OH2) 31 13/11e/8f 0.9/0.7e/ 0.5f
10c bpydcO−-IrCp*OH 18.4 0 0
11d bpydcO−-IrCp*OH 18.4 256 ± 10 17.0 ± 0.7
a
Reaction conditions: 5 mg of catalysts in reflux-condensing system,
0.1 MPa H2/CO2 = 1, 1 M NaHCO3(aq) in bottom flask, temp = 85
°C, time = 15 h. bTON and TOF are both based on Ir. c10 mg of
catalysts, 10 mL of 1 M NaHCO3(aq) in a flask, 0.1 MPa H2/CO2 = 1,
temp = 85 °C, time = 15 h. d10 mg of catalysts, 10 mL of 1 M
NaHCO3(aq) in a bomb, 1 MPa H2/CO2 = 1, temp = 85 °C, time =
15 h. eWith NaHBEt3. fWith AgNO3.

Figure 1. (a) Preparation of mbpyOH-[IrIII]-UiO and mbpy-[IrIII]-UiO


catalysts. (b) Local structure of mbpyOH-[IrIII]-UiO. (c) EXAFS
spectra of mbpyOH-[IrIII]-UiO at the Ir LIII-edge together with the afforded six-coordinated Ir centers with two nitrogen atoms from
fittings in R-space, showing the magnitude (hollow squares, red) and the
real (hollow squares, blue) components. The fitting range is 1.0−3.0 Å
bipyridine moiety, three chlorides, and one oxygen atom from a
in R-space (within the black line). THF molecule in the first coordination sphere (Figures S21 and
S22).
The catalytic activities of mbpyOH-IrCl3-UiO and mbpy-
mbpyOH-UiO and mbpy-UiO MOFs with [Zr6(μ3-O)4(μ3- IrCl3-UiO in CO2 hydrogenation were evaluated in the Soxhlet-
OH)4]12+ as secondary building units (SBUs) (Figure 1a,b). type reflux-condensing system. The bottom flask contained a
The UiO-type MOF structures of mbpyOH-UiO and mbpy- basic aqueous solution of 1 M NaHCO3. After a reaction of 15 h,
UiO were revealed by the similarity of their powder X-ray formate was detected in the bottom flask by 1H NMR and
diffraction (PXRD) patterns to that of UiO-67 (Figure S12).12 quantified by high performance liquid chromatography (HPLC)
Transmission electron microscopy (TEM) showed octahedral (Table 1, entries 1 and 2; Figures S28−S30). No gaseous product
nanocrystals of the MOFs (Figures S13 and S14). Chemical was detected in the headspace by gas chromatography (Figure
formulas of these MOFs were established by a combination of S31). mbpyOH-IrCl3-UiO gave an Ir-based turnover number (Ir-
thermogravimetric analyses to determine mass ratios of metal/ TON) of 543 ± 35 in 15 h with an average Ir-TOF of 36 ± 2 h−1
organics (Figures S15 and S16) and proton nuclear magnetic under 1 atm of H2/CO2 = 1/1, while mbpy-IrCl3-UiO afforded
resonance (1H NMR) spectra of digested MOFs in NaOD/D2O an Ir-TON of 23 ± 4 and an Ir-TOF of 1.5 ± 0.2. Control
to quantify amounts of ligands and capping formates (Figures experiments with non-metalated mbpyOH-UiO or mbpy-UiO
S17 and S18). We obtained chemical formulas of Zr6(μ3-O)4(μ3- gave no products (Table 1, entries 3 and 4), confirming the role
OH)4(bpdc)2.67(bpydcOH)2.67(HCO2)1.33 for mbpyOH-UiO of Ir as the catalytic center. The temperature of the condensed
and Zr6(μ3-O)4(μ3-OH)4(bpdc)2.07(bpydc)3.10(HCO2)1.66 for water at the catalyst bed was measured to be 85 °C.
mbpy-UiO. Monitoring the Ir-TONs of mbpyOH-IrCl3-UiO over time
Postsynthetic metalation of the MOFs with IrCl3·3H2O in a revealed an induction period of approximately 8 h, suggesting the
mixture of tetrahydrofuran (THF) and DMF gave metalated need to transform the bpyOH-IrIIICl3(solvent) species to an
MOFs mbpyOH-IrCl3-UiO and mbpy-IrCl3-UiO with bpyOH- active form, which possibly involves breaking Ir−Cl bonds. To
IrCl3(THF) and bpy-IrCl3(THF) moieties, respectively. Ir facilitate chloride removal, mbpyOH-IrCl3-UiO and mbpy-IrCl3-
loading levels determined by inductively coupled plasma-mass UiO were treated with NaBHEt3 followed by air oxidation to
spectrometry (ICP-MS) correspond to incorporation of 10.5 yield mbpyOH-[IrIII]-UiO and mbpy-[IrIII]-UiO (Figure 1a). A
mol% of Ir on the bpyOH sites in mbpyOH-IrCl3-UiO and 13.7 mixture of [bpyO-IrIIICl2(OH2)2], [bpyO-IrIIICl(OH2)3]+, and
mol% of Ir on the bpy sites in mbpy-IrCl3-UiO. PXRD patterns [bpyO-IrIII(OH)(OH2)3]+ were present in mbpyOH-[IrIII]-UiO
showed that both metalated MOFs remained crystalline (Figure with different extents of chloride removal (Figures S23−S26). In
S12). Nitrogen sorption experiments gave BET surface areas of the following discussions, we focus on the [bpyO−-IrIII(OH)-
705 m2/g for mbpyOH-Ir-UiO and 843 m2/g for mbpy-Ir-UiO (OH2)3]+ species as a representative site. The degree of
(Figure S19). deprotonation of [bpyO−-IrIII(OH)(OH2)3]+ was based on
Local coordination environments of Ir in MOFs were probed DFT calculation of pKa values of protons in each coordinated
by X-ray absorption spectroscopy (XAS) at the Ir LIII-edge. X-ray water and the hydroxyl group on the ligand (Figures S53−S55).
absorption near edge structure (XANES) of mbpyOH-IrCl3-UiO The activated mbpyOH-[IrIII]-UiO gave much higher catalytic
confirmed the oxidation state of IrIII by a rising edge similar to activity than the unactivated one, achieving an Ir-TON of 6149 ±
that of IrCl3 (main peak at 11217 eV) (Figure S20).4e Fittings of 50 in 15 h and an Ir-TOF of 410 ± 3 h−1 (Table 1, entry 5, Figure
the extended X-ray absorption fine structure (EXAFS) regions S32) with a shorter induction period of ∼3 h (Figure 2a). These
17748 DOI: 10.1021/jacs.7b10922
J. Am. Chem. Soc. 2017, 139, 17747−17750
Journal of the American Chemical Society Communication

Figure 2. Catalytic data of the Ir-based catalysts: (a) TONs as a function


of reaction time and (b) KIE values under different conditions.

TON and TOF values compare favorably with other catalysts for
CO2 hydrogenation at ambient pressure (Table S6).
mbpyOH-[IrIII]-UiO was recovered and reused for at least
three times without obvious loss of catalytic activity (Figure
S34). The recovered mbpyOH-[IrIII]-UiO after catalysis
remained crystalline, as shown by PXRD patterns (Figure
S12). The oxidation state of IrIII in mbpyOH-[IrIII]-UiO
remained unchanged based on XANES and XPS studies (Figures
S20 and S26). ICP-MS analyses of the aqueous solution in the
refluxing flask showed negligible metal leaching (0.18% of Ir and
0.05% of Zr). Importantly, a second 15 h reaction of this solution
produced no additional products, proving that the mbpyOH-
[IrIII]-UiO instead of the leached species was the active catalyst.
Control experiments using Ir nanoparticles (Ir-NP) did not give
any CO2 hydrogenation products (Table 1, entry 7; Figures S35
and S36), ruling out Ir-NPs as catalysts.
The reflux-condensing system for placing the catalysts at
liquid−gas interface greatly enhanced their TOFs compared to
catalysis with the same catalysts in solutions. For example,
mbpyOH-[IrIII]-UiO in a 1 M NaHCO3 solution generated Figure 3. (a) Proposed mechanism for the hydrogenation of CO2 by
negligible amounts of formate at 85 °C (Table 1, entry 8), in part mbpyOH-[IrIII]-UiO with activation enthalpies (see SI). (b) Calculated
due to the decomposition of the MOF in basic solutions. free-energy profiles for the proposed catalytic cycle for CO2 hydro-
Furthermore, homogeneous molecular bpydcOH-IrCl3(OH2) genation to formic acid.
complex with or without treatment with NaBHEt3 or AgNO3
gave low activities with TOFs of <1 h−1 at gas pressures of 1 MPa possible when quantum tunneling of proton is involved in the
(Table 1, entry 9). The homogeneous catalysts may form inactive RDS. In contrast, mbpy-[IrIII]-UiO gave a KIE (H2O/H2-D2O/
dimers or oligomers, which is prevented in MOFs due to active H2) value of 7.8.
site isolation on the framework. Consistent with this hypothesis, On the other hand, when D2 was used in the gas phase and
introduction of a bulky pentamethylcyclopentadiene (Cp*) H2O was used as the solvent, DCO2Na was exclusively produced
group in the complex bpydcO−-IrCp*OH slowed catalyst for both mbpyOH-[IrIII]-UiO and mbpy-[IrIII]-UiO (Figures
deactivation to afford a moderate Ir-TON of 256 ± 10 in 15 h S43 and S44), again confirming that the deuterium in DCO2−
at a gas pressure of 1 MPa and a temperature of 85 °C (Table 1, originated from D2. The KIE (H2O/H2−H2O/D2) value ([rate
entry 11, Figure S37), which is still much lower than that of with H2]/[rate with D2]) was 19 for mbpyOH-[IrIII]-UiO and
mbpyOH-[IrIII]-UiO in the reflux-condensing setup at a gas 5.9 for mbpy-[IrIII]-UiO. Such a KIE value over 10 for mbpyOH-
pressure of 0.1 MPa. [IrIII]-UiO also indicates nuclear tunneling during the hydride
The enhanced activity of mbpyOH-[IrIII]-UiO over that of transfer to CO2.
mbpy-[IrIII]-UiO by 11 times (Table1, entry 6, Figure S33) When the reaction was carried out using a combination of D2
demonstrated the importance of hydroxyl group on the bpyOH in the gas phase and D2O in the aqueous phase, negligible
ligand.13 To further probe the mechanism, we measured solvent amount of formate was produced with mbpyOH-[IrIII]-UiO
KIEs (Figures 2b, S38−S40) by comparing the reaction rates (Figures S45 and S46), affording an overall KIE (H2/H2O−D2/
using D2O vs H2O in two separate experiments. HCO2Na but D2O) value of over 3000, which is no less than the product of KIE
not DCO2Na was produced with D2O in the bottom flask and H2 (H2O/H2-D2O/H2) and KIE (H2O/H2−H2O/D2). This is only
in the headspace for both mbpyOH-[IrIII]-UiO and mbpy-[IrIII]- possible when proton transfer and hydride transfer to CO2 take
UiO as revealed by 1H NMR and 2H NMR (Figures S41 and place in the same RDS.14 Similarly, the double deuterium
S42). These results suggest that hydrogen in formate comes from experiment for mbpy-[IrIII]-UiO gave a KIE (H2/H2O−D2/
H2 in the gas phase. As the heterolytic splitting of H2 generates D2O) value over 45, larger than the product of KIE (H2O/H2-
iridium hydride species, formates must be produced from CO2 D2O/H2) and KIE (H2O/H2−H2O/D2).
via a hydride transfer step. Importantly, a dramatic reduction of Based on these experimental observations, we propose a
reaction rate was observed using D2O as the solvent with a large catalytic cycle which is supported by DFT calculations as shown
KIE (H2O/H2-D2O/H2) value ([rate in H2O]/[rate in D2O]) of in Figure 3. The transition states of H2 heterolysis (A*) and
140 for mbpyOH-[IrIII]-UiO. Such a large KIE value is only proton−hydride transfer (C*) gave activation energies of Ea1 =
17749 DOI: 10.1021/jacs.7b10922
J. Am. Chem. Soc. 2017, 139, 17747−17750
Journal of the American Chemical Society Communication

0.34 kcal·mol−1 and Ea2 = 16.26 kcal·mol−1, supporting the Thousand Talents Program of the P. R. China, Xiamen
proton−hydride transfer to be the RDS. The corresponding University, and the U.S. National Science Foundation (DMR-
energy profiles are shown in Figure 3b. The equilibria between 1308229).
B/C and D/A depend on the reaction pH and the pKa’s of B and
A (Figure S57). In addition, calculated H/D KIE value
considering only zero-point energies gave a small number of
■ REFERENCES
(1) (a) Klankermayer, J.; Leitner, W. Science 2015, 350, 629−630.
∼2, suggesting that the quantum tunneling effect plays a key role (b) Jessop, P. G.; Joo, F.; Tai, C.-C. Coord. Chem. Rev. 2004, 248, 2425−
in the experiment. Similar concerted proton and hydride transfer 2442. (c) Klankermayer, J.; Wesselbaum, S.; Beydoun, K.; Leitner, W.
has been previously reported for the alcohol dehydrogenase and Angew. Chem., Int. Ed. 2016, 55, 7296−7343.
an Os complex.15 (2) (a) Munshi, P.; Main, A. D.; Linehan, J. C.; Tai, C.-C.; Jessop, P. G.
It is worth noting that the homogeneous catalyst bpydcO−- J. Am. Chem. Soc. 2002, 124, 7963−7971. (b) Jessop, P. G.; Ikariya, T.;
Noyori, R. Nature 1994, 368, 231.
IrCp*OH in basic solution gave different isotope distributions in (3) (a) Kornienko, N.; Zhao, Y.; Kley, C. S.; Zhu, C.; Kim, D.; Lin, S.;
products. With both H2O/D2 and D2O/H2, the homogeneous Chang, C. J.; Yaghi, O. M.; Yang, P. J. Am. Chem. Soc. 2015, 137, 14129−
catalyst produced a mixture of DCO2Na and HCO2Na with the 14135. (b) Chen, X.; Jiang, H.; Hou, B.; Gong, W.; Liu, Y.; Cui, Y. J. Am.
DCO2Na/HCO2Na ratios of 0.80 and 0.98, respectively (Figures Chem. Soc. 2017, 139, 13476−13482. (c) Zeng, L.; Guo, X.; He, C.;
S47−S52). This D/H exchange likely results from base-assisted Duan, C. ACS Catal. 2016, 6, 7935−7947. (d) Fei, H.; Cohen, S. M.
reductive elimination of the slightly protonic Ir−H(D) bond to Chem. Commun. 2014, 50, 4810−4812. (e) Hong, D.-Y.; Hwang, Y. K.;
produce IrI species, which can be further protonated by Serre, C.; Ferey, G.; Chang, J.-S. Adv. Funct. Mater. 2009, 19, 1537−
D2O(H2O) to regenerate Ir-D(H). The D/H exchange masked 1552. (f) Ramos-Fernandez, E. V.; Pieters, C.; van der Linden, B.; Juan-
the KIE to give moderate apparent values of 1.1 and 1.6 with Alcaniz, J.; Serra-Crespo, P.; Verhoeven, M. W. G. M.;
Niemantsverdriet, H.; Gascon, J.; Kapteijn, F. J. Catal. 2012, 289, 42−
H2O/D2 and D2O/H2, respectively.
52.
In summary, we performed efficient catalytic CO2 hydro- (4) (a) Xu, H.-Q.; Hu, J.; Wang, D.; Li, Z.; Zhang, Q.; Luo, Y.; Yu, S.-
genation to form formic acid/formate by placing MOF catalysts H.; Jiang, H.-L. J. Am. Chem. Soc. 2015, 137, 13440−13443. (b) Feng,
in a Soxhlet-type reflux-condensing system. Replacing H2 and/or D.; Chung, W.-C.; Wei, Z.; Gu, Z.-Y.; Jiang, H.-L.; Chen, Y.-P.;
H2O by D2 and/or D2O in the reaction revealed ultralarge H/D Darensbourg, D. J.; Zhou, H.-C. J. Am. Chem. Soc. 2013, 135, 17105−
KIEs by preventing base-catalyzed H/D exchange in homoge- 17110. (c) Korzyński, M. D.; Dincă, M. ACS Cent. Sci. 2017, 3, 10−12.
neous reactions. The ultralarge KIEs and DFT calculations (d) Yang, D.; Odoh, S. O.; Wang, T. C.; Farha, O. K.; Hupp, J. T.;
suggest concerted proton−hydride transfer in the rate- Cramer, C. J.; Gagliardi, L.; Gates, B. C. J. Am. Chem. Soc. 2015, 137,
determining step of CO2 hydrogenation. This work highlights 7391−7396. (e) Wang, C.; Xie, Z.; de Krafft, K. E.; Lin, W. J. Am. Chem.
opportunities to use MOFs as solid molecular catalysts in Soc. 2011, 133, 13445−13454. (f) Wang, C.; Liu, D.; Lin, W. J. Am.
designer reactors to improve catalytic conversions and to Chem. Soc. 2013, 135, 13222−13234.
(5) (a) Kajiwara, T.; Fujii, M.; Tsujimoto, M.; Kobayashi, K.; Higuchi,
uncover interesting mechanistic details.


M.; Tanaka, K.; Kitagawa, S. Angew. Chem., Int. Ed. 2016, 55, 2697−
2700. (b) Rezayee, N. M.; Huff, C. A.; Sanford, M. S. J. Am. Chem. Soc.
ASSOCIATED CONTENT 2015, 137, 1028−1031.
*
S Supporting Information (6) Himeda, Y.; Miyazawa, S.; Hirose, T. ChemSusChem 2011, 4, 487−
The Supporting Information is available free of charge on the 493.
ACS Publications website at DOI: 10.1021/jacs.7b10922. (7) (a) Hull, J. F.; Himeda, Y.; Wang, W.-H.; Hashiguchi, B.; Periana,
R.; Szalda, D. J.; Muckerman, J. T.; Fujita, E. Nat. Chem. 2012, 4, 383−
Experimental procedures; spectroscopic characterization 388. (b) Tanaka, R.; Yamashita, M.; Nozaki, K. J. Am. Chem. Soc. 2009,
of precatalysts and catalysts; X-ray absorption spectro- 131, 14168−14169.
scopic analysis; evaluation of catalytic performances; KIEs (8) Jeletic, M. S.; Mock, M. T.; Appel, A. M.; Linehan, J. C. J. Am. Chem.
and quantifications of HCOOH and DCOOH; and DFT Soc. 2013, 135, 11533−11536.
calculations (PDF) (9) (a) Sato, S.; Koike, K.; Inoue, H.; Ishitani, O. Photochem. Photobiol.


Sci. 2007, 6, 454−461. (b) Morris, A. J.; Meyer, G. J.; Fujita, E. Acc.
Chem. Res. 2009, 42, 1983−1994.
AUTHOR INFORMATION (10) Boddien, A.; Mellmann, D.; Gärtner, F.; Jackstell, R.; Junge, H.;
Corresponding Authors Dyson, P. J.; Laurenczy, G.; Ludwig, R.; Beller, M. Science 2011, 333,
*chengjun@xmu.edu.cn 1733−1736.
*wangchengxmu@xmu.edu.cn (11) Wang, W.-H.; Hull, J. F.; Muckerman, J. T.; Fujita, E.; Himeda, Y.
*wenbinlin@uchicago.edu Energy Environ. Sci. 2012, 5, 7923.
(12) Cavka, J. H.; Jakobsen, S.; Olsbye, U.; Guillou, N.; Lamberti, C.;
ORCID Bordiga, S.; Lillerud, K. P. J. Am. Chem. Soc. 2008, 130, 13850−13851.
Cheng Wang: 0000-0002-7906-8061 (13) (a) Dub, P. A.; Gordon, J. C. ACS Catal. 2017, 7, 6635−6655.
Wenbin Lin: 0000-0001-7035-7759 (b) Belkova, N. V.; Epstein, L. M.; Filippov, O. A.; Shubina, E. S. Chem.
Rev. 2016, 116, 8545−8587.
Author Contributions
§ (14) O’Leary, M. H. Annu. Rev. Biochem. 1989, 58, 377−401.
B.A., L.Z., and M.J. contributed equally. (15) (a) Agarwal, P. K.; Webb, S. P.; Hammes-Schiffer, S. J. Am. Chem.
Notes Soc. 2000, 122, 4803−4812. (b) Nagorski, R. W.; Richard, J. P. J. Am.
The authors declare no competing financial interest. Chem. Soc. 2001, 123, 794−802. (c) Cui, Q.; Elstner, M.; Karplus, M. J.


Phys. Chem. B 2002, 106, 2721−2740. (d) Ess, D. H.; Schauer, C. K.;
ACKNOWLEDGMENTS Meyer, T. J. J. Am. Chem. Soc. 2010, 132, 16318−16320.
We acknowledge funding support from the National Natural
Science Foundation and Ministry of Science and Technology of
the P. R. China (NSFC21671162, 2016YFA0200702, NSFC-
21471126, NSFC21373166, NSFC21621091), the National
17750 DOI: 10.1021/jacs.7b10922
J. Am. Chem. Soc. 2017, 139, 17747−17750

Вам также может понравиться