Skip to main content
ACS AuthorChoice logoLink to ACS AuthorChoice
. 2020 Sep 30;142(41):17384–17392. doi: 10.1021/jacs.0c05913

Quantitative Electro-Reduction of CO2 to Liquid Fuel over Electro-Synthesized Metal–Organic Frameworks

Xinchen Kang , Bin Wang , Kui Hu , Kai Lyu , Xue Han , Ben F Spencer , Mark D Frogley §, Floriana Tuna †,, Eric J L McInnes , Robert A W Dryfe , Buxing Han ⊥,*, Sihai Yang †,*, Martin Schröder †,*
PMCID: PMC7586324  PMID: 32997941

Abstract

graphic file with name ja0c05913_0005.jpg

Efficient electro-reduction of CO2 over metal–organic framework (MOF) materials is hindered by the poor contact between thermally synthesized MOF particles and the electrode surface, which leads to low Faradaic efficiency for a given product and poor electrochemical stability of the catalyst. We report a MOF-based electrode prepared via electro-synthesis of MFM-300(In) on an indium foil, and its activity for the electrochemical reduction of CO2 is assessed. The resultant MFM-300(In)-e/In electrode shows a 1 order of magnitude improvement in conductivity compared with that for MFM-300(In)/carbon-paper electrodes. MFM-300(In)-e/In exhibits a current density of 46.1 mA cm–2 at an applied potential of −2.15 V vs Ag/Ag+ for the electro-reduction of CO2 in organic electrolyte, achieving an exceptional Faradaic efficiency of 99.1% for the formation of formic acid. The facile preparation of the MFM-300(In)-e/In electrode, coupled with its excellent electrochemical stability, provides a new pathway to develop efficient electro-catalysts for CO2 reduction.

Introduction

Efficient conversion of CO2 into chemical feedstocks and fuels is a highly desirable but extremely challenging target.13 Reduction of CO2 into useful chemicals via thermo-catalysis and photocatalysis has been studied very widely.47 However, the former often requires both high temperature and pressure to activate CO2, while the latter relies heavily upon the use of sacrificial agents, typically organic amines, thus limiting long-term applications. Electro-catalysis enables the storage of intermittent renewable energy into chemical energy,810 and there are powerful drivers for the development of active, selective and stable electro-catalysts for the efficient reduction of anthropogenic CO2 emission via conversion to valuable chemicals.1114 Electrocatalytic reduction of CO2 normally relies upon the use of active transition metals (e.g., Cu, Co, and Pd) to reduce the high overpotentials required to activate CO2 to the CO2•− radical anion or other intermediates that can be converted further.1517

Metal–organic frameworks (MOFs) are crystalline hybrid materials constructed from metal ions or clusters bridged by polydentate organic ligands.1820 Compared with conventional electro-catalysts, MOFs have unique features for the electrochemical reduction of CO2.21,22 Metal sites (e.g., Co, Cu, Fe, and Ni) that show activity for CO2 electro-reduction2326 can be readily incorporated into MOF structures via single-site dispersion. Additionally, the intrinsic and tunable porosity of MOFs can lead to high capacity adsorption and selectivity to CO2,27,28 thus promoting activation and further conversion. Gaseous products obtained from electro-reduction of CO2 (e.g., CO and CH4) often show lower adsorption in MOFs than CO2 and can thus be readily recovered and the MOF electrode regenerated.21 MOFs have inherent design flexibility via choice of metal ions and organic ligands, and therefore, their function and activity for electro-reduction of CO2 can be optimized.29

To date, universal approaches to fabricate efficient MOF-based electrodes showing high charge-transfer capacity have been rarely reported.3033 Normal methods are based upon the doping of MOF materials onto an electrode substrate.34 Although various electrodes loaded with MOFs have been studied for the electro-reduction of CO2, the poor contact between the catalyst and substrate surface limits their performance.21 Indium-based materials have demonstrated excellent performance for the electro-reduction of CO2.35,36 Although electrodeposition of MOFs, particularly Cu- and Zn-based systems, has been described,3740 the electro-synthesis of In-based MOFs has not been reported previously. We therefore sought to develop an In electrode decorated with a MOF film that might incorporate active defect sites and show high charge transfer capacity.

Here, we report a facile route to the fabrication of such decorated electrodes via electro-synthesis of MOFs onto a metal foil substrate. The MFM-300(In)-e/In electrode (“e” is short for electro-synthesis) prepared in this way shows a 1 order of magnitude enhancement in electrical conductivity compared with electrodes prepared using MOFs prepared by thermo-chemical methods. More importantly, exceptional activities for the electro-reduction of CO2 are observed with a current density of 46.1 mA cm–2 under an applied potential of −2.15 V vs Ag/Ag+. A Faradaic efficiency of 99.1% for formic acid after 2 h of electrolysis is observed. We also found that the electro-synthesized MOF incorporates structural defects in the form of additional framework In3+ sites, which, coupled with improved charge transfer capacity, greatly promote the activation of CO2 to the radicals, consistent with the observed excellent electrocatalytic activity and stability of the catalyst. Density functional theory (DFT) calculations are applied to reveal the mechanism of catalysis.

Results and Discussion

Materials Preparation and Characterizations

Samples of MFM-300(In)-t (“t” is short for thermo-synthesis) were obtained by solvothermal reaction of In(NO3)3·5H2O and biphenyl-3,3′,5,5′-tetracarboxylic acid (H4L; Figure S1) in a mixture of DMF and MeCN at 85 °C for 3 days.41 The crystal structure of MFM-300(In) is shown in Figure S2. The electro-synthesis of MFM-300(In)-e was achieved using indium foil (In-foil) as both the cathode and anode in a solvent mixture of DMF/dioxane/water containing H4L with 1-ethyl-3-methylimidazolium acetate (EmimOAc, structure shown in Figure S3) as the supporting electrolyte (Figure 1a). The formation of MOF occurred rapidly at an applied potential of 10 V and 60 °C, and the In-foil anode was covered by particles of MFM-300(In)-e within 200 s (Figure 1b). The electro-synthesis was then terminated as the surface of In-foil became fully covered with MOF particles thus limiting the supply of further In3+ ions from the anode for MOF synthesis. No MOF particle was detected in the electrolyte, indicating that particles of MFM-300(In)-e have a strong interaction with the In-foil substrate. The MFM-300(In)-e/In electrode was then immersed in acetone for 2 h and dried before being used as the MFM-300(In)-e/In electrode. Scanning electron microscopy confirms that the In-foil metal surface is coated uniformly with MFM-300(In)-e, which has an octahedral morphology with a narrow particle size distribution centered at ∼500 nm (Figures 1c–e). By comparison, MFM-300(In)-t shows cube-shaped crystals with an average particle size of ∼2.5 μm (Figure S4). These differences in morphology result only in small differences in the relative intensity of powder X-ray diffraction (PXRD) peaks for MFM-300(In)-t and MFM-300(In)-e (Figure 2a).

Figure 1.

Figure 1

(a) Schematic diagram for the electro-synthesis of MFM-300(In)-e. (b) Photograph of the as-synthesized MFM-300(In)-e on indium foil (0.5 × 1 cm2). (c and d) SEM images for MFM-300(In)-e. (e) Particle size distribution of MFM-300(In)-e. The scale bars of panels c and d are 200 and 2 μm, respectively.

Figure 2.

Figure 2

Characterization of as-synthesized MFM-300(In)-t (black line) and MFM-300(In)-e (red line). (a) PXRD patterns. (b) FTIR spectra. (c) N2 adsorption/desorption isotherms at 77 K. (d) Pore size distribution derived from NLDFT calculations.

Powder crystalline samples of MFM-300(In)-e were carefully scraped off the In-foil for characterisations. The crystal structure and phase purity of MFM-300(In)-e were confirmed by PXRD (Figure 2a); the increased peak width of MFM-300(In)-e compared with MFM-300(In)-t is consistent with their particle size distributions. Full chemical analysis suggests that the indium contents are 35.3 and 32.5 wt % in MFM-300(In)-e and MFM-300(In)-t, respectively. Negligible N content was detected in MFM-300(In)-e, indicating that the ionic liquid cations were removed completely during the solvent exchange process. The higher content of In3+ in MFM-300(In)-e suggests that acetate anions OAc introduced in the synthesis with EmimOAc are bound to the framework to balance the charge. The slightly higher residual solid observed in the thermogravimetric analysis (TGA) is consistent with the presence of additional In3+ in the framework of MFM-300(In)-e (Figure S5), while the slightly lower stability of MFM-300(In)-e compared with MFM-300(In)-t originates most likely from defects in the framework structure. The band observed near 1550 cm–1 by Fourier-transform infrared (FTIR) spectroscopy is assigned to the vibration of the carboxylate group, but has an apparent red shift (20 cm–1) for MFM-300(In)-e (Figure 2b), consistent with the presence of some coordinated acetate in this material. The Brunauer–Emmett–Teller (BET) surface areas of MFM-300(In)-t and MFM-300(In)-e are determined to be 1043 and 863 m2 g–1, respectively (Figure 2c). The slightly reduced surface area of MFM-300(In)-e is attributed to the partial collapse of micropores during the rapid electro-synthesis. Pore size distributions of MFM-300(In)-t and MFM-300(In)-e estimated by nonlocal density functional theory (NLDFT) calculations (Figure 2d) confirm a micropore profile centered at 6.5 Å. In addition, there are two small peaks centered at 9.5 and 11.5 Å in MFM-300(In)-e owing to the presence of defects in the framework. During the electro-synthesis, we propose that the In-foil is oxidized to In3+ which binds OAc anions from the EmimOAc electrolyte at the surface of the electrode thus preventing full dissolution of In3+. Binding of the tetracarboxylate ligand to In3+ occurs at the surface either directly or via replacement of OAc to assemble the framework structure uniformly across the In-foil surface.42 Since there are abundant In3+ ions at and around the anode, the as-synthesized MFM-300(In)-e contains excess In3+ sites with some associated OAc. Both MFM-300(In)-t and MFM-300(In)-e show typical features for In3+ at the surface as characterized by X-ray photoelectron spectroscopy (XPS) (Figure S6 and Table S1).

Electrochemical Reduction of CO2

Carbon paper (CP) has a very coarse surface that can support materials to fabricate electrodes. Samples of MFM-300(In)-t and MFM-300(In)-e were loaded onto CP substrates using Nafion D-521 as a binder to fabricate MFM-300(In)-t/CP and MFM-300(In)-e/CP electrodes, respectively. Indium foil has a very smooth surface, and we were thus unable to fabricate MFM-300(In)-t/In since loading thermally prepared MFM-300(In)-t onto this metal surface was unsuccessful. The surface structures of MFM-300(In)-e/In, MFM-300(In)-t/CP and MFM-300(In)-e/CP were characterized by XPS (Figure S7 and Table S1). These spectra show typical features of In3+ cations, and the shifts in binding energy [+0.5 and −0.5 eV for MFM-300(In)-t and MFM-300(In)-e, respectively] compared to powder samples originate from the different conductivities of the surface substrates.

All three electrodes were investigated for electrochemical reduction of CO2 in an H-type cell (Figure S8) with 0.5 M EmimBF4 (1-ethyl-3-methylimidazolium tetrafluoroborate) in MeCN and 0.5 M H2SO4 as catholyte and anolyte, respectively. Cyclic voltammetry (CV) of MFM-300(In)-e/In was studied as a function of gas loading (Figure 3a). Large current densities were generated in the CO2-saturated electrolyte, while negligible densities were observed for the electrolyte saturated with N2. Controlled potential electrolysis at −2.0∼−2.3 V vs Ag/Ag+ was conducted at room temperature under a flow of CO2 into the electrolyte. All liquid and gas-phase products were quantified by 1H NMR spectroscopy and gas chromatography (GC), respectively. Interestingly, only H2 was observed in the gas phase and formic acid was identified as the sole product in the liquid phase. The electrolysis was performed for 2 h and results are summarized in Figure 3b, 3c. The current density increases with increasing applied potential in all cases and the Faradaic efficiency for formic acid (FEHCOOH) reaches 99.1% with a current density of 46.1 mA cm–2 at −2.15 V vs Ag/Ag+ for MFM-300(In)-e/In. Significantly, MFM-300(In)-e/In shows a higher current density compared with MFM-300(In)-e/CP and MFM-300(In)-t/CP at −2.15 V vs Ag/Ag+; the latter gives the lowest FEHCOOH at all potentials. The Faradaic efficiency for hydrogen (FEH2) has an inverse trend over these electrodes (Figure S9) with MFM-300(In)-t/CP higher than MFM-300(In)-e/CP, which is higher than MFM-300(In)-e/In. In-foil, as a smooth metallic electrode (Figure S10), exhibits a current density of 19.7 mA cm–2 and FEHCOOH of 50.7% in 0.5 M EmimBF4/MeCN at −2.15 V vs Ag/Ag+, lower than that observed for MFM-300(In)-e/In under the same conditions (Table 1).

Figure 3.

Figure 3

Electrochemical reduction of CO2 over MFM-300(In)-t/CP (black), MFM-300(In)-e/In (red), and MFM-300(In)-e/CP (blue) in 0.5 M EmimBF4/MeCN. (a) CV of MFM-300(In)-e/In in N2- and CO2-saturated electrolytes. (b) Plots of current density vs applied potential. (c) Plots of FEHCOOH vs applied potential. (d) Plots of current density vs time before and after the introduction of CO2. (e) Plots of current density vs time at −2.15 V vs Ag/Ag+. (f) Plots of FEHCOOH vs time at −2.15 V vs Ag/Ag+. (g) Current density for five cycles of electrolysis at −2.15 V vs Ag/Ag+ over 2 h. (h) FEHCOOH for five cycles of electrolysis at −2.15 V vs Ag/Ag+ over 2 h.

Table 1. Catalytic Performance for Electrochemical Reduction of CO2 to Formic Acid Using Different Cathodes in the Organic Electrolyte.

electrode electrolyte potential (V vs Ag/Ag+) j (mA cm–2) FEHCOOH (%) ref
MFM-300(In)-e/In 0.5 M EmimBF4/MeCN –2.15 46.1 99.1 this work
MFM-300(In)-t/CP 0.5 M EmimBF4/MeCN –2.15 28.4 68.1 this work
MFM-300(In)-e/CP 0.5 M EmimBF4/MeCN –2.15 28.0 81.5 this work
In-foil 0.5 M EmimBF4/MeCN –2.15 19.7 50.7 this work
MoP@In-PC BmimPF6/MeCN/H2O (30/65/5) –2.2 43.8 96.5 (43)
Pb 700 mM BmimTrizu/MeCN/5 wt %H2O –2.4 24.5 95.2 (44)
Pb BmimPF6/MeCN/H2O (30/65/5) –2.2 17.8 95.3 (45)
Sn BmimPF6/MeCN/H2O (30/65/5) –2.2 15.8 95.0 (45)
Pb-PhyA BzmimBF4/MeCN/H2O (12.8/77.3/9.9) –2.25 30.5 92.7 (46)
SnO2@N-PC 0.5 M BmimPF6/MeCN –2.2 28.4 94.1 (47)
Bi 250 mM (DBU-H)PF6/acetonitrile/0.1 M TBAPF6 –1.95a 27 77 (56)
Sn powder 0.5 M EmimN(CN)2/H2O –1.2b 0.633 81.9 (57)
a

vs SCE.

b

vs RHE.

The dependence of current density and FEHCOOH on time was studied further (Figure 3d). N2 was first charged into the electrolyte, and an imperceptible current density was observed for all electrodes investigated. Significant increases in current density were observed over all electrodes on charging CO2 into the electrolyte, with MFM-300(In)-e/In showing the strongest response to CO2 and reaching a maximum current density most rapidly. No obvious inflection point was observed for the MFM-300(In)-t/CP electrode, indicating that H2 evolution occurred initially, accompanied by CO2 reduction throughout. The electro-reduction of CO2 was conducted for 6 h at −2.15 V vs Ag/Ag+ to assess the long-term electrochemical stability of the electrodes (Figure 3e, 3f). After 6 h of electrolysis, the current density of MFM-300(In)-e/In had increased slowly to 57.6 mA cm–2 with a slightly decrease in FEHCCOH to 91.2%. The average rate of production of formic acid is estimated to be 46 mg cm–2 h–1 for MFM-300(In)-e/In under these conditions (Figure S11). In contrast, the current density increased rapidly for MFM-300(In)-t/CP from 24.1 mA cm–2 (1 h) to 52.3 mA cm–2 (6 h) and for MFM-300(In)-e/CP from 24.1 mA cm–2 (1 h) to 49.0 mA cm–2 (6 h) but with notable decreases in FEHCCOH to 52.9% and 66.8%, respectively. No carbon-containing product was detected in the gas phase product. The increased current density is assigned to H2 evolution over the extended period of electrolysis, but MFM-300(In)-e/In retains a high selectivity toward the activation of CO2 vs H2 evolution (Figure S12). A comparison of the electrolytic performance for the formation of formic acid for MFM-300(In)-based electrodes and other state-of-the-art electrodes in organic electrolytes is given in Table 1, with MFM-300(In)-e/In showing the best performance.4347

Reusability of Electrodes

All three electrodes were reused over five cycles for electro-reduction of CO2 (Figure 3g,h). Both the current density and Faradaic efficiency for formation of formic acid show excellent stability for all three electrodes. During the electrolysis over MFM-300(In)-e/In, the surface of the electrode is rendered flat as confirmed by SEM (Figures 1d and S13). PXRD studies confirm retention of the structure of MFM-300(In)-e in cycled samples after electrolysis (Figure S14). XPS was used to study the surface properties of used MFM-300(In)-t/CP and MFM-300(In)-e/In electrodes (Figure S15). During the electrolysis, both MFM-300(In)-t/CP and MFM-300(In)-e/In electrodes evolved and the change of surface structure was accompanied by shifts of the In 3d peak (−0.5 eV and +0.5 eV, respectively), moving to the energy of the as-prepared MOF sample (Table S1). Before the electrolysis, MOF particles are relatively isolated (Figure 1d), and intercrystallite charge-transfer is thus restricted, resulting in the difference observed in XPS spectra. As the MOF surface evolves during the electrolysis, the surface becomes flatter and increasingly uniform (Figure S13), and the energy difference is therefore minimized. The absence of In0 at the electrode surface suggests that In3+ sites in MFM-300(In) are not reduced during electrolysis,48 consistent with the excellent electrochemical stability of MFM-300(In)-e/In.

Mechanistic Studies

Density functional theory (DFT) calculations afford the Gibbs free energy for the electro-reduction of CO2 to formic acid over the pristine and defective MFM-300(In), representing MFM-300(In)-t and MFM-300(In)-e, respectively (Figure 4a). The formation of *COOH over pristine MFM-300(In) (* indicates an adsorption site on the MOF) involves a high energy barrier, indicating that this process is the rate-determining step. In contrast, the formation of *COOH is spontaneous over defective MFM-300(In)-e. The In3+ defects promote the adsorption of intermediate species, and therefore, higher Gibbs free energies are required for the desorption of *COOH and *HCOOH. The calculated reaction pathway of the electro-reduction of CO2 over defective MFM-300(In) is shown in Figure S16. Furthermore, the Gibbs free energy for the formation of *COOH over In2O3, a benchmark indium-based catalyst, is very high (Figure S17), indicating that defective MFM-300(In) is a better candidate for the electro-reduction of CO2. The adsorbed *COOH intermediate could be further reduced to *HCOOH or *CO.10 The calculation of Gibbs free energy of electro-reduction of CO2 to CO over defective MFM-300(In) has also been conducted, and we find that the *HCOOH intermediate is more readily formed than *CO due to the lower energy barrier for the former. This rationalizes the formation of formic acid as the main product over MFM-300(In)-based electrodes.

Figure 4.

Figure 4

(a) Gibbs free energy profile derived from DFT calculations for electro-reduction of CO2 to formic acid over pristine MFM-300(In) and defective MFM-300(In) (* indicates an adsorption site on the MOF). (b) EPR spectra of spin adducts of radicals formed during the electro-reduction of CO2 using the three electrodes (0.5 M EmimBF4/MeCN). (c) Comparison of ·DMPO–COOH radicals generated at different electrodes. (d) Tafel plots using different electrodes in this study. (e) Nyquist plots for electrodes using indium foil as substrate. (f) Nyquist plots for electrodes using CP as substrates. (g) Adsorption isotherms for CO2 in MFM-300(In)-t and MFM-300(In)-e at 298 K. (h) In situ FTIR spectra of MFM-300(In)-e as a function of CO2 loading.

We have used 5,5-dimethyl-1-pyrroline N-oxide (DMPO)49 as a spin trap to identify any radical species involved in the catalytic reaction. Figure 4b shows the EPR spectra of the reaction solution measured after electrolysis at −2.15 V vs Ag/Ag+ using the different decorated electrodes. The EPR spectrum after electrolysis using MFM-300(In)-e/In shows three sets of radicals (Figure S19): two weak signals are assigned to oxidized DMPO radical ·DMPO-OX, aN = 1.5 mT, and to ·DMPO–OH radicals, aN = 1.43 mT, aH = 1.33 mT.5052 Based upon established pathways for CO2 reduction, ·DMPO–COOH radicals are the intermediate products (Figure S16).53,54 In the above electrolysis we observe the formation of ·COOH with aH = 2.21 mT.52 The EPR spectra for the electrolysis reactions using MFM-300(In)-t/CP and MFM-300(In)-e/CP electrodes and their simulations are shown in Figures S20 and S21. The concentration of ·DMPO–COOH radicals in solution using MFM-300(In)-e/In is notably higher than in reactions at the MFM-300(In)-t/CP and MFM-300(In)-t/CP electrodes under the same conditions, consistent with the higher activity of MFM-300(In)-e/In (Figure 4c).

The Tafel plot is an indicator of the reaction pathway with the ease of reduction reflected in a lower overpotential.8 The Tafel slope for reduction of CO2 at MFM-300(In)-e/In, MFM-300(In)-t/CP and MFM-300(In)-e/CP electrodes are 138, 173, and 154 mV dec–1, respectively (Figure 4d), indicating that the initial electron-transfer to generate surface-adsorbed CO2•– species is the rate-determining step.12 Electrolysis using the MFM-300(In)-e/In electrode gives the lowest slope for the Tafel plot, indicating a higher absolute current density for a given overpotential compared with the other two electrodes. The MFM-300(In)-e/In electrode also shows more rapid kinetics for generation of radicals, consistent with the higher current density and higher FEHCOOH for the electrolysis.

Electrochemical impedance spectroscopy (EIS) was conducted to investigate the properties of the electrode/electrolyte interface (Figures 4e, 4f).43 Values for the interfacial charge-transfer resistance (Rct) were obtained by fitting the experimental impedance data using Randles’ equivalent circuit (Figure S22). The interface between MOF particles and the substrate is a significant barrier to charge transfer. The electro-synthesized MOF on In-foil gives a better interfacial contact, resulting in a lower Rct for MFM-300(In)-e/In (9.5 Ω cm2) compared with MFM-300(In)-e/CP (74.1 Ω cm2). The lower Rct for MFM-300(In)-e/CP (74.1 Ω cm2) compared to MFM-300(In)-t/CP (178.2 Ω cm2) is consistent with the presence of In3+ defect sites, which enhances the charge transfer of the former electrode. The higher double-layer capacitance value of MFM-300(In)-e/In (1.06 mF cm–2) compared with MFM-300(In)-t/CP (0.31 mF cm–2) and MFM-300(In)-e/CP (0.46 mF cm–2), as measured by CV curves at −0.6 to −0.65 V vs Ag/Ag+ at different scan rate (Figure S23), indicates that MFM-300(In)-e/In has higher electrochemical surface area.19

CO2 adsorption isotherms of desolvated MFM-300(In)-t and MFM-300(In)-e show uptakes of 87.6 and 78.3 cm3 g–1, respectively, at 1 bar and 298 K (Figure 4g). The isosteric heat (Qst) and entropies (ΔS) of adsorption (Figure S24) were determined by fitting of the Van’t Hoff isochore from the adsorption isotherms of CO2 at different temperatures (Figure S25). The value of Qst of MFM-300(In)-e is higher than that of MFM-300(In)-t suggesting a stronger interaction of CO2 with MFM-300(In)-e.

Synchrotron FTIR microspectroscopy of MFM-300(In)-e was recorded as a function of CO2 loading at 298 K (Figure 4h). The change in the stretching mode ν(μ2–OH) at 3600 cm–1 in the presence of CO2 can provide direct insights into the strength of the host–guest binding. MFM-300(In)-t exhibits a red-shift of 5 cm–1 of the ν(μ2–OH) stretching band on binding of CO2.55 In contrast, a red-shift of 25 cm–1 is observed for this stretching vibration on CO2-loading into MFM-300(In)-e under the same conditions, suggesting that MFM-300(In)-e interacts more strongly with CO2. This is consistent with the DFT calculation and Qst analysis.

Conclusions

We have described a simple and effective strategy to prepare MOF-based electrodes via rapid electro-synthesis template by an ionic liquid. The as-prepared MFM-300(In)-e/In electrode incorporates active defect In3+ sites and shows high capacity for charge transfer and high electrochemical stability. This decorated electrode catalyzes the electro-reduction of CO2 to formic acid with a current density of 46.1 mA cm–2 and a Faradaic efficiency, FEHCOOH of 99.1% in organic electrolyte. To the best of our knowledge, this performance exceeds all other MOF systems. Overall, the excess In3+ sites in MFM-300(In)-e combined with a strong contact between the MOF and the indium foil leads to reduction in interfacial resistance resulting in exceptional activity, selectivity and stability for the electro-reduction of CO2.

Acknowledgments

We thank EPSRC (EP/I011870), the Royal Society (IC170327), National Natural Science Foundation of China (21733011 and 21890761), University of Manchester and Institute of Chemistry, Chinese Academy of Sciences for funding, and EPSRC for funding of the EPSRC National EPR Facility at Manchester. This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant Agreement No. 742401, NANOCHEM). X.K. is supported by a Royal Society Newton International Fellowship. We are grateful to Diamond Light Source for access to the Beamline B22. K.H. and K.L. acknowledge the China Scholarship Council for funding.

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacs.0c05913.

  • Additional experimental details, results, Figures S1-S25, and Table S1 (PDF)

The authors declare no competing financial interest.

Supplementary Material

References

  1. Diercks C. S.; Liu Y. Z.; Cordova K. E.; Yaghi O. M. The Role of Reticular Chemistry in the Design of CO2 Reduction Catalysts. Nat. Mater. 2018, 17, 301–307. 10.1038/s41563-018-0033-5. [DOI] [PubMed] [Google Scholar]
  2. Voiry D.; Shin H. S.; Loh K. P.; Chhowalla M. Low-Dimensional Catalysts for Hydrogen Evolution and CO2 Reduction. Nat. Rev. Chem. 2018, 2, 0105. 10.1038/s41570-017-0105. [DOI] [Google Scholar]
  3. Li X.; Yu J. G.; Jaroniec M.; Chen X. B. Cocatalysts for Selective Photoreduction of CO2 into Solar Fuels. Chem. Rev. 2019, 119, 3962–4179. 10.1021/acs.chemrev.8b00400. [DOI] [PubMed] [Google Scholar]
  4. An B.; Zhang J. Z.; Cheng K.; Ji P. F.; Wang C.; Lin W. B. Confinement of Ultrasmall Cu/Znox Nanoparticles in Metal-Organic Frameworks for Selective Methanol Synthesis from Catalytic Hydrogenation of CO2. J. Am. Chem. Soc. 2017, 139, 3834–3840. 10.1021/jacs.7b00058. [DOI] [PubMed] [Google Scholar]
  5. Kattel S.; Ramírez P. J.; Chen J. G.; Rodriguez J. A.; Liu P. Active Sites for CO2 Hydrogenation to Methanol on Cu/ZnO Catalysts. Science 2017, 355, 1296–1299. 10.1126/science.aal3573. [DOI] [PubMed] [Google Scholar]
  6. Dong C. Y.; Lian C.; Hu S. C.; Deng Z. S.; Gong J. Q.; Li M. D.; Liu H. L.; Xing M. Y.; Zhang J. L. Size-Dependent Activity and Selectivity of Carbon Dioxide Photocatalytic Reduction over Platinum Nanoparticles. Nat. Commun. 2018, 9, 1252. 10.1038/s41467-018-03666-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Yu H. J.; Li J. Y.; Zhang Y. H.; Yang S. Q.; Han K. L.; Dong F.; Ma T. Y.; Huang H. W. Three-in-One Oxygen Vacancies: Whole Visible-Spectrum Absorption, Efficient Charge Separation, and Surface Site Activation for Robust CO2 Photoreduction. Angew. Chem., Int. Ed. 2019, 58, 3880–3884. 10.1002/anie.201813967. [DOI] [PubMed] [Google Scholar]
  8. Zhang L.; Zhao Z. J.; Gong J. L. Nanostructured Materials for Heterogeneous Electrocatalytic CO2 Reduction and their Related Reaction Mechanisms. Angew. Chem., Int. Ed. 2017, 56, 11326–11353. 10.1002/anie.201612214. [DOI] [PubMed] [Google Scholar]
  9. Jin H. Y.; Guo C. X.; Liu X.; Liu J. L.; Vasileff A.; Jiao Y.; Zheng Y.; Qiao S. Z. Emerging Two-Dimensional Nanomaterials for Electrocatalysis. Chem. Rev. 2018, 118, 6337–6408. 10.1021/acs.chemrev.7b00689. [DOI] [PubMed] [Google Scholar]
  10. Zhu D. D.; Liu J. L.; Qiao S. Z. Recent Advances in Inorganic Heterogeneous Electrocatalysts for Reduction of Carbon Dioxide. Adv. Mater. 2016, 28, 3423–3452. 10.1002/adma.201504766. [DOI] [PubMed] [Google Scholar]
  11. Zhu H. J.; Lu M.; Wang Y. R.; Yao S. J.; Zhang M.; Kan Y. H.; Liu J.; Chen Y. F.; Li S. L.; Lan Y. Q. Efficient Electron Transmission in Covalent Organic Framework Nanosheets for Highly Active Electrocatalytic Carbon Dioxide Reduction. Nat. Commun. 2020, 11, 497. 10.1038/s41467-019-14237-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Gu J.; Hsu C. S.; Bai L. C.; Chen H. M.; Hu X. L. Atomically Dispersed Fe3+ Sites Catalyze Efficient CO2 Electroreduction to CO. Science 2019, 364, 1091–1094. 10.1126/science.aaw7515. [DOI] [PubMed] [Google Scholar]
  13. Zheng X. L.; Ji Y. F.; Tang J.; Wang J. Y.; Liu B. F.; Steinruck H. G.; Lim K.; Li Y. Z.; Toney M. F.; Chan K.; Cui Y. Theory-Guided Sn/Cu Alloying for Efficient CO2 Electroreduction at Low Overpotentials. Nat. Catal. 2019, 2, 55–61. 10.1038/s41929-018-0200-8. [DOI] [Google Scholar]
  14. Pang Y.; Li J.; Wang Z.; Tan C.-S.; Hsieh P.-L.; Zhuang T.-T.; Liang Z.-Q.; Zou C.; Wang X.; De Luna P.; Edwards J. P.; Xu Y.; Li F.; Dinh C.-T.; Zhong M.; Lou Y.; Wu D.; Chen L.-J.; Sargent E. H.; Sinton D. Efficient Electrocatalytic Conversion of Carbon Monoxide to Propanol Using Fragmented Copper. Nat. Catal. 2019, 2, 251–258. 10.1038/s41929-019-0225-7. [DOI] [Google Scholar]
  15. Wang Y. R.; Huang Q.; He C. T.; Chen Y. F.; Liu J.; Shen F. C.; Lan Y. Q. Oriented Electron Transmission in Polyoxometalate-Metalloporphyrin Organic Framework for Highly Selective Electroreduction of CO2. Nat. Commun. 2018, 9, 4466. 10.1038/s41467-018-06938-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Gao S.; Lin Y.; Jiao X. C.; Sun Y. F.; Luo Q. Q.; Zhang W. H.; Li D. Q.; Yang J. L.; Xie Y. Partially Oxidized Atomic Cobalt Layers for Carbon Dioxide Electroreduction to Liquid Fuel. Nature 2016, 529, 68–71. 10.1038/nature16455. [DOI] [PubMed] [Google Scholar]
  17. Kwak K.; Lee D. Electrochemistry of Atomically Precise Metal Nanoclusters. Acc. Chem. Res. 2019, 52, 12–22. 10.1021/acs.accounts.8b00379. [DOI] [PubMed] [Google Scholar]
  18. Li J.; Wang X. X.; Zhao G. X.; Chen C. L.; Chai Z. F.; Alsaedi A.; Hayat T.; Wang X. K. Metal-Organic Framework-Based Materials: Superior Adsorbents for the Capture of Toxic and Radioactive Metal Ions. Chem. Soc. Rev. 2018, 47, 2322–2356. 10.1039/C7CS00543A. [DOI] [PubMed] [Google Scholar]
  19. Yang Q. H.; Xu Q.; Jiang H. L. Metal-Organic Frameworks Meet Metal Nanoparticles: Synergistic Effect for Enhanced Catalysis. Chem. Soc. Rev. 2017, 46, 4774–480. 10.1039/C6CS00724D. [DOI] [PubMed] [Google Scholar]
  20. Yu J. M.; Xie L. H.; Li J. R.; Ma Y. G.; Seminario J. M.; Balbuena P. B. CO2 Capture and Separations Using MOFs: Computational and Experimental Studies. Chem. Rev. 2017, 117, 9674–9754. 10.1021/acs.chemrev.6b00626. [DOI] [PubMed] [Google Scholar]
  21. Kang X. C.; Zhu Q. G.; Sun X. F.; Hu J. Y.; Zhang J. L.; Liu Z. M.; Han B. X. Highly Efficient Electrochemical Reduction of CO2 to CH4 in an Ionic Liquid Using a Metal-Organic Framework Cathode. Chem. Sci. 2016, 7, 266–273. 10.1039/C5SC03291A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Kornienko N.; Zhao Y.; Kley C. S.; Zhu C.; Kim D.; Lin S.; Chang C. J.; Yaghi O. M.; Yang P. Metal-Organic Frameworks for Electrocatalytic Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2015, 137, 14129–14135. 10.1021/jacs.5b08212. [DOI] [PubMed] [Google Scholar]
  23. Lin S.; Diercks C. S.; Zhang Y. B.; Kornienko N.; Nichols E. M.; Zhao Y. B.; Paris A. R.; Kim D.; Yang P.; Yaghi O. M.; Chang C. J. Covalent Organic Frameworks Comprising Cobalt Porphyrins for Catalytic CO2 Reduction in Water. Science 2015, 349, 1208–1213. 10.1126/science.aac8343. [DOI] [PubMed] [Google Scholar]
  24. Smith P. T.; Benke B. P.; Cao Z.; Kim Y.; Nichols E. M.; Kim K.; Chang C. J. Iron Porphyrins Embedded into a Supramolecular Porous Organic Cage for Electrochemical CO2 Reduction in Water. Angew. Chem., Int. Ed. 2018, 57, 9684–9688. 10.1002/anie.201803873. [DOI] [PubMed] [Google Scholar]
  25. Weng Z.; Wu Y. S.; Wang M. Y.; Jiang J. B.; Yang K.; Huo S. J.; Wang X. F.; Ma Q.; Brudvig G. W.; Batista V. S.; Liang Y. Y.; Feng Z. X.; Wang H. L. Active Sites of Copper-Complex Catalytic Materials for Electrochemical Carbon Dioxide Reduction. Nat. Commun. 2018, 9, 415. 10.1038/s41467-018-02819-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Yang H. B.; Hung S. F.; Liu S.; Yuan K. D.; Miao S.; Zhang L. P.; Huang X.; Wang H. Y.; Cai W. Z.; Chen R.; Gao J. J.; Yang X. F.; Chen W.; Huang Y. Q.; Chen H. M.; Li C. M.; Zhang T.; Liu B. Atomically Dispersed Ni(I) as the Active Site for Electrochemical CO2 Reduction. Nat. Energy 2018, 3, 140–147. 10.1038/s41560-017-0078-8. [DOI] [Google Scholar]
  27. Chaemchuen S.; Kabir N. A.; Zhou K.; Verpoort F. Metal-Organic Frameworks for Upgrading Biogas Via CO2 Adsorption to Biogas Green Energy. Chem. Soc. Rev. 2013, 42, 9304–9332. 10.1039/c3cs60244c. [DOI] [PubMed] [Google Scholar]
  28. Park J.; Yuan D. Q.; Pham K. T.; Li J. R.; Yakovenko A.; Zhou H. C. Reversible Alteration of CO2 Adsorption upon Photochemical or Thermal Treatment in a Metal-Organic Framework. J. Am. Chem. Soc. 2012, 134, 99–102. 10.1021/ja209197f. [DOI] [PubMed] [Google Scholar]
  29. Huang Q.; Li Q.; Liu J.; Wang Y. R.; Wang R.; Dong L. Z.; Xia Y. H.; Wang J. L.; Lan Y. Q. Disclosing CO2 Activation Mechanism by Hydroxyl-Induced Crystalline Structure Transformation in Electrocatalytic Process. Matter 2019, 1, 1656–1668. 10.1016/j.matt.2019.07.003. [DOI] [Google Scholar]
  30. Hod I.; Bury W.; Karlin D. M.; Deria P.; Kung C. W.; Katz M. J.; So M.; Klahr B.; Jin D. N.; Chung Y. W.; Odom T. W.; Farha O. K.; Hupp J. T. Directed Growth of Electroactive Metal-Organic Framework Thin Films Using Electrophoretic Deposition. Adv. Mater. 2014, 26, 6295–6300. 10.1002/adma.201401940. [DOI] [PubMed] [Google Scholar]
  31. Hauser J. L.; Tso M.; Fitchmun K.; Oliver S. R. J. Anodic Electrodeposition of Several Metal Organic Framework Thin Films on Indium Tin Oxide Glass. Cryst. Growth Des. 2019, 19, 2358–2365. 10.1021/acs.cgd.9b00054. [DOI] [Google Scholar]
  32. Li M. Y.; Dincă M. Selective Formation of Biphasic Thin Films of Metal-Organic Frameworks by Potential-Controlled Cathodic Electrodeposition. Chem. Sci. 2014, 5, 107–111. 10.1039/C3SC51815A. [DOI] [Google Scholar]
  33. Antonio A. M.; Rosenthal J.; Bloch E. D. Electrochemically Mediated Syntheses of Titanium(III)-Based Metal-Organic Frameworks. J. Am. Chem. Soc. 2019, 141, 11383–11387. 10.1021/jacs.9b05035. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Dong B. X.; Qian S. L.; Bu F. Y.; Wu Y. C.; Feng L. G.; Teng Y. L.; Liu W. L.; Li Z. W. Electrochemical Reduction of CO2 to CO by a Heterogeneous Catalyst of Fe-Porphyrin-Based Metal-Organic Framework. ACS Appl. Energy Mate. 2018, 1, 4662–4669. 10.1021/acsaem.8b00797. [DOI] [Google Scholar]
  35. Bohlen B.; Wastl D.; Radomski J.; Sieber V.; Vieira L. Electrochemical CO2 Reduction to Formate on Indium Catalysts Prepared by Electrodeposition in Deep Eutectic Solvents. Electrochem. Commun. 2020, 110, 106597. 10.1016/j.elecom.2019.106597. [DOI] [Google Scholar]
  36. Luo W.; Xie W.; Li M.; Zhang J.; Züttel A. 3D Hierarchical Porous Indium Catalyst for Highly Efficient Electroreduction of CO2. J. Mater. Chem. A 2019, 7, 4505–4515. 10.1039/C8TA11645H. [DOI] [Google Scholar]
  37. Ameloot R.; Stappers L.; Fransaer J.; Alaerts L.; Sels B. F.; De Vos D. E. Patterned Growth of Metal-Organic Framework Coatings by Electrochemical Synthesis. Chem. Mater. 2009, 21, 2580–2582. 10.1021/cm900069f. [DOI] [Google Scholar]
  38. Li M. Y.; Dincă M. Reductive Electrosynthesis of Crystalline Metal-Organic Frameworks. J. Am. Chem. Soc. 2011, 133, 12926–12929. 10.1021/ja2041546. [DOI] [PubMed] [Google Scholar]
  39. Martinez Joaristi A.; Juan-Alcañiz J.; Serra-Crespo P.; Kapteijn F.; Gascon J. Electrochemical Synthesis of Some Archetypical Zn2+, Cu2+, and Al3+ Metal Organic Frameworks. Cryst. Growth Des. 2012, 12, 3489–3498. 10.1021/cg300552w. [DOI] [Google Scholar]
  40. Ji H.; Hwang S.; Kim K.; Kim C.; Jeong N. C. Direct in situ Conversion of Metals into Metal-Organic Frameworks: A Strategy for the Rapid Growth of MOF Films on Metal Substrates. ACS Appl. Mater. Interfaces 2016, 8, 32414–32420. 10.1021/acsami.6b12755. [DOI] [PubMed] [Google Scholar]
  41. Qian J. J.; Jiang F. L.; Yuan D. Q.; Wu M. Y.; Zhang S. Q.; Zhang L. J.; Hong M. C. Highly Selective Carbon Dioxide Adsorption in a Water-Stable Indium-Organic Framework Material. Chem. Commun. 2012, 48, 9696–9698. 10.1039/c2cc35068h. [DOI] [PubMed] [Google Scholar]
  42. Sang X. X.; Zhang J. L.; Xiang J. F.; Cui J.; Zheng L. R.; Zhang J.; Wu Z. H.; Li Z. H.; Mo G.; Xu Y.; Song J. L.; Liu C. C.; Tan X. N.; Luo T.; Zhang B. X.; Han B. X. Ionic Liquid Accelerates the Crystallization of Zr-Based Metal-Organic Frameworks. Nat. Commun. 2017, 8, 175. 10.1038/s41467-017-00226-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Sun X. F.; Lu L.; Zhu Q. G.; Wu C. Y.; Yang D. X.; Chen C. J.; Han B. X. MoP Nanoparticles Supported on Indium-Doped Porous Carbon: Outstanding Catalysts for Highly Efficient CO2 Electroreduction. Angew. Chem., Int. Ed. 2018, 57, 2427–2431. 10.1002/anie.201712221. [DOI] [PubMed] [Google Scholar]
  44. Feng J. P.; Zeng S. J.; Liu H. Z.; Feng J. Q.; Gao H. S.; Bai L.; Dong H. F.; Zhang S. J.; Zhang X. P. Insights into Carbon Dioxide Electroreduction in Ionic Liquids: Carbon Dioxide Activation and Selectivity Tailored by Ionic Micro-Habitat. ChemSusChem 2018, 11, 3191–3197. 10.1002/cssc.201801373. [DOI] [PubMed] [Google Scholar]
  45. Zhu Q. G.; Ma J.; Kang X. C.; Sun X. F.; Liu H. Z.; Hu J. Y.; Liu Z. M.; Han B. X. Efficient Reduction of CO2 into Formic Acid on a Lead or Tin Electrode Using an Ionic Liquid Catholyte Mixture. Angew. Chem., Int. Ed. 2016, 55, 9012–9016. 10.1002/anie.201601974. [DOI] [PubMed] [Google Scholar]
  46. Wu H. R.; Song J. L.; Xie C.; Hu Y.; Ma J.; Qian Q. L.; Han B. X. Design of Naturally Derived Lead Phytate as an Electrocatalyst for Highly Efficient CO2 Reduction to Formic Acid. Green Chem. 2018, 20, 4602–4606. 10.1039/C8GC02457J. [DOI] [Google Scholar]
  47. Lu L.; Sun X. F.; Ma J.; Zhu Q. G.; Wu C. Y.; Yang D. X.; Han B. X. Selective Electroreduction of Carbon Dioxide to Formic Acid on Electrodeposited SnO2@N-Doped Porous Carbon Catalysts. Sci. China: Chem. 2018, 61, 228–235. 10.1007/s11426-017-9118-6. [DOI] [Google Scholar]
  48. Larrazábal G. O.; Martín A. J.; Mitchell S.; Hauert R.; Pérez-Ramirez J. Enhanced Reduction of CO2 to CO Over Cu-In Electrocatalysts: Catalyst Evolution Is the Key. ACS Catal. 2016, 6, 6265–6274. 10.1021/acscatal.6b02067. [DOI] [Google Scholar]
  49. Villamena F. A.; Locigno E. J.; Rockenbauer A.; Hadad C. M.; Zweier J. L. Theoretical and Experimental Studies of the Spin Trappingof Inorganic Radicals by 5,5-Dimethyl-1-Pyrroline N-Oxide (DMPO). 1. Carbon Dioxide Radical Anion. J. Phys. Chem. A 2006, 110, 13253–13258. 10.1021/jp064892m. [DOI] [PubMed] [Google Scholar]
  50. Feng G. D.; Cheng P.; Yan W. F.; Boronat M.; Li X.; Su J. H.; Wang J. Y.; Li Y.; Corma A.; Xu R. R.; Yu J. H. Accelerated Crystallization of Zeolites Via Hydroxyl Free Radicals. Science 2016, 351, 1188–1191. 10.1126/science.aaf1559. [DOI] [PubMed] [Google Scholar]
  51. Fontmorin J.M.; Burgos Castillo R.C.; Tang W.Z.; Sillanpaa M. Stability of 5,5-Dimethyl-1-Pyrroline-N-Oxide as a Spin-Trap for Quantification of Hydroxyl Radicals in Processes Based on Fenton Reaction. Water Res. 2016, 99, 24–32. 10.1016/j.watres.2016.04.053. [DOI] [PubMed] [Google Scholar]
  52. Buettner G. R. Spin Trapping: ESR Parameters of Spin Adducts. Free Radical Biol. Med. 1987, 3, 259–303. 10.1016/S0891-5849(87)80033-3. [DOI] [PubMed] [Google Scholar]
  53. Feaster J. T.; Shi C.; Cave E. R.; Hatsukade T. T.; Abram D. N.; Kuhl K. P.; Hahn C.; Norskøv J. K.; Jaramillo T. F. Understanding Selectivity for the Electrochemical Reduction of Carbon Dioxide to Formic Acid and Carbon Monoxide on Metal Electrodes. ACS Catal. 2017, 7, 4822–4827. 10.1021/acscatal.7b00687. [DOI] [Google Scholar]
  54. Sun Z. Y.; Ma T.; Tao H. C.; Fan Q.; Han B. X. Fundamentals and Challenges of Electrochemical CO2 Reduction Using Two-Dimensional Materials. Chem. 2017, 3, 560–587. 10.1016/j.chempr.2017.09.009. [DOI] [Google Scholar]
  55. Savage M.; Cheng Y.; Easun T. L.; Eyley J. E.; Argent S. P.; Warren M. R.; Lewis W.; Murray C.; Tang C. C.; Frogley M. D.; Cinque G.; Sun J.; Rudic S.; Murden R. T.; Benham M. J.; Fitch A. N.; Blake A. J.; Ramirez-Cuesta A. J.; Yang S.; Schroder M. Selective Adsorption of Sulfur Dioxide in a Robust Metal-Organic Framework Material. Adv. Mater. 2016, 28, 8705–8711. 10.1002/adma.201602338. [DOI] [PubMed] [Google Scholar]
  56. Atifi A.; Boyce D. W.; DiMeglio J. L.; Rosenthal J. Directing the Outcome of CO2 Reduction at Bismuth Cathodes Using Varied Ionic Liquid Promoters. ACS Catal. 2018, 8, 2857–2863. 10.1021/acscatal.7b03433. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Zhang X.; Zhao Y.; Hu S.; Gliege M. E.; Liu Y.; Liu R.; Scudiero L.; Hu Y.; Ha S. Electrochemical Reduction of Carbon Dioxide to Formic Acid in Ionic Liquid [Emim][N(CN)2]/Water System. Electrochim. Acta 2017, 247, 281–287. 10.1016/j.electacta.2017.06.112. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials


Articles from Journal of the American Chemical Society are provided here courtesy of American Chemical Society

RESOURCES