You are on page 1of 15

Journal of Energy Chemistry 40 (2020) 156–170

Contents lists available at ScienceDirect

Journal of Energy Chemistry


journal homepage: www.elsevier.com/locate/jechem

Review

Metal-organic frameworks for electrochemical reduction of carbon


dioxide: The role of metal centers
Ping Shao a, Luocai Yi a, Shumei Chen b, Tianhua Zhou a,∗, Jian Zhang a
a
State Key Laboratory of Structural Chemistry, Fujian Institute of Research on the Structure of Matter, Chinese Academy of Sciences, Fuzhou 350002, Fujian,
China
b
College of Chemistry, Fuzhou University, Fuzhou 350002, Fujian, China

a r t i c l e i n f o a b s t r a c t

Article history: Direct electrochemical reduction of CO2 into valuable chemicals and fuel is one of the most promising ap-
Received 16 December 2018 proaches to address the current energy crisis and lower CO2 emission. Recently, numerous metal-organic
Revised 3 April 2019
framework (MOF) and their derived materials have extensively been developed as electrocatalysts for
Accepted 8 April 2019
CO2 reduction owing to their unique structure including porosity, large specific surface area, and tunable
Available online 13 April 2019
chemical structures. In this review, the recent progress of MOF-based electrocatalysts for CO2 reduction
Keywords: was summarized and discussed. Detailed discussions mainly focus on the synthesis and mechanism of
Metal organic framework pristine MOFs and MOF-derived materials for electrocatalytic CO2 reduction. These examples are expected
Electrocatalyst to provide clues to rational design and synthesis of stable and high-performance MOFs-based electrocat-
Carbon dioxides alysts for CO2 reduction.
Reduction reaction © 2019 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published
Nanomaterials
by Elsevier B.V. and Science Press. All rights reserved.

Ping Shao received her M.S. degree from Nanchang 1. Introduction


Hangkong University in 2018. From 2016 to 2018, she
studied as a jointly student in Prof. Zhenhai Wen’s group
at the Fujian Institute of Research on the Structure of There is a growing awareness of the severe environmental prob-
Matter, Chinese Academy of Sciences. She is currently lem associated with the use of foil fuel, which has caused a
working as a research assistant in Fujian Institute of Re-
large amount of CO2 emission and global climate change [1–7].
search on the Structure of Matter, Chinese Academy of
Sciences. Her research is focused on electrochemical func- To address this issue, various greenhouse gas reduction strate-
tional materials and their application in energy storage gies are developed. Short-term greenhouse gas reduction strategies
and conversion.
are likely to improve the current energy supply, such as cogen-
eration and combined cycle gas turbines, and enhanced efficien-
cies of coal-based systems. In the long-term, however, energy ef-
Luocai Yi received his B.S. degree (2014) and M.S. de- ficiency and renewable will become a crucial part of the energy
gree (2018) in Jingdezhen Ceramic Institute and Nan-
chang Hangkong University, respectively. Since July 2016, supply and use, which will effectively reduce the CO2 emission.
he studied as a jointly student in Prof. Zhenhai Wen’s One promising measure is the development of CO2 capture and
group at the Fujian Institute of Research on the Structure conversion, including biomass, physical, and chemical solutions [8–
of Matter, Chinese Academy of Sciences. He is pursuing
his Ph.D. degree in FJIRSMCAS. His current research fo-
11]. Nevertheless, CO2 capture not only consumes a lot of energy
cuses on the design and the synthesis of nano-materials and costs, but also is easy to reach the capacity limits [5]. Accord-
and their applications, such as desalination, photothermal ingly, chemical conversion could be an ideal solution, which not
conversion, photoelectrocatalysis, and device design and
fabrication. Mechanics and electronics are his hobbies.
only reduces CO2 emission but also produces valuable chemicals
[12–15].
A promising way of atmospheric CO2 conversion is to catalyze
the chemical reduction of CO2 to useful chemicals, such as carbon
monoxide, methane, formic acid, ethane, ethylene [16–23]. In
the past decades, various CO2 reduction approaches have been

Corresponding author. developed, including electrocatalytic reduction [24–30], photo-
E-mail address: thzhou@fjirsm.ac.cn (T. Zhou). catalytic reduction [31–41], chemical and biochemical reduction

https://doi.org/10.1016/j.jechem.2019.04.013
2095-4956/© 2019 Science Press and Dalian Institute of Chemical Physics, Chinese Academy of Sciences. Published by Elsevier B.V. and Science Press. All rights reserved.
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 157

[42]. Among these methods, electrocatalytic reduction CO2 as a


promising method has attracted the wide attention of researchers,
since it can not only be conducted in ambient conditions, but
also can be driven by renewable energy (e.g., solar, wind, and
tide) [2,43]. However, extremely stable chemical structure of
CO2 molecule requires large kinetic and thermodynamic energy
barrier to be overcome to finish the reduction of CO2 . However,
chemical conversion of CO2 involves multiple-electron and proton
transfer. The process presents many complex intermediates, which
determines the reaction pathway and the selectivity as well as
Faradaic efficiency of the products [27,44,45]. The activity and
selectivity of the catalysts depend on the number of surface active
site, atom distribution, surface electronic structure, coordination
geometry, and chemical composition of catalysts [46]. Therefore,
rational design and synthesis of high-performance electrocatalysts
is of great importance for elucidating the relationship between
catalyst structure and the reaction mechanism. To this end, it
is crucial for the creation of well-defined catalysts with tunable Scheme 1. Schematic diagram of electrocatalytic reduction of CO2 .
structure.
As a class of porous materials, MOFs are constructed by self-
assembly of metal ions or clusters and organic ligand [47–55]. They 2. Mechanism of electrocatalytic reduction of CO2
can be prepared through rational design including pre-assembly
[56] and post-synthesis modification method [57,58] to realize spe- Electrocatalytic reduction of CO2 as a kind of promising way
cific functional applications [59]. Compared with homogeneous to convert CO2 into value-added fuel, has attracted the wide at-
molecular catalysts, MOF materials possess the recyclability of het- tention of researchers. Generally, the electrocatalytic reduction of
erogeneous crystalline catalysts and permanent porosity. On the CO2 is performed in aqueous electrolytes, in a two-compartment
other hand, compared with conventional inorganic porous ma- three-electrode system. As shown in Scheme 1, the reduction of
terials, MOF-based materials are well-defined crystalline mate- CO2 occurs in the cathode compartment, the anode compartment
rials with highly periodic arrangements of the organic ligands is generally coupled to the oxygen evolution reaction [62,79], and
and metal nodes. Furthermore, their physical and chemical prop- the cathode compartments are separated from the anode by Nafion
erties can be regulated by tuning the composition, metal node proton exchange membrane. The cathode compartment consists
and organic ligand at the molecular level. These advantages make of working electrode and reference electrode (Ag/AgCl), while the
them attractive porous materials for application in gas capture anode compartment is composed of counter electrode (Platinum
and separation, heterogeneous catalysis, sensing and drug deliv- plate or net).
ery [60–65]. Especially, MOF-based materials have attracted grow- Electrocatalytic reduction of CO2 is multi-electron and -proton
ing interests in the field of energy conversion and storage includ- reduction reaction involving a number of different surface reac-
ing photo- and electrocatalytic water splitting, and CO2 reduction tion intermediates. As a result, there are a large number of pos-
[36,38,59,66–70]. However, poor electronic conductivity and struc- sible reduction productions, including hydrogen, carbon monox-
ture stability of pristine MOFs to some extend hamper their di- ide, formaldehyde, formic acid, methane, methanol, and ethene,
rect application as electrocatalysts. Subsequently, the pristine MOFs through two, four, six, eight, and twelve-electron reduction in
have been used as promising precursor/templates to prepare MOF- aqueous solution. Their reduction potential reference to standard
derived porous nanomaterials including metal nanoparticle or sin- hydrogen electrode (SHE) at pH7 are shown in the following
gle atom encapsulated in-situ formed carbon hybrid and metal- Eqs. (1)–(7) [1].
free nanocarbon for electrocatalysis applications [71–74]. Recent
CO2 + e− → CO·−
2 E 0 = −1.9 V (1)
studies indicate that these MOF derivatives show obvious excel-
lent performance for application in energy conversion and stor-
age such as hydrogen evolution reaction (HER), oxygen reduc- CO2 + 2H+ + 2e− → CO + H2 O E 0 = −0.52 V (2)
tion reaction (ORR), oxygen evolution reaction (OER), CO2 reduc-
tion reaction (CO2 RR), supercapacitors, and lithium-ion Batteries
CO2 + 2H+ + 2e− → HCOOH E 0 = −0.61 V (3)
[36,59,66–69].
In this review, we have systematically summarized current
progress of MOF-based materials for electrocatalysis reduction of CO2 + 4H+ + 4e− → HCHO + H2 O E 0 = −0.51 V (4)
CO2 . Although a few of reviews have described MOFs for photo-
and electro-catalytic CO2 conversion [43–45,60–62,75–78], these
reviews mainly focused on the pristine MOFs for reduction of CO2 . CO2 + 6H+ + 6e− → CH3 OH + H2 O E 0 = −0.38 V (5)
We mainly focus on discussing the recent progresses of MOF com-
posites (e.g. carbon nitride, graphene, carbon nanotube-containing CO2 + 8H+ + 8e− → CH4 + 2H2 O E 0 = −0.24 V (6)
MOF materials) and MOF-derived metal nanoparticles as well as
single-atom catalysts for electrocatalytic reduction of CO2 . The re-
lationships between synthetic methods, different metal centers and 2CO2 + 12H+ + 12e− → C2 H4 + 4H2 O E 0 = −0.34 V (7)
the electrocatalytic performance for CO2 reduction have also been
considered. Especially, how the metal center of the MOF-based ma-
2H+ + 2e− → H2 O E 0 = −0.42 V (8)
terials influences the selectivity and activity of electrocatalytic CO2
reduction. We hope these examples can provide predictive guid- However, effective electrochemical reduction of CO2 requires
ance for the rational design and synthesis of high-performance more negative potentials compared to the equilibrium potentials.
electrocatalysts for CO2 reduction. The reasons are chiefly as follows. Firstly, as shown in Eq. (1), the
158 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

the intermediates in situ during CO2 reduction. To further under-


stand the mechanism of CO2 reduction, the design and synthesis
of electrocatalysts with high activity, high selectivity and stability
is highly desirable.

3. MOFs-based materials for electrochemical reduction of CO2

Owing to their tunable porous structure and properties, a num-


ber of MOFs and MOF-derived materials were developed as poten-
tial electrocatalysts for CO2 reduction [16,60,61,72,84–87]. To con-
struct MOF materials towards CO2 reduction, selection of suitable
metal centers is of importance. So far, numerous metal nodes are
developed as the catalytic active sites of electrocatalytic CO2 re-
duction, such as Ag, Ni, Co, or Cu [61,71,73,88]. For example, Fe, Co,
Zn and noble-metal based MOFs materials have been used as elec-
trocatalysts for CO2 conversion to CO [73,87,89], whereas Cu based
MOFs materials have also been shown for electrochemistry CO2 re-
duction to other hydrocarbons such as C2 and C3 [86]. The syn-
thetic methods and chemical composition of MOFs-based catalysts
for electrochemistry CO2 reduction have been listed in Table 1.

3.1. MOF materials

To prepare high-performance MOF-based materials for electro-


catalytic reduction of CO2 , various strategies have been adopted
to improve their conductivity, increase the number of exposed
surface-active sites, and enhance the stability. One of these strate-
gies is to build MOF-composites as electrocatalysts by in situ
Fig. 1. Possible mechanism of electrocatalytic reduction of CO2 on transition-metals
and molecular catalysts. Reproduced with permission [83]. Copyright 2015, ACS. growth of MOFs on conductive substrates such as carbon pa-
per, carbon nanotube (CNT), graphene, or modified Al2 O3 nanorod
[89,90,93,99,107]. The strong interactions between MOFs and sub-
CO2 molecules are adsorbed onto the surface of the catalyst and strates contribute to increasing the number of exposed active sites
transformed into CO·− 2 by electron transfer, which require a large and facilitating the charge transfer from the electrode to catalysts.
negative potential to break down the stable linear structure of CO2 As a result, the composites exhibit high activity and current den-
molecules. Secondly, CO·− 2 is further converted into different re- sity for electrocatalytic reduction of CO2 . Another strategy is to fab-
action intermediates at the surface of the catalyst and desorbed ricate MOF-derived electrocatalysts by means of direct pyrolysis of
from the surface of the catalyst (Fig. 1). During this process, be- MOF and MOF-based composites [17,111]. The strategy contributes
sides overcoming complex kinetics obstacles, the electrolyte sys- to the formation of homogeneously distributed metal nanoparticles
tem of aqueous solution in the reactor also needs to consume or/and singe atom throughout the whole skeleton, together with
energy to facilitate the mass and electron transport. Thirdly, the abundant defects within graphitic carbon matrix, which promote
electrolytes for most of the CO2 reduction reactions are aqueous the contact between dissolved CO2 and active sites [112–114].
solution, which makes hydrogen evolution reaction (HER) a com-
petitive reduction reaction (Eq. (8)).
3.1.1. Iron-based MOF
In fact, there are different theoretic mechanisms for various
Iron (Fe) can form accessible reduced or oxidized states. Owing
products (C1, C2, C3,…) generated from the CO2 RR. For example,
to the good chemical stability with Fe(III)-based MOF, they are
the reduction of CO2 to CO proceeds via two-electron transfer pro-
potential candidates for heterogeneous catalysis. So far, many
cess [80]. Firstly, the first electron is transferred to the surface
Fe-based MOFs have been used as photocatalyst for water splitting
of CO2 molecule to form CO·− ·−
2 radical; then the CO2 radical ab- or dye degradation [36,59,67]. Considering their Lewis acidity
sorbed on surface of catalysts leads to extra negative charge which
and a redox activity [115], Fe-based MOFs are also considered
promote protonation to form ∗ COOH intermediate (Eq. (9)). Sub-
as potential electrocatalysts for CO2 conversion. For example,
sequently, further proton-coupled-electron-transfer reactions with
∗ COOH intermediate form CO and one H O molecule; the CO Hod and co-workers reported the electrophoretic deposition of
2
Fe_MOF-525 thin film on fluorine doped tin oxide (FTO) glass
molecules desorb from the surface of catalysts (Eq. (10)), ∗ denote
[93]. Through anchoring homogenous catalyst Fe-porphyrins on
surface-bound sites.
MOF-525, Fe_MOF-525 presented large accessible active sites
CO·− +
2 + H → ∗COOH (9) (∼1015 sites/cm2 ). The MOF can electrocatalytic CO2 reduction
to the production of CO and H2 in closely equal amount, with
∼100% Faradaic efficiency at a CO2 /CO overpotential of about
∗COOH + e− + H+ → CO + H2 O (10)
650 mV. The Faradaic efficiencies for CO and H2 are 54% and 45%
However, to produce multi-carbon products, the catalyst surface without 2,2,2-trifluoroethanol (TFE), respectively. On the contrary,
requires strong adsorption capacity for the CO intermediate dur- Fe_MOF-525 exhibits slightly improved Faradaic efficiency for
ing CO2 reduction, therefore, it promotes CO further dimerization CO (60%) in the presence of TFE (Fig. 2). They revealed that the
to generated multi-carbon species via multiple proton-coupled- Fe(0)-porphyrin species are catalytic active sites for CO2 reduction.
electron-transfer reactions (Fig. 1) through different reaction in- The work provides an effective strategy to construct Fe-based
termediates formed on the surface of different electrocatalysts MOFs electrocatalysts for CO2 reduction by immobilizing molecu-
[81–83]. Up to date, there is no advanced technology for detecting lar catalysts on MOF. This strategy not only avoids the aggregation
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 159

Table 1. Electrocatalytic CO2 reduction on MOF-based materials.

Synthetic method Electrode Potential Products FE (%) Ref.

Electrophoretic deposition Zn-BTC MOF/CP −2.2 V vs. Ag/AgCl CH4 80.1 ± 6.6 [90]
CO 7.9 ± 2.6
Vigorous stirring ZIF–8 −1.8 V vs. SCE CO 65 [91]
Liquid-phase epitaxy Re-SURMOF/FTO −1.6 V vs. NHE CO 93 ± 5 [92]
Electrophoretic deposition Fe-MOF-525/FTO – CO 41 ± 8 [93]
H2 60 ± 4
Refluxed mbpyOH-[IrIII ]-UiO – formate TON:543 ± 35 [94]
Heat treatment Ni-containing cyclic porphyrin dimer (Ni2 -CPDPy ) −0.8 V vs. RHE CO 94 [95]
Pyrolysis Tris(2-benzimidazolyl-methyl)amine (NTB) ligand −0.8 V vs. RHE CO 97 [96]
directed synthesis Ni-N-RGO
Pyrolysis ZIF-8 derived Fe-N-C −0.6 V vs. RHE CO 91 [87]
Pyrolysis ZIF-8 derived −1.0 V vs. RHE CO 71.9 [97]
Ni SAs/N-C
Refluxed at high temperature Ag2 O/layered ZIF −1.2 V vs. RHE CO 80.5 [73]
Pyrolysis ZIF-Fe-CNT-FA-p –0.86 V vs. RHE CO 100 [98]
Pyrolysis C-AFC©ZIF-8 −0.43 V vs. RHE CO 93 [72]
Microwave Al2 (OH)2 TCPP-Co −0.7 V vs. RHE CO 76 [89]
Sonication and magnetic stirring CoPc/CNT −0.63 V vs. RHE CO 92 [99]
Hydrothermal M-PMOF −0.8 V vs. RHE CO 98.7 [100]
Pyrolysis Co/Zn ZIFs derived Co-N2 overpotential 520 mV CO 94 [88]
Pyrolysis CoPc derived Co−N5 /HNPCSs −0.73 V vs. RHE CO 99.2 [101]
Refluxed CoPP@CNT −0.49 V vs. RHE CO 98.3 [102]
— CoPP-PG −0.5 V vs. RHE CO — [85]
CH3 OH
CH4
Refluxed and sonication CCG/CoPc-A −0.59 V vs. RHE CO 75 [103]
Solution self-assembly Copper rubeanate MOF −1.2 V vs. SHE HCOOH 65 [104]
Electrochemistry Cu3 (BTC)2 – oxalic acid 51 [105]
Solvent-free synthetic Cu-based HKUST-1 10 mA.cm−2 CH3 OH 5.6 [84]
C2 H5 OH 10.3
Recrystallization crystallized CuPc −1.6 V vs.Ag/AgCl C2 H4 25 [106]
Thermal treatment HKUST-1 mediate Cu −1.07 V vs. RHE C2 H4 45 [86]
Ultrasonic disperse GDE-Cu-MOF −1.8 V vs. SCE C2 H4 16 [107]
Carbonization OD Cu/C-10 0 0 (Cu-based HKUST-1) −0.1 V vs. RHE CH3 OH 45.2 [108]
C2 H5 OH
Commercial CuPc −1.06 V vs. RHE CH4 66 [109]
Refluxed PorCu −0.976 V vs. RHE CH4 44 [110]
C2 H 4
Zn-BTC MOF/CP: Zn-1,3,5-benzenetricarboxylic acid metal-organic frameworks deposited on carbon paper; TCPP-H2 : 4,4 ,4 ,4-(porphyrin-5,10,15,20-tetrayl)tetrabenzoate;
M-PMOF: polyoxometalate-metalloporphyrin organic frameworks; HNPCSs: hollow N-doped porous carbon spheres; GDE-Cu-MOF: Cu3 (BTC)2 (Cu-MOF) was introduced into
a carbon paper based gas diffusion electrode (GDE); ZIF-Fe-CNT-FA-p: pyrolyzed the Fe-Zn bimetallic ZIF-8 on MWCNTs hydirds were loaded with furfuryl alcohol (FA); C-
AFC©ZIF-8: high temperature carbonization surface functionalization of ZIF-8 mixed with ammonium ferric citrate after vigorously stirring; PorCu: cop-per(II)−5,10,15,20-
tetrakis(2,6-dihydroxyphenyl)-porphyrin; CoPP-PG: Co protoporphyrin-coated pyrolytic graphite; CCG/CoPc-A: cobalt (II) octaalkoxyphthalocyanine was immobilized on
chemically converted graphene via π -π stacking; CuPc/C: copper (II)−5,10,15,20-tetrakis(2,6-dihydroxyphenyl) porphyrin supported on carbon black.

and deactivation process of homogeneous molecular catalysts, but ciently electrochemistry converting CO2 into CO. As an example,
also increases the number of catalytic active sites. Yang and co-workers in situ fabricated a kind of atomically de-
In order to further improve the performance of Fe-based MOF, fined and nanosized cobalt-porphyrin MOFs [89], Al2 (OH)2 TCPP-Co
introduction of carbon black into porphyrin-based MOF fabricated (TCPP-H2 = 4,4 ,4 ’,4 ’’-(porphyrin-5,10,15,20-tetrayl) tetrabenzoate)
the composite catalysts PCN-222(Fe)/C. Dong and co-workers ob- atomic layer deposition (ALD) method in microwave reactor. Co-
served that the composite catalysts (mass ratio = 1:2) on carbon porphyrin MOFs with thickness optimization displayed the excel-
paper electrode exhibited high catalytic activity for electrocatalytic lent activity and selectivity of CO about 76% and retained stabil-
CO2 reduction to CO with the overpotential of 494 mV [116]. The ity exceed 7 h in controlled potential electrolysis at –0.7 V versus
maximum Faradaic efficiency is 91% toward CO in a CO2 -saturated RHE in CO2 -saturation buffer. In the low-overpotential region, the
0.5 M KHCO3 solution. The high performance of the composite MOF composite showed a Tafel slope of 165 mV/decade, implying
PCN-222(Fe)/C could be attributed to the synergistic effects of the that one-electron reduction of CO2 to form the CO2  radical could
single active sites of Fe-porphyrin molecule, the high CO2 adsorp- be the rate-determining step. To gain deeper insight into the re-
tion capacity of porous structural MOF, and the improved conduc- action process, in situ spectroelectrochemical measurements were
tivity from introduced carbon black. Furthermore, the catalyst ex- performed on the as-prepared sample for a series of applied poten-
hibits high stability after the reaction for 10 h at –0.60 V versus tials and revealed the reduction of CO2 absorbed on a Co porphyrin
RHE. to CO probably proceeded via Co(I) with a one-electron reduction
(Fig. 3). The work suggested that the well-defined MOFs with tun-
3.1.2. Cobalt-based MOFs able structure provided a good platform to uncover the reaction
Like iron, cobalt also exhibits several oxidation states, such mechanism of CO2 reduction. The activity of Co-based MOFs can
as Co(II), Co(III), and Co(IV). It offers a wide field for the cre- not only be regulated by optimize the catalytic active sites, but also
ation of Co-based MOF with various coordination geometries in- be promoted by modulation of metal nodes.
cluding square planar, tetrahedral, pyramidal, and octahedral. As More recently, Lan and co-workers prepared polyoxometalate
above-mentioned, to achieve electrochemical reduction of CO2 , (POM)-metalloporphyrin organic frameworks (M-PMOF) with hy-
several examples of immobilized cobalt-based MOFs thin film drothermal method [100]. They tested the electrocatalytic proper-
on conductive substrate have been built. They displayed effi- ties toward the CO2 RR in CO2 -saturated 0.5 M KHCO3 electrolyte
160 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

Fig. 2. (a) The crystal structure of MOF-525 and the chemical structure of metal node; (b) The PXRD patterns of MOF-525 and Fe_MOF-525 film; (c) Current density versus
time for Fe_MOF-525 under different conditions; (d) Faradaic efficiency over about 4 h of electrolysis. Reproduced with permission [93]. Copyright 2015, ACS.

Fig. 3. (a) Molecular structure of the MOF catalyst; (b) The organic building units of the 3D MOF Al2 (OH)2 TCPP-Co. Color code: Co, orange; O, red; C, black; N, blue;
Al, light-blue octahedral; and pyrrole ring, blue; (c) The MOF is integrated with a conductive substrate to build electrochemistry system for reduction of CO2 ; (d) The
product selectivity at the potential range of −0.5–0.9 V vs. RHE; (e) The long-term stability test of the MOF catalyst and the corresponding faradaic efficiency measurement.
Reproduced with permission [89]. Copyright 2015, ACS.
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 161

Fig. 4. (a) The free energy diagrams of CO2 reduction to CO for different catalysts, ∗ denotes a surface active site; (b) The free energy for ∗ COOH formation (G1 ), ∗ CO
formation (G2 ), and CO desorption process (G3 ) during CO2 RR; (c,d) The proposed mechanism for the CO2 RR over Co-PMOF catalysts (POM represents electron-donating
species, porphyrin is charge transfer ligand and transition metals represent electron-collecting species). Reproduced with permission [100]. Copyright 2018, springer.

and observed that Co-PMOF exhibited remarkable faradaic effi- trode (1.1 μmol/cm2 /h) at a low potential of –1.2 V versus SHE. The
ciency (99%) of CO production. The authors attributed the excel- high selectivity of CR-MOF could be attributed to the lower density
lent electrocatalytic performance of Co-PMOF electrocatalysts to of electrons on the metallic sites. This structure results in weaker
the synergistic effects of the POM and Co-porphyrin in the CO2 re- adsorption of CO2 to the CR-MOF, compared with Cu metal elec-
duction reaction. Further DFT calculations demonstrated Co-PMOF trode (Fig. 5c). The weak adsorption contributes to the production
displayed lower adsorbed free energies for the intermediates of of HCOOH and thus enhances the selectivity of CR-MOF electrode
∗ COOH and ∗ CO in rate-determining steps (RDS) for CO reduction
2 toward HCOOH. However, it remains unclear whether the porous
(Fig. 4a,b). The cobalt in Co-porphyrin was proposed as a favorable structure affects the activity or/and selectivity of CO2 reduction.
active site whereas POM served as electron-rich aggregates and ef- Besides the structure of MOF, the solvent also influences the ac-
ficiently promote electronic transport. tivity and selectivity of CO2 reduction. For instance, Kumar and co-
workers investigated the electrocatalytic CO2 reduction of a water
3.1.3. Copper-based MOF stable and porous copper MOF Cu3 (BTC)2 with tetra-coordinated
Cu-MOF materials were first used as electrocatalysts for re- Cu site in N,N-dimethylformamide (DMF) containing tetrabutylam-
duction of CO2 in 2012, Hinogami and co-workers fabricated a monium tetrafluoroborate (TBATFB) [105]. They demonstrated that
copper rubeanate MOF (CR-MOF) by mixing the ethanol solution Cu3 (BTC)2 on glass carbon electrode can effectively catalyze the re-
of rubeanic acid and the aqueous solution of CuSO4 [104]. The duction of CO2 to oxalic acid with 90% purity and the Faradaic ef-
authors found that the CR-MOF electrode on carbon paper (CP) ficiency of 51%. The enhanced activity was attributed to the Cu(I)
showed a highly selective CO2 reduction to HCOOH in 0.5 M KHCO3 formed adduct with CO2 which promoted the production of ox-
solution with a current efficiency of 30% at every potential and alic acid (Fig. 6). They proposed that the reaction proceeds through
a HCOOH selectivity of above 98%. Compared with the Cu metal two-electron reduction and dimerization process which abstracts
electrode (Fig. 5a), the onset potential of CR-MOF electrode was proton from non-aqueous solvent DMF to form the oxalic acid.
about 0.2 V more positive than the Cu metal electrode under the However, investigations of the influence of solvent and ionic liq-
same conditions. The amount of HCOOH on CR-MOF electrode uid on the activity and selectivity of Cu-MOF are further needed.
(13.4 μmol/cm2 /h) is 13-fold greater than that on Cu metal elec- The solvent and electrolytes could affect not only the reaction
162 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

Fig. 5. The cyclic voltammetry curves for (1) CR-MOF, (2) bare CP, and (3) Cu metal electrode in CO2 -saturated (a) and N2 -saturated (b) 0.5 M KHCO3 electrolyte; (c) At
different potentials, the production distributions of Cu metal and CR-MOF electrodes. Reproduced with permission [104]. Copyright 2012, ECS.

pathways and the final products, but also the CO2 adsorption ca- generate a CO molecule which can be further reduced to CH4 by
pacity of catalysts. Enhancing the CO2 concentration around active six electrons due to the stronger absorption capacity of CO than
sites can help to enable the reactants and products to diffuse effi- CH4 on the surface of Zn-MOF. The synergistic effect of ILs and
ciently, facilitating the CO2 reduction process. Zn-MOF plays a crucial role in highly selective catalyzing CO2
reduction to CH4 . In the absence of Zn-MOF, silver electrode in ILs
3.1.4. Zinc-based MOF mainly catalyzes reduction of CO2 to CO [117].
Zinc (II) with an electronic configuration of [Ar]3d10 has a com- In addition to organic electrolyte, it was found that inorganic
pletely filled shell orbital and is a borderline hard acid. It readily counterions also affected the activity and selectivity of CO2 re-
coordinates with the oxygen, nitrogen or sulfur atoms without a duction on Zn-MOF electrode. Wang and co-workers synthesized
strong preference for organic ligands. Hence, zinc-based metal- a series of Zn-based zeolitic imidazolate framework (ZIF-8) nano-
organic frameworks (MOFs) with different coordination numbers materials by introduction of different zinc sources as reactant and
from 4 to 6. Zn-based MOFs have extensively been investigated investigated the effects of different zinc sources and electrolytes
because of their attractive properties in many potential applica- on the electrocatalytic CO2 reduction properties in aqueous so-
tions such as gas adsorption and separation, sensing, catalysis, lution [91]. They observed that ZIF-8 synthesized using ZnSO4
and CO2 capture. Recently, Zn-based MOFs have also been used presented the highest catalytic activity towards CO2 reduction to
as catalysts for electrocatalytic CO2 reduction. Han and co-workers CO with Faradaic efficiency of 65% at –1.8 V versus Hg/Hg2 Cl2 (SCE)
investigated the CO2 reduction properties of experimental Zn- (Fig. 8a). The high activity and selectivity of ZIF-8 having SO4 2−
1,3,5-benzenetricarboxylic acid (Zn-BTC MOFs) ionic liquids (ILs) can be attributed to the less interactions between the SO4 2− and
electrolytes and revealed that the morphology of MOFs and the Zn nodes, which contributes to facilitating the anion exchange for
type of electrodes influences the current density and selectivity to charge balance when the oxidation state of Zn node is changed.
CH4 . Furthermore, they systematically studied the effect of ILs on Meanwhile, a decrease in current density was observed with
the CO2 reduction performance [90]. For imidazolium-based ionic reducing the hydrated anionic radius of the electrolyte (NaCl),
liquid, the performance depended on the type of counterion and along with an increase in faradaic efficiency toward CO2 reduction
viscosity of ionic liquid. They proposed that fluorine-containing to CO (Fig. 8b,c). It was proposed that the small hydrated anionic
ionic liquid showed high current density due to its strong inter- radius contributed to the anion exchange for charge balance and
actions with CO2 adsorbed (Fig. 7a, b). The work demonstrated depressed the hydrogen production. Cyclic voltammetry (CV) re-
that the structure and property of the electrolytes affected the vealed that ZIF-8 showed a redox peak corresponding to the redox
reaction pathway as well as the activity and selectivity of the final reaction of Zn center (ZnII /ZnI ) (Fig. 8d). The metal nodes of ZIF–8
product. The Zn-MOF plays a role in facilitating electron transfer were thus regarded as the electrocatalytic active centers. Notably,
during CO2 reduction. As Fig. 7(c) shown, during electrolysis this assignment should be carefully confirmed because Zn(II) being
the imidazolium cations were firstly absorbed on the surface a d10 system has no tendency to enable redox reactions under the
of Zn-MOF. Then one electron was transferred to CO2 molecule current reaction condition. However, the work suggested that with
captured by the ILs via conductive Zn-MOF substrates and formed the use of a suitable solvent, the construction of electrocatalytic
CO·−2 intermediates. The intermediates take another electron and CO2 reduction on Zn-MOFs was demonstrated to be possible.
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 163

Fig. 6. (a) Cyclic voltammogram curves of Cu3 (BTC)2 coated GC (black) and GC background (red line); Inset: Cu electrode in 0.1 M KCl solution. (b) CV curves of (1) bare
GC, (2) bare GC in presence of CO2 , (3) Cu3 (BTC)2 coated GC and (4) Cu3 (BTC)2 coated GC in CO2 saturated 0.01 M TBATFB/DMF electrolyte; (c) FT-IR spectra of oxalic acid
(I, authentic) and oxalic acid (II, synthesized); (d) GC–MS spectrum verifying the product of electrocatalytic reduction of CO2 over Cu3 (BTC)2 electrode. Reproduced with
permission [105]. Copyright 2012, Elsevier.

3.1.5. Noble-metal-based MOF demonstrate that the orientation of SURMOFs is in favor of charge
Like porphyrin-based catalysts, noble-metal-based molecular transport between the interface of electrode and the Re-based re-
catalysts displayed high activity and selectivity but they are un- dox catalyst (Fig. 9b).
stable. Thus, to further enhance the catalytic performance, in gen- So far, considerable silver-based catalysts have been inves-
eral the grafting of molecular catalyst on an electrode is required. tigated as electrocatalysts for reducing CO2 to CO due to the
For example, it incorporated a molecular catalyst (Fig. 9a) [92], weak adsorption of CO on the surface of silver catalyst. However,
ReL(CO)3 Cl (L = 2,2 -bipyridine-5,5 -dicarboxylic acid), into highly these nanoparticles easily aggregate in large nanoparticles, leading
oriented SURMOF thin films grown on FTO along the [001] di- to reducing the accessible active sites. To further improve the
rection through liquid-phase epitaxy. The Re-SURMOF-based elec- performance of silver-based nanomaterials, various strategies were
trocatalysts display excellent electrocatalytic performance with the adopted such as tuning the size and shape of nanoparticles. MOFs
high Faradic efficiency of 93% ± 5% for electrocatalytic reducing as a porous material with high specific surface area provide a
CO2 to CO. The reduction mechanism was proposed to be two- better platform to construct functional materials, and act as not
electron pathway including the reduction of the Re+ center of only a support for immobilization noble-metal molecular catalysts
the [Re(bpy)(CO)3]− to Re0 . The superior performance of Re-based but also a template which promotes the formation of uniform
MOF was attributed to the high oriented Re-SURMOF with the uni- distributed small noble-metal nanoparticles for CO2 reduction.
form thickness and a low defect density. Theoretical calculations Jiang and co-authors in-situ fabricated Ag2 O/zeolitic imidazolate
164 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

Fig. 7. CV curves (a) and chronoamperometry curves (b) at –2.2 V vs. Ag/Ag+ in different ILs: (1) BmimBF4 ; (2) BmimOTf; (3) BmimPF6 ; (4) BmimClO4 ; (c) The proposed
mechanism of electrochemical conversion CO2 into CH4 for Zn-MOF/CP electrocatalyst. Reproduced with permission [90]. Copyright 2017, Wiley.

framework (ZIF) composites via one-step hydrothermal treat- cated by high temperature treatment. These MOFs-derived carbon
ment of ZIF-7 in AgNO3 aqueous solution (Fig. 10a) [73]. It was based materials including metal nanoparticles or single-atom
found that small Ag2 O nanoparticles were homogenously dis- embedded in in-situ formed graphitic carbon materials exhibited
tributed on ZIF layer with large specific surface area. It was found distinct advantages for electrocatalysis reduction of CO2 , such as
that Ag2 O/ZIF showed more positive onset potential of –0.6 V good electronic conductivity, large specific surface area, uniform
versus RHE for CO2 reduction than that on the Ag/ZIF-8 in CO2 - distributed nanoparticles [7,59,67–69,71–74,88,101,108,119,120],
saturated 0.25 M K2 SO4 solution. Compared with ZIF or Ag/C alone, which affected the activity and selectivity of electrochemical CO2
Ag2 O/lZIF-8 exhibited the higher CO faradaic efficiency of 80.6% reduction.
at –1.2 V versus RHE and larger current density. The excellent
performance was considered as the synergistic effect of the small
3.2.1. MOF-derived nanoparticles as electrocatalysts
size of nanoparticles and high specific surface area (Fig. 10b,c).
MOF derived metal nanoparticles have been developed as
oxygen reduction reaction, oxygen evolution reaction, hydrogen
3.2. MOF-derived materials evolution reaction, and CO2 reduction catalysts. Their catalytic per-
formances to some extent depend on the structure of nanoparti-
MOF derivatives materials have also attracted increas- cles [59,75]. For example, Guo and co-workers fabricated ZIF-8 de-
ing interest due to the promising electrocatalytic properties rived hybrids materials [98], Pyrolysis of ZIF-8/multi-walled car-
[16,89,93,95,118]. Previous reports have pointed out that the metal bon nanotubes (MWCNTs) composites loaded with furfuryl alco-
nodes of MOFs materials act as the active centers for electrocat- hol (FA) (ZIF-CNT-FA-p) acquired metal-free porous carbon nano-
alytic reduction of CO2 , and the electrochemistry performance of materials. This material displayed high Faradaic efficiency of 100%
MOFs can be regulated by tailoring the structure of catalytic active and current density of 7.7 mA cm−2 at an overpotential of 740 mV.
sites and organic ligands. As to the structure, pristine MOFs are a To further improve the performance, iron was introduced into
kind of crystalline materials that consist of flexible metal ions or the ZIF-CNT-FA-p composites to form ZIF-Fe-CNT-FA-p. Compare
clusters and organic ligands, which with principle advantages of to ZIF-CNT-FA-p, ZIF-Fe-CNT-FA-p demonstrated lower overpoten-
controllable pore size distribution and high specific surface area. tial of 440 mV at a current density of 2 mA cm−2 and a higher CO
However, the limited conductivity and insufficient structure stabil- Faradaic efficiency of 97% (Fig. 11a–c). The efficient catalytic per-
ity of the pristine MOFs inhibit the development and utilization in formance is regarded as the synergistic effect of the intrinsic ac-
electrocatalytic CO2 reduction. Subsequently, MOF-derived materi- tive sites consisting of the Fe-Nx and high content of pyridinic-
als have been developed as electrocatalysts for reduction of CO2 to N in the hybrid as well as the improved electron and mass
overcome the drawbacks of pristine MOFs, which could be fabri- transport derived from the introduced CNT supports (Fig. 11d,e).
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 165

Fig. 8. (a) Faradaic efficiencies of ZIF–8SO4 catalysts; (b) CV curves for ZIF–8SO4 catalysts in CO2 -saturated 0.5 M NaHCO3 , NaCl and NaClO4 electrolyte and (c) the correspond-
ing faradaic efficiencies at different applied potentials; (d) CV curves of ZIF–8SO4 catalysts in Ar and CO2 -saturated 0.5 M NaCl electrolyte. Reproduced with permission [91].
Copyright 2017, Wiley.

of 190 mV versus RHE, and long-term durability in 0.5 M NaHCO3


aqueous solution. By contrast, the Fe-based materials mainly acted
as HER catalysts. A combination of X-ray absorption spectroscopy,
electrochemical measurements, and 57 Fe Mössbauer spectroscopy
revealed that the selectivity and activity of Fe-N-C toward electro-
chemical reduction of CO2 to CO can be regulated by changing the
ratio of isolated FeN4 and Fe-based nanoparticles.
A large number of MOF-derived meta-nitrogen doping carbon
hybrids have been investigated for electrocatalytic reduction of
Fig. 9. (a) Illustration of the fabrication process of Re-SURMOF on modified FTO
substrate; (b) The schematic diagram of charge transfer route of epitaxial Re- CO2 . However, they produce majorly C1 products such as CO,
SURMOF on FTO substrate along [001] direction. Reproduced with permission [92]. HCOOH. Correspondingly, only a few of MOFs are capable of cat-
Copyright 2016, RSC. alyzing reduction of CO2 to multi-carbon products. By contrast,
copper-based catalysts are a promising metal material which can
further convert CO intermediate into multi-carbon product [7]. For
However, numerous studies have shown that this type of Fe-N-C instance, Zhao and co-worker prepared oxide-derived Cu/carbon
materials are often inhomogeneous and contain a large amount of (OD Cu/C) catalysts via a facile carbonization of Cu-based MOF
aggregated Fe-based nanoparticles along with atomically-dispersed (HKUST-1). They found that OD Cu/C effectively converted CO2 to
single-atom Fe components. This complicated material leads to the methanol and ethanol at lower potential, and proposed the ex-
difficulty of identifying the active sites during catalysis. Recently, traordinary performance derived from the coupled interactions be-
Fontecave and co-workers have synthesized a series of Fe-N-C hy- tween the surficial copper oxides/copper and the matrix of con-
brids having a variable ratio of nanoparticles and single-Fe atom ductive porous carbon [108]. The in situ infrared reflectance (IR)
to uncover the relationship of the structure and selectivity of the spectroscopy and theoretical calculations revealed that the ∗ CO
electrocatalytic CO2 reduction [87]. They found that the materi- intermediate was adsorbed on the surface of OD Cu/C catalysts,
als containing sole Fe-N4 moieties displayed highest activity and and then porous carbon supports promoted C–C coupling which
selectivity toward electrocatalytic CO2 reduction to CO, producing is a crucial process in the formation of C2 species (Fig. 12).
CO with the Faradaic efficiency of above 90%, low overpotential In addition, Nam and co-worker also reported another HKUST-1
166 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

Fig. 10. (a) The schematic diagram of the construction of Ag2 O/layered ZIF catalysts; (b) the faradaic efficiency and (c) current density for CO production over different
catalysts at different applied potentials. Reproduced with permission [73]. Copyright 2017, RSC.

(MOF)-regulated Cu cluster for electrocatalytic reduction CO2 to


C2 H4 [86]. They demonstrated that Cu-based catalysts with a low
surface Cu coordination number can promote the high selectivity
for C2 hydrocarbons, since MOF-derived undercoordinated Cu sites
are more active for C–C coupling. Therefore, these findings may
provide guideline for the design and synthesis of potential catalysts
towards electrochemical reduction CO2 to multi-carbon products.

3.2.2. MOF-derived single-atom catalysts


So far, various MOFs have been adopted as precursors to fabri-
cate in-situ formed graphitic carbon encapsulated metal nanopar-
ticles as CO2 reduction electrocatalysts [75]. In general, direct
pyrolysis of MOF precursors readily forms aggregated metal
nanoparticles which show low exposed active sites. To expose the
active sites as much as possible, the single atom metal anchored
on the carbon matrix by pyrolysis of MOF precursors has been
identified as a promising electrocatalyst toward CO2 reduction, due
to the high surface area, high conductivity, and the possibility
of introducing different heteroatoms (N, P, S and B). So far, vari-
ous strategies have been adopted to prepare single-atom catalysts
by pyrolysis of MOF precursor. Zhao and co-workers reported a
new strategy to construct single-atom Ni-based catalysts by in-
troduction of Zn into metal nodes to form bimetal MOFs by ex-
changing partial Zn nodes with adsorbed Ni ions within the cav-
ity of the ZIF-8 [97]. The strategy avoids the aggregation of Ni
species in the hybrid. Finally, a Ni single atom catalyst encapsu-
lated in nitrogen-doped porous carbons (Ni SAs/N–C) was synthe-
sized through pyrolysis of the bimetal ZIF-9 for efficient electrore-
duction of CO2 (Fig. 13a). Electrochemical CO2 reduction exper-
iments revealed that Ni SAs/N–C exhibited an excellent electro- Fig. 11. (a) Illustration of the MWCNT support improved the interparticle con-
catalytic performance with the current density of 10.48 mA cm−2 ductivity and mass transport for pyrolyzed ZIFs towards electrochemical reduc-
tion of CO2 . FEs for CO (b) and H2 (c) over different catalysts at different poten-
at the overpotential of 0.89 V and the Faradaic efficiency of over
tials, respectively. LSV curves of (d) ZIF-CNT-FA-p and (e) ZIF-Fe-CNT-FA-p in CO2 -
71.9% for CO production (Fig. 13h). This designed single-atom cata- saturated 0.1 M NaHCO3 (dashed lines) and N2 -saturated 0.1 M NaH2 PO4 /Na2 HPO4
lysts facilitate the mass and electron transfer, because of the large (solid lines). Reproduced with permission [98]. Copyright 2017, RSC.
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 167

number of exposed reactive active sites. The high performance was


considered as the synergistic effect of the increased number of
surface-active sites, the strong bonding with CO·− 2 intermediates
on single Ni sites, and better electrical conductivity. However, the
Ni SAs/N–C catalyst displayed a competitive activity toward hydro-
gen evolution reaction (HER), resulting in unexpected Faradaic ef-
ficiency.
In order to further improve the performance of ZIF-8 for CO2 re-
duction, Wang and co-workers developed another strategy to pre-
pare single-atom catalysts with high performance by means of reg-
ulating the coordination number of single-atom cobalt [88]. They
found that the pyrolysis temperature can effectively affect the co-
ordination number of Co atom. During the pyrolysis, Co/Zn ZIFs
broken to release C–N fragments at higher temperature, leading
to the decrease of Co–N bonds (Fig. 14a). As a result, the single-
atom cobalt with two coordinate nitrogen atoms (Co-N2 ) exhib-
ited the higher electrocatalytic activity and selectivity for reduc-
tion of CO2 to CO than the Co center with Co-N4 site (Fig. 14h).
The catalyst with Co-N2 sites displayed the best activity and se-
lectivity among these tested samples, with the current density of
approximately 18.1 mA cm−2 and the Faradaic efficiency of 94% at
a low overpotential of 520 mV in CO2 -saturated 0.5 M KHCO3 so-
lution. Experiments and theoretical calculations indicated that the
high performance of catalysts with low coordination number was
Fig. 12. The proposed mechanism for converting CO2 into hydrocarbon products on
attributed to low charge transfer resistances, low activated energy
OD Cu/C-10 0 0. Reproduced with permission [108]. Copyright 2017, ACS.
to form CO·−∗
2 intermediates and the strong adsorption capacity of
CO·−∗
2 on active sites (Fig. 14i).

Fig. 13. (a) Illustration of the synthesis of Ni SAs/N-C. (b) TEM, (c) HAADF-STEM images of Ni SAs/N-C, and (d) the corresponding SAED pattern of an individual rhomb-
dodecahedron. (e, f) Magnified HAADF-STEM images of Ni SAs/N-C; The Ni single atoms are marked with red circles. (g) EDS mapping of Ni SAs/N-C. (h) The FEs of CO
produced on Ni SAs/N-C and Ni NPs/N-C at different applied potentials; (i) the proposed mechanism for CO2 reduction over Ni SAs/N-C. Reproduced with permission [97].
Copyright 2017, ACS.
168 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

Fig. 14. (a) The schematic fabrication process of Co-N4 and Co-N2 . (b) SEM, (c) TEM images and (d) corresponding EDS mapping of Co-N2 . (e, f) HAADF-STEM images of
Co-N2 , the Co atom is marked with red lines. (g) Corresponding SAED pattern of Co-N2 ; (h) CO Faradaic efficiencies at different applied potentials over different catalysts; (i)
Gibbs free energy diagrams for electrocatalytic CO2 to CO over Co-N2 and Co-N4 catalysts. Reproduced with permission [88]. Copyright 2017, Wiley.

In addition to Co and Ni-based single-atom catalysts, single- duction, which exhibited excellent electrocatalytic activity and se-
atom Fe-based catalysts can also be prepared using ZIF-8 as lectivity due to the tunable metal nodes and organic linkers. Com-
template through confining Fe species inside the cages of ZIF-8 pared with homogeneous molecular catalysts, heterogeneous MOFs
[72]. Ye and co-worker found that this strategy can avoid the not only possess the merits of molecular catalysts and heteroge-
aggregation of iron species and contributes to the formation of neous porous catalysts, but also display the periodic arrangement
isolated single-Fe atom catalysts due to the confinement effect of of organic ligands and metal notes. This type of molecular hetero-
ZIF-8. The isolated single-Fe atom catalysts displayed high activity geneous catalysts is helpful for understanding the relationship be-
and selectivity of CO2 reduction in KHCO3 solution compared tween the structure and properties at the molecular level. So far,
with the Fe-base nanoparticles. Furthermore, the performance of different MOFs have been developed as electrocatalysts for elec-
single-atom catalysts is also dependent of the synthesis method trochemical reduction of CO2 . However, only a few of MOFs ex-
of single-atom catalysts. The single-atom Fe species confined on hibited electrocatalytic reduction of CO2 . Furthermore, most of re-
the surface of the ZIF-8 provided more exposed isolated Fe-N ports mainly focused on metal-porphyrin MOFs due to their unique
active sites compared with the single-atom Fe confined in the square plane coordinated geometry, which favors the CO2 adsorp-
cage of ZIF-8, which contributed to high CO Faradaic efficiency and tion on active sites. On the other hand, owing to the improved
current density toward CO2 reduction. To further understand the electron transfer and mass transport, MOF-derived materials have
mechanism of CO2 reduction on Fe-N-C catalysts, a combination been developed as electrocatalysts for CO2 reduction which ex-
of experimental and theoretical computations would be effective hibited unprecedented electrocatalytic activity, selectivity. Various
method to further clarify the reaction pathway on the M-N4 sites. strategies have been employed to fabricate MOF-derived nanocar-
Pan and co-worker demonstrated that both Fe-N-C and Co-N-C bon hybrids for application in electrocatalytic CO2 reduction, but
single-atom catalysts can effectively catalyze the reduction of the MOF precursors reported are mainly imidazolate-containing ZIF
CO2 to CO, while the Fe-N-C catalysts displayed the highest CO materials.
Faradaic efficiency of 93% at an overpotential 0.47 V among the Although most of MOFs and MOF-derived electrocatalysts have
as-prepared samples [111]. The experimental observations together been shown promising performance, great challenges remain to
with theoretical calculations revealed that the high activity and se- develop MOFs and MOF-derived catalysts with high activity and
lectivity toward CO2 reduction were attributed to the edge-hosted selectivity toward electrochemical reduction of CO2 . Firstly, only
M-N2+2 -C8 sites which bridge in situ formed two armchair-like a few of non-precious metal active centers are available to con-
graphitic layers. The metal centers having dangling bonds neighbor struct MOF-based catalysts including Fe, Co, Ni, Zn, and Cu. Other
to N are the real active sites which absorb the intermediates ∗CO non-precious metals could also be introduced into MOF, such as
and ∗OH after the cleavage of C–O bond during catalysis. Sn, Mo, and W which are considered as promising active centers
for CO2 reduction [121]. Secondly, MOFs and MOF-derived materi-
4. Conclusions and outlooks als were focused on single metal used as metal active sites. Devel-
opment of mixed metal-based catalysts could contribute to under-
In this review, we summarized and discussed an overview of stand the reaction mechanism of CO2 reduction through regulating
progress about different strategies used to prepare MOF-based the electronic structure, the d-band centre, and surface charge den-
electrocatalysts for the electrocatalytic CO2 reduction. Pristine sity of metal centers by incorporation of second active or inactive
MOFs materials were often used as the electrocatalysts for CO2 re- metal [46]. Thirdly, besides porphyrin- and ZIF-based MOFs, more
P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170 169

efforts are needed to design and synthesize different types of MOFs [7] B. Zhang, J. Zhang, J. Energy. Chem. 26 (2017) 1050–1066.
which provide more opportunities to search the potential MOF- [8] J.D. Figueroa, T. Fout, S. Plasynski, H. McIlvried, R.D. Srivastava, Int. J. Greenh.
Gas Control. 2 (2008) 9–20.
based electrocatalysts with high activity and selectivity toward [9] M.H. Beyzavi, R.C. Klet, S. Tussupbayev, J. Borycz, N.A. Vermeulen, C.J. Cramer,
CO2 reduction. Additionally, limited research has been reported on J.F. Stoddart, J.T. Hupp, O.K. Farha, J. Am. Chem. Soc. 136 (2014) 15861–15864.
the utilization of MOF with different morphologies for CO2 re- [10] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm,
T.H. Bae, J.R. Long, Chem. Rev. 112 (2012) 724–781.
duction, such as ultrathin two-dimensional nanosheet, nanowire, [11] J.-P. Zou, D.-D. Wu, J. Luo, Q.-J. Xing, X.-B. Luo, W.-H. Dong, S.-L. Luo,
and nanocage. These unique morphologies could facilitate the elec- H.-M. Du, S.L. Suib, ACS Catal. 6 (2016) 6861–6867.
tronic transfer and mass transport during electrocatalysis. Fourth, [12] H. Mistry, A.S. Varela, S. Kühl, P. Strasser, B.R. Cuenya, Nat. Rev. Mater. 1
(2016) 16009.
the products of MOF-based catalysts reported are main C1 chem-
[13] H. Rao, J. Bonin, M. Robert, J. Phys. Chem. C 122 (2018) 13834–13839.
ical such as CO and HCOOH. It is well known that Cu-based cat- [14] S. Dou, L. Tao, R. Wang, S. El Hankari, R. Chen, S. Wang, Adv. Mater. 30 (2018)
alysts are potential candidates towards converting carbon dioxide 1705850.
[15] H. Huang, J. Lin, G. Zhu, Y. Weng, X. Wang, X. Fu, J. Long, Angew. Chem. 128
to multi-carbon chemical [7,122,123]. However, the efficiency and
(2016) 8454–8458.
selectivity of CO2 reduction over Cu-based MOF materials still re- [16] W. Zhu, C. Zhang, Q. Li, L. Xiong, R. Chen, X. Wan, Z. Wang, W. Chen, Z. Deng,
main a lot of room for improvement. To obtain higher Faraday ef- Y. Peng, Appl. Catal., B 238 (2018) 339–345.
ficiency of multi-carbon products, the rational design and synthe- [17] Y. Pan, R. Lin, Y. Chen, S. Liu, W. Zhu, X. Cao, W. Chen, K. Wu, W.C. Cheong,
Y. Wang, L. Zheng, J. Luo, Y. Lin, Y. Liu, C. Liu, J. Li, Q. Lu, X. Chen, D. Wang,
sis of Cu-based MOFs materials is highly desirable such as con- Q. Peng, C. Chen, Y. Li, J. Am. Chem. Soc. 140 (2018) 4218–4221.
structing single-atom Cu-based catalysts and adjusting the coor- [18] X. Liu, H. Yang, J. He, H. Liu, L. Song, L. Li, J. Luo, Small 14 (2018) 1704049.
dination geometry of Cu sites. Fifth, although the reported MOF- [19] X. Li, W. Bi, M. Chen, Y. Sun, H. Ju, W. Yan, J. Zhu, X. Wu, W. Chu, C. Wu,
Y. Xie, J. Am. Chem. Soc. 139 (2017) 14889–14892.
based materials demonstrated the electrochemical stability by a [20] Z. Miao, P. Hu, C. Nie, H. Xie, W. Fu, Q. Li, J. Energy. Chem. 38 (2019) 114–118.
constant current method, the recyclability of the MOF-based ma- [21] H. Xie, J. Wang, K. Ithisuphalap, G. Wu, Q. Li, J. Energy. Chem. 26 (2017)
terials is scarcely discussed. The surface and bulky structures of 1039–1049.
[22] H. Xie, T. Wang, J. Liang, Q. Li, S. Sun, Nano Today 21 (2018) 41–54.
MOF-based materials after long-term test should be further inves- [23] Y. Lum, J.W. Ager, Nat. Catal. 2 (2019) 86–93.
tigated by a wide range of characterizations, such as X-ray photo- [24] K. Jiang, S. Siahrostami, T. Zheng, Y. Hu, S. Hwang, E. Stavitski, Y. Peng,
electron spectroscopy (XPS), X-ray powder diffraction (XRD), and J. Dynes, M. Gangisetty, D. Su, K. Attenkofer, H. Wang, Energy Environ. Sci.
11 (2018) 893–903.
high-resolution transmission electron microscopy (HRTEM). No-
[25] L. Fan, Z. Xia, M. Xu, Y. Lu, Z. Li, Adv. Funct. Mater. 28 (2018) 1706289.
tably, XRD measurements reflect the bulky structure of materials, [26] Y.X. Duan, F.L. Meng, K.H. Liu, S.S. Yi, S.J. Li, J.M. Yan, Q. Jiang Adv. Mater. 30
and cannot provide accurate information about the surface struc- (2018) 1706194.
[27] X.-F. Bai, W. Chen, B.-Y. Wang, G.-H. Feng, W. Wei, Z. Jiao, Y.-H. Sun, Acta Phys.
ture of materials which is a crucial for photo- and electro-catalysts.
Chim. Sin. 33 (2017) 2388–2403.
On the contrast, HRTEM is an effective method to govern the struc- [28] D. Gao, F. Cai, G. Wang, X. Bao, Curr. Opin. Green Sustain. Chem. 3 (2017)
tural changes in the surface and interface of catalysts [124]. Sixth, 39–44.
although single-atom catalysts derived from MOFs exhibit excellent [29] C.D. Malonzo, S.M. Shaker, L.M. Ren, S.D. Prinslow, A.E. Platero-Prats,
L.C. Gallington, J. Borycz, A.B. Thompson, T.C. Wang, O.K. Farha, J.T. Hupp,
CO2 reduction performance, this type of catalysts probably face the C.C. Lu, K.W. Chapman, J.C. Myers, R.L. Penn, L. Gagliardi, M. Tsapatsis,
diffusivity of sub-nanometer species and high mobility. Thus, fur- A. Stein, J. Am. Chem. Soc. 138 (2016) 2739–2748.
ther efforts are needed to develop effective strategies for design [30] Q. Lu, F. Jiao, Nano Energy 29 (2016) 439–456.
[31] H.Q. Xu, J. Hu, D. Wang, Z. Li, Q. Zhang, Y. Luo, S.H. Yu, H.L. Jiang, J. Am. Chem.
and synthesis of single-atom catalysts with controllable structure Soc. 137 (2015) 13440–13443.
anchoring single-atom active sites. Finally, mechanistic understand- [32] C. Wang, Z. Xie, K.E. deKrafft, W. Lin, J. Am. Chem. Soc. 133 (2011)
ing of CO2 electroreduction is of great importance to design and 13445–13454.
[33] L. Shi, T. Wang, H. Zhang, K. Chang, J. Ye, Adv. Funct. Mater. 25 (2015)
prepare high-performance catalysts. A combination of theoretical 5360–5367.
calculations and in-situ experimental technologies is an effectively [34] N.C. Thacker, Z.K. Lin, T. Zhang, J.C. Gilhula, C.W. Abney, W.B. Lin, J. Am. Chem.
approach to exploring the intermediates and reveal the structural Soc. 138 (2016) 3501–3509.
[35] H.-Q. Xu, J. Hu, D. Wang, Z. Li, Q. Zhang, Y. Luo, S.-H. Yu, H.-L. Jiang, J. Am.
change of catalysts during electrocatalysis. These knowledges are
Chem. Soc. 137 (2015) 13440–13443.
helpful for rational design of new catalysts and avoiding a large [36] T. Zhang, W. Lin, Chem. Soc. Rev. 43 (2014) 5982–5993.
amount of trial-and-error experiments. In summary, creation of [37] J. Low, J. Yu, W. Ho, J. Phys Chem. Lett. 6 (2015) 4244–4251.
[38] I.I. Alkhatib, C. Garlisi, M. Pagliaro, K. Al-Ali, G. Palmisano, Catal. Today (2018),
stable high-performance MOF-based electrocatalysts for CO2 reduc-
doi:10.1016/j.cattod.2018.1009.1032.
tion is still a significant challenge. Hopefully, the recent progress [39] A. Corma, H. Garcia, J. Catal. 308 (2013) 168–175.
on MOF-based materials for electrocatalytic CO2 reduction could [40] K. Li, B. Peng, T. Peng, ACS Catal. 6 (2016) 7485–7527.
not only promote the development of the MOFs in the field of en- [41] S.-L. Xie, J. Liu, L.-Z. Dong, S.-L. Li, Y.-Q. Lan, Z.-M. Su, Chem. Sci. 10 (2019)
185–190.
ergy conversion, but also provide clues to design and synthesis of [42] A.M. Appel, J.E. Bercaw, A.B. Bocarsly, H. Dobbek, D.L. DuBois, M. Dupuis,
functional MOF-based materials for CO2 reduction. J.G. Ferry, E. Fujita, R. Hille, P.J. Kenis, C.A. Kerfeld, R.H. Morris, C.H. Peden,
A.R. Portis, S.W. Ragsdale, T.B. Rauchfuss, J.N. Reek, L.C. Seefeldt, R.K. Thauer,
G.L. Waldrop, Chem. Rev. 113 (2013) 6621–6658.
Acknowledgments [43] A.S. Varela, W. Ju, P. Strasser, Adv. Energy Mater. 8 (2018) 1703614.
[44] J. Qiao, Y. Liu, F. Hong, J. Zhang, Chem. Soc. Rev. 43 (2014) 631–675.
[45] F.N. Al-Rowaili, A. Jamal, M.S. Ba Shammakh, A. Rana, ACS Sustain. Chem. Eng.
This work was supported by the National Natural Science Foun- 6 (2018) 15895–15914.
dation of China (51772291, 21673238, 21773242) and the Strate- [46] R.M. Arán-Ais, D. Gao, B. Roldan Cuenya, Acc. Chem. Res. 51 (2018)
gic Priority Research Program of the Chinese Academy of Sciences 2906–2917.
[47] L.F. Song, J. Zhang, L.X. Sun, F. Xu, F. Li, H.Z. Zhang, X.L. Si, C.L. Jiao, Z.B. Li,
(XDB20 0 0 0 0 0 0).
S. Liu, Y.L. Liu, H.Y. Zhou, D.L. Sun, Y. Du, Z. Cao, Z. Gabelica, Energy Environ.
Sci. 5 (2012) 7508–7520.
References [48] W. Xia, A. Mahmood, R.Q. Zou, Q. Xu, Energy Environ. Sci. 8 (2015)
1837–1866.
[1] D.D. Zhu, J.L. Liu, S.Z. Qiao, Adv Mater. 28 (2016) 3423–3452. [49] Z. Chang, D.H. Yang, J. Xu, T.L. Hu, X.H. Bu, Adv. Mater. 27 (2015) 5432–5441.
[2] T. Zheng, K. Jiang, H. Wang, Adv Mater. (2018) e1802066. [50] S.T. Meek, J.A. Greathouse, M.D. Allendorf, Adv. Mater. 23 (2011) 249–267.
[3] A. Vasileff, Y. Zheng, S.Z. Qiao, Adv. Energy Mater. 7 (2017) 1700759. [51] A. Morozan, F. Jaouen, Energy Environ. Sci. 5 (2012) 9269–9290.
[4] M. Ma, W.A. Smith, Anisotropic and Shape-Selective Nanomaterials: Struc- [52] D. Zhao, D.Q. Yuan, H.C. Zhou, Energy Environ. Sci. 1 (2008) 222–235.
ture-Property Relationships, Springer International Publishing, Cham, 2017, [53] Y.-X. Tan, F. Wang, J. Zhang, Chem. Soc. Rev. 47 (2018) 2130–2144.
pp. 337–373. [54] M.R. Liu, Q.L. Hong, Q.H. Li, Y. Du, H.X. Zhang, S. Chen, T. Zhou, J. Zhang, Adv.
[5] X. Duan, J. Xu, Z. Wei, J. Ma, S. Guo, S. Wang, H. Liu, S. Dou, Adv. Mater. 29 Funct. Mater. 28 (2018) 1801136.
(2017) 1701784. [55] F. Wang, Z.-S. Liu, H. Yang, Y.-X. Tan, J. Zhang, Angew. Chem. Int. Ed. 50 (2011)
[6] S. Chu, A. Majumdar, Nature 488 (2012) 294–303. 450–453.
170 P. Shao, L. Yi and S. Chen et al. / Journal of Energy Chemistry 40 (2020) 156–170

[56] C. Wang, D. Liu, W. Lin, J. Am. Chem. Soc. 135 (2013) 13222–13234. [94] B. An, L. Zeng, M. Jia, Z. Li, Z. Lin, Y. Song, Y. Zhou, J. Cheng, C. Wang, W. Lin,
[57] S.M. Cohen, Chem. Sci. 1 (2010) 32–36. J. Am. Chem. Soc. 139 (2017) 17747–17750.
[58] T. Zhou, Y. Du, A. Borgna, J. Hong, Y. Wang, J. Han, W. Zhang, R. Xu, Energy [95] H. Nishihara, T. Hirota, K. Matsuura, M. Ohwada, N. Hoshino, T. Akutagawa,
Environ. Sci. 6 (2013) 3229–3234. T. Higuchi, H. Jinnai, Y. Koseki, H. Kasai, Y. Matsuo, J. Maruyama, Y. Hayasaka,
[59] S.-L. Li, Q. Xu, Energy Environ. Sci. 6 (2013) 1656–1683. H. Konaka, Y. Yamada, S. Yamaguchi, K. Kamiya, T. Kamimura, H. Nobukuni,
[60] H. Zhang, J. Li, Q. Tan, L. Lu, Z. Wang, G. Wu, Chem. Eur. J. 24 (2018) F. Tani, Nat. Commun. 8 (2017) 109.
18137–18157. [96] H.Y. Jeong, M. Balamurugan, V.S.K. Choutipalli, J. Jo, H. Baik, V. Subramanian,
[61] Z. Lei, Y. Xue, W. Chen, W. Qiu, Y. Zhang, S. Horike, L. Tang, Adv. Energy Mater. M. Kim, U. Sim, K.T. Nam, Chem. Eur. J. 24 (2018) 18444–18454.
8 (2018) 1801587. [97] C. Zhao, X. Dai, T. Yao, W. Chen, X. Wang, J. Wang, J. Yang, S. Wei, Y. Wu, Y. Li,
[62] C.S. Diercks, Y. Liu, K.E. Cordova, O.M. Yaghi, Nat Mater 17 (2018) 301–307. J. Am. Chem. Soc. 139 (2017) 8078–8081.
[63] Z.-G. Gu, J. Zhang, Coord. Chem. Rev. 378 (2019) 513–532. [98] Y. Guo, H. Yang, X. Zhou, K. Liu, C. Zhang, Z. Zhou, C. Wang, W. Lin, J. Mater.
[64] Z.-G. Gu, C. Zhan, J. Zhang, X. Bu, Chem. Soc. Rev. 45 (2016) 3122–3144. Chem. A 5 (2017) 24867–24873.
[65] H.-X. Zhang, M. Liu, T. Wen, J. Zhang, Coord. Chem. Rev. 307 (2016) 255–266. [99] X. Zhang, Z. Wu, X. Zhang, L. Li, Y. Li, H. Xu, X. Li, X. Yu, Z. Zhang, Y. Liang,
[66] C.-C. Hou, Q. Xu, Adv. Energy Mater. (2018), doi:10.1002/aenm.201801307. H. Wang, Nat. Commun. 8 (2017) 14675.
[67] Q.-L. Zhu, Q. Xu, Chem. Soc. Rev. 43 (2014) 5468–5512. [100] Y.R. Wang, Q. Huang, C.T. He, Y. Chen, J. Liu, F.C. Shen, Y.Q. Lan, Nat. Commun.
[68] H. Zhang, J. Nai, L. Yu, X.W. Lou, Joule 1 (2017) 77–107. 9 (2018) 4466.
[69] J. Lee, O.K. Farha, J. Roberts, K.A. Scheidt, S.T. Nguyen, J.T. Hupp, Chem. Soc. [101] Y. Pan, R. Lin, Y. Chen, S. Liu, W. Zhu, X. Cao, W. Chen, K. Wu, W.C. Cheong,
Rev. 38 (2009) 1450–1459. Y. Wang, L. Zheng, J. Luo, Y. Lin, Y. Liu, C. Liu, J. Li, Q. Lu, X. Chen, D. Wang,
[70] Y. Xu, Q. Li, H. Xue, H. Pang, Coord. Chem. Rev. 376 (2018) 292–318. Q. Peng, C. Chen, Y. Li, J. Am. Chem. Soc. 140 (2018) 4218–4221.
[71] G. Liu, X. Meng, H. Zhang, G. Zhao, H. Pang, T. Wang, P. Li, T. Kako, J. Ye, [102] M. Zhu, J. Chen, L. Huang, R. Ye, J. Xu, Y.F. Han, Angew. Chem. Int. Ed. (2019),
Angew. Chem. Int. Ed. 56 (2017) 5570–5574. doi:10.10 02/anie.20190 0499.
[72] Y. Ye, F. Cai, H. Li, H. Wu, G. Wang, Y. Li, S. Miao, S. Xie, R. Si, J. Wang, X. Bao, [103] J. Choi, P. Wagner, S. Gambhir, R. Jalili, D.R. MacFarlane, G.G. Wallace, D.L. Of-
Nano Energy 38 (2017) 281–289. ficer, ACS Energy Lett. 4 (2019) 666–672.
[73] X. Jiang, H. Wu, S. Chang, R. Si, S. Miao, W. Huang, Y. Li, G. Wang, X. Bao, J. [104] R. Hinogami, S. Yotsuhashi, M. Deguchi, Y. Zenitani, H. Hashiba, Y. Yamada,
Mater. Chem. A 5 (2017) 19371–19377. ECS Electrochem. Lett. 1 (2012) H17–H19.
[74] J. Su, Y. Yang, G. Xia, J. Chen, P. Jiang, Q. Chen, Nat. Commun. 8 (2017) 14969. [105] R. Senthil Kumar, S. Senthil Kumar, M. Anbu Kulandainathan, Electrochem.
[75] A. Mahmood, W. Guo, H. Tabassum, R. Zou, Adv. Energy Mater. 6 (2016) Commun. 25 (2012) 70–73.
1600423. [106] S. Kusama, T. Saito, H. Hashiba, A. Sakai, S. Yotsuhashi, ACS Catal. 7 (2017)
[76] A.A. Olajire, Renew. Sustain. Energy Rev. 92 (2018) 570–607. 8382–8385.
[77] D.D. Zhu, J.L. Liu, S.Z. Qiao, Adv. Mater. 28 (2016) 3423–3452. [107] Y.L. Qiu, H.X. Zhong, T.T. Zhang, W.B. Xu, P.P. Su, X.F. Li, H.M. Zhang, ACS Appl.
[78] Z. Tian, C. Priest, L. Chen, Adv. Theory Simul. 1 (2018) 180 0 0 04. Mater. Interfaces 10 (2018) 2480–2489.
[79] S. Aoi, K. Mase, K. Ohkubo, T. Suenobu, S. Fukuzumi, ACS Energy Lett. 2 (2017) [108] K. Zhao, Y. Liu, X. Quan, S. Chen, H. Yu, ACS Appl. Mater. Interfaces 9 (2017)
532–536. 5302–5311.
[80] P. Shao, S. Ci, L. Yi, P. Cai, P. Huang, C. Cao, Z. Wen, ChemElectroChem 4 (2017) [109] Z. Weng, Y. Wu, M. Wang, J. Jiang, K. Yang, S. Huo, X.F. Wang, Q. Ma,
2593–2598. G.W. Brudvig, V.S. Batista, Y. Liang, Z. Feng, H. Wang, Nat. Commun. 9 (2018)
[81] K.J.P. Schouten, Y. Kwon, C.J.M. van der Ham, Z. Qin, M.T.M. Koper, Chem. Sci. 415.
2 (2011) 1902. [110] Z. Weng, J. Jiang, Y. Wu, Z. Wu, X. Guo, K.L. Materna, W. Liu, V.S. Batista,
[82] A.A. Peterson, J.K. Nørskov, J. Phy. Chem. Lett. 3 (2012) 251–258. G.W. Brudvig, H. Wang, J. Am. Chem. Soc. 138 (2016) 8076–8079.
[83] R. Kortlever, J. Shen, K.J.P. Schouten, F. Calle-Vallejo, M.T.M. Koper, J. Phy. [111] F. Pan, H. Zhang, K. Liu, D. Cullen, K. More, M. Wang, Z. Feng, G. Wang, G. Wu,
Chem.Lett. 6 (2015) 4073–4082. Y. Li, ACS Catal. 8 (2018) 3116–3122.
[84] J. Albo, D. Vallejo, G. Beobide, O. Castillo, P. Castano, A. Irabien, ChemSusChem [112] C. Gao, S. Chen, Y. Wang, J. Wang, X. Zheng, J. Zhu, L. Song, W. Zhang, Y.
10 (2017) 1100–1109. Xiong, 30 (2018) 1704624.
[85] J. Shen, R. Kortlever, R. Kas, Y.Y. Birdja, O. Diaz-Morales, Y. Kwon, I. Ledez- [113] M.-M. Millet, G. Algara-Siller, S. Wrabetz, A. Mazheika, F. Girgsdies,
ma-Yanez, K.J. Schouten, G. Mul, M.T. Koper, Nat. Commun. 6 (2015) 8177. D. Teschner, F. Seitz, A. Tarasov, S.V. Levchenko, R. Schlögl, E. Frei, J. Am.
[86] D.H. Nam, O.S. Bushuyev, J. Li, P. De Luna, A. Seifitokaldani, C.T. Dinh, Chem. Soc. 141 (2019) 2451–2461.
F.P. Garcia de Arquer, Y. Wang, Z. Liang, A.H. Proppe, C.S. Tan, P. Todorovic, [114] X. Su, X.-F. Yang, Y. Huang, B. Liu, T. Zhang, Acc. Chem. Res. 52 (2019)
O. Shekhah, C.M. Gabardo, J.W. Jo, J. Choi, M.J. Choi, S.W. Baek, J. Kim, D. Sin- 656–664.
ton, S.O. Kelley, M. Eddaoudi, E.H. Sargent, J. Am. Chem. Soc. 140 (2018) [115] I. Bhugun, D. Lexa, J.-M. Saveant, J. Am. Chem. Soc. 116 (1994) 5015–5016.
11378–11386. [116] B.-X. Dong, S.-L. Qian, F.-Y. Bu, Y.-C. Wu, L.-G. Feng, Y.-L. Teng, W.-L. Liu,
[87] T.N. Huan, N. Ranjbar, G. Rousse, M. Sougrati, A. Zitolo, V. Mougel, F. Jaouen, Z.-W. Li, ACS Appl. Energy Mater. 1 (2018) 4662–4669.
M. Fontecave, ACS Catal. 7 (2017) 1520–1525. [117] B.A. Rosen, A. Salehi-Khojin, M.R. Thorson, W. Zhu, D.T. Whipple, P.J.A. Kenis,
[88] X. Wang, Z. Chen, X. Zhao, T. Yao, W. Chen, R. You, C. Zhao, G. Wu, J. Wang, R.I. Masel, Science 334 (2011) 643–644.
W. Huang, J. Yang, X. Hong, S. Wei, Y. Wu, Y. Li, Angew. Chem. Int. Ed. 57 [118] H. Furukawa, K.E. Cordova, M. O’Keeffe, O.M. Yaghi, Science 341 (2013) 974.
(2018) 1944–1948. [119] J. Yang, Z. Qiu, C. Zhao, W. Wei, W. Chen, Z. Li, Y. Qu, J. Dong, J. Luo, Z. Li,
[89] N. Kornienko, Y. Zhao, C.S. Kley, C. Zhu, D. Kim, S. Lin, C.J. Chang, O.M. Yaghi, Y. Wu, Angew. Chem. Int. Ed. 57 (2018) 14095–14100.
P. Yang, J. Am. Chem. Soc. 137 (2015) 14129–14135. [120] Y. Qian, Z. Hu, X. Ge, S. Yang, Y. Peng, Z. Kang, Z. Liu, J.Y. Lee, D. Zhao, Carbon
[90] X. Kang, Q. Zhu, X. Sun, J. Hu, J. Zhang, Z. Liu, B. Han, Chem. Sci. 7 (2016) 111 (2017) 641–650.
266–273. [121] J. Wu, Y. Huang, W. Ye, Y. Li, Adv. Sci. 4 (2017) 1700194.
[91] Y. Wang, P. Hou, Z. Wang, P. Kang, ChemPhysChem 18 (2017) 3142–3147. [122] A.J. Garza, A.T. Bell, M. Head-Gordon, ACS Catal. 8 (2018) 1490–1499.
[92] L. Ye, J. Liu, Y. Gao, C. Gong, M. Addicoat, T. Heine, C. Wöll, L. Sun, J. Mater. [123] D. Gao, R.M. Arán-Ais, H.S. Jeon, B. Roldan Cuenya, Nat. Catal. 2 (2019)
Chem. A 4 (2016) 15320–15326. 198–210.
[93] I. Hod, M.D. Sampson, P. Deria, C.P. Kubiak, O.K. Farha, J.T. Hupp, ACS Catal. 5 [124] S.W. Lee, C. Carlton, M. Risch, Y. Surendranath, S. Chen, S. Furutsuki, A. Ya-
(2015) 6302–6309. mada, D.G. Nocera, Y. Shao-Horn, J. Am. Chem. Soc. 134 (2012) 16959–16962.

You might also like