Copper-based metal-organic frameworks for electrochemical reduction of CO2

Xiaomin Kang Guodong Fu Xian-Zhu Fu Jing-Li Luo

Citation:  Xiaomin Kang, Guodong Fu, Xian-Zhu Fu, Jing-Li Luo. Copper-based metal-organic frameworks for electrochemical reduction of CO2[J]. Chinese Chemical Letters, 2023, 34(6): 107757. doi: 10.1016/j.cclet.2022.107757 shu

Copper-based metal-organic frameworks for electrochemical reduction of CO2

English

  • The increasing of CO2 emission in the earth's atmosphere is nowadays an urgent and critical issue facing humanity [1]. It further causes the rapid accumulation of greenhouse gas, melting of floating icebergs, a rise in sea level and so forth, threatening the existence and development of human race. Thus, the electrochemical conversion of electricity and CO2 into chemicals and fuels has drawn great attention [2,3]. Electrochemical CO2 reduction reaction (CO2ER) is energetically and atomically inefficient caused by large kinetic overpotentials and lack in selectivity of the chemical product, especially the competition with the hydrogen evolution reaction in aqueous solutions. The CO2 molecules get to the catalytic sites through diffusion in the aqueous phase, and several valuable products can be generated, such as carbon monoxide (CO), ethylene (C2H4), formic acid (HCOOH), oxalic acid (H2C2O4) and alcohols (CH3OH, C2H5OH, etc.) [4]. In CO2ER, the cathodic reaction is of general form:

    (1)

    while the anodic reaction in CO2ER is oxygen evolution reaction (OER), like in water splitting in order to sum to the overall reaction [5]:

    (2)

    In other words, CO2 + 2H+ + 2e → HCOOH, water is the only renewable and scalable source of electrons and protons [6]. Besides, the equilibrium potentials for CO2ER have been listed in Table 1 as follows.

    Table 1

    Table 1.  Electrochemical reactions with equilibrium potentials.
    DownLoad: CSV

    All the CO2ER standard potentials are calculated via the Gibbs free energy of reactions using gas-phase thermochemistry and the Henry law data from National Institute of Standards and Technology (NIST) for aqueous products. Furthermore, the CO2ER products exhibit similar reduction potentials. This explains why CO2ER and hydrogen-evolution reaction (HER) compete since their reduction potentials are similar [7,8].

    Among all the catalysts for CO2ER, Cu and its oxides have been well developed for the electroreduction of CO2 to high-value hydrocarbons, fuels, alcohol products, and multicarbon chemicals [9-12]. In 1989, Hori et al. first analyzed the gaseous and liquid products of a CO2ER catalyzed by polycrystalline metal electrodes (Table 2) [13]. Among various metal electrodes, Cu exhibits outstanding selectivity toward hydrocarbon, aldehyde, and alcohol products. Moreover, as a pure metal, Cu can reduce CO2 to products that require the transfer of more than two electrons with substantial Faradaic efficiencies (FEs).

    Table 2

    Table 2.  Summary of properties of various metals towards the electrochemical reduction of CO2.
    DownLoad: CSV

    Furthermore, metals are divided into four groups based on their selectivity, which depends on their binding energy to key CO2ER and HER intermediates, including *H, *OCHO, *COOH, and *CO [14]. Specifically, Cu alone can reduce CO2 to > 2e products [15]. This is attributed to its negative adsorption energy for *CO and positive adsorption energy for *H [16]. In addition, the multiple crystal facets, grain boundaries, and defects on the surface of metal catalysts influence their selectivity for CO2ER products [17].

    Nevertheless, various Cu(hkl) surfaces have different effects on the product distribution [18]. Thus far, according to experimental data, the surface crystal facet has been determined to have a slight effect on the catalytic activity. However, crystal defects have been shown to significantly impact the CO2ER selectivity [19]. For instance, the Cu(111) surface shows considerably higher selectivity toward C2 products than the Cu(751) surface, which has abundant highly undercoordinated sites [20]. Among the three Cu single-crystal facets with low indexes, i.e., Cu(111), Cu(110), and Cu(100), the Cu(100) surface exhibits the highest selectivity toward C2H6, while Cu(111) and Cu(110) exhibit the highest selectivity toward CH4. The Cu(100) surfaces more easily form alcohols (CH3OH, C2H5OH, etc.) compared to formic acid at −1.00 V vs. reversible hydrogen electrode (RHE), while the opposite is observed for Cu(111). The Cu(110) surface exhibits high selectivity toward alcohols and carboxylates at similar rates. The C2H4/CH4 ratio is plotted as a function of the crystal orientation, and the results clearly demonstrate that steps can enhance the selectivity toward C2H4 formation in comparison with planar single crystals [21]. Further, defects enhance the selectivity toward ethanol (C2H5OH) and C2H4.

    Hori et al. during their tests, found that the single-crystal electrode could be reduced and reconstructed, thus leading to the formation of defects. Specifically, the reconstruction on the surface of a Cu electrode would alter the activity and product selectivity [4]. To date, the mechanisms of the CO2ER on Cu electrodes remain unclear. However, early experiments revealed that the reduction of CO on Cu results in similar product distribution as that for reduction of CO2 (CO2R) on Cu, and the following research on CO reduction on Cu further supported this proposal [22]. Therefore, CO is a key intermediate in the CO2R to hydrocarbons, aldehydes, and alcohols on Cu [23]. In detail, *CO intermediates (adsorbed CO) are bound on the Cu surface and converted into alcohols by *CO dimerization or hydrocarbons via *COH or *CHO intermediates [24]. Furthermore, CO adsorption on Cu can suppress the competing HER because of site-blocking effects and the changes in the *H-binding energy. This explains the high FE for CO2R in aqueous electrolytes [25]. Overall, Cu is a transition metal capable of catalyzing the conversion of CO2 directly into various value-added products.

    In electrocatalysis, the surface area of the catalyst significantly influences its CO2ER activity. Cu-based catalysts have been fabricated into various shapes to enlarge their active surface areas. Among these Cu-based electrocatalysts, Cu-based metal-organic frameworks (MOFs), as porous crystalline materials consisting of a 3D framework of Cu ions or Cu clusters linked by organic linkers, have attracted significant interest for their application as electrocatalysts in CO2ER [26-31].

    Cu-based MOFs exhibit many intriguing properties, including high crystallinity and porosity, and strong metal-ligand coordination bonds; thus, they exhibit a high adsorption capacity for reaction intermediates [32-36]. Moreover, the formation of undercoordinated Cu sites in MOFs can be promoted to enhance the CO2ER efficiency; This is attributed to the chemical tunability and the porosity/surface-topography controllability of MOFs. Thus, Cu-based MOFs can be explored as a catalytic material for better understanding the CO2ER. However, owing to the textural properties and poor stability of MOFs, Cu-based MOFs are employed as active supports and catalyst precursors, which are either treated by pyrolysis or electrochemical reaction to afford stable and efficient electrocatalysts for the CO2ER (Fig. 1).

    Figure 1

    Figure 1.  Cu-MOF related catalysts engineering for CO2ER.

    The combination of the catalytic properties of Cu with the unique structural features of MOFs has been explored in CO2ER research (Table 3). In particular, Cu ions coordinate with organic ligands, forming 3D network-like structures. Subsequently, the Cu centers participate in electrochemical redox processes. The coordinated Cu2+ can shift into transient reduced states, i.e., Cu+ and Cu0, or transient oxidized states, i.e., Cu3+ or Cu4+.

    Table 3

    Table 3.  Summary of electrochemical reduction of CO2 (CO2ER) performance of Cu-MOF catalysts.
    DownLoad: CSV

    In 2010, Angamuthu et al. reported a dinuclear Cu(Ⅰ) complex as a reductant for CO2ER (Fig. 2a). The Cu(Ⅱ) complex could be reduced to regenerate the Cu(Ⅰ) complex by treatment in acetonitrile with a soluble lithium salt at a relatively accessible potential [37].

    Figure 2

    Figure 2.  Copper ions coordinated by organic ligands, resulting in 3D network structures. (a) Schematic overview of the formation and reactivity of the Cu complexes. Reproduced with permission [37]. Copyright 2010, Science. (b) Optimized structure of the models for Cu-MOF-5, and CO2 adsorbed on Cu-MOF-5 (Ads). Reproduced with permission [40]. Copyright 2013, American Chemical Society. (c) Structural schematic diagram of GDE with Cu3(BTC)2 as CO2 capture agent. Reproduced with permission [44]. Copyright 2018, American Chemical Society. (d) Molecular structures of three Cu-complex materials and their electrocatalytic performance for CO2 reduction. Molecular structures of CuPc, HKUST-1, and [Cu(cyclam)]Cl2. Reproduced with permission [46]. Copyright 2018, Springer Nature. (e) Hierarchical tuning of the performance of electrochemical carbon dioxide reduction using conductive two-dimensional metallophthalocyanine based metal-organic framework. Reproduced with permission [49]. Copyright 2020, American Chemical Society. (f) Cu3(BTC)2 metal-organic framework (Cu-MOF) and graphene oxide composite in electroreduction of CO2 to formic acid. Reproduced with permission [51]. Copyright 2020, American Chemical Society.

    For the application of Cu-MOF as a catalyst for CO2ER, which was firstly investigated in 2012, Hinogami et al. synthesized a copper rubeanate MOF (CR-MOF) to enhance the selectivity of the CO2ER. At the onset potential of ~ 0.20 V, which was more positive than that for a Cu electrode in an aqueous electrolyte, formic acid was the only CO2R product obtained on the CR-MOF electrode (FE ≈ 100.00%). However, the Cu electrode generated several products. Meanwhile, the quantity of products and the partial current of formic acid on the CR-MOF electrode were considerably greater than those on the Cu electrode [38].

    Later in 2012, Kumar et al. prepared copper benzene-1,3,5-tricarboxylate (Cu-BTC) porous films, which exhibited excellent stability and uniform dispersion in an aqueous system as electrocatalysts loaded on glassy carbon electrodes in 0.10 mol/L KCl for CO2ER. Cyclic voltammetric studies indicated that Cu(Ⅰ) would react in-situ with CO2. Moreover, by adding N, N-dimethylformamide (DMF) to saturated CO2, oxalic acid can be obtained (FE of ~ 51.00% and a current density of 19.00 mA/cm2). During the CO2ER process, protons are taken from DMF through a two-electron pathway and dimerization process, thus reducing CO2 to oxalic acid [39].

    Maihom et al. further investigated the reaction mechanisms of CO2 hydrogenation to formic acid over Cu-alkoxide-functionalized MOFs (Cu-MOF-5) by M06-L density functional calculations (Fig. 2b). The results revealed that the CO2ER has two pathways: concerted and stepwise. Compared with the concerted mechanism, the stepwise mechanism firstly proceeds with hydrogen abstraction by CO2 to yield a formate intermediate, after which another hydrogen is obtained to yield formic acid. The activation energies are approximately 24.20 and 18.30 kcal/mol for the first and second steps, respectively. These energies are considerably smaller than that for the concerted mechanism (67.20 kcal/mol). This result indicated that the stepwise mechanism was more efficient than the concerted mechanism. Moreover, the catalytic effect of Cu-MOF-5 was comparable to that achieved in a gas-phase uncatalyzed reaction with an activation energy of 73.00 kcal/mol. This further indicated that Cu-MOF-5 can be utilized as an electrocatalyst for the CO2ER [40].

    Following these pioneer works, Cu-based MOF-loaded gas-diffusion electrodes (GDE), including 1-(2-methyl-4-(2-oxopyrrolidin-1-yl)phenyl)-3-morpholino-5,6-dihydropyridin-2(1H)-one (HKUST-1) MOF, copper(Ⅱ)–adeninate–acetate coordination framework (CuAdeAce) MOF, copper bis-bidentate dithiooxamidate (CuDTA) mesoporous metal-organic aerogel (MOA), and CuZnDTA MOA, were employed in CO2ER by Albo et al. These electrode materials exhibited excellent catalytic potentials because of their high surface areas and abundant, exposed Cu active centers, which were beneficial for CO2ER. Furthermore, the production of methanol (CH3OH) and C2H5OH in the liquid phase can be achieved with a high efficiency. The FEs for the CO2ER on HKUST-1, CuAdeAce-, CuDTA- and CuZnDTA-based electrodes were 15.90%, 1.20%, 6.00% and 9.90%, respectively, at the current density of 10.00 mA/cm2. Among the sample electrode materials, HKUST-1- and CuZnDTA-based electrodes exhibited outstanding and stable electrocatalytic performances [41].

    In 2016, Rungtaweevoranit et al. utilized Zr-based MOFs to promote the activity and selectivity of a Cu catalyst. Specifically, an 18.00 nm-Cu nanocrystal was encapsulated in a UiO-66 single crystal, and it exhibited strong interaction with the Zr oxide of the MOF support. The Cu⊂UiO-66 structure induced an 8-fold enhancement in the yield and 100.00% selectivity for CH3OH. This indicated that Cu nanocrystals embedded in MOFs can be applied to achieve high catalytic activities [42].

    Kung et al. reported a method to anchor Cu nanoparticles (NPs) on NU-1000 thin films by installing single-site Cu(Ⅱ) into the NU-1000 thin film. In the subsequent electrochemical reduction reaction, Cu(Ⅱ) was reduced to metallic Cu. Furthermore, the Cu nanoparticles exhibited moderate electrocatalytic activity, with an HCOO FE of 28.00% at −1.20 mA/cm2 and −0.82 V vs. RHE [43].

    Moreover, Qiu et al. utilized Cu3(BTC)2 (Cu-MOFs) as electrocatalysts for CO2 capture. This material was employed in the fabrication of a GDE (Fig. 2c). The FEs of CH4 on the GDE with Cu-MOFs were 2–3 fold higher than those on the GDE without Cu–MOFs at negative potentials (−2.30 ~ −2.50 V vs. saturated calomel electrode (SCE)). Meanwhile, the FE of the competitive HER was reduced by 30%. The enhanced CO2ER performance on the GDE with Cu-MOFs was attributed to the accumulation of CO2 on the interface of the GDE and the electrolyte, which was attributed to the CO2-capture properties of Cu-MOFs [44].

    In 2019, An et al. put forward a catalyst composed of cooperative Cu(Ⅰ) sites on a Zr12 cluster of a MOF for CO2ER to C2H5OH. By introducing the alkali cation, the spatially proximate Zr12-supported Cu(Ⅰ) centers activate hydrogen via bimetallic oxidative addition and promote C-C coupling to yield C2H5OH. Notably, Cs+-modified MOF catalysts yielded C2H5OH with > 99.00% selectivity after a 10 h reaction. This demonstrated that the catalyst can be employed as a tunable platform to design non-precious metal catalysts for future CO2ER [45].

    Apart from the above Cu-MOFs, three Cu complex materials for CO2ER were investigated by Weng et al. (Fig. 2d). Among these three materials, copper(Ⅱ) phthalocyanine exhibited the highest selectivity toward CH4 with a FE of ~ 66.00% at 13.00 mA/cm2 and −1.06 V vs. RHE. In-situ X-ray absorption spectroscopy analysis highlighted that copper(Ⅱ) phthalocyanine underwent reversible structural and oxidation state changes to form 2.00 nm-metallic Cu clusters. Density functional calculations further confirmed the molecular structure of the metal-ion ligand coordination and the dispersed Cu clusters, which contributed to the CO2ER selectivity toward CH4 [46].

    Nevertheless, Xia et al. utilized Cu/adeninato/carboxylato metal-biomolecule frameworks (Cu/ade-MOFs) for CO2ER toward hydrocarbon (CH4, and C2H4) generation. The cathodized Cu/ade-MOFs exhibited excellent catalytic performance with a total hydrocarbon FE of over 73.00%. C2H4 was collected and measured with a maximum FE of 45.00% at 8.50 mA/cm2 and −1.40 V vs. RHE, while CH4 was obtained with a FE of 50.00% at ~15.00 mA/cm2 and −1.60 V vs. RHE. This indicated that the reconstruction of cathodized Cu/ade-MOFs and the formed Cu nanoparticles functionalized by N-containing ligands contributed to the excellent CO2ER performance [47].

    In 2019, Wu et al. utilized Cu(Ⅱ) porphyrinic MOF nanosheets for CO2ER and compared the results with those obtained with CuO, Cu2O, Cu, a porphyrin-Cu(Ⅱ) complex, and a CuO/complex composite. Among them, Cu-MOF nanosheets exhibited excellent formate selectivity with a FE of 68.40% at −1.55 V vs. Ag/Ag+. Furthermore, the C-C coupling product, acetate, and formate were obtained in a rather wide voltage range from –1.40 V to −1.65 V with the total liquid product FE ranging from 38.00% to 85.20%. Thus, the high selectivity and catalytic activity of Cu-MOFs for CO2R to yield formate and acetate should be attributed to the synergistic enhancement of the porphyrin-Cu(Ⅱ) complex [48].

    In addition to the above structures, four systematic structural analogs of conductive 2D MOFs composed of metallophthalocyanine (MPc) ligands linked by Cu nodes for CO2ER to CO were investigated by Meng et al. (Fig. 2e). The catalytic activity and selectivity of the MOFs were determined hierarchically by two structural factors: the metal within the MPc (M = Co vs. Ni) catalytic subunit and the identity of the heteroatomic cross-linker (X) between the subunits (X = O vs. NH). Furthermore, these properties were governed by the metal within the MPcs and further modulated by the heteroatomic linkages. Among these MOFs, CoPc-Cu-O exhibited the highest selectivity toward CO (as a composite with carbon black in 1:1 mass ratio; FECO = 85.00% at the current density of −17.30 mA/cm2) at −0.63 V. Without any conductive additives, CoPc-Cu-O, when employed directly as an electrode material, exhibited excellent performance at a current density of −9.50 mA/cm2 with a FECO of 79.00% [49].

    In 2021, copper tetrahydroxyquinone (Cu-THQ), which is a 2D Cu-based conductive MOF with an average lateral size of 140 nm, was fabricated by Salehi-Khojin et al. for aqueous CO2ER at low overpotentials. It exhibited a high current density of ~173.00 mA/cm2 at −0.45 V vs. RHE, an average FE of ~ 91.00% toward CO production, and an excellent turnover frequency of ~20.82 s−1. When Cu-THQ was employed for CO2ER to yield CO, it exhibited 35 and 25 times higher activities than those of state-of-the-art MOFs and MOF-derived catalysts, respectively. The experimental and DFT results both revealed the existence of reduced Cu (Cu+) during the CO2ER, which reversibly converts into Cu2+ after the reaction [50].

    Jeong et al. designed a Cu-based MOF (Cu-benzene-1,3,5-tricarboxylic acid) along with graphene oxide (GO) for CO2ER. The Cu-MOF/GO electrocatalyst was employed in different supporting electrolytes, including KHCO3/H2O, tetrabutylammonium bromide (TBAB)/DMF, KBr/CH3OH, CH3COOK/CH3OH, TBAB/CH3OH, and tetrabutylammonium perchlorate (TBAP)/ CH3OH (Fig. 2f). After conducting CO2ER at various polarization potentials, the results showed that HCOOH was the main product. The highest concentrations of HCOOH formed were 0.14 mmol/L (at −0.10 V), 66.57 mmol/L (at −0.60 V), 0.27 mmol/L (at −0.50 V), 0.24 mmol/L (at −0.50 V), 0.78 mmol/L (at −0.40 V), and 0.31 mmol/L (at −0.45 V) in KHCO3/H2O, TBAB/DMF, KBr/CH3OH, CH3COOK/CH3OH, TBAB/CH3OH, and TBAP/CH3OH supporting electrolyte systems, respectively. A high FE of 58.00% was obtained in 0.10 mol/L TBAB/DMF, whereas with Cu-MOF alone, the efficiency was 38.00%. Jeong et al. demonstrated that the synergistic effect of GO sheets at a concentration of 3.00 wt% and the Cu+-OH interaction both contributed to the formation of formic acid in various electrolytes [51].

    Recently, the facile electrodeposition of an HKUST-1 thin film on carbon paper was investigated by Wang et al. Their findings indicated that the CO2ER activity was highly affected by the structural reconstruction of HKUST-1. The HKUST-1 film was reconstructed into two structures during the CO2ER process. Three-dimensional nanospheres (numerous small fragments) and 3D nanonetworks (cross-linked nanobelts in different directions) were formed during the CO2ER over time. The 3D nanosphere structure exhibited enhanced catalytic activity, attributed to its abundant active sites, relatively low charge-transfer resistance, and high Cu+/Cu0 ratio. The FEs of C2 product (C2H4 and C2H5OH) were 58.60% at −0.98 V vs. RHE [52].

    Studies have shown that the doping amount of Cu plays a key role in the catalyst structure and catalytic activity for CO2ER. Thus, a series of Cu-doped ZIF-8 catalysts with different Cu2+ doping amounts, obtained by solvothermal synthesis method, was investigated by Iqbal et al. Among the various Cu-ZIF-8 catalysts, Cu30%ZIF-8 exhibited the highest current density of −40.00 mA/cm2 at −2.10 V vs. Ag/AgCl and selectivity toward CH4 and CO compared to previously obtained results on Cu electrodes. The excellent CO2R performance of the as-prepared catalyst was attributed to its crystalline nanostructure with abundant Cu active metal sites and N, micro‑meso dual porosity, and the broad surface area of the ZIF structure [53].

    Recently, Peng et al. investigated semiconductive MOF-Cu3(HITP)2 (HITP = 2,3,6,7,10,11-hexaiminotriphenylene) as a catalyst for the CO2ER. In comparison with the MOF alone, the combination with Ketjen Black would promote the selectivity toward C2H4 with FEs of 60.00%-70.00% in a wide potential range. Multicrystalline Cu nanocrystallites were induced in the reconstructed MOF and stabilized by the conducting support via current shock and charge delocalization. The action mechanism of this catalyst is similar to that of dendrite prevention through conductive scaffolds in metal-ion batteries. DFT calculations revealed that the multifacets and rich grain boundaries both promoted C-C coupling while suppressing the HER [54].

    Above all, Xiao et al. investigated an electrocatalyst derived from a helical Cu-porphyrinic MOF meso‑tetra (4-carboxyphenyl) porphyrin (TCPP) on Cu(OH)2 nanoarrays (H-CuTCPP). The catalysts exhibited an extraordinary excellent acetic acid production whose FE was of 26.10% at −1.60 V vs. Ag/Ag+, which was much higher than that of non-helical CuTCPP whose FE was of 19.80%. The high efficiency was attributed to the efficient utilization of the space derived from helical MOF nanoarrays. It contributed to the large amount of the active catalytic sites. Besides, the results also indicated that H-CuTCPP exhibited stronger CO linear adsorption which was in favor of producing acetic acid [55].

    Attributed to their poor stability, particularly at high negative potentials, Cu-MOF, as precursors, undergo further treatment (pyrolysis or electrochemical reaction) to fabricate more stable and efficient electrocatalysts (Table 4).

    Table 4

    Table 4.  Cu-MOF as electrocatalyst precursors.
    DownLoad: CSV
    2.2.1   Cu-MOF treated by pyrolysis

    Cu-MOF can be pyrolyzed to fabricate more stable and active electrocatalysts. In 2017, Zhao et al. utilized a Cu-based MOF (HKUST-1) as a precursor, and after a facile carbonization process, oxide-derived Cu/carbon (OD Cu/C) catalysts were obtained (Fig. 3a). The catalysts exhibited high selectivity toward alcohols in the CO2ER with a total FE of 45.20%-71.20% at −0.10 ~ −0.70 V vs. RHE. This indicated that the onset potential for C2H5OH formation was approximately −0.10 V vs. RHE, which is among the lowest overpotentials reported. Thus, the enhanced activity and selectivity of the oxide-derived Cu/carbon might be attributed to the synergistic effect between the highly dispersed Cu and the porous carbon matrix [56].

    Figure 3

    Figure 3.  Cu-MOF as precursor under pyrolysis treatment in CO2ER. (a) The synthesis process of oxide-derived Cu/carbon catalysts. Reproduced with permission [56]. Copyright 2017, American Chemical Society. (b) Comparison of paddle-wheel structured HKUST-1 vs. CuAc. Structural investigations of as-fabricated HKUST-1 by SEM. Reproduced with permission [57]. Copyright 2018, American Chemical Society. (c) Illustration of the synthesis procedure of the Cu-MOF/NP catalysts. Reproduced with permission [59]. Copyright 2019, American Chemical Society. (d) Schematic illustration fabrication process for the Cu-NC composite. Reproduced with permission [61]. Copyright 2019, Royal Society of Chemistry. (e) Schematic illustration of the catalyst synthesis process (red oxygen atoms relate to CuO, pink oxygen atoms relate to Cu2O). Structural investigations of H-265. Reproduced with permission [63]. Copyright 2020, Royal Society of Chemistry. (f) Schematic of the procedure to synthesize the Cu-N-C-T catalysts. Cu(BTC)(H2O)3 MOF precursor. Cu-N-C-800 synthesized by pyrolysis of Cu(BTC)(H2O)3 MOF and dicyandiamide at 800 ℃, which favors the CO2ER toward C2H4. Cu-N-C-900 derived from the pyrolysis of Cu(BTC)(H2O)3 MOF and dicyandiamide at 900 ℃, which favors the CO2ER toward CH4. Reproduced with permission [64]. Copyright 2020, American Chemical Society.

    Later, Nam et al. designed a strategy that involved the formation of MOF-regulated Cu clusters, which shifted the CO2ER selectivity toward multiple-carbon-atom-containing products. The symmetrical Cu dimer was distorted into an asymmetric motif by the separation of adjacent benzene tricarboxylate moieties under thermal treatment (Fig. 3b). Cu clusters with low coordination numbers were formed from distorted Cu dimers in HKUST-1 during the CO2ER, thus resulting in a FE of 45.00% for C2H4 [57].

    In addition to Cu-MOF, which are applied alone as precursors, porous porphyrinic triazine frameworks with atomically isolated M-N4 (M = Ni, Cu, Fe and Co) sites were fabricated via pyrolysis for the CO2ER to CO. Among them, Ni single atoms/porphyrinic triazine framework (NiSAs/PTF) exhibited the highest FE of 98.00% at a mild potential of −0.80 V vs. RHE. Moreover, the mechanisms of the catalytic CO2-CO conversion reaction over the different active M-N4 sites were unraveled by the combination of DFT calculations and experiments [58].

    Inspired by previous studies, Liu et al. designed a Cu-MOF-derived nanoparticle as a catalyst for CO2ER (Fig. 3c), in which Cu/Cu2O particles formed a porous octahedral structure containing tunable Cu and CuI catalytic active sites. The CO2ER could be realized with a high current density of 25.15 mA/cm2 at −0.79 V vs. RHE, which is attributed to the high surface area derived from the porous structures with exposed metal ions. Moreover, a new flow electrochemical reactor integrated with a membrane electrode assembly was designed to reduce the applied potential by approximately 200.00 mV and promote the sensitivity of the reactor for identifying and quantifying products [59].

    Guo et al. introduced a method for tuning the selectivity of the CO2ER to yield CO, using a MOF-derived bimetallic oxide catalyst. MOF-derived In-Cu bimetallic oxides were synthesized through the pyrolysis of an In-Cu bimetallic MOF. By altering the In/Cu ratio, the FE of CO could reach up to 92.10% at a current density of ~ 11.20 mA/cm2. Guo et al. believed that the outstanding performance of the catalyst was attributed to the strong CO2 adsorption, high electrochemical surface area, and low charge-transfer resistance of the bimetallic catalyst [60].

    Furthermore, Wei et al. developed a MOF-derived Cu@NC catalyst by the calcination of N-containing benzimidazole-modified Cu-BTC MOFs (BEN-Cu-BTC) at various temperatures (Fig. 3d). The Cu-NC samples exhibited N-species- and content-dependent catalytic activities for the highly selective electroreduction of CO2 to C2 products. The results indicated that the high content of the pyrrolic-N and Cu-N species formed by the pyrolysis process at 400 ℃ can facilitate the production of C2 on the Cu surface, with reaction rates and FEs for C2H4 and C2H5OH of rCH = 5.38 µmol m−2 s−1 (FE = 11.20%) and rC = 8.83 µmol m−2 s−1 (FE = 18.40%) at −1.01 V (vs. RHE), respectively. This further demonstrated that the catalytic activity can be controlled by N-type species, which are affected by the corresponding annealing temperature. Furthermore, the Cu-NC catalysts exhibited excellent electrochemical stability for over 8 h [61].

    Inspired by previous studies, a layer-stacked, bimetallic 2D conjugated MOF (2D c-MOF) was designed with copper phthalocyanine as the ligand (CuN4) and zinc-bis(dihydroxy) complex (ZnO4) as the linker (PcCu-O8-Zn). The PcCu-O8-Zn complex exhibited a high CO selectivity of 88.00%. The catalytic process followed a synergistic mechanism, in which ZnO4 complexes acted as CO2ER catalytic sites, while CuN4 centers promoted the protonation of adsorbed CO2 during the CO2ER [62].

    Furthermore, a catalyst derived from Cu-MOFs (Fig. 3e) for CO2 conversion into C2H4 conversion was proposed by Yao et al., which exhibited enhanced performance because of its porous morphology, complex oxidation states, and strong lattice strain. The Cu@CuxO core-shell structure promoted the activation of CO2, attributed to the formation of Cu+/Cu0 interfaces via stabilized Cu+to facilitate *CO-CO dimerization. This accelerated the conversion of CO2 into C2 products but suppressed the conversion to C1 products. The catalyst exhibited a 51.00% FE for C2H4 with a current density of 150.00 mA/cm2 in a flow cell and a 70.00% FE for C2 products in an H-cell with excellent stability [63].

    Inspired by the significant research interest received by single-atom catalysis, a single-atom Cu catalyst dispersed on N-doped carbon by a N-coordination strategy was proposed by Guan et al. They highlighted that Cu-Nx configurations have a high catalytic activity for CO2ER. Furthermore, the Cu-doping concentration and Cu-Nx configurations were adjusted as a function of the pyrolysis temperature (Fig. 3f). When the Cu concentration reached 4.90%, the distance between neighboring Cu-Nx species decreased, further promoting the C-C coupling and the production of C2H4. However, the concentration of Cu was lower than 2.40%, and the distance between the Cu and Nx species was large, which led to the formation of CH4. DFT results further verified that two CO intermediates binding on two adjacent Cu-N2 sites promoted the production of C2H4, while the isolated Cu-N4, neighboring Cu-N4, and isolated Cu-N2 sites facilitated the formation of CH4 [64].

    Moreover, Cao et al. fabricated a N-doped Cu-NC composite electrocatalyst using a N-rich Cu-BTT MOF precursor (H3BTT = 1,3,5-benzenetris tetrazolate) by pyrolysis. Among the various pyrolysis temperatures applied, 1100 ℃ was the optimal temperature at which Cu-NC (Cu-NC@1100 ℃) exhibited the best catalytic activity for CO (−0.60 V vs. RHE, jCO = 0.40 mA/cm2) and HCOOH production (−0.90 V vs. RHE, jHCOOH = 1.40 mA/cm2). This abnormal activity might be due to the synergistic effect of the low-density-dispersed Cu nanoparticles (~ 27 nm), large porous volume, rich pyrrolic-N and Cu-Nx active sites, and stable CO2 adsorption [65].

    Furthermore, Teng et al. prepared an Fe-Cu-BTT composite by introducing the Fe ion into a microporous N-rich MOF, and they employed it as a precursor. After the pyrolysis process, the FeXCu-NC catalyst exhibited an excellent selectivity toward CO. The FE of the catalyst pyrolyzed at 800 ℃ reached 48.50%, which was attributed to the Brunauer-Emmett-Teller surface area, total pore volume, Fe-Nx sites, and the low density of the Cu nanoparticles in the carbon matrix [66].

    Moreover, a Cu/Cu2O nanocomposite loaded on the surface of carbon derived from ZIF-L coated vertically on GO (Cu-GNC-VL) by pyrolysis was proposed by Zhang et al. The catalyst exhibited an excellent FE of 70.52% for C2H5OH production at a current density of 10.40 mA/cm2 and −0.87 V vs. RHE, which benefited from the synergy between the CO2 asymmetric chemical adsorption on Cu(111) and the suitable kinetics and thermodynamics of the C–C coupling on Cu2O(111) [67].

    Recently, Schuhmann et al. developed a series of CuxOyCz electrocatalysts obtained from a Cu-based MOF as a porous self-sacrificial template. By modifying the GDEs with Polytetrafluoroetylene (PTFE), 25.00–50.00 wt% teflonized GDEs were obtained, which exhibited a FE of 54.00% at −80.00 mA/cm2 for C2 products. The local OH activity of the PTFE-modified GDEs was evaluated using a closely positioned Pt nanoelectrode. The OH/H2O activity ratio increased with the current density because of the locally generated OH ions, irrespective of the PTFE amount [68].

    As mentioned above, single-atom catalysis of CO2ER has been widely investigated. Lu et al. fabricated atomically dispersed Cu species on N-doped carbon nanosheet composite materials (Cu-NC) obtained by MOF derivatization. The Cu-NC materials exhibited better catalytic performance for the synthesis of methyl N-phenylcarbamate in 71.00% yield, compared with the results obtained with traditional Cu bulk electrodes at ambient temperature and normal pressure. The as-prepared catalyst exhibited excellent stability, and its catalytic activity did not decrease even after 10 consecutive cycles. Furthermore, with the Cu-NC material, a variety of amines could be turned into corresponding carbamates smoothly after the reaction [69].

    By incorporating alloys into the structure of Cu-MOFs, it is possible to enhance their catalytic properties. Thus, Zhang et al. fabricated a Cu-In/C bimetallic catalyst by the pyrolysis of Cu-In MOF materials. The Cu-In/C catalyst with a molar ratio of 9:1 exhibited a high CO selectivity of up to 85.00% at −0.75 V vs. RHE, compared with the 3.10% and 10.80% achieved on Cu/C and In/C, respectively. The excellent performance was attributed to the formation of Cu4In derived from In distributed on the surface of Cu nanoparticles, which changed the geometric and electronic structures of the Cu surface. These modifications affected the adsorption properties of the catalyst for *H, CO2, and the intermediates during the CO2ER; thus, the electrochemical conversion of CO2 to CO was accelerated, and hydrogen evolution was suppressed. The large electrochemically active surface area and accelerated charge-transfer and reaction kinetics of this Cu-In bimetallic catalyst also promoted the CO2ER activity [70].

    Apart from the Cu-In/C bimetallic catalyst, a CuZn bimetallic material embedded on carbon electrocatalysts through a one-step MOF carbonization was fabricated and investigated by Rungtaweevoranit et al. The CuZn alloy exhibited the highest improvement (5-fold) in FE for the C2 products (C2H4 and C2H5OH) at −1.00 V vs. RHE, compared with that for its Cu counterpart, far outperforming the catalyst with segregated Cu and Zn phases. This excellent performance was attributed to the influence of Zn on the electronic structure of the Cu sites [71].

    Peng et al. designed a Cu2O/Cu@NC catalyst derived from a Cu-NBDC MOF (a Cu-based MOF synthesized using 2-aminoterephthalic acid (NBDC) as the ligand) by annealing at different temperatures. The results indicated that Cu2O/Cu@NC exhibited better CO2R activity and multielectron product selectivity than Cu2O/Cu@C. The FE increased with increasing temperature, while those of C2H4 and CH4 decreased with increasing temperature. Upon pyrolysis at 400 ℃, the FE of Cu2O/Cu@NC-400 was over 86.00% at −1.40 and −1.60 V vs. RHE, including a FEC2H4 of 20.40% at −1.40 V vs. RHE and FECH4 of 23.90% at −1.60 V vs. RHE. However, the FE for CH4 (−1.60 V vs. RHE) of Cu2O/Cu@C-400 without N doping was only approximately 2.33%, and no C2H4 was detected. These differences in the catalytic behavior were derived from the fact that Cu-N was conducive for the stable adsorption of the *CH2 intermediate during the CO2ER, thus inhibiting the evolution of H2 [72].

    Yang et al. proposed a novel Cu@Cu2O NC electrocatalyst which exhibited excellent performance in converting CO2 into methanol derived from Cu-BTC after pyrolysis at 400 ℃. The optimized catalyst exhibited excellent FE of 45.00% at −0.70 V vs. RHE. It was attributed to the synergistic effect between Cu0 and Cu+ ions on the surface which contributed to the adsorption of CO2 and the following formation of methanol [73].

    A Cu/Bi bimetal catalyst was obtained by calcination of the MOF precursor under inert atmosphere. The Bimetal structure contributed to the stronger adsorption for CO2 intermediate in comparison with Bi/Bi2O3@C without Cu species. The Cu/Bi catalyst exhibited extraordinary selectivity towards HCOOH with FE of about 93.00% at −0.94 V vs. RHE [74].

    Sikdar et al. developed an Ag/Cu bimetallic catalyst via a redox replacement process on the Cu/C catalyst which was fabricated by pyrolysis of Cu-MOF. The catalyst exhibited good performance in producing C2 product (C2H4, C2H6, C2H5OH, C3H7OH) with FE of 21.00% and HCOOH with FE of 40.00%. The results in the work indicated that the formation of adsorbed *CO followed by consumption of CO in the successive cascade steps contributed to yielding HCOOH and C-C coupled products [75].

    2.2.2   Cu-MOF treated by an electrochemical reaction

    Apart from applying the Cu-MOF as a precursor to yield the Cu-NC catalyst via further pyrolysis, Cu-MOFs can also be converted into Cu-nanoparticle-embedded carbon catalysts via an electrochemical reaction. Kim et al. proposed a method using Cu-based MOF-74 as the precursor to produce Cu nanoparticles via an electrochemical reduction process (Fig. 4a). The MOF, due to its porous structure, acted as a template for the fabrication of isolated Cu-nanoparticle clusters with high catalytic activities and high efficiencies for the CO2ER to CH4. The MOF-derived Cu nanoparticles demonstrated high FEs of > 50.00% for CH4 at −1.30 V vs. RHE [76].

    Figure 4

    Figure 4.  Cu-MOF as precursor undergoes electrochemical treatment in CO2ER. (a) Schematic illustration of the hydrothermal synthesis of Cu-MOF-74 and preparation of Cu NPs from Cu-MOF-74 by electroreduction. Reproduced with permission [76]. Copyright 2019, Elsevier. (b) Electrodeposited hollow metal organic framework mediated in-situ synthesis of copper dendrites with abundant active sites show outstanding performance for electroreduction of CO2 to formate. Reproduced with permission [77]. Copyright 2020, John Wiley and Sons. (c) The preparation and schematic CO2R of Cu/Cu2O@NG from MOF-199 and NG. Reproduced with permission [78]. Copyright 2021, Elsevier. (d) Schematic diagram of electrochemical CO2R of HE-Cu and p-Cu. (Catholyte: 0.10 mol/L CO2 bubbled KHCO3, anolyte: 0.50 mol/L NaOH and 25 ℃, CO2R activity and HCOOH selectivity potential: −1.03 V, C2 products selectivity and CH4 selectivity potential: −1.19 V). Reproduced with permission [80]. Copyright 2021, Elsevier.

    Moreover, Han et al. reported a strategy for fabricating heterogeneous electrocatalysts composed of 3D hierarchical Cu dendrites (Fig. 4b). Traditional electrode materials (i.e., Sn and In) often suffer from a low activity for CO2 conversion into formate. Moreover, the CO2ER yields a single product at high current densities and efficiencies. By an in-situ electrosynthesis method, a hollow-structured Cu-MOF film could be obtained in only 5 min. This indicated that more active sites could be exposed using this synthesis strategy, which is beneficial for the reduction of CO2 to formate. Notably, the current density could reach 102.10 mA/cm2 with a selectivity of 98.20% in an ionic-liquid-based electrolyte [77].

    Apart from the in-situ electrosynthesis method, Zin et al. reported a simple MOF-derivatization strategy to fabricate electrocatalysts (Fig. 4c). Using MOF-199 and N-doped graphene (NG) as precursors, they carried out electroreduction to obtain Cu/Cu2O@N-doped graphene (Cu/Cu2O@NG) materials for efficient CO2ER to produce C2-C3 products (C2H4, C2H5OH and n-propanol). A series of Cu/Cu2O@NG materials, particularly Cu/Cu2O@NG-2 with a FE of 56.00% and a current density of 19.00 mA/cm2, were produced to afford C2-C3 products. The excellent performance was ascribed mostly to the synergistic effect between Cu/Cu2O and NG, which resulted in good CO2 adsorption, rapid mass transfer, and abundant active sites [78].

    Moreover, an electrochemical oxidation–reduction method to prepare Cu clusters from MOFs was put forward by Zeng et al. The as-prepared Cu clusters exhibited a FE of 51.20% for CH4 at a current density of > 150.00 mA/cm2. These results suggested that the distinctive CH4 selectivity is attributed to the sub-nanometer size of the derived materials; moreover, the stabilization of the clusters by the residual ligands of the pristine MOF also contributed to the distinctive selectivity [79].

    A Cu-nanosheet-structured catalyst derived from the HKUST-1 MOF by an in-situ electrochemical reaction was proposed by Song et al. (Fig. 4d). The as-prepared electrochemical derivation from HKUST-1catalyst (HE-Cu) exhibited a superior FE of 56.00% compared with that of the pristine Cu foil (p-Cu, FE of 32.30%) at −1.03 V vs. RHE. HE-Cu exhibited a relatively low CO2R onset potential, a higher CO2 adsorption capacity (1.58-fold), and a larger electrochemical active surface area (1.24-fold) compared to those of p-Cu. The FEs of HCOOH and the C2 products for HE-Cu increased by 1.57-fold and 10.6-fold at −1.19 V vs. RHE, respectively, compared with those of p-Cu. The excellent CO2R activity and HCOOH and C2 product selectivity for HE-Cu were attributed to its stepped Cu(211) surfaces, (200) facets, and Cu edge atoms [80].

    2.2.3   Cu-MOFs as active supports

    Apart from direct application as electrocatalysts, Cu-MOFs can also be used as active supports for CO2ER, attributed to the textural properties and active sites of MOFs (Table 5). In 2019, Tan et al. prepared a catalyst by the in-situ etching of a Cu2O sphere with H3BTC to obtain a Cu-MOF shell as a support to further synthesize a Cu2O@Cu-MOF electrocatalyst (Fig. 5a). The catalyst exhibited outstanding performances for the formation of hydrocarbons from CO2, with a FE of 79.40% for CH4 and C2H4. Moreover, the FE for CH4 was determined to be up to 63.20% at −1.71 V [81]. As shown in Fig. 5b, Cu-MOFs used as supports contribute to the productivity of and selectivity toward CH4 and other valuable products.

    Table 5

    Table 5.  Cu-MOF as active support.
    DownLoad: CSV

    Figure 5

    Figure 5.  Cu-based MOF as active supports. (a) Schematic illustration of the process to synthesize Cu2O@Cu-MOF. Reproduced with permission [81]. Copyright 2019, American Chemical Society. (b) FEs of CH4 and C2H4 and the ratio of CH4 to C2H4 for Cu2O@Cu-MOF, Cu-MOF, and Cu2O at −1.71 V vs. RHE in CO2-saturated 0.10 mol/L KHCO3 solution. Reproduced with permission [81]. Copyright 2019, American Chemical Society. (c) Schematic illustration of preparation and CO2ER of M-TCPP@Cu electrode. Reproduced with permission [82]. Copyright 2021, American Chemical Society. (d) Possible reaction route of the CO2ER to C2H4. Reproduced with permission [82]. Copyright 2021, American Chemical Society.

    In 2021, Sun et al. proposed a molecular encapsulation strategy to enrich intermediates to facilitate the electrochemical conversion of CO2 to C2H4. They employed metal (FeCl, Co and Ni) tetrakis(4-carboxyphenyl) porphyrin (M-TCPP) with a Cu-MOF to create a series of metalloporphyrin-decorated Cu catalysts with a coral-like shape (Fig. 5c). As shown in Fig. 5d, the possible reaction route suggests that the M-TCPP in the catalysts provides more CO intermediates to the Cu sites, enhancing the selectivity toward C2H4 with a FE of 33.42% at −1.17 V vs. RHE on the Fe-TCPP@Cu electrode compared to that on the Cu electrode (16.85%, at −1.27 V vs. RHE). In addition, the metalloporphyrin-decorated Cu catalysts exhibited better performance than the physical mixture of Cu-MOFs and metal (FeCl, Co, and Ni) 5,10,15,20-tetraphenylporphyrin (M-TPPs) [82].

    Sun et al. prepared a novel catalyst using CuO NPs anchored on Cu-MOF nanosheets via a facile solvothermal method. The catalyst delivered excellent performances with a FE of 50.00% for C2H4 at −1.10 V vs. RHE and outstanding stability even after long-term experiment. The marvelous performances can be attributed to the interface between CuO and Cu-MOF which could adsorb and activate CO2 molecules [83].

    The atomic layer infiltration (ALI) technique was used to fabricate an HKUST-1–5C-Zn catalyst which used Cu-MOF as active support with highly dispersed Zn atoms. Attributed to the uniform distribution of Zn-O-Zn sites on HKUST-1 which contributed to the CO2 adsorption and formation of COOH*, it exhibited excellent performance in production of CO with a FE of ~80.00% at −1.80 V [84].

    As a promising technology, the CO2ER is essential in solving the global climate and environmental problems caused by the increasing CO2 emissions. As interesting electrochemical catalysts, Cu-based MOFs have been receiving widespread interest from scholars owing to their porous structures and properties, and recently, the research interest and number of studies have been increasing.

    In this minireview, we summarized the recent works on Cu-based MOFs and their derivatives as electrocatalysts for the CO2ER. Cu-MOF catalysts offer several advantages for the CO2ER process, which are attributed to their facile synthetic route, porous structure, and multiple active sites. When Cu-MOFs are directly used as catalysts, the discrete Cu atoms can provide active sites, and organic linkers can be modified to form active sites or charge-transfer agents. Further, their porous structure, formed by metal particles and the organic phase, facilitates the reaction with CO2 or their application as catalyst supports. In addition, carbon-based catalysts with more highly dispersed metal particles can be obtained, using Cu-MOFs or their derivatives as precursors. The metal active sites inherited from Cu-MOFs maintain their catalytic activity and conductivity, facilitating the realization of efficient single-point catalysis. The MOF structure, which is tunable, can be employed to further construct bimetallic or even polymetallic structures, which would enable the facile synthesis of metal alloys and the design of efficient catalysts for the CO2ER.

    However, Cu-MOFs catalysts, similar to traditional MOF catalysts, still suffer from poor chemical stability. Thus, further improvement is required, as it is still difficult to conduct detection and analysis processes after the reaction. The Cu-MOF catalysts obtained by pyrolysis and the electrochemical method have significantly stable structures and exhibit excellent catalytic activities. Future studies are suggested to address these properties in literatures. In addition, new methods and strategies are recommended to fabricate Cu-MOF with more stable structures and excellent chemical activity. For instance, incorporated with highly conductive and active materials (graphene, carbon-based catalysts, etc.) will help enhance the electrochemical activity and stability. Besides, due to the complexity and diversities of the Cu-MOF based structures, machine learning can be applied to design efficient Cu-MOF and its derivatives as electrocatalysts for CO2ER. These properties and recommendations highlight their significant untapped potential for future application in environmental remediation.

    The authors declare no competing financial interest.

    This work was financially supported by the National Natural Science Foundation of China (Nos. 22003041 and 51902204) and Shenzhen Innovative Research Team Program (No. KQTD20190929173914967).


    1. [1]

      C. Costentin, M. Robert, J.M. Savéant, Chem. Soc. Rev. 42 (2013) 2423–2436. doi: 10.1039/C2CS35360A

    2. [2]

      J. Qiao, Y. Liu, F. Hong, J. Zhang, Chem. Soc. Rev. 45 (2014) 631–675.

    3. [3]

      H.R.M. Jhong, S. Ma, P.J. Kenis, Curr. Opin. Chem. Eng. 2 (2013) 191–199. doi: 10.1016/j.coche.2013.03.005

    4. [4]

      Y. Hori, Electrochemical CO2 Reduction on Metal Electrodes, Springer, New York, 2008.

    5. [5]

      H. Dau, C. Limberg, T. Reier, et al., ChemCatChem 2 (2010) 724–761. doi: 10.1002/cctc.201000126

    6. [6]

      S.S. Chadwick, Ref. Serv. Rev. 16 (1988) 31–34.

    7. [7]

      K.P. Kuhl, E.R. Cave, D.N. Abram, T.F. Jaramillo, Energy Environ. Sci. 5 (2012) 7050–7059. doi: 10.1039/c2ee21234j

    8. [8]

      C. Oloman, H. Li, ChemSusChem 1 (2008) 385–391. doi: 10.1002/cssc.200800015

    9. [9]

      S. Nitopi, E. Bertheussen, S.B. Scott, X. Liu, I. Chorkendorff, Chem. Rev. 119 (2019).

    10. [10]

      M.B. Gawande, A. Goswami, F.O.X. Felpin, et al., Chem. Rev. 116 (2016) 3722. doi: 10.1021/acs.chemrev.5b00482

    11. [11]

      Y. Zhou, X.J. Guo, X.J. Li, et al., J. CO2 Util. 37 (2020) 188–194. doi: 10.1016/j.jcou.2019.12.009

    12. [12]

      H. Lv, F. Lv, H. Qin, X. Min, B. Liu, CCS Chem. 3 (2021) 1435–1444.

    13. [13]

      Y. Hori, A. Murata, R. Takahashi, J. Chem. Soc. Faraday Trans. 85 (1989) 2309–2326. doi: 10.1039/f19898502309

    14. [14]

      K.P. Kuhl, T. Hatsukade, E.R. Cave, et al., J. Am. Chem. Soc. 136 (2014) 14107–14113. doi: 10.1021/ja505791r

    15. [15]

      A.A. Peterson, F. Abild-Pedersen, F. Studt, J.R. Osmiumsl, J.K. Nerskov, Energy Environ. Sci. 3 (2010) 1311–1315. doi: 10.1039/c0ee00071j

    16. [16]

      J.M. Spurgeon, B. Kumar, Energy Environ. Sci. 11 (2018) 1536–1551. doi: 10.1039/C8EE00097B

    17. [17]

      C.W. Li, J. Ciston, M.W. Kanan, Nature 508 (2014) 504–507. doi: 10.1038/nature13249

    18. [18]

      I. Takahashi, O. Koga, N. Hoshi, Y. Hori, J. Electroanal. Chem. 533 (2002) 135–143. doi: 10.1016/S0022-0728(02)01081-1

    19. [19]

      Y. Hori, I. Takahashi, O. Koga, N. Hoshi, J. Phys. Chem. B 106 (2002) 15–17. doi: 10.1021/jp013478d

    20. [20]

      C. Hahn, T. Hatsukade, Y.G. Kim, A. Vailionis, T.F. Jaramillo, Proc. Natl. Acad. Sci. 114 (2017) 5918–5923. doi: 10.1073/pnas.1618935114

    21. [21]

      Y. Hori, I. Takahashi, O. Koga, N. Hoshi, J. Mol. Catal. A: Chem. 199 (2003) 39–47. doi: 10.1016/S1381-1169(03)00016-5

    22. [22]

      E. Bertheussen, T.V. Hogg, Y. Abghoui, et al., ACS Energy Lett. 3 (2018) 634–640. doi: 10.1021/acsenergylett.8b00092

    23. [23]

      A. Bagger, W. Ju, A.S. Varela, P. Strasser, J. Rossmeisl, ChemPhysChem 18 (2017) 3266–3273. doi: 10.1002/cphc.201700736

    24. [24]

      H. Ooka, M.C. Figueiredo, M.T.M. Koper, Langmuir 33 (2017) 9307–9313. doi: 10.1021/acs.langmuir.7b00696

    25. [25]

      Y. Hori, A. Murata, Y. Yoshinami, J. Chem. Soc. Faraday Trans. 87 (1991) 125–128. doi: 10.1039/ft9918700125

    26. [26]

      P. Shao, L. Yi, S. Chen, T. Zhou, J. Zhang, J. Energy Chem. 29 (2020) 156–170.

    27. [27]

      A.S. Varela, W. Ju, P. Strasser, Adv. Energy Mater. 8 (2018) 1703614–1703635. doi: 10.1002/aenm.201703614

    28. [28]

      X. Liu, T. Yue, K. Qi, et al., Chin. Chem. Lett. 9 (2019) 2189–2201.

    29. [29]

      D.D. Ma, S.G. Han, C. Cao, et al., Energy Environ. Sci. 14 (2021) 1544–1552. doi: 10.1039/D0EE03731A

    30. [30]

      M.K. Hu, S. Zhou, D.D. Ma, Q.L. Zhu, Appl. Catal. B: Environ. 310 (2022) 121324. doi: 10.1016/j.apcatb.2022.121324

    31. [31]

      C. Cao, Q. Xu, Q.L. Zhu, Chem. Catal. 2 (2022) 693–723. doi: 10.1016/j.checat.2022.01.021

    32. [32]

      B. Zhang, J. Zhang, J. Energy Chem. 26 (2017) 1050–1066. doi: 10.1016/j.jechem.2017.10.011

    33. [33]

      J. Albo, M. Perfecto-Irigaray, G. Beobide, A. Irabien, J. CO2 Util. 33 (2019) 157–165. doi: 10.1016/j.jcou.2019.05.025

    34. [34]

      J. Hu, Y. Liu, J. Liu, C. Gu, AIChE J. 66 (2020) 16835.

    35. [35]

      C.L. Rooney, Y. Wu, D.J. Gallagher, H. Wang, Nat. Sci. 1 (2022) 1–8.

    36. [36]

      K. Kim, P. Wagner, K. Wagner, A.J. Mozer, Fuel 36 (2022) 4653–4676. doi: 10.1021/acs.energyfuels.2c00400

    37. [37]

      R. Angamuthu, P. Byers, M. Lutz, A.L. Spek, E. Bouwman, Science 327 (2010) 313–315. doi: 10.1126/science.1177981

    38. [38]

      R. Hinogami, S. Yotsuhashi, M. Deguchi, ECS Electrochem. Lett. 1 (2012) H17–H19. doi: 10.1149/2.001204eel

    39. [39]

      R.S. Kumar, S.S. Kumar, M.A. Kulandainathan, Electrochem. Commun. 25 (2012) 70–73. doi: 10.1016/j.elecom.2012.09.018

    40. [40]

      T. Maihom, S. Wannakao, B. Boekfa, J. Limtrakul, J. Phyc. Chem. C 117 (2013) 17650–17658. doi: 10.1021/jp405178p

    41. [41]

      J. Albo, D. Vallejo, G. Beobide, ChemSusChem 10 (2017) 1100–1109. doi: 10.1002/cssc.201600693

    42. [42]

      B. Rungtaweevoranit, J. Baek, J.R. Araujo, Nano Lett. 16 (2016) 7645–7649. doi: 10.1021/acs.nanolett.6b03637

    43. [43]

      C.W. Kung, C.O. Audu, A.W. Peters, ACS Energy Lett. 2 (2017) 2394–2401. doi: 10.1021/acsenergylett.7b00621

    44. [44]

      Y.L. Qiu, H.X. Zhong, T.T. Zhang, ACS Appl. Mater. Interfaces 10 (2018) 2480–2489. doi: 10.1021/acsami.7b15255

    45. [45]

      B. An, Z. Li, Y. Song, Nat. Catal. 2 (2019) 709–717. doi: 10.1038/s41929-019-0308-5

    46. [46]

      Z. Weng, Y.S. Wu, M.Y. Wang, Nat. Commun. 9 (2018) 415. doi: 10.1038/s41467-018-02819-7

    47. [47]

      F. Yang, A.L. Chen, P.L. Deng, Chem. Sci. 10 (2019) 7975–7981. doi: 10.1039/C9SC02605C

    48. [48]

      J.X. Wu, S.Z. Hou, X.D. Zhang, Chem. Sci. 10 (2019) 2199–2205. doi: 10.1039/C8SC04344B

    49. [49]

      Z. Meng, J. Luo, W. Li, K.A. Mirica, J. Am. Chem. Soc. 142 (2020) 21656–21669. doi: 10.1021/jacs.0c07041

    50. [50]

      L. Majidi, A. Ahmadiparidari, N.N. Shan, Adv. Mater. 33 (2021) 2004393–2004401. doi: 10.1002/adma.202004393

    51. [51]

      S.M. Hwang, S.Y. Choi, M.H. Youn, ACS Omega 5 (2020) 23919–23930. doi: 10.1021/acsomega.0c03170

    52. [52]

      Y. Han, S. Zhu, S. Xu, ChemElectroChem 8 (2021) 3174–3180. doi: 10.1002/celc.202100942

    53. [53]

      A. Ahmad, N. Iqbal, T. Noor, J. CO2 Util. 48 (2021) 101523. doi: 10.1016/j.jcou.2021.101523

    54. [54]

      H. Sun, L. Chen, L.K. Xiong, Nat. Commun. 12 (2021) 6823–6834. doi: 10.1038/s41467-021-27169-9

    55. [55]

      Y.H. Xiao, Y.X. Zhang, R. Zhai, Z.G. Gu, J. Zhang, Sci. China Mater. 1 (2021) 1–7.

    56. [56]

      K. Zhao, Y. -M. Liu, X. Quan, S. Chen, H.T. Yu, ACS Appl. Mater. Interfaces 9 (2017) 5302–5311. doi: 10.1021/acsami.6b15402

    57. [57]

      D. Nam, O.S. Bushuyev, J. Li, J. Am. Chem. Soc. 140 (2018) 11378–11386. doi: 10.1021/jacs.8b06407

    58. [58]

      Y. Hou, Y.B. Huang, Y.L. Liang, G.L. Chai, R. Cao, CCS Chem. 1 (2019) 384–395. doi: 10.31635/ccschem.019.20190011

    59. [59]

      J. Liu, L. Peng, Y. Zhou, ACS Sustain. Chem. Eng. 7 (2019) 15739–15746. doi: 10.1021/acssuschemeng.9b03892

    60. [60]

      W. Guo, X. Sun, C. Chen, Green Chem. 21 (2019) 503–508. doi: 10.1039/C8GC03261K

    61. [61]

      Y.S. Cheng, X.P. Chu, M. Ling, Catal. Sci. Technol. 9 (2019) 5668–5675. doi: 10.1039/C9CY01131E

    62. [62]

      H. Zhong, M. Ghorbani-Asl, K.H. Ly, J. Zhang, X. Feng, Nat. Commun. 11 (2020) 1409–1419. doi: 10.1038/s41467-020-15141-y

    63. [63]

      K. Yao, Y. Xia, J. Li, J. Mater. Chem. A 8 (2020) 11117–11123. doi: 10.1039/D0TA02395G

    64. [64]

      A.X. Guan, Z. Chen, Y.L. Quan, ACS Energy Lett. 5 (2020) 1044–1053. doi: 10.1021/acsenergylett.0c00018

    65. [65]

      S.M. Cao, H.B. Chen, B.X. Dong, J. Energy Chem. 54 (2021) 555–563. doi: 10.1016/j.jechem.2020.06.038

    66. [66]

      S.M. Cao, H.B. Chen, M.J. Liu, J. CO2 Util. 44 (2021) 101418. doi: 10.1016/j.jcou.2020.101418

    67. [67]

      Y.Y. Zhang, K. Li, M.M. Chen, ACS Appl. Nano Mater. 3 (2020) 257–263. doi: 10.1021/acsanm.9b01935

    68. [68]

      N. Sikdar, J.R.C. Junqueira, S. Dieckhofer, Angew. Chem. Int. Ed. 60 (2021) 23427–23434. doi: 10.1002/anie.202108313

    69. [69]

      S.M. Li, Y. Shi, J.J. Zhang, ChemSusChem 14 (2021) 2050–2055. doi: 10.1002/cssc.202100342

    70. [70]

      X. Ma, J. Tian, M. Wang, Catal. Sci. Technol. 11 (2021) 6096–6102. doi: 10.1039/D1CY00843A

    71. [71]

      S. Juntrapirom, J. Santatiwongchai, A. Watwiangkham, Catal. Sci. Technol. 11 (2021) 8065–8078. doi: 10.1039/D1CY01839F

    72. [72]

      H.D. Jin, L.K. Xiong, X. Zhang, Acta Phys. Chim. Sin. 37 (2021) 2006017.

    73. [73]

      X. Yang, J. Cheng, X. Yang, Chem. Eng. J. 431 (2022) 134171. doi: 10.1016/j.cej.2021.134171

    74. [74]

      Y. Xue, C. Li, X. Zhou, ChemElectroChem 9 (2022) e202101648.

    75. [75]

      N. Sikdar, J.R.C. Junqueira, D. Öhl, Chem. Eur. J. 28 (2022) e202104249.

    76. [76]

      M.K. Kim, H.J. Kim, H. Lim, Y. Kwon, H.M. Jeong, Electrochim. Acta 306 (2019) 28–34. doi: 10.1016/j.electacta.2019.03.101

    77. [77]

      Q.G. Zhu, D.X. Yang, H.Z. Liu, Angew. Chem. Int. Ed. 59 (2020) 8896–8901. doi: 10.1002/anie.202001216

    78. [78]

      W.Y. Zhi, Y.T. Liu, S.L. Shan, J. CO2 Util. 50 (2021) 101594. doi: 10.1016/j.jcou.2021.101594

    79. [79]

      H. Zhang, Y. Yang, Y.X. Liang, ChemSusChem 14 (2021) 1–8. doi: 10.1002/cssc.202002788

    80. [80]

      D. Wang, J.L. Xu, Y. Zhu, Chemosphere 278 (2021) 130408. doi: 10.1016/j.chemosphere.2021.130408

    81. [81]

      X.Y. Tan, C. Yu, C.T. Zhao, ACS Appl. Mater. Interfaces 11 (2019) 9904–9910. doi: 10.1021/acsami.8b19111

    82. [82]

      T. Yan, J.H. Guo, Z.Q. Liu, W.Y. Sun, ACS Appl. Mater. Interfaces 13 (2021) 25937–25945. doi: 10.1021/acsami.1c03557

    83. [83]

      L. Wang, X. Li, L. Hao, Chin. J. Catal. 43 (2022) 1049–1057. doi: 10.1016/S1872-2067(21)63947-5

    84. [84]

      X. Han, Z. Liu, M. Cao, Chem. Mater. 1 (2022) 1–8.

  • Figure 1  Cu-MOF related catalysts engineering for CO2ER.

    Figure 2  Copper ions coordinated by organic ligands, resulting in 3D network structures. (a) Schematic overview of the formation and reactivity of the Cu complexes. Reproduced with permission [37]. Copyright 2010, Science. (b) Optimized structure of the models for Cu-MOF-5, and CO2 adsorbed on Cu-MOF-5 (Ads). Reproduced with permission [40]. Copyright 2013, American Chemical Society. (c) Structural schematic diagram of GDE with Cu3(BTC)2 as CO2 capture agent. Reproduced with permission [44]. Copyright 2018, American Chemical Society. (d) Molecular structures of three Cu-complex materials and their electrocatalytic performance for CO2 reduction. Molecular structures of CuPc, HKUST-1, and [Cu(cyclam)]Cl2. Reproduced with permission [46]. Copyright 2018, Springer Nature. (e) Hierarchical tuning of the performance of electrochemical carbon dioxide reduction using conductive two-dimensional metallophthalocyanine based metal-organic framework. Reproduced with permission [49]. Copyright 2020, American Chemical Society. (f) Cu3(BTC)2 metal-organic framework (Cu-MOF) and graphene oxide composite in electroreduction of CO2 to formic acid. Reproduced with permission [51]. Copyright 2020, American Chemical Society.

    Figure 3  Cu-MOF as precursor under pyrolysis treatment in CO2ER. (a) The synthesis process of oxide-derived Cu/carbon catalysts. Reproduced with permission [56]. Copyright 2017, American Chemical Society. (b) Comparison of paddle-wheel structured HKUST-1 vs. CuAc. Structural investigations of as-fabricated HKUST-1 by SEM. Reproduced with permission [57]. Copyright 2018, American Chemical Society. (c) Illustration of the synthesis procedure of the Cu-MOF/NP catalysts. Reproduced with permission [59]. Copyright 2019, American Chemical Society. (d) Schematic illustration fabrication process for the Cu-NC composite. Reproduced with permission [61]. Copyright 2019, Royal Society of Chemistry. (e) Schematic illustration of the catalyst synthesis process (red oxygen atoms relate to CuO, pink oxygen atoms relate to Cu2O). Structural investigations of H-265. Reproduced with permission [63]. Copyright 2020, Royal Society of Chemistry. (f) Schematic of the procedure to synthesize the Cu-N-C-T catalysts. Cu(BTC)(H2O)3 MOF precursor. Cu-N-C-800 synthesized by pyrolysis of Cu(BTC)(H2O)3 MOF and dicyandiamide at 800 ℃, which favors the CO2ER toward C2H4. Cu-N-C-900 derived from the pyrolysis of Cu(BTC)(H2O)3 MOF and dicyandiamide at 900 ℃, which favors the CO2ER toward CH4. Reproduced with permission [64]. Copyright 2020, American Chemical Society.

    Figure 4  Cu-MOF as precursor undergoes electrochemical treatment in CO2ER. (a) Schematic illustration of the hydrothermal synthesis of Cu-MOF-74 and preparation of Cu NPs from Cu-MOF-74 by electroreduction. Reproduced with permission [76]. Copyright 2019, Elsevier. (b) Electrodeposited hollow metal organic framework mediated in-situ synthesis of copper dendrites with abundant active sites show outstanding performance for electroreduction of CO2 to formate. Reproduced with permission [77]. Copyright 2020, John Wiley and Sons. (c) The preparation and schematic CO2R of Cu/Cu2O@NG from MOF-199 and NG. Reproduced with permission [78]. Copyright 2021, Elsevier. (d) Schematic diagram of electrochemical CO2R of HE-Cu and p-Cu. (Catholyte: 0.10 mol/L CO2 bubbled KHCO3, anolyte: 0.50 mol/L NaOH and 25 ℃, CO2R activity and HCOOH selectivity potential: −1.03 V, C2 products selectivity and CH4 selectivity potential: −1.19 V). Reproduced with permission [80]. Copyright 2021, Elsevier.

    Figure 5  Cu-based MOF as active supports. (a) Schematic illustration of the process to synthesize Cu2O@Cu-MOF. Reproduced with permission [81]. Copyright 2019, American Chemical Society. (b) FEs of CH4 and C2H4 and the ratio of CH4 to C2H4 for Cu2O@Cu-MOF, Cu-MOF, and Cu2O at −1.71 V vs. RHE in CO2-saturated 0.10 mol/L KHCO3 solution. Reproduced with permission [81]. Copyright 2019, American Chemical Society. (c) Schematic illustration of preparation and CO2ER of M-TCPP@Cu electrode. Reproduced with permission [82]. Copyright 2021, American Chemical Society. (d) Possible reaction route of the CO2ER to C2H4. Reproduced with permission [82]. Copyright 2021, American Chemical Society.

    Table 1.  Electrochemical reactions with equilibrium potentials.

    下载: 导出CSV

    Table 2.  Summary of properties of various metals towards the electrochemical reduction of CO2.

    下载: 导出CSV

    Table 3.  Summary of electrochemical reduction of CO2 (CO2ER) performance of Cu-MOF catalysts.

    下载: 导出CSV

    Table 4.  Cu-MOF as electrocatalyst precursors.

    下载: 导出CSV

    Table 5.  Cu-MOF as active support.

    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  12
  • 文章访问数:  400
  • HTML全文浏览量:  40
文章相关
  • 发布日期:  2023-06-15
  • 收稿日期:  2022-05-27
  • 接受日期:  2022-08-16
  • 修回日期:  2022-07-22
  • 网络出版日期:  2022-08-21
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章