Next Article in Journal
Facile Synthesis and Environmental Applications of Noble Metal-Based Catalytic Membrane Reactors
Next Article in Special Issue
Electrocatalytic Reduction of CO2 to C1 Compounds by Zn-Based Monatomic Alloys: A DFT Calculation
Previous Article in Journal
London Disperse Interactions Assist Chiral Induction in the Soai Autoamplifying Reaction Provoked by 1- and 2-Aza[6]helicenes
Previous Article in Special Issue
Efficient CO2 Electroreduction over Silver Hollow Fiber Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Heterogeneous Electroreduction of CO2 on Copper-Based Catalysts

1
Department of Chemistry, Laboratory of Advanced Materials, Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, 2005 Songhu Road, Shanghai 200438, China
2
Radiopharmacy and Molecular Imaging Center, School of Pharmacy, Key Laboratory of Smart Drug Delivery, Ministry of Education, Fudan University, 826 Zhangheng Road, Shanghai 201203, China
*
Author to whom correspondence should be addressed.
Catalysts 2022, 12(8), 860; https://doi.org/10.3390/catal12080860
Submission received: 15 June 2022 / Revised: 25 July 2022 / Accepted: 29 July 2022 / Published: 4 August 2022
(This article belongs to the Special Issue Heterogeneous Electrocatalysts for CO2 Reduction)

Abstract

:
Facing greenhouse effects and the rapid exhaustion of fossil fuel, CO2 electrochemical reduction presents a promising method of environmental protection and energy transformation. Low onset potential, large current density, high faradaic efficiency (FE), and long-time stability are required for industrial production, due to economic costs and energy consumption. This minireview showcases the recent progress in catalyst design and engineering technology in CO2 reduction reaction (CO2RR) on copper based-catalysts. We focus on strategies optimizing the performance of copper-based catalysts, such as single-atom catalysts, doping, surface modification, crystal facet engineering, etc., and reactor design including gas diffusion layer, membrane electrode assembly, etc., in enhancing target electroreduction products including methane, methanol, ethylene, and C2+ oxygenates. The determination of the correlation and the developed technology might be helpful for future applications in the industry.

Graphical Abstract

1. Introduction

In recent years, the increasing usage of fossil fuels has not only accelerated the exhaustion of limited natural energy, but has also led to enormous CO2 emissions, resulting in energy shortages and greenhouse effects. A great deal of research has been undertaken to solve these problems. Among many suggestions, electrochemical CO2 reduction reaction (CO2RR) appears a promising way to produce value-added chemicals and simultaneously decrease CO2 concentration in the atmosphere [1,2]. Powered by intermittent renewable electricity, and using water as the proton donor, CO2RR converts CO2 molecules into a wide range of reduced carbon compounds, including carbon monoxide (CO), methane (CH4), formic acid (HCOOH), methanol (CH3OH), ethylene (C2H4), ethanol (C2H5OH), and acetate (CH3COOH) [3]. The collected products are either directly used as fuels or further converted into value-added chemicals [4,5]. With these advantages, CO2RR has become an important research area and many scientists have tried to put the associated ideas into practice.
The low current density, low faradaic efficiency (FE), and high overpotential occurring with CO2RR remain a challenge for product generation. The disadvantages of the stable C=O bond (about 750 kJ mol−1) [6], and the complex pathway of proton-coupled electron transfer, contribute to the chemical sluggishness of CO2RR [7]. The similar range of redox potentials for different CO2RR products and the hydrogen evolution reaction (HER) contribute to the poor Faradaic efficiency (FE) of specific products [8]. Although many electrochemical catalysts have been designed to enhance the activity and selectivity, and good catalytic performance was obtained, another issue of catalytic stability was revealed in further application. Stability should be maintained for at least 4000 h for industrial application, and more than 20,000 h are required to make production appealing economically [9,10]. However, various factors in electroreduction, including surface reconstruction, flooding behavior, etc., caused unsatisfactory stability only lasting for tens of hours, far from industrial requirements [11,12]. Thus, research was carried out targeting large current density, high FE, low overpotential, and long stability.
Since Hori’s first work in 1985 [13], many efforts in fields of mechanism exploration and catalyst synthesis have been made to disclose the reaction process and to overcome the obstacles mentioned above [14,15,16,17,18,19,20,21,22]. Methods for catalyst synthesis including morphology [23], facet engineering [24,25], metal doping [26,27], and surface modification [28,29] could modulate the binding strength between the intermediates and active centers, thus favoring reaction rates and rearranging product distribution. Copper is one of the few metals that can reduce CO2 to hydrocarbons and alcohols with decent efficiency, and copper-based catalysts have received much attention [30,31]. The uniqueness of Cu as a CO2RR electrocatalyst is explained by the fact that it is the only metal that has negative adsorption energy for *CO and positive adsorption energy for *H [2]. Furthermore, the moderate binding of *CO to Cu provides a balance between activation of CO2 and hydrogenation of *CO, which is the key step towards hydrocarbons. In view of the unique advantage of Cu, we will focus on Cu-based catalysts and strategies to improve their catalytic performance.
In addition to catalyst design, other improvements to the catalytic system have been presented in recent years. Gas diffusion electrodes (GDE) were widely adopted to feed reactant gas directly to the interface between the catalyst and electrolyte, enhancing the CO2 concentration on the catalysts’ surface and promoting the current density. Other methods including changing the CO2 concentration and electrolyte engineering were also reported to enhance catalytic performance [32,33].
In this review, the advances in efficient copper-based catalysts and CO2RR systems in recent years are summarized. First, the reaction mechanisms for generating C1 and C2+ products are briefly introduced. Since low overpotential is required to minimize the energy cost, we next introduce catalyst synthesis and engineering to reduce overpotential. Single-atom catalysts (SACs) with maximum atom utilization efficiency present unique structures for the structure-function relationship. The electronic structures can be finely tuned through changing their near-range coordination environment and long-range interactions, lowering the overpotential of CO2RR products [34]. Then, strategies involving catalyst and system design to achieve high FEs and current densities are summarized. Finally, the degradation factors affecting CO2RR are discussed and methods for stability improvement are presented. Based on the interaction between the catalyst surface and the intermediates, valuable catalyst design and engineering technologies are proposed to achieve low overpotential, high FE, large current density, and long stability of copper-containing catalysts.

2. Electroreduction Pathways

CO2RR activity is carried out through multiple proton and electron transfer steps involving many possible intermediates. The catalytic performance can be tuned and determined largely by key reaction intermediates. Advances in operando spectroscopy and computational techniques provide significant scope for exploring the evolution of surface-bound species and rationalizing a pathway to the desired product. The reaction pathways of CO2 reduction with key intermediates on copper-based catalysts are summarized (Figure 1).
The reaction pathways for the formation of C1 products are relatively simple. C1 products from CO2RR include HCOOH, CO, CH3OH, and CH4. Generally, the formate pathway, which involves the *OCHO intermediate, is considered a dead-end, and the *OCHO is bound to the catalyst surface via one or two oxygen atoms [19,35]. For CO, *COOH is first generated with carbon atom binding to the catalyst surface, followed by *CO formation through dehydrogenation of *COOH, which is finally desorbed from the catalyst to release CO [36]. *CO is the initial key intermediate for further reduction, and the selectivity is determined by the binding energy of *CO. If the binding energy is too weak, the *CO will be desorbed from the catalyst surface to generate CO. If the catalyst binds *CO very strongly, the catalytically active sites are poisoned and the HER becomes dominant. Only the catalysts with the moderate binding energy of *CO can produce muti-electron reduction products [37]. *CO is further reduced to CH4 through the *CHO pathway on these catalysts, forming *CHOH and then *CH2OH and finally branching into two routes (CH3OH and CH4) [38]. It has also been observed that *C is formed by dehydration of *COH and constitutes another pathway to generate CH4 [39]. Furthermore, *CH3O formed by protonation of *CHO is also an important intermediate for CH4 generation [40,41].
The C2+ pathways are much more complex and many controversies still exist. The electroreduction pathway towards the main products includes C2H4, C2H5OH, CH3COOH, and n-C3H7OH. In the formation of the multi-carbon compounds, C–C coupling is the most important step and starts between different C-containing intermediates. *CO dimerization is the most widely accepted step, having been evidenced by theoretical calculation and in situ spectroscopic observation [42,43]. Then, the *CO dimer is further reduced to *OCHCH2, which serves as the key intermediate in determining the selectivity of ethylene and ethanol [44,45]. It is proposed that the *CO dimer follows another pathway to produce CH3COOH via *COCHOH intermediate [3]. Additionally, some studies suggest the possibility of coupling between *CHO and *CO, especially at higher overpotentials. After coupling, the intermediate *COCHO is further reduced to ethanol, ethylene, and acetate [18]. The coupling occurs between *CO and *CH2 via *CO insertion, which produces acetate through multiple proton and electron transfer steps [34,46]. Although it was suggested that n-propanol forms by C–C coupling between CO and C2H4 precursors [47], the elementary steps for C3 production have not yet been well proposed.

3. Advances in CO2 Electroreduction

The recent research into CO2 electroreduction focuses on reducing overpotential, enhancing the FE of the target product, enlarging current density, and maintaining stability. Different synthesis methods of the catalysts and catalysis engineering have been introduced to promote catalytic performance.

3.1. Low Overpotential

Overpotential is defined as the additional potential (above the thermodynamic requirement) to drive a reaction at a specific current density [48]. It can be simulated by theoretical calculations such as density functional theory (DFT). The adsorption energies of the evolving intermediates were calculated to build up the thermodynamic energy diagram for CO2RR, and the overpotential was attributed to the most unfavorable step in the electroreduction pathway [27,40], which was called the rate-determining step (RDS). Although some methods of calculation correction have been developed to assess the kinetic barrier [49,50], the difference in adsorption energy of intermediates was found to influence the overpotential, and other effects such as local CO concentration and charge transfer resistance also played a vital role. Recent research on reducing the onset potential for CO2RR products is shown in Table 1. The potential for the target product is quite different. Low overpotential is required for CO and formate, while high overpotential is needed for deep electroreduction products of methane, ethanol, and ethylene.

3.1.1. Two-Electron Electroreduction Products

Generally, two-electron electroreduction products with a relatively simple process require lower overpotential, compared with other reduction products such as methane and ethylene. Since the properties of active metal centers in SACS can be finely tuned through changing the near-range coordination environment and long-range interactions, SACs could effectively lower the overpotential of CO2RR products. Considering the promotion effect of unsaturated coordination on catalytic activity, Zheng et al. fabricated coordinatively unsaturated single-atom with nitrogen sites anchored on graphene (Cu-N2/CN) [51]. Aberration corrected high-angle annular dark field scanning transmission electron microscopy (AC HAADF-STEM) and X-ray absorption spectroscopy (XAS) demonstrated that the single-atom Cu species were uniformly distributed and coordinated with two N atoms (Figure 2a,b), and inductively coupled plasma mass spectrometry (ICP-MS) determined the Cu content of 1.45 wt% in Cu–N2/GN nanosheets. The coordinatively unsaturated Cu not only promoted the adsorption of CO2 on the catalyst surface, but also accelerated the electron transfer from Cu–N2 sites to *CO2. The Cu–N2/CN catalyst produced CO at a low overpotential, with a maximum FE of 81% at −0.50 V vs. RHE, and onset potential −0.33 V vs. RHE (Table 1). They found that the electronegative N atoms near the coordinatively unsaturated metal center further reduced the energy barrier, lowering the onset potential to −0.30 V vs. RHE [66]. Though we focus on Cu-based catalysts in this review, their effect on reducing the overpotential of CO2RR, especially for two-electron products, is inferior to catalysts with other metals. Zhang et al. prepared singly dispersed FeN5 sites supported on N-doped graphene with an additional axial ligand coordinated to FeN4 through thermal pyrolysis [52]. The AC HAADF-STEM images showed the atomically dispersed Fe and an absence of larger clusters in as-prepared catalysts. In electrochemical tests, the over-coordinated catalyst exhibited a high FE of 97% for CO production at low overpotential. DFT calculation disclosed that the additional coordination number weakened the *CO binding strength of FeN5, and facilitated the desorption of *CO, which changed the RDS and lowered the overpotential. The coordinated adjustment was also an effective way to modify the electronic structures of the metal center and enhance the catalytic performance. Yang et al. synthesized an S-doped Ni-SAC (Ni–N3S) by pyrolysis treatment [53]. The single-Ni-atom catalyst was prepared by pyrolysing a mixture of the amino acid (l-alanine or l-cysteine), melamine, and nickel acetate in argon, with the addition of a sulfur precursor (l-cysteine)The catalytic results showed that the S doping reduced the onset overpotential at only 70 mV, 100 mV lower compared to the Ni–N4 catalyst without S doping. Ni K-edge X-ray absorption near-edge structure (XANES) spectra indicated that the non-centrosymmetric ligand strength of Ni–N3S highly distorted the geometry, which was considered to promote the adsorption of reactants and intermediates, and reduce the overpotential. Cl and N dual-coordinated Mn-SAC ((Cl, N)-Mn/G) synthesized by Zhang et al. improved CO2RR, and the catalytic activity was obtained at low overpotential [67]. The d-band center of (Cl, N)-Mn/G was lower than that of MnN4, thus weakening the strong adsorption for *CO. DFT calculation proved that the energy barrier of the RDS (desorption of *CO) decreased from 1.64 eV for MnN4 to 0.65 eV for (Cl, N)-Mn/G (Figure 2c). As there were four types of N atoms existing in SACs, different types of N atoms coordinated with the metal center and resulted in the variation of catalytic performance. Gu et al. prepared Fe3+-N-C coordinated with pyrrolic N, which displayed a CO partial current density of 94 mA cm−2 at an overpotential of 340 mV [54]. In this catalyst, the pyrrolic N coordination rendered the Fe3+/2+ reduction potential more negative than the Fermi level of the carbon support, leading to the stabilization of the Fe3+ ions in CO2RR conditions. Then, the stabilized Fe3+ induced faster CO2 adsorption and weaker CO absorption, which contributed to the high activity with low overpotential.
Other strategies to boost the catalytic performance of metal centers were also proposed. Pei et al. reported a catalyst with Co-Ni bimetallic sites connected by an N bridge, namely Co–N–Ni/NPCNSs, achieving CO Faradaic efficiency of 96.4% at −0.48 V vs. RHE, with lower overpotential than the catalysts with single metal sites (Figure 2d) [55]. Experimental characterization and DFT calculation suggested that the charge transfer between Co and Ni via the N bridge promoted the formation of *COOH intermediates, thus accelerating the electroreduction. Similarly, anchoring Fe-N sites with the molecular catalyst cobalt phthalocyanine (CoPc) (denoted as CoPc©Fe-N-C) exhibited a low onset potential of −0.13 V vs. RHE (Table 1), with enhanced CO current density and a broadened potential window of high CO FE [56]. Furthermore, the overpotential of CoPc©Fe-N-C was shown to be lower than other samples at the density above 300 mA cm−2, suggesting that the synergistic effect between different metal centers may act as an effective strategy in future applications at large current density. By introducing the electron-withdrawing cyano group, the CoPc anchored on carbon nanotubes exhibited exceptional selectivity and activity for CO at a low overpotential of 0.52 V, ascribing to the promotion of the active sites Co (I) for the reduction of CO2 [68]. For the product of HCOOH, the overpotential was reduced by modifying the metal center. The positively charged single-atom Sn was anchored on N-doped graphene, lowering the energy barriers of CO2 activation and protonation. The reduction potential of −0.74 V vs. SCE corresponding to an overpotential of only 60 mV was observed [57]. Exclusive Bi-N4 sites were fabricated on porous carbon networks by Zhang et al., exhibiting high intrinsic activity for CO2 conversion at a low overpotential of 0.39 V [58].

3.1.2. Multi-Electron Products

In general, Cu-based catalysts have been used in deep CO2 electroreduction. Although Cu SACs produced CH4 and other deep reduction products, the overpotential was relatively high [69,70]. Xu et al., found that Cu metallic clusters were synthesized through reversible transformation by an amalgamated Cu–Li method, converting the CO2 to ethanol with an extraordinary 91% FE at −0.7 V vs. RHE (Figure 3a) [62]. Operando XAS identified the transformation from atomically dispersed Cu atoms to Cun clusters (n = 3 and 4) under CO2RR conditions. According to the operando measurements and DFT calculation, the proposed reaction mechanism is illustrated in Figure 3b. Under open-circuit potential, the uniformly dispersed Cu ion was anchored by oxygen atoms from the hydroxyl group and water. After imposing a negative bias potential, the Cu2+ was reduced to Cu3 or Cu4 ligated by the surface hydroxyl group. Zhu et al., reported that a 3D dendritic cooper-cuprous oxide composite catalyst (Cu–Cu2O/Cu) was obtained via in situ growth and decomposition of the corresponding complex film [65]. The catalyst formed in situ exhibited a high C2 FE of 80% under a low applied potential of −0.4 V vs. RHE, attributing to the low charge transfer resistance and favorable C–C coupling.
Among the evolving intermediates of CO2RR, *CO was an important intermediate and its binding strength and surface coverage influenced the overpotential for deep reduction products. The electric field was considered to stabilize adsorption species such as *CO, and to promote C–C coupling [71]. Fan et al. modified commercial Cu nanoparticles (NPs) with water-insoluble organosuperbases to enhance C2+ production [60]. The modifier with high pKa would be protonated to a “proton sponge”, presenting a positive charge under CO2RR. The positive species near the double-layer surface induced a large local electric field to stabilize *CO and promote C–C coupling, decreasing the onset potential of C2H4 from −0.57 V to −0.43 V. Furthermore, overpotential is reduced at the large current density, revealing the future potential for realizing the large-scale electrosynthesis of C2+ compounds from CO2. Morales-Guio et al. fabricated the copper catalyst with tandem gold to increase the local CO concentration on the surface, and the reduction products (>2e) were generated more than 100 times faster than on Cu catalyst. The detection overpotential for alcohol was reduced by 260 mV (Figure 3c,d) [63]. Theoretical calculation suggested that the estimated local CO concentration on the tandem catalysts’ surface was above solubility and much larger than that on Cu, due to the CO spillover from gold, contributing to the enhanced performance at low overpotential. Similar work was reported by Li et al, who functionalized the Cu surface with porphyrin-based metallic complexes for CO2RR, and a low onset potential of −0.42 V vs. RHE for ethanol was observed [64].
Reducing the energy barrier to the following steps of *CO was taken into account to lower the overpotential for multi-electron electroreduction products. Nellaiappan et al. fabricated a nanocrystalline high-entropy alloy (HEA: AuAgPtPdCu) to convert CO2 into gaseous products [59]. The main products of CH4 and C2H4 were obtained at a low applied potential of −0.3 V vs. RHE, with FEs of 38.0% and 29.5%, respectively. DFT calculation suggested that the step *OCH3 to *O was the RDS, and the HEA catalyst showed excellent performance at low overpotential. Ma et al. modified Cu with halogen and found that the onset potential of C2H4 decreased obviously as the electronegativity of the halogen increased (Figure 3e) [61]. Although the enhancement of CO adsorption may play a role due to the presence of Cu+, isotopic experiments and DFT calculation coupled with previous research indicated that electronegative fluorine enhanced the dissociation of H2O into *H, thus lowering the barrier of hydrogenation of *CO, the RDS of CO2RR to C2H4 on Cu given by DFT. In general, the overpotentials of multi-electron electroreduction products are difficult to reduce, due to the sluggish kinetics of the evolving reaction following *CO. Many researchers have recently focused on FEs and current densities and have made great progress, but the overpotential of CO2RR should be paid more attention to reduce the cost of electricity to drive the reaction, to make CO2RR competitive economically. Though researchers achieved limited activities at low overpotential, these provided insights into the mechanisms of CO2RR and gave valuable information to support electrocatalyst design with high effciency at low overpotential.

3.2. High Faradaic Efficiency

Faradaic efficiency (FE) is the ratio between the amount of product detected and the amount of product theoretically formed, based on the charge passed through the circuit, and it expresses the selectivity of an electrocatalytic reaction. A high FE for CO2RR is desirable to minimize the total current required for a target production rate and save the downstream separation costs. In the following section, we summarize recent strategies to raise Fes, according to the products shown in Table 2. High FE was achieved for CO and formate, even at a large current density scale, but it remains a challenge for deep electroreduction products. GDE plays a positive role in achieving current density as well as FE for the target products.

3.2.1. CO and Formate

High FEs for CO and formate in CO2RR were comparatively easy to reach, even in early studies. Several recent studies have addressed catalyst design for these two-electron electroreduction products. It is noteworthy that non-copper catalysts such as Ag, Au, Sn, etc., are inclined to obtain two-electron electroreduction products.
Metal-based electrocatalysts usually possess high conductivity and easily facilitate electron transfer to the active surface sites, thus achieving good performance. Liu et al. synthesized five-fold twinned Ag nanowires through a facile bromide-mediated polyol method, which efficiently produced CO and achieved a maximum FE of 99.3% at −0.956 V vs. RHE [72]. The DFT calculation revealed that the improved performance over the catalysts originated from the diameter and length effects. as well as the special Ag (100) enclosed nanostructure and the increased edge-to-corner ratio. Gao et al. prepared a series of cadmium sulfide (CdS) nanostructures using a simple microwave heating strategy [83]. The enhancement of the electric field at the tips of CdS nanoneedles increased the neighbouring K+ concentration, which interacted with CO2 via a non-covalent feature and facilitated the reaction. As a result, the CO formation FE for CdS nanoneedles (91.1%) greatly surpassed that for CdS nanorods (42.4%) and nanoparticles (25.1%). Fan et al. reported a chemical lithium tuning method to create abundant active grain boundaries (GB) on Bi nanoparticles [74]. Due to the small particle size and enriched GBs, the catalysts achieved a maximal formate FE of 97%, and maintained above 90% over a wide electrochemical window from −0.72 V to −1.05 V.
As a promising alternative to metal catalysts, carbon-based materials also showed high FEs for CO and formate production. Ren et al. reported that the isolated diatomic Ni–Fe sites supported on N-doped carbon achieved a CO FE above 90% over a wide potential range [84]. Theoretical DFT studies revealed that the diatomic site underwent a distinct structural change after adsorption of CO2, which reduced the energy barrier for the formation of *COOH and desorption of CO. Zhang et al. observed a 90.9% FE for formate production in Bi nanorods with nitrogen-doped carbon nanotubes (Bi-NRs@NCNTs) [75]. The favorable nanocapillary and nanoconfinement effects of hollow Bi-NRs@NCNTs synergistically facilitated the mass transfer, adsorption, and concentration of reactant CO2 molecules onto the active sites, promoting the formation of HCOO. Chen et al. reported a 2D/0D composite catalyst composed of bismuth oxide nanosheets and nitrogen-doped graphene quantum dots (Bi2O3-NGQDs) for the efficient electrochemical reduction of CO2 to formate [85]. The electrochemical tests demonstrated that the Bi2O3-NGQDs exhibited almost 100% FE for formate at -0.9 V vs. RHE. The DFT calculation disclosed that the high FE of the catalyst was attributed to the increased adsorption energy of *CO2 and *OCHO intermediate after combination with NGQDs. Therefore, non-copper catalysts play a significant role in the generation of two-electron electroreduction products of CO and formate.

3.2.2. Methane

To raise the FE of CH4, C–C coupling should be prohibited to produce C2 products. SAC was shown to be an effective strategy for producing CH4 exclusively, because the large distance between isolated Cu sites resulted in poor C–C coupling. Guan et al. demonstrated this effect in experiments by tuning the Cu concentration in SAC [69]. Through changing the calcination temperature, Cu-SACs with different Cu concentrations were fabricated. A high Cu concentration of 4.9%mol was obtained, and the distance of neighboring Cu sites enabled C–C coupling to produce C2H4, with FE of CH4 of only 13.9%. After raising the calcination temperature and lowering the Cu concentration to 2.4%mol, the distance became so large that C–C coupling was prohibited, which resulted in a high FE of 38.6% for CH4 production. A regular pattern was seen in Figure 4a. With the temperature elevated, the Cu concentration decreased, decreasing the current density of C2H4 and continuously increasing the ratio of CH4 to C2H4. To further enhance the FE of methane on SAC, Cai et al. fabricated carbon-dots-based SAC margined with CuN2O2 sites, using the partial-carbonization method [70]. The introduction of oxygen ligands modified the electronic structures of the center atoms, leading to a high FE of 78% for methane production. SACs based on other metals also effectively enhanced the FE of methane. Han et al. chose Zn as the metal center and anchored it onto microporous N-doped carbon [76]. When Zn was the metal center, the formation route of CH4 involved the formation of *OCHO instead of *COOH, which was crucial for CO generation, and the CO2-to-CO pathway was largely suppressed. As a result, the catalyst exhibited a high FE of 85%, and a partial current density of 31.8 mA cm−2 for methane at a potential of −1.8 V vs. SCE.
Catalysts with single Cu sites promoted the FE of methane. Wang et al. designed a single atomic Cu substituted CeO2 catalyst with multiple oxygen vacancies to optimize the CO2RR to CH4 [87]. The single-atomic Cu on CeO2 (110) stabilized three oxygen vacancies around each Cu site, which was evidenced by X-ray photoelectron spectroscopy (XPS) and XAS. These features were beneficial for CO2 adsorption and activation, enabling an FE of 58% towards methane. Zhang et al. fabricated a conductive metal-organic framework with Cu-O4 sites, which showed a high FE of 80% and a current density of 203 mA·cm−2 when tested in the flow cell (Figure 4b) [77]. Due to the smaller charge transfer resistance and favorable adsorption of *H on the O atom, the Cu-O4 site was feasible to accept an (H+/e) pair reduced to +1 valence, which was beneficial to the CO2 adsorption and reduction. Recently, our group fabricated ultrathin rhombohedral-phase CuGaO2 nanosheets by an induced anisotropic growth strategy, to convert CO2 to CH4 efficiently [41]. On the surface of CuGaO2 nanosheets, the (001) surface was highly exposed, which possessed abundant single-interlayered Cu(Ⅰ) with a large Cu–Cu distance of 3.013Å, and prohibited C2H4 formation. The DFT calculation indicated that the CO pathway was energetically unfavorable due to the high CO desorption barrier. The DFT suggested that water dissociation to generate the first adsorbed *H on CuGaO2 was comparatively easier than CuAlO2. Considering that the pathway towards CH4 involved multiple hydrogenation steps, the simpler process of water dissociation enhanced CH4 production. With these features, the CuGaO2 nanosheets exhibited efficient CO2-to-CH4 electroreduction with FE of 71.7% at a high CH4 partial current density of 717 mA cm−2, substantially outperforming the previous catalysts.
While the mechanism of single-site Cu catalysts prohibited C–C coupling, lowering the intermediate concentration also decreased C–C coupling and enhanced the FE towards methane. Wang et al. directly changed the partial pressure of the CO2 gas stream towards Cu and observed that methane FEs in dilute CO2 increased at high reaction rates [32]. According to Henty’s law, local CO2 concentration was proportional to the partial pressure, and the local *CO2 coverage was positively related to CO2 concentration.Therefore, tuning the CO2 partial pressure effectively changed the *CO2 coverage, which further decreased *CO coverage according to DFT studies. Finally, the value of (ΔECHO-ΔEOCCOH), considered as the descriptor for the *CHO vs. C–C coupling, decreased when *CO coverage was reduced, thus comparatively enhancing the methane FE. In experiments, the CO2RR performance at various CO2 concentrations (25%, 50%, 75%, 100%) was evaluated by tuning the volume ratios of CO2 to N2 in the gas streams. At high current densities, the ratio between CH4 and C2H4 increased with decreasing CO2 concentration (Figure 4c), in accord with the theoretic prediction mentioned above. Although low CO2 concentration was unfavorable for CO2RR, the CO2 concentration of 75% exhibited the highest methane FE of 48%, with a partial current density of 108 mA·cm−2 at −1.416 V vs. RHE, obviously higher than the pure CO2 gas stream which mainly produced ethylene. Xiong et al. fabricated yolk-shell nanocell structures which involved an Ag core and a Cu2O shell resembling a tandem nanoreactor [88]. By fixing the Ag core and changing the Cu2O envelope size, the CO flux arriving at the shell was regulated, which further modulated the *CO coverage and *H adsorption and facilitated CH4 production. As a result, the best catalyst among structures with different envelope sizes exhibited a high CH4 Faraday efficiency of 74% and a partial current density of 178 mA·cm−2 at −1.2 V vs. RHE.
The electrolyte component changes the selectivity of CO2RR, and the organic additives in the aqueous electrolyte have been reported to enhance the FE of methane, partly attributed to the increase of CO2 solubility. Qui et al. added methyl carbamate (MC) containing an –NH2 group into 0.5 M aqueous solution of NaHCO3, with Cu foil as the catalyst [33]. The –NH2 group in MC is prone to generate an –NH3 group in an aqueous solution, and the affinity effect of the Lewis acid (–NH3+) with the Lewis base (CO2) increased the solubility of CO2. Furthermore, the generated –NH3+ group promoted the adsorption of *CO and *CHO and stabilized the *CHO intermediate via the facilitation of partial electron transfer from Cu to C, and the formation of H-bonds. As a result, the FE of CH4 increased by 30% with the addition of MC. Organic additives influenced the morphology of catalysts. Han et al. used ethylenediamine tetramethylenephosphonic acid (EDTMPA) to promote CO2 electroreduction to methane [89]. Experimental and theoretical studies have shown that EDTMPA molecules were preferentially adsorbed on Cu (110) during the CO2RR, which not only induced the selective generation of Cu (110) to shave an inherently high *CO binding strength, but also formed a local environment that promoted proton transfer from water to the catalyst surface. After adding EDTMPA, the FE of methane increased to 61% and 64% in H cell and flow cell, respectively, demonstrating the effect of EDTMPA in CH4 promotion at current density with a different order of magnitude.

3.2.3. Methanol

As the most widely used materials in CO2RRM, Cu based materials were also effective for enhancing the FE of methanol. Yang et al. reported that Cu selenide NPs exhibited outstanding performance for CO2RR to methanol [78]. At a low overpotential of 285 mV, the current density was up to 41.5 mA cm−2, and FE reached 77.6%. The control experiment suggested that Se in the catalysts was crucial for efficient methanol production, compared with Cu oxide and sulfide, and the Se atom in the catalysts was unsaturated, further enhancing the performance. Zhao et al. fabricated single atom immobilized MXene for methanol production by selective etching of hybrid A layers (Al and Cu) in quaternary MAX phase (Ti3(Al1−xCux)C2), illustrated in Figure 4d [86]. According to XAS and DFT calculation, this etching process resulted in Cu atoms with unsaturated electronic structures, which delivered a low energy barrier for the RDS (formation of *CHO), thus achieving a high FE of 59.1% towards methanol, with good stability. Li et al. studied the influence of dual doping in a Cu2O/Cu host on the CO2-to-methanol process, using cation (Ag, Au, Zn, Cd) and anion (S, Se, I) [79]. The results indicated the Ag and S dual doping exhibited good performance, achieving a methanol FE of 67.4% at potential of -1.18 V vs. RHE, with a high current density of 122.7 mA·cm−2 in an ionic liquid/H2O electrolyte (Figure 4e). The DFT calculation indicated that the anion S modified the electronic structures of neighboring Cu atoms, which reduced the energy barrier to *CHO formation, and the cation Ag mainly increased the reaction barrier to HER.
Other non-Cu catalysts have been reported to obtain high FE for methanol. Zhang et al. designed ultrathin Pd nanosheets partially capped by SnO2 NPs to selectively produce methanol [90]. The combination of Pd and SnO2 not only improved the adsorption capacity of CO2 on the catalyst surface, but also weakened CO poisoning on Pd, facilitating the formation pathway of CH3OH. By tuning the Pd/Sn ratio, the 2D hierarchical structure achieved a maximum FE of 54.8% for methanol at −0.24 V vs. RHE. Ji et al. reported the Fe2P2S6 nanosheets efficiently catalyzed the CO2RR to alcohols in aqueous electrolyte, with a total FEmethanol+ethanol of 88.3%, and 65.2% for methanol FE [91]. An Fe atom on the Fe2P2S6 surface was regarded as the active site for alcohol formation, according to the DFT calculation. In addition to metal catalysts, other catalysts without metal have also been reported. For example, boron phosphine NPs were demonstrated to efficiently convert CO2 to methanol, achieving a high FE of 92.0% for CH3OH at −0.5 V vs. RHE [92]. Based on DFT calculations, it was demonstrated that the outermost B sites strongly adsorbed CO2, after which P donated electrons to B, which synergistically activated CO2. In addition, the desorption of CO and OCH2 were energetically unfavorable, which contributed to the high selectivity of the CO2-to-CH3OH conversion process.

3.2.4. Ethylene

The RDS of C2H4 formation was C–C coupling between *CO or *CHO. After C–C coupling, C2+ products were generated. Therefore, facilitating C–C coupling was crucial for raising the FE of ethylene in electrochemical CO2RR [18,93]. Hori et al. studied CO2RR on Cu single crystal and showed that the product distribution was dependent on the facet of the catalyst surface [94]. Specifically, CH4 and C2H4 were the main products on Cu (111) and Cu (100), respectively, while Cu (110) suppressed FEs of CH4 and C2H4 but favored the formation of C2+ oxygenates. Facet tuning became an effective strategy for promoting electrochemical CO2RR. Recently, Gao et al. prepared Cu2O NPs with different crystal facets and found that the Cu2O NPs enclosed with both (111) and (100) facets exhibited excellent performance, with a maximum FE of ethylene of 59% [24]. The XPS indicated that these Cu2O NPs were well preserved after the stability test, with only tiny NPs on the surfaces ascribed to the Cu NPs formed during the electroreduction process. The DFT calculation suggested that the surface of t-Cu2O combined the strong CO adsorption capacity of Cu2O (100) and the easy C2H4 desorption of Cu2O (111), thus promoting ethylene production. Besides, the difference from the Fermi level of Cu2O facilitated the charge transfer between the two facets, and further promoted the multi-electron kinetics involved in ethylene production. Previous studies on facet dependence were mostly performed in H-cells with slow reaction rate, while Gregorio et al. demonstrated that facet-dependent selectivity was retained in gas-fed flow cells at large current densities [95], which suggests that the facet dependence found in the lab could provide useful information for future applications at large current densities. Furthermore, hydrogen production was suppressed due to the high alkaline conditions in the flow cell, which further increased the conversion of CO2RR. In addition to these low-index surfaces, high-index surfaces with abundant active sites changed the product distribution. Zhong et al. prepared the Cu catalyst derived from Cu(OH)2, CuO, and Cu2O, respectively, and the Cu(OH)2-derived catalyst (Cu(OH)2-D/Cu) exhibited the best performance [25]. Transmission electron microscopy (TEM) images showed the Cu(OH)2-D/Cu with a high-index stepped surface, such as Cu(310) = 3(100) × (110) and Cu(210) = 2(110) × (100), composed of Cu(100) and (110) (Figure 5a). In situ spectroscopy revealed the Cu(OH)2-D/Cu with a high-index facet promoted the adsorption of CO intermediate, while the DFT calculation demonstrated that the activation energy of CO dimerization on Cu(210) and Cu(310) was lower than that on the low-index facet. Performance testing in H-cell showed that the Cu(OH)2-D/Cu exhibited higher selectivity for C2+ products than CuO-D/Cu and Cu2O-D/Cu, achieving ≈59% at −0.98 V, compared with ≈45% for CuO-D/Cu and about 21% for Cu2O-D/Cu at ca. −1.03 V and −1.23 V, respectively. To overcome the solubility limitation of CO2 in the aqueous solution and meet the requirements for industry application, the performance of Cu(OH)2-D/Cu was assessed with a flow-cell electrolyzer at 1 M KOH. Resulting from the alkaline environment, the FE of H2 was suppressed, and the Cu(OH)2-D/Cu catalyst achieved a high FE of 87% for C2+ product, with 58% for ethylene at −0.54 V vs. RHE at total current density of 250 mA·cm−2. The strong performance of Cu(OH)2-D/Cu at current densities with different orders of magnitude suggests that facet engineering is an effective strategy to enhance the FEs of C2+ products in future applications.
Surface modification by organic addiction was proved effective to enhance the FE of ethylene. Wakerley et al. modified the Cu dendrites with 1-octadecanethiol to tune the selectivity of CO2RR [97]. Contact angle measurements illustrated a drastic improvement in hydrophobicity after modification, with contact angles increasing from 17° to 153°.This hydrophobic feature coupled with hierarchical structures resulted in gas trapping at the interface between the catalyst surface and electrolyte, which facilitated the mass transport of CO2 and decreased that of H+. Therefore, the FEs for electrochemical CO2RR were obviously enhanced after modification, with methane going from 0% to 7%, ethylene from 9% to 56%, and ethanol from 4% to 17%, but with slightly lower current density. Wei et al. coated the Cu surface with a 50 nm thick film of polyaniline (PANI), enhancing the selectivity of CO2RR for both Cu foil and Cu NPs [28]. The hydrophobic nature of the polymer suppressed the HER, and the in situ spectroscopy revealed that the Cu/PANI interface improved the coverage of *CO and strengthened interactions with *CO, leading to a higher probability of C–C coupling. Hence, when PANI coating was applied to Cu nanoparticles, the FE of C2+ hydrocarbon reached 80% and the FE of ethylene over 40%. The modification of the function group of the organic additive further enhanced the selectivity of C2H4. Li et al. investigated the influence of eleven kinds of molecules derived by the electro-dimerization of arylpyridiniums on Cu catalysts [29]. The in situ Raman spectroscopic interrogation indicated that different functional groups caused differences in the electron-donating abilities of molecules, and further caused the change of the relative populations of atop-bound CO (COatop) and bridge-bound CO (CObridge). Combined with CO2RR performance, a volcano-shaped trend between the FE of ethylene and the ratio of Cuatop and Cubridge was found, with the ratio near 1 obtaining the highest C2H4 selectivity (Figure 5b), suggesting that C–C coupling between Cuatop and Cubridge was favorable. Accordingly, the DFT calculation indicated that in the presence of the organic additive, the adsorption of CO on atop sites increased compared with that on bridge sites. Furthermore, the CO dimerization between Cuatop and Cubridge possessed the lowest energy barrier, according to with the experimental results. To further raise the FE of ethylene, the researchers synthesized the molecule with optimal electro-donating properties and many electro-donating N atoms. Hence, the ethylene FE on this catalyst (named Cu-12) reached 72% at −0.83 V, with a current density of 232 mA·cm−2 (Figure 5c), surpassing Cu and the eleven catalysts with other additives.
While previous research demonstrated that Cu-based alloy enhanced C2H4 production, various guest elements and active sites made the discovery of the efficient alloy catalysts time-consuming. Recently, Zhong et al. used DFT calculation in combination with active machine learning and found that Cu-Al electrocatalysts efficiently reduced CO2 to C2H4 [26]. 12,229 surfaces and 228,969 adsorption sites from 244 different copper-containing intermetallic crystals were obtained from The Materials Project, and then a subset of these sites was assessed using DFT simulation to calculate their CO adsorption energies (ΔEco). These data were used to train a machine learning model which was employed to predict ΔEco on the adsorption sites. Combined with volcano scaling relations which have optimal ΔEco near −0.67 eV, adsorption sites with optimal ΔEco were predicted and then simulated using DFT to provide additional training data. In total, about 4000 DFT simulations were carried out to yield a set of candidates for experimental testing. Among these, Cu-Al exhibited the highest abundance of sites and site types with near-optimal ΔEco, illustrated in Figure 5d. Therefore, Cu-Al catalysts were prepared and tested experimentally, and the de-alloyed Cu-Al catalysts on polytetrafluoroethylene (PTFE) substrates showed excellent performance. Under electrochemical CO2RR conditions, the FE of ethylene achieved 80% at a current density of 400 mA·cm−2. This research illustrated that computation and machine learning were useful tools to guide the experimental exploration of efficient CO2RR catalysts, which may be widely used in the future.

3.2.5. C2+ Oxygenates

As mentioned above, ethylene was usually the major product after C–C coupling, while C2+ oxygenates, including C2H5OH, CH3COO, and n-C3H7OH, generally accounted for a small part of the C2+ product. Therefore, if C2+ oxygenates were the target products in the electrocatalytic CO2RR, elaborate catalyst design was required to improve the FE of C2+ oxygenates.
It is usually believed that bimetallic catalysts composed of Cu and a metal capable of catalyzing CO2 to CO, such as Ag, Au, or Zn, would enhance the selectivity of C2+ oxygenates in the electrocatalytic CO2RR [80,98]. Among these, the Cu-Ag catalysts have shown a great potential for obtaining high FEs of C2+ oxygenates. Li et al. fabricated bimetallic Cu-Ag catalysts with different ratios via co-sputtering [80]. The X-ray diffraction (XRD) and XAS characterization revealed that Ag and Cu were in an alloy phase. For the good catalysts (Ag0.14/Cu0.86), the FE of ethanol reached 41% with a current density of 250 mA cm−2 at −0.67 V vs. RHE, higher than the pure copper control (29%). According to the DFT calculation, the introduction of Ag disrupted the uniform Cu sites, resulting in a broader peak of CO adsorption during in situ Raman spectrum analysis, and destabilized the ethylene intermediates, thus promoting ethanol formation. Lv et al. developed a Cu3Ag1 electrocatalyst by galvanic replacement of an electrodeposited Cu matrix [27]. From XPS and XAS results, it was concluded that the introduction of Ag in this catalyst facilitated the electron transfer from Cu atoms to Ag atoms, leading to electron deficiency in the Cu matrix. DFT calculation revealed that an alcohol-pathway intermediate like *OCH2CH3 was more stabilized on electron-deficient Cu atoms than on pure Cu. As a result, the Cu3Ag1 catalyst enabled a 63% FE for CO2-to-alcohol production at −0.95 V vs. RHE. For acetate production, Wang et al. synthesized ultrasmall CunAgm bimetallic NPs to enhance the acetate FE [81]. They fabricated a series of these bimetallic NPs through electroreduction on electrochemically polymerized poly-Fe(vbpy)3(PF6)2 film electrodes on glassy carbon. It was found that the FE of CH3COO increased with the addition of Ag to the nanoparticles, reaching a maximum of 5.5% for Cu2Ag3 catalysts (Figure 5e). Through further optimization of the reaction conditions, the FE of acetate reached 21.2% at −1.33 V vs. RHE in a −0.5 M KHCO3 solution with 8 ppm benzotriazole at 0 ℃.
Combining Cu catalysts with carbon-based materials raises the FE of C2+ oxygenate. Chen et al. used N-doped graphene quantum dots (NGQ) as the second component to modify CuO-derived nanorod (Cu-nr) [99]. Characterizations such as TEM and XAS indicated the uniform dispersion of the NGQ on the Cu-nr. During the electrochemical reduction, the structures were retained, and the FEs of C2+ alcohols reached 52.4% with a current density of 282 mA cm−2 at −0.9 V vs. RHE, obviously higher than bare NGQ and Cu-nr. Control experiments and DFT studies showed that the combination of materials exhibited the synergistic effect and stabilized *C2H3O, the intermediate to C2+ alcohols, thus enhancing activity and selectivity. Wang et al. reported that confinement using a nitrogen-doped carbon (N-C) layer on Cu catalysts promoted ethanol production [100]. They fabricated N-C/Cu by sputter deposition of a layer of N-C on the surface of sputtered Cu NPs, achieving an ethanol FE of 52% at −0.68 V vs. RHE. The different plots of electron density generated by DFT showed that the N-C layer was beneficial for electron transfer to adsorbed *CO on Cu, compared to the carbon layer, facilitating the generation of the C–C coupled intermediate. Furthermore, the capping layer stabilized the C-O bond of *CHCOH, suppressing the deoxygenation process and favoring the ethanol pathway instead of ethylene. The carbon-based materials not only stabilized the reaction intermediates, but also influenced the dispersion of active sites. Using reversible transformation, the Cu SAC synthesized by an amalgamated Cu–Li method converted the CO2 to ethanol with an extraordinary 91% FE at −0.7 V vs. RHE (Figure 3a) [62]. During the process of catalyst synthesis, the carbon surface formed hydroxyl groups due to the highly basic environment caused by the dissolved LiOH. These hydroxyl groups were found to play a crucial role in stabilizing the transiently formed Cu clusters, which were responsible for the high selectivity towards ethanol.
For C3 products, it remains difficult to achieve high FEs in electrochemical CO2RR. The reason was that the formation of C3 product requires the stabilization of C2 intermediate, and a number of complex active sites not only for C1–C1 coupling but also for C1–C2 coupling. Furthermore, the mechanism towards C3 products has been far less well understood compared to C2 products. Nonetheless, some strategies have been proposed to raise the FEs of C3 oxygenates in electrochemical CO2RR, especially for n-C3H7OH. Rahaman et al. reported the fabrication of a dendritic Cu by a stepwise method, which enhanced the FE of n-C3H7OH by a process involving the electrodeposition of dendritic Cu catalysts under mass transfer control using Cu(Ⅱ) ions followed by 3 h thermal annealing at 300 ℃ [96]. The initial electrodeposition catalysts mainly produced HCOO and C2H4, but the annealed samples directed the major product towards C2+ alcohols (Figure 5f), with the FE of n-C3H7OH reaching 13.1%. It was proposed that the presence of smaller nm-sized Cu NPs/Cu crystals and nanocavities on the surface of the dendritic Cu played a key role in promoting C–C coupling and C2+ alcohol production. Using an electrochemical lithium method, Peng et al. modulated the density of sulfur vacancies on CuS by tuning the cycle number, and synthesized CuSX catalysts with a double-sulfur vacancy (DSV) [82]. The DFT calculation revealed that the DSV formed on hexagonal copper sulfide allowed enriched negative charges near two adjacent co-planar vacancies and stabilized the *OCCO dimer. In addition, the coupling energy between a third *CO and the dimer was decreased in the presence of the DSV. As a result, after 10 cycles the CuSx catalysts involving the rich DSVs, exhibited an FE towards n-propanol of 15.4% at −1.05 V vs. RHE, higher than other samples (Figure 5g). The decrease in FEn-propanol with numerous cycles was ascribed to the damage of the formed DSV sites.

3.3. Current Density

The current density is another key parameter in electrochemical CO2RR, describing the activity of the reaction. However, as mentioned in the introduction, the stable C=O bonds and sluggish kinetics result in poor current density in CO2 reduction. Therefore, to enable electrochemical CO2RR applicability, strategies based on reactors and catalysts were needed to promote current density to an industrial-relevant level (>200 mA·cm−2) [101].
Early studies operated in H-cells showed low current density (≤100 mA·cm−2) due to the mass-transport limitation of CO2. Flow cells with GDE configuration were widely used to promote the transport of the reactant and significantly raise the current density of CO2RR.
Carbon-based gas diffusion layers (GDLs) were often used in previous studies on gas diffusion electrodes (GDEs) for CO2RR [102]. As shown in Figure 6a, recently developed GDEs typically contain GDL consisting of a microporous substrate (MPS), a microporous layer (MPL), and a catalyst layer (CL) [103]. The MPS is usually made of conductive carbon fibers or titanium foam coated with PTFE, providing mechanical stability and electrical contact, as well as the distribution of CO2 gas through its macro-scale pores. MPLs are comprised of carbon nanoparticles and polytetrafluoroethylene (PTFE), which enhance the interfacial electrical connection and prevent flooding in the GDE. CLs are prepared by coating MPLs with a mixture of catalyst particles and ionomers, as binders. Ionomers hold catalyst particles together and concurrently offer ionic conductivity to CLs.
The flow cell consists of three flow channels for the CO2 gas, catholyte, and anolyte, as shown in Figure 6b [104]. The catholyte and anolyte liquid streams are separated by an ion-exchange membrane, which prevents CO2RR products from crossing over to the anode where they can be oxidized back into CO2, and restricts the evolved O2 at the anode from crossing over to the cathode and stealing electrons for ORR. The GDE separates the catholyte and gas channel. The catalyst layer, located on the liquid facing the front of the GDE, is in contact with the electrolyte, while gas-phase CO2 is continuously delivered to the catalyst through the back of the GDE, thus lowering the CO2 diffusion to the surface of the catalyst by three orders of magnitude and achieving a high current density above 200 mA·cm−2 [29,41,77,105]. SACs were found to achieve a large current density with the use of GDE [106,107,108]. Recently, the self-supported single-atom nickel-decorated porous carbon membrane catalyst was directly used as a gas diffusion electrode, combining gas diffusion and catalyst layers into a single architecture [109].
The improvement in current density not only satisfies the basic demand of the industrial process, but also provides ways to compare the performance of catalysts at current densities of different orders of magnitude within the laboratory, thus giving more guidance for future applications. Additionally, the high feeding rate of CO2 combined with a short diffusion distance, allowed highly alkaline electrolytes, increased conductivity, and suppressed HER, further promoting electrocatalytic performance.
Figure 6. (a) Schematic illustration of the structure of gas diffusion electrode [103]; copyright Royal Society of Chemistry, 2020. (b) Schematic illustration of the structure of a flow cell [104]; copyright John Wiley & Son, 2019. (c) Illustration of the promotion of gas and ion transport by the hydrophobic -CF2 layer and hydrophilic −SO3 layer, respectively [105]; copyright AAAS, 2020. (d) Current density on F-Cu catalysts compared by other typical electrocatalysts (squares and circles represent the electrocatalytic reactions conducted in an H-shape cell and flow cell, respectively) [61]; copyright Springer Nature, 2020. (e) The stability of Cu-12 with N-aryl-dihydropyridine-based oligomer [29]; copyright Springer Nature, 2019. (f) Long-term catalytic performance test of CO2 reduction to ethylene in 7 M KOH catalyzed by the graphite/carbon NPs/Cu/PTFE electrode, compared with a traditional carbon-based gas diffusion electrode (GDE) (insets show the cross-section SEM and energy-dispersive X-ray spectroscopy mapping of the sample after 150 h of continuous CO2 reduction operation) [110]; copyright AAAS, 2018. (g) Schematic illustration of the structure of a membrane electrode assembly (MEA) cell [111]; copyright MDPI, 2020.
Figure 6. (a) Schematic illustration of the structure of gas diffusion electrode [103]; copyright Royal Society of Chemistry, 2020. (b) Schematic illustration of the structure of a flow cell [104]; copyright John Wiley & Son, 2019. (c) Illustration of the promotion of gas and ion transport by the hydrophobic -CF2 layer and hydrophilic −SO3 layer, respectively [105]; copyright AAAS, 2020. (d) Current density on F-Cu catalysts compared by other typical electrocatalysts (squares and circles represent the electrocatalytic reactions conducted in an H-shape cell and flow cell, respectively) [61]; copyright Springer Nature, 2020. (e) The stability of Cu-12 with N-aryl-dihydropyridine-based oligomer [29]; copyright Springer Nature, 2019. (f) Long-term catalytic performance test of CO2 reduction to ethylene in 7 M KOH catalyzed by the graphite/carbon NPs/Cu/PTFE electrode, compared with a traditional carbon-based gas diffusion electrode (GDE) (insets show the cross-section SEM and energy-dispersive X-ray spectroscopy mapping of the sample after 150 h of continuous CO2 reduction operation) [110]; copyright AAAS, 2018. (g) Schematic illustration of the structure of a membrane electrode assembly (MEA) cell [111]; copyright MDPI, 2020.
Catalysts 12 00860 g006
Based on the flow cell with GDE, modifying the catalyst’s surface with ionomer further enhanced the CO2 mass transfer, thus increasing the current density. Arquer et al. spray-coated a solution of perfluorinated sulfonic acid (PFSA) ionomers onto hydrophilic metal catalysts to establish a catalyst: ionomer planar heterojunction [105]. The PFSA ionomers containing -SO3 (hydrophilic) and -CF2 (hydrophobic) groups, formed colloids with hydrophilic -SO3 groups exposed to solvent in a polar electrolyte, which was proved by ex situ and in situ surface-enhanced Raman spectroscopy (SERS). Therefore, this controlled assembly into distinct hydrophobic and hydrophilic layers offered differentiated pathways whereby gas transport was promoted through the hydrophobic layers, and ion transport was enhanced by the hydrophilic layers (Figure 6c), thus promoting CO2RR due to the benefit from a three-phase reaction interface. Enhanced gas transport with decent water availability was evidenced by the oxygen reduction reaction (ORR) and HER. To maximize the three-phase reaction interface, they fabricated the 3D catalysts by ionomer bulk heterojunction consisting of Cu NPs and PFSA, and spray-casted on a PTFE/Cu/ionomer gas diffusion support. Hence, the catalyst achieved CO2 electroreduction in 7 M KOH with a peak ethylene partial current density of 1.3 A·cm−2.
Other strategies for overpotential decrease or selectivity enhancement also increased the current density. For example, the fluorine-modified Cu (F-Cu) catalysts mentioned earlier achieved high current density in an alkaline flow cell [61]. The modification of F enhanced CO adsorption, water activation, and the hydrogenation of *CO to *CHO, and the kinetics of C2+ formation was promoted. Hence, the current density reached 1.6 A·cm−2 with an FE of 80% for C2+ products, surpassing typical electrocatalysts (Figure 6d). The electrochemical lithium tuning strategy to form vacancies was used to raise the FE of n-propanol, as already mentioned, and also enhanced the current density [82,112,113]. During the process, the Li+ intercalated into the lattice of synthesized Cu3N to form Li3N, based on the partial conversion reaction of Cu3N + Li+ + e → Cu3Nx + Li3N (0 < x < 1). The nitrogen vacancy density was adjusted by controlling the charge–discharge current. With an optimal current of 50 μA, the Cu3Nx catalysts exhibited a partial current density towards C2 products of 307 mA·cm−2, with a high FE of 81.7%. The DFT calculation revealed that the nitrogen-vacancy-rich Cu3N promoted *CO adsorption and C–C coupling, thus enhancing the electrocatalytic performance. Since the Li2CuO2 involved Li species, the Li vacancies were directly formed in the discharging process. It was suggested by DFT that the lithium vacancies (VLi) resulted in the short distance to the adjacent [CuO4] layer and reduced the coordination number of Li+ around Cu, thus lowering the energy barrier of C–C coupling and enhancing the activity and selectivity. Therefore, with the VLi percentage around 1.6%, the Li2-xCuO2 catalysts showed a high current density of 706 mA·cm−2 with an FE of 90.6% towards C2+ products. A similar strategy was adopted to obtain a high current density for formate production. Yan et al. developed a surface-lithium-doped tin (s-SnLi) catalyst by controlled electrochemical lithiation [114]. DFT calculations indicated that the Li dopants introduced electron localization and lattice strains on the Sn surface, thus enhancing the activity and selectivity of the CO2 electroreduction to formate. As a result, the s-SnLi electrocatalyst exhibited strong performance, with a partial current density of 1.0 A·cm−2 for producing formate, and a corresponding FE of 92 %. For CO production, Wang et al. developed a porous Ni-N-C catalyst containing atomically dispersed NiN4 sites and nanostructured zirconium oxide (ZrO2@Ni-NC) via a post-synthetic coordination coupling carbonization strategy [73]. In a flow cell, the ZrO2@Ni-NC delivered a current density of 200 mA·cm−2 with CO selectivity of 98.6% at −1.58 V vs. RHE. The experimental results demonstrated that the introduced nanostructured ZrO2 accelerated the transfer rate of the first electron and altered the RDS from conventional CO2 adsorption to the protonation of *CO2 intermediate to form *COOH intermediate, thus achieving a high reaction rate.

3.4. Stability

Considering the demand for large-scale application of electrochemical CO2RR, stability is also an important parameter, as it determines whether cost efficiency in mass production can be achieved. It has been suggested that the short life of the system causes an exponential increase in maintenance costs and the infeasibility of the practical economy [12]. However, the stability issue affecting the reaction has often been overlooked compared to activity and selectivity. The stability issue involves the maintenance of catalysts and electrochemical systems. In this section, the factors causing poor stability are introduced, followed by strategies to inhibit effectively these deactivation processes, which include surface capping, cutting-edge design, and improved cell configuration.
The catalysts’ surface morphology has been acknowledged as one of the most important factors determining catalytic activity and selectivity. However, metal-based catalysts are susceptible to surface reconstruction, leading to unsatisfying stability. For example, the high-index facets of Cu showing high FE for ethylene usually underwent reconstruction due to their high surface energy during electrocatalysis. An increase in the average size of crystalline NPs occurred due to the tendency for the smaller nanosized particles to adjust their surface energy via various growth mechanisms such as Ostwald ripening and coalescence [115,116]. An effective measure to preserve the morphological change was the usage of surface capping ligands as surface passivating reagents. Ma et al. prepared Au NPs on carbon support modified by polyvinyl alcohol (PVA), exhibiting a consistent FE of CO above 95% and a current density of 27 mA·cm−2 for 24 h reaction time [117]. It was suggested the longer durability compared to that of bare Au/C arose from the Au NPs’ immobilization on carbon support by the PVA surface capping ligands. Zhang et al. showed that N-heterocyclic carbene ligands enhanced the stability of CO2RR [118]. In addition to these surface capping ligands, other organic additives also extended the operational time. Li et al. fabricated an N-aryl-substituted tetrahydro-bipyridine as an additive on sputtered Cu catalysts, by the electrochemical dimerization of N-arylpyridinium additives [29]. The catalysts exhibited an ethylene FE of 65% at −0.83 V vs. RHE, and after further optimization by increasing the number of electro-donating N atoms, the sputtered Cu catalysts modified by electro-oligomerized N,N’-(1,4-phenylne) bispyridinium showed remarkable stability over 190 h, with high current density and C2H4 selectivity (Figure 6e).
Although alloyed metals were observed to enhance the electrocatalytic performance in CO2RR, they were restricted by the disadvantage that each metal tended to aggregate into a single metal state during the reaction, and the metal with low surface energy preferred to compose an outer layer by segregating from the alloy [119,120]. To solve these issues, a cutting-edge design was more recently proposed. Sun et al. fabricated AuFe@Fe core-shell NPs by leaching out Fe from prepared AuFe alloyed NPs, which showed stable CO production for 90 h [121]. The DFT calculation suggested that the high stability arose from dimpled Au surfaces, which mitigated Au diffusion and aggregation. Zhu et al. synthesized Au3Cu alloy with abundant vacancies by acid etching from as-prepared Au3Cu alloy NPs [122]. The prepared alloy NPs were stable under CO2RR condition for 30 h, with a high CO FE of 94.3% at −0.43 V vs. RHE.
In addition to the deactivation of catalysts, the overall stability of CO2RR was also dependent on the reactors. As mentioned earlier, the flow cell with GDE configuration enhanced the mass transfer and increased the current density, but faced severe problems such as flooding and salt accumulation. It was a challenging task to achieve the long-term stability of CO2RR.
Previous studies of GDEs for CO2RR mainly used carbon-based GDLs [102], and the issue of flooding caused by the penetration of H2O into GDE resulted from the surface of carbon in the GDE turning hydrophilic during long-term operation. The diffusion of gaseous reactant towards the catalysts was prohibited and the performance was impaired. To overcome this issue, the modification of GDE was adopted. For example, Ma et al. modified the GDE by coating 20 μL 1H,1H,2H,2H-perfluorooctyltrichlorosilane (5 wt% in toluene) onto the gas diffusion layer to enhance hydrophobicity [61]. The modified catalyst was stable for 40 h at a fixed current density of 400 mA·cm−2, towards ethylene and ethanol. Dinh et al. fabricated GDL with separate PTFE and carbon NP layers to decouple the hydrophobic and current collection requirements of traditional carbon-based GDLs [110]. The pure PTFE layer acted as a stable hydrophobic GDL that prevented flooding, and the Cu catalysts were sandwiched by the two layers stabilized by the carbon NPs and graphite. After the fabrication, the graphite–carbon NPs/Cu/PTDE electrode operated for 150 h without loss of ethylene selectivity, demonstrating a 300-fold increase in the reaction time compared with that of the Cu–carbon GDL (Figure 6f). In fact, the GDL has often been overlooked because of the availability of a variety of commercial GDLs for fuel-cell application. It is worth noting that most of the GDLs tested so far are not stable for CO2 reduction, mainly because of salt accumulation [123,124,125]. The creation of next-generation GDLs with mechanical and chemical properties suited for CO2 reduction is necessary to enhance the current density and stability of CO2RR.
The salt accumulation issue on the GDE was mainly attributed to the formation of carbonates by the reaction of CO2 with OH, blocking the pores of GDE and hindering the permeability and conductivity. The side reaction resulted in carbon loss and the decrease of pH, which was unbeneficial to C–C coupling, and increased energy costs. Regularly refreshing the electrolyte suppressed this process, but with elevated operating costs and energy consumption. Operating in acid conditions seemed to be an effective way to alleviate the salt accumulation, as carbonate formation was largely suppressed. It was reported that the acidic fuel cell was effective for at least an order of magnitude longer than the alkaline fuel cell [126]. However, the dominant HER in an acidic medium could lead to low selectivity towards the target product. Although some research has suggested utilizing the cation effect to suppress HER and made some progress on the issue of selectivity during catalysis, the FE towards multi-electron electroreduction product remained unsatisfying and did not meet the expectations for stability [127,128]. Therefore, a number of efforts are still needed to enhance the selectivity and stability of CO2RR in the acidic flow cell.
The recent development of membrane electrode assembly (MEA) cells may be a potential strategy to enhance stability. The configuration of MEA is shown in Figure 6g [111]. In an MEA cell, the cathodic GDE and anodic catalyst were directly pressed on both sides of the ion exchange membrane, thus the flooding of the GDE and the carbonate formation were circumvented because there was no catholyte. Li et al. reported that the MEA achieved a stable operation for 190 h during electrochemical CO2RR to C2H4, with an FE of ethylene of around 60% and a current density of 120 mA·cm−2 in neutral medium, demonstrating the long durability of the MEA cell [29]. As there is no electrolyte in the cathode, the humidity of the CO2 input stream was found to be important as the proton in the solution is needed for electroreduction [111]. Increasing reactant humidification resulted not only in high current density operation but also low power requirements for reducing CO2 to CO [129]. However, despite excellent stability, the MEA system suffered from low current density and high cell voltage due to low local pH and conductivity. Recently, the direct electrolysis of carbonate led to stable operation under CO2RR conditions. The catalysts were loaded on hydrophilic carbon paper without carbonate formation, thus avoiding flooding and salt accumulation. Inspired by this concept, Li et al. constructed a system and achieved continuous operation for 145 h with an Ag catalyst, outperforming a great deal of previous research [130]. However, other problems existed in this system. Only two electron reduction products were obtained, and the FEs towards them decreased rapidly as the current density increased. The cell voltage was relatively large due to the bipolar membrane inside the device, far from a mature technique [131,132,133].

4. Conclusions and Prospects

In this minireview, we have summarized recent progress in CO2RR on copper based-catalysts, including catalyst synthesis and reactor design. The drawbacks of high overpotential, low faradaic efficiency, small current density, and short-time stability have prevented further industrial application due to economic costs. Catalyst synthesis and the catalysis engineering have been improved to obtain target products, with high current density at low overpotential.
To produce electroreduction products at low overpotential, SACs have been employed in CO2RR due to their good catalytic performance. Single-atom fabrication by the stabilized ligands enhanced the catalytic reduction reaction at decreased overpotential. For multi-electron electroreduction products, Cu-based catalysts were used in metal clusters, special morphology, and tandem catalysts, or modified by base chemicals. Thus, the binding and coverage of the *CO intermediates were enhanced, resulting in multi-electron electroreduction at low overpotential.
To obtain the target product economically, high FE in CO2RR is necessary. For CH4, traditional SACs are favorable for the prohibition of C–C coupling. Cu sites were stabilized by CeO2, MOF, and CuGaO2, exhibiting high CH4 partial current density. The nanocell catalysts combining an Ag core and Cu2O shell resembled a tandem nanoreactor, achieving high CH4 FE by changing the Cu2O envelope size. For methanol, Cu-based materials were also discussed; strategies that boosted the FE of methanol included fabrication of other NPs, MXene immobilization, and doping. For ethylene, the strategies for the catalyst design were concentrated on C–C coupling. Crystal facet tuning and surface modification were found to enhance the generation of ethylene. For C2+ oxygenates, bimetallic catalysts containing Cu and another metal were suggested for stabilizing oxygenate intermediates, contributing to the product of C2+ oxygenates. It may be possible to obtain C3 products of n-propanol, using Cu materials with further fabrication and the introduction of co-planar vacancies.
To enlarge the current density, a flow cell with GDE technology was introduced to overcome the limitation of CO2 dissolution. A highly alkaline electrolyte was reported to increase the conductivity and promote electroreduction. Modifying the catalyst’s surface with ionomer was thought to enhance the CO2 mass transfer, thus increasing the current density. It was suggested that the strategies used for low overpotential and high FE are also effective to increase the current density.
To maintain the stability of the reaction, surface reconstruction should be prohibited. Stabilized ligands and alloyed metals were used to protect the active metal species of the catalysts during electroreduction. Fabrication of the GDE and GDL was adopted to minimize the flooding effect and an acidic electrolyte was used to limit the effect of salt accumulation on the catalytic performance in the flow cell. A developed MEA was achieved to avoid flooding and salt accumulation, but improvements remain necessary to address overpotential, FE, and current density.
Recent strategies including single-atom catalysts, crystal facet engineering, bimetallics, doping, surface modification, metal–carbon composition, and the developed technologies of flow cells, GDL, and MEA have brought improvements in the overpotential, FE, current density, and stability of CO2RR. Thus, the combined route of catalyst synthesis and operational technology presents a promising pathway for obtaining the target products of methane, methanol, ethylene, and C2+ oxygenates at low economic cost. It can be expected that the practical application of CO2RR will be achieved in industry in the future.

Author Contributions

Writing-original draft preparation, B.W.; Funding acquisition, J.C. and L.Q.; Writing—review and editing, L.Q. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (81971666), the Shanghai Pujiang Program (18PJ1401300), and Fudan Undergraduate Research Opportunities Program (FDUROP 21067).

Acknowledgments

The authors thank the following funding agencies for supporting this work: the Natural Science Foundation of China (81971666), the Shanghai Pujiang Program (18PJ1401300), and Fudan Undergraduate Research Opportunities Program (FDUROP 21067).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Olah, G.A.; Prakash, G.K.S.; Goeppert, A. Anthropogenic Chemical Carbon Cycle for a Sustainable Future. J. Am. Chem. Soc. 2011, 133, 12881–12898. [Google Scholar] [CrossRef] [PubMed]
  2. Nitopi, S.; Bertheussen, E.; Scott, S.B.; Liu, X.; Engstfeld, A.K.; Horch, S.; Seger, B.; Stephens, I.E.L.; Chan, K.; Hahn, C.; et al. Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte. Chem. Rev. 2019, 119, 7610–7672. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar] [CrossRef]
  4. Khodakov, A.Y.; Chu, W.; Fongarland, P. Advances in the Development of Novel Cobalt Fischer−Tropsch Catalysts for Synthesis of Long-Chain Hydrocarbons and Clean Fuels. Chem. Rev. 2007, 107, 1692–1744. [Google Scholar] [CrossRef]
  5. Zheng, T.; Zhang, M.; Wu, L.; Guo, S.; Liu, X.; Zhao, J.; Xue, W.; Li, J.; Liu, C.; Li, X.; et al. Upcycling CO2 into energy-rich long-chain compounds via electrochemical and metabolic engineering. Nat. Catal. 2022, 5, 388–396. [Google Scholar] [CrossRef]
  6. Glockler, G. Carbon–Oxygen Bond Energies and Bond Distances. J. Phys. Chem. 1958, 62, 1049–1054. [Google Scholar] [CrossRef]
  7. Siewert, I. Proton-Coupled Electron Transfer Reactions Catalysed by 3 d Metal Complexes. Chem. A Eur. J. 2015, 21, 15078–15091. [Google Scholar] [CrossRef]
  8. Ross, M.B.; De Luna, P.; Li, Y.; Dinh, C.-T.; Kim, D.; Yang, P.; Sargent, E.H. Designing materials for electrochemical carbon dioxide recycling. Nat. Catal. 2019, 2, 648–658. [Google Scholar] [CrossRef] [Green Version]
  9. Burdyny, T.; Smith, W.A. CO2 reduction on gas-diffusion electrodes and why catalytic performance must be assessed at commercially-relevant conditions. Energy Environ. Sci. 2019, 12, 1442–1453. [Google Scholar] [CrossRef] [Green Version]
  10. Choi, W.; Won, D.H.; Hwang, Y.J. Catalyst design strategies for stable electrochemical CO2 reduction reaction. J. Mater. Chem. A 2020, 8, 15341–15357. [Google Scholar] [CrossRef]
  11. Jin, S.; Hao, Z.; Zhang, K.; Yan, Z.; Chen, J. Advances and Challenges for the Electrochemical Reduction of CO2 to CO: From Fundamentals to Industrialization. Angew. Chem. Int. Ed. 2021, 60, 20627–20648. [Google Scholar] [CrossRef]
  12. Park, S.; Wijaya, D.T.; Na, J.; Lee, C.W. Towards the Large-Scale Electrochemical Reduction of Carbon Dioxide. Catalysts 2021, 11, 253. [Google Scholar] [CrossRef]
  13. Yoshio, H.; Katsuhei, K.; Shin, S. Production of CO and CH4 in electrochemical reduction of CO2 at metal electrodes in aqueous hydrogencarbonate solution. Chem. Lett. 1985, 14, 1695–1698. [Google Scholar] [CrossRef]
  14. Díaz-Sainz, G.; Alvarez-Guerra, M.; Irabien, A. Continuous electroreduction of CO2 towards formate in gas-phase operation at high current densities with an anion exchange membrane. J. CO2 Util. 2022, 56, 101822. [Google Scholar] [CrossRef]
  15. Cheng, D.; Zhao, Z.-J.; Zhang, G.; Yang, P.; Li, L.; Gao, H.; Liu, S.; Chang, X.; Chen, S.; Wang, T.; et al. The nature of active sites for carbon dioxide electroreduction over oxide-derived copper catalysts. Nat. Commun. 2021, 12, 395. [Google Scholar] [CrossRef]
  16. Kuo, L.; Dinh, C.-T. Toward efficient catalysts for electrochemical CO2 conversion to C2 products. Curr. Opin. Electrochem. 2021, 30, 100807. [Google Scholar] [CrossRef]
  17. Xue, L.; Zhang, C.; Wu, J.; Fan, Q.-Y.; Liu, Y.; Wu, Y.; Li, J.; Zhang, H.; Liu, F.; Zeng, S. Unveiling the reaction pathway on Cu/CeO2 catalyst for electrocatalytic CO2 reduction to CH4. Appl. Catal. B 2022, 304, 120951. [Google Scholar] [CrossRef]
  18. Garza, A.J.; Bell, A.T.; Head-Gordon, M. Mechanism of CO2 Reduction at Copper Surfaces: Pathways to C2 Products. ACS Catal. 2018, 8, 1490–1499. [Google Scholar] [CrossRef] [Green Version]
  19. Feaster, J.T.; Shi, C.; Cave, E.R.; Hatsukade, T.; Abram, D.N.; Kuhl, K.P.; Hahn, C.; Nørskov, J.K.; Jaramillo, T.F. Understanding Selectivity for the Electrochemical Reduction of Carbon Dioxide to Formic Acid and Carbon Monoxide on Metal Electrodes. ACS Catal. 2017, 7, 4822–4827. [Google Scholar] [CrossRef]
  20. Vijay, S.; Ju, W.; Brückner, S.; Tsang, S.-C.; Strasser, P.; Chan, K. Unified mechanistic understanding of CO2 reduction to CO on transition metal and single atom catalysts. Nat. Catal. 2021, 4, 1024–1031. [Google Scholar] [CrossRef]
  21. Li, S.; Dong, X.; Chen, W.; Song, Y.; Li, G.; Wei, W.; Sun, Y. Efficient CO2 Electroreduction over Silver Hollow Fiber Electrode. Catalysts 2022, 12, 453. [Google Scholar] [CrossRef]
  22. Wang, Z.; Li, T.; Wang, Q.; Guan, A.; Cao, N.; Al-Enizi, A.M.; Zhang, L.; Qian, L.; Zheng, G. Hydrophobically made Ag nanoclusters with enhanced performance for CO2 aqueous electroreduction. J. Power. Sources 2020, 476, 228705. [Google Scholar] [CrossRef]
  23. Kim, J.; Choi, W.; Park, J.W.; Kim, C.; Kim, M.; Song, H. Branched Copper Oxide Nanoparticles Induce Highly Selective Ethylene Production by Electrochemical Carbon Dioxide Reduction. J. Am. Chem. Soc. 2019, 141, 6986–6994. [Google Scholar] [CrossRef] [PubMed]
  24. Gao, Y.; Wu, Q.; Liang, X.; Wang, Z.; Zheng, Z.; Wang, P.; Liu, Y.; Dai, Y.; Whangbo, M.-H.; Huang, B. Cu2O Nanoparticles with Both {100} and {111} Facets for Enhancing the Selectivity and Activity of CO2 Electroreduction to Ethylene. Adv. Sci. 2020, 7, 1902820. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  25. Zhong, D.; Zhao, Z.-J.; Zhao, Q.; Cheng, D.; Liu, B.; Zhang, G.; Deng, W.; Dong, H.; Zhang, L.; Li, J.; et al. Coupling of Cu(100) and (110) Facets Promotes Carbon Dioxide Conversion to Hydrocarbons and Alcohols. Angew. Chem. Int. Ed. 2021, 60, 4879–4885. [Google Scholar] [CrossRef] [PubMed]
  26. Zhong, M.; Tran, K.; Min, Y.; Wang, C.; Wang, Z.; Dinh, C.-T.; De Luna, P.; Yu, Z.; Rasouli, A.S.; Brodersen, P.; et al. Accelerated discovery of CO2 electrocatalysts using active machine learning. Nature 2020, 581, 178. [Google Scholar] [CrossRef]
  27. Lv, X.; Shang, L.; Zhou, S.; Li, S.; Wang, Y.; Wang, Z.; Sham, T.-K.; Peng, C.; Zheng, G. Electron-Deficient Cu Sites on Cu3Ag1 Catalyst Promoting CO2 Electroreduction to Alcohols. Adv. Energy Mater. 2020, 10, 2001987. [Google Scholar] [CrossRef]
  28. Wei, X.; Yin, Z.; Lyu, K.; Li, Z.; Gong, J.; Wang, G.; Xiao, L.; Lu, J.; Zhuang, L. Highly Selective Reduction of CO2 to C2+ Hydrocarbons at Copper/Polyaniline Interfaces. ACS Catal. 2020, 10, 4103–4111. [Google Scholar] [CrossRef]
  29. Li, F.; Thevenon, A.; Rosas-Hernandez, A.; Wang, Z.; Li, Y.; Gabardo, C.M.; Ozden, A.; Cao Thang, D.; Li, J.; Wang, Y.; et al. Molecular tuning of CO2-to-ethylene conversion. Nature 2020, 577, 509. [Google Scholar] [CrossRef] [Green Version]
  30. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective Electrochemical Reduction of Carbon Dioxide to Ethylene and Ethanol on Copper(I) Oxide Catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  31. Popović, S.; Smiljanić, M.; Jovanovič, P.; Vavra, J.; Buonsanti, R.; Hodnik, N. Stability and Degradation Mechanisms of Copper-Based Catalysts for Electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2020, 59, 14736–14746. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, X.; Xu, A.N.; Li, F.W.; Hung, S.F.; Nam, D.H.; Gabardo, C.M.; Wang, Z.Y.; Xu, Y.; Ozden, A.; Rasouli, A.S.; et al. Efficient Methane Electrosynthesis Enabled by Tuning Local CO2 Availability. J. Am. Chem. Soc. 2020, 142, 3525–3531. [Google Scholar] [CrossRef] [PubMed]
  33. Qiu, Y.L.; Zhong, H.X.; Xu, W.B.; Zhang, T.T.; Li, X.F.; Zhang, H.M. Tuning the electrocatalytic properties of a Cu electrode with organic additives containing amine group for CO2 reduction. J. Mater. Chem. A 2019, 7, 5453–5462. [Google Scholar] [CrossRef]
  34. Zhang, J.; Cai, W.; Hu, F.X.; Yang, H.; Liu, B. Recent advances in single atom catalysts for the electrochemical carbon dioxide reduction reaction. Chem. Sci. 2021, 12, 6800–6819. [Google Scholar] [CrossRef] [PubMed]
  35. Chernyshova, I.V.; Somasundaran, P.; Ponnurangam, S. On the origin of the elusive first intermediate of CO2 electroreduction. Proc. Natl. Acad. Sci. USA 2018, 115, E9261. [Google Scholar] [CrossRef] [Green Version]
  36. Zhang, S.; Fan, Q.; Xia, R.; Meyer, T.J. CO2 Reduction: From Homogeneous to Heterogeneous Electrocatalysis. Acc. Chem. Res. 2020, 53, 255–264. [Google Scholar] [CrossRef]
  37. Bagger, A.; Ju, W.; Varela, A.S.; Strasser, P.; Rossmeisl, J. Electrochemical CO2 Reduction: A Classification Problem. ChemPhysChem 2017, 18, 3266–3273. [Google Scholar] [CrossRef] [Green Version]
  38. Cheng, T.; Xiao, H.; Goddard, W.A. Free-Energy Barriers and Reaction Mechanisms for the Electrochemical Reduction of CO on the Cu(100) Surface, Including Multiple Layers of Explicit Solvent at pH 0. J. Phys. Chem. Lett. 2015, 6, 4767–4773. [Google Scholar] [CrossRef]
  39. Nie, X.; Esopi, M.R.; Janik, M.J.; Asthagiri, A. Selectivity of CO2 Reduction on Copper Electrodes: The Role of the Kinetics of Elementary Steps. Angew. Chem. Int. Ed. 2013, 52, 2459–2462. [Google Scholar] [CrossRef]
  40. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  41. Peng, C.; Xu, Z.; Luo, G.; Yan, S.; Zhang, J.; Li, S.; Chen, Y.; Chang, L.Y.; Wang, Z.; Sham, T.-K.; et al. Highly-Exposed Single-Interlayered Cu Edges Enable High-Rate CO2-to-CH4 Electrosynthesis. Adv. Energy Mater. 2022, 12, 2200195. [Google Scholar] [CrossRef]
  42. Calle-Vallejo, F.; Koper, M.T.M. Theoretical Considerations on the Electroreduction of CO to C2 Species on Cu(100) Electrodes. Angew. Chem. Int. Ed. 2013, 52, 7282–7285. [Google Scholar] [CrossRef] [PubMed]
  43. Lum, Y.; Cheng, T.; Goddard, W.A.; Ager, J.W. Electrochemical CO Reduction Builds Solvent Water into Oxygenate Products. J. Am. Chem. Soc. 2018, 140, 9337–9340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Yang, K.D.; Lee, C.W.; Jin, K.; Im, S.W.; Nam, K.T. Current Status and Bioinspired Perspective of Electrochemical Conversion of CO2 to a Long-Chain Hydrocarbon. J. Phys. Chem. Lett. 2017, 8, 538–545. [Google Scholar] [CrossRef] [PubMed]
  45. Kortlever, R.; Shen, J.; Schouten, K.J.; Calle-Vallejo, F.; Koper, M.T. Catalysts and Reaction Pathways for the Electrochemical Reduction of Carbon Dioxide. J. Phys. Chem. Lett. 2015, 6, 4073–4082. [Google Scholar] [CrossRef]
  46. Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electrochemical Reduction of CO at a Copper Electrode. J. Phys. Chem. B 1997, 101, 7075–7081. [Google Scholar] [CrossRef]
  47. Ren, D.; Wong, N.T.; Handoko, A.D.; Huang, Y.; Yeo, B.S. Mechanistic Insights into the Enhanced Activity and Stability of Agglomerated Cu Nanocrystals for the Electrochemical Reduction of Carbon Dioxide to n-Propanol. J. Phys. Chem. Lett. 2016, 7, 20–24. [Google Scholar] [CrossRef]
  48. Ma, W.; He, X.; Wang, W.; Xie, S.; Zhang, Q.; Wang, Y. Electrocatalytic reduction of CO2 and CO to multi-carbon compounds over Cu-based catalysts. Chem. Soc. Rev. 2021, 50, 12897–12914. [Google Scholar] [CrossRef]
  49. Montoya, J.H.; Peterson, A.A.; Nørskov, J.K. Insights into C-C Coupling in CO2 Electroreduction on Copper Electrodes. ChemCatChem 2013, 5, 737–742. [Google Scholar] [CrossRef]
  50. Bagger, A.; Arnarson, L.; Hansen, M.H.; Spohr, E.; Rossmeisl, J. Electrochemical CO Reduction: A Property of the Electrochemical Interface. J. Am. Chem. Soc. 2019, 141, 1506–1514. [Google Scholar] [CrossRef] [PubMed]
  51. Zheng, W.; Yang, J.; Chen, H.; Hou, Y.; Wang, Q.; Gu, M.; He, F.; Xia, Y.; Xia, Z.; Li, Z.; et al. Atomically Defined Undercoordinated Active Sites for Highly Efficient CO2 Electroreduction. Adv. Funct. Mater. 2020, 30, 1907658. [Google Scholar] [CrossRef]
  52. Zhang, H.; Li, J.; Xi, S.; Du, Y.; Hai, X.; Wang, J.; Xu, H.; Wu, G.; Zhang, J.; Lu, J.; et al. A Graphene-Supported Single-Atom FeN5 Catalytic Site for Efficient Electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 14871–14876. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, H.B.; Hung, S.-F.; Liu, S.; Yuan, K.; Miao, S.; Zhang, L.; Huang, X.; Wang, H.-Y.; Cai, W.; Chen, R.; et al. Atomically dispersed Ni(i) as the active site for electrochemical CO2 reduction. Nat. Energy 2018, 3, 140–147. [Google Scholar] [CrossRef]
  54. Gu, J.; Hsu, C.-S.; Bai, L.; Chen, H.M.; Hu, X. Atomically dispersed Fe3+ sites catalyze efficient CO2 electroreduction to CO. Science 2019, 364, 1091. [Google Scholar] [CrossRef]
  55. Pei, J.; Wang, T.; Sui, R.; Zhang, X.; Zhou, D.; Qin, F.; Zhao, X.; Liu, Q.; Yan, W.; Dong, J.; et al. N-Bridged Co-N-Ni: New bimetallic sites for promoting electrochemical CO2 reduction. Energy Environ. Sci. 2021, 14, 3019–3028. [Google Scholar] [CrossRef]
  56. Lin, L.; Li, H.; Yan, C.; Li, H.; Si, R.; Li, M.; Xiao, J.; Wang, G.; Bao, X. Synergistic Catalysis over Iron-Nitrogen Sites Anchored with Cobalt Phthalocyanine for Efficient CO2 Electroreduction. Adv. Mater. 2019, 31, 1903470. [Google Scholar] [CrossRef]
  57. Zu, X.; Li, X.; Liu, W.; Sun, Y.; Xu, J.; Yao, T.; Yan, W.; Gao, S.; Wang, C.; Wei, S.; et al. Efficient and Robust Carbon Dioxide Electroreduction Enabled by Atomically Dispersed Sn delta+ Sites. Adv. Mater. 2019, 31, 1808135. [Google Scholar] [CrossRef]
  58. Zhang, E.; Wang, T.; Yu, K.; Liu, J.; Chen, W.; Li, A.; Rong, H.; Lin, R.; Ji, S.; Zhene, X.; et al. Bismuth Single Atoms Resulting from Transformation of Metal-Organic Frameworks and Their Use as Electrocatalysts for CO2 Reduction. J. Am. Chem. Soc. 2019, 141, 16569–16573. [Google Scholar] [CrossRef]
  59. Nellaiappan, S.; Katiyar, N.K.; Kumar, R.; Parui, A.; Malviya, K.D.; Pradeep, K.G.; Singh, A.K.; Sharma, S.; Tiwary, C.S.; Biswas, K. High-Entropy Alloys as Catalysts for the CO2 and CO Reduction Reactions: Experimental Realization. ACS Catal. 2020, 10, 3658–3663. [Google Scholar] [CrossRef]
  60. Fan, L.; Liu, C.-Y.; Zhu, P.; Xia, C.; Zhang, X.; Wu, Z.-Y.; Lu, Y.; Senftle, T.P.; Wang, H. Proton sponge promotion of electrochemical CO2 reduction to multi-carbon products. Joule 2022, 6, 205–220. [Google Scholar] [CrossRef]
  61. Ma, W.; Xie, S.; Liu, T.; Fan, Q.; Ye, J.; Sun, F.; Jiang, Z.; Zhang, Q.; Cheng, J.; Wang, Y. Electrocatalytic reduction of CO2 to ethylene and ethanol through hydrogen-assisted C–C coupling over fluorine-modified copper. Nat. Catal. 2020, 3, 478–487. [Google Scholar] [CrossRef]
  62. Xu, H.; Rebollar, D.; He, H.; Chong, L.; Liu, Y.; Liu, C.; Sun, C.-J.; Li, T.; Muntean, J.V.; Winans, R.E.; et al. Highly selective electrocatalytic CO2 reduction to ethanol by metallic clusters dynamically formed from atomically dispersed copper. Nat. Energy 2020, 5, 623–632. [Google Scholar] [CrossRef]
  63. Morales-Guio, C.G.; Cave, E.R.; Nitopi, S.A.; Feaster, J.T.; Wang, L.; Kuhl, K.P.; Jackson, A.; Johnson, N.C.; Abram, D.N.; Hatsukade, T.; et al. Improved CO2 reduction activity towards C2+ alcohols on a tandem gold on copper electrocatalyst. Nat. Catal. 2018, 1, 764–771. [Google Scholar] [CrossRef]
  64. Li, F.; Li, Y.C.; Wang, Z.; Li, J.; Nam, D.-H.; Lum, Y.; Luo, M.; Wang, X.; Ozden, A.; Hung, S.-F.; et al. Cooperative CO2-to-ethanol conversion via enriched intermediates at molecule–metal catalyst interfaces. Nat. Catal. 2020, 3, 75–82. [Google Scholar] [CrossRef]
  65. Zhu, Q.; Sun, X.; Yang, D.; Ma, J.; Kang, X.; Zheng, L.; Zhang, J.; Wu, Z.; Han, B. Carbon dioxide electroreduction to C2 products over copper-cuprous oxide derived from electrosynthesized copper complex. Nat. Commun. 2019, 10, 3851. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zheng, W.; Wang, Y.; Shuai, L.; Wang, X.; He, F.; Lei, C.; Li, Z.; Yang, B.; Lei, L.; Yuan, C.; et al. Highly Boosted Reaction Kinetics in Carbon Dioxide Electroreduction by Surface-Introduced Electronegative Dopants. Adv. Funct. Mater. 2021, 31, 2008146. [Google Scholar] [CrossRef]
  67. Zhang, B.; Zhang, J.; Shi, J.; Tan, D.; Liu, L.; Zhang, F.; Lu, C.; Su, Z.; Tan, X.; Cheng, X.; et al. Manganese acting as a high-performance heterogeneous electrocatalyst in carbon dioxide reduction. Nat. Commun. 2019, 10, 2980. [Google Scholar] [CrossRef] [PubMed]
  68. Zhang, X.; Wu, Z.; Zhang, X.; Li, L.; Li, Y.; Xu, H.; Li, X.; Yu, X.; Zhang, Z.; Liang, Y.; et al. Highly selective and active CO2 reduction electrocatalysts based on cobalt phthalocyanine/carbon nanotube hybrid structures. Nat. Commun. 2017, 8, 14675. [Google Scholar] [CrossRef] [Green Version]
  69. Guan, A.; Chen, Z.; Quan, Y.; Peng, C.; Wang, Z.; Sham, T.-K.; Yang, C.; Ji, Y.; Qian, L.; Xu, X.; et al. Boosting CO2 Electroreduction to CH4 via Tuning Neighboring Single-Copper Sites. ACS Energy Lett. 2020, 5, 1044–1053. [Google Scholar] [CrossRef]
  70. Cai, Y.; Fu, J.; Zhou, Y.; Chang, Y.-C.; Min, Q.; Zhu, J.-J.; Lin, Y.; Zhu, W. Insights on forming N,O-coordinated Cu single-atom catalysts for electrochemical reduction CO2 to methane. Nat. Commun. 2021, 12, 586. [Google Scholar] [CrossRef]
  71. Resasco, J.; Chen, L.D.; Clark, E.; Tsai, C.; Hahn, C.; Jaramillo, T.F.; Chan, K.; Bell, A.T. Promoter Effects of Alkali Metal Cations on the Electrochemical Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2017, 139, 11277–11287. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Liu, S.; Wang, X.-Z.; Tao, H.; Li, T.; Liu, Q.; Xu, Z.; Fu, X.-Z.; Luo, J.-L. Ultrathin 5-fold twinned sub-25nm silver nanowires enable highly selective electroreduction of CO2 to CO. Nano Energy 2018, 45, 456–462. [Google Scholar] [CrossRef]
  73. Wang, X.; Feng, S.; Lu, W.; Zhao, Y.; Zheng, S.; Zheng, W.; Sang, X.; Zheng, L.; Xie, Y.; Li, Z.; et al. A New Strategy for Accelerating Dynamic Proton Transfer of Electrochemical CO2 Reduction at High Current Densities. Adv. Funct. Mater. 2021, 31, 2104243. [Google Scholar] [CrossRef]
  74. Fan, L.; Xia, C.; Zhu, P.; Lu, Y.; Wang, H. Electrochemical CO2 reduction to high-concentration pure formic acid solutions in an all-solid-state reactor. Nat. Commun. 2020, 11, 3633. [Google Scholar] [CrossRef] [PubMed]
  75. Zhang, W.; Yang, S.; Jiang, M.; Hu, Y.; Hu, C.; Zhang, X.; Jin, Z. Nanocapillarity and Nanoconfinement Effects of Pipet-like Bismuth@Carbon Nanotubes for Highly Efficient Electrocatalytic CO2 Reduction. Nano Lett. 2021, 21, 2650–2657. [Google Scholar] [CrossRef]
  76. Han, L.; Song, S.; Liu, M.; Yao, S.; Liang, Z.; Cheng, H.; Ren, Z.; Liu, W.; Lin, R.; Qi, G.; et al. Stable and Efficient Single-Atom Zn Catalyst for CO2 Reduction to CH4. J. Am. Chem. Soc. 2020, 142, 12563–12567. [Google Scholar] [CrossRef]
  77. Zhang, Y.; Dong, L.-Z.; Li, S.; Huang, X.; Chang, J.-N.; Wang, J.-H.; Zhou, J.; Li, S.-L.; Lan, Y.-Q. Coordination environment dependent selectivity of single-site-Cu enriched crystalline porous catalysts in CO2 reduction to CH4. Nat. Commun. 2021, 12, 6390. [Google Scholar] [CrossRef]
  78. Yang, D.; Zhu, Q.; Chen, C.; Liu, H.; Liu, Z.; Zhao, Z.; Zhang, X.; Liu, S.; Han, B. Selective electroreduction of carbon dioxide to methanol on copper selenide nanocatalysts. Nat. Commun. 2019, 10, 677. [Google Scholar] [CrossRef]
  79. Li, P.; Bi, J.; Liu, J.; Zhu, Q.; Chen, C.; Sun, X.; Zhang, J.; Han, B. In situ dual doping for constructing efficient CO2-to-methanol electrocatalysts. Nat. Commun. 2022, 13, 1965. [Google Scholar] [CrossRef]
  80. Li, Y.C.; Wang, Z.; Yuan, T.; Nam, D.-H.; Luo, M.; Wicks, J.; Chen, B.; Li, J.; Li, F.; de Arguer, F.P.G.; et al. Binding Site Diversity Promotes CO2 Electroreduction to Ethanol. J. Am. Chem. Soc. 2019, 141, 8584–8591. [Google Scholar] [CrossRef]
  81. Wang, Y.; Wang, D.; Dares, C.J.; Marquard, S.L.; Sheridan, M.V.; Meyer, T.J. CO2 reduction to acetate in mixtures of ultrasmall (Cu)n,(Ag)m bimetallic nanoparticles. Proc. Natl. Acad. Sci. USA 2018, 115, 278–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  82. Peng, C.; Luo, G.; Zhang, J.; Chen, M.; Wang, Z.; Sham, T.-K.; Zhang, L.; Li, Y.; Zheng, G. Double sulfur vacancies by lithium tuning enhance CO2 electroreduction to n-propanol. Nat. Commun. 2021, 12, 1580. [Google Scholar] [CrossRef] [PubMed]
  83. Gao, F.-Y.; Hu, S.-J.; Zhang, X.-L.; Zheng, Y.-R.; Wang, H.-J.; Niu, Z.-Z.; Yang, P.-P.; Bao, R.-C.; Ma, T.; Dang, Z.; et al. High-Curvature Transition-Metal Chalcogenide Nanostructures with a Pronounced Proximity Effect Enable Fast and Selective CO2 Electroreduction. Angew. Chem. Int. Ed. 2020, 59, 8706–8712. [Google Scholar] [CrossRef] [PubMed]
  84. Ren, W.; Tan, X.; Yang, W.; Jia, C.; Xu, S.; Wang, K.; Smith, S.C.; Zhao, C. Isolated Diatomic Ni-Fe Metal–Nitrogen Sites for Synergistic Electroreduction of CO2. Angew. Chem. Int. Ed. 2019, 58, 6972–6976. [Google Scholar] [CrossRef] [PubMed]
  85. Chen, Z.; Mou, K.; Wang, X.; Liu, L. Nitrogen-Doped Graphene Quantum Dots Enhance the Activity of Bi2O3 Nanosheets for Electrochemical Reduction of CO2 in a Wide Negative Potential Region. Angew. Chem. Int. Ed. 2018, 57, 12790–12794. [Google Scholar] [CrossRef] [PubMed]
  86. Zhao, Q.; Zhang, C.; Hu, R.; Du, Z.; Gu, J.; Cui, Y.; Chen, X.; Xu, W.; Cheng, Z.; Li, S.; et al. Selective Etching Quaternary MAX Phase toward Single Atom Copper Immobilized MXene (Ti3C2Clx) for Efficient CO2 Electroreduction to Methanol. ACS Nano 2021, 15, 4927–4936. [Google Scholar] [CrossRef] [PubMed]
  87. Wang, Y.; Chen, Z.; Han, P.; Du, Y.; Gu, Z.; Xu, X.; Zheng, G. Single-Atomic Cu with Multiple Oxygen Vacancies on Ceria for Electrocatalytic CO2 Reduction to CH4. ACS Catal. 2018, 8, 7113–7119. [Google Scholar] [CrossRef]
  88. Xiong, L.; Zhang, X.; Chen, L.; Deng, Z.; Han, S.; Chen, Y.; Zhong, J.; Sun, H.; Lian, Y.; Yang, B.; et al. Geometric Modulation of Local CO Flux in Ag@Cu2O Nanoreactors for Steering the CO2RR Pathway toward High-Efficacy Methane Production. Adv. Mater. 2021, 33, 2101741. [Google Scholar] [CrossRef]
  89. Han, Z.; Han, D.; Chen, Z.; Gao, J.; Jiang, G.; Wang, X.; Lyu, S.; Guo, Y.; Geng, C.; Yin, L.; et al. Steering surface reconstruction of copper with electrolyte additives for CO2 electroreduction. Nat. Commun. 2022, 13, 3158. [Google Scholar] [CrossRef] [PubMed]
  90. Zhang, W.; Qin, Q.; Dai, L.; Qin, R.; Zhao, X.; Chen, X.; Ou, D.; Chen, J.; Chuong, T.T.; Wu, B.; et al. Electrochemical Reduction of Carbon Dioxide to Methanol on Hierarchical Pd/SnO2 Nanosheets with Abundant Pd-O-Sn Interfaces. Angew. Chem. Int. Ed. 2018, 57, 9475–9479. [Google Scholar] [CrossRef]
  91. Ji, L.; Chang, L.; Zhang, Y.; Mou, S.; Wang, T.; Luo, Y.; Wang, Z.; Sun, X. Electrocatalytic CO2 Reduction to Alcohols with High Selectivity over a Two-Dimensional Fe2P2S6 Nanosheet. ACS Catal. 2019, 9, 9721–9725. [Google Scholar] [CrossRef]
  92. Mou, S.; Wu, T.; Xie, J.; Zhang, Y.; Ji, L.; Huang, H.; Wang, T.; Luo, Y.; Xiong, X.; Tang, B.; et al. Boron Phosphide Nanoparticles: A Nonmetal Catalyst for High-Selectivity Electrochemical Reduction of CO2 to CH3OH. Adv. Mater. 2019, 31, 1903499. [Google Scholar] [CrossRef]
  93. Guan, A.; Yang, C.; Wang, Q.; Qian, L.; Cao, J.; Zhang, L.; Wu, L.; Zheng, G. Atomic-Level Copper Sites for Selective CO2 Electroreduction to Hydrocarbon. ACS Sustain. Chem. Eng. 2021, 9, 13536–13544. [Google Scholar] [CrossRef]
  94. Hori, Y.; Takahashi, I.; Koga, O.; Hoshi, N. Electrochemical reduction of carbon dioxide at various series of copper single crystal electrodes. J. Mol. Catal. A Chem. 2003, 199, 39–47. [Google Scholar] [CrossRef]
  95. De Gregorio, G.L.; Burdyny, T.; Loiudice, A.; Iyengar, P.; Smith, W.A.; Buonsanti, R. Facet-Dependent Selectivity of Cu Catalysts in Electrochemical CO2 Reduction at Commercially Viable Current Densities. ACS Catal. 2020, 10, 4854–4862. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Rahaman, M.; Dutta, A.; Zanetti, A.; Broekmann, P. Electrochemical Reduction of CO2 into Multicarbon Alcohols on Activated Cu Mesh Catalysts: An Identical Location (IL) Study. ACS Catal. 2017, 7, 7946–7956. [Google Scholar] [CrossRef]
  97. Wakerley, D.; Lamaison, S.; Ozanam, F.; Menguy, N.; Mercier, D.; Marcus, P.; Fontecave, M.; Mougel, V. Bio-inspired hydrophobicity promotes CO2 reduction on a Cu surface. Nat. Mater. 2019, 18, 1222–1227. [Google Scholar] [CrossRef] [PubMed]
  98. Vasileff, A.; Xu, C.; Jiao, Y.; Zheng, Y.; Qiao, S.-Z. Surface and Interface Engineering in Copper-Based Bimetallic Materials for Selective CO2 Electroreduction. Chem 2018, 4, 1809–1831. [Google Scholar] [CrossRef] [Green Version]
  99. Chen, C.; Yan, X.; Liu, S.; Wu, Y.; Wan, Q.; Sun, X.; Zhu, Q.; Liu, H.; Ma, J.; Zheng, L.; et al. Highly Efficient Electroreduction of CO2 to C2+ Alcohols on Heterogeneous Dual Active Sites. Angew. Chem. Int. Ed. 2020, 59, 16459–16464. [Google Scholar] [CrossRef]
  100. Wang, X.; Wang, Z.; de Arquer, F.P.G.; Cao-Thang, D.; Ozden, A.; Li, Y.C.; Nam, D.-H.; Li, J.; Liu, Y.-S.; Wicks, J.; et al. Efficient electrically powered CO2-to-ethanol via suppression of deoxygenation. Nat. Energy 2020, 5, 478–486. [Google Scholar] [CrossRef]
  101. Chen, C.; Li, Y.; Yang, P. Address the “alkalinity problem” in CO2 electrolysis with catalyst design and translation. Joule 2021, 5, 737–742. [Google Scholar] [CrossRef]
  102. Furuya, N.; Yamazaki, T.; Shibata, M. High performance RuPd catalysts for CO2 reduction at gas-diffusion electrodes. J. Electroanal. Chem. 1997, 431, 39–41. [Google Scholar] [CrossRef]
  103. Nguyen, T.N.; Dinh, C.-T. Gas diffusion electrode design for electrochemical carbon dioxide reduction. Chem. Soc. Rev. 2020, 49, 7488–7504. [Google Scholar] [CrossRef]
  104. Kibria, M.G.; Edwards, J.P.; Gabardo, C.M.; Dinh, C.-T.; Seifitokaldani, A.; Sinton, D.; Sargent, E.H. Electrochemical CO2 Reduction into Chemical Feedstocks: From Mechanistic Electrocatalysis Models to System Design. Adv. Mater. 2019, 31, 1807166. [Google Scholar] [CrossRef]
  105. de Arquer, F.P.G.; Cao-Thang, D.; Ozden, A.; Wicks, J.; McCallum, C.; Kirmani, A.R.; Dae-Hyun, N.; Gabardo, C.; Seifitokaldani, A.; Wang, X.; et al. CO2 electrolysis to multicarbon products at activities greater than 1 A cm−2. Science 2020, 367, 661–666. [Google Scholar] [CrossRef]
  106. Xie, H.; Wan, Y.; Wang, X.; Liang, J.; Lu, G.; Wang, T.; Chai, G.; Adli, N.M.; Priest, C.; Huang, Y.; et al. Boosting Pd-catalysis for electrochemical CO2 reduction to CO on Bi-Pd single atom alloy nanodendrites. Appl. Catal. B 2021, 289, 119783. [Google Scholar] [CrossRef]
  107. Jeong, H.-Y.; Balamurugan, M.; Choutipalli, V.S.K.; Jeong, E.-s.; Subramanian, V.; Sim, U.; Nam, K.T. Achieving highly efficient CO2 to CO electroreduction exceeding 300 mA cm−2 with single-atom nickel electrocatalysts. J. Mater. Chem. A 2019, 7, 10651–10661. [Google Scholar] [CrossRef]
  108. Abbas, S.A.; Song, J.T.; Tan, Y.C.; Nam, K.M.; Oh, J.; Jung, K.-D. Synthesis of a Nickel Single-Atom Catalyst Based on Ni–N4–xCx Active Sites for Highly Efficient CO2 Reduction Utilizing a Gas Diffusion Electrode. ACS Appl. Energ. Mater. 2020, 3, 8739–8745. [Google Scholar] [CrossRef]
  109. Yang, H.; Lin, Q.; Zhang, C.; Yu, X.; Cheng, Z.; Li, G.; Hu, Q.; Ren, X.; Zhang, Q.; Liu, J.; et al. Carbon dioxide electroreduction on single-atom nickel decorated carbon membranes with industry compatible current densities. Nat. Commun. 2020, 11, 593. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. Dinh, C.-T.; Burdyny, T.; Kibria, M.G.; Seifitokaldani, A.; Gabardo, C.M.; Arquer, F.P.G.d.; Kiani, A.; Edwards, J.P.; Luna, P.D.; Bushuyev, O.S.; et al. CO2 electroreduction to ethylene via hydroxide-mediated copper catalysis at an abrupt interface. Science 2018, 360, 783–787. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  111. Lin, R.; Guo, J.; Li, X.; Patel, P.; Seifitokaldani, A. Electrochemical Reactors for CO2 Conversion. Catalysts 2020, 10, 473. [Google Scholar] [CrossRef]
  112. Peng, C.; Luo, G.; Xu, Z.; Yan, S.; Zhang, J.; Chen, M.; Qian, L.; Wei, W.; Han, Q.; Zheng, G. Lithiation-Enabled High-Density Nitrogen Vacancies Electrocatalyze CO2 to C2 Products. Adv. Mater. 2021, 33, 2103150. [Google Scholar] [CrossRef] [PubMed]
  113. Peng, C.; Zhu, X.; Xu, Z.; Yan, S.; Chang, L.Y.; Wang, Z.; Zhang, J.; Chen, M.; Sham, T.-K.; Li, Y.; et al. Lithium Vacancy-Tuned [CuO4] Sites for Selective CO2 Electroreduction to C2+ Products. Small 2022, 18, 2106433. [Google Scholar] [CrossRef]
  114. Yan, S.; Peng, C.; Yang, C.; Chen, Y.; Zhang, J.; Guan, A.; Lv, X.; Wang, H.; Wang, Z.; Sham, T.-K.; et al. Electron Localization and Lattice Strain Induced by Surface Lithium Doping Enable Ampere-Level Electrosynthesis of Formate from CO2. Angew. Chem. Int. Ed. 2021, 60, 25741–25745. [Google Scholar] [CrossRef]
  115. Duan, Y.-X.; Meng, F.-L.; Liu, K.-H.; Yi, S.-S.; Li, S.-J.; Yan, J.-M.; Jiang, Q. Amorphizing of Cu Nanoparticles toward Highly Efficient and Robust Electrocatalyst for CO2 Reduction to Liquid Fuels with High Faradaic Efficiencies. Adv. Mater. 2018, 30, 1706194. [Google Scholar] [CrossRef]
  116. Spöri, C.; Kwan, J.T.H.; Bonakdarpour, A.; Wilkinson, D.P.; Strasser, P. The Stability Challenges of Oxygen Evolving Catalysts: Towards a Common Fundamental Understanding and Mitigation of Catalyst Degradation. Angew. Chem. Int. Ed. 2017, 56, 5994–6021. [Google Scholar] [CrossRef]
  117. Ma, L.; Hu, W.; Pan, Q.; Zou, L.; Zou, Z.; Wen, K.; Yang, H. Polyvinyl alcohol-modified gold nanoparticles with record-high activity for electrochemical reduction of CO2 to CO. J. CO2 Util. 2019, 34, 108–114. [Google Scholar] [CrossRef]
  118. Zhang, L.; Wei, Z.; Thanneeru, S.; Meng, M.; Kruzyk, M.; Ung, G.; Liu, B.; He, J. A Polymer Solution To Prevent Nanoclustering and Improve the Selectivity of Metal Nanoparticles for Electrocatalytic CO2 Reduction. Angew. Chem. Int. Ed. 2019, 58, 15834–15840. [Google Scholar] [CrossRef] [PubMed]
  119. Xiao, S.; Hu, W.; Luo, W.; Wu, Y.; Li, X.; Deng, H. Size effect on alloying ability and phase stability of immiscible bimetallic nanoparticles. Eur. Phys. J. B 2006, 54, 479–484. [Google Scholar] [CrossRef]
  120. Papaefthimiou, V.; Diebold, M.; Ulhaq-Bouillet, C.; Doh, W.H.; Blume, R.; Zafeiratos, S.; Savinova, E.R. Potential-Induced Segregation Phenomena in Bimetallic PtAu Nanoparticles: An In Situ Near-Ambient-Pressure Photoelectron Spectroscopy Study. ChemElectroChem 2015, 2, 1519–1526. [Google Scholar] [CrossRef]
  121. Sun, K.; Cheng, T.; Wu, L.; Hu, Y.; Zhou, J.; Maclennan, A.; Jiang, Z.; Gao, Y.; Goddard, W.A.; Wang, Z. Ultrahigh Mass Activity for Carbon Dioxide Reduction Enabled by Gold–Iron Core–Shell Nanoparticles. J. Am. Chem. Soc. 2017, 139, 15608–15611. [Google Scholar] [CrossRef] [PubMed]
  122. Zhu, W.; Zhang, L.; Yang, P.; Hu, C.; Dong, H.; Zhao, Z.-J.; Mu, R.; Gong, J. Formation of Enriched Vacancies for Enhanced CO2 Electrocatalytic Reduction over AuCu Alloys. ACS Energy Lett. 2018, 3, 2144–2149. [Google Scholar] [CrossRef]
  123. Rabinowitz, J.A.; Kanan, M.W. The future of low-temperature carbon dioxide electrolysis depends on solving one basic problem. Nat. Commun. 2020, 11, 5231. [Google Scholar] [CrossRef]
  124. Overa, S.; Feric, T.G.; Park, A.-H.A.; Jiao, F. Tandem and Hybrid Processes for Carbon Dioxide Utilization. Joule 2021, 5, 8–13. [Google Scholar] [CrossRef]
  125. Luo, Y.; Zhang, K.; Li, Y.; Wang, Y. Valorizing carbon dioxide via electrochemical reduction on gas-diffusion electrodes. InfoMat 2021, 3, 1313–1332. [Google Scholar] [CrossRef]
  126. Mustain, W.E.; Chatenet, M.; Page, M.; Kim, Y.S. Durability challenges of anion exchange membrane fuel cells. Energy Environ. Sci. 2020, 13, 2805–2838. [Google Scholar] [CrossRef]
  127. Huang, J.E.; Li, F.; Ozden, A.; Sedighian Rasouli, A.; García de Arquer, F.P.; Liu, S.; Zhang, S.; Luo, M.; Wang, X.; Lum, Y.; et al. CO2 electrolysis to multicarbon products in strong acid. Science 2021, 372, 1074–1078. [Google Scholar] [CrossRef]
  128. Gu, J.; Liu, S.; Ni, W.; Ren, W.; Haussener, S.; Hu, X. Modulating electric field distribution by alkali cations for CO2 electroreduction in strongly acidic medium. Nat. Catal. 2022, 5, 268–276. [Google Scholar] [CrossRef]
  129. Shafaque, H.W.; Lee, C.; Fahy, K.F.; Lee, J.K.; LaManna, J.M.; Baltic, E.; Hussey, D.S.; Jacobson, D.L.; Bazylak, A. Boosting Membrane Hydration for High Current Densities in Membrane Electrode Assembly CO2 Electrolysis. ACS Appl. Mater. Interfaces 2020, 12, 54585–54595. [Google Scholar] [CrossRef]
  130. Li, Y.C.; Lee, G.; Yuan, T.; Wang, Y.; Nam, D.-H.; Wang, Z.; García de Arquer, F.P.; Lum, Y.; Dinh, C.-T.; Voznyy, O.; et al. CO2 Electroreduction from Carbonate Electrolyte. ACS Energy Lett. 2019, 4, 1427–1431. [Google Scholar] [CrossRef]
  131. Li, T.; Lees, E.W.; Zhang, Z.; Berlinguette, C.P. Conversion of Bicarbonate to Formate in an Electrochemical Flow Reactor. ACS Energy Lett. 2020, 5, 2624–2630. [Google Scholar] [CrossRef]
  132. Zhang, Z.; Lees, E.W.; Habibzadeh, F.; Salvatore, D.A.; Ren, S.; Simpson, G.L.; Wheeler, D.G.; Liu, A.; Berlinguette, C.P. Porous metal electrodes enable efficient electrolysis of carbon capture solutions. Energy Environ. Sci. 2022, 15, 705–713. [Google Scholar] [CrossRef]
  133. Lees, E.W.; Goldman, M.; Fink, A.G.; Dvorak, D.J.; Salvatore, D.A.; Zhang, Z.; Loo, N.W.X.; Berlinguette, C.P. Electrodes Designed for Converting Bicarbonate into CO. ACS Energy Lett. 2020, 5, 2165–2173. [Google Scholar] [CrossRef]
Figure 1. CO2RR pathways towards C1 and C2+ products on metal-based catalysts.
Figure 1. CO2RR pathways towards C1 and C2+ products on metal-based catalysts.
Catalysts 12 00860 g001
Figure 2. (a) AC HAADF-STEM image of Cu-N2/GN (the isolated Cu single atoms dispersed on graphene as labeled by the red circle). (b) FT-EXAFS of Cu–N2/GN [51]; copyright John Wiley & Son, 2019. (c) Calculated free energy of CO2RR on (Cl, N)−Mn/G and N−Mn/G, an * denotes the catalytic site [67]; copyright Springer Nature, 2019. (d) FECO of the Co–N–Ni/NPCNSs, Co–N/NPCNSs, and Ni–N/NPCNSs catalysts under electrochemical bias [55]; copyright Royal Society of Chemistry, 2021.
Figure 2. (a) AC HAADF-STEM image of Cu-N2/GN (the isolated Cu single atoms dispersed on graphene as labeled by the red circle). (b) FT-EXAFS of Cu–N2/GN [51]; copyright John Wiley & Son, 2019. (c) Calculated free energy of CO2RR on (Cl, N)−Mn/G and N−Mn/G, an * denotes the catalytic site [67]; copyright Springer Nature, 2019. (d) FECO of the Co–N–Ni/NPCNSs, Co–N/NPCNSs, and Ni–N/NPCNSs catalysts under electrochemical bias [55]; copyright Royal Society of Chemistry, 2021.
Catalysts 12 00860 g002
Figure 3. (a) FEs and product distribution at different polarization potentials catalyzed by Cu/C with the optimum loading amount. (b) The hypothesized reaction mechanism involving the formation of Cu clusters through the operando measurements, the Cu2+ was reduced to Cun ligated by the surface hydroxyl group in II, and CO2 is reduced to ethanol via III, IV, V. [62]; copyright Springer Nature, 2020. (c) Catalytic performance of Au/Cu, copper, and gold catalysts (rate of CO2 reduction to >2e products). (d) Partial current densities to alcohols [63]; copyright Springer Nature, 2018. (e) Onset potentials for CO and C2H4 on X-Cu catalysts [61]; copyright Springer Nature, 2020.
Figure 3. (a) FEs and product distribution at different polarization potentials catalyzed by Cu/C with the optimum loading amount. (b) The hypothesized reaction mechanism involving the formation of Cu clusters through the operando measurements, the Cu2+ was reduced to Cun ligated by the surface hydroxyl group in II, and CO2 is reduced to ethanol via III, IV, V. [62]; copyright Springer Nature, 2020. (c) Catalytic performance of Au/Cu, copper, and gold catalysts (rate of CO2 reduction to >2e products). (d) Partial current densities to alcohols [63]; copyright Springer Nature, 2018. (e) Onset potentials for CO and C2H4 on X-Cu catalysts [61]; copyright Springer Nature, 2020.
Catalysts 12 00860 g003
Figure 4. (a) FEs (top panel, left y-axis), partial current densities of CH4 and C2H4 (top panel, right y-axis), and ratios of CH4/C2H4 (bottom panel) on Cu−N−C−T (T = 800, 900, 1000, and 1100 °C) catalysts at −1.6 V vs RHE [69]; copyright American Chemical Society, 2020. (b) FEs of CO2RR products at different applied potentials catalyzed by Cu-DBC [77]; copyright Springer Nature, 2021. (c) Comparison of the ratios of FEmethane to FEC2+ on Cu catalysts at various CO2 concentrations [32]; copyright American Chemical Society, 2020. (d) Schematic illustration of the fabrication of SA-Cu-MXene via selective etching quaternary MAX-Ti3(Al1−xCux)C2 (Gray, blue, red, yellow, brown, and green balls represent Al, Ti, Cu, O, C, and Cl atoms, respectively) [86]; copyright American Chemical Society, 2021. (e) Performance of CO2RR catalyzed by different dual-doping catalysts at the potential of −1.18 V vs. RHE [79]; copyright Springer Nature, 2021.
Figure 4. (a) FEs (top panel, left y-axis), partial current densities of CH4 and C2H4 (top panel, right y-axis), and ratios of CH4/C2H4 (bottom panel) on Cu−N−C−T (T = 800, 900, 1000, and 1100 °C) catalysts at −1.6 V vs RHE [69]; copyright American Chemical Society, 2020. (b) FEs of CO2RR products at different applied potentials catalyzed by Cu-DBC [77]; copyright Springer Nature, 2021. (c) Comparison of the ratios of FEmethane to FEC2+ on Cu catalysts at various CO2 concentrations [32]; copyright American Chemical Society, 2020. (d) Schematic illustration of the fabrication of SA-Cu-MXene via selective etching quaternary MAX-Ti3(Al1−xCux)C2 (Gray, blue, red, yellow, brown, and green balls represent Al, Ti, Cu, O, C, and Cl atoms, respectively) [86]; copyright American Chemical Society, 2021. (e) Performance of CO2RR catalyzed by different dual-doping catalysts at the potential of −1.18 V vs. RHE [79]; copyright Springer Nature, 2021.
Catalysts 12 00860 g004
Figure 5. (a) Aberration-corrected HAADF-TEM images of Cu(OH)2-D [25]; copyright John Wiley & Son, 2021. (b) The relationship between the ethylene FE and the ratio of atop CO and bridge CO on Cu–x electrodes. (c) FE of ethylene on Cu and Cu–12 using CO2-saturated 1 M KHCO3 as the supporting electrolyte [29]; copyright Springer Nature, 2019. (d) t-SNE19 representation of approximately 4000 adsorption sites performed by DFT calculations with Cu-containing alloys (Cu-Al clusters labeled numerically) [26]; copyright Springer Nature, 2020. (e) Efficiency of acetate formation as a function of the atom ratios of Cu and Ag in CumAgn/polymer/GCE in CO2-saturated 0.5 M KHCO3, −1.53 V vs. RHE after 3600 s at 0 °C [81]; copyright National Academy of Sciences, 2018. (f) Schematic illustration of the morphology and product selectivity of a dendritic Cu catalyst before and after thermal annealing for electrocatalytic CO2RR [96]; copyright American Chemical Society, 2017. (g) FEn-PrOH and the ratio of FEn-PrOH/FEC1+C2+C3 of the CuS, CuSx-1-cycle, CuSx-DSV, and CuSx-100-cycle catalysts at −1.05 V vs. RHE [82]; copyright Springer Nature, 2021.
Figure 5. (a) Aberration-corrected HAADF-TEM images of Cu(OH)2-D [25]; copyright John Wiley & Son, 2021. (b) The relationship between the ethylene FE and the ratio of atop CO and bridge CO on Cu–x electrodes. (c) FE of ethylene on Cu and Cu–12 using CO2-saturated 1 M KHCO3 as the supporting electrolyte [29]; copyright Springer Nature, 2019. (d) t-SNE19 representation of approximately 4000 adsorption sites performed by DFT calculations with Cu-containing alloys (Cu-Al clusters labeled numerically) [26]; copyright Springer Nature, 2020. (e) Efficiency of acetate formation as a function of the atom ratios of Cu and Ag in CumAgn/polymer/GCE in CO2-saturated 0.5 M KHCO3, −1.53 V vs. RHE after 3600 s at 0 °C [81]; copyright National Academy of Sciences, 2018. (f) Schematic illustration of the morphology and product selectivity of a dendritic Cu catalyst before and after thermal annealing for electrocatalytic CO2RR [96]; copyright American Chemical Society, 2017. (g) FEn-PrOH and the ratio of FEn-PrOH/FEC1+C2+C3 of the CuS, CuSx-1-cycle, CuSx-DSV, and CuSx-100-cycle catalysts at −1.05 V vs. RHE [82]; copyright Springer Nature, 2021.
Catalysts 12 00860 g005
Table 1. Recent research on onset potential for CO2RR products (* The onset potential is not clearly determined and this value is the lowest applied bias with decent FE).
Table 1. Recent research on onset potential for CO2RR products (* The onset potential is not clearly determined and this value is the lowest applied bias with decent FE).
Major ProductCatalystOnset Potential
(V vs. RHE)
Reference
COCu-N2/CN−0.33[51]
FeN5−0.2 *[52]
Ni–N3S−0.17[53]
Fe3+-N-C−0.2[54]
Co–N–Ni/NPCNSs−0.2[55]
CoPc©Fe-N-C−0.13[56]
Formatesingle-atom Snδ+ on N-doped graphene−0.18[57]
BiN4/C−0.51[58]
MethaneAuAgPtPdCu−0.3 *[59]
EthyleneOrganosuperbases modified Cu-NC−0.43[60]
F-Cu~−0.2[61]
EthanolCun (n = 3 and 4) cluster−0.3~−0.4[62]
Au/Cu−0.7 *[63]
FeTPP[Cl]/Cu−0.42[64]
AcetateCu–Cu2O/Cu~−0.2 *[65]
Table 2. Recent progress on promoting current density and FE for CO2RR products.
Table 2. Recent progress on promoting current density and FE for CO2RR products.
Major ProductCatalystsCurrent Density
(mA·cm−2)
Faradaic Efficiency
(%)
Electrode ConfigurationReference
CO5-fold twinned Ag NWs~2.299.3Non-GDE electrode[72]
ZrO2@Ni-NC20098.6GDE[73]
FormatenBuLi-Bi50092[74]
Bi-NRs@NCNTs690.9Non-GDE electrode[75]
MethaneSA-Zn/MNC31.885[76]
Cu-DBC20380GDE[77]
CuGaO271771.7[41]
MethanolCu1.63Se(1/3)41.577.6Non-GDE electrode[78]
Ag,S-Cu2O/Cu122.767.4[79]
EthyleneCu(OH)2-D/Cu25058GDE[25]
F-Cu160065[61]
Cu-Al40080[26]
EthanolCu3Ag12563Non-GDE electrode[27]
Cun (n = 3 and 4) cluster~291[62]
Ag0.14/Cu0.8625041GDE[80]
AcetateCu2Ag3~0.921.2Non-GDE electrode[81]
n-PropanolCuSX-DSV9.915.4[82]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wu, B.; Chen, J.; Qian, L. Recent Advances in Heterogeneous Electroreduction of CO2 on Copper-Based Catalysts. Catalysts 2022, 12, 860. https://doi.org/10.3390/catal12080860

AMA Style

Wu B, Chen J, Qian L. Recent Advances in Heterogeneous Electroreduction of CO2 on Copper-Based Catalysts. Catalysts. 2022; 12(8):860. https://doi.org/10.3390/catal12080860

Chicago/Turabian Style

Wu, Bowen, Jian Chen, and Linping Qian. 2022. "Recent Advances in Heterogeneous Electroreduction of CO2 on Copper-Based Catalysts" Catalysts 12, no. 8: 860. https://doi.org/10.3390/catal12080860

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop