Download PDF
Review  |  Open Access  |  17 Jan 2025

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Views: 122 |  Downloads: 42 |  Cited:  0
Microstructures 2025, 5, 2025003.
10.20517/microstructures.2024.69 |  © The Author(s) 2025.
Author Information
Article Notes
Cite This Article

Abstract

Carbon dioxide reduction reaction (CO2RR) is an efficacious method to mitigate carbon emissions and simultaneously convert CO2 into high-value carbon products. The efficiency of CO2RR depends on the development of highly active and selective catalysts. Copper (Cu)-based catalysts can effectively reduce CO2 to hydrocarbons and oxygen-containing compounds because of their unique geometric and electronic structures. Most importantly, Cu can reduce CO2 to multiple carbon products (C2+). Therefore, this review aims to outline recent research progress in Cu-based catalysts for CO2RR. After introducing the mechanism of this electroreduction reaction, we summarize the influence of the size, morphology, and coordination environment of single component Cu-based catalysts on their performance, especially the performance control of catalysts that contain nano Cu or Cu single-atom sites. Then, the synergistic regulation strategies of doping other metals into Cu-based catalysts are summarized. Finally, the research on the supports used for Cu-based catalysts is reviewed. The prospects and challenges of Cu-based catalysts are discussed.

Keywords

CO2RR, copper, nanoparticles, single atoms, electrocatalyst

INTRODUCTION

The world is currently facing a severe energy crisis. The problems of air pollution, plastic pollution, water pollution, and carbon dioxide (CO2) emissions become increasingly serious and need to be urgently addressed[1-3]. In the past century, the concentration of CO2 has increased due to rapid energy consumption, leading to serious global warming problems[4]. At present, many methods have been proposed to alleviate climate change, including the capture and sequestration of CO2, which is still a big challenge due to the relative chemical inertness of CO2[5]. Among them, the electrochemical method can convert CO2 into more valuable carbon products[6-8]. Metal catalysts, such as Au, Ag, Cu, Pd, etc., can effectively reduce CO2. Cu-based catalysts are the focus because they can efficiently catalyze CO2 into multiple carbon products[9]. The formation of *C2 intermediates requires sufficient surface CO2 concentration and *C1 coverage. However, since these intermediates need to be able to move freely, the adsorption strength of the catalyst used for *C1 intermediates should not be too high. Cu has moderate adsorption strength for these intermediates, so it is considered to be the most favorable heterogeneous catalyst for the electroreduction of CO2[10-15]. Although Cu has a high binding strength to *CO, it has disadvantages of high overpotential and low selectivity for C2+ products, such as ethylene (C2H4), ethanol (CH3CH2OH), acetic acid (CH3COOH), acetone (CH3COCH3) and propanol (CH3CH2CH2OH)[16,17].

In 1870, it was first discovered that CO2 could be electro-reduced to formic acid (HCOOH) in aqueous media[18,19]. In 1985, Hori et al. were the first person to observe that Cu can reduce CO2 to methane (CH4) and C2H4[20]. In particular, C2H4 is only produced on Cu electrodes. They found that the selectivity of CH4 and C2H4 is related to the pH and cation of the electrolyte. A higher pH value is conducive to the formation of C2H4, and the selectivity of C2H4 improves with the increase in cation size[21,22]. Since then, many researchers have conducted research on Cu-based catalysts for the reduction reaction of CO2 (CO2RR).

Improving the selectivity of Cu-based catalysts to CO2RR products is a current research focus. However, due to the wide variety of CO2RR reaction pathways and product types, the high-selective formation of a certain product is challenging. Cu-based electrocatalysts have made significant progress as key materials for CO2RR technology in recent years. Modifying Cu-based catalysts by doping or changing crystal planes, sizes, and morphologies can affect the bonding strength of important electroreduction intermediates such as *CO and *OCHO, which enables to improve the function of Cu-based catalysts. In addition, electrolytes and applied currents can alter the structure of Cu-based catalyst, causing changes in catalytic performance. In recent years, many researchers have found that the selectivity of Cu-based catalysts to certain reduction products in CO2RR can be effectively improved through nanoconfinement effects, composite catalysts, introducing hydrophilic metals, and increasing surface hydrophobicity of catalysts.

This article reviews recently important progress on Cu-based catalysts for CO2RR. Firstly, the synergistic control strategies of the morphology, size, and chemical environment of single-component Cu-based catalysts and their effects on catalyst performance are discussed based on the active components of the catalysts, especially the performance control of nano Cu and Cu single-atom sites. Subsequently, the synergistic regulation strategies of Cu-based catalysts doped with other metals are summarized. Finally, the supports of Cu-based catalysts are summed up. Additionally, prospects and challenges are discussed.

REACTION MECHANISM OF CO2RR

The catalytic processes of CO2RR typically involve the following three steps[6]:

(1) CO2 first adsorbs on the catalyst, forming *CO2[19,23];

(2) The transfer of electrons and protons causes the dissociation of C=O bonds, resulting in the formation of C-H and C-O bonds;

(3) Product desorbs from the catalyst surface[24].

At the thermodynamic level, transferring the first electron to the CO2 molecule requires large recombination energy (750 kJ mol-1) to activate CO2[25,26]. On Cu, CO is a key intermediate for CO2RR to produce hydrocarbons[27].

As shown in Figure 1, the different pathways of CO2RR reactions result in the production of various intermediate products, such as C1, C2, and C3, and involve multi-electron or proton transfer[28].

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 1. Possible mechanistic pathways of ECO2RR towards C1 and C2 products. (Reproduced with permission[19]. Copyright 2022, Royal Society of Chemistry). ECO2RR: Electrocatalytic reduction of CO2.

Table 1 lists the semi-reactions of different products and their standard reduction potentials relative to reversible hydrogen electrodes (RHE) under alkaline conditions[28,29]. Among them, the hydrogen evolution reaction (HER) is a competitive reaction[30-33]. However, a variety of products are usually detected on the surface of Cu, resulting in poor selectivity of products. Moreover, competitive HER and inefficient *CO dimerization in aqueous solution still limit the performance of CO2RR[34]. Therefore, it is necessary to regulate the selectivity and suppress HER through various means, e.g., by regulating catalyst morphology, adding additives, and modifying the catalyst surface.

Table 1

The semi-reactions of CO2RR and their standard reduction potentials versus RHE

ProductsEquationPotential (V)
H22H2O + 2e- → H2 + 2OH--0.828
COCO2 + 2H2O + 2e- → CO + 2OH--0.932
HCOOHCO2 + H2O + 2e- → HCOO- + OH--0.639
CH3OHCO2 + 5H2O + 6e- → CH3OH + 6OH--0.812
CH4CO2 + 6H2O + 8e-  → CH4 + 8OH--0.659
C2H42CO2 + 8H2O + 12e- → C2H4 + 12OH--0.743
C2H62CO2 + 10H2O + 14e- → C2H6 + 14OH--0.685
CH3CH2OH2CO2 + 9H2O + 12e- → CH3CH2OH + 12OH--0.744
CH3CHO2CO2 + 7H2O + 10e- → CH3CHO + 10OH--0.775
CH3COOH2CO2 +  5H2O +  8e- → CH3COO- + 7OH--0.653
n-C3H7OH3CO2 + 13H2O + 18e- → n-C3H7OH + 18OH--0.733
CH3COCH33CO2 + 11H2O + 16e- → CH3COCH3 + 16OH--0.726

SINGLE-COMPONENT CU-BASED CATALYSTS

The active components of a catalyst refer to the species that play a catalytic role in the catalyst. At present, transition metals are mainly the active component of CO2RR catalysts. As shown in Figure 2, metals can be divided into four categories according to surface-bound intermediates[35]. Ni, Fe, Pt, and Ti easily catalyze H2 formation via HER during the CO2RR process[36]. The surfaces of Au, Ag, Zn, and Pd mainly produce CO[37]. Pb, In, Sn, and other metals preferably generate HCOOH[25,38-40]. Cu can further reduce the intermediate *CO in the CO2RR process to produce high-value carbon products[35,41-44]. Accordingly, Cu is the focus of many researchers. In this section, we mainly summarized some representative Cu-based nanoparticles (NPs) and single-atom catalysts (SACs), the control functions of additives for Cu-based catalysts to regulate CO2RR catalysis, and the corresponding selective catalytic mechanisms.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 2. The classification of metal catalysts for CO2 reduction. (Reproduced with permission[35]. Copyright 2020, Wiley Materials).

Cu nanoparticles

Cu NPs refer to the catalysts with size of Cu particles in the nanometer range. The size, morphology, chemical environment, and crystal plane of Cu NPs have a significant influence on the products and performance of CO2RR. When the size of Cu particles decreases to the nanometer level[45,46], the catalyst exhibits higher surface activity and richer surface defects than the Cu particles with larger sizes[47-49]. The surface properties and structures of Cu NPs can be regulated by appropriate synthesis methods[50]. Thus, the selectivity and activity of the catalytic reaction can be adjusted[51]. Modifying the surface of Cu NPs with hydrophobic materials, such as Nafion[52,53], polytetrafluoroethylene (PTFE)[54], and alkyl thiol[55], can inhibit HER and promote CO2 mass transfer[56,57], thereby improving the efficiency of CO2RR and enhancing C2+ selectivity. Cu NPs are attractive catalysts for CO2RR to produce valuable chemicals[58]. Next, we will summarize some Cu NP catalysts with different morphologies, sizes, and chemical environments to regulate the catalytic effect of CO2RR and the corresponding selective catalytic mechanism.

Reske et al. synthesized six spherical Cu NP catalysts of different average sizes (2-15 nm) and coordination numbers (CNs) by stirring CuCl2 loaded micelles and changing the molecular weight of polyvinylpyrrolidone (PVP) heads or metal salt/PVP ratios [Figure 3A-F][59]. As shown in Figure 3G and H, the proportion of atoms with low CNs (Cu-CNs < 8) significantly increased in Cu NPs with particle sizes less than 6 nm. During CO2RR, the catalytic activity increased but tended to promote HER, mainly producing H2 and CO. Manthiram et al. prepared Cu NPs with a diameter of 7.0 ± 0.4 nm terminated with tetradecyl phosphonate by reducing cupric acetate Cu(Ac)2 in trioctylamine[60]. The Faraday efficiency (FE) of CH4 is 76% at -1.35 V vs. RHE. The current density of CH4 was four times that of Cu foil electrodes. They found that more isolated NPs exposed more catalytic active sites to form CH4. However, when the Cu NPs aggregated, CH4 would be lost in the products. Jung et al. prepared a 20 nm Cu2O NP/C with cubic morphology. Cu NPs were grown on the carbon carrier using cysteine molecules[61]. Under negative potential, the 20 nm cubic Cu2O crystal particles disintegrated into 2-4 nm particles, and FEs of C2H4 (FE of C2H4) and C2+ (FE C2+) of the shattered Cu-based NP/C catalyst reached 57.3% and 74%, respectively. Yang et al. reduced Cu(Ac)2 precursors with tetradecyl phosphonic acid at high temperatures to prepare Cu NPs, and controlled the size of Cu NPs by adjusting the ratio[62]. The authors established that the selectivity of C2+ increased with the proportion of metallic Cu nanograins. The 7 nm Cu NP with a unity fraction of Cu nanograins exhibited six times higher FE C2+ than the 18 nm Cu NP ensemble with one-third of Cu nanograins [Figure 3I and J]. As shown in Figure 3K, Zhang et al. adsorbed a mixture of Cu phthalocyanine (CuPc) and carbon black (CB) onto a glassy carbon wafer and electroreduced CuPc to Cu NPs (20 nm) under CO2RR conditions[63]. The size of Cu NPs was influenced by the reaction time, reduction potential, oxidation degree of the carbon support, and the loading amount of CuPc. Furthermore, due to the abundance of grain boundaries in large NPs, the selectivity of C2+ products also increased from 0% to 40% during the slow growth of Cu NPs (2 nm to 20 nm). The highest FE C2+ was 70%, and the current density for C2+ was 800 mA cm-2.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 3. (A-F) Tapping-mode AFM images of micellar Cu NPs; Particle size dependence of (G) the composition of gaseous reaction products (balance is CO2) during catalytic CO2 electroreduction over Cu NPs; (H) the faradic selectivities of reaction products during the CO2 electroreduction on Cu NPs. (Reproduced with permission[59]. Copyright 2014, American Chemical Society); (I) Structure-activity correlation of relative fraction of active Cu nanograins and C2+ FE of Cu NP ensembles with three different NP sizes; (J) FE for all CO2RR products of three different Cu NP sizes (7, 10 and 18 nm) at -0.8 V. (Reproduced with permission[62]. Copyright 2023, The Author(s), under exclusive license to Springer Nature Limited); (K) Structure-activity correlation of relative fraction of active Cu nanograins and C2+ FE of Cu NP ensembles with three different NP sizes.(Reproduced with permission[63]. Copyright 2023, American Chemical Society). NPs: Nanoparticles; CO2RR: Carbon dioxide reduction reaction; FE: Faraday efficiency; AFM: Atomic force microscopy.

Cu NPs with various morphologies and crystal faces have different effects on CO2RR. Zi et al. found that Cu nanoneedles can induce ultra-high local potassium concentration (4.22 M)[64]. High concentrations of potassium can promote the C-C coupling, achieving efficient CO2 reduction in 3-M KCl electrolytes of pH = 1. FE C2+ reached up to 90.69% ± 2.15% at 1,400 mA cm-2. Fu et al. prepared a series of Cu2O nanocrystals (NCs) with diverse crystal faces and morphologies[65]. Among them, the o-Cu2O NCs with high-index facets reached the highest FE C2+ (48.3%). They investigated the structural alterations of Cu2O NCs after stability tests and discovered that abundant crystal defects and high-index facets were observed, which may be the active sites. As shown in Figure 4A, Luo et al. modified the surface of Cu2O particles using reducing agents[66], and found that the star-shaped Cu2O NPs (F-Cu2O) with exposed (322) facets achieved the highest C2H4 selectivity (FE = 74.1%). They found that the Cu2O (332) surface can significantly reduce the free energy during the coupling process of *CHO intermediates and promote the production of C2H4. As shown in Figure 4B, they further modified the surface with ZIF-8. The ZIF-8 shell provided abundant pores, substance exchange channels and appropriate hydrophobic microenvironments, which facilitated the desorption of *CO intermediates. Periasamy et al. prepared stable Cu2O/polypyrrole (Cu2O/Ppy) particles on linen-textured (LT) paper[67]. The Cu2O/Ppy with octahedral and micro flower shapes had exposed (111) facets, (311) and (211) facets, exhibiting high selectivity towards CH3OH products. At -0.85 V vs. RHE, the FE of methanol (FE of CH3OH) was 93% ± 1.2%. Wu et al. prepared a highly stable mixed Cu2O-Cu nanocube catalyst [Cu2O (CO) ] by thermal reduction of CuO nanocubes under CO atmosphere[68]. The material exhibited a high-density CuO/Cu interface and abundant Cu (100) crystal planes. At an industrial current density of 500 mA cm-2, the FE of C2+ products was 77.4% (FE of C2H4 was 56.6%). Recently, as shown in Figure 4C-E, Geng et al. prepared hydrophobic porous Cu2O spheres (pore sizes 140/240/340 nm) using polystyrene (PS) spheres as pore guides by wet chemical methods[69]. The hydrophobic porous morphology of the catalyst facilitated the transportation and capture of CO2 and promoted CO2 activation and the formation of *OCCOH intermediates, thereby promoting C-C coupling. The P-Cu2O-240 sample had the highest CO2 mass transfer efficiency. As shown in Figure 4E, at -1.0 A cm-2, FE C2 was 75.3% ± 3.1%.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 4. (A) Synthesis of Cu2O NPs by the reductant-controlling method and Cu2O@ZIF-8 composites; (B) HAADF-STEM image and elemental mappings. (Reproduced with permission[66]. Copyright 2022, Wiley-VCH GmbH); (C) Scheme, SEM, and enlarged HAADF-STEM images of P-Cu2O-140, P-Cu2O-240, and P-Cu2O-340; (D) Water contact angle (CA) of P-Cu2O-140, P-Cu2O-240, P-Cu2O-340 (from top to bottom); (E) FEs for C2+ products under different applied current densities over porous Cu2O samples. (Reproduced with permission[69]. Copyright 2024, American Chemical Society). NPs: Nanoparticles; HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; FE: Faraday efficiency.

The grain boundaries of Cu NPs affect the selectivity of CO2RR. Frese et al. first discovered an increase in CH4 production on Cu (100), Cu (110), and Cu (111)[70]. Hori et al. found that Cu (100) and Cu (111) tended to generate C2H4 and CH4, respectively, while Cu (110) preferably produced acetate and acetaldehyde. They also reported that CO is a key intermediate in CO2RR[71]. Schouten et al. observed that the Cu (100) surface tended to form C2H4 at a relatively low overpotential, while the Cu (111) surface preferentially generated CH4 and only a small amount of C2H4[72]. Gao et al. found that the crystal facets exposed by Cu2O NPs greatly affected the selectivity of CO2-C2H4 [Figure 5A-F][73]. Cu2O NPs with both (111) and (100) crystal faces demonstrated a stronger selectivity for CO2-C2H4 with an FE of C2H4 of 59% compared to Cu2O NPs with only one crystal face. Wu et al. further proved this point[74]. They found that the Cu (100)/Cu (111) interface had a good localized electronic structure, which enhanced CO adsorption and C-C coupling, and its performance was better than that of the Cu (100) and Cu (111). Ma et al. reconstructed Cu NPs on vertical graphene [plasma-enhanced chemical vapor deposition (PECVD)][75]. They constructed incompatible sites and abundant oxygen vacancies on Cu active sites through reduction-oxidation-reduction (ROR), enhancing CO2RR activity. This catalyst had abundant grain boundaries, producing *COOH-derived products (CO) at high oxidation potentials (+1.2 V vs. RHE), and *OCHO-derived products (HCOOH) at low oxidation potentials (+0.8 V vs. RHE). Recently, Chen et al. utilized tannic acid (TA) molecules in situ to regulate and reconstruct Cu-based materials and prepared a Cu electrocatalyst (TA-Cu)[76]. Cu (111) NPs and partially oxidized CuOx were uniformly distributed on TA-Cu nanorods [Figure 5G]. The hydroxyl groups in TA could stabilize the key intermediate *COH through hydrogen bonding, thereby promoting the generation of C2H4. At -0.7 V vs. RHE, the process reached an FE of C2H4 of 63.6%, a stability of ten h, and a current density exceeding 497.2 mA cm-2[Figure 5H and I].

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 5. SEM images of (A, B) c-Cu2O NPs; (C, D) o-Cu2O NPs; and (E, F) t-Cu2O NPs. (Reproduced with permission[73]. Copyright 2020, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim); (G) The preparation of TA-Cu via electrochemical reconstruction of CuTA; (H) FEs of TA-Cu obtained at different applied potentials; (I) Stability for CO2 electrolysis with the TA-Cu. (Reproduced with permission[76]. Copyright 2023, Wiley-VCH GmbH). NPs: Nanoparticles; TA: Tannic acid ; FE: Faraday efficiency; SEM: Scanning electron microscope.

Carbon intermediates can be effectively limited by forming nano-sized confined spaces in Cu-based materials, thus effectively promoting the formation of C2+. For example, Liu et al. prepared a porous Cu nanosphere catalyst (P-Cu)[77], which can enrich the intermediate *CO in the pore structure. As shown in Figure 6A-C, Zhang et al. found a volcano curve relationship between the selectivity of C2H5OH and the nanocavity size of porous CuO in the range of 0 to 20 nm[78]. The increase in *OH coverage associated with the nanocavity size-dependent confinement effect was believed to account for the remarkable CH3CH2OH selectivity. Increased *OH coverage may facilitate the hydrogenation of *CHCOH to *CHCHOH (CH3CH2OH pathway). As shown in Figure 6D and E, p-CuO-(12.5 nm) has an FE of CH3CH2OH of 44.1% ± 1.0% at a high C2H5OH partial current density of 501.0 ± 15.0 mA cm-2, while FE C2+ was 90.6% ± 3.4%. Liu et al. increased the coverage of local CO by introducing a Cu hollow multi-shell structure (HoMSs) and utilizing the nano-constrained effect[79]. The presence of Cu in the cavity stabilized *CO and promoted dimerization. As the shell layer increased, the selectivity and activity of C2+ products significantly improved. The Cu HoMSs with three shells exhibited the largest FE C2+ in neutral electrolyte of 77.0% ± 0.3%, and the current density was 513.7 ± 0.7 mA cm-2. As shown in Figure 6F-H, based on mature CO2-CO electroreduction technology, Yang et al. reported porous Cu2O catalysts with nanocavities[80]. They found that fragmented and solid Cu2O catalysts mainly produced CO, while the FE of C2+ production on porous Cu2O catalysts was much higher, indicating that the porous structure could promote C-C bonding. At -0.61 Vvs. RHE, the FE C2+ was 75.2% ± 2.7%, and the partial current density was 267 ± 13 mA cm-2. They found that pore cavities could restrict the in-situ generation of carbon intermediates effectively, which bound with Cu sites [Figure 6I]. In addition, this restricted intermediate promoted C-C coupling within the reactive nanocavity.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 6. (A) TEM images of p-CuO-(12.5 nm); (B) HAADF-STEM images of p-CuO-(12.5 nm); (C) High-resolution HAADF-STEM images of p-CuO-(12.5 nm); (D) FEethanol and (E) FE C2+ over Cu NPs, p-CuO-(7.0 nm), p-CuO-(12.5 nm), and p-CuO-(19.4 nm) catalysts. (Reproduced with permission[78]. Copyright 2023, PNAS); Multihollow Cu2O catalyst imaged by SEM (F), TEM (G) and HRTEM (H); The inset in (G) shows the corresponding SAED pattern; (I) Schematic of carbon intermediates that are confined in the nanocavities. White: H; gray: C; red: O; violet: Cu. (Reproduced with permission[80]. Copyright 2020, American Chemical Society). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; NPs: Nanoparticles; FE: Faraday efficiency; TEM: Transmission electron microscope; HAADF-STEM: High-angle annular dark-field scanning transmission electron microscope; HRTEM: High-resolution transmission electron microscope; SAED: Selected-area electron diffraction.

Modifying the Cu NPs with different materials can enhance C-C coupling, thereby improving the production of multiple carbon products. For example, Zhao et al. reported an in-situ polymerization strategy of encapsulating hydrophobic polymers onto Cu NPs to synthesize fluorinated polymer-functionalized Cu NPs (Cu poly-1/2)[58]. Increasing the hydrophobicity of the catalyst surface can suppress HER, which is beneficial for reducing the thickness of the CO2 gas diffusion layer and promoting CO2 mass transfer[56,57]. Compared with bare Cu (FE C2+ = 61.59%), the use of Cu poly-1/2 increased the FE C2+ to 71.08% at -3.98 V vs. RHE. Pellessier et al. covered commercial Cu NPs with porous PTFE nanocoats[54]. The physical adsorption of PTFE on Cu NPs generated a large interface surface area, which could improve the binding energy of CO intermediates on Cu and enhance C-C coupling. The FE C2+ was 78% within the current density range of 400-500 mA cm-2. Using ammonia and Cu(NO3)2 as raw materials, Zhou et al. prepared AN-Cu(OH)2 by a one-step wet chemical method. Then, the acidic ionomer Nafion was used to regulate the surface hydrophobicity and local alkalinity of the catalyst [AN-Cu(OH)@Nafion][52]. The FE of C2H4 of optimal Nafion-activated AN-Cu(OH)@Nafion was 44%, and HER was effectively inhibited. Recently, Su et al. used Nafion to modify Cu NP catalyst (Cu@Nafion)[53]. Nafion can inhibit HER and enhance CO2 mass transfer, which is beneficial for C-C coupling. At -1.2V vs. RHE, FE C2+ was about 73.5%, and HER decreased from 40.6% to 16.8%. As shown in Figure 7A and B, Chen et al. developed a Cu polyamine hybrid catalyst by co-electroplating polyamines onto the surface of Cu electrodes (Cu-Pi, i =1-5)[81]. The amine functional group causes higher pH values and has higher intermediate stability, as well as the ability to adsorb more CO, thereby improving the selectivity of C2H4. The Cu-P1 catalyst reached the highest FE of C2H4 of 72%. As shown in Figure 7C and D, Lin et al. modified alkyl thiols with different alkyl chain lengths to regulate the interfacial wettability[55]. The FE C2+ of the Cu-12C catalyst was the highest, reaching 86.1%. At 400 mA cm-2, the catalyst achieved an FE C2+ of 80.3% [Figure 7E and F], which is one of the best performances at this current density. Increasing the interfacial hydrophobicity of the catalyst effectively prevented the absorption of H2O, facilitated CO2 transportation, increased *CO coverage, and thus enhanced C2+ selectivity[82].

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 7. (A) Schematic illustration of P1 and Cu co-electroplating on the GDL; (B) FE for all products on Cu-P1. [Reproduced with permission[81] .Copyright 2020, The Author(s), under exclusive license to Springer Nature Limited]; (C) HRTEM image of Cu-12C, revealing a 2-3 nm continuous and conformal alkanethiol layer; (D) Illustration shows Cu catalysts via hydrophobic treatment by alkanethiol; (E) FE and partial current densities of C2+ on Cu-12C under different current densities; (F) The variety of product selectivity with different interfacial wettability. [Reproduced with permission[55].Copyright 2023, The Author(s)]. FE: Faraday efficiency; GDL: Gas diffusion layer; HRTEM: High-resolution transmission electron microscope.

The above results indicate that enhancing the C-C coupling process in CO2RR is significant for the formation of C2+ products, and the efficiency depends on the coverage of *CO intermediates on the catalyst surface. For smaller-sized Cu-based NPs, adjusting the chemical environment of catalytically active sites can adjust the adsorption capacity for *CO, while for slightly larger catalyst particles, the adsorption capacity of *CO can be adjusted through nano-confined-space strategies. Enhancing the concentration of *CO can effectively promote the formation of C2+ products. Modifying the surface of Cu NPs with different materials will influence the selectivity of CO2RR products. When using hydrophobic materials, HER can be suppressed, thereby enhancing the CO2 reduction efficiency. When modified with hydrophilic salt materials, C1 products can be promoted, thereby reducing the generation of C2+. The synthesis of nanocavity structures using nanoconfinement effects can enrich and stabilize CO, greatly increasing the concentration of CO on Cu NPs and promoting the generation of C2+ products. Adjusting the Cu surface through different molecules is beneficial for reducing the energy barrier of C-C coupling and stabilizing intermediates, thereby improving CO2RR performance. However, the stability of Cu NPs is generally poor, and using appropriate carriers to anchor Cu NPs and developing new synthesis strategies are feasible methods to increase stability.

Cu-based single atoms

In Cu-based SACs, Cu atoms exist in the form of single atoms and can effectively participate in the reaction process. SACs have maximum atomic utilization efficiency and high catalytic activity[83], and can promote the catalytic reaction[84]. Compared with other Cu-based catalysts, Cu SACs have a higher utilization rate of active sites[85]. These SACs can maximize the utilization of Cu resources through reasonable design and preparation methods[86]. Due to their monodispersed catalytic sites, some Cu SACs cannot overcome the energy barrier of C-C coupling, inhibiting such reactions and thereby the formation of C2+ products and improving the selectivity of C1 products such as CH4, CO, CH3OH, and HCOOH[87,88]. The key to improving the performance of SACs is to regulate the interaction between catalytically active sites and reaction intermediates[89]. The CO2RR performance of Cu SACs can be improved by precisely adjusting the coordination environment and electronic structure of the central metal[90,91]. Most single-atom Cu is loaded onto C-based or NC materials through coordination with C or N. A strong correlation exists between catalytic activity and the coordination environment (such as CN and coordinating atoms) of Cu metal centers[92]. For example, Cu-C and Cu-N can promote the generation of CH4 and CO, respectively, and Cu-N2, Cu-N3 and Cu-N4 also form different catalytic effects. Metal and nitrogen-doped carbon (MNC) catalyst can activate and reduce CO2 to various products. The separation of electrons and geometric sites of a single metal atom will make the poor adsorption of H, thus eliminating the Tafel-type reaction and inhibiting the HER effectively[93]. In this section, we will summarize in detail the impact of the coordination environment of Cu SACs on their performance in CO2RR.

When Cu SACs coordinate with C in graphene to form Cu-C bonds, the activation sites of CO2 molecules are changed to promote the formation of *OCHO intermediates, thereby promoting the production of CH4[94-98]. In the process of CO2RR, the Cu-C bond formed by loading Cu atoms on carbon materials is usually conducive to the formation of *OCHO, which is converted to CH4. Shi et al. immobilized Cu single atoms on graphdiyne (GDY) and constructed Cu-C bonds for the first time[99], which not only stabilized the single Cu atoms but also enhanced the electronic control effect. They found that the Cu-C bond could induce the formation of *OCHO. Furthermore, the catalyst had a stronger binding ability to CO2 than H2O, thus suppressing HER. The maximum FE of CH4 was 81% at a stable electroreduction of ten h. Zhao et al. prepared Cu SACs with Cu-C2 coordination structure by accurately anchoring single Cu atoms onto hydrogenated graphdiyne (HGDY), using the special structure of 1,3,5-triethynylbenzene through the Sonokashira reaction Cu single atoms (Cu-SAs)/HGDY[98]. They confirmed that the low-coordination Cu-C2 provided the Cu single-atom center with more positive charges and promoted the formation of H• and improved the selectivity of CH4. At -1.1 Vvs. RHE, FE of CH4 was 44%, and the CH4 current density was 230.7 mA cm-2. Importantly, the catalyst achieved a turnover frequency (TOF) of 2,756 h-1.

When Cu coordinates with N in graphite carbon nitride (g-C3N4), it tends to produce CH4[100], and it can stabilize Cu single atoms[99]. Li et al. incorporated single-atom Cu into the nitrogen cavity of the host g-C3N4[100]. The catalyst exhibited an FE of CH4 of 49.04% and a maximum CH4/C2H4 ratio of 35.03. The Cu-N-C structure enhances the adsorption of *COOH and promotes the desorption of *CO, which is beneficial for the production of CO. Chen et al. anchored single-atom Cu onto a carbon nitride structure (Cu-N-C)[101], which enhanced the adsorption of *COOH and promoted the desorption of *CO. A high FE of CO of 98% with a local current density of 131.3 mA cm-2 was achieved at -0.67 V vs. RHE. Cai et al. prepared a carbon dot-supported SAC (Cu-CDs) with a unique CuN2O2 site [Figure 8A-E][87]. Since the generation of hydrocarbons requires to transfer more electrons and thereby a higher energy barrier, the product of CO2RR is usually CO. Therefore, the authors used partial carbonization to adjust the electronic structure of the Cu atom, thus reducing the total endothermic energy of *OCHO. Due to the introduction of oxygen ligands, FE of CH4 increased to 78% [Figure 8F and G]. Furthermore, they believed that pyridinic N was the anchor site for CO2 capture, due to the oxidation of Cu2+ in Cu-CD which had more pyridinic N than CD. Atomic-dispersed Cu-N3-C catalysts also exhibit excellent selectivity for CO. Chen et al. used biomass guanosine to directly synthesize a two-dimensional (2D) atom-dispersed Cu-N3-C catalyst (CuG-1000)[102]. The electrons were transferred from Cu coordinated with N to *COOH, thus regulating the adsorption and desorption of key intermediates and improving the selectivity of CO. At -0.65 V vs. RHE, the FE of CO of CuG-1000 reached 99% with a current density of 6.53 mA cm-2. Recently, as shown in Figure 8H-K, Purbia et al. synthesized Cu-SAC-N-doped carbon quantum dots (Cu SAC-N-CQDs) catalysts using CuCl2 as the Cu source and dopamine hydrochloride as the carrier through a hydrothermal reaction at 160 °C in the presence of citric acid for six h[103]. Figure 8H-K displays the structural schematic diagram and high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image of the Cu SAC N-CQD. The selectivity of CH3CH2OH is highly related to the coordination of Cu SACs, and changes with the variation of Cu loading. During the CO2 reduction progress, the in-situ generated Cu is the active site, which is beneficial for the production of CH3CH2OH. This catalyst had a unique Cu-N-C site and achieved an FE of 70% for CH3CH2OH at -0.2 V vs. RHE [Figure 8I], as well as a stability exceeding 50 h. Pan et al. designed C-supported Cu catalysts with atomic-dispersed N, OH-Cu3 sites Cu-N/interconnected mesoporous carbon fiber (IPCF)[104]. Atomically dispersed and N-coordinated Cu moieties were loaded onto a mesoporous N-doped carbon fiber carrier. The N, OH-Cu3 sites had moderate *CO adsorption affinity and low barrier for *CO hydrogenation. Cu-N/IPCF could enhance nano confinement to prevent CO from escaping mesochannels, which greatly increased the residence time of *CO, and continuously reduced to CH4 on Cu sites. At 300 mA cm-2, this catalyst showed a high FE of CH4 of 74.2%.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 8. (A) Scheme of the low-temperature calcining procedure for Cu-CD catalysts; (B) Large-field of view and (C) magnified view of TEM images; (D) Relatively large-field of view and e typical view of HAADF-STEM images of distributed single Cu atoms in carbon dots. Yellow circles in (E) indicate typical single Cu atoms; (F) FE and current density of Cu-CDs; (G) Stability test of Cu-CDs and CuPc at their highest ECR FE potentials. [Reproduced with permission[87]. Copyright 2021, The Author(s)]; (H) Structural schematic diagram of Cu SACs N CQDs electrocatalyst; (I) Faraday efficiency at different potential of Cu-SACs N CQDs; (J, K) HAADF-STEM image of all the Cu species on the carbon-dots Cu-N-CQDs (scale bar: 2nm). (Reproduced with permission[103]. Copyright 2024, Elsevier B.V.). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; FE: Faraday efficiency; SACs: Single-atom catalysts; TEM: Transmission electron microscope; CQDs: Carbon quantum dots.

The coordination of Cu single atoms with N to form Cu-Nx sites results in various CO2RR products during electroreduction. Dong et al. synthesized locally planar, symmetrically fractured planar-symmetry-broken CuN3 (PSB-CuN3) SACs[92]. The catalyst exhibited an FE of HCOOH of 94.3% at -0.73 V vs. RHE and could run stably in a flow cell for 100 h. The active centers of formate products were highly concentrated in the CuN3 region. In the CO2RR process, CuN4C4 sites tended to produce *COO-, while CuN3C3 and CuN2C2 tended to produce *CHO. Compared with highly symmetrical CuN4C4, the ΔG (0.23 eV) of the HCOOH on CuN3C3 was much lower than that of CO (0.68 eV), CH4/CH3OH (0.78 eV), and HER (1.01 eV). Xia et al. prepared ultra-high density Cu single atoms on thin-walled N-doped carbon nanotubes (TWN)[105], named TWN-Cu13.35-600-SACs, with a Cu content of 13.35 wt%. The researchers established that the adjacent Cu-N3 site was the active site. The FE of CH3CH2OH improved with the increase of the density of Cu-N3 sites. As shown in Figure 9A-C, in H-cell, the highest FE of CH3CH2OH was about 81.9%. In addition, the stability of the catalyst exceeded 25 h. Xu et al. synthesized a Cu SAC (Cu-N4-NG) based on N-doped graphene (NG) using a two-step pyrolysis method[106]. Cu atoms coordinated with four adjacent N atoms. They found that Cu-N4 portion was beneficial for the CO2 activation step. The CO2RR on Cu-N4-NG was less thermodynamically hindered, effectively limiting HER. Cu-N4-NG achieved an FE of CO of 80.6% at -1.0 V vs. RHE. Cheng et al. prepared Cu-N4-C/1100 containing Cu-N4[107]. They adjusted the coordination environment and electronic structure of Cu through pyrolysis. The edge-hosted Cu-N4 was the key active site for generation CO, which strongly interacted with *COOH and facilitated the desorption of *CO. Compared with Cu-N3-C/800, which contained Cu-N3 sites, Cu-N4-C/1100 exhibited excellent performance for CO production. At -0.9 V vs. RHE, the catalyst achieved an FE of CO of 98%. Karapinar et al. prepared Cu-N-C SACs using CuCl2 as a Cu source through simple pyrolysis. Cu atoms were coordinated by four N atoms to form CuN4 sites dispersed in an N-doped conductive carbon matrix[108]. During electrolysis, the separated sites instantaneously transformed into Cu NPs, which may be the active species. At -1.2 V vs. RHE, the FE of CH3CH2OH was 55%. As shown in Figure 9D and E, Zhao et al. encapsulated single-atom Cu on N-doped porous carbon (Cu-SA/NPC)[109]. They found that the reaction mainly occurred at the coordination sites between Cu and four pyrrole-N atoms (Cu-pyrrolic-N4). The coordination of Cu with four pyrrole-N atoms reduced the free energies for CO2 activation. Cu-SA/NPC could reduce CO2 to CH3COOH, CH3CH2OH, and CH3COCH3 products, and CH3COCH3 was the major product (FE = 36.7%). Generally speaking, Cu SACs with Cu-N4 sites (CN = 4) have the best catalytic effect on CO2RR and exhibit superior stability. Recently, Roy et al. used CuCl2 solution as the Cu source to form Cu-N2 sites by loading Cu single atoms onto a carbon nitride (CN) matrix through the metal ion exchange method [Figure 9F-H][110]. This catalyst had high-density (1.5 at%) single-atom sites. The cooperative Cu-Cu sites generated on the 9N pores in Cu-PTI synergistically enhanced the activity of the CO2 to CH4. As shown in Figure 9I and J, at -0.84 V vs. RHE, the FE of CH4 was 68%.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 9. (A) Preparation and characterization of the catalysts; (B) HAADF-STEM images of TWN-Cu13.35-600-SACs; (C) FE and the product distribution at different potentials. (Reproduced with permission[105]. Copyright 2023, American Chemical Society); (D) TEM images of Cu-SA/NPC. (E) FE of CO2RR products on Cu-SA/NPC. [Reproduced with permission[109]. Copyright 2020, The Author(s)]; (F) Synthetic scheme of Na-PHI, Li-PTI; (G) HRTEM images of Na-PHI; (H) HRTEM images of Li-PTI; (I) FE of CH4 comparison of Cu-PHI/PTI; (J) Partial current density of CH4 for the catalysts. (Reproduced with permission[110]. Copyright 2024, Wiley-VCH GmbH). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; FE: Faraday efficiency; SACs: Single-atom catalysts; TWN: Thin-walled N-doped carbon nanotubes; Cu-SA/NPC: Single-atom Cu on N-doped porous carbon; CO2RR: Carbon dioxide reduction reaction; TEM: Transmission electron microscope; PHI: Poly (heptazine imide); PTI: Poly (triazine imide); HRTEM: High-resolution transmission electron microscope.

Furthermore, Cu coordinated by heteroatoms can promote the formation of C2+ products. Wu et al. broke the coordination symmetry of Cu sites to form Cu-S2N1 sites in atomic precision in Cu6 clusters [Cu6(MBD)6, MBD = 2-mercaptobenzimidazole][111], which was conducive to the generation of *COOH instead of *OCHO, thereby inducing the formation of more valuable hydrocarbons. At -1.4 V vs. RHE, the catalyst exhibited an FE of 65.5% for hydrocarbons, where FE of CH4 and FE of C2H4 accounted for 42.5% and 23%, respectively. Recently, Lv et al. anchored Cu single atoms onto F, O, N-co-doped carbon composites (FONCDs) via precipitation reaction, and the prepared CuFONC catalyst had a stable CuN2O1 configuration and a high density of Cu SAs[112]. They established that the addition of F and O resulted in excellent stability of CuN2O1 during CO2RR. The addition of F adjusted the 3d orbitals charge of Cu atom in CuFONC. The Cu coordinated with N/O atoms, and the Cu-N/Cu-O ratio was 2:1. The Cu-N site in CuFONC was found to be an adsorption site for *CO. At -1.3 V vs. RHE, FE C2 was 80.5%.

The catalytic activity of Cu SACs is connected with the CN and degree of distortion of the Cu metal. Cu-N and Cu-C are effective active sites for CO2RR. Cu-C and Cu-N sites generally generate hydrocarbons and oxygen-containing compounds, respectively. Researchers have found that a single Cu site usually produces C1 products, and adjacent Cu can promote C-C coupling. It is difficult to precisely control the active sites, as they usually generate C1 products, making it challenging to obtain C2+ products. Therefore, the key is to design the active site reasonably. Cu SACs with high CO2RR performance can be designed by regulating the CN of Cu and N (Cu-N4 being optimal), adjusting the loading density of Cu single atoms, and doping other heteroatoms to coordinate with Cu to saturate the Cu coordination sites.

CU-BASED CATALYST ADDITIVES

Cu has moderate *CO adsorption energy, which can form chemical bonds between adjacent CO molecules[113]. However, it is limited in the initial CO2RR step (CO2-CO), resulting in low CO surface coverage. A second metal component can be introduced to take advantage of its higher CO2-CO reduction performance[114,115]. Au, Ag, Zn, and Ni have strong catalytic effects on the production of CO from CO2RR. Additives can also adjust the acidity, alkalinity, and electronic structure of Cu-based electrocatalysts[116]. Regarding catalyst composition, the construction of two-component or multi-component catalysts has been shown to effectively improve CO2RR efficiency[117,118] and simultaneously inhibit HER[119,120]. Combining Cu with other metals[113,121] usually exhibits in the characteristics of high selectivity, stable performance, and low overpotential[122,123].

CuAu

Ouyang et al. found that on Cu alloy catalysts such as CuAu and CuAg, the surface coupling hydrogenation activity was enhanced and indirectly inhibited solvent hydrogenation of intermediates, resulting in higher CH3CH2OH selectivity than Cu (100)[124]. Morales-Guio et al. deposited Au NPs by physical vapor deposition on top of polycrystalline Cu foil (Au/Cu)[125]. Au/Cu efficiently reduced CO2 to CH3CH2OH. They observed that CO2 was reduced to CO on Au NPs, which was then enriched and further reduced to alcohols on nearby Cu. Shen et al. synthesized AuCu alloy embedded in Cu submicron conical arrays (AuCu/Cu SCA)[114]. Cu atoms were replaced by larger Au atoms, causing lattice expansion and forming AuCu alloys. Altering the content of Au affected CO2RR activity and the selectivity of C2H5OH/C2H4 products. The CO generated at Au sites coupled with *CH2 intermediates at Cu sites to form *COCH2, which further generated CH3CH2OH. In addition, the catalyst could maintain electrocatalytic activity for 24 h at high current densities. Wei et al. prepared an Au-doped Cu nanowire (Cu Au NWs)[126]. They modified a small amount of Au NPs on the surface of Cu NWs, using the homonuclear method. The addition of Au NPs resulted in a rougher surface, exposing more active sites. In addition, the *CO intermediates generated on Au NPs aggregate on Cu NWs, promoting the adsorption of *CO. The interface electrons transferred from Cu to Au induced electron-deficient Cu, which facilitates the adsorption of *CO. At -1.25 V vs. RHE, the FE C2+ increased from 39.7% of Cu NWs without Au to 65.3%. Zheng et al. successfully synthesized an Au-Cu Janus nanostructure catalyst (Au-Cu Janus NSs) using a seed growth strategy[127]. Cu atoms were deposited on one side of concave cubic Au seeds. The Au-Cu domain boundaries in the spatial separated Janus nanostructures facilitated synergistic catalysis. At -0.75 V vs. RHE, the FE C2+ was 67%.

CuAg

Ma et al. prepared Ag65-Cu35 Janus nanostructure (JNS-100) catalysts with (100) facets [Figure 10A-C][128]. Compared to Cu, Ag65-Cu35 JNS-100 had higher electron abundance. This catalyst exhibited the highest selectivity for C2H4 and C2+ products with FE values of 54% and 72%, respectively [Figure 10D-F]. As mentioned earlier, Cu (100) had a lower energy barrier for C-C coupling[49]. As shown in Figure 10G, Ag65-Cu35 JNS-100 exposed Cu (100), Janus nanostructure induced tandem catalysis and electron transfer between Ag and Cu. Wei et al. improved the FE C2+ in CO2RR by concatenating nano electrocatalysts[129]. As shown in Figure 10H, they designed a three-dimensional (3D) catalyst electrode with Ag NPs, which were deposited at the bottom of a Cu nanoneedle array (Cu needle-Ag) to generate the intermediate product CO. The dense Cu needles would prevent CO from flowing out of the Ag NPs, causing further dimerization of CO on the Cu nanoneedles to form C2+ products. Through this nanostructure design, FE C2+ was 70% in a flow cell. Choi et al. synthesized a tight atomic Cu-Ag interface on the Cu NW surface[130]. The tight atomic Cu-Ag interface promoted the synergistic effect of CO-Ag* and H-Cu*, greatly improving the selectivity of CH4. The maximum FE of CH4 at -1.17 V vs. RHE was 72%. Hoang et al. electrodeposited 3,5-diamino-1,2,4-triazole inhibitors on thin CuAg alloy films[123]. The CO2RR performance of the alloy wire with an Ag content of 6% was the best, achieving an FE of nearly 60% and 25% for C2H4 and C2H5OH, respectively, at low potentials of -0.7 V vs. RHE. CuAg wire (6%) exhibits Ag-Cu bonds. Also, the Ag-Ag distance in the CuAg film is slightly reduced, indicating that Ag atoms were at least partially alloyed with smaller Cu atoms. As shown in Figure 10I, Li et al. synthesized Ag,S-Cu2O/Cu catalysts through an in-situ dual doping strategy[131]. This catalyst had a typical 3D porous architecture which resulted in more active sites. The Cu atoms near heteroatoms in Cu2O/Cu served as efficient active centers for CH3OH production. S- could regulate the electronic structure and morphology, while Ag+ inhibited HER. FE of CH3OH reached 67.4%. Du et al. cascaded Ag NPs and AgCu single-atom alloy (SAA) (AgCu SANP)[132]. The Ag NPs selectively reduced CO2 to CO, followed by the formation of multiple carbon products on AgCu SAA. Owing to the asymmetric binding of Cu atom to the adjacent Ag atom, the incorporation of the single Ag atom enhanced the adsorption energy of CO on the Cu sites. At -0.65 V vs. RHE, FE C2+ reached 94% ± 4% under a working current density of about 720 mA cm-2. Qi et al. synthesized CuAg alloy catalysts (CO2-10-Cu94Ag6) by the co-electrodeposition method[133]. Under CO2-supersaturated conditions, the co-deposition of Ag and Cu forms a CuAg alloy, with the Cu (100) facets being preferentially exposed. The dispersed Ag atoms on Cu strengthened C-O bond of C3 chain. The FE 2-propanol in the H-cell with 1.0 mol L-1 CsHCO3 was 56.7%. They observed that C3 products can only be detected when the electrolyte cation is Cs+ and CO2 is supersaturated. This was attributed to the larger radius of Cs+ that can interact with two *CO. Recently, Li et al. synthesized Cu3N-Ag NCs by doping Ag into Cu3N nanocubes (NCs)[134]. The Ag sites generated CO, while the Cu sites promoted C-C coupling to *COCHO, which enhanced the production of C2H4. As shown in Figure 10J, the cascade Ag-Cu dual sites in Cu3N-Ag NCs could promote C-C coupling and generate C2H4.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 10. (A) HAADF-STEM image of Ag/Cu interface in Ag65-Cu35 JNS-100; (B) HAADF-STEM images of Ag65-Cu35 JNS-100; (C) EDS elemental mappings of Ag65-Cu35 JNS-100; (D) FE of major CO2-reduction products obtained on Ag65-Cu35 JNS-100; (E) Comparison of C2H4 FE between Ag65-Cu35 JNS-100 and Cu NCs at different potentials; (F) Comparison of C2+/C1 product ratios between Ag65-Cu35 JNS-100, Ag + Cu mixture and Cu NCs; (G) Schematic illustration of a plausible CO2RR mechanism on Ag65-Cu35 JNS-100. (Reproduced with permission[128]. Copyright 2022, Wiley-VCH GmbH); (H) Schematic illustration of the synthesis of Cu needle-Ag catalyst. (Reproduced with permission[129]. Copyright 2023, Advanced Functional Materials published by Wiley-VCH GmbH); (I) Schematic diagram of the in situ dual-doping process for preparing the x, y-Cu2O/Cu catalysts. [Reproduced with permission[131]. Copyright 2022, The Author(s)]; (J) Introduction of Ag into Cu3N with weaker binding strength with reaction intermediates of CO2RR could increase local CO coverage and facilitate *CO protonation to *CHO. (Reproduced with permission[134] . Copyright 2024, American Chemical Society). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; JNS: Janus nanostructure; FE: Faraday efficiency; NCs: Nanocrystals; CO2RR: Carbon dioxide reduction reaction; EDS: Energy dispersive spectrometer.

CuZn

In 2017, Kattel et al. established that the ZnO/Cu (111) interface sites were the real active sites of CuZn catalysts that could reduce CO2 to CH3OH[135]. The synergistic effect of Cu and ZnO facilitated the synthesis of CH3OH. Zhu et al. prepared Cu/ZnO-CeO2 catalysts by the flame spray pyrolysis method[136], and the catalysts exhibited better CH3OH selectivity (70%) than Cu/ZnO and Cu/CeO2. In 2022, Amann et al. found that the hydrogenation of CO2 preferentially occurred on ZnO[137]. Wan et al. found that phase-separated bimetallic Cu-Zn catalysts had a lower energy barrier to form the *COOH than the core-shell one[138]. The FE of CO was 94% and stability exceeded 15 h. Zhen et al. found that the synergistic effect between Cu and Zn can result in effective CO adsorption. Both balanced Cu-Zn sites and Zn-rich Cu-Zn sites promoted C-C coupling[139]. Recently, as shown in Figure 11A-D, Zhang et al. developed CuZn alloy/CuZn aluminate oxide (Al2O4) composite electrocatalyst (Cu9Zn1/Cu0.8Zn0.2Al2O4)[140]. The Al2O4 triggered a robust electron interaction among Cu, Zn, and Al, leading to the generation of numerous highly reactive interfaces characterized by exceptional activity. Supported by interface effects, the optimized catalyst achieved an FE C2+ of 88.5% and a current density of up to 400 mA cm-2 [Figure 11E-J].

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 11. (A) Schematic illustration for the preparation of the CuZn/CuZnAl2O4 catalyst; (B) TEM image; (C) HR-TEM image; (D) HAADF-STEM image; (E) LSV curves toward CO2RR on CuZn/CuZnAl2O4; Product distributions and corresponding FE produced by CuZn/CuZnAl2O4 (F); Cu/CuAl2O4 (G); and CuZn (H); (I) C2+ partial current density; (J) Stability test of CuZn/CuZnAl2O4. (Reproduced with permission[140]. Copyright 2023, ELSEVIER B.V. and Science Press). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; CO2RR: Carbon dioxide reduction reaction; TEM: Transmission electron microscope; HR-TEM: High-resolution transmission electron microscope; LSV: Linear sweep voltammetry.

Others

Liu et al. loaded Ni SAC onto Cu catalyst using the electrostatic self-assembly method[141]. This catalyst promoted the dimerization of CO, reaching an FE of C2H4 of approximately 62% at -1.4 V vs. RHE. At 500 mA cm-2, the catalyst could sustain stability for 14 h. Recently, Song et al. prepared a double-layer hollow spherical nanosphere catalyst (Ni-SA@Cu-NP)[142]. The inner layer of Ni-SA@Cu-NP was composed of dispersed Ni atoms, and the outer carbon layer consisted of Cu NPs. CO was generated on the Ni site in the inner layer, accumulated in the cavity, and then overflew to the outer Cu site for CO dimerization, achieving FE C2+ of 74.4%. The addition of In often reduced CO2 to HCOOH. Wei et al. prepared a series of bimetallic InxCuy NP electrocatalysts[143]. They adjusted the growth direction of the crystal face by changing the molar ratio of In/Cu. Among the catalysts, In1.5Cu0.5 NPs had an FE of HCOOH of 90% at -1.2 V vs. RHE. Wang et al. prepared an In2O3/Cu catalyst with 3D succulent plant morphology[144]. The surface tensile strain and interfacial electronic interaction reduced the energy barrier for the formation of *HCOO. At -1.4 V vs. RHE, FE of HCOOH was 87.5%. Li et al. prepared a metal Cu/Pd-1% catalyst[145]. The Pd element was uniformly distributed on the Cu NPs. The doping of Pd revealed the d-band center of Cu toward the Fermi level. The higher d-band center location favors the adsorption of intermediates. This catalyst enhanced the affinity for *CO, exhibiting an FE C2+ of 66.2%. Zheng et al. introduced a single-atom Pb alloy Cu catalyst (Pb1Cu) [Figure 12A and B][146], which could completely convert CO2 to HCOOH with an FE of nearly 96%. The activation of Cu sites on Pb1Cu catalyst directed CO2RR towards the HCOO* instead of the COOH*. As shown in Figure 12C, the highest FE of HCOOH was 95.7% at -0.72 V vs. RHE, and an FE exceeding 80% could still be maintained even when the partial current density was up to -1200 mA cm-2. Yang et al. prepared rod-like bimetallic CuBi75 catalysts[122]. Bi, C, and O were uniformly dispersed on CuBi75 catalysts. At -0.77 V vs. RHE, FE of HCOOH was 100%. As shown in Figure 12D, Cao et al. introduced single Bi atoms on Cu NPs to form BiCu-SAA[147]. The single-atom Bi on the Cu (111) surface resulted in significant charge redistribution. Local electrons can be transferred to the antibonding orbitals of CO2 molecules, thereby activating and further reducing CO2. As shown in Figure 12E and F, the FE C2+ of the BiCu-SAA catalyst was 73.4 %, and the catalyst remained stable at the current density of 400 mA cm-2.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 12. (A) Schematic illustration of CO2 conversion into HCOOH over a Pb1Cu SAA; (B) HAADF-STEM image and (C) FEs of all reduction products and j-V curves of Pb1Cu catalyst. [Reproduced with permission[146]. Copyright 2021, The Author(s), under exclusive license to Springer Nature Limited]; (D) Schematic diagram of BiCu-SAA (E) The FEs of CO2RR products on the catalysts with different Bi content at an applied potential of -1.10 V; (F) FE of C2+ products on the BiCu-SAA and control Cu-Nano catalysts at different applied potentials. (Reproduced with permission[147]. Copyright 2023, Wiley-VCH GmbH). HAADF-STEM: High-angle annular dark-field scanning transmission electron microscopy; CO2RR: Carbon dioxide reduction reaction; FE: Faraday efficiency; SAA: Single-atom alloy.

Hu et al. synthesized La(OH)3/Cu by modifying the Cu catalyst with La(OH)3[148]. The modification of La(OH)3 caused Cu to be in an electron-deficient state, which favored *CO adsorption and *CO-*COH coupling, thus increasing C2+ selectivity. The main component was C2H4. This catalyst achieved an FE C2+ of 71.2%, which was 1.2 times higher than that of pure Cu, and exhibited a current density of up to 1,000 mA cm-2. Li et al. doped the p-block metal atom Ga into Cu and prepared a CuGa catalyst[149]. At -1.07 V vs. RHE, the CuGa catalyst exhibited an excellent FE C2+ of 81.5% at a current density of 0.9 A cm-2. When the current density reached 1100 mA cm-2, the catalyst still maintained a high FE C2+ of 76.9%. It was attributed to the p-d hybrid interaction between Ga and Cu, which redistributed the electronic structure and improved the bonding strength of intermediates. This finesse can also be expanded to other p-block metals, such as Al and Ge. Xie et al. synthesized a bimetallic Mg-Cu catalyst[150]. The modification with Mg not only enhanced the activation of CO2 but also stabilized the Cu sites. Compared to Cu catalysts, Mg-Cu catalysts had abundant Cu+ sites and stronger adsorption capacity for CO2 and CO. The FE C2+ of the catalyst reached 80% (mainly C2H4), and the current density reached 1 A cm-2.

Generally speaking, introducing other metals as additives can effectively improve CO2 reduction efficiency. The metals used for doping and the levels of doping significantly influence both the products and the efficiency. For instance, introducing hydrophilic metals may result in the preferential formation of oxygen-containing compounds. The interaction between two metals usually improves CO2 reduction efficiency and leads to high selectivity for C2+ products. When Cu forms an alloy with other metals, the diversity in electronegativity can lead to charge transfer, resulting in a change in the d-band center of active sites. This electronic effect can regulate the adsorption of key intermediates. The addition of various metals often leads to different products. The Cu-Au catalysts are often beneficial to the production of CH3CH2OH. The Cu-Ag catalysts usually facilitate the generation of oxygen-containing C2+ products. The Cu-Zn catalysts are favorable to the production of C2H4. The addition of In, Bi and Sn usually leads to HCOOH. At the same time, doping with other metals can also achieve high current densities, some of which can reach industrial-grade current densities, and there is hope for commercialization. However, most of the doped metals are precious metals, and their use is not very economical. Furthermore, strong interactions between alloy components can lead to better catalytic performance of alloy catalysts than single-component NPs[151,152].

CU-BASED CATALYST SUPPORTS

When preparing Cu-based catalysts, it is necessary to load Cu NPs or Cu SAs on conductive materials to enhance catalytic activity. Choosing a suitable carrier can stabilize and enhance the electrocatalytic effect of these catalysts[9]. Common carriers include C-based materials (such as CB[153], carbon nanotubes, and graphene), N-doped carbon (NC) materials, and oxide carriers[154,155]. There are also some other types of carriers, such as MXene and metal-organic frameworks (MOFs).

C-based materials

C-based materials are commonly used to disperse SACs due to their advantages such as low cost, good stability, and good conductivity[156]. The properties of the substrate can change the electronic state of the active sites, thereby altering the reaction pathway and catalytic mechanism. To prepare commercial carbon-supported Cu SACs, Xu et al. mixed bulk Cu with molten Li and subsequently added CB as the support (Vulcan XC-72), followed by leaching LiOH by thoroughly mixing with water, in which process Cu atoms transferred to the carbon surface[153]. On this support, Cu single atoms can be highly dispersed. Cu was still atomically dispersed after prolonged electrocatalysis. At -0.7 V vs. RHE, FE of CH3CH2OH achieved 91%. The stability exceeded 16 h. Recently, Pan et al. developed a dual continuous mesoporous carbon carrier (IPCF) to support the synthesis of Cu-N/IPCF catalysts from Cu single atoms[104]. They designed carbon supported Cu catalysts by regulating the atomic scale structure of Cu active sites and designing the mesoscale structure of carbon supports. The catalyst was derived from block copolymers and had interconnected mesopores. Its unique long-range channel provided a microenvironment lacking H2O, prolonging the transport pathway of CO intermediates. And this catalyst had N, OH-Cu3 sites, which effectively inhibited HER and achieved an FE of CH4 of 74.2% at 300 mA cm-2. As shown in Figure 13A-C, Yang et al. prepared a through-hole carbon nanofiber catalyst with a large and uniformly distributed Cu single-atom modification CuSAs/through-hole carbon nanofibers (TCNFs)[157]. Due to the self-supporting and through-hole structure of the catalyst, a large number of Cu SAs were fully exposed on the surface, which was conducive to capturing CO2. At -0.9 V vs. RHE, FE of CH3OH was 44%, and the CH3OH partial current density was -93 mA cm-2.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 13. (A) Pyridine Cu-N4 structure; (B) LSV curves and (C) FE of all products at CuSAs/TCNFs. (Reproduced with permission[157]. Copyright 2019, American Chemical Society); (D) TEM image and (E) corresponding enlargement of the sites 1-4 and intensity profiles of Cu-1/hNCNC; (F) FE, current density (j), and product distribution for Cu-1/hNCNC at different polarization potentials; (G) Ethanol FE and j as a function of time for Cu-1/hNCNC during a chronoamperometric test at -0.30 V. (Reproduced with permission[161]. Copyright 2024, American Chemical Society). FE: Faraday efficiency; LSV: Linear sweep voltammetry; CuSAs/TCNFs: Cu Single Atoms/through-hole carbon nanofibers; TEM: Transmission electron microscope; hNCNCs: NC nanocages.

NC materials

NC materials are ideal carriers due to their tunable pore structure, intense metal-heteroatom interactions, and stability[104,158]. Compared with C-based materials, NC materials contain N species, a higher N content, and a uniform N arrangement, which can provide abundant and more accurate coordination sites for single atoms[159]. Feng et al. used B- and N-doped graphene materials (N-doped GDY and B-doped GDY) as supports to anchor individual Cu atoms[160]. They found that Cu@doped GDY can spontaneously capture CO2. Cu atoms had strong interactions with adjacent C, B, or N atoms. The electronegativity of coordinated elements is a crucial factor in improving catalytic performance. They discovered that the catalytic performance of Cu@N-doped GDY was better. Recently, as shown in Figure 13D-G, Xu et al. prepared Cu-1/hNCNC by vapor deposition on multi-stage NC nanocages (hNCNCs) using commercial Cu2O powder as a Cu source[161]. This catalyst had CuOCu-N4-oxygen-bridged binuclear Cu sites on hNCNCs. Cu-1/hNCNC exhibited the most advanced low overpotential for C2H5OH at 0.19 V vs. RHE, with FE of CH3CH2OH reaching 56.3% at -0.30 V vs. RHE, and the Cu active site was very steady during the CO2RR process.

Cu NPs usually employ C-based materials and oxides as supports, and the interaction between appropriate supports and Cu NPs can improve the selectivity of the product[156]. For example, the g-C3N4 supports can not only elevate the d-orbital of Cu to the Fermi level, but also serve as an additional active center for CO2RR, enhancing the selectivity of CH4 and C2H4[162]. As shown in Figure 14A and B, Li et al. synthesized CuNCN by the precipitation method, using CuCl as a Cu source and adding NH3•H2O and cyanamide (H2NCN), and then prepared Cu-based/CxNy catalysts by pyrolysis of CuNCN[163]. The state of Cu species on the g-C3N4 carrier was regulated by the pyrolysis temperature. Among the prepared catalysts, CuNCN-300 obtained by pyrolysis of CuNCN at 300 °C had the highest FE of C2H4 of 48.5% and a current density of 500 mA cm-2. The good dispersibility of Cu3N in CuNCN-300 promoted the selectivity of C2H4, while the tris-s-triazine structure (g-C3N4 fragment) in CxNy enhanced the reaction of *CO - *CHO on the Cu surface. Li et al. prepared a p-NG-Cu-7 catalyst by loading 7-nm Cu NPs on pyridine-rich N-graphene (p-NG)[164]. As shown in Figure 14C, due to the rich Lewis base sites of p-NG, the formation of *COOH was promoted, which was then enriched on Cu NPs and converted into CO and CHxO with subsequent C-C coupling. Compared with the catalytic performance of GO-Cu-7 catalyst deposited on conventional graphene (FE of C2H4 = 7.5%), the p-NG-Cu-7 catalyst had an FE of C2H4 of 19%, and the selectivity of C2H4 in hydrocarbons reached 79%. The p-NG-Cu-7 catalyst could also produce C2H5OH with 63% FE.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 14. (A) The schematic illustration of synthesis of CuNCN and preparation of Cu-based/CxNy by CuNCN pyrolysis; (B) HRTEM of CuNCN-300. (Reproduced with permission[163]. Copyright 2022, Elsevier B.V. All rights reserved); (C) Product selectivity of the hydrocarbons generated from the p-NG-Cu-7 catalyzed reduction at -0.9 V. (Reproduced with permission[164]. Copyright 2016, Elsevier Ltd). NCN: Elements N and C; HRTEM: High-resolution transmission electron microscope.

Oxides

The synergistic effect between metal oxide carriers and Cu can modulate the electronic structure of the catalyst. Loading Cu on oxide (ZnO/Al2O3) can promote the reduction of CO2 to CH3OH[165]. The Lewis acid sites in metal oxides (such as Al2O3 and Cr2O3) have been proven to activate CO2 molecules and promote CO2 methanation[166-169]. As shown in Figure 15 A, Chen et al. anchored Cu SAs to ultra-thin porous Al2O3 rich in Lewis acid centers (Cu/Al2O3 SAC)[170]. When Cu single atoms were supported by ultra-thin porous Al2O3, the Cu/Al2O3 SAC achieved an FE of C2H4 of 62%. As shown in Figure 15B and C, Cu atoms on Al2O3 were in a higher oxidation state (electrons transferred from Cu atoms to the carrier). In addition, the Cu atoms on Al2O3 exhibited higher d-band centers, indicating an improved electron transfer ability. The electron-accepting properties of Al2O3 were beneficial for stabilizing methanation intermediates and reducing the energy barrier. TiO2 could stabilize Cu NPs and provide more active sites. As shown in Figure 15D and E, Yuan et al. prepared CuO/TiO2 catalysts by the hydrothermal method using CuO and TiO2 and then reduced them in situ to Cu/TiO2 catalysts[171]. TiO2, as a semiconductor material, can serve as an oxidation-reduction electron carrier and assist in CO2 adsorption[172-174]. It can stabilize the CO2RR intermediates and reduce overpotential. The CuO/TiO2 significantly increased the adsorption capacity of CO2. Firstly, a large amount of CO2 was adsorbed on TiO2. Then, the adsorbed CO2 obtained an electron from the Cu/TiO2 and was converted to CO2-, which could be dimerized to *C2O2-. As shown in Figure 15F and G, the Cu/TiO2 catalyst effectively reduced CO2 to multiple oxygen-containing carbon compounds such as CH3CH2OH, CH3COCH3, and CH3CH2CH2OH. The maximum total FE was 47.4%.

Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms

Figure 15. (A) Schematic diagram of CO2RR on Cu/Al2O3 SAC; (B) Electronic structure and (C) projected densities of states (pDOS) of d-orbitals with an aligned Fermi level of Cu/Al2O3 SAC and Cu/Cr2O3 SAC. Color code: Cu, brick red; Al, purple; Cr, gray; O, red. (Reproduced with permission[170]. Copyright 2021, American Chemical Society); (D) TEM images of CuO/TiO2-5 catalyst in low magnification; (E) Schematic diagram of CO2RR on CuO/TiO2-5 catalyst; (F) FEs for different products over various CuO/TiO2 catalysts at -0.85 V vs. RHE in CO2-saturated 0.5 M KHCO3 aqueous solution; (G) FEs for different products over CuO/TiO2-5 catalyst at various potentials. (Reproduced with permission[171]. Copyright 2018, MDPI, Basel, Switzerland). CO2RR: Carbon dioxide reduction reaction; SAC: Single-atom catalysts; FE: Faraday efficiency; RHE: Reversible hydrogen electrodes; TEM: Transmission electron microscope.

Others

Abdinejad et al. synthesized Cu-Pd/MXene catalysts using MXene-based (Ti3C2Tx) materials as carriers[175]. They paired 2D MXene with bimetallic Cu-Pd. Compared with Cu-Pd, the Cu-Pd/MXene had a larger active surface area and electron transfer rate. This stemmed from MXene improving electron transfer and having a larger electrochemically active surface area (EASA). Due to the unique multi-layer composition of MXene, the catalytic surface area and conductivity increased. At -0.5 V vs. RHE, FE of HCOOH reached 93%, and overall battery energy efficiency (EE) reached 47%. MOFs are considered as the ideal catalyst supports. The Cu-N coordination bond formed by Cu(II) and N-heterocyclic ligands has moderate strength, located between Cu-O coordination and Cu porphyrin bond[176]. Thus, Cu-N coordination is more stable. Therefore, Chen et al. designed a MOF (2Bn-Cu@UiO-67) of encapsulated N-heterocyclic carbene (NHC) ligand linked to Cu SAC[177], which exhibited an FE of CH4 of 81%. Due to the interaction between Cu and NHCs, the electron occupancy rate on the d orbital was much lower than that of Cu foil. The σ donation from NHC enriched the electron density of Cu single atoms, promoting the adsorption of CHO*. The high porosity promoted the spread of CO2 molecules towards 2Bn-Cu, remarkably improving the catalytic efficiency.

The above results indicate that the appropriate carrier can improve the performance and selectivity of the catalyst, and some supports can enhance catalyst stability.

The following Table 2 lists the catalytic performance data of different types of Cu catalysts.

Table 2

A summary of reported performance data for Cu-based CO2RR electrocatalysts

CategoryCatalystCO2RR conditionsPotential/
V vs. RHE
FEj/ mA cm-2Ref.
Cu NPs7 nm n-Cu/C0.1 M NaHCO3, H-cell-1.35CH4 = 76%9.5[60]
20-nm Cu2O NP/C0.1 M KHCO3, H-cell-1.1
C2+ = 74%, C2H4 = 57.3%27.5[61]
20 nm Cu NPs5 M KOH, flow-cell-0.73C2+ = 70%800[63]
Cu nanoneedles3 M KCl, pH = 1, flow-cell-2.3C2+ = 90.69% ± 2.15%1400[64]
o-Cu2O with high-index facets1 M KCl, H-cell-1.1C2+ = 48.3%17.7[65]
F-Cu2O with exposed (322) facets0.1 M KHCO3, H-cell-1.2C2H4 = 74.1%, for 12 h[66]
Cu2O/Ppy with high refractive index (311) and (211) facets0.5 M KHCO3, H-cell-0.85CH3OH = 93% ± 1.2%0.223[67]
Cu2O (CO) with abundant Cu (100) crystal planes0.1 M KHCO3, flow-cell-1C2+ = 77.4%, C2H4 = 56.6%500[68]
P-Cu2O-2401 M KOH, flow-cell-2.5C2+ = 75.3% ± 3.1%1000[69]
P-Cu0.1 M KHCO3, H-cell-1.3C2+ = 57.22%, C2H4 = 30%40[77]
Cu-HoMSs0.5 M KHCO3, flow-cell-0.88C2+ = 77.0% ± 0.3%, (mainly C2H4 and CH3CH2OH)513.7 ± 0.7[79]
Multihollow Cu2O2 M KOH, flow-cell-0.61C2+ = 75.2% ± 2.7%342[80]
p-CuO-(12.5 nm)3 M KOH, flow-cell-0.87C2H5OH = 44.1 ± 1%, C2+ = 90.6% ± 3.4%501 ± 15[78]
fluorinated polymer-functionalized Cu-poly-11 M KHCO3, flow-cell-3.98C2+ = 71.08%500[58]
AN-Cu(OH)@Nafion1 M KOH, flow-cell-0.76C2H4 = 44%300[52]
Cu@Nafion-40.1 M KHCO3, H-cell-1.2C2+ = 73.5%13[53]
Cu-P1 with polyamines10 M KOH, flow-cell-0.47C2H4 = 87% ± 3%[81]
Cu-12C1 M KOH, flow-cell-1.2C2+ = 80.3%321[55]
t-Cu2O0.5 M KHCO3, H-cell-1.1C2H4 = 59%22[73]
Cu (100)/Cu (111)1 M KHCO3, flow-cell-0.6C2+ = 74.9 ± 1.7%, for 50 h300[74]
TA-Cu1 M KOH, flow-cell-0.7C2H4 = 63.6%497.2[76]
Cu SAs-GDY0.1 M KHCO3, flow-cell-1.2CH4 = 81%, for 10 h243[99]
Cu SACsCu-SAs/HGDY1 M KOH, flow-cell-1.1CH4 = 44%230.7[98]
Cu-CN0.1 M KHCO3, H-cell-1.2CH4 = 49.04%7.97[100]
Cu-N-C0.5 M KHCO3, H-cell-0.67CO = 98%4.5[101]
Cu-CDs0.5 M KHCO3, H-cell-1.44CH4 = 78%40[87]
CuG-10001 M KOH, H-cell-0.65CO = 99%6.53[102]
Cu SACs N-CQDs0.1 M KHCO3, H-cell-0.2CH3CH2OH = 70%, for 50 h6.53[103]
Cu-N/IPCFAlkaline electrolyte, flow-cell-1.21CH4 = 74.2%300[104]
PSB-CuN3 SACs0.5 M KHCO3, flow-cell-0.73HCOOH = 94.3%, for 10 h94.4[92]
TWN-Cu-600-SACs0.5 M CsHCO3, H-cell-1.1CH3CH2OH = 81.9%, for 25 h35.6[105]
Cu-N4-NG0.1 M KHCO3-1FE of CO = 80.6%[106]
Cu-N4-C/11000.1 M KHCO3, H-cell-0.9CO = 98%, for 40 h3.3[107]
Cu-SA/NPC0.1 M KHCO3-0.36Acetone = 36.7%7[109]
Cu-NC1 M KOH, flow-cell-0.84CH4 = 68%, for 12 h348[110]
Cu/C-Al2O3 SAC1 M KOH, flow-cell-1.2CH4 = 62%153[170]
CuSAs/TCNFs0.1 M KHCO3, flow-cell-0.9CH3OH = 44%93[157]
Cu6(MBD)61 M KOH, flow-cell-1.4CH4 = 42.5%, C2H4 = 23%183.4[111]
CuFONC0.1 M KHCO3, H-cell-1.3C2+ = 80.5%60[112]
AuCu/Cu SCA0.5 M KHCO3, H-cell-0.8C2H4 = 16%, C2H5OH = 29%, for 24 h4.9[114]
AdditiveCu Au NWs0.1 M KHCO3 and 0.1 M KCl, H-cell-1.25C2+ = 65.3%, for 10 h12.1[126]
Au-Cu Janus NSs3 M KOH, flow-cell-0.75C2+ = 67%290[127]
Cu needle-Ag0.1 M KHCO3, flow-cell-1C2+ = 70%350[129]
Cu9Ag1NWs0.1 M KHCO3-1.17CH4 = 72%[130]
CuAg-DAT0.1 M KHCO3, flow-cell-0.7C2H4 = 60%, C2H5OH = 25%300[123]
Ag,S-Cu2O/CuBMImBF4/H2O = 1:3, flow-cell-1.18CH3OH = 67.4%122.7[131]
Ag65-Cu35 JNS-1000.1 M KHCO3, H-cell-1.4C2H4 = 54%, C2+ = 72%15.14[128]
AgCu SANP1 M KOH, flow-cell-0.65C2+ = 94%720[132]
Cu94Ag61 M CsHCO3, H-cell-0.732-propanol = 56.7%, for 24 h59.3[133]
Cu4Zn0.1 M KHCO3, H-cell-1.05CH3CH2OH = 29.1%8.2[11]
Cu-Zn0.1 M KHCO3-1CO = 94%, for 15 h[138]
Cu9Zn1/Cu0.8Zn0.2Al2O42 M KOH, H-cell-1.15C2+ = 88.5%25[140]
Ni SAC + Cu-R1 M KHCO3, flow-cell-1.05 ~ -1.4C2H4 = 62%, for 14 h370[141]
Ni-SA@Cu-NP1 M KOH, flow-cell-1.2C2+ = 74.4%, for 15 h337.4[142]
In1.5Cu0.5NPs0.1 M KHCO3, H-cell-1.2HCOOH = 90%3.59[143]
In2O3/Cu0.5 M KHCO3-1.4HCOOH = 87.5%[144]
Cu-In2O30.1 M KHCO3, H-cell-0.7CO = 95%3.7[175]
Cu/Pd-1%1 M KOH, flow-cell-1.8C2+ = 66.2%463.2[145]
Pb1Cu0.5 M KHCO3, flow-cell-0.72HCOOH = 95.7%500[146]
CuBi750.5 M KHCO3-0.77HCOOH = 100%[122]
BiCu-SAA0.1 M KHCO3, flow-cell-1.1C2+ = 74.3%, for 11 h400[147]
La(OH)3/Cu0.1 M KHCO3, flow-cell-1.25C2+ = 71.2%, for 8 h1000[148]
CuGa1 M KOH, flow-cell-1.07C2+ = 81.5%900[149]
Mg-Cu1 M KOH, flow-cell-0.77C2+ = 80%1000[150]

CONCLUSION AND OUTLOOK

In conclusion, with the development of catalyst synthesis and characterization methods, we gained a deeper comprehension of the catalytic CO2RR process, the interaction mechanism with the reactants, and the dynamic evolution of the active sites. Cu NPs exhibit higher selectivity towards C2+ products. However, the C2+ products are often the sum of multiple products, and improving the selectivity of individual C2+ products remains crucial. The size and morphology of Cu NPs exert a considerable influence on the efficiency of CO2RR, while the grain boundaries of Cu NPs affect the selectivity of CO2RR products. Cu NPs with varying sizes and morphologies possess different crystal planes, which tend to generate different products. Surface modification of Cu NPs with different materials can increase CO2RR activity. Cu SACs have a higher selectivity for single carbon products, especially CO and CH4. The CO2RR on Cu SACs is mainly influenced by the coordination environment and the support of Cu SACs. The selectivity and production efficiency of CO2RR can be enhanced by precise regulation of the coordination environment. Cu SACs have a higher CO2RR efficiency with N-coordination than with C-coordination and are the best coordinated by four N atoms, and the corresponding catalysts also exhibit the best stability. Especially the FE of CO products has reached nearly 100%. In terms of catalyst stability, the stability of Cu NPs is often inferior to that of Cu SACs. The possible reason is that coordination bonds can be formed between the support and single-atom Cu, thereby improving the stability of Cu SACs. Furthermore, adding additives and selecting appropriate carriers can also enhance catalytic performance and improve stability, providing ideas for designing Cu-based CO2RR catalysts in the future.

Cu-based catalysts are not stable enough, and future CO2RR catalysts being able to precisely control the active sites to improve their comprehensive performance are the focus of research, mainly in the following areas:

Precise regulation of the coordination environment, electronic structure, and the unique interaction between the supports and the Cu sites of SACs is very important to achieve superior catalytic activities. For example, a second metal atom can be introduced to form dual-atom-site catalysts based on SACs while adjusting the type of coordination atoms and the spatial configuration to change the adsorption energy of Cu sites for CO2 and adjust their catalytic performance.

NP-based single-atom sites, dual-atom sites, or clusters can be introduced to form nano-multiple-site catalysts and coexist with them. In this way, the respective advantages of these four species can be exploited in the same catalyst. The synergy in catalysis of different catalyst forms will be the future direction of CO2RR since single-atom sites, dual-atom sites, clusters, and NPs have different roles in the reaction. In short, the process of CO2RR is not singular. To obtain an efficient Cu-based catalyst, it is necessary to precisely control each catalyst component to perform its specific function, including components that enhance the adsorption energy of CO2, components that reduce the reaction energy barrier, and components that efficiently remove reaction products.

The current density of Cu-based catalysts in CO2RR is mostly low, and only a few can achieve industrial-grade current densities at high FE. Therefore, to achieve industrial commercialization, it is essential to identify appropriate strategies to increase the reaction current density. In addition to selectivity and current density, stability and one-way conversion efficiency are also crucial. It is necessary to study the dynamic changes on the surface of Cu-based catalysts and the evolution of active sites. Industrialization can be realized by developing industrial electrolytic cell devices.

DECLARATIONS

Authors’ contributions

Contributed equally to this work: Li, Q.; Jiang, J.

Literature search and organization and manuscript drafting: Li, Q.; Jiang, J.

Provided administrative and software technical: Liu, D.; Xu, D.

Manuscript revision: Jiang, S.

Supervision and suggestion: Chen, Y.

Project supervision: Liu, X.; Zhu, D.

Availability of data and materials

Not applicable.

Financial support and sponsorship

The authors gratefully acknowledge the support of the National Natural Science Foundation of China (No. 22101150; No. 22101029; No. 52201261). This work was also supported by the Beijing Municipal Natural Science Foundation (2222006), the Scientific Research Program of BJAST (23CB020, 24CB003-11), and Beijing Municipal Financial Project BJAST Young Scholar Programs B (YS202202).

Conflicts of interest

All authors declared that there are no conflicts of interest.

Ethical approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Copyright

© The Author(s) 2025.

REFERENCES

1. Ringe, S.; Clark, E. L.; Resasco, J.; et al. Understanding cation effects in electrochemical CO2 reduction. Energy. Environ. Sci. 2019, 12, 3001-14.

2. Kim, D.; Resasco, J.; Yu, Y.; Asiri, A. M.; Yang, P. Synergistic geometric and electronic effects for electrochemical reduction of carbon dioxide using gold-copper bimetallic nanoparticles. Nat. Commun. 2014, 5, 4948.

3. Yang, D.; Zhu, Q.; Sun, X.; et al. Nanoporous Cu/Ni oxide composites: efficient catalysts for electrochemical reduction of CO2 in aqueous electrolytes. Green. Chem. 2018, 20, 3705-10.

4. Lu, L.; Zhong, H.; Wang, T.; Wu, J.; Jin, F.; Yoshioka, T. A new strategy for CO2 utilization with waste plastics: conversion of hydrogen carbonate into formate using polyvinyl chloride in water. Green. Chem. 2020, 22, 352-8.

5. Sun, H.; Xu, X.; Yan, Z.; et al. Superhydrophilic amorphous Co-B-P nanosheet electrocatalysts with Pt-like activity and durability for the hydrogen evolution reaction. J. Mater. Chem. A. 2018, 6, 22062-9.

6. Wang, Y.; Han, P.; Lv, X.; Zhang, L.; Zheng, G. Defect and interface engineering for aqueous electrocatalytic CO2 reduction. Joule 2018, 2, 2551-82.

7. Zhang, H.; Wang, J.; Zhao, Z.; et al. The synthesis of atomic Fe embedded in bamboo-CNTs grown on graphene as a superior CO2 electrocatalyst. Green. Chem. 2018, 20, 3521-9.

8. Liu, M.; Pang, Y.; Zhang, B.; et al. Enhanced electrocatalytic CO2 reduction via field-induced reagent concentration. Nature 2016, 537, 382-6.

9. Xie, H.; Wang, T.; Liang, J.; Li, Q.; Sun, S. Cu-based nanocatalysts for electrochemical reduction of CO2. Nano. Today. 2018, 21, 41-54.

10. Woldu, A. R.; Huang, Z.; Zhao, P.; Hu, L.; Astruc, D. Electrochemical CO2 reduction (CO2RR) to multi-carbon products over copper-based catalysts. Coord. Chem. Rev. 2022, 454, 214340.

11. Ren, D.; Ang, B. S.; Yeo, B. S. Tuning the selectivity of carbon dioxide electroreduction toward ethanol on oxide-derived CuxZn catalysts. ACS. Catal. 2016, 6, 8239-47.

12. 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.

13. Iijima, G.; Inomata, T.; Yamaguchi, H.; Ito, M.; Masuda, H. Role of a hydroxide layer on Cu electrodes in electrochemical CO2 reduction. ACS. Catal. 2019, 9, 6305-19.

14. Carroll T, Yang X, Gordon KJ, Fei L, Wu G. Ethylene electrosynthesis via selective CO2 reduction: fundamental considerations, strategies, and challenges. Adv. Energy. Mater. 2024, 14, 2401558.

15. Qin, Q.; Suo, H.; Chen, L.; et al. Emerging Cu-based tandem catalytic systems for CO2 electroreduction to multi-carbon products. Adv. Mater. Interfaces. 2024, 11, 2301049.

16. Zheng, W.; Yang, X.; Li, Z.; et al. Designs of tandem catalysts and cascade catalytic systems for CO2 upgrading. Angew. Chem. Int. Ed. 2023, 62, e202307283.

17. Chen, H.; Mo, P.; Zhu, J.; et al. Anionic coordination control in building Cu-based electrocatalytic materials for CO2 reduction reaction. Small 2024, 20, e2400661.

18. Machado AS, Nunes da Ponte M. CO2 capture and electrochemical conversion. Curr. Opin. Green. Sustain. Chem. 2018, 11, 86-90.

19. Li, L.; Li, X.; Sun, Y.; Xie, Y. Rational design of electrocatalytic carbon dioxide reduction for a zero-carbon network. Chem. Soc. Rev. 2022, 51, 1234-52.

20. Hori, Y.; Kikuchi, K.; Suzuki, S. Production of CO and CH4 in electrochemical reduction of CO2 at metal electrodes in aqueous hydrogencarbonate solution. Chem. Lett. 1985, 14, 1695-8.

21. Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electrochemical reduction of CO at a copper electrode. J. Phys. Chem. B. 1997, 101, 7075-81.

22. Murata, A.; Hori, Y. Product selectivity affected by cationic species in electrochemical reduction of CO2 and CO at a Cu electrode. Bull. Chem. Soc. Jpn. 1991, 64, 123-7.

23. 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, 4073-82.

24. Zhang, L.; Zhao, Z. J.; Gong, J. Nanostructured materials for heterogeneous electrocatalytic CO2 reduction and their related reaction mechanisms. Angew. Chem. Int. Ed. 2017, 56, 11326-53.

25. Jones, J.; Prakash, G. K. S.; Olah, G. A. Electrochemical CO2 reduction: recent advances and current trends. Isr. J. Chem. 2014, 54, 1451-66.

26. Benson, E. E.; Kubiak, C. P.; Sathrum, A. J.; Smieja, J. M. Electrocatalytic and homogeneous approaches to conversion of CO2 to liquid fuels. Chem. Soc. Rev. 2009, 38, 89-99.

27. Hori, Y.; Murata, A.; Takahashi, R.; Suzuki, S. Electroreduction of carbon monoxide to methane and ethylene at a copper electrode in aqueous solutions at ambient temperature and pressure. J. Am. Chem. Soc. 1987, 109, 5022-3.

28. Handoko, A. D.; Wei, F.; Jenndy; Yeo, B. S.; Seh, Z. W. Understanding heterogeneous electrocatalytic carbon dioxide reduction through operando techniques. Nat. Catal. 2018, 1, 922-34.

29. Li, C.; Ji, Y.; Wang, Y.; et al. Applications of metal-organic frameworks and their derivatives in electrochemical CO2 reduction. Nanomicro. Lett. 2023, 15, 113.

30. Wang, K.; Liu, D.; Liu, L.; et al. Tuning the local electronic structure of oxygen vacancies over copper-doped zinc oxide for efficient CO2 electroreduction. eScience 2022, 2, 518-28.

31. Albo, J.; Vallejo, D.; Beobide, G.; Castillo, O.; Castaño, P.; Irabien, A. Copper-based metal-organic porous materials for CO2 electrocatalytic reduction to alcohols. ChemSusChem 2017, 10, 1100-9.

32. Merino-garcia, I.; Albo, J.; Solla-gullón, J.; Montiel, V.; Irabien, A. Cu oxide/ZnO-based surfaces for a selective ethylene production from gas-phase CO2 electroconversion. J. CO2. Util. 2019, 31, 135-42.

33. Nitopi, S.; Bertheussen, E.; Scott, S. B.; et al. Progress and perspectives of electrochemical CO2 reduction on copper in aqueous electrolyte. Chem. Rev. 2019, 119, 7610-72.

34. Christensen, O.; Zhao, S.; Sun, Z.; et al. Can the CO2 reduction reaction be improved on Cu: selectivity and intrinsic activity of functionalized Cu surfaces. ACS. Catal. 2022, 12, 15737-49.

35. Tomboc, G. M.; Choi, S.; Kwon, T.; Hwang, Y. J.; Lee, K. Potential link between Cu surface and selective CO2 electroreduction: perspective on future electrocatalyst designs. Adv. Mater. 2020, 32, e1908398.

36. Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrocatalytic process of CO selectivity in electrochemical reduction of CO2 at metal electrodes in aqueous media. Electrochim. Acta. 1994, 39, 1833-9.

37. Lum, Y.; Cheng, T.; Goddard, W. A. I. I. I.; Ager, J. W. Electrochemical CO reduction builds solvent water into oxygenate products. J. Am. Chem. Soc. 2018, 140, 9337-40.

38. Feaster, J. T.; Shi, C.; Cave, E. R.; et al. Understanding selectivity for the electrochemical reduction of carbon dioxide to formic acid and carbon monoxide on metal electrodes. ACS. Catal. 2017, 7, 4822-7.

39. Göttle, A. J.; Koper, M. T. M. Proton-coupled electron transfer in the electrocatalysis of CO2 reduction: prediction of sequential vs. concerted pathways using DFT. Chem. Sci. 2017, 8, 458-65.

40. Han, J.; Bai, X.; Xu, X.; et al. Advances and challenges in the electrochemical reduction of carbon dioxide. Chem. Sci. 2024, 15, 7870-907.

41. Li, Y. C.; Wang, Z.; Yuan, T.; et al. Binding site diversity promotes CO2 electroreduction to ethanol. J. Am. Chem. Soc. 2019, 141, 8584-91.

42. Farrell, A. E.; Plevin, R. J.; Turner, B. T.; Jones, A. D.; O'Hare, M.; Kammen, D. M. Ethanol can contribute to energy and environmental goals. Science 2006, 311, 506-8.

43. 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, 2814-21.

44. Yang, H.; Li, S.; Xu, Q. Efficient strategies for promoting the electrochemical reduction of CO2 to C2+ products over Cu-based catalysts. Chin. J. Catal. 2023, 48, 32-65.

45. Wang, Y.; Shen, H.; Livi, K. J. T.; et al. Copper nanocubes for CO2 reduction in gas diffusion electrodes. Nano. Lett. 2019, 19, 8461-8.

46. Song, Y.; Peng, R.; Hensley, D. K.; et al. High-selectivity electrochemical conversion of CO2 to ethanol using a copper nanoparticle/N-doped graphene electrode. ChemistrySelect 2016, 1, 6055-61.

47. Birdja, Y. Y.; Pérez-gallent, E.; Figueiredo, M. C.; Göttle, A. J.; Calle-vallejo, F.; Koper, M. T. M. Advances and challenges in understanding the electrocatalytic conversion of carbon dioxide to fuels. Nat. Energy. 2019, 4, 732-45.

48. Roberts, F. S.; Kuhl, K. P.; Nilsson, A. High selectivity for ethylene from carbon dioxide reduction over copper nanocube electrocatalysts. Angew. Chem. Int. Ed. 2015, 54, 5179-82.

49. Jiang, K.; Sandberg, R. B.; Akey, A. J.; et al. Metal ion cycling of Cu foil for selective C-C coupling in electrochemical CO2 reduction. Nat. Catal. 2018, 1, 111-9.

50. Feijóo, J.; Yang, Y.; Fonseca, G. M. V.; et al. Operando high-energy-resolution X-ray spectroscopy of evolving Cu nanoparticle electrocatalysts for CO2 reduction. J. Am. Chem. Soc. 2023, 145, 20208-13.

51. Yin, Z.; Yu, C.; Zhao, Z.; et al. Cu3N nanocubes for selective electrochemical reduction of CO2 to ethylene. Nano. Lett. 2019, 19, 8658-63.

52. Zhou, L.; Li, C.; Lv, J.; et al. Synergistic regulation of hydrophobicity and basicity for copper hydroxide‐derived copper to promote the CO2 electroreduction reaction. Carbon. Energy. 2023, 5, e328.

53. Su, Y.; Cheng, Y.; Li, Z.; et al. Exploring the impact of nafion modifier on electrocatalytic CO2 reduction over Cu catalyst. J. Energy. Chem. 2024, 88, 543-51.

54. Pellessier, J.; Gong, X.; Li, B.; et al. PTFE nanocoating on Cu nanoparticles through dry processing to enhance electrochemical conversion of CO2 towards multi-carbon products. J. Mater. Chem. A. 2023, 11, 26252-64.

55. Lin, Y.; Wang, T.; Zhang, L.; et al. Tunable CO2 electroreduction to ethanol and ethylene with controllable interfacial wettability. Nat. Commun. 2023, 14, 3575.

56. Rabiee, H.; Ge, L.; Zhao, J.; et al. Regulating the reaction zone of electrochemical CO2 reduction on gas-diffusion electrodes by distinctive hydrophilic-hydrophobic catalyst layers. Appl. Catal. B:. Environ. 2022, 310, 121362.

57. Zhang, Y.; Zhang, R.; Chen, F.; et al. Mass-transfer-enhanced hydrophobic Bi microsheets for highly efficient electroreduction of CO2 to pure formate in a wide potential window. Appl. Catal. B:. Environ. 2023, 322, 122127.

58. Zhao, T.; Zong, X.; Liu, J.; et al. Functionalizing Cu nanoparticles with fluoric polymer to enhance C2+ product selectivity in membraned CO2 reduction. Appl. Catal. B:. Environ. 2024, 340, 123281.

59. Reske, R.; Mistry, H.; Behafarid, F.; Roldan, C. B.; Strasser, P. Particle size effects in the catalytic electroreduction of CO2 on Cu nanoparticles. J. Am. Chem. Soc. 2014, 136, 6978-86.

60. Manthiram, K.; Beberwyck, B. J.; Alivisatos, A. P. Enhanced electrochemical methanation of carbon dioxide with a dispersible nanoscale copper catalyst. J. Am. Chem. Soc. 2014, 136, 13319-25.

61. Jung, H.; Lee, S. Y.; Lee, C. W.; et al. Electrochemical fragmentation of Cu2O nanoparticles enhancing selective C-C coupling from CO2 reduction reaction. J. Am. Chem. Soc. 2019, 141, 4624-33.

62. Yang, Y.; Louisia, S.; Yu, S.; et al. Operando studies reveal active Cu nanograins for CO2 electroreduction. Nature 2023, 614, 262-9.

63. Zhang, J.; My, P. T. H.; Gao, Z.; et al. Electrochemical CO2 reduction over copper phthalocyanine derived catalysts with enhanced selectivity for multicarbon products. ACS. Catal. 2023, 13, 9326-35.

64. Zi, X.; Zhou, Y.; Zhu, L.; et al. Breaking K+ concentration limit on Cu nanoneedles for acidic electrocatalytic CO2 reduction to multi-carbon products. Angew. Chem. Int. Ed. 2023, 62, e202309351.

65. Fu, W.; Liu, Z.; Wang, T.; et al. Promoting C2+ production from electrochemical CO2 reduction on shape-controlled cuprous oxide nanocrystals with high-index facets. ACS. Sustain. Chem. Eng. 2020, 8, 15223-9.

66. Luo, H.; Li, B.; Ma, J. G.; Cheng, P. Surface modification of nano-Cu2O for controlling CO2 electrochemical reduction to ethylene and syngas. Angew. Chem. Int. Ed. 2022, 61, e202116736.

67. Periasamy, A. P.; Ravindranath, R.; Senthil, K. S. M.; Wu, W. P.; Jian, T. R.; Chang, H. T. Facet- and structure-dependent catalytic activity of cuprous oxide/polypyrrole particles towards the efficient reduction of carbon dioxide to methanol. Nanoscale 2018, 10, 11869-80.

68. Wu, Q.; Du, R.; Wang, P.; et al. Nanograin-boundary-abundant C2O-Cu nanocubes with high C2+ selectivity and good stability during electrochemical CO2 reduction at a current density of 500 mA/cm2. ACS. Nano. 2023, 17, 12884-94.

69. Geng, Q.; Fan, L.; Chen, H.; et al. Revolutionizing CO2 electrolysis: fluent gas transportation within hydrophobic porous Cu2O. J. Am. Chem. Soc. 2024, 146, 10599-607.

70. Frese, K. W. Electrochemical reduction of CO2 at solid electrodes. In: Sullivan BP, Krist K, Guard HE, editors. Electrochemical and electrocatalytic reactions of carbon dioxide.Amsterdam: Elsevier; 1993.pp.145-216.

71. 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.

72. Schouten, K. J.; Qin, Z.; Pérez, G. E.; Koper, M. T. Two pathways for the formation of ethylene in CO reduction on single-crystal copper electrodes. J. Am. Chem. Soc. 2012, 134, 9864-7.

73. Gao, Y.; Wu, Q.; Liang, X.; et al. Cu2O nanoparticles with both {100} and {111} facets for enhancing the selectivity and activity of CO2 electroreduction to ethylene. Adv. Sci. 2020, 7, 1902820.

74. Wu, Z. Z.; Zhang, X. L.; Niu, Z. Z.; et al. Identification of Cu(100)/Cu(111) interfaces as superior active sites for CO dimerization during CO2 electroreduction. J. Am. Chem. Soc. 2022, 144, 259-69.

75. Ma, Z.; Tsounis, C.; Toe, C. Y.; et al. Reconstructing Cu nanoparticle supported on vertical graphene surfaces via electrochemical treatment to tune the selectivity of CO2 reduction toward valuable products. ACS. Catal. 2022, 12, 4792-805.

76. Chen, S.; Ye, C.; Wang, Z.; et al. Selective CO2 reduction to ethylene mediated by adaptive small-molecule engineering of copper-based electrocatalysts. Angew. Chem. Int. Ed. 2023, 62, e202315621.

77. Liu, B.; Cai, C.; Yang, B.; et al. Intermediate enrichment effect of porous Cu catalyst for CO2 electroreduction to C2 fuels. Electrochim. Acta. 2021, 388, 138552.

78. Zhang, J.; Zeng, G.; Zhu, S.; et al. Steering CO2 electroreduction pathway toward ethanol via surface-bounded hydroxyl species-induced noncovalent interaction. Proc. Natl. Acad. Sci. U. S. A. 2023, 120, e2218987120.

79. Liu, C.; Zhang, M.; Li, J.; et al. Nanoconfinement engineering over hollow multi-shell structured copper towards efficient electrocatalytical C-C coupling. Angew. Chem. Int. Ed. 2022, 61, e202113498.

80. Yang, P. P.; Zhang, X. L.; Gao, F. Y.; et al. Protecting copper oxidation state via intermediate confinement for selective CO2 electroreduction to C2+ fuels. J. Am. Chem. Soc. 2020, 142, 6400-8.

81. Chen, X.; Chen, J.; Alghoraibi, N. M.; et al. Electrochemical CO2-to-ethylene conversion on polyamine-incorporated Cu electrodes. Nat. Catal. 2021, 4, 20-7.

82. Wakerley, D.; Lamaison, S.; Ozanam, F.; et al. Bio-inspired hydrophobicity promotes CO2 reduction on a Cu surface. Nat. Mater. 2019, 18, 1222-7.

83. Qu, Y.; Li, Z.; Chen, W.; et al. Direct transformation of bulk copper into copper single sites via emitting and trapping of atoms. Nat. Catal. 2018, 1, 781-6.

84. Lang, R.; Du, X.; Huang, Y.; et al. Single-atom catalysts based on the metal-oxide interaction. Chem. Rev. 2020, 120, 11986-2043.

85. Ji, S.; Chen, Y.; Wang, X.; Zhang, Z.; Wang, D.; Li, Y. Chemical synthesis of single atomic site catalysts. Chem. Rev. 2020, 120, 11900-55.

86. Zhao, Z.; Lu, G. Cu-based single-atom catalysts boost electroreduction of CO2 to CH3OH: first-principles predictions. J. Phys. Chem. C. 2019, 123, 4380-7.

87. Cai, Y.; Fu, J.; Zhou, Y.; et al. Insights on forming N,O-coordinated Cu single-atom catalysts for electrochemical reduction CO2 to methane. Nat. Commun. 2021, 12, 586.

88. Li, Z.; Qiu, S.; Song, Y.; et al. Engineering single-atom active sites anchored covalent organic frameworks for efficient metallaphotoredox CN cross-coupling reactions. Sci. Bull. 2022, 67, 1971-81.

89. Li, X.; Yu, X.; Yu, Q. Research progress on electrochemical CO2 reduction for Cu-based single-atom catalysts. Sci. China. Mater. 2023, 66, 3765-81.

90. Li, X.; Rong, H.; Zhang, J.; Wang, D.; Li, Y. Modulating the local coordination environment of single-atom catalysts for enhanced catalytic performance. Nano. Res. 2020, 13, 1842-55.

91. Zhu, Y.; Sokolowski, J.; Song, X.; He, Y.; Mei, Y.; Wu, G. Engineering local coordination environments of atomically dispersed and heteroatom-coordinated single metal site electrocatalysts for clean energy-conversion. Adv. Energy. Mater. 2020, 10, 1902844.

92. Dong, J.; Liu, Y.; Pei, J.; et al. Continuous electroproduction of formate via CO2 reduction on local symmetry-broken single-atom catalysts. Nat. Commun. 2023, 14, 6849.

93. Wang, X.; Ju, W.; Liang, L.; et al. Electrochemical CO2 activation and valorization on metallic copper and carbon-embedded N-coordinated single metal MNC catalysts. Angew. Chem. Int. Ed. 2024, 63, e202401821.

94. Han, L.; Song, S.; Liu, M.; et al. Stable and efficient single-atom Zn catalyst for CO2 reduction to CH4. J. Am. Chem. Soc. 2020, 142, 12563-7.

95. Zhao, J.; Chen, Z.; Zhao, J. Metal-free graphdiyne doped with sp-hybridized boron and nitrogen atoms at acetylenic sites for high-efficiency electroreduction of CO2 to CH4 and C2H4. J. Mater. Chem. A. 2019, 7, 4026-35.

96. He, F.; Zhuang, J.; Lu, B.; et al. Ni-based catalysts derived from Ni-Zr-Al ternary hydrotalcites show outstanding catalytic properties for low-temperature CO2 methanation. Appl. Catal. B:. Environ. 2021, 293, 120218.

97. Liu, X.; Wang, Z.; Tian, Y.; Zhao, J. Graphdiyne-supported single iron atom: a promising electrocatalyst for carbon dioxide electroreduction into methane and ethanol. J. Phys. Chem. C. 2020, 124, 3722-30.

98. Zhao, P.; Jiang, H.; Shen, H.; et al. Construction of low-coordination Cu-C2 single-atoms electrocatalyst facilitating the efficient electrochemical CO2 reduction to methane. Angew. Chem. Int. Ed. 2023, 62, e202314121.

99. Shi, G.; Xie, Y.; Du, L.; et al. Constructing Cu-C bonds in a graphdiyne-regulated Cu single-atom electrocatalyst for CO2 reduction to CH4. Angew. Chem. Int. Ed. 2022, 61, e202203569.

100. Li, M.; Zhang, F.; Kuang, M.; et al. Atomic Cu sites engineering enables efficient CO2 electroreduction to methane with high CH4/C2H4 ratio. Nanomicro. Lett. 2023, 15, 238.

101. Chen, S.; Li, Y.; Bu, Z.; et al. Boosting CO2-to-CO conversion on a robust single-atom copper decorated carbon catalyst by enhancing intermediate binding strength. J. Mater. Chem. A. 2021, 9, 1705-12.

102. Chen, S.; Xia, M.; Zhang, X.; et al. Guanosine-derived atomically dispersed Cu-N3-C sites for efficient electroreduction of carbon dioxide. J. Colloid. Interface. Sci. 2023, 646, 863-71.

103. Purbia, R.; Choi, S. Y.; Woo, C. H.; et al. Highly selective and low-overpotential electrocatalytic CO2 reduction to ethanol by Cu-single atoms decorated N-doped carbon dots. Appl. Catal. B:. Environ. 2024, 345, 123694.

104. Pan, F.; Fang, L.; Li, B.; et al. N and OH-immobilized Cu3 clusters in situ reconstructed from single-metal sites for efficient CO2 electromethanation in bicontinuous mesochannels. J. Am. Chem. Soc. 2024, 146, 1423-34.

105. Xia, W.; Xie, Y.; Jia, S.; et al. Adjacent copper single atoms promote C-C coupling in electrochemical CO2 reduction for the efficient conversion of ethanol. J. Am. Chem. Soc. 2023, 145, 17253-64.

106. Xu, C.; Zhi, X.; Vasileff, A.; et al. Highly selective two-electron electrocatalytic CO2 reduction on single-atom Cu catalysts. Small. Struct. 2021, 2, 2000058.

107. Cheng, H.; Wu, X.; Li, X.; et al. Construction of atomically dispersed Cu-N4 sites via engineered coordination environment for high-efficient CO2 electroreduction. Chem. Eng. J. 2021, 407, 126842.

108. Karapinar, D.; Huan, N. T.; Ranjbar, S. N.; et al. Electroreduction of CO2 on single-site copper-nitrogen-doped carbon material: selective formation of ethanol and reversible restructuration of the metal sites. Angew. Chem. Int. Ed. 2019, 58, 15098-103.

109. Zhao, K.; Nie, X.; Wang, H.; et al. Selective electroreduction of CO2 to acetone by single copper atoms anchored on N-doped porous carbon. Nat. Commun. 2020, 11, 2455.

110. Roy, S.; Li, Z.; Chen, Z.; et al. Cooperative copper single-atom catalyst in 2D carbon nitride for enhanced CO2 electrolysis to methane. Adv. Mater. 2024, 36, e2300713.

111. Wu, Q. J.; Si, D. H.; Sun, P. P.; et al. Atomically precise copper nanoclusters for highly efficient electroreduction of CO2 towards hydrocarbons via breaking the coordination symmetry of Cu site. Angew. Chem. Int. Ed. 2023, 62, e202306822.

112. Lv, Z.; Wang, C.; Liu, Y.; et al. Improving CO2-to-C2 conversion of atomic CuFONC electrocatalysts through F, O-codrived optimization of local coordination environment. Adv. Energy. Mater. 2024, 14, 2400057.

113. Chen, C.; Li, Y.; Yu, S.; et al. Cu-Ag tandem catalysts for high-rate CO2 electrolysis toward multicarbons. Joule 2020, 4, 1688-99.

114. Shen, S.; Peng, X.; Song, L.; et al. AuCu alloy nanoparticle embedded Cu submicrocone arrays for selective conversion of CO2 to ethanol. Small 2019, 15, e1902229.

115. Cao, B.; Li, F.; Gu, J. Designing Cu-based tandem catalysts for CO2 electroreduction based on mass transport of CO intermediate. ACS. Catal. 2022, 12, 9735-52.

116. Ji, Y.; Guan, A.; Zheng, G. Copper-based catalysts for electrochemical carbon monoxide reduction. Cell. Rep. Phys. Sci. 2022, 3, 101072.

117. Luc, W.; Collins, C.; Wang, S.; et al. Ag-Sn bimetallic catalyst with a core-shell structure for CO2 reduction. J. Am. Chem. Soc. 2017, 139, 1885-93.

118. Ross, M. B.; Dinh, C. T.; Li, Y.; et al. Tunable Cu enrichment enables designer syngas electrosynthesis from CO2. J. Am. Chem. Soc. 2017, 139, 9359-63.

119. Peng, L.; Wang, Y.; Wang, Y.; et al. Separated growth of Bi-Cu bimetallic electrocatalysts on defective copper foam for highly converting CO2 to formate with alkaline anion-exchange membrane beyond KHCO3 electrolyte. Appl. Catal. B:. Environ. 2021, 288, 120003.

120. Hou, C.; Wang, H.; Li, C.; Xu, Q. From metal-organic frameworks to single/dual-atom and cluster metal catalysts for energy applications. Energy. Environ. Sci. 2020, 13, 1658-93.

121. Xu, Y.; Li, C.; Xiao, Y.; et al. Tuning the selectivity of liquid products of CO2RR by Cu-Ag alloying. ACS. Appl. Mater. Interfaces. 2022, 14, 11567-74.

122. Yang, Z.; Wang, H.; Fei, X.; et al. MOF derived bimetallic CuBi catalysts with ultra-wide potential window for high-efficient electrochemical reduction of CO2 to formate. Appl. Catal. B:. Environ. 2021, 298, 120571.

123. Hoang, T. T. H.; Verma, S.; Ma, S.; et al. Nanoporous copper-silver alloys by additive-controlled electrodeposition for the selective electroreduction of CO2 to ethylene and ethanol. J. Am. Chem. Soc. 2018, 140, 5791-7.

124. Ouyang, Y.; Shi, L.; Bai, X.; Ling, C.; Li, Q.; Wang, J. Selectivity of electrochemical CO2 reduction toward ethanol and ethylene: the key role of surface-active hydrogen. ACS. Catal. 2023, 13, 15448-56.

125. Morales-guio, C. G.; Cave, E. R.; Nitopi, S. A.; et al. Improved CO2 reduction activity towards C2+ alcohols on a tandem gold on copper electrocatalyst. Nat. Catal. 2018, 1, 764-71.

126. Wei, Z.; Yue, S.; Gao, S.; Cao, M.; Cao, R. Synergetic effects of gold-doped copper nanowires with low Au content for enhanced electrocatalytic CO2 reduction to multicarbon products. Nano. Res. 2023, 16, 7777-83.

127. Zheng, Y.; Zhang, J.; Ma, Z.; et al. Seeded growth of gold-copper janus nanostructures as a tandem catalyst for efficient electroreduction of CO2 to C2+ products. Small 2022, 18, e2201695.

128. Ma, Y.; Yu, J.; Sun, M.; et al. Confined growth of silver-copper janus nanostructures with {100} facets for highly selective tandem electrocatalytic carbon dioxide reduction. Adv. Mater. 2022, 34, e2110607.

129. Wei, C.; Yang, Y.; Ma, H.; et al. Nanoscale management of CO transport in CO2 electroreduction: boosting faradaic efficiency to multicarbon products via nanostructured tandem electrocatalysts. Adv. Funct. Mater. 2023, 33, 2214992.

130. Choi, C.; Cai, J.; Lee, C.; Lee, H. M.; Xu, M.; Huang, Y. Intimate atomic Cu-Ag interfaces for high CO2RR selectivity towards CH4 at low over potential. Nano. Res. 2021, 14, 3497-501.

131. Li, P.; Bi, J.; Liu, J.; et al. In situ dual doping for constructing efficient CO2-to-methanol electrocatalysts. Nat. Commun. 2022, 13, 1965.

132. Du, C.; Mills, J. P.; Yohannes, A. G.; et al. Cascade electrocatalysis via AgCu single-atom alloy and Ag nanoparticles in CO2 electroreduction toward multicarbon products. Nat. Commun. 2023, 14, 6142.

133. Qi, K.; Zhang, Y.; Onofrio, N.; et al. Unlocking direct CO2 electrolysis to C3 products via electrolyte supersaturation. Nat. Catal. 2023, 6, 319-31.

134. Li, J.; Chen, Y.; Yao, B.; et al. Cascade dual sites modulate local CO coverage and hydrogen-binding strength to boost CO2 electroreduction to ethylene. J. Am. Chem. Soc. 2024, 146, 5693-701.

135. Kattel, S.; Ramírez, P. J.; Chen, J. G.; Rodriguez, J. A.; Liu, P. Active sites for CO2 hydrogenation to methanol on Cu/ZnO catalysts. Science 2017, 355, 1296-99.

136. Zhu, J.; Ciolca, D.; Liu, L.; Parastaev, A.; Kosinov, N.; Hensen, E. J. M. Flame synthesis of Cu/ZnO-CeO2 catalysts: synergistic metal-support interactions promote CH3OH selectivity in CO2 hydrogenation. ACS. Catal. 2021, 11, 4880-92.

137. Amann, P.; Klötzer, B.; Degerman, D.; et al. The state of zinc in methanol synthesis over a Zn/ZnO/Cu(211) model catalyst. Science 2022, 376, 603-8.

138. Wan, L.; Zhang, X.; Cheng, J.; et al. Bimetallic Cu-Zn catalysts for electrochemical CO2 reduction: phase-separated versus core-shell distribution. ACS. Catal. 2022, 12, 2741-8.

139. Zhen, S.; Zhang, G.; Cheng, D.; et al. Nature of the active sites of copper zinc catalysts for carbon dioxide electroreduction. Angew. Chem. Int. Ed. 2022, 61, e202201913.

140. Zhang, Z.; Tian, H.; Bian, L.; Liu, S.; Liu, Y.; Wang, Z. Cu-Zn-based alloy/oxide interfaces for enhanced electroreduction of CO2 to C2+ products. J. Energy. Chem. 2023, 83, 90-7.

141. Liu, M.; Wang, Q.; Luo, T.; et al. Potential alignment in tandem catalysts enhances CO2-to-C2H4 conversion efficiencies. J. Am. Chem. Soc. 2024, 146, 468-75.

142. Song, J.; Zhang, H.; Sun, R.; et al. Local CO generator enabled by a CO-producing core for kinetically enhancing electrochemical CO2 reduction to multicarbon products. ACS. Nano. 2024, 18, 11416-24.

143. Wei, B.; Xiong, Y.; Zhang, Z.; Hao, J.; Li, L.; Shi, W. Efficient electrocatalytic reduction of CO2 to HCOOH by bimetallic In-Cu nanoparticles with controlled growth facet. Appl. Catal. B:. Environ. 2021, 283, 119646.

144. Wang, Z.; Li, Z.; Liu, S.; et al. Enhancing the selectivity of CO2-to-HCOOH conversion by constructing tensile-strained Cu catalyst. Mater. Today. Phys. 2023, 38, 101247.

145. Li, X.; Qin, M.; Wu, X.; et al. Enhanced CO affinity on Cu facilitates CO2 electroreduction toward multi-carbon products. Small 2023, 19, e2302530.

146. Zheng, T.; Liu, C.; Guo, C.; et al. Copper-catalysed exclusive CO2 to pure formic acid conversion via single-atom alloying. Nat. Nanotechnol. 2021, 16, 1386-93.

147. Cao, Y.; Chen, S.; Bo, S.; et al. Single atom Bi decorated copper alloy enables C-C coupling for electrocatalytic reduction of CO2 into C2+ products. Angew. Chem. Int. Ed. 2023, 62, e202303048.

148. Hu, S.; Chen, Y.; Zhang, Z.; et al. Ampere-level current density CO2 reduction with high C2+ selectivity on La(OH)3-modified Cu catalysts. Small 2024, 20, e2308226.

149. Li, P.; Bi, J.; Liu, J.; et al. p-d orbital hybridization induced by p-block metal-doped Cu promotes the formation of C2+ products in ampere-level CO2 electroreduction. J. Am. Chem. Soc. 2023, 145, 4675-82.

150. Xie, M.; Shen, Y.; Ma, W.; et al. Fast screening for copper-based bimetallic electrocatalysts: efficient electrocatalytic reduction of CO2 to C2+ products on magnesium-modified copper. Angew. Chem. Int. Ed. 2022, 61, e202213423.

151. Hao, J.; Zhu, H.; Li, Y.; et al. Tuning the electronic structure of AuNi homogeneous solid-solution alloy with positively charged Ni center for highly selective electrochemical CO2 reduction. Chem. Eng. J. 2021, 404, 126523.

152. Liu, A.; Yang, Y.; Ren, X.; et al. Current progress of electrocatalysts for ammonia synthesis through electrochemical nitrogen reduction under ambient conditions. ChemSusChem 2020, 13, 3766-88.

153. Xu, H.; Rebollar, D.; He, H.; et al. Highly selective electrocatalytic CO2 reduction to ethanol by metallic clusters dynamically formed from atomically dispersed copper. Nat. Energy. 2020, 5, 623-32.

154. Pan, F.; Yang, X.; O'carroll, T.; Li, H.; Chen, K.; Wu, G. Carbon catalysts for electrochemical CO2 reduction toward multicarbon products. Adv. Energy. Mater. 2022, 12, 2200586.

155. Yan, Y.; Ke, L.; Ding, Y.; et al. Recent advances in Cu-based catalysts for electroreduction of carbon dioxide. Mater. Chem. Front. 2021, 5, 2668-83.

156. Su, X.; Wang, C.; Zhao, F.; Wei, T.; Zhao, D.; Zhang, J. Size effects of supported Cu-based catalysts for the electrocatalytic CO2 reduction reaction. J. Mater. Chem. A. 2023, 11, 23188-210.

157. Yang, H.; Wu, Y.; Li, G.; et al. Scalable production of efficient single-atom copper decorated carbon membranes for CO2 electroreduction to methanol. J. Am. Chem. Soc. 2019, 141, 12717-23.

158. Wu, Y.; Jiang, Z.; Lu, X.; Liang, Y.; Wang, H. Domino electroreduction of CO2 to methanol on a molecular catalyst. Nature 2019, 575, 639-42.

159. Zheng, W.; Yang, J.; Chen, H.; et al. Atomically defined undercoordinated active sites for highly efficient CO2 electroreduction. Adv. Funct. Mater. 2020, 30, 1907658.

160. Feng, Z.; Tang, Y.; Ma, Y.; et al. Theoretical investigation of CO2 electroreduction on N (B)-doped graphdiyne mononlayer supported single copper atom. Appl. Surf. Sci. 2021, 538, 148145.

161. Xu, F.; Feng, B.; Shen, Z.; et al. Oxygen-bridged Cu binuclear sites for efficient electrocatalytic CO2 reduction to ethanol at ultralow overpotential. J. Am. Chem. Soc. 2024, 146, 9365-74.

162. Jiao, Y.; Zheng, Y.; Chen, P.; Jaroniec, M.; Qiao, S. Z. Molecular scaffolding strategy with synergistic active centers to facilitate electrocatalytic CO2 reduction to hydrocarbon/alcohol. J. Am. Chem. Soc. 2017, 139, 18093-100.

163. Li, H.; Cao, S.; Sun, H.; et al. CuNCN derived Cu-based/CxNy catalysts for highly selective CO2 electroreduction to hydrocarbons. Appl. Catal. B:. Environ. 2023, 320, 121948.

164. Li, Q.; Zhu, W.; Fu, J.; Zhang, H.; Wu, G.; Sun, S. Controlled assembly of Cu nanoparticles on pyridinic-N rich graphene for electrochemical reduction of CO2 to ethylene. Nano. Energy. 2016, 24, 1-9.

165. Waugh, K. Methanol synthesis. Catal. Today. 1992, 15, 51-75.

166. Chen, S.; Abdel-Mageed, A. M.; Dyballa, M.; et al. Raising the COx methanation activity of a Ru/γ-Al2O3 catalyst by activated modification of metal-support interactions. Angew. Chem. Int. Ed. 2020, 59, 22763-70.

167. Chen, J.; Falivene, L.; Caporaso, L.; Cavallo, L.; Chen, E. Y. Selective reduction of CO2 to CH4 by tandem hydrosilylation with mixed Al/B catalysts. J. Am. Chem. Soc. 2016, 138, 5321-33.

168. Sampson, M. D.; Kubiak, C. P. Manganese electrocatalysts with bulky bipyridine ligands: utilizing lewis acids to promote carbon dioxide reduction at low overpotentials. J. Am. Chem. Soc. 2016, 138, 1386-93.

169. Kim, Y. E.; Kim, J.; Lee, Y. Formation of a nickel carbon dioxide adduct and its transformation mediated by a lewis acid. Chem. Commun. 2014, 50, 11458-61.

170. Chen, S.; Wang, B.; Zhu, J.; et al. Lewis acid site-promoted single-atomic Cu catalyzes electrochemical CO2 methanation. Nano. Lett. 2021, 21, 7325-31.

171. Yuan, J.; Zhang, J.; Yang, M.; Meng, W.; Wang, H.; Lu, J. CuO nanoparticles supported on TiO2 with high efficiency for CO2 electrochemical reduction to ethanol. Catalysts 2018, 8, 171.

172. Chu, D.; Qin, G.; Yuan, X.; Xu, M.; Zheng, P.; Lu, J. Fixation of CO2 by electrocatalytic reduction and electropolymerization in ionic liquid-H2O solution. ChemSusChem 2008, 1, 205-9.

173. Thompson, T. L.; Diwald, O.; Yates, J. T. CO2 as a probe for monitoring the surface defects on TiO2(110)-temperature-programmed desorption. J. Phys. Chem. B. 2003, 107, 11700-4.

174. Cueto, L. F.; Hirata, G. A.; Sánchez, E. M. Thin-film TiO2 electrode surface characterization upon CO2 reduction processes. J. Sol-Gel. Sci. Technol. 2006, 37, 105-9.

175. Abdinejad, M.; Subramanian, S.; Motlagh, M. K.; et al. Insertion of MXene-based materials into Cu-Pd 3D aerogels for electroreduction of CO2 to formate. Adv. Energy. Mater. 2023, 13, 2300402.

176. Lin, L.; Liu, T.; Xiao, J.; et al. Enhancing CO2 electroreduction to methane with a cobalt phthalocyanine and zinc-nitrogen-carbon tandem catalyst. Angew. Chem. Int. Ed. 2020, 59, 22408-13.

177. Chen, S.; Li, W. H.; Jiang, W.; et al. MOF encapsulating N-heterocyclic carbene-ligated copper single-atom site catalyst towards efficient methane electrosynthesis. Angew. Chem. Int. Ed. 2022, 61, e202114450.

Cite This Article

Review
Open Access
Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms
Qianwen Li, ... Xiangwen LiuXiangwen Liu

How to Cite

Li, Q.; Jiang, J.; Jiang, S.; Liu, D.; Xu, D.; Chen, Y.; Zhu, D.; Liu, X. Catalyst design for the electrochemical reduction of carbon dioxide: from copper nanoparticles to copper single atoms. Microstructures 2025, 5, 2025003. http://dx.doi.org/10.20517/microstructures.2024.69

Download Citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click on download.

Export Citation File:

Type of Import

Tips on Downloading Citation

This feature enables you to download the bibliographic information (also called citation data, header data, or metadata) for the articles on our site.

Citation Manager File Format

Use the radio buttons to choose how to format the bibliographic data you're harvesting. Several citation manager formats are available, including EndNote and BibTex.

Type of Import

If you have citation management software installed on your computer your Web browser should be able to import metadata directly into your reference database.

Direct Import: When the Direct Import option is selected (the default state), a dialogue box will give you the option to Save or Open the downloaded citation data. Choosing Open will either launch your citation manager or give you a choice of applications with which to use the metadata. The Save option saves the file locally for later use.

Indirect Import: When the Indirect Import option is selected, the metadata is displayed and may be copied and pasted as needed.

About This Article

Special Issue

© The Author(s) 2025. Open Access This article is licensed under a Creative Commons Attribution 4.0 International License (https://creativecommons.org/licenses/by/4.0/), which permits unrestricted use, sharing, adaptation, distribution and reproduction in any medium or format, for any purpose, even commercially, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made.

Data & Comments

Data

Views
122
Downloads
42
Citations
0
Comments
0
0

Comments

Comments must be written in English. Spam, offensive content, impersonation, and private information will not be permitted. If any comment is reported and identified as inappropriate content by OAE staff, the comment will be removed without notice. If you have any queries or need any help, please contact us at support@oaepublish.com.

0
Download PDF
Share This Article
Scan the QR code for reading!
See Updates
Contents
Figures
Related
Microstructures
ISSN 2770-2995 (Online)

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/

Portico

All published articles are preserved here permanently:

https://www.portico.org/publishers/oae/