www.acsaem.org Spotlight on Applications Heterogeneous Electrocatalysts for CO2 Reduction Chao Yang, Yuhang Wang, Linping Qian, Abdullah M. Al-Enizi, Lijuan Zhang, and Gengfeng Zheng* Cite This: ACS Appl. Energy Mater. 2021, 4, 1034−1044 Downloaded via FORMOSA PLASTICS CORP on April 24, 2023 at 08:06:21 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles. ACCESS Read Online Metrics & More Article Recommendations ABSTRACT: The electrocatalytic CO2 reduction reaction (CO2RR) has been a fast developing and innovative topic in recent years. It may enable efficient reduction of carbon emission, as well as store renewable electricity into chemical bonds in fuels or other chemicals. However, due to the complexity of reaction factors and mechanisms, the activity and selectivity of CO2RR have yet to be further optimized in order to realize commercial applications. Specifically, the high selectivity of CO2 reduction products (especially for multi-electron-transfer and multicarbon reduction products), lower overpotentials, high energy efficiencies, good stability, and low fabrication cost are required. In this Spotlight, we focus on the design of active sites in heterogeneous catalysts for electrochemical CO2 reduction, and we provide several strategies to target those challenges in this field. First, as the major side reaction in aqueous solutions, the hydrogen evolution reaction can be inhibited by increasing the energy barrier for H2 formation and tuning the electrolyte diffusion toward the electrocatalyst surface, thus enhancing the Faradaic efficiencies of CO2RR products. Second, the proximity of adjacent catalytic sites suggests a strong capability for tuning the coupling efficiency of multiple carbon atoms. Third, the active sites of heterogeneous electrocatalysts can be generated by different defects, including single-atom doping, anion vacancy, alloy formation, and lattice defects, which can lead to distinctively different production distributions. Finally, a solar-driven electrocatalytic CO2 reduction system was developed, enabling a high solar-to-fuel photoconversion efficiency. Further investigations of heterogeneous electrocatalysts and reaction systems are expected to enhance the electrochemical CO2RR performances toward the application level. KEYWORDS: electrochemical CO2 reduction, heterogeneous catalyst, adsorption, copper, Faradaic efficiency 1. INTRODUCTION Influenced by anthropogenic activities, the content of carbon dioxide (CO2) in the atmosphere is continuously increasing.1 Among many different CO2 utilization pathways, the electrocatalytic conversion of CO2 into chemical products represents a unique and promising strategy.2 In this process, using electricity energy as the input, CO2 can be potentially transformed into a range of carbon-based molecules, capable of reducing CO2 emission and enabling storage of chemical energy.3 In addition, it is ideal to use renewable resources such as solar or wind energies to produce electricity for potential scale-up applications.4 Using technoeconomic analysis methods,5 the feasibility and limitations of the electrocatalytic CO2RR (CO2 reduction reaction) have been studied. Carbon monoxide (CO) and formate (HCOOH) are among the highest efficiency CO2RR products under the current electrolysis conditions.5,6 Nonetheless, the low electrochemical performances toward C2 products, such as ethylene, ethanol, and n-propanol, should be improved drastically to fulfill a reasonably profitable level, including high partial current densities of >300 mA cm−2, high Faradaic efficiencies (FEs) of >80−90%, low full cell operating © 2021 American Chemical Society voltage, and excellent stability (>1000 h). Recently, the increasing use of gas-diffusion-layer (GDL)-based flow-cell systems and membrane−electrode assemblies (MEAs) has suggested attractive capabilities for realizing these metrics.7 In contrast to the conventional H-type electrolysis cells, a solid− liquid−gas triphase interface formed near the GDL and electrocatalysts helps to solve the challenge of CO2 mass transfer in aqueous solutions and achieve high current densities that can be implemented commercially.8 2. ELECTROLYTE EFFECT In spite of the attractive potential in the field of CO2RR, the complexity of this reaction still restricts an understanding of the overall process. The electrocatalytic CO2RR is typically Received: October 28, 2020 Accepted: December 24, 2020 Published: January 21, 2021 1034 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Figure 1. Schematic of several possible reaction pathways of electrochemical CO2 reduction reaction toward methane, ethylene, and ethanol, respectively. conducted in aqueous electrolytes, which is attributed to their abundant proton resources and excellent conductivity at room temperature.9 However, the understanding of local environments near the interface of the solid−liquid−gas triphase remains quite limited.10 First, protons are consumed by both CO2RR and hydrogen evolution reaction (HER), resulting in an increasing local pH near the electrode surface. Second, the reaction of CO2 and water generation are also influenced by the multiequilibrium of CO2/bicarbonate/carbonate, and simultaneously, the formation of carbonates/bicarbonates at the interface leads to a change in catalyst performance.11 The difference in pH can lead to the change in specific product selectivity through the interactions with the H* intermediate, which is regarded as the Langmuir−Hinshelwood mechanism (i.e., surface recombination reaction).12 In this process, a general pH-dependence of onset potentials for methane formation has been observed.13 A higher Faradaic efficiency for producing CH4 was also demonstrated in a concentrated bicarbonate electrolyte, such as in 1 M KHCO3 (where the surface pH was ca. 8.9) versus in 0.05 M KHCO3 (with pH 9.8).14 Indeed, the role of the CO2/bicarbonate/carbonate equilibrium reaction may be not only as a pH buffer, but also as a carrier of reaction species. Shao et al. proposed the role of bicarbonate anions in CO2RR by fast real-time infrared (IR) spectroscopy.15 When the potential was stepped from 0.2 to −0.6 V in 12CO2-saturated KH12CO3 and KH13CO3 solutions, the IR band from 12CO was increased gradually due to the conversion of H13CO32− into H12CO32−. Thus, it was suggested that, under reaction conditions, CO2 molecules cannot reach the outer Helmholtz plane densely packed by the hydrate anions; instead, carbonate anions can release CO2 molecules near the surface for CO2RR. The mechanism was proposed as the fast equilibrium below: CO2 + H2O + OH− ⇌ HCO3− + H2O ⇌ H2CO3 + OH−. In another study, the bicarbonate anion can also serve as a proton donor.16 Thus, it is critical to investigate the role of the electrolyte on the electrochemical CO2RR. In addition to the equilibrium of CO2/HCO3−/H2CO3, the hydrated cations can also affect the interfacial interactions at the surface.17 For example, K+ can stabilize the adsorbed intermediates to lower the thermodynamic energy barrier. Sargent and co-workers reported the field-induced K + enrichment effect by a synthesized gold (Au) nanoneedle catalyst.18 With the increasing size of the alkali cations, the pKa of the cations decreases, leading to the increase in CO2 concentration and decrease in pH.19 Controlling the transport process to tune the reaction pathways is another useful method. A membrane coated electrode can generally enhance the activity and selectivity of electrochemical CO2RR; this is attributed to both their functions, such as the hydrophilic,20 gas adsorption capability, and the capability for stabilizing intermediates for multielectron-transfer reduction products.21 3. COMPLEXITY OF HETEROGENEOUS CATALYSTS FOR REACTION PATHWAYS The complexity of CO2RR is attributed to a variety of heterogeneous catalysts, reaction mechanisms, electrolysis conditions, and target products.22,23 The thermodynamic potentials of electrochemical CO2RR to different products are comparable to that of HER.24,25 For a two-electron-transfer reaction with a suitable electrocatalyst, the CO2RR typically starts to occur at low overpotentials. The two-electron-transfer products, CO or HCOOH, can go through different pathways. *COOH is an important intermediate for CO formation by proton-coupled electron-transfer (PCET) reactions, while *OCOH is likely to be the intermediate for formate.26 The reaction pathway selectivity of these two species is strongly determined by the catalysts. For example, the major CO2RR product of Bi and In is formate, while for Au and Ag it is CO.27 However, for multi-electron (>2e)-transferred products, the kinetics of CO2RR becomes more sluggish, leading to larger overpotentials and lower selectivity. In addition, the multielectron-transfer pathways can have various reaction mechanisms and subsequent product selectivities.28,29 Various theoretical methods, such as density functional theory (DFT) and molecular dynamics (MD) simulations, have been utilized to guide CO2RR experiments.30 Using a solvation method, the pathways of C1 and C2 products have been studied.31 *CO−COH is formed by the coupling of *CO and *COH at a neutral pH. However, the dimerization of *CO to form *CO−CO is a critical pathway toward C2 products at alkaline pHs. Further selectivity on ethylene and ethanol can involve the surface-adsorbed H2O (Figure 1). In combination with the isotope labeling and spectroscopy methods, multiple CO2RR pathways have been suggested.32 Particularly, *CO has been proposed as an indispensable intermediate in the multielectron-transfer reductions.33 The density and capability of the CO adsorption active sites in the electrocatalysts are 1035 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Figure 2. Schematic of different research methods on electrochemical CO2 reduction. Figure 3. (a) Schematic illustration of further passivation of HER activity by reduction of the number of active sites via Li+ association. The brown, green, red, blue, and purple spheres represent C, H, O, N, and Li atoms. (b) CV curves of PNFE/CNT networks in 0.5 M Li2SO4 and 0.5 M H2SO4 aqueous solutions, respectively. Reproduced with permission from ref 37. Copyright 2016 Wiley. (c) CO partial current densities versus KHCO3 concentration at different constant potentials for the Sn-OD-Cu in CO2-saturated KHCO3 solution with different concentrations. Reproduced with permission from ref 38. Copyright 2019 American Chemical Society. (d) Summary of N atomic contents in CN-CNTs and CNH−CNTs. (e) Faradaic efficiency for CO, H2, and CH4 versus potential on CN-H−CNTs. Reproduced with permission from ref 39. Copyright 2017 Wiley Publications. (f) The FEs for electrochemical reduction of CO2 measured on MNC-D. Reproduced with permission from ref 40. Copyright 2019 Springer Nature Publications. (g) XRD patterns of TNS-0.9-SnO2, TNS-2.0-SnO2, and TNS-3.0-SnO2. (h) Faradaic efficiencies for HCOOH production of TNS-0.9-SnO2, TNS-2.0-SnO2, and TNS-3.0-SnO2 at each applied potential for 2 h. Reproduced with permission from ref 41. Copyright 2018 Wiley Publications. 1036 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications In addition to the interaction between cations and electrocatalysts, the concentration effect of KHCO3 electrolyte in CO2RR was also systematically investigated. Cu2O@SnOx was used as an example to first show a stable selectivity toward CO with good FECO (>85%). The concentration effect of KHCO3 was then studied.38 The activity for H2 production was initially suppressed with the increasing KHCO3 concentration over low overpotentials, while the production of CO was promoted. The Cu2O@SnOx catalyst presented a maximum FECO of 98.1% in 0.7 M KHCO3 at a moderate overpotential of 490 mV. A linear correlation was also found between log(jCO) and log[HCO3−], which suggested that the CO2RR rate had a first-order dependence on [HCO3−] (Figure 3c). Combined with the Tafel analysis, the CO2RR mechanism in this system was proposed as CO2 + HCO3− + e− + * → COOH* + CO32−, where * denotes an active site. Thus, the enhancement of CO2RR to produce CO should be attributed not only to the pH effect, but also to the concentration effect. Furthermore, the catalyst design can also influence the competitivity between CO2RR and HER. For instance, metalfree carbon materials have been reported by many research groups to be one of the most promising catalysts for CO2RR, due to their distinct inherent characteristics, such as porous structures, high surface area, and tunable structural configurations.55 Although pristine (i.e., undoped) carbon materials show limited catalytic activity toward CO2 reduction, introducing nitrogen atoms is a promising approach. A nitrogen atom has a larger electronegativity than carbon, and it can create electroactive sites when it is doped into carbon materials. Nitrogen exists in carbon materials in four types: pyridinic, pyrrolic/pyridonic, quaternary or graphitic, and pyridine-N-oxide.56 Among them, pyrrolic nitrogen presents a good activity for CO2RR but is sluggish toward HER, while pyridinic and graphitic nitrogen atoms are active HER catalysts. In 2017, our group reported a steam-etching method to modulate the ratio of different nitrogen doping effects.39 Due to their different affinities for water molecules, the carbon atoms adjacent to pyridinic and graphitic nitrogen atoms were more prone to be etched away by steam. As a result, the nitrogen-doped carbon network exhibited an increasing content of pyrrolic nitrogen from 22.1% to 55.9% among the total N species, which were designated as CN-CNT and CNH-CNT, respectively (Figure 3d). Cyclic voltammetry (CV) tests conducted in both Ar-saturated and CO2-saturated electrolytes showed that, compared to an unetched CN-CNT sample, the CN-H-CNT (i.e., after steam etching) presented a higher overpotential for HER and a lower overpotential for CO2RR. The FE of CO formation (FECO) reached 88% for CN-H-CNT, while it is only ∼60% for unetched CN-CNT (Figure 3e). Our observation indicated that the HER inhibition can benefit the performance of the CO2RR. In a subsequent study, we further demonstrated that nitrogen atom configurations were also tuned by the secondary doping process. The edge sites of mesoporous N-doped carbon frameworks were easier to attack by a second nitrogen doping source (such as N,N-dimethylformamide, DMF) than inner doping sites. This secondary N-doping process led to an increase in the content of the nitrogen dopant level and FECO of ∼92% (Figure 3f).40 It is also important to note that, in the studies of metal-free catalysts, trace metal contaminants may dominate the performance of CO2RR, instead of the intrinsic activity of metal-free catalysts. Meanwhile, the trace amount of contaminants may important. Several experimental results have reported the combined effects of both *CO surface density and adsorption energy.34−36 Thus, it is critical to rationally design and tune the active sites on heterogeneous electrocatalysts for CO2RR. 4. OUR WORK The CO2RR research in our group can be generalized in the following aspects (Figure 2). First of all, the competition with the major side reaction, HER, restricts the performance of many catalysts. By limiting the surface adsorption site density and adsorption capability of the H atoms, both thermodynamics and kinetics of HER can be inhibited,37−41 thus potentially leading to the enhancement of CO2RR products. Second, the CO2RR selectivity can be tuned by the surface density of the active sites,42−44 in which the adjacent two active sites promote the carbon bond formation between two adsorbed *CO intermediates.45 Third, the nanostructured catalysts have been designed and fabricated with different defect engineering, strain, and alloy effects, which allow high performances for C2 products to be exhibited.42,46,47 Last but not least, pushing the activity of electrocatalytic CO2 reduction reaction by system engineering has also been investigated. In addition to the use of flow-cell and MEA systems,48 we developed a redox-mediated electrolysis system to efficiently convert solar-generated electricity to CO2RR products with an attractive energy efficiency and solar-to-fuel efficiency.49 4.1. Competitive Reaction of CO2RR and HER. As a simple 2e−-transfer reaction, HER (i.e., the water reduction) is universal for electrochemical reductions in aqueous electrolytes.50 Water reduction generally takes place at a lower (less negative) thermodynamic potential compared to that of CO2RR. To improve the performance of CO2RR, the inhibition of HER should be an effective strategy, which is correlated with two aspects: the electrolyte environment and the electrocatalyst structure.51−53 Combined with the aspects mentioned above, a series of studies have been reported. For instance, polyimides are generally regarded as an anode candidate for Li and Na batteries due to their reversible redox capability,54 while the voltage outputs are still far from optimum without sufficient surface engineering for HER passivation. In 2016, our group demonstrated the work of inhibiting HER on a polyimide surface, i.e., poly(naphthalene four formyl ethylenediamine) (PNFE) as an example.37 The four oxygen atoms in each repeating unit of PNFE were first verified as the major adsorption sites for hydrogen atoms, with the lowest adsorption energy barrier (Had) compared to the carbon and nitrogen atoms. DFT calculations demonstrated the intrinsic sluggish HER of PNFE in a 0.5 M H2SO4 electrolyte without Li+ (Figure 3a). To further passivate the HER activity, a Li+containing electrolyte was used. The strong affinity for Li+ ions led to the blocking of the adsorption sites (i.e., O atoms) for hydrogen atoms, leading to a higher energy barrier and more sluggish kinetics for H2 formation. Thus, both the Tafel (2Hadsorption → H2) and Heyrovsky (Hadsorption + H2O + e− → H2 + OH−) pathways were inhibited. When coating on a carbon nanotube (CNT) substrate, the PNFE/CNT electrode exhibited a large HER onset overpotential of 820 mV vs a reversible hydrogen electrode (RHE) in 0.5 M Li2SO4 aqueous solution (Figure 3b). Li+ ions blocking the H-adsorption sites may further improve the concentration of cations to tune the performance of CO2RR. 1037 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Figure 4. (a) Structure of synthesized Cu-N-C-800, -900. (b) Faradaic efficiencies (top panel, left y-axis), partial current densities of CH4 and C2H4 (top panel, right y-axis), and the ratios of CH4/C2H4 (bottom panel) for Cu-N-C-T (T = 800, 900, 1000, and 1100 °C) catalysts at −1.6 V vs RHE. Reproduced with permission from ref 45. Copyright 2020 American Chemical Society. (c) Structure models of CeO2(110) doped with one single Cu site, with three oxygen vacancies. (d) Faradaic efficiency comparison of samples with different Cu doping levels. Reproduced with permission from ref 63. Copyright 2018 American Chemical Society. not be realized, as it is hard to be quantified by XPS or EDX techniques.57 In an electrochemical system, the source of impurities can be quite diverse, such as from reactants, electrolyte, counter electrode, reference electrode, and so on.58 In our work, in order to avoid challenges from impurities, a series of rational control experiments should be conducted to exclude the effects from potential contamination sources during synthesis or electrochemical measurements. Moreover, from the view of material design, the construction of a hydrophobic surface has been a common way to inhibit electrolyte diffusion,20 while at the cost of limiting gas diffusion. An alternative approach to regulate the mass transfer is to tune the confined space of the electrocatalyst environment. In 2018, our group developed a two-dimensional (2D) assembly of active catalyst and inert spacer components, with tunable interspacing distances.41 In brief, atomically thin titania nanosheets (TNS) and SnO2 nanoparticles were assembled into highly ordered 2D nanosheets via molecular interaction between TNS and cationic surfactants. The interspace distance between adjacent TNS layers was tuned by different types of cationic surfactants, ranging from 1 to 3 nm. The X-ray diffraction patterns of (00k) planes confirmed the structure of lamella assemblies with different interspacing distances (i.e., 0.9, 2.0, and 3.0 nm, respectively) and the existence of the SnO2 nanoparticles inside adjacent TNS layers (Figure 3g). The electrochemical CO2 performances were analyzed for these samples. Due the capability of differentiating masstransfer rates between CO2 and H2O, the TNS-2.0-SnO2 (with 2.0 nm interlayer spacing) presented the highest FEHCOO− of 73% at −1.6 V vs RHE, while TNS-0.9-SnO2 (without ordered assembly of SnO2 nanoparticles inside 2D confined space) had the lowest FEHCOO− of 39% (Figure 2h). In addition, the largest interlayer spacing in TNS-3.0-SnO2 (i.e., 3.0 nm) was unable to differentiate the diffusion rates of CO2 and H2O molecules, so the HER production was still high, thus substantiating the effect of the confined space of the 2D assembly on CO2RR. 4.2. Promotion of CO2 Deep Reduction. Copper (Cu) is known as a unique electrocatalyst to obtain deep reduction (i.e., multi-electron-transfer) products. The common deep reduction products are methane (CH4), ethylene (C2H4), ethanol (C2H5OH), and propanol (C3H7OH).27 This unique feature of Cu can be attributed to its moderate adsorption for both carbon and hydrogen.59 The current research on Cubased CO2RR electrocatalyst designs involves various aspects, such as doping,60 surface facets,61 defect effects,42 and so on. However, many challenges still exist with a variety of different 1038 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Figure 5. (a) Schematic illustration of intermediate adsorption controlling which copper facets are exposed, with the chemisorbed intermediates *CO acting as capping agents, leading to a high portion of Cu(100) in the forming catalyst. Reproduced with permission from ref 64. Copyright 2019 Springer Nature. (b) Schematic of the synthesis of CuO nanoparticles at different cooling rates. Reproduced with permission from ref 46. Copyright 2020 Elsevier. (c) FEH2 and FEalcohol for ECR on a Cu matrix and Cu3Ag1 at different working potentials. (d) Ratio of alcohols among Cx products at different potentials for Cu3Ag1. Reproduced with permission from ref 70. Copyright 2020 Wiley Publications. majority of Cu species coordinated with nitrogen atoms remained in the form of Cu(I)−N2 pairs, while at a higher temperature of 900 °C (named as Cu-N-C-900), single-atomic Cu(II)−N4 was the major form of Cu dopants. Fouriertransformed extended X-ray absorption fine structure (FTEXAFS) confirmed both structures and indicated that the distance between adjacent Cu(I)−N2 species was significantly shorter than that of individual Cu(II)−N4 species. The electrochemical CO2RR performance showed that the Cu(I)−N2 pairs promoted the C−C coupling and subsequent formation of C2H4, while single-atomic Cu(II)−N4 sites catalyzed CO2RR toward CH4 (Figure 4b). In addition to the regulation of nitrogen-doped carbon, metal oxides are also good substrate candidates. For example, ceria (CeO2) is an attractive substrate by providing strong metal−support interactions and enhancing the dispersion of loaded metal species.62 In 2018, we reported a unique structure of single-atomic Cu-substituted CeO2, which demonstrated mechanistic pathways. In our study, we have conducted research on copper-based catalysts from two aspects. 4.2.1. Single-Atomic Cu Configuration for CO2RR. CO is a crucial intermediate in the deep reduction and C−C coupling step. The coverage and density of CO can strongly affect the product selectivity. 8 In catalysts with a single-atomic distribution of Cu sites, the limited active sites and relatively far distance between adjacent sites can hinder the coupling of two adsorbed *CO species. Therefore, CH4 can be the main product in this process. For example, we investigated the relationship of CO2RR selectivity and the surface Cu density.45 Here, nitrogen-doped carbon was used as the framework to promote the dispersion of copper species. The distance between adjacent Cu species was adjusted by a different thermal pyrolysis of a Cu-containing metal−organic framework (MOF). The contents of Cu decreased gradually with the increase of pyrolysis temperature (Figure 4a). At a lower calcined temperature of 800 °C (named as Cu-N-C-800), a 1039 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Figure 6. (a) Energy diagram of each part in the redox-medium-assisted system. (b) Photovoltaic and electrocatalytic current−voltage curves of the GaAs solar cell (blue curve) and two-electrode O2 evolution (red curve). The red dot indicates the maximum power conversion efficiency (PCE) point of the designed artificial light reaction, based on photoelectrochemical O2 evolution and Zn deposition. (c) FECO (red bars) and concentration of CO production (blue bars) at different current densities of 2, 5, and 8 mA cm−2. (d) The voltage-versus-time profile, FECO, and solar-to-CO PCE of the redox-medium-assisted system at a current density of 10 mA cm−2. Reproduced with permission from ref 49. Copyright 2018, Wang et al., under a Creative Commons CC BY license. A high Faradaic efficiency toward producing C2H4 (FEC2H4) of 63% was reported, substantially exceeding pure Cu or CuOx without oxygen vacancies. Designing and engineering electrocatalysts with specific crystal planes or defects have been the focus of many studies.42,64 For instance, Cu catalysts with preferentially exposed Cu(100) facets were generated for CO2 electroreduction, which achieved ∼90% Faradaic efficiency for C2+ products at a partial current density of 520 mA cm−2 (Figure 5a).64 In addition, when a Cu catalyst was electrochemically deposited under an environment with a high concentration of CO, defects with a high density were formed, which was distinctively different from the Cu catalyst synthesized under otherwise identical conditions but without the CO environment.65 Theoretical calculations suggested that defects were favored to be formed on the Cu surface with a high density of surface-adsorbed *CO during catalyst synthesis, which, in turn, helped to maintain a high surface *CO coverage when this defect-rich CuDS catalyst was used for CO2RR. As a result, the CuDS catalyst achieved a high CO2RR selectivity toward ethanol in both H-type electrochemical cells and flow cells. In another study, a fast cooling approach was utilized to create a high density of microstrains and grain boundaries in copper oxide, which presented a substantial chemisorption capability toward CO2 and CO (Figure 5b).55 Compared to other CuO catalysts formed under slower cooling rates, the fast cooled, strain-rich CuO catalyst enabled an excellent total current density of >300 mA cm−2, as well as an FE for the combined C2 products (majority as alcohols) of 78% at −1.0 V vs RHE. that oxygen vacancies in CeO2 could be spontaneously enriched upon the introduction of single-atomic Cu sites.63 DFT calculations revealed a stable structure with up to 3 oxygen vacancies that were neighboring to a single Cu site, which further facilitated the adsorption and activation of CO2 (Figure 4c). Experimentally, different mass loadings of Cu were synthesized by a facile immersion−precipitation method. At a low (<5%) doping density, the Cu species remained a highly dispersive feature of single-atomic sites on CeO2. The low oxidation states of [Cu0 + Cu1+] further supported the existence of multiple oxygen vacancies around the Cu sites. Such a configuration of both single-atomic Cu sites and multiple oxygen vacancies played a beneficial role in electrochemical CO2RR performance. The Cu-doped CeO2 showed a high activity of electrochemical CO2 reduction compared to Cu nanoparticles or undoped CeO2, with an excellent CH4 production FE of 58% at −1.8 V vs RHE (Figure 3d). 4.2.2. Cu-Based Nanomaterials for CO2RR. Distinctive from single-atomic Cu catalysts, Cu-based nanomaterials can facilitate the coupling of C−C, ascribed to their adjacent Cu sites. Oxygen vacancies can play an important role in the Cucontaining oxides as well, which serve as Lewis base sites for enhanced adsorption of CO2 and subsequent reaction intermediates. For instance, our group reported the fabrication of oxygen vacancy-rich CuOx nanodendrites by an in situ electrochemical conversion method.47 The rich oxygen vacancies provided a high density of electrons to activate and reduce CO2, while the high density of Cu sites facilitated the coupling of two *CO intermediates adsorbed in adjacent sites. 1040 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Alloying copper with other metals has also been proposed as a potential approach to tuning the electrocatalytic behaviors.59 For most bimetallic catalysts, several effects are potentially involved to present the structure−performance relationship, namely, the ensemble effect, ligand effect, and strain effect.66 Cu and Ag are catalyst candidates widely used in CO2RR, but the selective products can be substantially altered when they form alloys.67−69 For instance, Cu3Ag1 catalyst was identified with electron-deficient Cu catalytic sites, which were attributed to the interphase electron transfer from Cu to Ag.70 Compared to the pure Cu(100), the electron-deficient Cu sites helped to stabilize the alcohol pathway intermediates, such as CH3CH2O*, and thus, the production of alcohol was enhanced. The FE of alcohols (including both ethanol and propanol) reached 63% at −0.95 V vs RHE (Figure 5c), and the selectivity of alcohols among all the CO2RR production was as high as 92% (Figure 5d). 4.3. Solar-Driven CO2 Electroreduction. To further promote this technology into large-scale applications, using renewable energy-produced electricity should be the reasonable path toward reducing the carbon footprint. For example, Grätzel and co-workers demonstrated that, by combining a tandem device from a commercial photovoltaic cell and SnO2modified CuO nanowires as the CO2RR electrocatalyst, the solar-to-CO conversion efficiency was reported to be 13.4%.71 Nonetheless, the sunlight fluctuation places a challenge when applying this direct coupling of photovoltaics and electrolyzers. Inspired by the process of natural photosynthesis where an ATP/ADP redox medium is used to transfer the photoenergy to the dark step of carbon fixation, our group developed a Zn/ Zn(OH)42− redox medium to decouple the solar-driven water oxidation (i.e., the light reaction) and the electrochemical CO2RR (i.e., the dark reaction) (Figure 6a).49 Functioning like a Zn battery, this Zn/Zn(OH)42− redox medium assisted to transfer electrical energy from the light reaction toward the dark reaction, with >99% Faradaic efficiency of reversibility. The electrolyte was 1 M KOH with Zn(OH)42−, and the reaction was driven by a triple-junction InGaP/GaAs/Ge solar cell. In the light reaction, NiFe hydroxide was used as the water oxidation electrocatalyst in the anode, and a Zn plate was used as a cathode (Figure 6c). The electrochemical energy was then stored at the Zn/Zn(OH)42− redox medium, which allowed spontaneous reduction of CO2 to CO using a nano-Au catalyst. The yield and selectivity of CO2RR were tuned by applying different electrolysis current densities (i.e., 2, 5, and 8 mA cm−2) under a constant current mode (Figure 6b). The electrical energy conversion efficiency for CO2RR was up to 63%, and the total solar-to-CO efficiency was maintained at ∼15.6% for over 24 h (Figure 6d). developed and implemented to investigate surface reactive intermediates and reaction mechanisms. From the technoeconomic analysis results, a promising option for the fuel production by electrocatalytic CO2RR should require rational consideration as an integral view, which includes but is not limited to improving highly active/selective electrocatalytic sites, enhancing mass transfer, and optimizing the energy efficiency of the whole system from the electricity input toward the product formation. Last but not least, the stability of the electrocatalytic systems should also be reasonably evaluated and improved to guide the rational design of this potential technology toward large-scale implementation. 5. CONCLUSIONS This Spotlight has briefly introduced the electrochemical CO2RR and classified this into two aspects, including the electrolyte effect and the complexity of heterogeneous catalysts for reaction pathways. Several examples of recent progress in our research group are provided to show some strategies in tuning and understanding both the electrocatalyst and electrolyte effects. In spite of the progress by our group and many other researchers, many more efforts are still needed to understand and push forward this field of electrochemical CO 2 RR. In addition, more in situ/operando analysis techniques and multiscale theoretical models should also be Notes ■ AUTHOR INFORMATION Corresponding Author Gengfeng Zheng − Laboratory of Advanced Materials, Department of Chemistry, and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China; orcid.org/00000002-1803-6955; Email: gfzheng@fudan.edu.cn Authors Chao Yang − Laboratory of Advanced Materials, Department of Chemistry, and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China Yuhang Wang − Laboratory of Advanced Materials, Department of Chemistry, and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China Linping Qian − Laboratory of Advanced Materials, Department of Chemistry, and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China; orcid.org/00000003-3412-5453 Abdullah M. Al-Enizi − Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia; orcid.org/0000-0002-3967-5553 Lijuan Zhang − Laboratory of Advanced Materials, Department of Chemistry, and Shanghai Key Laboratory of Molecular Catalysis and Innovative Materials, Fudan University, Shanghai 200438, China Complete contact information is available at: https://pubs.acs.org/10.1021/acsaem.0c02648 Author Contributions The manuscript was written through contributions of all authors. All authors have given approval to the final version of the manuscript. The authors declare no competing financial interest. ■ ACKNOWLEDGMENTS We thank the following funding agencies for supporting this work: the National Key Research and Development Program of China (2017YFA0206901, 2018YFA0209401), the National Science Foundation of China (22025502, 21773036, 21975051), the Science and Technology Commission of Shanghai Municipality (19XD1420400), and the Innovation Program of Shanghai Municipal Education Commission (2019-01-07-00-07-E00045). The authors also extend their 1041 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Reduction via Field-Induced Reagent Concentration. Nature 2016, 537, 382−386. (19) Resasco, J.; Chen, L. D.; Clark, E.; Tsai, C.; Hahn, C.; Jaramillo, T. F.; Chan, K.; Bell, A. T. Promoter Effects of Alkali Metal Cations on the Electrochemical Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2017, 139, 11277−11287. (20) Wakerley, D.; Lamaison, S.; Ozanam, F.; Menguy, N.; Mercier, D.; Marcus, P.; Fontecave, M.; Mougel, V. Bio-Inspired Hydrophobicity Promotes CO2 Reduction on a Cu Surface. Nat. Mater. 2019, 18, 1222−1227. (21) Wei, X.; Yin, Z.; Lyu, K.; Li, Z.; Gong, J.; Wang, G.; Xiao, L.; Lu, J.; Zhuang, L. Highly Selective Reduction of CO2 to C2+ Hydrocarbons at Copper/Polyaniline Interfaces. ACS Catal. 2020, 10, 4103−4111. (22) Nitopi, S.; Bertheussen, E.; Scott, S. B.; Liu, X.; Engstfeld, A. K.; Horch, S.; Seger, B.; Stephens, I. E. L.; Chan, K.; Hahn, C.; Norskov, J. K.; Jaramillo, T. F.; Chorkendorff, I. Progress and Perspectives of Electrochemical CO2 Reduction on Copper in Aqueous Electrolyte. Chem. Rev. 2019, 119, 7610−7672. (23) Seifitokaldani, A.; Gabardo, C. M.; Burdyny, T.; Dinh, C. T.; Edwards, J. P.; Kibria, M. G.; Bushuyev, O. S.; Kelley, S. O.; Sinton, D.; Sargent, E. H. Hydronium-Induced Switching between CO2 Electroreduction Pathways. J. Am. Chem. Soc. 2018, 140, 3833−3837. (24) Ooka, H.; Figueiredo, M. C.; Koper, M. T. M. Competition between Hydrogen Evolution and Carbon Dioxide Reduction on Copper Electrodes in Mildly Acidic Media. Langmuir 2017, 33, 9307−9313. (25) Lum, Y.; Cheng, T.; Goddard, W. A., 3rd; Ager, J. W. Electrochemical CO Reduction Builds Solvent Water into Oxygenate Products. J. Am. Chem. Soc. 2018, 140, 9337−9340. (26) Cheng, T.; Xiao, H.; Goddard, W. A., 3rd Reaction Mechanisms for the Electrochemical Reduction of CO2 to CO and Formate on the Cu(100) Surface at 298 K from Quantum Mechanics Free Energy Calculations with Explicit Water. J. Am. Chem. Soc. 2016, 138, 13802−13805. (27) White, J. L.; Baruch, M. F.; Pander, J. E.; Hu, Y.; Fortmeyer, I. C.; Park, J. E.; Zhang, T.; Liao, K.; Gu, J.; Yan, Y.; Shaw, T. W.; Abelev, E.; Bocarsly, A. B. Light-Driven Heterogeneous Reduction of Carbon Dioxide: Photocatalysts and Photoelectrodes. Chem. Rev. 2015, 115, 12888−12935. (28) Zhuang, T.-T.; Liang, Z.-Q.; Seifitokaldani, A.; Li, Y.; De Luna, P.; Burdyny, T.; Che, F.; Meng, F.; Min, Y.; Quintero-Bermudez, R.; Dinh, C. T.; Pang, Y.; Zhong, M.; Zhang, B.; Li, J.; Chen, P.-N.; Zheng, X.-L.; Liang, H.; Ge, W.-N.; Ye, B.-J.; Sinton, D.; Yu, S.-H.; Sargent, E. H. Steering Post C−C Coupling Selectivity Enables High Efficiency Electroreduction of Carbon Dioxide to Multi-Carbon Alcohols. Nat. Catal. 2018, 1, 421−428. (29) Luc, W.; Fu, X.; Shi, J.; Lv, J.-J.; Jouny, M.; Ko, B. H.; Xu, Y.; Tu, Q.; Hu, X.; Wu, J.; Yue, Q.; Liu, Y.; Jiao, F.; Kang, Y. TwoDimensional Copper Nanosheets for Electrochemical Reduction of Carbon Monoxide to Acetate. Nat. Catal. 2019, 2, 423−430. (30) Tian, Z.; Priest, C.; Chen, L. Recent Progress in the Theoretical Investigation of Electrocatalytic Reduction of CO2. Adv. Theory Simul. 2018, 1, 1800004. (31) Xiao, H.; Cheng, T.; Goddard, W. A., III Atomistic Mechanisms Underlying Selectivities in C(1) and C(2) Products from Electrochemical Reduction of CO on Cu(111). J. Am. Chem. Soc. 2017, 139, 130−136. (32) 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, 922−934. (33) 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, 5022−5023. (34) De Gregorio, G. L.; Burdyny, T.; Loiudice, A.; Iyengar, P.; Smith, W. A.; Buonsanti, R. Facet-Dependent Selectivity of Cu appreciation to the Researchers Supporting Program (RSP2020/55) at King Saud University, Riyadh, Saudi Arabia, for partial funding of this work. ■ Spotlight on Applications REFERENCES (1) Sanz-Perez, E. S.; Murdock, C. R.; Didas, S. A.; Jones, C. W. Direct Capture of CO2 from Ambient Air. Chem. Rev. 2016, 116, 11840−11876. (2) Hepburn, C.; Adlen, E.; Beddington, J.; Carter, E. A.; Fuss, S.; Mac Dowell, N.; Minx, J. C.; Smith, P.; Williams, C. K. The Technological and Economic Prospects for CO2 Utilization and Removal. Nature 2019, 575, 87−97. (3) Verma, S.; Lu, S.; Kenis, P. J. A. Co-Electrolysis of CO2 and Glycerol as a Pathway to Carbon Chemicals with Improved Technoeconomics Due to Low Electricity Consumption. Nature Energy 2019, 4, 466−474. (4) Sriramagiri, G. M.; Ahmed, N.; Luc, W.; Dobson, K. D.; Hegedus, S. S.; Jiao, F. Toward a Practical Solar-Driven CO2 Flow Cell Electrolyzer: Design and Optimization. ACS Sustainable Chem. Eng. 2017, 5, 10959−10966. (5) Jouny, M.; Luc, W.; Jiao, F. General Techno-Economic Analysis of CO2 Electrolysis Systems. Ind. Eng. Chem. Res. 2018, 57, 2165− 2177. (6) Chen, A.; Lin, B.-L. A Simple Framework for Quantifying Electrochemical CO2 Fixation. Joule 2018, 2, 594−606. (7) Gabardo, C. M.; O’Brien, C. P.; Edwards, J. P.; McCallum, C.; Xu, Y.; Dinh, C.-T.; Li, J.; Sargent, E. H.; Sinton, D. Continuous Carbon Dioxide Electroreduction to Concentrated Multi-Carbon Products Using a Membrane Electrode Assembly. Joule 2019, 3, 2777−2791. (8) Li, J.; Chen, G.; Zhu, Y.; Liang, Z.; Pei, A.; Wu, C.-L.; Wang, H.; Lee, H. R.; Liu, K.; Chu, S.; Cui, Y. Efficient Electrocatalytic CO2 Reduction on a Three-Phase Interface. Nat. Catal. 2018, 1, 592−600. (9) Song, Y.; Zhang, X.; Xie, K.; Wang, G.; Bao, X. HighTemperature CO2 Electrolysis in Solid Oxide Electrolysis Cells: Developments, Challenges, and Prospects. Adv. Mater. 2019, 31, 1902033. (10) Burdyny, T.; Smith, W. A. CO2 Reduction on Gas-Diffusion Electrodes and Why Catalytic Performance Must Be Assessed at Commercially-Relevant Conditions. Energy Environ. Sci. 2019, 12, 1442−1453. (11) Jouny, M.; Hutchings, G. S.; Jiao, F. Carbon Monoxide Electroreduction as an Emerging Platform for Carbon Utilization. Nat. Catal. 2019, 2, 1062−1070. (12) Schreier, M.; Yoon, Y.; Jackson, M. N.; Surendranath, Y. Competition between H and CO for Active Sites Governs CopperMediated Electrosynthesis of Hydrocarbon Fuels. Angew. Chem., Int. Ed. 2018, 57, 10221−10225. (13) Schouten, K. J. P.; Pérez Gallent, E.; Koper, M. T. M. The Influence of pH on the Reduction of CO and CO2 to Hydrocarbons on Copper Electrodes. J. Electroanal. Chem. 2014, 716, 53−57. (14) Billy, J. T.; Co, A. C. Experimental Parameters Influencing Hydrocarbon Selectivity During the Electrochemical Conversion of CO2. ACS Catal. 2017, 7, 8467−8479. (15) Zhu, S.; Jiang, B.; Cai, W. B.; Shao, M. Direct Observation on Reaction Intermediates and the Role of Bicarbonate Anions in CO2 Electrochemical Reduction Reaction on Cu Surfaces. J. Am. Chem. Soc. 2017, 139, 15664−15667. (16) Wuttig, A.; Yoon, Y.; Ryu, J.; Surendranath, Y. Bicarbonate Is Not a General Acid in Au-Catalyzed CO2 Electroreduction. J. Am. Chem. Soc. 2017, 139, 17109−17113. (17) Chen, L. D.; Urushihara, M.; Chan, K.; Nørskov, J. K. Electric Field Effects in Electrochemical CO2 Reduction. ACS Catal. 2016, 6, 7133−7139. (18) Liu, M.; Pang, Y.; Zhang, B.; De Luna, P.; Voznyy, O.; Xu, J.; Zheng, X.; Dinh, C. T.; Fan, F.; Cao, C.; de Arquer, F. P.; Safaei, T. S.; Mepham, A.; Klinkova, A.; Kumacheva, E.; Filleter, T.; Sinton, D.; Kelley, S. O.; Sargent, E. H. Enhanced Electrocatalytic CO 2 1042 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Catalysts in Electrochemical CO2 Reduction at Commercially Viable Current Densities. ACS Catal. 2020, 10, 4854−4862. (35) Lum, Y.; Kwon, Y.; Lobaccaro, P.; Chen, L.; Clark, E. L.; Bell, A. T.; Ager, J. W. Trace Levels of Copper in Carbon Materials Show Significant Electrochemical CO2 Reduction Activity. ACS Catal. 2016, 6, 202−209. (36) Seh, Z. W.; Kibsgaard, J.; Dickens, C. F.; Chorkendorff, I.; Norskov, J. K.; Jaramillo, T. F. Combining Theory and Experiment in Electrocatalysis: Insights into Materials Design. Science 2017, 355, eaad4998. (37) Wang, Y.; Cui, X.; Zhang, Y.; Zhang, L.; Gong, X.; Zheng, G. Achieving High Aqueous Energy Storage Via Hydrogen-Generation Passivation. Adv. Mater. 2016, 28, 7626−32. (38) Li, T.; Yang, C.; Luo, J.-L.; Zheng, G. Electrolyte Driven Highly Selective CO2 Electroreduction at Low Overpotentials. ACS Catal. 2019, 9, 10440−10447. (39) Cui, X.; Pan, Z.; Zhang, L.; Peng, H.; Zheng, G. Selective Etching of Nitrogen-Doped Carbon by Steam for Enhanced Electrochemical CO2 Reduction. Adv. Energy Mater. 2017, 7, 1701456. (40) Kuang, M.; Guan, A.; Gu, Z.; Han, P.; Qian, L.; Zheng, G. Enhanced N-Doping in Mesoporous Carbon for Efficient Electrocatalytic CO2 Conversion. Nano Res. 2019, 12, 2324−2329. (41) Han, P.; Wang, Z.; Kuang, M.; Wang, Y.; Liu, J.; Hu, L.; Qian, L.; Zheng, G. 2D Assembly of Confined Space toward Enhanced CO2 Electroreduction. Adv. Energy Mater. 2018, 8, 1801230. (42) Wang, Y.; Han, P.; Lv, X.; Zhang, L.; Zheng, G. Defect and Interface Engineering for Aqueous Electrocatalytic CO2 Reduction. Joule 2018, 2, 2551−2582. (43) Gu, Z.; Shen, H.; Shang, L.; Lv, X.; Qian, L.; Zheng, G. Nanostructured Copper-based Electrocatalysts for CO2 Reduction. Small Methods 2018, 2, 1800121. (44) Guan, A.; Yang, C.; Quan, Y.; Shen, H.; Cao, N.; Li, T.; Ji, Y.; Zheng, G. One-Dimensional Nanomaterial Electrocatalysts for CO2 Fixation. Chem. - Asian J. 2019, 14, 3969−3980. (45) Guan, A.; Chen, Z.; Quan, Y.; Peng, C.; Wang, Z.; Sham, T.-K.; Yang, C.; Ji, Y.; Qian, L.; Xu, X.; Zheng, G. Boosting CO2 Electroreduction to CH4 Via Tuning Neighboring Single-Copper Sites. ACS Energy Lett. 2020, 5, 1044−1053. (46) Yang, C.; Shen, H.; Guan, A.; Liu, J.; Li, T.; Ji, Y.; Al-Enizi, A. M.; Zhang, L.; Qian, L.; Zheng, G. Fast Cooling Induced GrainBoundary-Rich Copper Oxide for Electrocatalytic Carbon Dioxide Reduction to Ethanol. J. Colloid Interface Sci. 2020, 570, 375−381. (47) Gu, Z.; Yang, N.; Han, P.; Kuang, M.; Mei, B.; Jiang, Z.; Zhong, J.; Li, L.; Zheng, G. Oxygen Vacancy Tuning toward Efficient Electrocatalytic CO2 Reduction to C2H4. Small Methods 2018, 3, 1800449. (48) Shen, H.; Gu, Z.; Zheng, G. Pushing the Activity of CO2 Electroreduction by System Engineering. Sci. Bull. 2019, 64, 1805− 1816. (49) Wang, Y.; Liu, J.; Wang, Y.; Wang, Y.; Zheng, G. Efficient SolarDriven Electrocatalytic CO2 Reduction in a Redox-Medium-Assisted System. Nat. Commun. 2018, 9, 5003. (50) Zhu, W.; Michalsky, R.; Metin, O.; Lv, H.; Guo, S.; Wright, C. J.; Sun, X.; Peterson, A. A.; Sun, S. Monodisperse Au Nanoparticles for Selective Electrocatalytic Reduction of CO2 to CO. J. Am. Chem. Soc. 2013, 135, 16833−6. (51) Dinh, C. T.; Burdyny, T.; Kibria, M. G.; Seifitokaldani, A.; Gabardo, C. M.; Garcia de Arquer, F. P.; Kiani, A.; Edwards, J. P.; De Luna, P.; Bushuyev, O. S.; Zou, C.; Quintero-Bermudez, R.; Pang, Y.; Sinton, D.; Sargent, E. H. CO2 Electroreduction to Ethylene via Hydroxide-Mediated Copper Catalysis at An Abrupt Interface. Science 2018, 360, 783−787. (52) Clark, E. L.; Hahn, C.; Jaramillo, T. F.; Bell, A. T. Electrochemical CO2 Reduction over Compressively Strained Cuag Surface Alloys with Enhanced Multi-Carbon Oxygenate Selectivity. J. Am. Chem. Soc. 2017, 139, 15848−15857. Spotlight on Applications (53) Hall, A. S.; Yoon, Y.; Wuttig, A.; Surendranath, Y. Mesostructure-Induced Selectivity in CO2 Reduction Catalysis. J. Am. Chem. Soc. 2015, 137, 14834−14837. (54) Zhang, L.; Qin, X.; Zhao, S.; Wang, A.; Luo, J.; Wang, Z. L.; Kang, F.; Lin, Z.; Li, B. Advanced Matrixes for Binder-Free Nanostructured Electrodes in Lithium-Ion Batteries. Adv. Mater. 2020, 32, 1908445. (55) Duan, X.; Xu, J.; Wei, Z.; Ma, J.; Guo, S.; Wang, S.; Liu, H.; Dou, S. Metal-Free Carbon Materials for CO2 Electrochemical Reduction. Adv. Mater. 2017, 29, 1701784. (56) Lee, W. J.; Lim, J.; Kim, S. O. Nitrogen Dopants in Carbon Nanomaterials: Defects or a New Opportunity? Small Methods 2017, 1, 1600014. (57) Thome, I.; Nijs, A.; Bolm, C. Trace Metal Impurities in Catalysis. Chem. Soc. Rev. 2012, 41, 979−87. (58) Wuttig, A.; Surendranath, Y. Impurity Ion Complexation Enhances Carbon Dioxide Reduction Catalysis. ACS Catal. 2015, 5, 4479−4484. (59) Vasileff, A.; Xu, C.; Jiao, Y.; Zheng, Y.; Qiao, S.-Z. Surface and Interface Engineering in Copper-Based Bimetallic Materials for Selective CO2 Electroreduction. Chem. 2018, 4, 1809−1831. (60) Zhou, Y.; Che, F.; Liu, M.; Zou, C.; Liang, Z.; De Luna, P.; Yuan, H.; Li, J.; Wang, Z.; Xie, H.; Li, H.; Chen, P.; Bladt, E.; Quintero-Bermudez, R.; Sham, T. K.; Bals, S.; Hofkens, J.; Sinton, D.; Chen, G.; Sargent, E. H. Dopant-Induced Electron Localization Drives CO2 Reduction to C2 Hydrocarbons. Nat. Chem. 2018, 10, 974−980. (61) Huang, Y.; Handoko, A. D.; Hirunsit, P.; Yeo, B. S. Electrochemical Reduction of CO2 Using Copper Single-Crystal Surfaces: Effects of CO* Coverage on the Selective Formation of Ethylene. ACS Catal. 2017, 7, 1749−1756. (62) Gao, D.; Zhang, Y.; Zhou, Z.; Cai, F.; Zhao, X.; Huang, W.; Li, Y.; Zhu, J.; Liu, P.; Yang, F.; Wang, G.; Bao, X. Enhancing CO2 Electroreduction with the Metal-Oxide Interface. J. Am. Chem. Soc. 2017, 139, 5652−5655. (63) Wang, Y.; Chen, Z.; Han, P.; Du, Y.; Gu, Z.; Xu, X.; Zheng, G. Single-Atomic Cu with Multiple Oxygen Vacancies on Ceria for Electrocatalytic CO2 Reduction to CH4. ACS Catal. 2018, 8, 7113− 7119. (64) Wang, Y.; Wang, Z.; Dinh, C.-T.; Li, J.; Ozden, A.; Golam Kibria, M.; Seifitokaldani, A.; Tan, C.-S.; Gabardo, C. M.; Luo, M.; Zhou, H.; Li, F.; Lum, Y.; McCallum, C.; Xu, Y.; Liu, M.; Proppe, A.; Johnston, A.; Todorovic, P.; Zhuang, T.-T.; Sinton, D.; Kelley, S. O.; Sargent, E. H. Catalyst Synthesis under CO2 Electroreduction Favours Faceting and Promotes Renewable Fuels Electrosynthesis. Nat. Catal. 2020, 3, 98−106. (65) Gu, Z.; Shen, H.; Chen, Z.; Yang, Y.; Yang, C.; Ji, Y.; Wang, Y.; Zhu, C.; Liu, J.; Li, J.; Sham, T.-K.; Xu, X.; Zheng, G. Efficient Electrocatalytic CO2 Reduction to C2+ Alcohols at Defect-Site-Rich Cu Surface. Joule 2021, DOI: 10.1016/j.joule.2020.12.011. (66) Luo, M.; Guo, S. Strain-Controlled Electrocatalysis on Multimetallic Nanomaterials. Nat. Rev. Mater. 2017, 2, 17059. (67) Wang, Y.; Wang, D.; Dares, C. J.; Marquard, S. L.; Sheridan, M. V.; Meyer, T. J. CO2 Reduction to Acetate in Mixtures of Ultrasmall (Cu)n,(Ag)m Bimetallic Nanoparticles. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 278−283. (68) Chen, C.; Li, Y.; Yu, S.; Louisia, S.; Jin, J.; Li, M.; Ross, M. B.; Yang, P. Cu-Ag Tandem Catalysts for High-Rate CO2 Electrolysis toward Multicarbons. Joule 2020, 4, 1688−1699. (69) Lee, S.; Park, G.; Lee, J. Importance of Ag−Cu Biphasic Boundaries for Selective Electrochemical Reduction of CO2 to Ethanol. ACS Catal. 2017, 7, 8594−8604. (70) Lv, X.; Shang, L.; Zhou, S.; Li, S.; Wang, Y.; Wang, Z.; Sham, T. K.; Peng, C.; Zheng, G. Electron-Deficient Cu Sites on Cu3Ag1 Catalyst Promoting CO2 Electroreduction to Alcohols. Adv. Energy Mater. 2020, 10, 2001987. (71) Schreier, M.; Héroguel, F.; Steier, L.; Ahmad, S.; Luterbacher, J. S.; Mayer, M. T.; Luo, J.; Grätzel, M. Solar Conversion of CO2 to CO 1043 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044 ACS Applied Energy Materials www.acsaem.org Spotlight on Applications Using Earth-Abundant Electrocatalysts Prepared by Atomic Layer Modification of CuO. Nat. Energy 2017, 2, 17087. 1044 https://dx.doi.org/10.1021/acsaem.0c02648 ACS Appl. Energy Mater. 2021, 4, 1034−1044