Next Article in Journal
Photodynamic Light-Triggered Release of Curcumin from Hierarchical FAU Zeolite
Next Article in Special Issue
Electrolessly Deposited Carbon-Supported CuNiSn Electrocatalysts for the Electrochemical Reduction of CO2
Previous Article in Journal
A Review on the Green Synthesis of Benzimidazole Derivatives and Their Pharmacological Activities
Previous Article in Special Issue
Electrocatalytic Reduction of CO2 to C1 Compounds by Zn-Based Monatomic Alloys: A DFT Calculation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Progress in Surface-Defect Engineering Strategies for Electrocatalysts toward Electrochemical CO2 Reduction: A Review

by
Sridharan Balu
1,2,
Abdul Hanan
3,
Harikrishnan Venkatesvaran
1,
Shih-Wen Chen
1,2,
Thomas C.-K. Yang
1,2,* and
Mohammad Khalid
3,4,*
1
Department of Chemical Engineering and Biotechnology, National Taipei University of Technology, Taipei 10608, Taiwan
2
Precision Analysis and Materials Research Center, National Taipei University of Technology, Taipei 10608, Taiwan
3
Graphene & Advanced 2D Materials Research Group (GAMRG), School of Engineering and Technology, Sunway University, No. 5, Jalan Universiti, Bandar Sunway, Petaling Jaya 47500, Malaysia
4
School of Applied and Life Sciences, Uttaranchal University, Dehradun 248007, Uttarakhand, India
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(2), 393; https://doi.org/10.3390/catal13020393
Submission received: 20 December 2022 / Revised: 6 February 2023 / Accepted: 7 February 2023 / Published: 11 February 2023
(This article belongs to the Special Issue Heterogeneous Electrocatalysts for CO2 Reduction)

Abstract

:
Climate change, caused by greenhouse gas emissions, is one of the biggest threats to the world. As per the IEA report of 2021, global CO2 emissions amounted to around 31.5 Gt, which increased the atmospheric concentration of CO2 up to 412.5 ppm. Thus, there is an imperative demand for the development of new technologies to convert CO2 into value-added feedstock products such as alcohols, hydrocarbons, carbon monoxide, chemicals, and clean fuels. The intrinsic properties of the catalytic materials are the main factors influencing the efficiency of electrochemical CO2 reduction (CO2-RR) reactions. Additionally, the electroreduction of CO2 is mainly affected by poor selectivity and large overpotential requirements. However, these issues can be overcome by modifying heterogeneous electrocatalysts to control their morphology, size, crystal facets, grain boundaries, and surface defects/vacancies. This article reviews the recent progress in electrochemical CO2 reduction reactions accomplished by surface-defective electrocatalysts and identifies significant research gaps for designing highly efficient electrocatalytic materials.

Graphical Abstract

1. Introduction

The continuous combustion of fossil fuels for energy production has led to the release of extreme concentrations of carbon dioxide (CO2) into the atmosphere. Furthermore, CO2 traps solar radiation within the atmospheric surface of the Earth (i.e., the greenhouse effect) and causes severe climate alterations (i.e., global warming). According to the 2022 United Nations Environment Programme (UNEP) report, the global temperature will be raised by 1.8 °C by the end of this century due to the greenhouse effect. According to the latest report (November 2022), the global average CO2 level has reached a new record of 417.3 ppm, which is about 1 ppm (0.24%) higher than the previous year.
The conversion of abundant CO2 into chemical and fuel energy using renewable energy sources is considered a promising alternative method for utilizing and converting renewables and environmental protection [1,2]. Several conversion techniques, including thermo-chemical, photo-chemical, electrochemical, and biological methods, have been used to convert CO2 into valuable feedstock chemicals. Among them, the heterogeneous electrocatalytic/electrochemical CO2 reduction reaction (EC-CO2-RR) has attracted keen interest owing to its enormous advantages, as it leads to the direct conversion of fuels and valuable chemical products [3,4]. Figure 1A presents a simplified pictorial representation of the anthropogenic carbon cycle. The CO2 emissions are controlled by capturing and converting CO2 gas by the abovementioned technical pathways. This efficient carbon cycling strategy solves energy and environmental issues and creates a highly profitable carbon economy platform [5]. In the presence of intermittent renewable electricity and water (proton donor), the oxygen evolution reaction (OER) occurs at the anode compartment, and the EC-CO2-RR takes place at the cathode compartment, converting CO2 into the various carbon products, such as carbon monoxide (CO), formic acid (HCOOH), methane (CH4), methanol (CH3OH), ethanol (CH3CH2OH), acetic acid (CH3COOH), and ethylene/olefins (CH2=CH2) (Figure 1B). Owing to these numerous advantages, the EC-CO2-RR has received keen attention from researchers. Among the different products, the EC-CO2-RR produces a large amount of gaseous CO due to its low overpotential of formation or multi-electron transfer reaction kinetics (i.e., two-electron transfer process) compared to that of other hydrocarbon products (i.e., multi-electron processes). However, the EC-CO2-RR has some practical downsides/challenges, such as the high overpotential requirements for the multi-electron transfer reaction, low current density, product selectivity, and poor Faradaic efficiency (FE). Furthermore, the current densities are still limited to <30 mA/cm2 owing to the longer diffusion length and low solubility of CO2(g) in aqueous/electrolyte media. Due to the high stability of the C=O bond (750 kJ/mol), the complexity of the charge transfer and the interference of HER overpotentials cause sluggish EC-CO2-RR kinetics and a low FE in the desired product, respectively.
Several efforts have been made to enhance electrochemical CO2 reduction and large-scale viability, including the design of catalysts and reactors, electrode modification, and electrolyte optimization, followed by the pioneering research conducted by Hori and his co-workers [7]. Much work is needed to develop novel technologies for efficiently capturing and converting CO2 into value-added chemicals. Thus, this mini-review discusses the main practical problems, challenges, and mechanistic pathways of the EC-CO2-RR and presents a clear overview. In particular, the recent electrocatalyst design processes/modifications via surface-defect engineering approaches in significant metallic, metal oxide/sulfide, and carbon-based alloy catalysts are discussed. Furthermore, this review presents several electrocatalyst design techniques for attaining higher current densities, low overpotentials, and higher Faradaic efficiencies, with the aim of obtaining the desired value-added products through the electrochemical CO2-RR.

2. Electrocatalytic CO2 Reduction Reaction (CO2-RR) Mechanism on Electrode Surface

Many research groups have published exciting results with novel electrocatalytic materials and are trying to understand the exact reaction mechanism of electrochemical CO2 reduction and the catalytic performance of various metal/metal-mediated catalysts. According to the thermodynamic prospects, the electrochemical reduction of CO2 into mono-carbon (C1) and multi-carbon (C2) products is generally a multi-electron transfer process associated with a proton-coupling reaction [8]. Furthermore, CO2 electroreduction depends on the electrocatalytic materials, electrolytes, pH, and related experimental conditions. Regardless of the carbon products formed, the CO2-RR is generally initiated by a single-electron injection to form unstable anionic CO2 radicals. This entails a high overpotential, and, more importantly, the designated reaction pathways (i.e., CO or formate) are determined by the coordination/adsorption of CO2 radicals on the electrode surface [9,10], since CO2 has a high energy (−1.9 V vs. SHE) and readily reacts with adsorbed water molecules, including other CO2 molecules in the solution, producing CO, formate, and other reduction products. The standard redox potentials (vs. SHE) of electrochemical CO2 reduction reactions for the formation of various products are shown in Figure 2B [11].
The transition-metal electrocatalysts are categorized into four groups according to their binding tendency (available orbitals) and active d-orbital electrons for binding with CO2 intermediate species (Figure 2A). The first group of metals (i.e., Ni, Pt, Fe, and Ti) are less active towards CO2-RR but mostly favor hydrogen (H2) generation due to their strong hydrogen binding tendency. The second group consists of metals such as Zn, Au, Ag, and Pd. These metals can bind with CO2•− intermediates to different degrees, producing carbon monoxide (CO) as a significant product, and are not able to reduce further owing to their lower binding affinity for surface CO. The group-three metals, including In, Sn, Pb, Hg, Tl, Cd, and Bi, can bind with intermediates and generate formic acid/formate species because of the greater negative formation potential of HCOO and the weak adsorption of the intermediates on the surface [12,13]. The fourth group consists of copper (Cu), which has been widely used as a CO2 electrocatalyst to produce various (C1) hydrocarbons and alcohol products owing to its excellent binding tendency with intermediates, including CO2 and CO [9,10,14]. In general, the heterogeneous electrocatalytic CO2-RR involves the following three significant steps: (i) the chemisorption of CO2 on the electrode surface, (ii) the electron injection process to cleave C–O bonds and the proton transfer reaction to form C–H bonds (hydrocarbons), and (iii) the desorption of the final carbon products from the electrode surface and their diffusion into the electrolyte medium.
Consequently, forming C2 reduction products is tedious compared to forming C1 products such as CO and formate, since it is a multi-electron transfer process. Additionally, in the formation of C2 hydrocarbons, such as ethyl alcohol and ethylene, the competing C–C and C–H bonds during the electrochemical CO2-RR depend on the electrolyte conditions and applied potential [15]. In the CO2-RR, the C–C bond formation is considered the rate-determining step (RDS) for the generation of C2 species [16]. Thus, the formation of C2 reduction species can be classified into the ethylene and ethanol pathways (Figure 2C). The ethylene pathway starts with the dimerization of two surface-adsorbed CO intermediates, generating a C2O2 intermediate in a bridging mode. After that, the C2O2 intermediate undergoes protonation reactions (*CO + H = *CHO; *CHO + CO = *COCHO; *COCO + H = *COCOH) and thus produces stable CO–COH and COCHO intermediates. Previous studies have stated that the formation of C–C through the two surface-adsorbed CO intermediates occurs at a low overpotential. At the same time, the reaction between the surface-bound CO and *CHO intermediate requires a high overpotential due to the enormous activation barrier to form a CO dimer. Furthermore, the subsequent reduction of CO to *COCHO, followed by CO to *CHO conversion, is more favorable than the CO dimerization reaction [17]. When comparing the tautomeric *COCOH and *COCHO structures, *COCHO is more stable and preferred due to its lower involvement in generating double bonds on the surface or forming free radicals on the carbon atom. Therefore, the *COCHO intermediate can be protonated or reduced into COCH2OH and undergoes electron injection and the proton transfer reaction to produce the final H2C=CH2 (ethylene) products.
On the other hand, the ethanol pathway starts with the intermediates (i.e., *COCHO and *COCOH), resulting in the ethylene pathway. However, forming *COCHO intermediates is considered a pivotal step in the selectivity of the C2 reduction products on the Cu surface. C–C bond formation is significant for generating C2H5OH and other C2+ reduction products [17]. The *COCHO further undergoes a reduction/protonation reaction for ethanol production and generates *C2H2O2 (cis-glyoxal). At this stage, the consumption of glyoxal is thermodynamically more favorable owing to the stable oxygen bonds of cis-glyoxal on the catalyst surface. Then, the cis-glyoxal is further reduced to form acetaldehyde (at a low potential) and ethanol (at a high potential). The cis-glyoxal to ethanol conversion consists in the formation of several intermediates, such as C2H6O2 (ethylene glycol), C2H4O (vinyl alcohol), and CH3CHO (acetaldehyde), in which the vinyl alcohol is quickly reduced to generate ethanol rather than that of the ethylene glycol owing to its negative –OH groups and sp2-hybridized oxygen atoms bound with the copper surface [18]. Moreover, a recent study conducted by Varandili et al. presented the concept of forming C2 products over a CuZn catalytic system for electrochemical CO2 reduction reactions. The as-reported catalysts, Cu-derived ZnO@Cu and ZnO-derived Cu@ZnO, showed selectivity towards ethanol and methane, respectively. The higher concentration of Zn immensely increased the concentration of CO and shifted the selectivity from CH4 to C2H5OH through a tandem conversion mechanism [19]. Additionally, Iyengar et al. established a high ethanol selectivity over a Cu–Ag tandem electrocatalyst. Their results suggested that the *CHx–*CO coupling pathway drastically enhanced the production of C2H5OH. This indicated that under a high CO concentration, the bifurcation reactions through *CHx–*CO coupling occurred prior to the –CO dimerization reaction on Cu(111), leading to ethanol production over ethylene [20]. The overall electrochemical process involving the reactor cell is expressed below (Equations (1) and (2)) [21]:
At   anode : H 2 O electrons ( e ) O 2 + H +
At   cathode : CO 2 + electrons ( e ) + H + Various   reduction   products   ( C 1 , C 2 )

2.1. Advantages and Economic Benefits of Electrochemical CO2 Conversion

Electrochemical CO2 reduction (CO2R) is considered one of the most effective methods of reducing CO2 emissions [22,23]. Electrochemical methods work in moderate conditions and may be tailored to specific products, enable modular systems, and allow the combination of renewable energy and CO2 conversion in energy-intensive industrial sectors, including iron and steel production [24]. The corresponding advantages of the electroreduction of CO2 for society are illustrated in Figure 3A. Previous research has largely found short-term, local, and provincial advantages for lowering greenhouse gas emissions, while global warming is a long-term, worldwide problem [25]. Because the advantages of lowering emissions will not be realized in their entirety for 50 to 100 years, the world has been sluggish in addressing climate change [26]. The CO2 electrochemical reduction reaction (CO2-RR) converts the greenhouse gas CO2 into value-added goods, reducing carbon emissions and the use of fossil fuels and feedstocks to manufacture fuel/chemical products [27]. This offers the possibility of combining electrocatalytic procedures using renewable energy to store renewable energy. The electrochemical reduction of CO2 seems to be a way to contribute to a less-carbon-based economy by providing value-added compounds from CO2 whilst employing minimal renewable energy [28]. Renewable electricity costs have dropped dramatically in recent years, so that it is now competitive with conventional power generation methods and may shortly be the lower-priced source of power [29]. By employing particular transition metals for electrocatalysts, synthetic chemicals may be produced from the electrochemical reduction of CO2. The products generated depend on the metal’s potential to absorb CO, the key step in the CO2 reduction cycle [30]. In this context, employing aqueous electrolytes and CO2 conversion to fuels and feedstocks using renewable energy is a highly appealing alternative to closing the carbon cycle [31]. The liquid electrolyte–gas diffusion electrode system also has several practical drawbacks, related to system stability and energy efficiency. The membrane electrode assembly (MEA)-based electrolyzer system for the EC-CO2-RR was recently introduced to overcome these issues. Moreover, the enhanced performance of the MEA system is associated with eliminating ohmic losses in the electrolyte, the fouling of electrolyte impurities with catalysts, and flooding through gas diffusion electrodes. Thus, the combined hybrid system strategies enhance the production of gaseous and liquid products with long-term (>100 h) uninterrupted operation stability. Several recent studies have reported the MEA-based electrolyzer system’s enhanced performance resulting in ~80% and ~50% FE for C2+ and ethylene products, respectively, with current density and operation stabilities over 100 h [32,33,34].
Conversely, due to its low cost and environmental compatibility, heterogeneous electrocatalysis has recently attracted great interest. In the past decade, these electrocatalysts have been widely explored with a combination of three strategies: the first is the electroreduction of CO2 to syngas, a combination of CO and H2, employing Ag and Au electrodes, which approximates closely to the real-world applications [35,36]; the second approach is to convert CO2 to formate on Sn electrodes with Faradaic efficiencies (FE) of more than 70% at current densities ranging from 100 mA/cm2 to 300 mA/cm2 [37]; and the third and most challenging approach is to reduce CO2 to C2 and C3 hydrocarbons such as ethylene, ethanol, and n-propanol. Among the reduction products, formate (formic acid) is widely used by several industries, including in the production of medicine, dyes, rubber, leather, and pesticides. Nevertheless, the industrial manufacturing of formate involves tedious methanol carbonylation and hydrolysis processes, which are costly processes, higher energy demanding, and yielded with unfavorable intermediates [38]. Likewise, the production cost of CO2-RR products such as CO, CH3OH, CH2=CH2, C2H5OH, CH4, and HCOOH is more viable than commercial and industrial manufacturing and operational costs (Figure 3B).
The commercial and economic viability of CO2-RR products (kg) is easily calculated, since the production is majorly depended on the input of electricity (kWh−1). Furthermore, life-cycle assessment (LCA) is considered an important evaluation method to quantify the environmental impacts and performance of a process or technology. A recent article published by Ong et al. described a prospective LCA of the EC–CO2-RR for the production of selective formic acid and ethylene products based on a standard laboratory-scale H-cell reactor design. The analysis found that the average production of formic acid and ethylene by the electrochemical CO2 reduction method (ECR) was around 57% and 52%, respectively. In addition, the formic acid production via the ECR process reduced the consumption of petrochemical resources by 24% and the ethylene production via the ECR processes significantly decreased the human health damage (67%), petrochemical consumption (110%), and ecosystem damage (94%) [39]. Khoo et al. presented a unique LCA case study assessing the potential environmental impacts of small-scale and large-scale models for the electrochemical reduction of CO2 to ethylene. The authors conducted an LCA for the CO2-RR with 1 g ethylene production for the small-scale experimental setup and 1 ton ethylene production for the large-scale industry-standard experiment with a similar CO2 capture and separation set up. Producing 0.98~3.7 g CO2 on a small scale was equivalent to producing 0.65~3.0 tons of CO2 in the large-scale model [40]. Formic acid consumed more electrical energy than the other C2–C4 products. Currently, the selectivity of the maximum electrocatalytic materials is as follows: CO or formic acid ~100% [41,42,43,44], ethylene ~60% [45,46,47], acetate ~40% [48,49,50], and n-propanol ~15% [51]. This necessitates additional purification costs. Thus, electrochemical reduction (ECR) often occurs under favorable reaction conditions at ambient temperatures and pressures, which substantially advantages large-scale applications.

2.2. Challenges of CO2 Electroreduction

Generally, the CO2 reactants are supplied to the active sites (CO2 mass transport) of the catalyst surface (i.e., electrode surface) by diffusion via the non-flowing bulk electrolyte medium. Hence, as discussed before, the reduction reaction begins with the formation of intermediate products. Additionally, the reaction rate or efficiency for obtaining the desired products depends on the binding tendency, electron–proton coupling reactions, and the solubility/mass-transportation efficiency of CO2. In brief, HCOO products are formed by electron transfer and proton injection. If the catalyst surface has only adsorbed protons, this leads to the formation of COOH and CO. In the case of the strong adsorption of CO on the metal surface (i.e., Cu), it will further reduce into the C1 and C2 products; otherwise, the weak adsorption (i.e., Zn, Ag, Au, Pd) of CO will lead to desorption from the catalyst surface without any further reduction processes [52]. Since the mass transportation of CO2 in the bulk electrolyte is limited, and the gas reactant at the catalyst surface is insufficient, several recent studies have tried to overcome this issue by designing gas-diffusion electrode (GDE) materials to achieve high and flawless CO2 flux at the cathode surface [11,53]. The GDE maintains a high CO2 (gas) concentration in the vicinity of the electrocatalyst through the porous-structured electrodes that can enhance gas mass transport and current densities. When compared to the static non-flowing CO2-RR cell, the GDE consists of the following electrode segments: a gas diffusion layer (GDL), a carbon paper/cloth or polytetrafluoroethylene (PTFE) membrane, and a catalyst layer (CL), followed by a microporous carbon/PTFE layer (MPL). In general, the catalyst inks are prepared and coated on the GDL during the fabrication of the GDE [34,54]. The electrode configuration of the GDE is relatively complicated and more efficient than that of laboratory electrochemical H-cell and flow-cell planar electrode systems. In addition, the other technological challenges related to the commercialization of the electrochemical CO2-RR pertain to its efficiency, which is mainly hindered by the voltage efficiency (η), Faradaic efficiency (FE), and electrode stability (Figure 4).
Even though the kinetics of the CO2-RR are thermodynamically more favorable than the HER, the electrocatalytic efficiency is mainly affected by sluggish reaction kinetics; high overpotentials (i.e., voltage efficiency); and poor selectivity, owing to the similar reduction potentials of the various reduction products [8]. The overpotential (η) is defined as the extra energy (i.e., potential/voltage) required to drive CO2 reduction in addition to its thermodynamically required energy. Previous studies have suggested that the standard overpotential for the CO2-RR is −1.90 V (vs. SHE), and the concept of overpotential originated from the sluggish reaction kinetics of the formation of CO2 intermediate products [9,55]. Furthermore, the problems related to voltage efficiency can be overcome by catalyst modifications (surface defects) to enhance the stabilization of the intermediate products on metal surfaces. Secondly, FE is defined as the efficiency with which charges (i.e., electrons) are transferred in a system facilitating an electrochemical reaction for the generation of a desired product. Usually, in electrochemical CO2 conversion reactions, the current is an essential component that is consumed to generate reduced products; in most reported cases, the loss of current (i.e., energy loss) was due to the occurrence/formation of undesirable reactions or products (e.g., H2). Thus, electrocatalytic materials with high H2 overpotentials characteristically present a promising FE for the CO2-RR [52]. The product selectivity of the different catalytic materials is measured by the generated products’ Faradaic efficiency. The FE of the desired products can be calculated by the following expression: FE = nFN/Q × 100%, where n is the number of electrons transferred per mole (products), N is the number of moles (products), Q is the total charge consumed during the electrochemical reaction, and F is the Faradaic constant [56].
The electrocatalyst stability or electrode stability is also considered a key performance indicator for evaluating a catalyst’s long-term applicability via the produced cathodic current density and FE of the desired products during the CO2-RR [5]. Due to the generation of byproducts, the electrolyte medium is considered a primary source of impurities in the working electrode. In some cases, the produced intermediates or byproducts are adsorbed on the electrode surface, restricting further reactions and affecting the electrocatalyst’s performance. In this regard, more research must be carried out to fabricate impurity-resistance working electrodes for long-term stability and to decrease the operational cost of the CO2-RR [52,57]. Figure 4 (right side) depicts various catalyst modification processes and designs for an enhanced electrochemical CO2 reduction reaction. The overpotential and selectivity of the CO2-RR are significantly influenced by the electronic structure (i.e., binding energies with intermediate species) of the catalytic materials. On the other hand, the geometric properties, such as the structure/morphology, surface roughness, porous nature, and particle size, are directly related to the number of catalytic active sites for catalyst–CO2 interactions. Thus, a large grain boundary can be created by highly nanoporous materials, quantum confinement can be achieved through a confined geometry, and electronic properties can be directly modified by alloying and doping processes between metals.

2.3. Factors Affecting CO2 Electroreduction

The efficiency of the CO2-RR is affected by several factors, mainly including the nature or type of electrocatalyst, the electrochemical cell parameters, and the electrolyte. Among these factors, the physicochemical properties of the electrocatalytic materials have a significant role in the CO2-RR efficiency and selectivity. Many strategies have been put forward for the rational design of novel electrocatalytic materials. Nanoscale material synthesis [58], specific facet-oriented material preparation [59], the construction of porous materials [60], the formation of alloys [61], doping [62], defect engineering [63], and surface modification [64] have been implemented to enhance CO2-RR efficiency through increased adsorption sites, enhanced intermediate product formation/binding capabilities, and altered geometric and electronic properties [5].
The activity or catalytic performance of electrocatalysts can be further confirmed through the total current density produced via an anodic CO2-RR. The current density (J) is determined via the normalization of the catalytic current (i.e., anodic current) by the geometric area or by the electrochemical active surface area (ECSA) of the working electrode [65]. On the other hand, the rate of the CO2-RR is indicated by the turnover frequency (TOF), which is derived from the active sites of the catalytic materials [66]. In addition to the above factors, the efficiency of the CO2-RR is mostly influenced by the potential-dependent structure of the solvent, the concentration, the potential-dependent behavior of the ions present in the electrolyte, the impact of the double-layer (adsorbed) species, and electrode–electrolyte ion interactions [5,67].

3. Effect of Surface Defect Sites on CO2-RR Electrocatalysts

The physicochemical properties and performance of all ‘solid-state’ materials are controlled by defects. In general, synthesizing a perfect crystal without any defect is difficult. The surface and electronic performance of catalysts are tuned according to several intrinsic properties, such as the morphology, particle size, crystal facets, lattice distortion, surface defects, surface functionalization, grain boundaries, doping, oxygen (Ov) or copper (Cuv) vacancies, and surface states of the element [9,68,69,70,71,72]. Surface defect sites or structural disorder can be rationally introduced via several engineering strategies, influencing site-specific activities in the CO2-RR [73,74]. Surface/crystal defects can be identified or characterized by several techniques, such as Raman spectroscopy [75], X-ray photoelectron spectroscopy (XPS) [76], X-ray absorption spectroscopy (XAS) [77], electron spin resonance (ESR) [78], aberration-corrected transmission electron microscopy (AC-TEM) [79], and density functional theory (DFT) methods [80,81], based on the changes in the vibrational modes of the crystal, the changes in the bonding energy, the electronic structure of the crystal or local geometric measurements, the determination of unpaired electrons, atomic-scale defect counting, and computational simulations, respectively.
Among the above-mentioned surface properties of the catalytic materials, the defect and interface properties influence the efficiency, selectivity, and performance. Electrochemical reactions (i.e., CO2-RR) mostly depend on the catalysts’ adsorption, charge distribution, and surface structure properties [73,82]. Moreover, defects or lattice imperfections are consistently observed in materials such as crystalline (low-dimensional) and amorphous solids [83,84]. Crystal defects can be classified into four types: line defects, point defects, bulk defects, and planar defects. However, most materials have native defects, which are intensely involved in catalytic activities. In addition to the native defects, other defects can be introduced to a catalyst (i.e., composition, etching, pretreatment, and doping) to increase the catalytic performance [85,86,87,88].
In general, the defects and interfaces have diverse functional scales. In brief, the defects at the edges can stabilize the interface, and the near-interface environment enables defect formation. Consequently, vacancies and doping focus on the 0D environment, while grain boundaries generally exist in 1D or 2D environments. Usually, the 2D and 3D boundaries are the most predominant active sites for electrochemical reactions [89]. Figure 5A shows the systematic strategies that have been applied for defect engineering and interface processes. Crystal defects and interfaces may have a robust synergistic effect on catalytic reactions by creating more active sites for CO2 adsorption for the surface reactions [90]. Figure 5B depicts the effect induced by defects/interfaces on the catalyst surface. The most evident advantages of interface/defect engineering processes are: (i) the increase in the active sites on the catalyst surface, (ii) the facilitating of different types of surface charges (i.e., negative and positive) for CO2 adsorption, (iii) the development of a vacancy-influenced unsaturated coordination environment, and (iv) the creation of a defect-localized non-uniform charge distribution over the catalyst surface. The performance of the EC-CO2-RR over the electrocatalyst mostly depends on the number of active sites on the catalyst surface. Additionally, it can be modified through electronic structure and surface structure regulations. The presence of intrinsic and extrinsic defects in metal-free materials, carbon materials, metal oxides, metal sulfides, and single/bimetallic materials can efficiently facilitate the EC-CO2-RR by modulating and improving the reaction active sites and the electronic structures for the effective adsorption and conversion of the reaction species via the enhancement of the local electron density, the binding energies of the reaction intermediate species, and the conductivity of the materials [91].

3.1. Single/Bimetallic Materials

Over the past several decades, numerous studies have focused on designing high-performance single and bimetallic materials (such as Sn, Zn, Fe, Cu, Bi, Tl, Pb, Au, Ag, Pd, and Pt and Cu-Zn, Sn-Cu, Au-Cu, PdCu, and In-Cu) for high product selectivity and stability [11,92,93]. Among the metallic electrocatalytic materials, copper (Cu) is widely reported as an excellent catalyst for CO2 reduction into the various C2 hydrocarbons [94,95,96], with appreciable current densities of up to 5~10 mA cm−2 [96,97,98]. However, ‘selectivity’ is an important concept that needs to be considered to overcome the practical application and commercialization challenges. The selectivity of the desired hydrogenated products is influenced by the proton concentration near the electrode, the crystal structure, the morphology, and the size of Cu-based catalytic materials [99,100]. On the other hand, single-atom catalysts have attracted much interest owing to their many attractive features and benefits, such as their metal-supportive solid chemical interactions, large atom utilization, low-coordination metal atoms, and specific electronic configurations [101]. Additionally, they have an excellent CO2 to CO selectivity, which indicates that the HER competition is extremely limited due to the large number of inherent isolated active sites, which facilitates the low adsorption of H2O on the catalyst surface [102] and diminishes the H2 generation. Jaramillo and his group reported the combined theoretical and experimental evaluation of the selectivity of different metal electrodes for electrochemical CO2 reduction. According to the well-defined volcano plots of several metals [13], the noble metals, including Au, Ag, Pt, and Cu, have weak binding tendencies and weak interactions with *OCHO intermediates and are thus not able to produce formatted species (low selectivity).
In contrast, Zn and Ni have a strong binding tendency and interact strongly with *OCHO intermediates; however, they also have a low selectivity towards formate, owing to the easy desorption of intermediates from the surface. Sn is located at the top of the volcano plot, demonstrating that it has a sufficient binding tendency with *OCHO and thus produces formate species. Moreover, Cu, Bi, Sn, Sb, Pd, In, and Pb demonstrated high selectivity [103] for HCOO products (Table 1). Specifically, a key drawback of these catalysts is their high overpotentials (i.e., low catalytic activity), which can be overcome by several techniques, such as catalyst modification (i.e., doping, alloying, rational arrangements, porous structures, and atomically thin 2D-engineering and defect-engineering processes) [104,105,106].
In the electrochemical CO2-RR, carbon monoxide is probably the most common intermediate species for producing formate, methanol, and methane. DFT calculations for electrochemical reduction on Cu (211) surfaces were conducted by Peterson et al., showing that the most thermodynamically favorable pathway involved the initial formation of *CO; the subsequent hydrogenation to *HCO, *H2CO (desorbed as HCHO), and *H3CO (methoxy) intermediates; and the formation of CH3OH by *H3CO reduction [107]. Liu et al. reported a new strategy to introduce defective sites by eliminating carbon deposition on the Cu electrode [108]. Since Cu is an excellent metal catalyst that converts CO2 into high-value-added hydrocarbon fuels, it is limited by poor stability due to the deposition of carbon on the electrode surface following the production of CH4, which leads to deactivation. Liu et al. reported the synthesis of 1D Cu-nanoneedles (CuNNs) and found that (1) crystal defects facilitated ethylene formation while suppressing methane production, and (2) crystal defects enhanced the catalytic stability, revealing that the ethylene formation pathway avoided the carbon deposition related to methane formation (Figure 6A,B). The carbon deposition on the Cu electrode favored H2 generation rather than CO, CH4, or C2H4 generation, which was confirmed by the FE% of the annealed (under pure Ar) CuNN electrodes (Figure 6C) and demonstrated that carbon deposition decreased the stability of the electrocatalytic reaction.
In contrast, the FE% of the C2H4 produced by pure CuNNs and OT-CuNNs (annealed under air) was 15% and 52% higher than that produced by the annealed CuNN electrode, respectively (Figure 6D,E). Additionally, the FE% of CH4 decreased by around 4% (CuNNs) and 1.8% (OT-CuNNs) compared to the annealed CuNNs. Furthermore, the instability of the annealed CuNNs was revealed by i-t measurements (Figure 6F), thus proving the long-term stability of CuNNs and OT-CuNN electrodes. Figure 6G,H presents the defect densities vs. the catalytic stability and the Raman measurements. From these results, it can be quickly understood that the different annealing treatments of the Cu electrodes led to various reduction product (C2H4) distributions and enhanced the FE% with excellent stability. Recently, Mi et al. demonstrated the selective electroreduction of CO2 to C2 reduction products in the presence of Cu nanowires derived from Cu3N. Compared to the Cu3N nanowires, the Cu3N-derived Cu nanowires exhibited many grain boundaries, which undoubtedly facilitated high selectivity, catalytic activity, and long-term durability. Furthermore, the Cu-NWs (single-metal catalyst) showed a high FE of 86% for C2 products at −1.0 V (vs. RHE). Additionally, the stability test revealed 28h of flawless durability for continuous electrolysis [109]. Another work reported by Hoang et al. revealed the excellent catalytic performance of a bimetallic Cu-based catalyst (Cu-Ag) synthesized by the electrodeposition technique for the CO2-RR. The obtained results showed that the Cu-Ag film containing 6% Ag exhibited a superior CO2 reduction performance, with high FEs of 60% and 25% for ethylene (C2H4) and ethanol (C2H5OH), respectively, at −0.7 V (vs. RHE) with a total current density of 300 mA/cm2 [110]. More recently, Fu et al. reported an atomically dispersed highly defective Ni-N3/C catalyst for superior electrochemical CO2 reduction reactions. Herein, the micro-mesoporous carbon substrate enabled high-charge transport and provided more active sites for CO2 adsorption and conversion. The as-reported Ni-SAs/HMMNC-800 electrocatalyst achieved excellent selectivity, with a high FE of 99.5% for the formation of CO at −0.7 V (vs. RHE), and exhibited a high current density of 13 mA/cm2 with long-term working stability. The electrochemical activity enhancement was mainly due to the Ni-N3 coordination sites coupled with defects, which favored the generation of large amounts of *COOH intermediates and suppressed competing H2 generation [111]. Moreover, Zhang et al. synthesized highly defective and amorphous Au-NCs@ MnO2 nanosheet core–shells (Au NCs@a-MnO2 NSs) for effective CO2 reduction into CO. Herein, the Au NCs@a-MnO2 NSs showed highly gas-permeable behavior due to their nanocrystal/defective nanosheet/core–shell structures, which helped to expand their CO2 adsorption capabilities. Additionally, the Au NC core promoted high electron conductivity and boosted the electron transport from the catalyst to the CO2 molecules. Thus, the Au NCs@ α-MnO2 NSs core–shell electrocatalyst exhibited a high FE of 90.5% for CO conversion at −0.7 V (vs. RHE), with a high current density of 14.3 mA/cm2. The above results revealed that the synergistic structural and electronic effect between Au and α-MnO2 was the major reason for the excellent electrocatalytic activity, selectivity, and stability [112]. Another recent work reported by Wu et al. elucidated the construction of single-atomic Ni-sites in CNT for the CO2-RR. The Ni/N-CNTs catalyst was constructed through a one-pot pyrolysis route, exhibiting many surface defects and single-atom Ni sites for highly effective CO2 capture and reduction reactions. Furthermore, the as-prepared Ni/N-CNT electrocatalyst exhibited excellent electrocatalytic performance with a 98% FE for CO at −0.65 V (vs. RHE) and an excellent TOF of 304.5 h−1 [113]. Furthermore, Ali et al. reported the synthesis of porous aza-doped unique 2D graphene analogs for the electrochemical conversion of CO2 to formic acid. The single-layered porous aza-fused/pi-conjugated graphene exhibited highly ordered specific porous structures, and the dispersed N-atoms helped to stabilize the CO [114]. In addition to the abovementioned research, a recent study conducted by Ni et al. investigated in-plane defective metal-nitrogen-carbon catalysts for the electroreduction of CO2 (CO2-RR). Herein, the authors demonstrated that the intrinsic carbon defects could be substantially enhanced by coupling single-atom Fe-N4 sites. The DFT calculations revealed that the energy barrier for CO2 reduction could be reduced by the intrinsic defects associated with the Fe-N4 sites, which could also suppress H2 production. Thus, the resulting electrocatalyst exhibited an FE of around 90% for CO, with a high current density of 33 mA/cm2 [115].
Table 1. Single- and bimetallic-material-based electrocatalysts for electrochemical CO2-RR.
Table 1. Single- and bimetallic-material-based electrocatalysts for electrochemical CO2-RR.
Single and Bimetallic Materials for CO2-RR
CatalystType of DefectElectrolyte/E vs. RHEProduct* FE (%)Ref.
BiOx@CO2 vacancy0.1 M KHCO3/−0.52 V HCOO89.3[116]
Bi-few layersExfoliation0.5 M KHCO3/−0.7 VHCOO85[117]
Exfoliated Bi nanosheetsIncreased edge sites/defects0.5 M KHCO3/−1.1 VHCOO86[118]
Bulk Bi nanosheets64.9
Sn foilHigh native O2 content0.1 M KHCO3/−1.36 VHCOO49.1[119]
Heat-treated Sn dendrites71.6
Few-layer Sb nanosheetsExfoliation/active edge sites0.5 M NaHCO3/−1.07 VHCOO84[120]
Cu dendriteOxide-derived surface
active sites
1.0 M Na2SO4/−0.8 VC2H455[121]
C2H6
Cu (100)Crystal orientations/
defects
0.1 M KHCO3/−5 mA/cm2C2H440.4[122]
C2H5OH9.7
CO0.9
CH430.4
Cu mesocrystalsFacets/edge defects0.1 M KHCO3/−0.99 VC2H427.2[123]
HCOO4.3
CO0.55
CH41.47
Bi/CuSurface oxide layer0.1 M KHCO3/−1.69 VHCOO91.3[124]
Pd-SnAlloying/tuning surface electronic structures0.5 M KHCO3/−0.26 VHCOO99[125]
Sn56.3-Pb43.7/
Carbon cloth
Alloying0.5 M KHCO3/−2.19 VHCOO79.8[126]
Core–shell Ag-Sn NPsO2 vacant sites0.5 M NaHCO3/−0.81 VHCOO87.2[127]
Cu/AuCore–shell structures0.5 M KHCO3/−0.65 VCO30[128]
Pd85Cu15Metal doping/alloying0.1 M KHCO3/−0.89 VCO86[129]
Cu38Cd62Alloying0.05 M KHCO3/−1.12 VHCOO10[130]
CO43
Cu-InAlloying/edge sites0.1 M KHCO3/−0.7 VCO95[131]
Ordered Cu-PdAlloying/tuned geometric effects1.0 M KOH/−0.55 VCO80[132]
Disordered Cu-Pd1.0 M KOH/−0.60 VCO60
CH41.0
C2H44.0
C2H5OH2.0
Phase-separated Cu-Pd1.0 M KOH/−0.71 VCO17
C2H448
C2H5OH15
* FE% = Faradaic efficiency (%); HCOO = formate/formic acid; CO = carbon monoxide; C2H5OH = ethanol; CH4 = methane; C2H4 = ethylene; H2 = hydrogen.

3.2. Metal Alloys

As seen above, numerous metal electrodes have been utilized/investigated for the electrochemical CO2-RR. Furthermore, according the abovementioned studies, the binding energy/tendency of the metals strongly influences the selectivity of the CO2 reduction products. Several recent studies demonstrate that alloying has received much attention due to its excellent catalytic performance, achieved by tuning the intermediate products’ binding energies and stabilization degrees [70] (Table 2). Investigations have also been conducted on ECR in Sn-based alloys through an aqueous medium of 0.5 M KHCO3. The Sn-Pb alloy with a surface composition of Sn56.3Pb43.7 had the greatest FE of 79.8% for HCOOH and the highest current density of 45.7 mA/cm2 [126]. A Cu alloy catalyst was also reported to be selective for CO in aqueous ECR, with a FE higher than 90% for CO [133]. According to a report by Xu et al., incorporating Au with highly stabilized Cu nanoparticles resulted in very low overpotentials in the CO2-RR [134]. Surface strain refers to the force in the mismatched lattice when doping a metal into other metal components. It plays a significant role in influencing catalytic performance. The compressive strain effect in a CuAg alloy facilitates the reduction of alcohol and acid [48,110,135]. It is believed that core–shell structures can provide more active sites due to the surface strain. A nanoporous core–shell AuCu3@Au electrode had an FE of 97.27% for CO production [136] (concave defect). Au-Pd core–shell nanoparticles reached a maximum FE of 96.7% for CO production [137]. The tensile strain in the core–shell structures makes the adsorption of *COOH easier than *CO [138]. A study on a heterolayer slab model of Cu-M (M = Rh, Ir, Pd, Pt) concluded that both the strain and ligand effects impacted the adsorption energies of the intermediates of the CO2-RR, and the adsorption of *COOH and *CHO was more strain-sensitive than that of *CO and *COH on the Cu surface [138]. Core–shell Cu@CuEu nanoparticles were found to have 3.4 times the CH4 FE of Cu nanoparticles due to the vacancy in the CuEu layers [139]. The compressive strain in core–shell Bi@Sn nanoparticles boosted the electroreduction of CO2 into formic acid, with an FE of 91% [140]. Furthermore, the micro-morphology of a core–shell porous-structured Cu@Sn electrode reached an FE of 100% for formate production [141]. The grain boundary is the interface between two crystallites. Grain boundaries in gold nanoparticles have been applied to control the product selectivity of the CO2-RR [142]. For example, Au nanoparticles rich in grain boundaries have an FE of 94% for CO production. Copper rich in grain boundaries was found to have a high FE of 73% for ethylene, ethanol, and propanol production in a wide potential range, which exceeded that of most copper electrocatalysts [143]. Bismuth grain boundaries were formed by galvanic-cell deposition to improve formate selectivity, with an FE of more than 92% [144].

3.3. Metal Oxides

The success of electrochemical CO2 reduction to value-added C1 and C2 chemicals/fuels may resolve the environmental issues brought on by excessive fossil fuel usage. An effective method for lowering the amount of metal used and for electro-reducing CO2 is to load transition metals onto metal oxides (Figure 7). Numerous recent publications in peer-reviewed journals have documented the adjustment (and general enhancement) of the selectivity and activity of materials employed in the electrochemical conversion of CO2 to value-added chemicals following oxide formation. Such species are known as oxide-derived (0D) materials, and improved CO2 reduction activity has been demonstrated for 0D-Cu [145], 0D-Au [94], and 0D-Pb [146] when compared to corresponding metal substrates. The other catalysts investigated for CO2 electroreduction can be divided into three categories. The first category, consisting of gold, silver, zinc, and (to a lesser degree) copper, has a higher Faradaic efficacy in producing CO [147]. The second group, which includes metals such as Pt, Ni, Fe, Ti, and Al, mainly forms hydrogen [148], while Pt may also produce hydrocarbons containing up to nine carbon atoms, though with a low yield [149]. Finally, the third category consists of metals with a higher hydrogen overpotential, such as Hg, Cd, Pb, In, and Sn, which convert CO2 exclusively to formic acid and formate salts. This third set of catalysts has good specificity, since they primarily create aqueous formate with hydrogen gas. As per Oloman and Li, these might provide the foundation for industrial formate synthesis by rapid CO2 electroreduction [37]. Lead and tin have already been widely explored in this regard; nonetheless, their stability under extended electrolysis remains restricted [150,151]. Recently, a bismuth oxide (BiOn) cluster catalyst was developed for an effective CO2-RR to formate. This BiOn cluster catalyst displayed good activity, selectivity, and stability during formate synthesis, with a formate Faradaic efficiency of more than 90% at current densities of up to 500 mA/cm2 in an alkaline membrane electrode assembly electrolyzer, equivalent to a mass action of 3750 AgBi. Tin (Sn) [152], bismuth (Bi) [43], indium (In) [153], lead (Pb) [154], and mercury (Hg) are common metal catalysts used for the CO2-RR to produce formate.
A fast-heating approach was used to create 1.72 and 3.51 nm thick Co3O4 films as a prototype. The atomic thickness provided the Co3O4 with an abundance of active sites, allowing the adsorption of a substantial quantity of CO2. The greater and more scattered charge densities around the Fermi level enabled improved electron transport. The electrocatalytic activity of the 1.72 nm thick Co3O4 layers was just 1.5 and 20 times greater than that of the 3.51 nm thick Co3O4 layers and the bulk analog. In addition, the 1.72 nm thick Co3O4 films demonstrated a formate Faradaic efficiency of more than 60% in 20 h [155]. Zn-based compounds were also investigated as ECR catalysts. ECR on such a highly purified Zn electrode is often preferential for producing HCOOH and CO in an aqueous environment [156]. Previously, the creation of a mixed metal-oxide CuInO2 nanostructure during CO2 electroreduction stabilized the active catalytic phase of copper(I). The introduction of a nanoporous Sn:In2O3 interlayer to a Cu2O pre-catalyst system resulted in the creation of CuInO2 nanoparticles with a much greater efficacy for CO2 electroreduction and a lower overpotential than traditional Cu nanoparticles obtained from solitary Cu2O [157]. Metal oxides may also be utilized directly as catalysts for the reduction of CO2 in addition to metal-based catalysts. Tin oxide nanocrystals were studied by Meyer et al. as effective electrocatalysts for the production of formate [158]. These nanostructured tin oxide catalysts had optimum Faradaic efficiencies of more than 93% and current densities of 10 mA/cm2 at overpotentials as low as 340 mV. Supporting this research, improved active CO2 electroreduction catalysts, including decreased SnO2 porous nanowires [159], hierarchical mesoporous SnO2 nanosheets [160], and hybrid SnOx/carbon nanotubes [161], have been produced. Thorough research by Bijandra et al. revealed that the presence of several grain boundaries inside this porous structure was the cause of the enhanced CO2 reduction performance on decreased SnO2 porous nanowires [159]. In a two-step procedure, Zhang et al. created multilayer mesoporous SnO2 nanosheets using carbon cloth after synthesizing SnS2 hydrothermally [162]. Subsequently, after further calcination, highly porous SnO2 nanosheets with a surface area of 93.6 m2/g were produced [160]. To further understand the mechanisms behind the higher electroreduction activity of Pt oxides, the species adsorbed onto the surface of Pt and Pt oxide were examined using SEIRAS. Following CO2 electroreduction, the primary adsorption species were methanol and HCOO on the Pt oxide, whereas they were and methanol and linear-CO on the Pt. It has been shown that HCOO on Pt oxide and CO on Pt are necessary for the CO2 electroreduction process. Although CO aggressively adsorbs on the active site and hinders subsequent reactions, the adsorption species significantly impact the CO2 electroreduction efficiency [163]. The Fermi level of ZnO may be tuned by substituting heteroatoms with various outer electron configurations (Mo and Cu) for the Zn site [164].
Recently, Zhong et al. synthesized Vo-CuO(Sn) nanosheets through a simple one-pot synthesis route (Figure 8A), and the as-synthesized Vo-CuO(Sn) consisting of oxygen vacancies with Sn dopants was utilized for CO2-RR applications [165]. The structural/morphological investigations confirmed that the Sn atom slightly affected the CuO nanosheet morphology. The amount of Sn dopants was measured using ICP-OES, indicating 2.97% atomic Sn. The experimental results showed that the Vo-CuO(Sn) exhibited higher selectivity for the CO2-RR in the CO2-saturated electrolyte compared to the Ar-saturated electrolyte medium and exhibited a higher current density of 38.65 mA/cm2. The CO2-RR resulted in CO as the reduction product, which was confirmed by the GC-MS results. Vo-CuO(Sn) presented a 42.1% FE for CO at −0.23 V (vs. RHE) and a 99% FE at −0.53 V (vs. RHE); thus, the selectivity of CO was around 95% over Vo-CuO(Sn). The substantial enhancement in the catalytic activity of Vo-CuO(Sn) was mainly due to the synergistic effect of Sn doping and enriched oxygen vacancies.
Furthermore, Huang et al. reported the preparation of O-vacancy-engineered indium oxide (InOx) nanoribbons (NRs) for a highly efficient CO2-RR (Figure 8B). They synthesized three InOx NRs with different VO concentrations by simply calcinating them under different atmospheres. High VO concentrations endowed the H–InOx NRs with an excellent FE% (HCOO) of over 80% at broad potential values and a total FE of 91.7% with a higher current density. The H–InOx NRs exhibited long-term stability for the conversion of CO2 to formate without substantial decay during 20 h of continuous reduction treatment. The enhanced CO2 adsorption, activation, and reduction performance of H–InOx were owed to the high VO concentrations, thus showing the importance of defect engineering on the catalyst surface [166]. Additionally, a recent work reported by Sun et al. described the enhanced CO2 electroreduction activity of oxygen vacancies/defects (Vo) on a 1D In-SnO2 hollow-nanofiber catalyst [167]. The authors prepared SnO2 hollow nanofibers through the electrospinning technique and with In as a deponent. The oxygen vacancies/defects were confirmed through XPS and EPR analyses (Figure 8C). As can be seen from the schematic diagram (e), the Vo in the In–SnO2 was considered an anion vacancy (n-type semiconductor) and could act as a quasi-free electron donor. The free electron readily available in the Vo on the surface contributed to the high electronic conductivity and enhanced the local surface electron densities for the electrocatalytic reactions. These electron-rich domains also served as highly efficient active sites to stabilize the reduction intermediates such as CO2 for O–bonding, which further protonated to generate *OCHO for HCOO production. Hence, the In–SnO2 Vo hollow-nanofiber electrocatalyst resulted in an FE of 86.2% in the conversion of CO2 to formate species at −1.34 V (vs. RHE).
More recently, Wang et al. reported the synthesis of Cu–ZnO electrocatalyst by adopting the low-valance Cu-doping and oxygen vacancy engineering techniques for electrochemical CO2 reduction reactions. The production or the selectivity of CO can be optimized by weakening the adsorption sites of *H through oxygen vacancies (Vos). Thus, the as-synthesized Cu-ZnO (Vos) resulted in a >80% FE for the conversion of CO2 to CO products within the potential range of −0.76 V to −1.06 V (vs. RHE), with a high current density of 45 mA/cm2 [168]. Geng et al. prepared ZnO nanosheets rich in oxygen defects through H2 plasma treatment for the enhanced electrochemical conversion of CO2 to CO. The Gibbs free energy for the CO2 activation process in the presence of oxygen-vacancy ZnO nanosheets decreased by 0.20 eV compared to the ZnO nanosheets without oxygen vacancies. The resultant VO-ZnO nanosheets produced a current density of −16.1 mA/cm2 with a high FE of 83% for the generation of CO from CO2. This proved that Vos facilitated the activation of CO2 with superior reaction kinetics [169]. Recently, Zhang et al. prepared porous indium oxynitride nanosheets with oxygen vacancies and N deponents (Vo-N–InON) to produce formic acid from CO2. The Vo-N–InON electrocatalyst exhibited 95.1% formate selectivity at −0.8 V (vs. RHE) in the flow cell, producing a current density of 121.1 mA/cm2 at −1.13 V (vs. RHE). In addition, the DFT calculations demonstrated that the activation energy for forming an intermediate product (*OCHO) was drastically reduced by about −0.812 eV after introducing Vo. This revealed the synergistic effect of oxygen vacancies and N-doping in the In2O3 catalyst during the CO2-RR [170]. Consequently, Du et al. reported the introduction of high-energy plasma-assisted oxygen vacancies (Vo) in In2O3 (Vo-In2O3) films for the electrocatalytic CO2 reduction reaction. The DFT calculation results indicated that the CO2 reduction reaction proceeded by forming *COOH as an intermediate product, and the Gibbs free energy for the *COOH decreased with the increase in the Vos. The oxygen richness inherited inside the In2O3 facilitated the efficient formation of HCOOH, and resulted in 70% and 83% FEs for the HCOOH and C1 reduction products, respectively [171]. A recent study by Zong et al. explored the synthesis of morphology-controlled ZnO by the electrospinning method with enriched O-vacancies to enhance the electrochemical reduction of CO2 to CO. The obtained p-ZnO-800 electrocatalyst resulted in a high current density of 150 mA/cm2 and an 82.93% FE for CO in a flow-cell configuration reactor. The authors concluded that the enhanced electrocatalytic activity of the p-ZnO-800 catalyst was due to several reasons, including the increased active sites from the ZnO clusters, the enhanced electron transfer process, and the high selectivity towards CO by the introduction of O-vacancies [172]. Additionally, Ren et al. demonstrated cerium-doped ZnO catalysts (Ce0.016Zn0.984O) for the enhancement of the CO2 electroreduction (CO2-RR) process. The oxygen vacancies were achieved by doping the Ce3+ deponent at different concentrations. The Ce0.016Zn0.984O with the highest Vo exhibited a high current density of 24 mA/cm2 at −1.0 V (vs. RHE), with a high FE of 88% for CO [173]. This study revealed that an increase in the electron transfer kinetics and a reduction in the thermodynamic energy barrier of the metal-oxide catalysts could be achieved through defect (vacancy) engineering approaches. Furthermore, EC-CO2-RR activities of various metal-oxide catalysts were shown in Table 3.

3.4. Metal Sulfides

Due to their favorable electronic band gaps, energy-band location exposed active sites, and potential catalytic activity, metal sulfides have gained much interest. The significant advances made possible by interfacial engineering are primarily attributable to its remarkable capacity for controlling electron and mass transport, adjusting intermediate adsorption, preventing severe catalyst aggregation, and offering sophisticated promoters for the logical design of highly effective catalysts [189]. Meanwhile, defect engineering in metal sulfides involving vacancies and metal/heteroatom doping enhances the electrocatalytic efficiency of the catalytic materials, owing to their increased current densities, Faradaic efficiencies, and active reaction sites and decreased work function compared to pristine catalyst materials [190,191]. Changes to the composition, morphology, and electrical structure of bimetallic sulfides, on the other hand, may be accomplished by metal doping and the fabrication of heterostructures that have significant promise in the electrocatalytic CO2-RR [192]. A hierarchical CuS hollow polyhedron (CuS-HP) was developed for the electrochemical CO2 reduction (E-CO2R) in neutral pH aqueous environments. The CuS-HP experienced structural rebuilding to form a sulfur-doped metallic Cu catalyst under E-CO2R conditions, encouraging formate synthesis with a Faradaic efficiency of >90% throughout a broad potential range [193]. Compared to E-MoS2 and other previously reported transition metal sulfide electrocatalysts, the improved N-MoS2@NCDs-180 for CO2 electroreduction demonstrated a high CO Faradaic efficiency of up to 90.2% and a low onset overpotential of 130 mV for CO production [194]. CuInS2 hollow nanoparticles were created by Xiong et al. with homogeneous bimetallic mixing, combining the benefits of hollow nanostructures with the synergy of bimetallic sulfides and exhibiting an outstanding electrochemical CO2-RR capability [195]. At −0.7 V, the FE for HCOOH was 72.8%, and at −1.0 V, the FE for CO was 82.3%. Amorphous Ag–Bi–S–O-altered Bi nanoparticles were produced by Zhang et al. via the electrocatalytic processing of Ag0.95BiS0.75O3.1 nanorods [188]. The catalysts produced a high FE for HCOOH of up to 94.3% and a significant fractional current density of 12.52 mA/cm2 when the overpotential was just 450 mV. Additionally, another recent report suggested that adding Mn atoms to In2S3 nanosheets enhanced the FE and the current density of the carbon products [196]. Mn–In2S3 nanosheets had a high FE of 92% and a significant current density of 20.1 mA/cm2 at −0.9 V [39]. In the CO2-RR process, Zn-based catalysts exhibited a high selectivity for CO. However, they often present a large overpotential and slow reaction kinetics. Due to Zn’s poor adsorption onto the chemical intermediate *COOH in the CO synthesis route caused by its electron-rich properties, a significant energy barrier must be overcome during the CO2-RR procedure [197]. In comparison to a reversible hydrogen electrode (RHE), a cadmium sulfide (CdS) nanoneedle grid catalyst could reduce CO2 to CO with a Faraday efficiency of up to 95.5:4.0% at an exceptional exchange rate of 212 mA/cm2 at @1.2 V [198]. A schematic illustration of metal oxides and metal sulfides and their structures for the electroreduction of CO2 is depicted in Figure 3. Furthermore, EC-CO2-RR activities of various metal-sulfide catalysts were shown in Table 4.

3.5. Carbon-Based Materials

In terms of defect-engineered materials, carbon-based nanomaterials (CBNs) have been identified as an excellent choice for the electroreduction of CO2 under liquid- and gas-phase reactions [1,207]. From the family of carbon nanostructures, all the allotropes, such as graphene, carbon nanotubes (CNTs), and carbon nanodiamond materials, have obtained paramount importance in many catalytic activities owing to their remarkable properties [208]. This is mainly due to the unhinged electron distribution and defect-assisted electronic structural distortion in the catalytic materials. Introducing defects can alter the carbon skeleton’s charge state and, thereby, the catalytic material’s conductivity and stability [209,210]. Defect engineering includes intrinsic effects (topological and edge defects, vacancies, and holes made without any dopants) and extrinsic defects (such as metal atom sites and non-metal heteroatom doping) [211,212,213]. Each of these techniques has its advantages for catalytic reactions in terms of catalytic stability, product selectivity, electron transfer, increasing the active reaction sites, product stability, and the stabilization of intermediates, which contribute to enhancing CO2 reduction reactions (CO2-RR) [214,215,216].
Dai’s research group was the first to shed light on the use of defect-induced CBNs in catalysis via their ground-breaking work on oxygen reduction using vertically aligned nitrogen-containing carbon nanotubes (VA-NCNTs) published in 2009 [217]. The authors demonstrated this metal-free catalyst for high-performance oxygen reduction reactions (ORRs) in fuel cells with better electrocatalytic activity and long-term stability. Following this, numerous methodologies and techniques have been employed to formulate CBNs as an efficient catalytic material for the electrocatalytic CO2-RR. The main techniques that have been employed to modify intrinsic defects, i.e., point defects and topological defects, include mechanical ball milling [218], chemical oxidation or etching [219], plasma etching [220,221], nitrogen removal [222], and in situ etching [223]. On the other hand, extrinsic defects such as heteroatom doping with varied electronegativity [85,224] and metal-atom dispersed active sites [225,226,227] are primarily implemented using various chemical vapor deposition (CVD) [228,229,230] and pyrolysis [231] methods for heteroatom doping and pyrolysis synthesis [232,233], with defect engineering [234], spatial confinement [235,236] and coordination design [237] used for the metal-atom dispersive site processes. The appropriate synthesis method yields distinctive electrocatalytic materials that can be employed in efficient CO2-RRs in unique ways to provide carbon fuels and other value-added products. Figure 9 depicts the various defect engineering methods, including the modification of intrinsic and extrinsic defects, applied to synthesize highly active electrocatalysts.
The utilization of CBNs in the electrocatalysis of carbon dioxide mainly results in the formation of CO, which is the primary intermediate (*CO) in most CO2 reduction pathways. For instance, Ajayan et al. [230] investigated the reduction of CO2 to CO using N-doped 3D graphene foam. In this work, pyrrole and pyridinic-N defects were modified to reduce the barrier height of the *COOH intermediate formation. This was supported by density functional theory (DFT) calculations. Meanwhile, Zhou et al. [85] synthesized an N-doped CNT array with more accessible pyridinic-N sites for the CO2 reactant and *COOH intermediate in the electroreduction of CO2 to CO. In another work by Meyer et al. [241], N-doped carbon nanotubes were employed as a robust catalytic material for the electroreduction of CO2 in aqueous media to generate formate as an end product. This metal-free catalyst was supported by a polyethyleneimine (PEI) co-catalyst to enhance the CO2 absorption, thereby increasing the Faradaic efficiency to 87% with a decreased catalytic overpotential for the formation of CO2●− from CO2 reactant. Wang et al. [242] synthesized another N-doped graphene from melamine and graphene oxide to catalyze CO2 into formate; the presence of over 5 atomic % nitrogen atoms triggered the CO2 conversion with a small overpotential of 0.24V, involving graphitic, pyrrolic, and pyridinic-N atoms. Nitrogen is the most commonly explored heteroatom with high electronegativity to carbon atoms, e.g., N-doped graphene, quantum dots (QDs), and CNTs [49,243]. Other heteroatoms, such as B, F, and S, have also been doped with various carbon materials and employed for electrocatalysis, showing unusual activity compared to metal catalysts [238,244,245]. These heteroatom-doped electrocatalysts present enhanced catalytic activity for the electroreduction process compared to pristine graphitic carbon materials, which have trouble interacting with CO2 molecules due to the inactive neutral carbon atoms in the sp2-conjugated structures (Figure 10).
It is a well-established fact that metal catalysts or the presence of metal atoms in composites substantially enhance catalytic activity [246,247]. Introducing metallic active sites on CBNs with uniform dispersion boosts the electrocatalytic activity. Recently, single-metal-atom electrocatalysts have attracted much attention due to their high atomic utilization rate [248,249,250]. Similar to the carbon effect, these metal atoms can strongly bind with C–N atoms in the network, achieving outstanding catalytic performance and stability. In particular, the M–N–C (metal-nitrogen-carbon) system not only has a high atomic utilization rate but also can effectively reduce the hydrogen adsorption on metal atom sites, thereby enhancing the selectivity and efficiency of the CO2-RR by greatly reducing the HER [251,252,253]. Among the works on the ORR using metal catalysts, Strasser et al. [225] first explored electrocatalytic reduction (ECR) using Fe–N–C, Mn–N–C, Fe, and Mn–N–C catalysts and showed high performance. Following this, many noble metals and other transition metals were paired in M–N–C systems; for instance, Kattel et al. [254] prepared N-doped carbon-assisted Pd single-atom catalysts possessing well-dispersed Pd-N4 sites that greatly stabilized the adsorbed CO2 reduction intermediates, enhancing the ECR at lower overpotentials. However, considering the expensiveness and rarity of noble metals, low-cost single-atom transition metals have also been extensively explored. Co-N-C, Fe-N-C, and Ni-N-C are the most commonly investigated catalysts for selective ECR reactions. Among these, compared to Co-N-C and Fe-N-C, Ni-N-C is considered to be an effective electrocatalyst due to its higher CO selectivity and improved catalytic activity. Jiang et al. [251] synthesized a highly porous microwave-peeled graphene oxide (MPGO) with major surface defects and single-atom Ni anchoring (Ni-N-MPGO) with a loading of about 6.9% Ni. The improved ECR performance of the Ni-N-MPGO was established by forming edge-anchoring unsaturated Ni–N active structures in the catalyst with a Faradaic efficiency of 92.1%. Various modifications and designs of CBNs with intrinsic and extrinsic defects have presented catalytic abilities with exceptional performance, selectivity, and stability for ECR and the CO2-RR (Table 5).

3.6. Metal Nitrides

Transition metal nitrides (TMNs) have shown excellent catalytic performance in electrochemical CO2 reduction reactions due to their high mechanical/electrochemical stability and metal-like properties [270]. Jiang et al. reported low-cost, metal-free, C-doped, line-defect embedded boron nitride (BN) nanoribbons for electrocatalytic CO2 reduction reactions. The boron and carbon dopants acted as dual active sites. These two active sites could strongly bind with CO intermediates and thus promote the selective conversion of CO2 to CH4. Additionally, the edges of B and C atoms enabled the coupling of *CHx and CO intermediates, which led to the formation of C2 reduction products such as ethylene and ethanol [271]. Yin et al. reported the synthesis of a new perovskite-type Cu(I) nitride (Cu3N) nanocube (NC) electrocatalyst at different sizes (10, 20, and 25 nm NCs) for the electrochemical CO2-RR. The experimental results showed that the 25 nm Cu3N-NCs exhibited significant catalytic activity and selectivity for the conversion of CO2 to ethylene, with an FE of ~60% at −1.6 V (vs. RHE) [272]. Moreover, Mi et al. achieved the selective electrochemical reduction of CO2 to C2 products using Cu3N-derived Cu nanowires (NWs) with high-density grain boundaries. The reported grain-boundary-induced Cu3N-derived Cu-NWs exhibited FEs for the C2 products of ethanol, ethane, and ethylene of around 8%, 12%, and 66% at −1.0 V (vs. RHE), respectively [109]. Furthermore, Liang et al. synthesized a Cu–Cu3N core–shell-like electrocatalyst to produce multi-carbon products from CO2. The Cu–Cu3N electrocatalyst exhibited an FE of around 64 ± 2% for C2+ reduction products, with an enhanced working stability over 30 h. The Cu3N acted as an active center and reduced the energy barrier of the CO dimerization reaction [273]. Conversely, Liu et al. prepared Pd-modified niobium nitride (Pd/NbN) and vanadium nitride (Pd/VN) electrocatalysts for the EC-CO2-RR for syngas production. The experimental and computational calculations resulted in a higher FE% for CO achieved by the Pd/NbN than by the Pd/VN and commercial Pd/C catalytic materials. This work suggested that NbN could be a good substrate for Pd modifications, and the 5 wt.% Pd/NbN exhibited higher CO2-R activities than the 10 wt.% Pd/C commercial electrocatalyst for the production of syngas [270]. Li et al. constructed unique isolated FeN4 active sites on a disordered carbon substrate (FeN4/C) for enhanced electrochemical CO2 reduction reactions. The FeN4/C catalyst achieved an FE of around ~93% for CO at −0.6 V (vs. RHE) with a high current density of 2.5 mA/cm2, improving upon that of pristine N/C (46%) and Fe/C (23%). Additionally, the FeN4/C demonstrated 24 h of excellent working stability in the EC-CO2-RR [274].

4. Conclusions and Perspectives

Over the next few decades, renewable energy may remain the foremost energy source due to the over-consumption of fossil fuels. At the same time, the impact of CO2 emissions is still a severe issue in our modern technological society. The electrochemical CO2 reduction reaction (CO2-RR) is an emerging alternative technique that provides a fascinating way to reduce CO2 into C1, C2, and various value-added chemicals/fuels. Numerous electrocatalysts are used for the CO2-RR, among which several metals with surface imperfections or defects have been identified as exceptional catalysts that can achieve the effective conversion of CO2 into valuable hydrocarbon products. Controlling, modulating, and tuning metal/catalyst surfaces is essential to achieve stable and highly selective CO2 reduction. With the ultimate purpose of establishing a carbon-neutral economy, various strategies have been proposed and executed at the laboratory level over the past decade to uncover techniques with technical and economic feasibility. Amongst others, defect engineering has found its place at the top of the list owing to its exceptional catalytic activity and increased Faradaic efficiency for the electrocatalytic conversion of CO2 into desired value-added carbonaceous fuels and chemicals that can be stored for future use. Defects in the catalytic surface provide more room to accommodate more CO2 molecules, with great increases in the active sites on the surface and porous levels. In this mini-review, we meticulously summarized the relevant aspects of this subject, beginning with the mechanistic pathways of the electrochemical CO2-RR and the advantages and economic benefits of the various conversion reactions, before moving onto the challenges surrounding the processes and the possible factors affecting the conversion of CO2 into the various value-added C1 and C2 reduction products.
Over the past several decades, numerous electrode materials (i.e. electrocatalysts) have been utilized/investigated for the electrochemical CO2-RR. More importantly, the binding energy/tendency of the metals strongly influences the selectivity of the CO2 reduction products. Thus, an alternative catalyst modification strategy is necessary to overcome the practical complications of the electrochemical CO2-RR. This mini-review discussed the main practical problems, challenges, and mechanistic pathways of the EC-CO2-RR and presented clear interpretations. In particular, the recently developed electrocatalyst design processes/modifications via surface-defect engineering approaches using significant metallic, metal-oxide/sulfide, carbon-based, alloy, and metal-nitride electrocatalysts were discussed. Furthermore, this review presented past, present, and future perspectives on catalyst design for the electrochemical CO2-RR with the aim of producing desired fuels/products.

Author Contributions

Conceptualization, S.B., T.C.-K.Y. and M.K.; resources, T.C.-K.Y. and S.-W.C.; writing—original draft preparation, S.B., A.H., H.V., S.-W.C., T.C.-K.Y. and M.K.; writing—review and editing, S.B., A.H., H.V., T.C.-K.Y. and M.K.; supervision, T.C.-K.Y. and M.K.; project administration, T.C.-K.Y.; funding acquisition, T.C.-K.Y. and S.-W.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Science and Technology Council, Taiwan (NSTC, project numbers: 110–2221-E-027–006-MY2; 110–2923-E-027–001-MY3).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Seh, Z.W.; Kibsgaard, J.; Dickens, C.F.; Chorkendorff, I.; Nørskov, J.K.; Jaramillo, T.F. Combining Theory and Experiment in Electrocatalysis: Insights into Materials Design. Science 2017, 355, eaad4998. [Google Scholar] [CrossRef]
  2. Whipple, D.T.; Kenis, P.J.A. Prospects of CO2 Utilization via Direct Heterogeneous Electrochemical Reduction. J. Phys. Chem. Lett. 2010, 1, 3451–3458. [Google Scholar] [CrossRef]
  3. Jouny, M.; Luc, W.; Jiao, F. General Techno-Economic Analysis of CO2 Electrolysis Systems. Ind. Eng. Chem. Res. 2018, 57, 2165–2177. [Google Scholar] [CrossRef]
  4. Wu, B.; Chen, J.; Qian, L. Recent Advances in Heterogeneous Electroreduction of CO2 on Copper-Based Catalysts. Catalysts 2022, 12, 860. [Google Scholar] [CrossRef]
  5. Zhang, X.; Guo, S.-X.; Gandionco, K.A.; Bond, A.M.; Zhang, J. Electrocatalytic Carbon Dioxide Reduction: From Fundamental Principles to Catalyst Design. Mater. Today Adv. 2020, 7, 100074. [Google Scholar] [CrossRef]
  6. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Wang, Y.; Xue, X.; Chen, R.; Yang, S.; Jin, Z. Progress and Perspective of Electrocatalytic CO2 Reduction for Renewable Carbonaceous Fuels and Chemicals. Adv. Sci. 2018, 5, 1700275. [Google Scholar] [CrossRef]
  7. Jin, S.; Hao, Z.; Zhang, K.; Yan, Z.; Chen, J. Advances and Challenges for the Electrochemical Reduction of CO2 to CO: From Fundamentals to Industrialization. Angew. Chem. Int. Ed. 2021, 60, 20627–20648. [Google Scholar] [CrossRef]
  8. Hoang, V.C.; Gomes, V.G.; Kornienko, N. Metal-Based Nanomaterials for Efficient CO2 Electroreduction: Recent Advances in Mechanism, Material Design and Selectivity. Nano Energy 2020, 78, 105311. [Google Scholar] [CrossRef]
  9. Jones, J.-P.; Prakash, G.K.S.; Olah, G.A. Electrochemical CO2 Reduction: Recent Advances and Current Trends. Isr. J. Chem. 2014, 54, 1451–1466. [Google Scholar] [CrossRef]
  10. 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, 1908398. [Google Scholar] [CrossRef]
  11. Han, N.; Ding, P.; He, L.; Li, Y.; Li, Y. Promises of Main Group Metal–Based Nanostructured Materials for Electrochemical CO2 Reduction to Formate. Adv. Energy Mater. 2020, 10, 1902338. [Google Scholar] [CrossRef]
  12. 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, 1833–1839. [Google Scholar] [CrossRef]
  13. Feaster, J.T.; Shi, C.; Cave, E.R.; Hatsukade, T.; Abram, D.N.; Kuhl, K.P.; Hahn, C.; Nørskov, J.K.; Jaramillo, T.F. Understanding Selectivity for the Electrochemical Reduction of Carbon Dioxide to Formic Acid and Carbon Monoxide on Metal Electrodes. ACS Catal. 2017, 7, 4822–4827. [Google Scholar] [CrossRef]
  14. Hori, Y.; Takahashi, R.; Yoshinami, Y.; Murata, A. Electrochemical Reduction of CO at a Copper Electrode. J. Phys. Chem. B 1997, 101, 7075–7081. [Google Scholar] [CrossRef]
  15. Goodpaster, J.D.; Bell, A.T.; Head-Gordon, M. Identification of Possible Pathways for C–C Bond Formation during Electrochemical Reduction of CO2: New Theoretical Insights from an Improved Electrochemical Model. J. Phys. Chem. Lett. 2016, 7, 1471–1477. [Google Scholar] [CrossRef] [PubMed]
  16. Zheng, Y.; Vasileff, A.; Zhou, X.; Jiao, Y.; Jaroniec, M.; Qiao, S.-Z. Understanding the Roadmap for Electrochemical Reduction of CO2 to Multi-Carbon Oxygenates and Hydrocarbons on Copper-Based Catalysts. J. Am. Chem. Soc. 2019, 141, 7646–7659. [Google Scholar] [CrossRef]
  17. Garza, A.J.; Bell, A.T.; Head-Gordon, M. Mechanism of CO2 Reduction at Copper Surfaces: Pathways to C2 Products. ACS Catal. 2018, 8, 1490–1499. [Google Scholar] [CrossRef]
  18. Schouten, K.J.P.; Kwon, Y.; van der Ham, C.J.M.; Qin, Z.; Koper, M.T.M. A New Mechanism for the Selectivity to C1 and C2 Species in the Electrochemical Reduction of Carbon Dioxide on Copper Electrodes. Chem. Sci. 2011, 2, 1902–1909. [Google Scholar] [CrossRef]
  19. Varandili, S.B.; Stoian, D.; Vavra, J.; Rossi, K.; Pankhurst, J.R.; Guntern, Y.T.; López, N.; Buonsanti, R. Elucidating the Structure-Dependent Selectivity of CuZn towards Methane and Ethanol in CO2 Electroreduction Using Tailored Cu/ZnO Precatalysts. Chem. Sci. 2021, 12, 14484–14493. [Google Scholar] [CrossRef]
  20. Iyengar, P.; Kolb, M.J.; Pankhurst, J.R.; Calle-Vallejo, F.; Buonsanti, R. Elucidating the Facet-Dependent Selectivity for CO2 Electroreduction to Ethanol of Cu–Ag Tandem Catalysts. ACS Catal. 2021, 11, 4456–4463. [Google Scholar] [CrossRef]
  21. Pei, Y.; Zhong, H.; Jin, F. A Brief Review of Electrocatalytic Reduction of CO2—Materials, Reaction Conditions, and Devices. Energy Sci. Eng. 2021, 9, 1012–1032. [Google Scholar] [CrossRef]
  22. Yi, J.; Li, Q.; Chi, S.; Huang, Y.; Cao, R. Boron-Doped Covalent Triazine Framework for Efficient CO2 Electroreduction. Chem. Res. Chin. Univ. 2022, 38, 141–146. [Google Scholar] [CrossRef]
  23. Costamagna, J.A.; Isaacs, M.; Aguirre, M.J.; Ramírez, G.; Azocar, I. Electroreduction of CO2 Catalyzed by Metallomacrocyclics BT-N4-Macrocyclic Metal Complexes; Zagal, J.H., Bedioui, F., Dodelet, J.-P., Eds.; Springer New York: New York, NY, USA, 2006; pp. 191–254. ISBN 978-0-387-28430-9. [Google Scholar]
  24. Bevilacqua, M.; Filippi, J.; Miller, H.A.; Vizza, F. Recent Technological Progress in CO2 Electroreduction to Fuels and Energy Carriers in Aqueous Environments. Energy Technol. 2015, 3, 197–210. [Google Scholar] [CrossRef]
  25. Ye, W.; Guo, X.; Ma, T. A Review on Electrochemical Synthesized Copper-Based Catalysts for Electrochemical Reduction of CO2 to C2+ Products. Chem. Eng. J. 2021, 414, 128825. [Google Scholar] [CrossRef]
  26. Bao, K.; Shi, J.; Liao, F.; Huang, H.; Liu, Y.; Kang, Z. The Advance and Critical Functions of Energetic Carbon Dots in Carbon Dioxide Photo/Electroreduction Reactions. Small Methods 2022, 6, 2200914. [Google Scholar] [CrossRef]
  27. Al-Tamreh, S.A.; Ibrahim, M.H.; El-Naas, M.H.; Vaes, J.; Pant, D.; Benamor, A.; Amhamed, A. Electroreduction of Carbon Dioxide into Formate: A Comprehensive Review. ChemElectroChem 2021, 8, 3207–3220. [Google Scholar] [CrossRef]
  28. Feng, D.-M.; Zhu, Y.-P.; Chen, P.; Ma, T.-Y. Recent Advances in Transition-Metal-Mediated Electrocatalytic CO2 Reduction: From Homogeneous to Heterogeneous Systems. Catalysts 2017, 7, 373. [Google Scholar] [CrossRef]
  29. Hou, M.; xia Shi, Y.; jun Li, J.; Gao, Z.; Zhang, Z. Cu-Based Organic-Inorganic Composite Materials for Electrochemical CO2 Reduction. Chem. Asian J. 2022, 17, e202200624. [Google Scholar] [CrossRef] [PubMed]
  30. Lü, F.; Bao, H.; Mi, Y.; Liu, Y.; Sun, J.; Peng, X.; Qiu, Y.; Zhuo, L.; Liu, X.; Luo, J. Electrochemical CO2 Reduction: From Nanoclusters to Single Atom Catalysts. Sustain. Energy Fuels 2020, 4, 1012–1028. [Google Scholar] [CrossRef]
  31. Wang, G.; Chen, J.; Ding, Y.; Cai, P.; Yi, L.; Li, Y.; Tu, C.; Hou, Y.; Wen, Z.; Dai, L. Electrocatalysis for CO2 Conversion: From Fundamentals to Value-Added Products. Chem. Soc. Rev. 2021, 50, 4993–5061. [Google Scholar] [CrossRef]
  32. Larrazábal, G.O.; Strøm-Hansen, P.; Heli, J.P.; Zeiter, K.; Therkildsen, K.T.; Chorkendorff, I.; Seger, B. Analysis of Mass Flows and Membrane Cross-over in CO2 Reduction at High Current Densities in an MEA-Type Electrolyzer. ACS Appl. Mater. Interfaces 2019, 11, 41281–41288. [Google Scholar] [CrossRef]
  33. 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. [Google Scholar] [CrossRef]
  34. Gawel, A.; Jaster, T.; Siegmund, D.; Holzmann, J.; Lohmann, H.; Klemm, E.; Apfel, U.-P. Electrochemical CO2 Reduction—The Macroscopic World of Electrode Design, Reactor Concepts & Economic Aspects. IScience 2022, 25, 104011. [Google Scholar] [CrossRef]
  35. Jeanty, P.; Scherer, C.; Magori, E.; Wiesner-Fleischer, K.; Hinrichsen, O.; Fleischer, M. Upscaling and Continuous Operation of Electrochemical CO2 to CO Conversion in Aqueous Solutions on Silver Gas Diffusion Electrodes. J. CO2 Util. 2018, 24, 454–462. [Google Scholar] [CrossRef]
  36. Verma, S.; Hamasaki, Y.; Kim, C.; Huang, W.; Lu, S.; Jhong, H.-R.M.; Gewirth, A.A.; Fujigaya, T.; Nakashima, N.; Kenis, P.J.A. Insights into the Low Overpotential Electroreduction of CO2 to CO on a Supported Gold Catalyst in an Alkaline Flow Electrolyzer. ACS Energy Lett. 2018, 3, 193–198. [Google Scholar] [CrossRef]
  37. Oloman, C.; Li, H. Electrochemical Processing of Carbon Dioxide. ChemSusChem 2008, 1, 385–391. [Google Scholar] [CrossRef] [PubMed]
  38. Subramanian, K.; Asokan, K.; Jeevarathinam, D.; Chandrasekaran, M. Electrochemical Membrane Reactor for the Reduction of Carbondioxide to Formate. J. Appl. Electrochem. 2007, 37, 255–260. [Google Scholar] [CrossRef]
  39. Ai, L.; Ng, S.-F.; Ong, W.-J. A Prospective Life Cycle Assessment of Electrochemical CO2 Reduction to Selective Formic Acid and Ethylene. ChemSusChem 2022, 15, e202200857. [Google Scholar] [CrossRef] [PubMed]
  40. Khoo, H.H.; Halim, I.; Handoko, A.D. LCA of Electrochemical Reduction of CO2 to Ethylene. J. CO2 Util. 2020, 41, 101229. [Google Scholar] [CrossRef]
  41. Yang, H.; Han, N.; Deng, J.; Wu, J.; Wang, Y.; Hu, Y.; Ding, P.; Li, Y.; Li, Y.; Lu, J. Selective CO2 Reduction on 2D Mesoporous Bi Nanosheets. Adv. Energy Mater. 2018, 8, 1801536. [Google Scholar] [CrossRef]
  42. Gong, Q.; Ding, P.; Xu, M.; Zhu, X.; Wang, M.; Deng, J.; Ma, Q.; Han, N.; Zhu, Y.; Lu, J.; et al. Structural Defects on Converted Bismuth Oxide Nanotubes Enable Highly Active Electrocatalysis of Carbon Dioxide Reduction. Nat. Commun. 2019, 10, 2807. [Google Scholar] [CrossRef] [PubMed]
  43. Han, N.; Wang, Y.; Yang, H.; Deng, J.; Wu, J.; Li, Y.; Li, Y. Ultrathin Bismuth Nanosheets from in Situ Topotactic Transformation for Selective Electrocatalytic CO2 Reduction to Formate. Nat. Commun. 2018, 9, 1320. [Google Scholar] [CrossRef] [PubMed]
  44. Han, N.; Wang, Y.; Ma, L.; Wen, J.; Li, J.; Zheng, H.; Nie, K.; Wang, X.; Zhao, F.; Li, Y.; et al. Supported Cobalt Polyphthalocyanine for High-Performance Electrocatalytic CO2 Reduction. Chem 2017, 3, 652–664. [Google Scholar] [CrossRef]
  45. Ren, D.; Deng, Y.; Handoko, A.D.; Chen, C.S.; Malkhandi, S.; Yeo, B.S. Selective Electrochemical Reduction of Carbon Dioxide to Ethylene and Ethanol on Copper(I) Oxide Catalysts. ACS Catal. 2015, 5, 2814–2821. [Google Scholar] [CrossRef]
  46. Reller, C.; Krause, R.; Volkova, E.; Schmid, B.; Neubauer, S.; Rucki, A.; Schuster, M.; Schmid, G. Selective Electroreduction of CO2 toward Ethylene on Nano Dendritic Copper Catalysts at High Current Density. Adv. Energy Mater. 2017, 7, 1602114. [Google Scholar] [CrossRef]
  47. 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. [Google Scholar] [CrossRef]
  48. 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. [Google Scholar] [CrossRef]
  49. Liu, Y.; Chen, S.; Quan, X.; Yu, H. Efficient Electrochemical Reduction of Carbon Dioxide to Acetate on Nitrogen-Doped Nanodiamond. J. Am. Chem. Soc. 2015, 137, 11631–11636. [Google Scholar] [CrossRef]
  50. Wang, Y.; Wang, D.; Dares, C.J.; Marquard, S.L.; Sheridan, M.V.; Meyer, T.J. CO2 Reduction to Acetate in Mixtures of Ultrasmall (Cu)n,(Ag)m Bimetallic Nanoparticles. Proc. Natl. Acad. Sci. USA 2018, 115, 278–283. [Google Scholar] [CrossRef]
  51. Ren, D.; Wong, N.T.; Handoko, A.D.; Huang, Y.; Yeo, B.S. Mechanistic Insights into the Enhanced Activity and Stability of Agglomerated Cu Nanocrystals for the Electrochemical Reduction of Carbon Dioxide to N-Propanol. J. Phys. Chem. Lett. 2016, 7, 20–24. [Google Scholar] [CrossRef]
  52. Wang, R.; Kapteijn, F.; Gascon, J. Engineering Metal–Organic Frameworks for the Electrochemical Reduction of CO2: A Minireview. Chem. Asian J. 2019, 14, 3452–3461. [Google Scholar] [CrossRef] [PubMed]
  53. Yang, K.; Kas, R.; Smith, W.A.; Burdyny, T. Role of the Carbon-Based Gas Diffusion Layer on Flooding in a Gas Diffusion Electrode Cell for Electrochemical CO2 Reduction. ACS Energy Lett. 2021, 6, 33–40. [Google Scholar] [CrossRef]
  54. Rabiee, H.; Ge, L.; Zhang, X.; Hu, S.; Li, M.; Yuan, Z. Gas Diffusion Electrodes (GDEs) for Electrochemical Reduction of Carbon Dioxide, Carbon Monoxide, and Dinitrogen to Value-Added Products: A Review. Energy Environ. Sci. 2021, 14, 1959–2008. [Google Scholar] [CrossRef]
  55. Lu, Q.; Jiao, F. Electrochemical CO2 Reduction: Electrocatalyst, Reaction Mechanism, and Process Engineering. Nano Energy 2016, 29, 439–456. [Google Scholar] [CrossRef]
  56. Kim, C.; Dionigi, F.; Beermann, V.; Wang, X.; Möller, T.; Strasser, P. Alloy Nanocatalysts for the Electrochemical Oxygen Reduction (ORR) and the Direct Electrochemical Carbon Dioxide Reduction Reaction (CO2RR). Adv. Mater. 2019, 31, 1805617. [Google Scholar] [CrossRef]
  57. He, J.; Huang, A.; Johnson, N.J.J.; Dettelbach, K.E.; Weekes, D.M.; Cao, Y.; Berlinguette, C.P. Stabilizing Copper for CO2 Reduction in Low-Grade Electrolyte. Inorg. Chem. 2018, 57, 14624–14631. [Google Scholar] [CrossRef] [PubMed]
  58. Mistry, H.; Varela, A.S.; Kühl, S.; Strasser, P.; Cuenya, B.R. Nanostructured Electrocatalysts with Tunable Activity and Selectivity. Nat. Rev. Mater. 2016, 1, 16009. [Google Scholar] [CrossRef]
  59. Prabhu, P.; Lee, J.-M. Metallenes as Functional Materials in Electrocatalysis. Chem. Soc. Rev. 2021, 50, 6700–6719. [Google Scholar] [CrossRef]
  60. Chen, H.; Liang, X.; Liu, Y.; Ai, X.; Asefa, T.; Zou, X. Active Site Engineering in Porous Electrocatalysts. Adv. Mater. 2020, 32, 2002435. [Google Scholar] [CrossRef]
  61. Rasul, S.; Pugnant, A.; Xiang, H.; Fontmorin, J.-M.; Yu, E.H. Low Cost and Efficient Alloy Electrocatalysts for CO2 Reduction to Formate. J. CO2 Util. 2019, 32, 1–10. [Google Scholar] [CrossRef]
  62. Huang, S.; Meng, Y.; Cao, Y.; He, S.; Li, X.; Tong, S.; Wu, M. N-, O- and P-Doped Hollow Carbons: Metal-Free Bifunctional Electrocatalysts for Hydrogen Evolution and Oxygen Reduction Reactions. Appl. Catal. B Environ. 2019, 248, 239–248. [Google Scholar] [CrossRef]
  63. Jia, Y.; Jiang, K.; Wang, H.; Yao, X. The Role of Defect Sites in Nanomaterials for Electrocatalytic Energy Conversion. Chem 2019, 5, 1371–1397. [Google Scholar] [CrossRef]
  64. Cabrera, C.R.; Abruña, H.D. Electrocatalysis of CO2 Reduction at Surface Modified Metallic and Semiconducting Electrodes. J. Electroanal. Chem. Interfacial Electrochem. 1986, 209, 101–107. [Google Scholar] [CrossRef]
  65. Sun, S.; Li, H.; Xu, Z.J. Impact of Surface Area in Evaluation of Catalyst Activity. Joule 2018, 2, 1024–1027. [Google Scholar] [CrossRef]
  66. Guo, S.-X.; Liu, Y.; Lee, C.-Y.; Bond, A.M.; Zhang, J.; Geletii, Y.V.; Hill, C.L. Graphene-Supported [{Ru4O4(OH)2(H2O)4}(γ-SiW10O36)2]10− for Highly Efficient Electrocatalytic Water Oxidation. Energy Environ. Sci. 2013, 6, 2654–2663. [Google Scholar] [CrossRef]
  67. Dunwell, M.; Yan, Y.; Xu, B. Understanding the Influence of the Electrochemical Double-Layer on Heterogeneous Electrochemical Reactions. Curr. Opin. Chem. Eng. 2018, 20, 151–158. [Google Scholar] [CrossRef]
  68. Wang, Y.; Han, P.; Lv, X.; Zhang, L.; Zheng, G. Defect and Interface Engineering for Aqueous Electrocatalytic CO2 Reduction. Joule 2018, 2, 2551–2582. [Google Scholar] [CrossRef]
  69. 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, 9337–9340. [Google Scholar] [CrossRef]
  70. Zhu, D.D.; Liu, J.L.; Qiao, S.Z. Recent Advances in Inorganic Heterogeneous Electrocatalysts for Reduction of Carbon Dioxide. Adv. Mater. 2016, 28, 3423–3452. [Google Scholar] [CrossRef] [PubMed]
  71. 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. [Google Scholar] [CrossRef] [PubMed]
  72. Lee, S.Y.; Jung, H.; Kim, N.-K.; Oh, H.-S.; Min, B.K.; Hwang, Y.J. Mixed Copper States in Anodized Cu Electrocatalyst for Stable and Selective Ethylene Production from CO2 Reduction. J. Am. Chem. Soc. 2018, 140, 8681–8689. [Google Scholar] [CrossRef]
  73. Yan, D.; Li, Y.; Huo, J.; Chen, R.; Dai, L.; Wang, S. Defect Chemistry of Nonprecious-Metal Electrocatalysts for Oxygen Reactions. Adv. Mater. 2017, 29, 1606459. [Google Scholar] [CrossRef]
  74. Jia, J.; Qian, C.; Dong, Y.; Li, Y.F.; Wang, H.; Ghoussoub, M.; Butler, K.T.; Walsh, A.; Ozin, G.A. Heterogeneous Catalytic Hydrogenation of CO2 by Metal Oxides: Defect Engineering—Perfecting Imperfection. Chem. Soc. Rev. 2017, 46, 4631–4644. [Google Scholar] [CrossRef]
  75. Eckmann, A.; Felten, A.; Mishchenko, A.; Britnell, L.; Krupke, R.; Novoselov, K.S.; Casiraghi, C. Probing the Nature of Defects in Graphene by Raman Spectroscopy. Nano Lett. 2012, 12, 3925–3930. [Google Scholar] [CrossRef]
  76. Estrade-Szwarckopf, H. XPS Photoemission in Carbonaceous Materials: A “Defect” Peak beside the Graphitic Asymmetric Peak. Carbon 2004, 42, 1713–1721. [Google Scholar] [CrossRef]
  77. Wang, M.; Árnadóttir, L.; Xu, Z.J.; Feng, Z. In Situ X-Ray Absorption Spectroscopy Studies of Nanoscale Electrocatalysts. Nano-Micro Lett. 2019, 11, 47. [Google Scholar] [CrossRef]
  78. Stevens, K.T.; Halliburton, L.E.; Setzler, S.D.; Schunemann, P.G.; Pollak, T.M. Electron Paramagnetic Resonance and Electron-Nuclear Double Resonance Study of the Neutral Copper Acceptor in ZnGeP2 Crystals. J. Phys. Condens. Matter 2003, 15, 1625. [Google Scholar] [CrossRef]
  79. Hemraj-Benny, T.; Banerjee, S.; Sambasivan, S.; Balasubramanian, M.; Fischer, D.A.; Eres, G.; Puretzky, A.A.; Geohegan, D.B.; Lowndes, D.H.; Han, W.; et al. Near-Edge X-Ray Absorption Fine Structure Spectroscopy as a Tool for Investigating Nanomaterials. Small 2006, 2, 26–35. [Google Scholar] [CrossRef] [PubMed]
  80. Zang, Y.; Liu, T.; Li, H.; Wei, P.; Song, Y.; Cheng, C.; Gao, D.; Song, Y.; Wang, G.; Bao, X. In Situ Reconstruction of Defect-Rich SnO2 through an Analogous Disproportionation Process for CO2 Electroreduction. Chem. Eng. J. 2022, 446, 137444. [Google Scholar] [CrossRef]
  81. Domínguez-Gutiérrez, F.J.; Byggmästar, J.; Nordlund, K.; Djurabekova, F.; von Toussaint, U. Computational Study of Crystal Defect Formation in Mo by a Machine Learning Molecular Dynamics Potential. Model. Simul. Mater. Sci. Eng. 2021, 29, 55001. [Google Scholar] [CrossRef]
  82. Tang, C.; Zhang, Q. Nanocarbon for Oxygen Reduction Electrocatalysis: Dopants, Edges, and Defects. Adv. Mater. 2017, 29, 1604103. [Google Scholar] [CrossRef] [PubMed]
  83. Moshe, M.; Levin, I.; Aharoni, H.; Kupferman, R.; Sharon, E. Geometry and Mechanics of Two-Dimensional Defects in Amorphous Materials. Proc. Natl. Acad. Sci. USA 2015, 112, 10873–10878. [Google Scholar] [CrossRef]
  84. Sun, Z.; Ma, T.; Tao, H.; Fan, Q.; Han, B. Fundamentals and Challenges of Electrochemical CO2 Reduction Using Two-Dimensional Materials. Chem 2017, 3, 560–587. [Google Scholar] [CrossRef]
  85. Sharma, P.P.; Wu, J.; Yadav, R.M.; Liu, M.; Wright, C.J.; Tiwary, C.S.; Yakobson, B.I.; Lou, J.; Ajayan, P.M.; Zhou, X.-D. Nitrogen-Doped Carbon Nanotube Arrays for High-Efficiency Electrochemical Reduction of CO2: On the Understanding of Defects, Defect Density, and Selectivity. Angew. Chem. Int. Ed. 2015, 54, 13701–13705. [Google Scholar] [CrossRef] [PubMed]
  86. 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. [Google Scholar] [CrossRef] [PubMed]
  87. 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. [Google Scholar] [CrossRef]
  88. Mistry, H.; Choi, Y.; Bagger, A.; Scholten, F.; Bonifacio, C.S.; Sinev, I.; Divins, N.J.; Zegkinoglou, I.; Jeon, H.S.; Kisslinger, K.; et al. Enhanced Carbon Dioxide Electroreduction to Carbon Monoxide over Defect-Rich Plasma-Activated Silver Catalysts. Angew. Chem. 2017, 129, 11552–11556. [Google Scholar] [CrossRef]
  89. Yu, X.; Han, P.; Wei, Z.; Huang, L.; Gu, Z.; Peng, S.; Ma, J.; Zheng, G. Boron-Doped Graphene for Electrocatalytic N2 Reduction. Joule 2018, 2, 1610–1622. [Google Scholar] [CrossRef]
  90. Gao, D.; Zhang, Y.; Zhou, Z.; Cai, F.; Zhao, X.; Huang, W.; Li, Y.; Zhu, J.; Liu, P.; Yang, F.; et al. Enhancing CO2 Electroreduction with the Metal–Oxide Interface. J. Am. Chem. Soc. 2017, 139, 5652–5655. [Google Scholar] [CrossRef]
  91. Zhang, W.; Jia, B.; Liu, X.; Ma, T. Surface and Interface Chemistry in Metal-Free Electrocatalysts for Electrochemical CO2 Reduction. SmartMat 2022, 3, 5–34. [Google Scholar] [CrossRef]
  92. Yang, K.D.; Ko, W.R.; Lee, J.H.; Kim, S.J.; Lee, H.; Lee, M.H.; Nam, K.T. Morphology-Directed Selective Production of Ethylene or Ethane from CO2 on a Cu Mesopore Electrode. Angew. Chem. 2017, 129, 814–818. [Google Scholar] [CrossRef]
  93. Zoubir, O.; Atourki, L.; Ait Ahsaine, H.; BaQais, A. Current State of Copper-Based Bimetallic Materials for Electrochemical CO2 Reduction: A Review. RSC Adv. 2022, 12, 30056–30075. [Google Scholar] [CrossRef]
  94. Hansen, H.A.; Shi, C.; Lausche, A.C.; Peterson, A.A.; Nørskov, J.K. Bifunctional Alloys for the Electroreduction of CO2 and CO. Phys. Chem. Chem. Phys. 2016, 18, 9194–9201. [Google Scholar] [CrossRef] [PubMed]
  95. Hirunsit, P.; Soodsawang, W.; Limtrakul, J. CO2 Electrochemical Reduction to Methane and Methanol on Copper-Based Alloys: Theoretical Insight. J. Phys. Chem. C 2015, 119, 8238–8249. [Google Scholar] [CrossRef]
  96. Kuhl, K.P.; Cave, E.R.; Abram, D.N.; Jaramillo, T.F. New Insights into the Electrochemical Reduction of Carbon Dioxide on Metallic Copper Surfaces. Energy Environ. Sci. 2012, 5, 7050–7059. [Google Scholar] [CrossRef]
  97. Hori, Y.; Kikuchi, K.; Murata, A.; Suzuki, S. Production of Methane and Ethylene in Electrochemical Reduction of Carbon Dioxide at Copper Electrode in Aqueous Hydrogencarbonate Solution. Chem. Lett. 1986, 15, 897–898. [Google Scholar] [CrossRef]
  98. Kuhl, K.P.; Hatsukade, T.; Cave, E.R.; Abram, D.N.; Kibsgaard, J.; Jaramillo, T.F. Electrocatalytic Conversion of Carbon Dioxide to Methane and Methanol on Transition Metal Surfaces. J. Am. Chem. Soc. 2014, 136, 14107–14113. [Google Scholar] [CrossRef]
  99. Ma, M.; Djanashvili, K.; Smith, W.A. Controllable Hydrocarbon Formation from the Electrochemical Reduction of CO2 over Cu Nanowire Arrays. Angew. Chem. Int. Ed. 2016, 55, 6680–6684. [Google Scholar] [CrossRef]
  100. Sen, S.; Liu, D.; Palmore, G.T.R. Electrochemical Reduction of CO2 at Copper Nanofoams. ACS Catal. 2014, 4, 3091–3095. [Google Scholar] [CrossRef]
  101. Wang, A.; Li, J.; Zhang, T. Heterogeneous Single-Atom Catalysis. Nat. Rev. Chem. 2018, 2, 65–81. [Google Scholar] [CrossRef]
  102. Li, M.; Wang, H.; Luo, W.; Sherrell, P.C.; Chen, J.; Yang, J. Heterogeneous Single-Atom Catalysts for Electrochemical CO2 Reduction Reaction. Adv. Mater. 2020, 32, 2001848. [Google Scholar] [CrossRef] [PubMed]
  103. Pander, J.E.I.I.I.; Baruch, M.F.; Bocarsly, A.B. Probing the Mechanism of Aqueous CO2 Reduction on Post-Transition-Metal Electrodes Using ATR-IR Spectroelectrochemistry. ACS Catal. 2016, 6, 7824–7833. [Google Scholar] [CrossRef]
  104. Bi, W.; Wu, C.; Xie, Y. Atomically Thin Two-Dimensional Solids: An Emerging Platform for CO2 Electroreduction. ACS Energy Lett. 2018, 3, 624–633. [Google Scholar] [CrossRef]
  105. Li, F.; MacFarlane, D.R.; Zhang, J. Recent Advances in the Nanoengineering of Electrocatalysts for CO2 Reduction. Nanoscale 2018, 10, 6235–6260. [Google Scholar] [CrossRef]
  106. Gao, D.; Cai, F.; Wang, G.; Bao, X. Nanostructured Heterogeneous Catalysts for Electrochemical Reduction of CO2. Curr. Opin. Green Sustain. Chem. 2017, 3, 39–44. [Google Scholar] [CrossRef]
  107. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How Copper Catalyzes the Electroreduction of Carbon Dioxide into Hydrocarbon Fuels. Energy Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
  108. Xiang, K.; Zhu, F.; Liu, Y.; Pan, Y.; Wang, X.; Yan, X.; Liu, H. A Strategy to Eliminate Carbon Deposition on a Copper Electrode in Order to Enhance Its Stability in CO2RR Catalysis by Introducing Crystal Defects. Electrochem. Commun. 2019, 102, 72–77. [Google Scholar] [CrossRef]
  109. Mi, Y.; Shen, S.; Peng, X.; Bao, H.; Liu, X.; Luo, J. Selective Electroreduction of CO2 to C2 Products over Cu3N-Derived Cu Nanowires. ChemElectroChem 2019, 6, 2393–2397. [Google Scholar] [CrossRef]
  110. Hoang, T.T.H.; Verma, S.; Ma, S.; Fister, T.T.; Timoshenko, J.; Frenkel, A.I.; Kenis, P.J.A.; Gewirth, A.A. Nanoporous Copper–Silver Alloys by Additive-Controlled Electrodeposition for the Selective Electroreduction of CO2 to Ethylene and Ethanol. J. Am. Chem. Soc. 2018, 140, 5791–5797. [Google Scholar] [CrossRef]
  111. Fu, X.; Zhang, P.; Sun, T.; Xu, L.; Gong, L.; Chen, B.; Xu, Q.; Zheng, T.; Yu, Z.; Chen, X.; et al. Atomically Dispersed Ni–N3 Sites on Highly Defective Micro-Mesoporous Carbon for Superior CO2 Electroreduction. Small 2022, 18, 2107997. [Google Scholar] [CrossRef]
  112. Zhang, J.; Sun, W.; Ding, L.; Wu, Z.; Gao, F. Au Nanocrystals@Defective Amorphous MnO2 Nanosheets Core/Shell Nanostructure with Effective CO2 Adsorption and Activation toward CO2 Electroreduction to CO. ACS Sustain. Chem. Eng. 2021, 9, 5230–5239. [Google Scholar] [CrossRef]
  113. Wu, S.; Yi, F.; Ping, D.; Huang, S.; Zhang, Y.; Han, L.; Wang, S.; Wang, H.; Yang, X.; Guo, D.; et al. Constructing Single-Atomic Nickel Sites in Carbon Nanotubes for Efficient CO2 Electroreduction. Carbon 2022, 196, 1–9. [Google Scholar] [CrossRef]
  114. Ali, S.; Yasin, G.; Iqbal, R.; Huang, X.; Su, J.; Ibraheem, S.; Zhang, Z.; Wu, X.; Wahid, F.; Ismail, P.M.; et al. Porous Aza-Doped Graphene-Analogous 2D Material a Unique Catalyst for CO2 Conversion to Formic-Acid by Hydrogenation and Electroreduction Approaches. Mol. Catal. 2022, 524, 112285. [Google Scholar] [CrossRef]
  115. Ni, W.; Liu, Z.; Zhang, Y.; Ma, C.; Deng, H.; Zhang, S.; Wang, S. Electroreduction of Carbon Dioxide Driven by the Intrinsic Defects in the Carbon Plane of a Single Fe–N4 Site. Adv. Mater. 2021, 33, 2003238. [Google Scholar] [CrossRef]
  116. Zhao, X.-H.; Chen, Q.-S.; Zhuo, D.-H.; Lu, J.; Xu, Z.-N.; Wang, C.-M.; Tang, J.-X.; Sun, S.-G.; Guo, G.-C. Oxygen Vacancies Enriched Bi Based Catalysts for Enhancing Electrocatalytic CO2 Reduction to Formate. Electrochim. Acta 2021, 367, 137478. [Google Scholar] [CrossRef]
  117. Zhang, Y.; Zhang, X.; Ling, Y.; Li, F.; Bond, A.M.; Zhang, J. Controllable Synthesis of Few-Layer Bismuth Subcarbonate by Electrochemical Exfoliation for Enhanced CO2 Reduction Performance. Angew. Chem. 2018, 130, 13467–13471. [Google Scholar] [CrossRef]
  118. Zhang, W.; Hu, Y.; Ma, L.; Zhu, G.; Zhao, P.; Xue, X.; Chen, R.; Yang, S.; Ma, J.; Liu, J.; et al. Liquid-Phase Exfoliated Ultrathin Bi Nanosheets: Uncovering the Origins of Enhanced Electrocatalytic CO2 Reduction on Two-Dimensional Metal Nanostructure. Nano Energy 2018, 53, 808–816. [Google Scholar] [CrossRef]
  119. Won, D.H.; Choi, C.H.; Chung, J.; Chung, M.W.; Kim, E.-H.; Woo, S.I. Rational Design of a Hierarchical Tin Dendrite Electrode for Efficient Electrochemical Reduction of CO2. ChemSusChem 2015, 8, 3092–3098. [Google Scholar] [CrossRef]
  120. Li, F.; Xue, M.; Li, J.; Ma, X.; Chen, L.; Zhang, X.; MacFarlane, D.R.; Zhang, J. Unlocking the Electrocatalytic Activity of Antimony for CO2 Reduction by Two-Dimensional Engineering of the Bulk Material. Angew. Chem. 2017, 129, 14910–14914. [Google Scholar] [CrossRef]
  121. Dutta, A.; Rahaman, M.; Luedi, N.C.; Mohos, M.; Broekmann, P. Morphology Matters: Tuning the Product Distribution of CO2 Electroreduction on Oxide-Derived Cu Foam Catalysts. ACS Catal. 2016, 6, 3804–3814. [Google Scholar] [CrossRef]
  122. Hori, Y.; Takahashi, I.; Koga, O.; Hoshi, N. Electrochemical Reduction of Carbon Dioxide at Various Series of Copper Single Crystal Electrodes. J. Mol. Catal. A Chem. 2003, 199, 39–47. [Google Scholar] [CrossRef]
  123. Chen, C.S.; Handoko, A.D.; Wan, J.H.; Ma, L.; Ren, D.; Yeo, B.S. Stable and Selective Electrochemical Reduction of Carbon Dioxide to Ethylene on Copper Mesocrystals. Catal. Sci. Technol. 2015, 5, 161–168. [Google Scholar] [CrossRef]
  124. Lv, W.; Zhou, J.; Bei, J.; Zhang, R.; Wang, L.; Xu, Q.; Wang, W. Electrodeposition of Nano-Sized Bismuth on Copper Foil as Electrocatalyst for Reduction of CO2 to Formate. Appl. Surf. Sci. 2017, 393, 191–196. [Google Scholar] [CrossRef]
  125. Bai, X.; Chen, W.; Zhao, C.; Li, S.; Song, Y.; Ge, R.; Wei, W.; Sun, Y. Exclusive Formation of Formic Acid from CO2 Electroreduction by a Tunable Pd-Sn Alloy. Angew. Chem. Int. Ed. 2017, 56, 12219–12223. [Google Scholar] [CrossRef] [PubMed]
  126. Choi, S.Y.; Jeong, S.K.; Kim, H.J.; Baek, I.-H.; Park, K.T. Electrochemical Reduction of Carbon Dioxide to Formate on Tin–Lead Alloys. ACS Sustain. Chem. Eng. 2016, 4, 1311–1318. [Google Scholar] [CrossRef]
  127. Luc, W.; Collins, C.; Wang, S.; Xin, H.; He, K.; Kang, Y.; Jiao, F. Ag–Sn Bimetallic Catalyst with a Core–Shell Structure for CO2 Reduction. J. Am. Chem. Soc. 2017, 139, 1885–1893. [Google Scholar] [CrossRef]
  128. Chen, K.; Zhang, X.; Williams, T.; Bourgeois, L.; MacFarlane, D.R. Electrochemical Reduction of CO2 on Core-Shell Cu/Au Nanostructure Arrays for Syngas Production. Electrochim. Acta 2017, 239, 84–89. [Google Scholar] [CrossRef]
  129. Yin, Z.; Gao, D.; Yao, S.; Zhao, B.; Cai, F.; Lin, L.; Tang, P.; Zhai, P.; Wang, G.; Ma, D.; et al. Highly Selective Palladium-Copper Bimetallic Electrocatalysts for the Electrochemical Reduction of CO2 to CO. Nano Energy 2016, 27, 35–43. [Google Scholar] [CrossRef]
  130. Watanabe, M.; Shibata, M.; Kato, A.; Azuma, M.; Sakata, T. Design of Alloy Electrocatalysts for CO2 Reduction: III. The Selective and Reversible Reduction of on Cu Alloy Electrodes. J. Electrochem. Soc. 1991, 138, 3382. [Google Scholar] [CrossRef]
  131. Rasul, S.; Anjum, D.H.; Jedidi, A.; Minenkov, Y.; Cavallo, L.; Takanabe, K. A Highly Selective Copper–Indium Bimetallic Electrocatalyst for the Electrochemical Reduction of Aqueous CO2 to CO. Angew. Chem. Int. Ed. 2015, 54, 2146–2150. [Google Scholar] [CrossRef]
  132. Ma, S.; Sadakiyo, M.; Heima, M.; Luo, R.; Haasch, R.T.; Gold, J.I.; Yamauchi, M.; Kenis, P.J.A. Electroreduction of Carbon Dioxide to Hydrocarbons Using Bimetallic Cu–Pd Catalysts with Different Mixing Patterns. J. Am. Chem. Soc. 2017, 139, 47–50. [Google Scholar] [CrossRef]
  133. Xu, Z.; Lai, E.; Shao-Horn, Y.; Hamad-Schifferli, K. Compositional Dependence of the Stability of AuCu Alloy Nanoparticles. Chem. Commun. 2012, 48, 5626–5628. [Google Scholar] [CrossRef]
  134. Tang, H.; Liu, Y.; Zhou, Y.; Qian, Y.; Lin, B.-L. Boosting the Electroreduction of CO2 to Ethanol via the Synergistic Effect of Cu–Ag Bimetallic Catalysts. ACS Appl. Energy Mater. 2022, 5, 14045–14052. [Google Scholar] [CrossRef]
  135. Ma, X.; Shen, Y.; Yao, S.; An, C.; Zhang, W.; Zhu, J.; Si, R.; Guo, C.; An, C. Core–Shell Nanoporous AuCu3@Au Monolithic Electrode for Efficient Electrochemical CO2 Reduction. J. Mater. Chem. A 2020, 8, 3344–3350. [Google Scholar] [CrossRef]
  136. Zhu, S.; Qin, X.; Wang, Q.; Li, T.; Tao, R.; Gu, M.; Shao, M. Composition-Dependent CO2 Electrochemical Reduction Activity and Selectivity on Au–Pd Core–Shell Nanoparticles. J. Mater. Chem. A 2019, 7, 16954–16961. [Google Scholar] [CrossRef]
  137. Liu, F.; Wu, C.; Yang, S. Strain and Ligand Effects on CO2 Reduction Reactions over Cu–Metal Heterostructure Catalysts. J. Phys. Chem. C 2017, 121, 22139–22146. [Google Scholar] [CrossRef]
  138. Shan, J.; Sun, K.; Li, H.; Xu, P.; Sun, J.; Wang, Z. Composition Regulation and Defects Introduction via Amorphous CuEu Alloy Shell for Efficient CO2 Electroreduction toward Methane. J. CO2 Util. 2020, 41, 101285. [Google Scholar] [CrossRef]
  139. Xing, Y.; Kong, X.; Guo, X.; Liu, Y.; Li, Q.; Zhang, Y.; Sheng, Y.; Yang, X.; Geng, Z.; Zeng, J. Bi@Sn Core–Shell Structure with Compressive Strain Boosts the Electroreduction of CO2 into Formic Acid. Adv. Sci. 2020, 7, 1902989. [Google Scholar] [CrossRef]
  140. Hou, X.; Cai, Y.; Zhang, D.; Li, L.; Zhang, X.; Zhu, Z.; Peng, L.; Liu, Y.; Qiao, J. 3D Core–Shell Porous-Structured Cu@Sn Hybrid Electrodes with Unprecedented Selective CO2-into-Formate Electroreduction Achieving 100%. J. Mater. Chem. A 2019, 7, 3197–3205. [Google Scholar] [CrossRef]
  141. Feng, X.; Jiang, K.; Fan, S.; Kanan, M.W. Grain-Boundary-Dependent CO2 Electroreduction Activity. J. Am. Chem. Soc. 2015, 137, 4606–4609. [Google Scholar] [CrossRef] [PubMed]
  142. Chen, Z.; Wang, T.; Liu, B.; Cheng, D.; Hu, C.; Zhang, G.; Zhu, W.; Wang, H.; Zhao, Z.-J.; Gong, J. Grain-Boundary-Rich Copper for Efficient Solar-Driven Electrochemical CO2 Reduction to Ethylene and Ethanol. J. Am. Chem. Soc. 2020, 142, 6878–6883. [Google Scholar] [CrossRef] [PubMed]
  143. Chen, J.; Chen, S.; Li, Y.; Liao, X.; Zhao, T.; Cheng, F.; Wang, H. Galvanic-Cell Deposition Enables the Exposure of Bismuth Grain Boundary for Efficient Electroreduction of Carbon Dioxide. Small 2022, 18, 2201633. [Google Scholar] [CrossRef] [PubMed]
  144. Chen, Y.; Li, C.W.; Kanan, M.W. Aqueous CO2 Reduction at Very Low Overpotential on Oxide-Derived Au Nanoparticles. J. Am. Chem. Soc. 2012, 134, 19969–19972. [Google Scholar] [CrossRef] [PubMed]
  145. Singh, C.; Mukhopadhyay, S.; Hod, I. Metal–Organic Framework Derived Nanomaterials for Electrocatalysis: Recent Developments for CO2 and N2 Reduction. Nano Converg. 2021, 8, 1. [Google Scholar] [CrossRef]
  146. Keerthiga, G.; Chetty, R. Electrochemical Reduction of Carbon Dioxide on Zinc-Modified Copper Electrodes. J. Electrochem. Soc. 2017, 164, H164. [Google Scholar] [CrossRef]
  147. Gattrell, M.; Gupta, N.; Co, A. A Review of the Aqueous Electrochemical Reduction of CO2 to Hydrocarbons at Copper. J. Electroanal. Chem. 2006, 594, 1–19. [Google Scholar] [CrossRef]
  148. Centi, G.; Perathoner, S.; Winè, G.; Gangeri, M. Electrocatalytic Conversion of CO2 to Long Carbon-Chain Hydrocarbons. Green Chem. 2007, 9, 671–678. [Google Scholar] [CrossRef]
  149. Li, H.; Oloman, C. Development of a Continuous Reactor for the Electro-Reduction of Carbon Dioxide to Formate—Part 2: Scale-Up. J. Appl. Electrochem. 2007, 37, 1107–1117. [Google Scholar] [CrossRef]
  150. Kwon, Y.; Lee, J. Formic Acid from Carbon Dioxide on Nanolayered Electrocatalyst. Electrocatalysis 2010, 1, 108–115. [Google Scholar] [CrossRef]
  151. Ye, K.; Zhou, Z.; Shao, J.; Lin, L.; Gao, D.; Ta, N.; Si, R.; Wang, G.; Bao, X. In Situ Reconstruction of a Hierarchical Sn-Cu/SnOx Core/Shell Catalyst for High-Performance CO2 Electroreduction. Angew. Chem. Int. Ed. 2020, 59, 4814–4821. [Google Scholar] [CrossRef]
  152. Birdja, Y.Y.; Vos, R.E.; Wezendonk, T.A.; Jiang, L.; Kapteijn, F.; Koper, M.T.M. Effects of Substrate and Polymer Encapsulation on CO2 Electroreduction by Immobilized Indium(III) Protoporphyrin. ACS Catal. 2018, 8, 4420–4428. [Google Scholar] [CrossRef]
  153. Lee, C.H.; Kanan, M.W. Controlling H+ vs. CO2 Reduction Selectivity on Pb Electrodes. ACS Catal. 2015, 5, 465–469. [Google Scholar] [CrossRef]
  154. Gao, S.; Jiao, X.; Sun, Z.; Zhang, W.; Sun, Y.; Wang, C.; Hu, Q.; Zu, X.; Yang, F.; Yang, S.; et al. Ultrathin Co3O4 Layers Realizing Optimized CO2 Electroreduction to Formate. Angew. Chem. Int. Ed. 2016, 55, 698–702. [Google Scholar] [CrossRef]
  155. Noda, H.; Ikeda, S.; Oda, Y.; Imai, K.; Maeda, M.; Ito, K. Electrochemical Reduction of Carbon Dioxide at Various Metal Electrodes in Aqueous Potassium Hydrogen Carbonate Solution. Bull. Chem. Soc. Jpn. 1990, 63, 2459–2462. [Google Scholar] [CrossRef]
  156. Christophe, J.; Doneux, T.; Buess-Herman, C. Electroreduction of Carbon Dioxide on Copper-Based Electrodes: Activity of Copper Single Crystals and Copper–Gold Alloys. Electrocatalysis 2012, 3, 139–146. [Google Scholar] [CrossRef]
  157. Imani, R.; Qiu, Z.; Younesi, R.; Pazoki, M.; Fernandes, D.L.A.; Mitev, P.D.; Edvinsson, T.; Tian, H. Unravelling In-Situ Formation of Highly Active Mixed Metal Oxide CuInO2 Nanoparticles during CO2 Electroreduction. Nano Energy 2018, 49, 40–50. [Google Scholar] [CrossRef]
  158. Zhang, S.; Kang, P.; Meyer, T.J. Nanostructured Tin Catalysts for Selective Electrochemical Reduction of Carbon Dioxide to Formate. J. Am. Chem. Soc. 2014, 136, 1734–1737. [Google Scholar] [CrossRef]
  159. Kumar, B.; Atla, V.; Brian, J.P.; Kumari, S.; Nguyen, T.Q.; Sunkara, M.; Spurgeon, J.M. Reduced SnO2 Porous Nanowires with a High Density of Grain Boundaries as Catalysts for Efficient Electrochemical CO2-into-HCOOH Conversion. Angew. Chem. Int. Ed. 2017, 56, 3645–3649. [Google Scholar] [CrossRef] [PubMed]
  160. Li, F.; Chen, L.; Knowles, G.P.; MacFarlane, D.R.; Zhang, J. Hierarchical Mesoporous SnO(2) Nanosheets on Carbon Cloth: A Robust and Flexible Electrocatalyst for CO(2) Reduction with High Efficiency and Selectivity. Angew. Chem. Int. Ed. Engl. 2017, 56, 505–509. [Google Scholar] [CrossRef]
  161. Chen, Z.; Wu, R.; Wang, H.; Zhang, K.H.L.; Song, Y.; Wu, F.; Fang, F.; Sun, D. Embedding ZnSe Nanodots in Nitrogen-Doped Hollow Carbon Architectures for Superior Lithium Storage. Nano Res. 2018, 11, 966–978. [Google Scholar] [CrossRef]
  162. Cao, Z.; Yin, Y.; Fu, P.; Li, D.; Zhou, Y.; Deng, Y.; Peng, Y.; Wang, W.; Zhou, W.; Tang, D. TiO2 Nanosheet Arrays with Layered SnS2 and CoOx Nanoparticles for Efficient Photoelectrochemical Water Splitting. Nanoscale Res. Lett. 2019, 14, 342. [Google Scholar] [CrossRef] [PubMed]
  163. Ohkubo, K.; Takahashi, H.; Watters, E.P.J.; Taguchi, M. In-Situ Analysis of CO2 Electroreduction on Pt and Pt Oxide Cathodes. Electrochemistry 2020, 88, 210–217. [Google Scholar] [CrossRef]
  164. Wang, J.; Wang, G.; Zhang, J.; Wang, Y.; Wu, H.; Zheng, X.; Ding, J.; Han, X.; Deng, Y.; Hu, W. Inversely Tuning the CO2 Electroreduction and Hydrogen Evolution Activity on Metal Oxide via Heteroatom Doping. Angew. Chem. Int. Ed. 2021, 60, 7602–7606. [Google Scholar] [CrossRef]
  165. Zhong, X.; Liang, S.; Yang, T.; Zeng, G.; Zhong, Z.; Deng, H.; Zhang, L.; Sun, X. Sn Dopants with Synergistic Oxygen Vacancies Boost CO2 Electroreduction on CuO Nanosheets to CO at Low Overpotential. ACS Nano 2022, 16, 19210–19219. [Google Scholar] [CrossRef]
  166. Zhang, J.; Yin, R.; Shao, Q.; Zhu, T.; Huang, X. Oxygen Vacancies in Amorphous InOx Nanoribbons Enhance CO2 Adsorption and Activation for CO2 Electroreduction. Angew. Chem. Int. Ed. 2019, 58, 5609–5613. [Google Scholar] [CrossRef]
  167. Sun, S.; Cheng, H.; Li, X.; Wu, X.; Zhen, D.; Wang, Y.; Jin, R.; He, G. Improving CO2 Electroreduction Activity by Creating an Oxygen Vacancy-Rich Surface with One-Dimensional In–SnO2 Hollow Nanofiber Architecture. Ind. Eng. Chem. Res. 2021, 60, 1164–1174. [Google Scholar] [CrossRef]
  168. Wang, K.; Liu, D.; Liu, L.; Liu, J.; Hu, X.; Li, P.; Li, M.; Vasenko, A.S.; Xiao, C.; Ding, S. Tuning the Local Electronic Structure of Oxygen Vacancies over Copper-Doped Zinc Oxide for Efficient CO2 Electroreduction. EScience 2022, 2, 518–528. [Google Scholar] [CrossRef]
  169. Geng, Z.; Kong, X.; Chen, W.; Su, H.; Liu, Y.; Cai, F.; Wang, G.; Zeng, J. Oxygen Vacancies in ZnO Nanosheets Enhance CO2 Electrochemical Reduction to CO. Angew. Chem. 2018, 130, 6162–6167. [Google Scholar] [CrossRef]
  170. Zhang, B.; Chang, Y.; Wu, Y.; Fan, Z.; Zhai, P.; Wang, C.; Gao, J.; Sun, L.; Hou, J. Regulating *OCHO Intermediate as Rate-Determining Step of Defective Oxynitride Nanosheets Enabling Robust CO2 Electroreduction. Adv. Energy Mater. 2022, 12, 2200321. [Google Scholar] [CrossRef]
  171. Du, X.; Qin, Y.; Gao, B.; Wang, K.; Li, D.; Li, Y.; Ding, S.; Song, Z.; Su, Y.; Xiao, C. Plasma-Assisted and Oxygen Vacancy-Engineered In2O3 Films for Enhanced Electrochemical Reduction of CO2. Appl. Surf. Sci. 2021, 563, 150405. [Google Scholar] [CrossRef]
  172. Zong, X.; Jin, Y.; Li, Y.; Zhang, X.; Zhang, S.; Xie, H.; Zhang, J.; Xiong, Y. Morphology-Controllable ZnO Catalysts Enriched with Oxygen-Vacancies for Boosting CO2 Electroreduction to CO. J. CO2 Util. 2022, 61, 102051. [Google Scholar] [CrossRef]
  173. Ren, X.; Gao, Y.; Zheng, L.; Wang, Z.; Wang, P.; Zheng, Z.; Liu, Y.; Cheng, H.; Dai, Y.; Huang, B. Oxygen Vacancy Enhancing CO2 Electrochemical Reduction to CO on Ce-Doped ZnO Catalysts. Surf. Interfaces 2021, 23, 100923. [Google Scholar] [CrossRef]
  174. Shao, Q.; Wang, P.; Huang, X. Opportunities and Challenges of Interface Engineering in Bimetallic Nanostructure for Enhanced Electrocatalysis. Adv. Funct. Mater. 2019, 29, 1806419. [Google Scholar] [CrossRef]
  175. Qin, B.; Li, Y.; Wang, H.; Yang, G.; Cao, Y.; Yu, H.; Zhang, Q.; Liang, H.; Peng, F. Efficient Electrochemical Reduction of CO2 into CO Promoted by Sulfur Vacancies. Nano Energy 2019, 60, 43–51. [Google Scholar] [CrossRef]
  176. Zhang, A.; He, R.; Li, H.; Chen, Y.; Kong, T.; Li, K.; Ju, H.; Zhu, J.; Zhu, W.; Zeng, J. Nickel Doping in Atomically Thin Tin Disulfide Nanosheets Enables Highly Efficient CO2 Reduction. Angew. Chem. Int. Ed. 2018, 57, 10954–10958. [Google Scholar] [CrossRef] [PubMed]
  177. Li, Q.; Wang, Y.; Zeng, J.; Zhao, X.; Chen, C.; Wu, Q.; Chen, L.; Chen, Z.-Y.; Lei, Y. Bimetallic Chalcogenides for Electrocatalytic CO2 Reduction. Rare Met. 2021, 40, 3442–3453. [Google Scholar] [CrossRef]
  178. Xing, Z.; Rongjian, S.; Feng, Z.; Yuan, R.; Ruixia, L.; Zhenhai, W.; Ruihu, W. Metal–Organic Framework-Derived CuS Nanocages for Selective CO2 Electroreduction to Formate. CCS Chem. 2021, 3, 199–207. [Google Scholar] [CrossRef]
  179. Lv, K.; Suo, W.; Shao, M.; Zhu, Y.; Wang, X.; Feng, J.; Fang, M.; Zhu, Y. Nitrogen Doped MoS2 and Nitrogen Doped Carbon Dots Composite Catalyst for Electroreduction CO2 to CO with High Faradaic Efficiency. Nano Energy 2019, 63, 103834. [Google Scholar] [CrossRef]
  180. He, C.; Chen, S.; Long, R.; Song, L.; Xiong, Y. Design of CuInS2 Hollow Nanostructures toward CO2 Electroreduction. Sci. China Chem. 2020, 63, 1721–1726. [Google Scholar] [CrossRef]
  181. Zhou, J.-H.; Yuan, K.; Zhou, L.; Guo, Y.; Luo, M.-Y.; Guo, X.-Y.; Meng, Q.-Y.; Zhang, Y.-W. Boosting Electrochemical Reduction of CO2 at a Low Overpotential by Amorphous Ag-Bi-S-O Decorated Bi0 Nanocrystals. Angew. Chem. Int. Ed. 2019, 58, 14197–14201. [Google Scholar] [CrossRef]
  182. Zhang, A.; Liang, Y.; Li, H.; Zhao, X.; Chen, Y.; Zhang, B.; Zhu, W.; Zeng, J. Harmonizing the Electronic Structures of the Adsorbate and Catalysts for Efficient CO(2) Reduction. Nano Lett. 2019, 19, 6547–6553. [Google Scholar] [CrossRef] [PubMed]
  183. Song, Y.; Wang, Y.; Shao, J.; Ye, K.; Wang, Q.; Wang, G. Boosting CO2 Electroreduction via Construction of a Stable ZnS/ZnO Interface. ACS Appl. Mater. Interfaces 2022, 14, 20368–20374. [Google Scholar] [CrossRef]
  184. Gao, F.-Y.; Hu, S.-J.; Zhang, X.-L.; Zheng, Y.-R.; Wang, H.-J.; Niu, Z.-Z.; Yang, P.-P.; Bao, R.-C.; Ma, T.; Dang, Z.; et al. High-Curvature Transition-Metal Chalcogenide Nanostructures with a Pronounced Proximity Effect Enable Fast and Selective CO2 Electroreduction. Angew. Chem. Int. Ed. 2020, 59, 8706–8712. [Google Scholar] [CrossRef]
  185. Zhao, C.; Wang, J.; Goodenough, J.B. Comparison of Electrocatalytic Reduction of CO2 to HCOOH with Different Tin Oxides on Carbon Nanotubes. Electrochem. Commun. 2016, 65, 9–13. [Google Scholar] [CrossRef]
  186. Guo, Y.; Liu, B.; Gao, Y.; Luo, Y.; Zhao, J.; Zhang, Z.; Zhao, C. Oxygen Vacancy and Facet Engineering of Cuprous Oxide by Doping Transition Metal Oxides for Boosting Alcohols Selectivity in Electrochemical CO2 Reduction. J. Power Sources 2023, 556, 232468. [Google Scholar] [CrossRef]
  187. Li, Q.; Fu, J.; Zhu, W.; Chen, Z.; Shen, B.; Wu, L.; Xi, Z.; Wang, T.; Lu, G.; Zhu, J.; et al. Tuning Sn-Catalysis for Electrochemical Reduction of CO2 to CO via the Core/Shell Cu/SnO2 Structure. J. Am. Chem. Soc. 2017, 139, 4290–4293. [Google Scholar] [CrossRef]
  188. Hussain, N.; Abdelkareem, M.A.; Alawadhi, H.; Begum, S.; Elsaid, K.; Olabi, A.G. Novel Ternary CuO–ZnO–MoS2 Composite Material for Electrochemical CO2 Reduction to Alcohols. J. Power Sources 2022, 549, 232128. [Google Scholar] [CrossRef]
  189. Wu, J.; Risalvato, F.G.; Ma, S.; Zhou, X.-D. Electrochemical Reduction of Carbon Dioxide III. The Role of Oxide Layer Thickness on the Performance of Sn Electrode in a Full Electrochemical Cell. J. Mater. Chem. A 2014, 2, 1647–1651. [Google Scholar] [CrossRef]
  190. Dong, H.; Zhang, L.; Li, L.; Deng, W.; Hu, C.; Zhao, Z.-J.; Gong, J. Abundant Ce3+ Ions in Au-CeOx Nanosheets to Enhance CO2 Electroreduction Performance. Small 2019, 15, 1900289. [Google Scholar] [CrossRef]
  191. Mi, Y.; Qiu, Y.; Liu, Y.; Peng, X.; Hu, M.; Zhao, S.; Cao, H.; Zhuo, L.; Li, H.; Ren, J.; et al. Cobalt−Iron Oxide Nanosheets for High-Efficiency Solar-Driven CO2−H2O Coupling Electrocatalytic Reactions. Adv. Funct. Mater. 2020, 30, 2003438. [Google Scholar] [CrossRef]
  192. Wen, G.; Ren, B.; Park, M.G.; Yang, J.; Dou, H.; Zhang, Z.; Deng, Y.-P.; Bai, Z.; Yang, L.; Gostick, J.; et al. Ternary Sn-Ti-O Electrocatalyst Boosts the Stability and Energy Efficiency of CO2 Reduction. Angew. Chem. Int. Ed. 2020, 59, 12860–12867. [Google Scholar] [CrossRef] [PubMed]
  193. Choi, Y.-W.; Scholten, F.; Sinev, I.; Roldan Cuenya, B. Enhanced Stability and CO/Formate Selectivity of Plasma-Treated SnOx/AgOx Catalysts during CO2 Electroreduction. J. Am. Chem. Soc. 2019, 141, 5261–5266. [Google Scholar] [CrossRef]
  194. Xie, H.; Chen, S.; Ma, F.; Liang, J.; Miao, Z.; Wang, T.; Wang, H.-L.; Huang, Y.; Li, Q. Boosting Tunable Syngas Formation via Electrochemical CO2 Reduction on Cu/In2O3 Core/Shell Nanoparticles. ACS Appl. Mater. Interfaces 2018, 10, 36996–37004. [Google Scholar] [CrossRef] [PubMed]
  195. Ren, D.; Gao, J.; Pan, L.; Wang, Z.; Luo, J.; Zakeeruddin, S.M.; Hagfeldt, A.; Grätzel, M. Atomic Layer Deposition of ZnO on CuO Enables Selective and Efficient Electroreduction of Carbon Dioxide to Liquid Fuels. Angew. Chem. Int. Ed. 2019, 58, 15036–15040. [Google Scholar] [CrossRef]
  196. Gao, J.; Zhang, H.; Guo, X.; Luo, J.; Zakeeruddin, S.M.; Ren, D.; Grätzel, M. Selective C–C Coupling in Carbon Dioxide Electroreduction via Efficient Spillover of Intermediates As Supported by Operando Raman Spectroscopy. J. Am. Chem. Soc. 2019, 141, 18704–18714. [Google Scholar] [CrossRef]
  197. Zhu, S.; Ren, X.; Li, X.; Niu, X.; Wang, M.; Xu, S.; Wang, Z.; Han, Y.; Wang, Q. Core-Shell ZnO@Cu2O as Catalyst to Enhance the Electrochemical Reduction of Carbon Dioxide to C2 Products. Catalysts 2021, 11, 535. [Google Scholar] [CrossRef]
  198. Wang, Y.; Chen, Z.; Han, P.; Du, Y.; Gu, Z.; Xu, X.; Zheng, G. Single-Atomic Cu with Multiple Oxygen Vacancies on Ceria for Electrocatalytic CO2 Reduction to CH4. ACS Catal. 2018, 8, 7113–7119. [Google Scholar] [CrossRef]
  199. He, R.; Yuan, X.; Shao, P.; Duan, T.; Zhu, W. Hybridization of Defective Tin Disulfide Nanosheets and Silver Nanowires Enables Efficient Electrochemical Reduction of CO2 into Formate and Syngas. Small 2019, 15, 1904882. [Google Scholar] [CrossRef] [PubMed]
  200. Ning, H.; Fei, X.; Tan, Z.; Wang, W.; Yang, Z.; Wu, M. In Situ-Fabricated In2S3-Reduced Graphene Oxide Nanosheet Composites for Enhanced CO2 Electroreduction to Formate. ACS Appl. Nano Mater. 2022, 5, 2335–2342. [Google Scholar] [CrossRef]
  201. Liu, F.; Ren, X.; Zhao, J.; Wu, H.; Wang, J.; Han, X.; Deng, Y.; Hu, W. Inhibiting Sulfur Dissolution and Enhancing Activity of SnS for CO2 Electroreduction via Electronic State Modulation. ACS Catal. 2022, 12, 13533–13541. [Google Scholar] [CrossRef]
  202. Zeng, L.; Shi, J.; Luo, J.; Chen, H. Silver Sulfide Anchored on Reduced Graphene Oxide as a High -Performance Catalyst for CO2 Electroreduction. J. Power Sources 2018, 398, 83–90. [Google Scholar] [CrossRef]
  203. Li, S.; Duan, H.; Yu, J.; Qiu, C.; Yu, R.; Chen, Y.; Fang, Y.; Cai, X.; Yang, S. Cu Vacancy Induced Product Switching from Formate to CO for CO2 Reduction on Copper Sulfide. ACS Catal. 2022, 12, 9074–9082. [Google Scholar] [CrossRef]
  204. Cheng, J.; Yang, X.; Xuan, X.; Zhou, J. Efficient Hybrid Solar-to-Alcohol System via Synergistic Catalysis between Well-Defined Cu–N4 Sites and Its Sulfide (CuS). Chem. Eng. J. 2020, 392, 123799. [Google Scholar] [CrossRef]
  205. Xu, J.; Li, X.; Liu, W.; Sun, Y.; Ju, Z.; Yao, T.; Wang, C.; Ju, H.; Zhu, J.; Wei, S.; et al. Carbon Dioxide Electroreduction into Syngas Boosted by a Partially Delocalized Charge in Molybdenum Sulfide Selenide Alloy Monolayers. Angew. Chem. Int. Ed. 2017, 56, 9121–9125. [Google Scholar] [CrossRef] [PubMed]
  206. Li, F.; Chen, L.; Xue, M.; Williams, T.; Zhang, Y.; MacFarlane, D.R.; Zhang, J. Towards a Better Sn: Efficient Electrocatalytic Reduction of CO2 to Formate by Sn/SnS2 Derived from SnS2 Nanosheets. Nano Energy 2017, 31, 270–277. [Google Scholar] [CrossRef]
  207. Xue, D.; Xia, H.; Yan, W.; Zhang, J.; Mu, S. Defect Engineering on Carbon-Based Catalysts for Electrocatalytic CO2 Reduction. Nano-Micro Lett. 2020, 13, 5. [Google Scholar] [CrossRef]
  208. Titirici, M.-M.; White, R.J.; Brun, N.; Budarin, V.L.; Su, D.S.; del Monte, F.; Clark, J.H.; MacLachlan, M.J. Sustainable Carbon Materials. Chem. Soc. Rev. 2015, 44, 250–290. [Google Scholar] [CrossRef]
  209. Jia, Y.; Chen, J.; Yao, X. Defect Electrocatalytic Mechanism: Concept{,} Topological Structure and Perspective. Mater. Chem. Front. 2018, 2, 1250–1268. [Google Scholar] [CrossRef]
  210. Amiinu, I.S.; Liu, X.; Pu, Z.; Li, W.; Li, Q.; Zhang, J.; Tang, H.; Zhang, H.; Mu, S. From 3D ZIF Nanocrystals to Co–Nx/C Nanorod Array Electrocatalysts for ORR, OER, and Zn–Air Batteries. Adv. Funct. Mater. 2018, 28, 1704638. [Google Scholar] [CrossRef]
  211. Xie, C.; Yan, D.; Chen, W.; Zou, Y.; Chen, R.; Zang, S.; Wang, Y.; Yao, X.; Wang, S. Insight into the Design of Defect Electrocatalysts: From Electronic Structure to Adsorption Energy. Mater. Today 2019, 31, 47–68. [Google Scholar] [CrossRef]
  212. Liu, S.; Yang, H.; Su, X.; Ding, J.; Mao, Q.; Huang, Y.; Zhang, T.; Liu, B. Rational Design of Carbon-Based Metal-Free Catalysts for Electrochemical Carbon Dioxide Reduction: A Review. J. Energy Chem. 2019, 36, 95–105. [Google Scholar] [CrossRef]
  213. Zhao, H.; Sun, C.; Jin, Z.; Wang, D.-W.; Yan, X.; Chen, Z.; Zhu, G.; Yao, X. Carbon for the Oxygen Reduction Reaction: A Defect Mechanism. J. Mater. Chem. A 2015, 3, 11736–11739. [Google Scholar] [CrossRef]
  214. Todorova, T.K.; Schreiber, M.W.; Fontecave, M. Mechanistic Understanding of CO2 Reduction Reaction (CO2RR) Toward Multicarbon Products by Heterogeneous Copper-Based Catalysts. ACS Catal. 2020, 10, 1754–1768. [Google Scholar] [CrossRef]
  215. Yang, H.; Lin, Q.; Zhang, C.; Yu, X.; Cheng, Z.; Li, G.; Hu, Q.; Ren, X.; Zhang, Q.; Liu, J.; et al. Carbon Dioxide Electroreduction on Single-Atom Nickel Decorated Carbon Membranes with Industry Compatible Current Densities. Nat. Commun. 2020, 11, 593. [Google Scholar] [CrossRef]
  216. Zhang, B.; Zhang, J.; Shi, J.; Tan, D.; Liu, L.; Zhang, F.; Lu, C.; Su, Z.; Tan, X.; Cheng, X.; et al. Manganese Acting as a High-Performance Heterogeneous Electrocatalyst in Carbon Dioxide Reduction. Nat. Commun. 2019, 10, 2980. [Google Scholar] [CrossRef]
  217. Wang, W.; Shang, L.; Chang, G.; Yan, C.; Shi, R.; Zhao, Y.; Waterhouse, G.I.N.; Yang, D.; Zhang, T. Intrinsic Carbon-Defect-Driven Electrocatalytic Reduction of Carbon Dioxide. Adv. Mater. 2019, 31, 1808276. [Google Scholar] [CrossRef]
  218. Liu, S.; Yang, H.; Huang, X.; Liu, L.; Cai, W.; Gao, J.; Li, X.; Zhang, T.; Huang, Y.; Liu, B. Identifying Active Sites of Nitrogen-Doped Carbon Materials for the CO2 Reduction Reaction. Adv. Funct. Mater. 2018, 28, 1800499. [Google Scholar] [CrossRef]
  219. Ju, W.; Bagger, A.; Hao, G.-P.; Varela, A.S.; Sinev, I.; Bon, V.; Roldan Cuenya, B.; Kaskel, S.; Rossmeisl, J.; Strasser, P. Understanding Activity and Selectivity of Metal-Nitrogen-Doped Carbon Catalysts for Electrochemical Reduction of CO2. Nat. Commun. 2017, 8, 944. [Google Scholar] [CrossRef]
  220. Gong, K.; Du, F.; Xia, Z.; Durstock, M.; Dai, L. Nitrogen-Doped Carbon Nanotube Arrays with High Electrocatalytic Activity for Oxygen Reduction. Science 2009, 323, 760–764. [Google Scholar] [CrossRef] [PubMed]
  221. Dong, Y.; Zhang, S.; Du, X.; Hong, S.; Zhao, S.; Chen, Y.; Chen, X.; Song, H. Boosting the Electrical Double-Layer Capacitance of Graphene by Self-Doped Defects through Ball-Milling. Adv. Funct. Mater. 2019, 29, 1901127. [Google Scholar] [CrossRef]
  222. Xue, L.; Li, Y.; Liu, X.; Liu, Q.; Shang, J.; Duan, H.; Dai, L.; Shui, J. Zigzag Carbon as Efficient and Stable Oxygen Reduction Electrocatalyst for Proton Exchange Membrane Fuel Cells. Nat. Commun. 2018, 9, 3819. [Google Scholar] [CrossRef]
  223. Dou, S.; Tao, L.; Wang, R.; El Hankari, S.; Chen, R.; Wang, S. Plasma-Assisted Synthesis and Surface Modification of Electrode Materials for Renewable Energy. Adv. Mater. 2018, 30, 1705850. [Google Scholar] [CrossRef] [PubMed]
  224. Tao, L.; Wang, Q.; Dou, S.; Ma, Z.; Huo, J.; Wang, S.; Dai, L. Edge-Rich and Dopant-Free Graphene as a Highly Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Chem. Commun. 2016, 52, 2764–2767. [Google Scholar] [CrossRef] [PubMed]
  225. Wang, Q.; Lei, Y.; Wang, D.; Li, Y. Defect Engineering in Earth-Abundant Electrocatalysts for CO2 and N2 Reduction. Energy Environ. Sci. 2019, 12, 1730–1750. [Google Scholar] [CrossRef]
  226. Zhu, J.; Huang, Y.; Mei, W.; Zhao, C.; Zhang, C.; Zhang, J.; Amiinu, I.S.; Mu, S. Effects of Intrinsic Pentagon Defects on Electrochemical Reactivity of Carbon Nanomaterials. Angew. Chem. Int. Ed. 2019, 58, 3859–3864. [Google Scholar] [CrossRef] [PubMed]
  227. Feng, W.; Long, P.; Feng, Y.; Li, Y. Two-Dimensional Fluorinated Graphene: Synthesis, Structures, Properties and Applications. Adv. Sci. 2016, 3, 1500413. [Google Scholar] [CrossRef]
  228. Varela, A.S.; Ranjbar Sahraie, N.; Steinberg, J.; Ju, W.; Oh, H.-S.; Strasser, P. Metal-Doped Nitrogenated Carbon as an Efficient Catalyst for Direct CO2 Electroreduction to CO and Hydrocarbons. Angew. Chem. Int. Ed. 2015, 54, 10758–10762. [Google Scholar] [CrossRef] [PubMed]
  229. Zhu, J.; Mu, S. Defect Engineering in Carbon-Based Electrocatalysts: Insight into Intrinsic Carbon Defects. Adv. Funct. Mater. 2020, 30, 2001097. [Google Scholar] [CrossRef]
  230. Cheng, W.; Yuan, P.; Lv, Z.; Guo, Y.; Qiao, Y.; Xue, X.; Liu, X.; Bai, W.; Wang, K.; Xu, Q.; et al. Boosting Defective Carbon by Anchoring Well-Defined Atomically Dispersed Metal-N4 Sites for ORR, OER, and Zn-Air Batteries. Appl. Catal. B Environ. 2020, 260, 118198. [Google Scholar] [CrossRef]
  231. Liu, T.; Ali, S.; Lian, Z.; Si, C.; Su, D.S.; Li, B. Phosphorus-Doped Onion-like Carbon for CO2 Electrochemical Reduction: The Decisive Role of the Bonding Configuration of Phosphorus. J. Mater. Chem. A 2018, 6, 19998–20004. [Google Scholar] [CrossRef]
  232. Nakata, K.; Ozaki, T.; Terashima, C.; Fujishima, A.; Einaga, Y. High-Yield Electrochemical Production of Formaldehyde from CO2 and Seawater. Angew. Chem. Int. Ed. 2014, 53, 871–874. [Google Scholar] [CrossRef] [PubMed]
  233. Wu, J.; Liu, M.; Sharma, P.P.; Yadav, R.M.; Ma, L.; Yang, Y.; Zou, X.; Zhou, X.-D.; Vajtai, R.; Yakobson, B.I.; et al. Incorporation of Nitrogen Defects for Efficient Reduction of CO2 via Two-Electron Pathway on Three-Dimensional Graphene Foam. Nano Lett. 2016, 16, 466–470. [Google Scholar] [CrossRef]
  234. Xie, J.; Zhao, X.; Wu, M.; Li, Q.; Wang, Y.; Yao, J. Metal-Free Fluorine-Doped Carbon Electrocatalyst for CO2 Reduction Outcompeting Hydrogen Evolution. Angew. Chem. Int. Ed. 2018, 57, 9640–9644. [Google Scholar] [CrossRef] [PubMed]
  235. Qin, R.; Liu, P.; Fu, G.; Zheng, N. Strategies for Stabilizing Atomically Dispersed Metal Catalysts. Small Methods 2018, 2, 1700286. [Google Scholar] [CrossRef]
  236. Huang, F.; Deng, Y.; Chen, Y.; Cai, X.; Peng, M.; Jia, Z.; Xie, J.; Xiao, D.; Wen, X.; Wang, N.; et al. Anchoring Cu1 Species over Nanodiamond-Graphene for Semi-Hydrogenation of Acetylene. Nat. Commun. 2019, 10, 4431. [Google Scholar] [CrossRef]
  237. Yoo, M.; Yu, Y.-S.; Ha, H.; Lee, S.; Choi, J.-S.; Oh, S.; Kang, E.; Choi, H.; An, H.; Lee, K.-S.; et al. A Tailored Oxide Interface Creates Dense Pt Single-Atom Catalysts with High Catalytic Activity. Energy Environ. Sci. 2020, 13, 1231–1239. [Google Scholar] [CrossRef]
  238. Chen, Y.; Ji, S.; Zhao, S.; Chen, W.; Dong, J.; Cheong, W.-C.; Shen, R.; Wen, X.; Zheng, L.; Rykov, A.I.; et al. Enhanced Oxygen Reduction with Single-Atomic-Site Iron Catalysts for a Zinc-Air Battery and Hydrogen-Air Fuel Cell. Nat. Commun. 2018, 9, 5422. [Google Scholar] [CrossRef]
  239. Zhao, C.; Dai, X.; Yao, T.; Chen, W.; Wang, X.; Wang, J.; Yang, J.; Wei, S.; Wu, Y.; Li, Y. Ionic Exchange of Metal–Organic Frameworks to Access Single Nickel Sites for Efficient Electroreduction of CO2. J. Am. Chem. Soc. 2017, 139, 8078–8081. [Google Scholar] [CrossRef]
  240. Wei, S.; Wang, Y.; Chen, W.; Li, Z.; Cheong, W.-C.; Zhang, Q.; Gong, Y.; Gu, L.; Chen, C.; Wang, D.; et al. Atomically Dispersed Fe Atoms Anchored on COF-Derived N-Doped Carbon Nanospheres as Efficient Multi-Functional Catalysts. Chem. Sci. 2020, 11, 786–790. [Google Scholar] [CrossRef]
  241. Zhang, S.; Kang, P.; Ubnoske, S.; Brennaman, M.K.; Song, N.; House, R.L.; Glass, J.T.; Meyer, T.J. Polyethylenimine-Enhanced Electrocatalytic Reduction of CO2 to Formate at Nitrogen-Doped Carbon Nanomaterials. J. Am. Chem. Soc. 2014, 136, 7845–7848. [Google Scholar] [CrossRef]
  242. Wang, H.; Chen, Y.; Hou, X.; Ma, C.; Tan, T. Nitrogen-Doped Graphenes as Efficient Electrocatalysts for the Selective Reduction of Carbon Dioxide to Formate in Aqueous Solution. Green Chem. 2016, 18, 3250–3256. [Google Scholar] [CrossRef]
  243. Chai, G.-L.; Guo, Z.-X. Highly Effective Sites and Selectivity of Nitrogen-Doped Graphene/CNT Catalysts for CO2 Electrochemical Reduction. Chem. Sci. 2016, 7, 1268–1275. [Google Scholar] [CrossRef] [PubMed]
  244. Yang, H.; Wu, Y.; Lin, Q.; Fan, L.; Chai, X.; Zhang, Q.; Liu, J.; He, C.; Lin, Z. Composition Tailoring via N and S Co-Doping and Structure Tuning by Constructing Hierarchical Pores: Metal-Free Catalysts for High-Performance Electrochemical Reduction of CO2. Angew. Chem. Int. Ed. 2018, 57, 15476–15480. [Google Scholar] [CrossRef]
  245. Wu, J.; Sharifi, T.; Gao, Y.; Zhang, T.; Ajayan, P.M. Emerging Carbon-Based Heterogeneous Catalysts for Electrochemical Reduction of Carbon Dioxide into Value-Added Chemicals. Adv. Mater. 2019, 31, 1804257. [Google Scholar] [CrossRef]
  246. Sreekanth, N.; Nazrulla, M.A.; Vineesh, T.V.; Sailaja, K.; Phani, K.L. Metal-Free Boron-Doped Graphene for Selective Electroreduction of Carbon Dioxide to Formic Acid/Formate. Chem. Commun. 2015, 51, 16061–16064. [Google Scholar] [CrossRef] [PubMed]
  247. Ouyang, T.; Ye, Y.-Q.; Wu, C.-Y.; Xiao, K.; Liu, Z.-Q. Heterostructures Composed of N-Doped Carbon Nanotubes Encapsulating Cobalt and β-Mo2C Nanoparticles as Bifunctional Electrodes for Water Splitting. Angew. Chem. Int. Ed. 2019, 58, 4923–4928. [Google Scholar] [CrossRef] [PubMed]
  248. Wang, Z.; Jin, H.; Meng, T.; Liao, K.; Meng, W.; Yang, J.; He, D.; Xiong, Y.; Mu, S. Fe, Cu-Coordinated ZIF-Derived Carbon Framework for Efficient Oxygen Reduction Reaction and Zinc–Air Batteries. Adv. Funct. Mater. 2018, 28, 1802596. [Google Scholar] [CrossRef]
  249. Pan, Y.; Lin, R.; Chen, Y.; Liu, S.; Zhu, W.; Cao, X.; Chen, W.; Wu, K.; Cheong, W.-C.; Wang, Y.; et al. Design of Single-Atom Co–N5 Catalytic Site: A Robust Electrocatalyst for CO2 Reduction with Nearly 100% CO Selectivity and Remarkable Stability. J. Am. Chem. Soc. 2018, 140, 4218–4221. [Google Scholar] [CrossRef]
  250. Yan, C.; Li, H.; Ye, Y.; Wu, H.; Cai, F.; Si, R.; Xiao, J.; Miao, S.; Xie, S.; Yang, F.; et al. Coordinatively Unsaturated Nickel–Nitrogen Sites towards Selective and High-Rate CO2 Electroreduction. Energy Environ. Sci. 2018, 11, 1204–1210. [Google Scholar] [CrossRef]
  251. Bi, W.; Li, X.; You, R.; Chen, M.; Yuan, R.; Huang, W.; Wu, X.; Chu, W.; Wu, C.; Xie, Y. Surface Immobilization of Transition Metal Ions on Nitrogen-Doped Graphene Realizing High-Efficient and Selective CO2 Reduction. Adv. Mater. 2018, 30, 1706617. [Google Scholar] [CrossRef]
  252. Cheng, Y.; Zhao, S.; Li, H.; He, S.; Veder, J.-P.; Johannessen, B.; Xiao, J.; Lu, S.; Pan, J.; Chisholm, M.F.; et al. Unsaturated Edge-Anchored Ni Single Atoms on Porous Microwave Exfoliated Graphene Oxide for Electrochemical CO2. Appl. Catal. B Environ. 2019, 243, 294–303. [Google Scholar] [CrossRef]
  253. Wang, X.; Chen, Z.; Zhao, X.; Yao, T.; Chen, W.; You, R.; Zhao, C.; Wu, G.; Wang, J.; Huang, W.; et al. Regulation of Coordination Number over Single Co Sites: Triggering the Efficient Electroreduction of CO2. Angew. Chem. 2018, 130, 1962–1966. [Google Scholar] [CrossRef]
  254. Lu, C.; Yang, J.; Wei, S.; Bi, S.; Xia, Y.; Chen, M.; Hou, Y.; Qiu, M.; Yuan, C.; Su, Y.; et al. Atomic Ni Anchored Covalent Triazine Framework as High Efficient Electrocatalyst for Carbon Dioxide Conversion. Adv. Funct. Mater. 2019, 29, 1806884. [Google Scholar] [CrossRef]
  255. He, Q.; Lee, J.H.; Liu, D.; Liu, Y.; Lin, Z.; Xie, Z.; Hwang, S.; Kattel, S.; Song, L.; Chen, J.G. Accelerating CO2 Electroreduction to CO Over Pd Single-Atom Catalyst. Adv. Funct. Mater. 2020, 30, 2000407. [Google Scholar] [CrossRef]
  256. Rogers, C.; Perkins, W.S.; Veber, G.; Williams, T.E.; Cloke, R.R.; Fischer, F.R. Synergistic Enhancement of Electrocatalytic CO2 Reduction with Gold Nanoparticles Embedded in Functional Graphene Nanoribbon Composite Electrodes. J. Am. Chem. Soc. 2017, 139, 4052–4061. [Google Scholar] [CrossRef]
  257. Liu, Y.; Zhang, Y.; Cheng, K.; Quan, X.; Fan, X.; Su, Y.; Chen, S.; Zhao, H.; Zhang, Y.; Yu, H.; et al. Selective Electrochemical Reduction of Carbon Dioxide to Ethanol on a Boron- and Nitrogen-Co-Doped Nanodiamond. Angew. Chem. Int. Ed. 2017, 56, 15607–15611. [Google Scholar] [CrossRef]
  258. Ni, W.; Xue, Y.; Zang, X.; Li, C.; Wang, H.; Yang, Z.; Yan, Y.-M. Fluorine Doped Cagelike Carbon Electrocatalyst: An Insight into the Structure-Enhanced CO Selectivity for CO2 Reduction at High Overpotential. ACS Nano 2020, 14, 2014–2023. [Google Scholar] [CrossRef]
  259. Dong, Y.; Zhang, Q.; Tian, Z.; Li, B.; Yan, W.; Wang, S.; Jiang, K.; Su, J.; Oloman, C.W.; Gyenge, E.L.; et al. Ammonia Thermal Treatment toward Topological Defects in Porous Carbon for Enhanced Carbon Dioxide Electroreduction. Adv. Mater. 2020, 32, 2001300. [Google Scholar] [CrossRef]
  260. Xue, X.; Yang, H.; Yang, T.; Yuan, P.; Li, Q.; Mu, S.; Zheng, X.; Chi, L.; Zhu, J.; Li, Y.; et al. N{,}P-Coordinated Fullerene-like Carbon Nanostructures with Dual Active Centers toward Highly-Efficient Multi-Functional Electrocatalysis for CO2RR{,} ORR and Zn-Air Battery. J. Mater. Chem. A 2019, 7, 15271–15277. [Google Scholar] [CrossRef]
  261. Wu, Q.; Gao, J.; Feng, J.; Liu, Q.; Zhou, Y.; Zhang, S.; Nie, M.; Liu, Y.; Zhao, J.; Liu, F.; et al. A CO2 Adsorption Dominated Carbon Defect-Based Electrocatalyst for Efficient Carbon Dioxide Reduction. J. Mater. Chem. A 2020, 8, 1205–1211. [Google Scholar] [CrossRef]
  262. Wu, J.; Yadav, R.M.; Liu, M.; Sharma, P.P.; Tiwary, C.S.; Ma, L.; Zou, X.; Zhou, X.-D.; Yakobson, B.I.; Lou, J.; et al. Achieving Highly Efficient, Selective, and Stable CO2 Reduction on Nitrogen-Doped Carbon Nanotubes. ACS Nano 2015, 9, 5364–5371. [Google Scholar] [CrossRef]
  263. Ye, L.; Ying, Y.; Sun, D.; Zhang, Z.; Fei, L.; Wen, Z.; Qiao, J.; Huang, H. Highly Efficient Porous Carbon Electrocatalyst with Controllable N-Species Content for Selective CO2 Reduction. Angew. Chem. Int. Ed. 2020, 59, 3244–3251. [Google Scholar] [CrossRef]
  264. Chen, C.; Sun, X.; Yan, X.; Wu, Y.; Liu, H.; Zhu, Q.; Bediako, B.B.A.; Han, B. Boosting CO2 Electroreduction on N,P-Co-Doped Carbon Aerogels. Angew. Chem. Int. Ed. 2020, 59, 11123–11129. [Google Scholar] [CrossRef]
  265. Wu, J.; Ma, S.; Sun, J.; Gold, J.I.; Tiwary, C.; Kim, B.; Zhu, L.; Chopra, N.; Odeh, I.N.; Vajtai, R.; et al. A Metal-Free Electrocatalyst for Carbon Dioxide Reduction to Multi-Carbon Hydrocarbons and Oxygenates. Nat. Commun. 2016, 7, 13869. [Google Scholar] [CrossRef]
  266. Zhu, W.; Zhang, L.; Liu, S.; Li, A.; Yuan, X.; Hu, C.; Zhang, G.; Deng, W.; Zang, K.; Luo, J.; et al. Enhanced CO2 Electroreduction on Neighboring Zn/Co Monomers by Electronic Effect. Angew. Chem. Int. Ed. 2020, 59, 12664–12668. [Google Scholar] [CrossRef] [PubMed]
  267. Pan, F.; Li, B.; Sarnello, E.; Fei, Y.; Gang, Y.; Xiang, X.; Du, Z.; Zhang, P.; Wang, G.; Nguyen, H.T.; et al. Atomically Dispersed Iron–Nitrogen Sites on Hierarchically Mesoporous Carbon Nanotube and Graphene Nanoribbon Networks for CO2 Reduction. ACS Nano 2020, 14, 5506–5516. [Google Scholar] [CrossRef] [PubMed]
  268. Ren, W.; Tan, X.; Yang, W.; Jia, C.; Xu, S.; Wang, K.; Smith, S.C.; Zhao, C. Isolated Diatomic Ni-Fe Metal–Nitrogen Sites for Synergistic Electroreduction of CO2. Angew. Chem. Int. Ed. 2019, 58, 6972–6976. [Google Scholar] [CrossRef] [PubMed]
  269. Gong, Y.; Jiao, L.; Qian, Y.; Pan, C.; Zheng, L.; Cai, X.; Liu, B.; Yu, S.; Jiang, H. Regulating the Coordination Environment of MOF-Templated Single-Atom Nickel Electrocatalysts for Boosting CO2 Reduction. Angew. Chem. 2020, 132, 2727–2731. [Google Scholar] [CrossRef]
  270. Liu, Y.; Tian, D.; Biswas, A.N.; Xie, Z.; Hwang, S.; Lee, J.H.; Meng, H.; Chen, J.G. Transition Metal Nitrides as Promising Catalyst Supports for Tuning CO/H2 Syngas Production from Electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2020, 59, 11345–11348. [Google Scholar] [CrossRef]
  271. Tang, S.; Zhou, X.; Zhang, S.; Li, X.; Yang, T.; Hu, W.; Jiang, J.; Luo, Y. Metal-Free Boron Nitride Nanoribbon Catalysts for Electrochemical CO2 Reduction: Combining High Activity and Selectivity. ACS Appl. Mater. Interfaces 2019, 11, 906–915. [Google Scholar] [CrossRef]
  272. Yin, Z.; Yu, C.; Zhao, Z.; Guo, X.; Shen, M.; Li, N.; Muzzio, M.; Li, J.; Liu, H.; Lin, H.; et al. Cu3N Nanocubes for Selective Electrochemical Reduction of CO2 to Ethylene. Nano Lett. 2019, 19, 8658–8663. [Google Scholar] [CrossRef] [PubMed]
  273. Liang, Z.-Q.; Zhuang, T.-T.; Seifitokaldani, A.; Li, J.; Huang, C.-W.; Tan, C.-S.; Li, Y.; De Luna, P.; Dinh, C.T.; Hu, Y.; et al. Copper-on-Nitride Enhances the Stable Electrosynthesis of Multi-Carbon Products from CO2. Nat. Commun. 2018, 9, 3828. [Google Scholar] [CrossRef] [PubMed]
  274. Li, X.; Xi, S.; Sun, L.; Dou, S.; Huang, Z.; Su, T.; Wang, X. Isolated FeN4 Sites for Efficient Electrocatalytic CO2 Reduction. Adv. Sci. 2020, 7, 2001545. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Depiction of the simplified overall anthropogenic carbon cycle (A) and electrochemical CO2 reduction reaction in an electrochemical cell with possible carbon products (B) [6] (copyright: John Wiley and Sons, 2017).
Figure 1. Depiction of the simplified overall anthropogenic carbon cycle (A) and electrochemical CO2 reduction reaction in an electrochemical cell with possible carbon products (B) [6] (copyright: John Wiley and Sons, 2017).
Catalysts 13 00393 g001
Figure 2. Representation of overall electrochemical reduction of CO2 (ER-CO2) and hydrogen evolution reaction (HER) over different groups of metal electrodes (A); standard redox potentials (vs. SHE) for the formation of different CO2-RR products (B) [11] (copyright: John Wiley and Sons, 2019); and reaction pathways from CO2 to ethylene and ethanol (C) [10] (copyright: John Wiley and Sons, 2020).
Figure 2. Representation of overall electrochemical reduction of CO2 (ER-CO2) and hydrogen evolution reaction (HER) over different groups of metal electrodes (A); standard redox potentials (vs. SHE) for the formation of different CO2-RR products (B) [11] (copyright: John Wiley and Sons, 2019); and reaction pathways from CO2 to ethylene and ethanol (C) [10] (copyright: John Wiley and Sons, 2020).
Catalysts 13 00393 g002
Figure 3. Illustration of the advantages of the electroreduction of CO2 (ER-CO2) (A) and a comparison of the market price values (USD per kg) of CO2-RR reduction products in terms of the electrical energy input (USD per kWh−1) (B) [11] (copyright: John Wiley and Sons, 2019).
Figure 3. Illustration of the advantages of the electroreduction of CO2 (ER-CO2) (A) and a comparison of the market price values (USD per kg) of CO2-RR reduction products in terms of the electrical energy input (USD per kWh−1) (B) [11] (copyright: John Wiley and Sons, 2019).
Catalysts 13 00393 g003
Figure 4. Schematic illustration of practical challenges and factors affecting CO2 electroreduction (ER-CO2) [5] (copyright: Elsevier, 2020).
Figure 4. Schematic illustration of practical challenges and factors affecting CO2 electroreduction (ER-CO2) [5] (copyright: Elsevier, 2020).
Catalysts 13 00393 g004
Figure 5. Representation of various defect/interface designs on the electrocatalyst surface (A) and illustration of the effects of these defects/interfaces on the electrocatalyst surface for the CO2-RR (B) [68] (copyright: Cell Press, 2018).
Figure 5. Representation of various defect/interface designs on the electrocatalyst surface (A) and illustration of the effects of these defects/interfaces on the electrocatalyst surface for the CO2-RR (B) [68] (copyright: Cell Press, 2018).
Catalysts 13 00393 g005
Figure 6. (A,B) Schematic illustration of a plausible mechanism for (I) deactivation process on a Cu surface without crystal defects and (II) stable catalysis on a Cu surface containing crystal defects during the CO2-RR. FEs of various Cu electrodes: (C) annealed CuNNs; (D) CuNNs; (E) OT-CuNNs; and (F) their corresponding current density vs. time curves. (G) Correlations between the standard deviations of the FEs for various C2 products on the different Cu electrodes and the number of defects. (H) Raman spectra of the different electrodes before and after electrolytic reactions [108] (copyright: Elsevier, 2019).
Figure 6. (A,B) Schematic illustration of a plausible mechanism for (I) deactivation process on a Cu surface without crystal defects and (II) stable catalysis on a Cu surface containing crystal defects during the CO2-RR. FEs of various Cu electrodes: (C) annealed CuNNs; (D) CuNNs; (E) OT-CuNNs; and (F) their corresponding current density vs. time curves. (G) Correlations between the standard deviations of the FEs for various C2 products on the different Cu electrodes and the number of defects. (H) Raman spectra of the different electrodes before and after electrolytic reactions [108] (copyright: Elsevier, 2019).
Catalysts 13 00393 g006
Figure 7. Overall structures of metal oxides and metal sulfides for CO2 electroreduction.
Figure 7. Overall structures of metal oxides and metal sulfides for CO2 electroreduction.
Catalysts 13 00393 g007
Figure 8. (1) Schematic diagram of the synthesis of Vo-CuO(Sn) (A); low-magnification (B) and high-magnification (C) SEM images of Vo-CuO(Sn) nanosheets. Representative (D) low-resolution and (E,F) high-resolution TEM image images of Vo-CuO(Sn) nanosheets. (G) The EDS mapping of as-prepared Vo-CuO(Sn) nanosheets. Comparison of LSV curves (H) of CO2-RR FE% for Vo-CuO(Sn) (I) and Vo-CuO, CuO electrodes (J). The FE and partial current density of Vo-CuO(Sn) compared with other CO2-to-CO catalysts (K), and the stability measurement of Vo-CuO(Sn), Vo-CuO, CuO(Sn), and CuO samples (L) [165] (copyright: American Chemical Society, 2022). (2) Linear sweep voltammogram curves of H-InOx NRs (A); total current densities (B); formate partial current densities (C); potential-dependent formate FEs of the P-InOx NRs, O-InOx NRs, and H-InOx NRs (D); prolonged stability (E); formate FEs of the H-InOx NRs (F); and amorphous InOx nanoribbons (NRs) with different concentrations of oxygen vacancies applied as CO2 electroreduction (CO2-RR) catalysts (G) [166] (copyright: John Wiley and Sons, 2019). (3) Evidence of Vo. Sn3d5/2 XPS spectra of electro-reduced, thermal-reduced, and pristine NF In–SnO2 (A); corresponding O1s XPS spectra (B); schematics of Vo decoration through pre-electro reduction (C); EPR spectra of pre-electro-reduced and pristine NF In–SnO2 (D); and schematics of CO2-to-formate catalysis process at Vo sites (E) [167] (copyright: American Chemical Society, 2021).
Figure 8. (1) Schematic diagram of the synthesis of Vo-CuO(Sn) (A); low-magnification (B) and high-magnification (C) SEM images of Vo-CuO(Sn) nanosheets. Representative (D) low-resolution and (E,F) high-resolution TEM image images of Vo-CuO(Sn) nanosheets. (G) The EDS mapping of as-prepared Vo-CuO(Sn) nanosheets. Comparison of LSV curves (H) of CO2-RR FE% for Vo-CuO(Sn) (I) and Vo-CuO, CuO electrodes (J). The FE and partial current density of Vo-CuO(Sn) compared with other CO2-to-CO catalysts (K), and the stability measurement of Vo-CuO(Sn), Vo-CuO, CuO(Sn), and CuO samples (L) [165] (copyright: American Chemical Society, 2022). (2) Linear sweep voltammogram curves of H-InOx NRs (A); total current densities (B); formate partial current densities (C); potential-dependent formate FEs of the P-InOx NRs, O-InOx NRs, and H-InOx NRs (D); prolonged stability (E); formate FEs of the H-InOx NRs (F); and amorphous InOx nanoribbons (NRs) with different concentrations of oxygen vacancies applied as CO2 electroreduction (CO2-RR) catalysts (G) [166] (copyright: John Wiley and Sons, 2019). (3) Evidence of Vo. Sn3d5/2 XPS spectra of electro-reduced, thermal-reduced, and pristine NF In–SnO2 (A); corresponding O1s XPS spectra (B); schematics of Vo decoration through pre-electro reduction (C); EPR spectra of pre-electro-reduced and pristine NF In–SnO2 (D); and schematics of CO2-to-formate catalysis process at Vo sites (E) [167] (copyright: American Chemical Society, 2021).
Catalysts 13 00393 g008
Figure 9. Schematic illustration of distinct types of defective CBN (A) [207]; XPS N1s spectra (B), N content (C), and Raman spectra (D) of intrinsic-defect engineered D-NC-X electrocatalyst [238]; synthesis process (E), Raman spectra (F), and N1s XPS spectra (G) of nitrogen-doped GM-X electrocatalysts [239]; typical SEM image (H) and material model and local schematic structure (I) of N-coordinated metal-doped (M-N-C) electrocatalysts [240].
Figure 9. Schematic illustration of distinct types of defective CBN (A) [207]; XPS N1s spectra (B), N content (C), and Raman spectra (D) of intrinsic-defect engineered D-NC-X electrocatalyst [238]; synthesis process (E), Raman spectra (F), and N1s XPS spectra (G) of nitrogen-doped GM-X electrocatalysts [239]; typical SEM image (H) and material model and local schematic structure (I) of N-coordinated metal-doped (M-N-C) electrocatalysts [240].
Catalysts 13 00393 g009
Figure 10. C K-edge NEXAFS spectra (A), expanded view of NEXAFS spectra (B), and DFT calculations for ECR activities of variously defective D-NC–X electrocatalysts (C) [238]. Multipotential curves of GM2 electrocatalyst (D), illustration of the CRR processes (E), and free energy diagram (F) of various N–GRW electrocatalysts [239]. The catalytic reactivity trends (G), CO2 physisorption isotherm with inset of the pore size distribution (H), and catalyst mass-normalized CO partial currents vs. applied potential (I) of distinct M-N-C electrocatalysts [240].
Figure 10. C K-edge NEXAFS spectra (A), expanded view of NEXAFS spectra (B), and DFT calculations for ECR activities of variously defective D-NC–X electrocatalysts (C) [238]. Multipotential curves of GM2 electrocatalyst (D), illustration of the CRR processes (E), and free energy diagram (F) of various N–GRW electrocatalysts [239]. The catalytic reactivity trends (G), CO2 physisorption isotherm with inset of the pore size distribution (H), and catalyst mass-normalized CO partial currents vs. applied potential (I) of distinct M-N-C electrocatalysts [240].
Catalysts 13 00393 g010
Table 2. Single-metal- and alloy-based electrocatalysts for electrochemical CO2-RR.
Table 2. Single-metal- and alloy-based electrocatalysts for electrochemical CO2-RR.
Single Metals and Alloys for CO2-RR
CatalystType of DefectElectrolyte/E vs. RHEProductFE (%)Ref.
AuNPs/CNTMetal doping0.25 M Na2CO3/−0.5 V CO94[142]
Cu NPs (solar-driven)Grain boundary1 M KOH/−1.0 V to −1.3 V C2 73.1[143]
Bi nanodendritesLow-angle grain boundary0.5 M KHCO3/−0.76 V HCOO92[144]
CuAg wireAlloying/intrinsic defects (point, line, and plane)1 M KOH/−0.7 V C2H460[110]
C2H5OH25
Au-Pd Core–shellCore–shell structures0.5 M KHCO3/−0.6 V CO96.7[137]
AuCu3@AuOxidative etching/alloying0.5 M KHCO3/−0.6 V CO97.27[136]
Cu@CuEu NPsPoint (Eu) defect0.5 M KHCO3/−1.2 V CH474.7[139]
Bi@Sn NPsCompressive strain effect0.5 M KHCO3/−1.1 V HCOOH91[140]
Cu@SnMetal doping0.5 M KHCO3/−0.93 V HCOO100[141]
Table 3. Metal-oxide-based electrocatalysts for electrochemical CO2-RR.
Table 3. Metal-oxide-based electrocatalysts for electrochemical CO2-RR.
Metal-Oxide-Based Electrocatalysts
CatalystsType of DefectElectrolyte/E vs. RHEProducts* FE%Ref.
SnOxThermal treatment/
annealing in O2 atm
0.1 M KHCO3/−0.76 VCOOH64[174]
Cu2O/ZnOOxygen vacancies0.5 M KHCO3/−0.3 VAlcohols64.27[175]
7.0 nm SnOx layer of Sn nanoparticlesCore–shell structures/
alloying
0.5 M KHCO3/−0.7 VCO35[176]
CuO/ZnO/MoS2Grain boundaries0.5 M KHCO3/−0.6 VCH3OH24.6[177]
3.5 nm SnOx layer of Sn nanoparticlesLayer thickness0.1 M KHCO3/−1.2 VCOOH64[178]
Sn-Cu/SnOxCore–shell structures1 M KOH/−0.7 VCOOH98[152]
Au-CeOx nanosheetMetal doping0.1 M KHCO3/−0.5 VCO90.1[179]
Co2FeO4 nanosheetsMetal (Co) incorporation0.1M KHCO3/−1.0 VCO92[180]
Sn-Ti–OO2 defect sites0.5M KHCO3/−0.54 VCO94.5[181]
SnOx/AgOxPlasma oxidation0.1M KHCO3/−0.80 VC191[182]
Cu/In2O3Core–shell structures0.5M KHCO3/−0.40 V to −0.90 VSyngas90[183]
ZnO shell/CuO coreCore–shell structures1M KOH/−0.68 VCH3CH=OH49[184]
Cu2O-AgGrain boundaries0.2M KHCO3/−0.60 V −1.2 VC2H452[185]
In-doped Cu@Cu2OMetal doping/core–shell structures0.1M KHCO3/−0.45 V to −0.84 VCO2.2[186]
Cu-CeO2–4%O2 vacancy0.1M KHCO3/−1.8 VCH458[187]
Ag-Bi-S–O-decorated BiOBimetal (Ag, Bi)
doping/decoration
0.5M KHCO3/−0.45 VHCOOH94.3[188]
* FE% = Faradaic efficiency (%); HCOOH = formate/formic acid; CO = carbon monoxide; C2H5OH = ethanol; CH3CH=OH = isopropyl alcohol.
Table 4. Metal-sulfide-based electrocatalysts for electrochemical CO2-RR.
Table 4. Metal-sulfide-based electrocatalysts for electrochemical CO2-RR.
Metal-Sulfide-Based Electrocatalysts
CatalystsType of DefectElectrolyte/ E vs. RHEProducts* FE%Ref.
Ag-SnS2Chemical-induced defective sites0.5M KHCO3/−1.0 VHCOOH65.5[199]
5%Ni-SnS2Metal (Ni) dopingKHCO3/−0.9 VCO93[191]
Mn-In2S3Metal (Mn) doping0.1M KHCO3/−0.9 VHCOOH86[196]
Nitrogen-doped MoS2The heteroatom (N)
doping
EMIM-BF4/−0.9 VCO90.2[194]
ZnS/ZnOOxide–sulfide interface1M KOH/−0.73 VCO91.9[197]
In2S3-rGOLayer thickness0.1M KHCO3/−1.2 V-91%[200]
SnSThe heteroatom (In)
doping
1M KOH/−0.60 VHCOOH96.6[201]
Ag2S/N, S-rGOHeteroatom (N, S)
co-doping
0.1M KHCO3/−0.76 VCO87.4[202]
Lattice defect/metal (Cu) vacancy0.5M KHCO3/−0.84 VCOOH-[203]
CuS/CuPorSurface structural
defects
0.5M NaHCO3/−2.0 VC2H5OH74.4[204]
MoSeSAlloying (TMD alloy)−/−1.15 VCO45[205]
SnS2/rGOO2 defects0.5M NaHCO3/−0.75 VCO84.5[206]
* FE% = Faradaic efficiency (%); HCOOH = formate/formic acid; CO = carbon monoxide; C2H5OH = ethanol; CH3CH=OH = isopropyl alcohol.
Table 5. Carbon-based electrocatalysts for electrochemical CO2-RR.
Table 5. Carbon-based electrocatalysts for electrochemical CO2-RR.
Carbon-Based Nanomaterials for CO2-RR
CatalystType of DefectElectrolyte/E vs. RHEProduct* FE (%)Ref.
GNR AuNPsMetal doping0.5M KHCO3/−0.2 VCO>90[255]
BNDB, N co-doping0.5M KHCO3/−1.0 VC2H5OH93.2[256]
B-grapheneB doping 0.1M KHCO3/−1.1 VHCOOH-[245]
N-grapheneN doping 0.5M KHCO3/-HCOO-[242]
F-CPCF doping 0.5M KHCO3/−1.0 VCO88.3[257]
DPC-NH3-950Thermal N removal, topological effects0.1M KHCO3/−0.6 VCO95.2[258]
N-GRWN doping 0.5M KHCO3/−0.49 VCO87.6[239]
NCNTs-ACN-850N doping 0.1M KHCO3/−1.05 VCO80[85]
P-OLCP doping 0.5M KHCO3/−0.9 VCO81[228]
FCF doping 0.1M KHCO3/−1.22 VCO93.1[259]
Ni-SAs/N-CMetal (Ni) doping0.5M KHCO3/−1.0 VCO71.9[236]
N, P-FCN, P co-doping0.5M KHCO3/−0.8 VCO83.3[260]
D-NC-1100Heteroatom doping, porous carbon0.1M KHCO3/−0.6 VCO94.5[238]
DHPCThermal treatment0.5M KHCO3/−0.5 VCO99.5[261]
NG-800N doping 0.1M KHCO3/−0.58 VCO85[230]
NCNTsN doping0.1M KHCO3/−0.78 VCO80[262]
NPC-1000N doping0.5M KHCO3/−0.55 VCO98.4[263]
NDDN doping0.5M NaHCO3/−1.0 VAcetate91.8[49]
NPCAN, P co-doping0.5M KHCO3/−2.49 VCO99.1[264]
NGQDsN doping1M KOH/−0.75 V(Total)90[265]
−0.75 VC2H431
−0.86 VCH415
−0.74 VC2H5OH11.8
Co-N5/HNPCSsN doping0.2M NaHCO3/−0.79 VCO99.4[248]
Ni-N-CMetal (Ni, Fe) and N co-doping0.1M KHCO3/−0.78 VCO85[240]
Fe-N-C−0.55V65
Co-N2Metal (Co) and N co-doping0.5M KHCO3/−0.63 VCO94[252]
Ni-N-MPGOMetal (Ni) doping0.5M KHCO3/−0.70 VCO92.1[251]
Zn/Co-N-CN doping 0.5M KHCO3/−0.50 VCO93.2[266]
Fe-N/CNT@GNRFe, N co-doping0.5M KHCO3/−0.76 VCO96[267]
C-Zn1Ni4ZIF-8Ni, N co-doping0.5M KHCO3/−1.03 VCO98[249]
Ni/Fe-NCNi, Fe, and N doping0.5M KHCO3/−0.70 VCO98[268]
NiSA-N2-CMetal (Ni) doping0.5M KHCO3/−0.80 VCO98[269]
* FE% = Faradaic efficiency (%); HCOOH = formate/formic acid; CO = carbon monoxide; C2H5OH = ethanol; CH3CH=OH = isopropyl alcohol.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Balu, S.; Hanan, A.; Venkatesvaran, H.; Chen, S.-W.; Yang, T.C.-K.; Khalid, M. Recent Progress in Surface-Defect Engineering Strategies for Electrocatalysts toward Electrochemical CO2 Reduction: A Review. Catalysts 2023, 13, 393. https://doi.org/10.3390/catal13020393

AMA Style

Balu S, Hanan A, Venkatesvaran H, Chen S-W, Yang TC-K, Khalid M. Recent Progress in Surface-Defect Engineering Strategies for Electrocatalysts toward Electrochemical CO2 Reduction: A Review. Catalysts. 2023; 13(2):393. https://doi.org/10.3390/catal13020393

Chicago/Turabian Style

Balu, Sridharan, Abdul Hanan, Harikrishnan Venkatesvaran, Shih-Wen Chen, Thomas C.-K. Yang, and Mohammad Khalid. 2023. "Recent Progress in Surface-Defect Engineering Strategies for Electrocatalysts toward Electrochemical CO2 Reduction: A Review" Catalysts 13, no. 2: 393. https://doi.org/10.3390/catal13020393

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

Article Metrics

Back to TopTop