Jian
Zhao
*a,
Song
Xue
a,
James
Barber
b,
Yiwei
Zhou
a,
Jie
Meng
a and
Xuebin
Ke
*c
aTianjin Key Laboratory of Organic Solar Cells and Photochemical Conversion, School of Chemistry and Chemical Engineering, Tianjin University of Technology, Tianjin 300384, China. E-mail: zjtjbd@hotmail.com
bDepartment of Life Sciences, Imperial College London, London, UK
cDepartment of Chemical Engineering, University of Hull, HU6 7RX, UK. E-mail: X.Ke@hull.ac.uk
First published on 5th February 2020
The electrochemical (EC) reduction of CO2 is a promising approach for value-added fuel or chemical production. Cu-based electrodes have been extensively used as a ‘star’ material for CO2 reduction to hydrocarbons. This review mainly focuses on the recent progress of Cu-based heterogeneous electrocatalysts for CO2 reduction from 2013 to 2019. Various morphologies of oxide-derived, bimetallic Cu species and their activity in EC CO2 reduction are reviewed, providing insights for the standardization of Cu-based heterogeneous systems. We also present a tutorial manual to describe parameters for the EC CO2 reduction process, especially for the pretreatment of the reaction system. This will offer useful guidance for newcomers to the field. Aqueous and non-aqueous electrolyte effects based on Cu electrodes are discussed. Finally, an overview of reaction systems of EC/PEC CO2 reduction and H2O oxidation for Cu-based heterogeneous catalysts is provided.
In recent years, CO2 reduction to fuels and chemical products has been investigated by various methods, such as biochemical approaches,4 building blocks for organic synthesis,5 thermal hydrogenation,6–11 photocatalysis,12–14 electrocatalysis15–18 and dry reforming with methane.19,20 Among them, electrochemical (EC) reduction has attracted much attention.21–23 However, as researchers have acknowledged, electrochemical CO2 reduction is strongly influenced by the pH24–27 and conductivity or concentration of electrolyte,23,28 applied potentials and currents,29 CO2 concentration30 and flow rate,30,31 and temperature,32,33 making a direct, quantitative comparison of data from different groups difficult. To help provide information and to push for a more standardized “benchmark” for EC CO2 reduction, we provide a tutorial manual for researchers who intend to start EC CO2 reduction and clearly state the experimental parameters when comparing different systems.
EC reduction of CO2 is involved in a variety of products ranging from CO, HCOO−, HCHO, CH4, CH3OH, C2+ hydrocarbons (e.g. C2H4, C2H6) and oxygenates, to higher hydrocarbons. The standard potentials for selected CO2 reduction reactions are listed in Table 1. Since Hori’s work in 1989 (ref. 34) and 1994,35 Cu electrodes have attracted much attention due to their unique advantages for hydrocarbon production compared with other pure metallic electrodes.36–38 However, a large overpotential is required and low selectivity is observed due to the wide range of products and the competing hydrogen evolution reaction. Numerous publications on Cu-based electrodes have reported a lower overpotential and/or improved product selectivity, especially since 2013, as shown in Fig. 1. Different selectivity is sometimes reported with similar Cu-based materials. The difference may be caused by the experimental conditions and the material itself, such as the oxidation state of Cu, dimensional structure or surface roughness. Therefore, there is a need to compare these results under certain experimental conditions to further understand their differences and to provide a guide for the development of Cu-based electrocatalysts for CO2 reduction.
Reactions | E 0 (V) vs. NHE (pH = 7) | E 0 (V) vs. RHE |
---|---|---|
CO2 + e− → CO2˙− | −1.90 | −1.48 |
2H+ + 2e− → H2(g) | −0.42 | 0.00 |
CO2 + 2H+ + 2e− → HCOOH(l) | −0.55 | −0.19 |
CO2 + 2H+ + 2e− → CO(g) + H2O | −0.52 | −0.10 |
CO2 + 4H+ + 4e− → HCHO(l) + H2O | −0.48 | −0.06 |
CO2 + 6H+ + 6e− → CH3OH(l) + H2O | −0.39 | +0.03 |
CO2 + 8H+ + 8e− → CH4(g) + 2H2O | −0.25 | +0.17 |
2CO2 + 12H+ + 12e− → C2H4(g) + 4H2O | −0.38 | 0.08 |
2CO2 + 12H+ + 12e− → C2H5OH(l) + 3H2O | −0.35 | 0.09 |
2CO2 + 14H+ + 14e− → C2H6(g) + 4H2O | −0.28 | 0.14 |
3CO2 + 18H+ + 18e− → C3H7OH(l) + 5H2O | −0.30 | 0.10 |
In this review, we first provide a tutorial guide for conducting EC CO2 reduction experiments. Also, various conditions such as the catalyst state, flow rate and type of electrolyte are specified when comparing studies from different research groups. CO2 electrocatalysts are usually divided into two categories: homogeneous and heterogeneous systems. Some reviews on molecular electrocatalysts have covered the advantages of homogeneous systems.39–43 Readers interested in Cu-based complexes could refer to other work.44–51 Cu-based heterogeneous photocatalysts have been reviewed for direct conversion into solar fuels.52 In contrast to the reviews of Cu-based heterogeneous catalysts reported for EC CO2 reduction (review of Cu-based nanocatalysts,53 review of Cu-binary alloys,54 review mainly focused on theoretical studies, nanostructured Cu and bimetallics55), our review includes a tutorial guide for newcomers, lists the parameters specified by different research groups for comparison (Table 2), summarizes the benchmark activity for special products (Section 3) and covers more types of Cu-based heterogeneous electrocatalysts. Different types of Cu-based heterogeneous electrocatalysts, including film and power systems, will be discussed in five main categories: (1) morphology. Morphology control allows us to improve the catalytic activity by tailoring the structure of active sites and increasing the surface area/number of active sites. (2) Oxide-derived Cu. One recent method to enhance CO2 reduction is the oxidation and subsequent reduction of Cu. CuxO formed by annealing, electrodeposition, Cl− and plasma induction all exhibited improved performance for EC CO2 reduction after in situ or ex situ redox processes. (3) Bimetallic species. Alloys are known to be able to tune the geometric and electronic properties of their parent metals. A bifunctional interface (separated composite) without forming an alloy could also improve the performance of each metal while showing fewer changes to the intrinsic electronic properties. The synergistic effects between different metals could create novel catalytic properties. (4) Surface modification. Modification with inorganic species to enhance durability and organic ligands to capture key intermediates is another strategy to improve the performance of Cu-based electrodes. (5) Supports. Supports or substrates are critical to uniformly deposit the catalyst and create novel catalytic features at the interface. Following these sections, as shown in Fig. 2, the electrolyte effects and the overall reaction systems coupling CO2 reduction and H2O oxidation will be discussed for Cu-based heterogeneous electrocatalysts.
Pre-electrolysis of the electrolyte is one method, in which two platinum gauzes could be used as the working and counter electrodes in a two-electrode configuration and a negative cathodic potential was applied for certain period (e.g. −2 V for 24 h,57,58 Pt black cathode at 0.025 mA cm−2 for more than 15 h (ref. 34)). The electrolyte could be also purified by electrolysis between two graphite rods,59 or two high purity Ti foils (99.99%).60 Here, the electrodes used should be of high purity and should be removed from the solution before pre-electrolysis stops to avoid recontamination of the electrolyte caused by the dissolution of electrodeposited impurities on the electrode.
Another method is treating the electrolyte with an ion exchange resin (e.g. Chelex).61–63 Generally the ion exchange resin was stirred with the electrolyte for at least 24 h to minimize the concentration of transition metal impurities. During electrolysis, trace metal ions coordinated in situ with ethylenediaminetetraacetic acid (EDTA) or ex situ with solid-supported iminodiacetate resin should also be considered.62
During electrolysis, there are three types of CO2 gas and electrolyte flow: (1) continuously bubbling CO2 during the process;31,60,62,63,67–84 (2) CO2 bubbling plus flow electrolyte58,85–87 (the CH4/C2H4 ratio could be tuned by varying the CO2 gas and KHCO3 electrolyte flow rates);58 (3) no report of any form of convection at all.33,88
The unit of flow rate is standard cubic centimetres per minute (sccm) or ml min−1. Certain flow rates of CO2 gas were chosen by different research groups, such as 5 sccm,63,67–72 10 sccm,73–76 20 sccm (ref. 60, 77–84) and 30 sccm. The cell design and catalyst itself should also be considered when choosing the flow rate. The Takanabe group73 chose 10 sccm to ensure sufficient CO2 supply to the electrode surface while preventing the catalyst dropping off the electrode by gas bubbles. The Koper group31 investigated the influence of CO2 flow rate on the activity of 3D porous hollow fibre Cu and observed a maximum FE of 75% CO at −0.4 V vs. RHE when flow rate >30 sccm. In a flow setup, the flow rate of the CO2 gas was set to 50 sccm and that of the electrolyte was 100 sccm.85 To suit a specific system, one could also choose to adjust the flow of electrolyte with the applied potential.86
With CO2 gas bubbling, the catholyte could be stirred.62,89,90 The rotating rate also has an influence on the catalyst activity.30 H2 evolution increased and the product selectivity switched from CH4 to CO when the rotating rate was increased, although there was increased availability of CO2 at the electrode surface. This was caused by the enhanced mass transfer of dissolved CO away from the electrode surface and then less adsorbed CO was left for further reduction.90
Fig. 3 100 ppm CO and 100 ppm H2 tested with TCD, and 50 ppm CH4, 50 ppm C2H4 and 50 ppm C2H6 tested with FID using an online GC system for one-time injection. |
Fig. 4 1D 1H NMR analysis of methanol, ethanol, formate and acetic acid products with DMSO as internal standard. |
Faradaic efficiency (FE) is the ratio between the amount of product actually detected by an analysis technique such as GC, HPLC or NMR, and the amount of product theoretically formed based on the charge passed during electrolysis. The faradaic efficiency or selectivity for each product in EC CO2 reduction could be calculated according to the following equation:
Without CO2 gas bubbling during electrolysis:
CO2 gas was continuously bubbled during electrolysis (the first and second measurements are not used to calculate faradaic efficiency to ensure that the data is from a system under equilibrium conditions):
As stated, therefore, the most important thing for newcomers before conducting EC CO2 reduction in aqueous electrolyte is the pre-treatment of the reaction system. Otherwise, H2 may be the sole product rather than CO2 reduction products, as the potential needed for water reduction to H2 is less negative than that for CO2 reduction. In addition, attention should be paid to the flow and convection state of CO2 gas and the electrolyte to make sure that all the experiments are conducted under certain conditions for comparison.
Fig. 5 The benchmark activity and progress of mechanism studies on Cu-based heterogeneous catalysts for EC CO2 reduction. |
Fig. 6 Most likely reaction pathways from adsorbed *CO on Cu surface for C1 and C2 products. Plain red, plain blue and dashed blue routes are for major, minor and trace C2 products. |
Particle size is a critical parameter in tuning the activity and selectivity of Cu nanoparticle catalysts. It may be difficult to control smaller particle sizes and interparticle distances directly via an electrochemical method. Therefore, although powder systems have to be assembled into electrodes for further application in CO2 reduction, many studies have been reported based on Cu powder systems with a controlled small particle size. ∼5 nm Cu nanoparticles embedded in a thin film of metal organic framework (MOF, used to restrict the particle size) on fluorine-doped SnO2 (FTO) exhibited 31% (major HCOO−, minor CO) CO2 reduction selectivity at −0.82 V vs. RHE (0.1 M NaClO4, pH = 4.6).116 B. R. Cuenya, P. Strasser and co-workers65 prepared Cu nanoparticles with a mean size range of 2–15 nm on glassy carbon. The hydrocarbon production decreased as the particle size decreased and vanished for sizes ≤2 nm, as shown in Fig. 7a. Cu nanoparticles (12 nm, 19 nm, 24 nm, 37 nm)/four types of carbon support also showed higher C2H4/CH4 production than smooth copper film, and smaller particle sizes were more favorable for C2H4 formation (pH = 6.8).117 These individual studies give different trends for product distribution, probably caused by different preparation methods, supports and pH of the electrolytes.
Fig. 7 (a) The faradaic efficiency of reaction products during EC CO2 reduction on Cu nanoparticles. (b) The faradaic efficiency as a function of interparticle distances over 4.7 nm Cu nanoparticles. (c) CO2 flux obtained for Cu nanoparticles with different sizes and interparticle distances based on diffusion equations. Experimental conditions: 0.1 M KHCO3, –1.1 V vs. RHE, 30 sccm CO2. Reproduced from ref. 65 for (a) and ref. 119 for (b) and (c), with permission from the American Chemical Society, 2014 and 2016. |
Another critical parameter for Cu nanoparticle catalysts is the interparticle spacing.118 B. R. Cuenya and co-workers119 designed 1.5 nm, 4.7 nm and 7.4 nm Cu nanoparticles with interparticle distances of 10–22 nm, 24–53 nm and 41–92 nm, respectively. Smaller interparticle spacing was favorable for re-adsorption of the CO intermediate and its further reduction to hydrocarbons, and with the increase of interparticle spacing the CO2 flux increased, as shown in Fig. 7b and c. P. Strasser and co-workers120 also showed that C2H4 selectivity could be tuned by particle density caused by diffusional interparticle coupling that controlled CO desorption/re-adsorption and in turn the effective COad coverage. Recently, it was also reported that stacked small Cu nanoparticles could be formed by in situ electrochemical fragmentation during the CO2 reduction, promoting C–C coupling reaction.121
The Cu nanoparticles loading level also has a significant influence on the morphology evolution and product selectivity.77 P. D. Yang and co-workers122 assembled densely packed Cu nanoparticles (6.7 nm) on carbon paper electrode. These densely packed nanoparticles changed to cube-like structures intermixed with smaller nanoparticles (Fig. 8a). Compared with spatially separated nanoparticles and ex situ prepared nanocubes, the in situ formation of cube-like particles from densely packed nanoparticles suppressed C1 products and improved C2–C3 formation (C2H4, C2H5OH, and n-C3H7OH 50% at −0.75 V vs. RHE, Fig. 8b).
Fig. 8 (a) Schematic illustration of Cu nanoparticle ensembles as an active catalyst for C2–C3 product formation. (b) Investigation of the parameters affecting structural transformation of Cu nanoparticle ensembles and their catalytic activity. (i) Separation of the NPs from their initial densely packed assembly, (ii) use of Cu nanocubes as starting materials, and (iii) change of support to a low surface area carbon plate. Experimental conditions: 0.1 M KHCO3, (i) −0.84 V, (ii) −0.86 V, and (iii) −0.81 V vs. RHE, respectively. Reproduced from ref. 122. |
Therefore, small interparticle distances and densely packed nanoparticles would lead to higher hydrocarbon formation. The above Cu nanoparticles are crystalline in form. Compared with the crystalline form, the amorphous form seems to be more favourable for C2 products. J. M. Yan and co-workers78 synthesized amorphous Cu nanoparticles (average size 3.3 nm), and achieved 37% HCOOH and 22% C2H5OH at −1.0 V vs. RHE (0.1 M KHCO3, 20 sccm CO2). The crystalline Cu nanoparticles (average size 3.4 nm) only showed 26% HCOOH and no C2H5OH at the same potential. They ascribed the enhanced activity to the high electrochemically active surface area (ECSA) and CO2 adsorption on the amorphous surface.
Similar to Cu nanoparticles formed on film in situ as an electrode, the morphology of powder nanoparticles could also have an effect on product selectivity, and a surface morphology with more defects and boundaries promotes C2 products. Star decahedron Cu nanoparticles123 and branched CuO nanoparticles106 are reported to achieve high faradaic efficiency of ethylene (C2H4) up to 52% and 70% at −1.0 V vs. RHE, respectively.
Fig. 9 (a) Schematic of the electrogrowth process of Cu from Cu2(OH)3Cl. (b) Corresponding SEM images and (c) structure evolution of the key structural features after applying various potentials for at least 1 h in 0.1 M KHCO3. Reproduced from ref. 82 with permission from Springer Nature, 2018. |
Similar trends in the role of morphology are observed for nanocube structures in powder systems, especially for C2H4 production. R. Buonsanti and co-workers129 fabricated different sizes of Cu nanocrystal spheres (7.5 nm and 27 nm) and Cu nanocrystal cubes (24 nm, 44 nm, and 63 nm). There was a monotonic size-dependent trend for both shapes – the smaller, the more active. Cube-shaped copper nanocrystals showed better performance than spheres. The overall CO2 reduction activity changed from 50% to 80% and 63% for 24 nm, 44 nm, and 63 nm, respectively, and the highest faradaic efficiency of C2H4 was 41% for 44 nm nanocubes at −1.1 V vs. RHE (Fig. 10). Edges and Cu(100) were responsible for maximizing C2H4 selectivity. However, edges were also reported to promote CH4 selectivity in nanowire structures. P. D. Yang and co-workers130 prepared ultrathin (∼20 nm) 5-fold twinned Cu nanowire on carbon black/glassy carbon plates. The catalyst could achieve 55% CH4 at −1.25 V vs. RHE with <5% other CO2 reduction products, likely due to the high density of edge sites. With the evolving morphology, CH4 decreased and C2H4 increased. Wrapping Cu nanowires with graphene oxide could prevent morphology changes and in turn prevent the decrease of CH4 selectivity.
Fig. 10 TEM images of Cu cubes with an average edge length of 24 nm (A), 44 nm (B), 63 nm (C), and corresponding faradaic efficiency after subtracting the substrate signals. Experimental conditions: 0.1 M KHCO3, −1.1 V vs. RHE, 5 sccm CO2. Reproduced from ref. 129 with permission from Wiley, 2016. |
Fig. 11 (a) Schematic representation of the key aspects of Cu electrodeposition using H2 as a template. (b) SEM images of Cu foams with different pore sizes on Cu substrate, electrodeposited for different times. (c) The faradaic efficiency of CO2 reduction products for Cu foam with various pore sizes (experimental conditions: 0.5 M NaHCO3, −0.8 V vs. RHE). Reproduced from ref. 131 for (a) and (c), and ref. 88 for (b) with permission from the American Chemical Society, 2014 and 2016. |
In addition to pore sizes on the micron scale, nano-porous structures could also change the local pH and retention time of key intermediates133,135 and, in turn, the product selectivity. Using a sputtering method on anodized aluminium oxide, as shown in Fig. 12a, the pore widths and depths of the Cu mesopore electrode could be precisely controlled. Compared with an electrode of 300 nm (width)/40 nm (depth), C2H4 formation was enhanced from 8% to 38% when the pore width was narrowed to 30 nm, whereas the major C2 product changed to C2H6 with a faradaic efficiency of 46% when the depth was increased to 70 nm at −1.3 V vs. RHE (Fig. 12b). A pH change with 3D morphology was also reported by other groups. Using Cu foil with a mixed solution of (NH3)2S2O8 and NaOH for different times, W. A. Smith and co-workers134,136 obtained Cu(OH)2 nanowires with various lengths and densities. The nanowires with high lengths and densities had an influence on the diffusion of HCO3−, leading to a high local pH, since HCO3− can neutralize OH− (HCO3− + OH− → CO32− + H2O), as shown in Fig. 12c. With higher lengths (≥2.4 ± 0.56 μm), n-C3H7OH was detected along with CO, HCOOH and C2H4. For even higher lengths (≥7.3 ± 1.3 μm), C2H6 and C2H5OH could be formed. In summary, high lengths or depths of 3D nanostructures favor higher hydrocarbon formations.
Fig. 12 (a) Scheme of Cu mesopore electrode preparation and corresponding SEM images; (i–iii) are 30 nm/40 nm, 30 nm/70 nm and 300 nm/40 nm width/depth, respectively. (b) The faradaic efficiencies of CO2 reduction products for the prepared Cu mesopores. Experimental conditions: 0.1 M KHCO3, −1.3 V vs. RHE. (c) Schematic illustration of the diffusion of electrolytes into Cu nanowire arrays. Reproduced from ref. 133 for (a) and (b), and ref. 134 for (c) with permission from Wiley, 2016. |
Fig. 13 Schematic illustration of reaction pathways on Cu(100) and (111) single crystals. Reproduced from ref. 138 with permission from the American Chemical Society, 2016. |
B. S. Yeo and co-workers141 studied Cu2O (hydrothermally prepared)-derived Cu and Cu single-crystal surfaces. Of the three single-crystal surfaces (100), (111) and (110), Cu(100) exhibited the lowest energy barrier for the dimerization of CO*. A. Nilsson and co-workers157 investigated single crystal copper (100), (111), and (211) for comparison with Cu nanocubes. The (100) surface was the most comparable to the Cu nanocube surface in terms of C2H4 production, whereas CH4 was not suppressed. One possibility is that it has the ideal terrace length or active sites for C2H4 formation. K. Chan, H. T. Wang and co-workers59 galvanostatically cycled Cu foil in Cu(NO3)2 to obtain a Cu2O nanocube layer with smooth (100) facets on the surface. They also suggested that Cu(100) and stepped (211) facets favored C2+ products over Cu(111). From the present results, Cu(100) crystal facets favor C2+ products compared with other facets for most reported systems.
As stated above, understanding the effects of Cu morphology on the selectivity is highly complex since there is a combined effect of properties, such as low-coordinated sites, catalyst density or dispersion, CO2 flux and electrical double layers in the electrolyte, on the activity of CO2 electrochemical reduction. Although there is wide variation in the activity trends for Cu with various morphologies, the selectivity for possible products may be adjusted by tuning the particle size and interparticle spacing/particle density of Cu nanoparticles/nanocubes, tuning the pore size and depth/length of 3D structures, or tuning the energy facets of Cu crystals. In addition, attention should be paid to morphology evolution during the electrochemical CO2 reduction process.
Although the true active site is still under debate, oxide-derived Cu has shown excellent performance in decreasing the potential required and enhancing selectivity for specific products. Most recent reports also confirmed that two (Cuδ+ and Cu0) were better than one.100,154 In this section, we will discuss oxide-derived Cu in detail based on the fabrication process, including annealed/oxide-derived Cu, electrodeposited/oxide-derived Cu, Cl−/oxide-derived Cu, plasma/oxide-derived Cu, and in situ/oxide-derived Cu. Since the wide variation in experimental conditions results in various results for similar materials, detailed experimental conditions are included for different groups.
Fig. 14 (a) The faradaic efficiency of CO2 reduction products for annealed Cu and polycrystalline Cu foil. 0.1 M NaHCO3, 5 sccm CO2. (b) The specific current density of CO reduction versus the grain boundary density at −0.3 V, −0.4 V and −0.5 V, respectively. Reproduced from ref. 69 for (a) and ref. 71 for (b) with permission from the American Chemical Society, 2016. |
Higher temperature annealing forms both Cu2O and CuO nanowires and the corresponding derived Cu also exhibits superior performance for CO2 reduction. C. Wang and co-workers80 annealed Cu gauze in air at 600 °C to get CuO nanowires (with Cu2O present inside the nanowires) with a diameter of 50–100 nm and length of 10–50 μm. Then they obtained high density Cu nanowires by electrochemical reduction (ECR) or forming gas reduction (FGR). The Cu nanowires obtained via the ECR method exhibited a higher surface roughness and a thin surface layer of Cu2O nanocrystals less than 10 nm, while FGR led to larger Cu crystal segments (>100 nm size) without Cu2O present on the surface. At low overpotentials (−0.3 V to −0.5 V vs. RHE), the total faradaic efficiency of CO and HCOOH was as high as 70–80% for ECR nanowires, whereas H2 was the dominant product for Cu gauze and FGR nanowires (0.1 M KCHCO3, 20 sccm CO2). They ascribed the high activity to the high-density grain boundaries and low-coordinated surface sites associated with the small crystalline features, as well as more open facets (e.g. (100) and (211)) on the surface. Later studies by C. Wang, T. Mueller and co-workers144,145 further suggested that the high activity and selectivity of the ECR nanowires could be ascribed to the (110) surface, high-angle grain boundaries, or some closely related metastable surface feature. Pre-reduction of annealed Cu in different solutions led to different activities, as also reported by J. J. Zhang’s group.155 Reduction of annealed Cu in 1 M NaOH formed only a layer of nanofibers with 30–100 nm diameters, whereas in 1 M H3PO4 the nanofibers were surrounded by kernels and achieved 43% HCOO−, which was much higher than for the pre-reduced annealed Cu in NaOH (1% HCOO−) in 0.5 M KHCO3.
Summarizing the above reports, the annealed/oxide-redox method could enhance the selectivity and lower the overpotential required for HCOO− and CO during CO2 reduction, while promoting the selectivity of C2+ products during CO reduction.
With related mechanism studies on enhanced HCOO− and CO formation, using ambient pressure X-ray photoelectron spectroscopy interpreted with quantum mechanical predictions of the structures and free energies, W. A. Goddard III, J. Yano, E. J. Crumlin and co-workers96 showed that the presence of sub-oxide species below the Cu surface played a crucial role in the adsorption and activation of CO2 on annealed/oxide-derived Cu. This thin layer of sub-oxide was essential for converting chemisorbed CO2 in the presence of H2O as the first step toward CO2 reduction products such as HCOO− and CO. Quantum mechanical calculations and experimental results also showed that there was an optimized amount of subsurface oxide. More or none at all would destabilize the bent (chemisorbed) CO2. (1) In the case of Cu(111) without subsurface oxide (Fig. 15a), the C atom of CO2 was chemically bonded to a Cu0. One of the O atoms was stabilized by hydrogen bonding to H2Oad on another Cu0. (2) When the subsurface oxide was increased to 1/4 ML (Fig. 15b), the C atom was chemically bonded to two surface Cu0. One O atom was chemically bonded to one Cu0, and the other O atom was stabilized by the Cu+ pulled up by H2Oad. (3) When the subsurface oxide was increased to 1/2 ML (Fig. 15c), the C atom was chemically bonded to a Cu+ that shares one O atom stabilized by hydrogen bonding to H2Oad on another Cu+. Later they153 found that only surface Cu+ itself could not improve the performance of CO2 reduction, and proposed Cu metal embedded in an oxidized matrix model as a partially oxidized Cu surface, where the synergy between the surface Cu+ and surface Cu0 was responsible for CO2 activation.
Fig. 15 Predicted structures for chemisorbed-CO2 with H2O on Cu(111) with different levels of subsurface oxide Osub: (a) 0 ML, (b) 1/4 ML and (c) 1/2 ML Osub. Reproduced from ref. 96. |
Fig. 16 (a) The faradaic efficiency of CO2 reduction products for electrodeposited Cu(I) and annealed Cu(II). (b) Proposed schemes for electrodeposited and annealed samples. Experimental conditions: −1.1 V vs. RHE, 0.5 M KHCO3. Reproduced from ref. 156 with permission from the American Chemical Society, 2017. |
There is a dependence of CO2 activity on film thickness for electrodeposited Cu2O film. J. Baltrusaitis, G. Mul and co-workers72 electrodeposited Cu2O film on a Cu plate. A C2H4/CH4 ratio of ∼8–12 was observed for thin film at −1.1 V vs. RHE, with a larger amount of CH4 for thicker film (0.1 M KHCO3, 5 sccm CO2). B. S. Yeo and co-workers81 electrodeposited Cu2O film with different film thicknesses on Cu discs. A maximum of 34–39% C2H4 with a ratio of C2H4/CH4 up to 100, and 9–16% C2H5OH was obtained at −1.0 V vs. RHE, as shown in Fig. 17a. Then they146 prepared Cu2O- and CuO-derived Cu with different thicknesses via a hydrothermal method (1.3 and 11.5 μm). In contrast to other works proposing the protonation of C2H4, they showed that C2H6 and C2H5OH were likely to form via the dimerization of –CH3 intermediates on thick oxide-derived Cu (Fig. 17b). They observed that Cu2O was rapidly reduced to metallic Cu during CO2 reduction by using in situ Raman spectroscopy. The surface reoxidized in tens of seconds after the cathodic potential was removed. This is in contrast to the Lee group’s150 work on electrodeposited Cu2O/GDE, where they found that Cu2O was only partially reduced on the basis of ex situ XRD and Auger electron spectroscopy. Later again, the Yeo group only observed signals belonging to CO adsorbed on Cu metallic sites rather than oxide sites for electrodeposited CuxZn.84
Fig. 17 (a) The faradaic efficiency of C2H4 and C2H5OH versus the thickness of electrodeposited Cu2O. (b) Proposed mechanism on thick oxide-derived Cu and thin oxide-derived Cu, respectively. Experimental conditions: −1.0 V vs. RHE, 0.1 M KHCO3, 20 sccm CO2. Reproduced from ref. 81 for (a) and ref. 146 for (b) with permission from the American Chemical Society, 2015 and 2017. |
Fig. 18 (a) Proposed mechanism of Cu nanoparticle growth during electrochemical cycling in the presence of Cl−. (b) Rapid oxidation of Cl−/oxide-derived Cu with a high abundance of grain boundaries. Reproduced from ref. 142 for (a) and ref. 143 for (b) with permission from Wiley, 2016 and 2017. |
In contrast to the Nilsson group’s study157 and the Yeo group’s study,159 J. Lee and co-workers160 observed the preferential formation of multicarbon fuels, especially n-C3H7OH (the first report over 10% C3–C4 products, 0.1 M KCl), using in situ prepared Cl−-induced bi-phasic Cu2O–Cu. They also found that the catalyst exhibited relatively higher Cu+ coverage than oxide-derived Cu and the use of a KCl electrolyte could prolong the preservation of the Cu2O phase compared to a KHCO3 electrolyte.
A. T. Bell, J. W. Ager and co-workers142 electrochemically cycled copper foil in the presence of halide anions KF, KCl, KBr, and KI. They observed an enhanced faradaic efficiency for C2H4 and C2H5OH (excluding KF). Their observation of C2H5OH was in contrast to the Nilsson group and Yeo group studies. Without electrochemical cycling in halide anions, the product distribution was not significantly changed even with the addition of halide anions in electrolyte for CO2 reduction. In situ Raman spectroscopy and SEM showed that during the oxidation sweep, anodic corrosion formed a Cu2O layer, which consisted of cubic crystals ∼150 nm. CuCl formed cubes when precipitated in the solution at pH > 4. In neutral and basic solutions with a low Cl− concentration, CuCl could convert to Cu2O. During the reduction sweep, irregular Cu nanoparticles (ca. 20 nm in diameter) with rounded edges were formed (Fig. 18a). They ascribed the enhancement in C2H4 formation to a large number of (100) facets (similar to the Nilsson group) and the formation of grain boundaries and defects (similar to the Kanan group). Later they64 also investigated four types of oxide-derived Cu electrocatalysts: “oxide-derived nanocrystalline Cu” developed by the Kanan group, “Cu nanowire arrays” developed by the Smith group, “electrodeposited Cu2O films” developed by the Yeo group, and “electrochemical oxidation–reduction cycled Cu” developed by the Nilsson group. There was an optimal roughness factor for the oxide-derived layers to have a high local pH and maintain a high concentration of dissolved CO2. More recently, J. W. Ager and co-workers143 prepared 18O-enriched oxide-derived Cu by cycling Cu foil in the presence of KCl in H218O. The selectivity of C2 and C3 products in 0.1 M KHCO3 was up to 60%. By analysis with ex situ secondary-ion mass spectrometry (SIMS), they found that <1% of the original 18O content remained in the sample after the CO2 reduction reaction and believed that the grain boundary was more likely responsible for the high activity, as proposed by the Kanan group. Similarly to the Yeo group, they also observed the rapid re-oxidation of oxide-derived Cu with in situ Raman spectroscopy. They believed that this rapid re-oxidation was possibly due to the grain boundaries (Fig. 18b), since Cu film with fewer grain boundaries did not oxidize quickly.
E. H. Sargent and co-workers82 applied constant potential in CO2-saturated 0.1 M KHCO3 to reduce Cu2(OH)3Cl precursor on carbon paper. This dissolution and electro-redeposition process could form Cu2O nanoneedles, which exhibited excellent activity for C2H4 formation (partial current density 160 mA cm−2) with a ratio of C2H4/CH4 of up to 200 at −1.0 V vs. RHE (20 sccm CO2). For the first time, they used in situ Cu L-edge XAS to demonstrate that the Cu+ surface species could direct C2+ product selectivity. The process of Cu2+ to Cu+ was quick (within 5 min), while Cu+ to Cu0 was much slower, and 23% Cu+ surface species still remained after 1 h electrolysis at −1.2 V vs. RHE.
In summary, Cl−/oxide-derived Cu shows excellent activity for C2 and even C3–C4 product formation, regardless of whether it is caused by the morphology evolution to nanocubes or the presence of Cu+.
Fig. 19 (a) The faradaic efficiency of hydrocarbon products for plasma-treated Cu foil. (b) Current density and C2H4 faradaic efficiency for Cu nanocubes with plasma treatment. Experimental conditions: 0.1 M KHCO3, −1.0 V vs. RHE, 30 sccm CO2. Reproduced from ref. 95 for (a) and ref. 149 for (b) with permission from Springer Nature, 2016, and the American Chemical Society, 2017. |
Fig. 20 Schematic diagram for CO2 reduction by pulsed-bias method and potentiostatic electrolysis. Reproduced from ref. 83 with permission from the American Chemical Society, 2016. |
Early last year, however, I. Chorkendorff, I. E. L. Stephens and co-workers164 investigated CO electroreduction on polycrystalline Cu foil in 0.1 M KOH at low overpotentials from −0.4 to −0.6 V vs. RHE. Compared with oxide-derived nanostructured Cu in the literature, the polycrystalline Cu foil exhibited higher selectivity for C2 and C3 aldehydes and ethylene. This indicated that future studies should focus on the intrinsic activity of Cu.
In summary, for oxide-derived Cu, various methods have been developed for the preparation of a Cu-oxide precursor (Cu2+, Cu+) and for its subsequent redox process (in situ and ex situ, electrochemical and H2 reduction). As stated at the beginning of this part, the true active site of Cu-based electrocatalysts fabricated from oxidation–reduction processes is still under discussion. Using Cu-based electrocatalysts with oxidation–reduction pretreatment, the activity and product selectivity were indeed improved, especially for the ratio of C2H4 and CH4, which has been proven by many research groups as above. Further mechanism investigations may focus on in situ and operando studies to gain more insight into the subsurface oxygen, chemical state or morphology of the CuxO catalyst under CO2 reduction conditions.
E. H. Sargent, P. D. Yang and co-workers89 prepared a Cu-enriched Au nanoneedle electrode via an underpotential deposition (UPD) method with various Cu content. Designed syngas ratios were obtained (0.5 M KHCO3, 20 sccm CO2). In situ SERS and DFT calculations illustrated how the surface electronic structure could be tuned by Cu enrichment to influence CO binding. Tuning the composition of CuAu alloys from Au-rich to Cu-rich resulted in a selectivity change from CO to hydrocarbon, which was also reported in other studies.167 For Cu-rich alloys in another study, Au addition led to the suppression of CH4 while increasing CO production.168 A. T. Bell and co-workers63 prepared strained CuAg surface alloys via melting physical mixtures of Cu and Ag under argon in a vacuum arc furnace. The incorporation of Ag atoms onto the Cu surface modified the Cu electronic structure, where the valence band density states shifted to deeper levels. As a result, the binding energies of H and O relative to CO decreased and the ratio of CO to H2 products increased (0.05 M Cs2CO3, 5 sccm CO2). An AgCu dendritic catalyst was also electrodeposited on Cu foil and Ag57Cu43 exhibited 2.2 times higher CO production than pure Ag in terms of Ag mass activity at −0.83 V vs. RHE (0.5 M KHCO3, 10 sccm CO2).76 These results indicate that by alloying Cu with Au or Ag, the ratio of CO to H2 could be tuned. Additionally, additives during the electrodeposition process and the supports applied could affect the morphology of the deposited alloy and its corresponding activity for CH4 or C2+ products. Through the addition of 3,5-diamino-1,2,4-triazole into the electroplating solution, A. A. Gewirth and co-workers102 obtained homogeneous CuAg wire samples, which exhibited higher selectivity than their counterparts of up to 60% C2H4 and 25% C2H5OH at −0.7 V vs. RHE (1.0 M KOH). T. Meyer and co-workers169 electrodeposited ∼6.6 nm CuPd nanoalloy on a polymeric film. 51% CH4 was obtained with Cu2Pd in organic electrolyte at an overpotential of −0.86 V. They believed that the enhanced CH4 formation was due to the synergistic interplay between Pd–H sites and Cu–CO sites with the polymer as a basis for local CO2 concentration. Later, they170 electrodeposited ∼6 nm CuAg nanoalloy on this polymer on glassy carbon. At 0 °C, 21% C2H3OO− was achieved with Cu2Ag3 at −1.33 V vs. RHE in 0.5 M KHCO3 with 8 ppm benzotriazole.
As stated before, we separately summarized the electrode used after fabrication without post-treatment and the powder materials used after assembly or drop casting to form films/electrodes. The following catalysts are Cu–M powder alloys, which sometimes have to be mixed with a Nafion binder and finally cast on a conductive substrate. Based on some groups’ work, C. P. Berlinguette and co-workers171 provided a general relationship between the primary product formed and the metal–CO bonding strength (Fig. 21). The best CO formation catalysts often had a CO heat of adsorption (ΔH) of 10 kcal mol−1. A lower value of ΔH is more favorable for HCOO− formation, whereas a higher value of ΔH favors H2 and hydrocarbon formation. A series of In–M on titanium substrate was prepared and followed the trend FeIn < CoIn < ZnIn < NiIn < CuIn for CO production.
Fig. 21 General relationship between the primary product formed and CO heat of adsorption (ΔH) for pure metal and mixed metal films. Reproduced from ref. 171 with permission from Wiley, 2017. |
After preparing different ratios of AuxCuy (x = 3, y = 1; x = 1, y = 1; x = 1, y = 3) alloy nanoparticles, P. D. Yang and co-workers172 obtained corresponding monolayer samples on various substrates via self-assembly by the Langmuir–Schaefer method. The monolayer samples showed great mass activity, achieving 67% CO with a partial current of −230 mA mg−1 at −0.73 V vs. RHE for Au3Cu. The activity for CO2 reduction exhibited a volcano shape, where Au3Cu represented the peak, as determined by the electronic and geometric effects (Fig. 22). These effects were associated with the local atomic arrangements at the active sites. In contrast to the AuCu alloys, alloying Pd with Cu could enhance CO and/or C2 products. M. Yamauchi, P. J. A. Kenis and co-workers86 designed CuPd nanoalloys with ordered, disordered, and phase-separated elemental arrangements (Fig. 23). With the same atomic ratio, phase-separated arrangements (more sites with neighbouring Cu atoms) favored the production of C2 products compared to the other two arrangements, with >60% (48% C2H4 and 15% C2H5OH) at −0.7 V vs. RHE. CuPd with ordered arrangements gave 75% CO at −0.7 V vs. RHE. Surface valence bond spectra suggested that geometric effects were key in determining the selectivity compared to electronic effects. N. Umezawa, J. H. Ye and co-workers173 electrodeposited CuPd on glassy carbon and optimized the ratio between Pd and Cu. 80% faradaic efficiency of CO was obtained with optimal Pd7Cu3 at −0.8 V vs. RHE. DFT calculations suggested that synergistic geometric and electronic effects were responsible for the high selectivity. D. Ma, G. X. Wang and co-workers174 loaded CuPd nanoparticles (3.3 nm) on carbon black, and Pd85Cu15/C achieved 86% CO at −0.89 V vs. RHE.
Fig. 22 Proposed mechanism for CO2 reduction on the surface of Au–Cu bimetallic nanoparticles. Larger arrows indicate higher turnover. Reproduced from ref. 172 with permission from Springer Nature, 2014. |
Fig. 23 (a) Illustration of prepared CuPd nanoalloys with different atomic mixing patterns: ordered, disordered, and phase separated. (b) XRD patterns of prepared CuPd nanoalloys, as well as previously reported pure Cu, pure Pd and CuPd alloys. Reproduced from ref. 86 with permission from the American Chemical Society, 2017. |
Fig. 24 (a) The stabilization of *COOH by introducing bi-functional effects. (b) The selectivity of C2H4 and C2H5OH for Cu2O and Ag-incorporated Cu2O. Experimental conditions: −1.2 V vs. RHE in 0.2 M KCl. (c) The proposed CO-insertion mechanism indicating the transfer of CO from Ag site that weakly binds CO to Cu site that binds residual C1 intermediates in the case of phase separated and phase blended samples in (b). Reproduced from ref. 175 for (a) and ref. 177 for (b) and (c) with permission from the American Chemical Society, 2016 and 2017. |
To design Cu–M bimetallic species, therefore, the first aspect to be considered is the group that the parent metals belong to. The preparation method also has an influence on the activity. The most important thing for mechanism investigation in this system is precise control of the composition, morphology and position of the bimetallic species.
Fig. 25 Schematic diagram of Pd-induced surface restructuring that avoids the accumulation of carbonaceous species on the Cu surface. Reproduced from ref. 60 with permission from Wiley, 2017. |
Besides modification with inorganic materials, organic ligands are coated on the Cu surface to enhance CO2 reduction, especially for CH4 and C2H4 products through the interactions between key intermediates and the functional groups of ligands. By properly modifying Cu(OH)2 nanowires with amino acid,182 H2 formation could be suppressed and CO2 reduction was promoted since the interaction between the key intermediate CHO* and –NH3+ of the adsorbed zwitterionic glycine stabilized this key intermediate during CO2 reduction. E. Andreoli and co-workers183 modified the Cu foam surface with polyamide and obtained enhanced C2H4 production with unaffected CH4 (0.1 M NaHCO3). They ascribed this enhancement to the charge transfer to the Cu surface, stabilization of the CO dimer by the –NH2 group, and the adsorption of CO near the polymer.
Generally, Cu foil, Cu plate, Cu mesh, or even FTO is used as a substrate for Cu-based electrocatalysts for EC CO2 reduction, while glassy carbon is used for Cu powder electrocatalysts. Gas diffusion layers (GDL) or gas diffusion electrodes (GDE) were also chosen as substrates to enhance the performance of the corresponding loaded Cu-based electrocatalysts due to their gas/electrolyte penetrability.86,150 Polymer on FTO169 or glassy carbon170 was used as a substrate to enhance the local concentration of CO2. Polymer-based diffusion layers or electrodes have also been fabricated recently to enhance the activity and stability of Cu electrocatalysts, where as high as 76% C2H4 was obtained at −0.55 V vs. RHE.103,185
Therefore, in order to improve the performance of Cu-based electrocatalysts, supports or substrates with high surface area and high gas and liquid penetrability should be considered. More recently, it was found that Cu3N support could act as an underlying stable Cu+ species to reduce the CO dimerization energy barrier.186 This might be another consideration when choosing supports for copper electrocatalysts.
The concentration of bicarbonate and cation size both play important roles during CO2 reduction. A high concentration of bicarbonate leads to a relatively high pH, which in turn favors CH4 formation, while a big cation size promotes C2H4 and other C2 products. G. Mul and co-workers24 investigated the influence of the electrolyte on the selectivity of Cu2O-derived Cu. An increased electrolyte concentration led to a decreased C2H4/CH4 ratio (0.05 M, 0.1 M, 0.2 M KHCO3, 50 sccm CO2). P. Strasser and co-workers25 also observed that a low concentration of bicarbonate electrolyte favored H2 and CH4 formation, while a high concentration of bicarbonate, which showed a high interfacial pH near the Cu surface, favored C2H4 formation (0.05 to 0.2 M KHCO3). For various different concentrations of electrolytes, E. H. Sargent, D. Sinton and co-workers79 predicted the pH and CO2 concentration at the electrode surface using a diffusion-based model (Fig. 26). CO2 limitation occurred under high local pH conditions. In 1991, Hori and co-workers found that cationic species (Li+, Na+, K+ and Cs+) in bicarbonate solution affected the product selectivity (C2H4/CH4).187 A. T. Bell and co-workers188,189 recently also reported that there was a decrease in faradaic efficiency for H2 and CH4, and an increase in faradaic efficiency for C2H4 and C2H5OH for Cu cathodes with increasing cation size (Li+, Na+, K+, Rb+, and Cs+) (Fig. 27a). They ascribed the effect of the electrolyte cation size on CO2 reduction to cation hydrolysis in the vicinity of the cathode. The pKa for hydrolysed cations decreased and they served as buffer agents to lower the local pH, leading to an increase in the local concentration of dissolved CO2 (Fig. 27b and c). The concentration of molecular CO2 decreased with increasing pH due to its rapid consumption by hydroxyl anions to form HCO3− and CO32−. This process occurred at much higher rates than the rate of CO2 reduction. In contrast to CO2 reduction, by using single-crystal Cu(100), single-crystal Cu(111), and polycrystalline Cu electrodes, M. T. Koper, F. C. Vallejo and co-workers190 found that the cation effects were potential- and structure-dependent in CO reduction (Fig. 27d); the onset potential for C2H4 was lower for the single crystals (−0.3 V and −0.4 V vs. RHE for Cu(100) and Cu(111)) than for the polycrystalline electrode, for which the overpotential increased with increasing cation size. The onset potential for CH4 (−0.65 V vs. RHE) was independent of both cation size and surface structure. When the applied potential was more negative than −0.65 V vs. RHE, larger cations enhanced CH4 formation. When the applied potential was from −0.65 V to −0.3 V vs. RHE, larger cations increased C2H4 selectivity.
Fig. 26 (a) Predicted local pH and (b) CO2 concentration at the electrode surface using a diffusion-based model. Reproduced from ref. 79 with permission from the Royal Society of Chemistry, 2017. |
Fig. 27 (a) The selectivity of CO2 reduction products in different bicarbonate electrolytes. (b) Calculated local pH. (c) CO2 concentration at the Cu surface. Schematics of structure- and potential-dependent cation effects. Experimental conditions: −1.0 V vs. RHE, 0.1 M LiHCO3, NaHCO3, KHCO3, RbHCO3, and CsHCO3. Reproduced from ref. 188 for (a) to (c) and ref. 190 for (d) with permission from the American Chemical Society, 2016 and 2017. |
Halides are sometimes directly added into aqueous electrolytes to enhance the CO2 reduction and suppress the competing H2 evolution. P. Strasser and co-workers191 added various concentrations of halides into KHCO3 electrolytes and observed that the addition of Cl− and Br− resulted in increased CO selectivity. The adsorbed I− exhibited a larger effect on CH4 production than C2H4. The probable reason was the induced negative charge possessed a remarkably positive effect favoring the protonation of CO. KCl was also used as a catholyte because it resulted in a higher local pH and the formation of bi-phasic Cu2O–Cu, favored for multicarbon fuel production.177
As indicated above, high concentrations, large cation sizes and halide additives could be considered for CO2 reduction in bicarbonate electrolytes in order to obtain more hydrocarbon products. Moreover, regarding the role of bicarbonate aqueous electrolytes, recently M. H. Shao’s group192 directly observed that CO2 molecules were mediated to the Cu surface via their equilibrium with bicarbonate anions rather than direct adsorption from the solution (Fig. 28), using attenuated total reflection surface-enhanced infrared absorption spectroscopy, isotopic labelling, and potential stepping techniques.
Fig. 28 The proposed role of bicarbonate anions in EC CO2 reduction. Reproduced from ref. 192 with permission from the American Chemical Society, 2017. |
Fig. 29 CO2 concentration in a mixture of CH3CN and H2O versus the H2O content. Reproduced from ref. 195 with permission from the Electrochemical Society, 2000. |
Therefore, organic solvents and ionic liquids are better choices for suppressing H2 evolution in catalyst systems not suitable for aqueous solutions. In aqueous systems, alkaline conditions could promote C–C coupling during CO2 reduction.207,208 For the ultimate goal of CO2 recycling, neutral aqueous solution is the best choice and various concentrations and cations could be applied to tune the activity.
Fig. 30 Schematic diagram for overall reaction of CO2 reduction and H2O oxidation using Cu electrode in H-type cell. Reproduced from ref. 209 with permission from the Royal Society of Chemistry, 2013. |
In a PEC CO2 reduction system, sunlight, visible light and UV light are three options for the light source. For the purpose of solar energy utilization, sunlight irradiation without bias potential is the ultimate goal. Cu2O with a direct band gap of ∼2.0 eV is a promising material for enhancing CO2 activity and inhibiting H2 evolution in PEC CO2 reduction systems. N. R. Tacconi, K. Rajeshwar and co-workers218,219 first reported the utilization of Cu2O for CO2 PEC reduction. They electrodeposited Cu2O nanocrystals on CuO nanorod arrays. 95% CH3OH was formed with a bias potential of +0.17 using this CuO@Cu2O nanorod array as a photocathode (AM 1.5, 70 mW cm−2). Modifying the Cu-based photocathode could lead to different enhanced CO2 reduction products. L. R. Baker and co-workers220 electrodeposited CuFeO2/CuO on FTO and used it as a photocathode. 80% CH3COO− was detected with a bias potential of +0.2 V under white-light LED illumination (100 mW cm−2). P. D. Yang and co-workers221 directly assembled Au3Cu nanoparticles on the surface of TiO2-protected silicon nanowire as a photocathode. Compared with the planar counterpart, lower overpotential or additional bias (∼120 mV lower) was needed to drive CO2 reduction to CO. 80% CO could be achieved at −0.2 V vs. RHE (LED light source with intensity 20 mW cm−2, wavelength λ = 740 nm).
When Cu-based electrocatalysts are used as a cathode, both the cathode and the photoanode could be manipulated to enhance the performance of the overall system. J. S. Lee and co-workers222 constructed a PEC system comprising a WO3 photoanode and Cu cathode for CO2 reduction. 71.6% carbonic products (65% CH4) were obtained at +0.65 V vs. RHE under visible light irradiation (>420 nm). M. Miyauchi, H. Abe and co-workers223 obtained 79% HCOO− when using Cu–Zn alloy film as a cathode in 0.1 M KHCO3 and SrTiO3 as a photoanode in 0.1 M KCl + 0.01 M NaOH under UV light illumination without applied bias potential. Y. S. Kang and co-workers224 engineered a (040) facet BiVO4 photoanode and integrated it with a Cu cathode for CO2 PEC reduction (Cu cathode|NaCl|BiVO4 photoanode). Different products were obtained by tuning the bias potential, such as 65.4% HCOO− (+0.75 V), 85.1% HCHO (+0.9 V), 6.89% CH3OH and 4.4% C2H5OH (+1.35 V) in 0.5 M NaCl (AM 1.5).
When using CuxO as a photocathode or cathode, corrosion will happen. Protective layers such as TiO2 were applied in studies done well by M. Grätzel’s group. J. L. Gong and co-workers225 introduced a simple strategy by using Cu2O as the cathode and TiO2 nanorods as the photoanode for PEC reduction of CO2. 92.6% carbonic products (54% CH4, 30% CO, 3% CH3OH) were obtained at +0.75 V vs. RHE bias potential (AM 1.5, 100 mW cm−2). Through cut-off filter experiments, they confirmed that the photogenerated electrons were not the main reason for Cu2O corrosion; instead, photogenerated holes were primarily responsible for the instability of Cu2O. As shown in Fig. 31, the photogenerated electrons were consumed at the electrode/electrolyte interface, while the holes moved to the back contact and the counter electrode. Back illumination (the travel distance of the electrons was longer than that of the holes) was favorable for the stability of Cu2O.
Fig. 31 Photocurrent–time curves of Cu2O photocathode under (a) different light sources and (b) front and back illumination. Experimental conditions: 0.5 M N2-saturated Na2SO4, 0 V vs. RHE, 100 mW cm−2. (c) The travel distances of electrons and holes under front and back illumination. Reproduced from ref. 225 with permission from Wiley, 2016. |
The photocurrent density is relatively low for the above reported systems, as listed in Table 2. One future aim is to improve the photocurrent density in PEC CO2 conversion. Constructing a hybrid catalyst consisting of a molecular catalyst and semiconductor material for PEC CO2 reduction could improve the selectivity and efficiency.226–228 For a hybrid system, careful design of the semiconductor is also important. As shown in Fig. 32a, M. T. Mayer, M. Grätzel and co-workers229 designed a heterogeneous catalyst system by covalently immobilizing molecular Re(bpy) (CO)3Cl onto a TiO2-protected Cu2O photocathode (Cu2O/AZO/TiO2) via phosphonate linkers. On the TiO2 layer there was also a mesoporous film of 4.5 to 5 μm thickness with 18 nm TiO2 particles. The catalyst system without mesoporous TiO2 did not show substantial photocurrents, while mesoporous TiO2 (enhanced roughness and catalyst loading) exhibited enhanced photocurrents (Fig. 32b). 80–95% CO was achieved under chopped light illumination with photocurrent density of 2.5 mA cm−2 at −1.9 V vs. Fc+/Fc.
Fig. 32 (a) Schematic of hybrid system of protected Cu2O photocathode with covalent-bound Re(bpy) (CO)3Cl. (b) Corresponding PEC performance with and without mesoporous TiO2 layer on the surface of protected Cu2O photocathode. Experimental conditions: CO2-saturated CH3CN (0.1 M Bu4NPF6), scan rate 10 mV s−1, 100 mW cm−2. Reproduced from ref. 229 with permission from the American Chemical Society, 2016. |
Fig. 33 (a) Schematic of PEC CO2 reduction with PV cell. (b) Solar-to-CO conversion efficiency and selectivity of CO versus time. Experimental conditions: 0.1 M CsHCO3‖0.25 M CsOH, −0.55 V vs. RHE, PV cell (0.5625 cm2, 100 mW cm−2, 2.38 V). Reproduced from ref. 61 with permission from Springer Nature, 2017. |
The use of other forms of sustainable energy such as solar energy to drive the overall reaction of CO2 reduction and water splitting is promising. CuxO could be used as both a cathode and photocathode after surface protection, as stated above. It is also promising to use Cu oxide species directly as cathodes and photocathodes, since the Schottky junction between Cu2O and Cu could facilitate electron and hole separation, leading to enhanced activity.236 The morphology and activity evolution of CuxO is worthy of further investigation because the chemical changes of CuxO due to photocorrosion and electroreduction processes were indeed pre-activation steps for CO2 reduction, as discussed in this review for oxide-derived Cu.
In this review, the latest progress on Cu-based heterogeneous electrocatalysts for EC CO2 reduction was discussed. We summarized the benchmark activity for specific products in Section 3 and various types of Cu-based materials reported by different research groups in Section 4. The H2 evolution reaction is inevitable because its equilibrium potential is lower than that of CO2 reduction. Apart from engineering the material, therefore, adjusting the electrolyte composition from aqueous to non-aqueous, adding ionic liquid or other additives, and careful pre-treatment of the system are also important. Through the discussions in this review, we hope we could provide useful information to newcomers to the field through detailed information about the experimental conditions, and to those already experienced in the topic through the comparison data in Table 2 and our focus on the more recent literature about Cu-based heterogeneous electrocatalysts (2013–2019).
Electrocatalysts | Electrolyte saturated with CO2, flow rate of CO2 gas, and reference electrode used | Faradaic efficiency (FE) and current density for major products, and products distribution | Ref. |
---|---|---|---|
All potentials are converted to versus RHE if not specified | |||
a The potential applied for EC CO2 reduction versus RHE is calculated according to the equation E (vs. RHE) = E0 (vs. Ag/AgCl) + E (vs. Ag/AgCl) + 0.0591pH − η(IRdrop) in aqueous solution. E0 (vs. Ag/AgCl) = 0.21 V, 0.197 V, and 0.21 V for 3 M KCl, saturated KCl, and 3 M NaCl, respectively. | |||
Morphology | |||
Cu sheet | 0.1 M KHCO3 | CO and HCOO− as main products (>−0.6 V); hydrocarbons and alcohols are favorably produced (<−0.7 V) | 1989 (ref. 34) |
0.1 M KHCO3 | 29% CH4, 30% C2H4, 7% EtOH, 9% HCOO−, 11% H2 (−0.81 V), current density 5.0 mA cm−2 | ||
0.1 M KClO4 | 10% CH4, 48% C2H4, 16% EtOH, 9% HCOO−, 7% H2 (−0.89 V), current density 5.0 mA cm−2 | ||
0.1 M K2HPO4 | 17% CH4, 5% HCOO−, 72% H2 (−0.64 V), current density 5.0 mA cm−2 | ||
Cycling-derived Cu nanoparticle (50–100 nm) covered Cu foil | 0.1 M KClO4 (pH = 6.0) | Electropolished (smooth): 14% C2H4 and 5% CH4 (at −1.1 V), LSV onset potential −1.0 V | 2012 (ref. 114) |
Nanoparticle covered: 36% C2H4 and 1% CH4 (at −1.1 V), LSV onset potential −0.6 V | |||
Sputtered: 26% C2H4 and 8% CH4 (at −1.1 V), LSV onset potential −0.8 V | |||
Cu nanoparticles on glassy carbon plate | 0.1 M KHCO3, 30 sccm CO2 | 5–15 nm: 20–25% CO, 60–70% H2, 10–15% CH4, 0–10% C2H4 (−1.1 V), current density −48 mA cm−2 for 2 nm | 2014 (ref. 65) |
Ag/AgCl | Cu foil: 5% CO, 20% H2, 57% CH4, 20% C2H4 (−1.1 V), current density −23 mA cm−2 | ||
Cu nanoparticle spin-coated on glassy carbon (7 nm evolved to 23 nm during reaction) | 0.1 M NaHCO3, 20 sccm CO2 | Cu nanoparticle (7–25 nm): 76% CH4 at −1.35 V with partial current density 9.5 mA cm−2, FE onset potential −0.95 V | 2014 (ref. 77) |
Ag/AgCl | Cu foil: 44% CH4, 35% H2 (−1.35 V) | ||
Cu nanoparticles (100 nm) on carbon paper by pulse electrodeposition | 0.5 M NaHCO3, 1 sccm CO2 | 85% CH4 at −2.1 V with partial current density 38 mA cm−2, FE onset potential −1.1 V | 2017 (ref. 115) |
SCE | |||
Amorphous Cu nanoparticles (3.3 nm) on carbon cloth | 0.1 M KHCO3, 20 sccm CO2 | Amorphous (3.3 nm): 37% HCOOH, 22% C2H5OH (−0.8 V), partial current density 6 mA cm−2, FE onset potential −0.6 V | 2018 (ref. 78) |
1 mg/1 × 1 cm2 | Ag/AgCl (saturated KCl) | Crystalline (3.4 nm): 26% HCOOH (−0.8 V), 6% C2H5OH (−1.1 V) | |
Cu nanoparticles (2–4 nm)/C on glassy carbon plate | 0.1 M KHCO3, 20 sccm CO2 | 57% C2H4 at −1.1 V, current density −17.5 mA cm−2 | 2019 (ref. 121) |
0.2 mg/0.5 cm2 | Ag/AgCl (3 M NaCl) | ||
Branched CuO nanoparticles drop-casted on glass carbon | 0.1 M KHCO3, 60 sccm CO2 | 70% C2H4 at −1.0 V with partial current density −17.0 mA cm−2 | 2019 (ref. 106) |
0.2 mg/0.5 cm2 | Ag/AgCl | ||
Cu nanocubes spin-coated on glassy carbon | 0.1 M KHCO3, 5 sccm CO2 | 63 nm: 63% for carboneous products: 25% C2H4, 11% CH4 | 2016 (ref. 129) |
0.2 mg/2.5 × 2.5 cm2 | Ag/AgCl | 44 nm: 80% for carboneous products: 41% C2H4, 20% CH4 | |
24 nm: 43% for carboneous products: 9% C2H4, 14% CH4, 15% HCOO− | |||
At −1.1 V, current density −0.5, −5.5, −7.5 mA/cmECSA2 respectively | |||
Cu foil: 20% C2H4, 20% CH4, 7% HCOO− (−1.1 V) | |||
Electrodeposited Cu nanofoam (pore sizes 20–50 μm) | 0.1 M KHCO3 (pH = 6.8) | Major products: HCOOH, H2, and CO | 2014 (ref. 88) |
Ag/AgCl | Minor products (<2%): C2H4, C2H6, CH4, and C3H6 | ||
Not quantified (≪1%): CH3OH, C2H5OH | |||
The onset potential for HCOOH is −0.4 V; with a FE of 26% at −0.5 V; with a maximum FE of 37% at −0.9 V | |||
The FE for smooth copper electrode is <1% at −0.5 V, 25% at −0.9 V | |||
Cu mesopore (pore size/depth, 300 nm/40 nm, 30 nm/40 nm, 30 nm/70 nm) | 0.1 M KHCO3 | Cu main products: C1(CO 2% + CH4 46%) 48%, C2(C2H4) 8% at −1.3 V, current density 10 mA cm−2, LSV onset potential −0.56 V | 2016 (ref. 133) |
Ag/AgCl (3 M KCl) | 300 nm/40 nm: C1(CO 2% + CH4 16%) 18%, C1(HCOOH)32%, C2(C2H4)9% | ||
30 nm/40 nm: C1(CO 5% + CH4 14%)19%, C2(C2H4)38% | |||
30 nm/70 nm: C1(CO 8% + CH4 6%)14%, C2(C2H4)3%, C2(C2H6)46% | |||
At −1.3 V, current density 14.3 mA cm−2, LSV onset potential −0.56 V | |||
Nanoporous CuDAT-wire | 1 M KOH (flow cell) | A maximum of C2H4 40% with partial current density 40 mA cm−2 and C2H5OH 20% with partial current density 15 mA cm−2 at −0.5 V, FE onset potential −0.35 V | 2017 (ref. 87) |
(DAT-3,5-diamino-1,2,4-triazole, additive during electrodeposition) | Ag/AgCl | ||
0.28 mg cm−2 | |||
3D porous hollow fibre Cu (hydrogenation of CuO) | 0.3 M KHCO3, 20 sccm CO2 | 72% CO at −0.4 V with partial current density 5.6 mA cm−2, FE onset potential −0.15 V | 2016 (ref. 31) |
Ag/AgCl (3 M NaCl) | |||
Oxide-derived Cu | |||
Annealed Cu foil (thick Cu2O)-derived Cu | 0.5 M NaHCO3 (pH = 7.2), 5 sccm CO2 | Cu foil: a maximum of CO 20% at −0.8 V, 33% HCOOH at −0.7 to −0.9 V | 2012 (ref. 69) |
(≥∼3 μm thickness) | Ag/AgCl (3 M KCl) | Annealed Cu foil: a peak of CO 45% at −0.3 to −0.5 V, 33% HCOOH at −0.45 to −0.65 V, current density 0.5–7 mA cm−2 at −0.3 to −0.65 V | |
Annealed Cu gauze (CuO nanowire)-derived Cu (size 50–100 nm, length 10–50 μm) | 0.1 M KHCO3 (pH = 6.8), 20 sccm CO2 | Electrochemical reduction: CO + HCOOH 70–80% at −0.3 to −0.5 V; a peak of CO 61.8% at −0.4 V, HCOOH 30.7% at −0.6 V, current density 0.5–10 mA cm−2 at −0.3 V to −0.6 V | 2015 (ref. 80) |
Ag/AgCl (saturated KCl) | Forming gas reduction: H2 >90% at −0.3 to −0.5 V, FE onset potential for CO2 reduction is −0.6 V | ||
Annealed Cu(OH)2 foil (CuO nanowire)-derived Cu (length 8.1 μm) | 0.1 M KHCO3 (pH = 6.8) | 8% C3H7OH (FE onset potential −0.9 V), 4% C2H5OH (FE onset potential −0.7 V), 2% C2H6 (FE onset potential −0.7 V), 17% C2H4, 18% HCOOH, 40% H2 at −1.1 V | 2016 (ref. 134) |
Ag/AgCl (saturated KCl + AgCl) | Cu foil: 3% C2H4, 12% HCOOH, 80% H2 at −1.1 V | ||
Electrodeposited Cu2O film on Cu plate | 0.1 M KHCO3 (pH = 6.8), 5 sccm CO2 | A maximum of C2H4 33% with partial current density 12 mA cm−2, ratio of C2H4/CH4 8–12 (−1.1 V), FE onset potential −0.8 V | 2014 (ref. 72) |
Ag/AgCl (3 M NaCl) | |||
Electrodeposited Cu2O film on GDE | 0.5 M KHCO3 | A maximum of C2H4 26%, ratio of C2H4/CH4 up to 26 (−1.2 V), current density 12.5 mA cm−2 | 2015 (ref. 150) |
Ag/AgCl | |||
Electrodeposited Cu2O film (1.7–3.6 μm) on Cu disc | 0.1 M KHCO3, 20 sccm CO2 | A maximum of C2H4 34–39% and C2H5OH 9–16%, ratio of C2H4/CH4 up to 100 (−1.0 V), current density −30 mA cm−2 | 2015 (ref. 81) |
Ag/AgCl (saturated) | Cu2O-derived (0.9 μm): 40% C2H4, 9% C2H5OH, 8% HCOO− (−1.0 V) | ||
Cu foil: 9% CO, 14% C2H4, 13% HCOO− (−1.0 V) | |||
Electrodeposited/oxide-derived Cu nanofoam (pore sizes 50–100 μm) | 0.5 M NaHCO3 (pH = 7.2) | Major products: H2, CO, CH4, C2H4, C2H6, HCOOH | 2016 (ref. 131) |
Ag/AgCl (3 M KCl) | Not quantified (<1%): CH3OH, C2H5OH | ||
FE HCOOH <6% at any potentials | |||
A maximum FE CO of 45% in potentials ranging between −0.6 V and −0.7 V. A maximum of FE C2(C2H4, C2H6) 55% at −0.8 V with partial current density 6 mA cm−2 | |||
20% C2H4, 35% C2H6, 15% CO, 6% HCOO−, 15% H2 (−0.8 V) | |||
15% C2H4, 8% C2H6, 25% CO, 8% HCOO−, 25% H2 (−0.8 V)–Cu wafer | |||
Skeleton Cu (annealed, Cu foam electrodeposition) | 0.5 M NaHCO3 (pH = 7.2), 10 sccm CO2 | Electropolished Cu skeleton: at all potentials from −0.6 V to −1.3 V, CO <3% | 2017 (ref. 132) |
Ag/AgCl (3 M KCl) | Annealed Cu skeleton: CO 15% (−0.6 V), a maximum C2H4 + C2H6 32% (−1.1 V), FE onset for C2 −0.7 V | ||
Electrodeposited Cu foam on Cu skeleton: CO 19% (−0.6 V), a maximum C2H4 + C2H6 29% (−1.1 V), FE onset for C2 −0.7 V | |||
Dendritic Cu (electrodeposition, annealed) | 0.5 M KHCO3 (pH = 7.2), 10 sccm CO2 | Electrodeposited dendritic Cu:HCOOH 49% (−0.7 V); C2H4 34% (−1.1 V) | 2017 (ref. 156) |
Ag/AgCl (3 M KCl) | Annealed dendritic Cu:C2H5OH + C3H7OH 25% (−1.0 V, C2H5OH 13%) with partial current density 2.8 mA/cmECSA2; C3H7OH 13% (−0.9 V) with partial current density 0.9 mA/cmECSA2; FE onset potential −0.8 V | ||
Galvanostatic cycling in Cu2+-(Cu2O) derived Cu nanocubes on Cu foil | 0.25 M KHCO3, 50 sccm CO2 | 60% C2+(C2H4, C2H5OH, n-C3H7OH) with partial current density 42.5 mA cm−2 at −0.963 V, FE onset for C2 −0.705 V and C3 −0.857 V | 2018 (ref. 59) |
SCE | |||
Hydrothermally prepared Cu2O (1.3 and 11.2 μm) on Cu disc | 0.1 M KHCO3, 20 sccm CO2 Ag/AgCl (saturated) | Cu2O derived (1.3 μm): a maximum of C2H4 27% (−1.0 V), C2H6 <3% (−0.5 to −1.1 V) | 2017 (ref. 146) |
Cu2O derived (11.2 μm): a maximum of C2H4 15% (−0.8 V), C2H6 8% (−0.8 V) with FE onset −0.6 V, current density 20 mA cm−2 | |||
CuO derived (11.5 μm): a maximum of C2H4 13% (−0.8 V), C2H6 11% (−0.8 V) with FE onset −0.6 V, current density 20 mA cm−2 | |||
Plasma-treated Cu foil | 0.1 M KHCO3, 30 sccm CO2 | Cu foil: C2H4 30% at −0.9 V, FE onset potential −0.85 V | 2016 (ref. 95) |
Ag/AgCl | O2 plasma activated Cu:C2H4 60% (−0.9 V, FE onset potential −0.5 V) | ||
Plasma-treated Cl−-induced Cu2O–Cu nanocube | 0.1 M KHCO3, 20 sccm CO2 | C2 + C3 73% at −1.0 V C2H4 45%, C2H5OH 20%, n-C3H7OH 9%, current density 37 mA cm−2 at −1.0 V | 2017 (ref. 149) |
Ag/AgCl | |||
Cl−-induced Cu2O–Cu | 0.1 M KCl (catholyte), 0.1 M KHCO3 (anolyte) | From Cu2O: 46% C2 (22% C2H4 and 24% C2H5OH), 10% C3–C4 (9% C3H7OH, 1% C4H10) at −1.6 V, FE onset for C2 −0.6 V and C3–C4 −1.0 V | 2015 (ref. 160) |
Ag/AgCl (3 M KCl) | From Cu: 18% C2H4, 6% C2H5OH, 25% CH4 at −1.6 V | ||
Halide-induced Cu2O–Cu | 0.1 M KHCO3 (pH = 6.8), 5 sccm CO2 | Cycling Cu foil in halide: 16% C2H4, 8% C2H5OH at −1.0 V, current density 12 mA cm−2, FE onset for C2H5OH −0.6 V and C2H4 −0.8 V | 2016 (ref. 142) |
Ag/AgCl (saturated KCl) | Cu foil: 9%C2H4, 3%C2H5OH at −1.0V | ||
Cl−-induced Cu218O–Cu on Si (1 μm) | 0.1 M KHCO3, 5 sccm CO2 | Cycling Cu foil in KCl:C2/C3 60% (C2H4 34% with partial current density −7.5 mA cm−2), C1 6.5% at −1.0 V, current density −10.9 mA cm−2 | 2017 (ref. 143) |
Ag/AgCl | Cu foil: C2/C3 37%, C1 25% at −1.0 V, current density 6.9 mA cm−2 | ||
Cu2(OH)3Cl-induced Cu2O–Cu nanoneedle on carbon paper (electrodeposited/oxide-derived) | 0.1 M KHCO3, 20 sccm CO2 | 54% C2+(C2H4, CH3COOH, C2H5OH), 18% C1+ (CH4, CO, HCOOH), C2H4/CH4 = 130 at −1.2 V with a C2H4 partial current density 22 mA cm−2, FE onset potential −0.7 V | 2018 (ref. 82) |
Ag/AgCl (saturated KCl) | Flow cell: C2H4/CH4 = 200 at −1.0 V with a C2H4 partial current density 160 mA cm−2 | ||
Surface modification | |||
Amino acid-modified Cu(OH)2 nanowire derived Cu (diameter 100–200 nm, length tens of μm) | 0.1 M KHCO3 | Amino acid modified Cu nanowire: 13% C2H4, 21% C2H6, 52H2% (−1.3 V), with partial current density for CxHy 3 mA cm−2 | 2016 (ref. 182) |
Ag/AgCl (saturated KCl) | Cu nanowire: 6% C2H4, 12% C2H6, 76H2% (−1.3 V) | ||
Pd atoms-modified Cu mesh | 0.5 M KHCO3, 20 sccm CO2 | Cu: after 4 h, H2(100%) at −0.96 V | 2017 (ref. 60) |
Ag/AgCl | Pd-decorated Cu: initial and after 4 h, H2 (40%), 50% (CH4 + C2H4) at −0.96 V, current density 50 mA cm−2 | ||
SnO-modified CuO nanowire | 0.1 M NaHCO3 (pH = 6.75), 10 sccm CO2 | CuO nanowire: 36% CO with a partial current density −0.25 mA cm−2 at −0.7 V | 2017 (ref. 61) |
Ag/AgCl (saturated KCl) | SnO2-modifed CuO nanowire: 97% CO with a partial current density −1.1 mA cm−2 at −0.7 V | ||
Bimetallic species | |||
Dendritic Cu–In electrode (dendrite tips of ∼50 nm width) | 0.1 M KHCO3 | Pure dendritic Cu: at all potentials from −0.8 V to −1.1 V, HCOOH <10% | 2017 (ref. 166) |
Ag/AgCl | Pure In: a maximum of HCOOH 40% at −1.0 V | ||
Cu–in alloy: At −0.8 and −0.9 V, HCOOH 60% (60 or 80 at% In) | |||
At −1.0 V, 50% HCOOH and the ratio of H2:CO2 is 2.6:1(40 at% In); a maximum of HCOOH 62% (80 at% In) | |||
Cu–In alloy (in situ reduction of annealed/oxide-derived Cu in In solution) | 0.1 M KHCO3, 10 sccm CO2 | Oxide-derived Cu: a maximum of CO (40%) at −0.6 V, current density −1.7 mA cm−2, FE onset potential −0.3 V | 2015 (ref. 73) |
Ag/AgCl (saturated KCl) | Cu–in alloy: 95% CO at −0.7 V, current density −1.7 mA cm−2, FE onset potential −0.3 V | ||
Cu–Sn alloy (in situ reduction of oxide-derived Cu in Sn solution) | 0.1 M KHCO3, 10 sccm CO2 | >80% CO at −0.4 to −0.8 V, 90% CO with current density at −0.6 V | 2016 (ref. 74) |
Ag/AgCl (saturated KCl) | |||
Electrodeposited CuxZn oxides film | 0.1 M KHCO3, 20 sccm CO2 | A maximum of C2H5OH 29% at −1.05 V with partial current density −8 mA cm−2, FE onset potential −0.95 V | 2016 (ref. 84) |
Ag/AgCl (saturated) | |||
AuxCuy alloy nanoparticles assembled on glassy carbon | 0.1 M KHCO3, 30 sccm CO2 | Au: 60% CO with a partial current −180 mA mg−1 at −0.73 V, current density 2.6 mA cm−2 | 2014 (ref. 172) |
Ag/AgCl (1 M KCl) | Au3Cu: 67% CO with a partial current −230 mA mg−1 at −0.73 V, current density 2.6 mA cm−2 | ||
AuCu3: 15% CO with a partial current −40 mA mg−1 at −0.73 V, current density 1.4 mA cm−2 | |||
Cu: 2–3% CO with a partial current −10 mA mg−1 at −0.73 V, current density 0.7 mA cm−2 | |||
Ag–Cu alloy (Ag–Cu2O derived) | 0.2 M KCl (catholyte)|0.1 M KHCO3 (anolyte) | Phase-separated Ag–Cu2O: 20% C2H5OH at −1.2 V, FE onset potential −1.0 V | 2017 (ref. 177) |
Ag/AgCl | Phase blended Ag–Cu2O: 34%C2H5OH at −1.2 V, FE onset potential −0.9V | ||
Pure Cu2O: 11% C2H5OH, FE onset potential −1.0 V | |||
Electrodeposited Cu–Ag alloy on carbon paper (GDL) with the addition of 3,5-diamino-1,2,4-triazole | 1 M KOH, 7 sccm CO2 | 6% Ag in CuAg alloy: 60% C2H4 with a partial current density −180 mA cm−2 and 25% C2H5OH with a partial current density −90 mA cm−2 at −0.7 V | 2018 (ref. 102) |
Ag/AgCl | |||
Electrodeposited Cu–Ag alloy (∼6 nm) on polymer on glassy carbon | 0 °C, 0.5 M KHCO3(8 ppm benzotriazole) | 21% CH3COO− at −1.33 V, FE onset potential −1.1 V | 2018 (ref. 170) |
SCE | |||
Electrodeposited Cu–Pd alloy (∼6.6 nm) on polymer on FTO | 0.1 M TBAPF6/CH3CN (1 M H2O), Ag/AgNO3; 0.1 M KHCO3, SCE | Cu2Pd: 51% CH4 at with an overpotential 0.86 V, current density −6.0 mA cm−2 | 2015 (ref. 169) |
Electrodeposited Cu–Pd alloy on glassy carbon | 0.1 M KHCO3 | Pd7Cu3 80% CO at −0.8 V, current density −1.0 mA cm−2, FE onset potential −0.7 V | 2016 (ref. 173) |
Ag/AgCl (saturated KCl) | |||
Cu–Pd alloy nanoparticles on Vulcan carbon coated on carbon paper | 0.1 M KHCO3 | Pd85Cu15, 86% CO with a partial current density −6.9 mA cm−2 at −0.9 V, FE onset potential −0.6 V | 2016 (ref. 174) |
2 mg/1 × 2 cm2 | Ag/AgCl | ||
Cu–Pd alloy nanoparticles on GDE | 0.1 M KOH, 7 sccm CO2, electrolyte flow rate 0.5 sccm at −2 V to −3.5 V and 0.1 sccm at −1.6 V to −2 V vs. Ag/AgCl | Phase-separated CuPd: 48% C2H4, 15% C2H5OH, 15% CO at −0.7 V, FE C2H4 onset −0.4 V, FE C2H5OH onset −0.5 V | 2016 (ref. 86) |
4 mg/5.0 × 0.8 cm2 | Disordered CuPd: 12% C2H4, 5% C2H5OH, 50% CO, 4% CH4 at −0.7 V, FE C2H4 onset −0.6 V, FE C2H5OH onset −0.6 V | ||
Ordered CuPd: 75%CO, 2% CH4 at −0.7 V, FE C2H4 onset −0.7 V, FE C2H5OH onset −0.9 V | |||
Cu–Pd alloy nanoparticles on Vulcan carbon coated on carbon paper | 0.1 M KHCO3, 10 sccm CO2, | 87% CO at −0.9 V with partial current density 30 mA mg−1 (30 mA cm−2) | 2019 (ref. 241) |
1 mg cm−2 | Ag/AgCl (saturated KCl) | ||
Aqueous electrolyte | |||
Cu foil | 0.1 M MHCO3 (M = Li, Na, K, Rb, Cs, pH = 6.8), 5 sccm CO2 | C2H4/CH4 (Li+0.2, Na+0.3, K+0.7, Rb+1.8, Cs+3.3), H2 (Li+85.9%, Na+66.3%, K+55.1%, Rb+32.2%, Cs+31.1%) | 2016 (ref. 188) |
Ag/AgCl (saturated KCl) | C2H5OH (Rb+9.6%, Cs+11.4%) | ||
−1.0 V, current density increased from 2.5 mA cm−2 to 5.5 mA cm−2 with the increase of cation size | |||
Cu foil | 0.1 M KHCO3, 30 sccm CO2 | 15% CH4, 10% C2H4, 8% CO, 22% HCOO−, 40% H2 (−0.95 V), current density 2.5 mA cm−2 | 2016 (ref. 191) |
0.3 M KBr + 0.1 M KHCO3, 30 sccm CO2 | 5% CH4, 7% C2H4, 27% CO, 18% HCOO−, 40% H2 (−0.95 V), current density 2.5 mA cm−2 | ||
0.3 M KI + 0.1 M KHCO3, 30 sccm CO2 | 41% CH4, 17% C2H4, 5% CO, 10% HCOO−, 30% H2 (−0.95 V), current density 9.5 mA cm−2 | ||
Non-aqueous electrolyte | |||
Cu–Cu2O-[Cu(cyclam)](ClO4)2 composite on FTO | 0.1 M n-Bu4BF4 in DMF/H2O (97:3 v/v) | 90% HCOO− at −2.0 V vs. Fc+/Fc, current density −1.15 mA cm−2 | 2015 (ref. 196) |
Ag/AgCl (3 M KCl) | |||
Electrodeposited Cu2O on FTO | 0.1 M n-Bu4BF4 in DMF/H2O (99:1 v/v) | 90% HCOO− at −2.0 V vs. Fc+/Fc, current density −1.5 mA cm−2 | 2016 (ref. 197) |
Ag/AgCl (3 M KCl) | |||
Cu foil | 0.1 M supporting electrolyte in MeCN | CO, carbonate, bicarbonate as main products | 2016 (ref. 198) |
Ag/Ag+(0.1 M AgClO4 in MeCN) | |||
Porous dendritic Cu nanofoam (pore size 30–40 μm, 10 wt% Cu2O) on Cu plate | Ionic liquid [EMIM](BF4)/H2O (98:2 v/v) | Porous dendritic Cu: 83% HCOO− at −1.55 V vs. Fc+/Fc, current density −5.0 mA cm−2 | 2017 (ref. 206) |
Ag/AgCl (3 M KCl) | Cu foil: 45% HCOO− at −1.55 V vs. Fc+/Fc | ||
Phosphate modified dendritic Cu | 0.1 M supporting electrolyte in MeCN | 81% HCOO− at −1.45 V vs. NHE, current density −1.35 mA cm−2 | 2017 (ref. 199) |
Ag/Ag+ (0.1 M AgNO3 in MeCN) | |||
PEC | |||
Cu2O nanocrystal/CuO nanorod (photocathode)‖Pt (anode) | 0.1 M Na2SO4, simulated AM 1.5 illumination (70 mW cm−2) | 95% CH3OH, +0.17 V bias potential, photocurrent density 0.2 mA cm−2 | 2013 (ref. 218) |
Cu2O/AZO/TiO2 (photocathode)‖Pt (anode) | MeCN, 0.1 M Bu4NPF6, under filtered 450 W Xe arc lamp (1 sun) | 80–95% CO at −1.9 V vs. Fc+/Fc, photocurrent density 2.5 mA cm−2 | 2016 (ref. 229) |
Ag/AgCl wire | |||
CuFeO2/CuO (photocathode)‖Pt (anode) | 0.1 M NaHCO3 + KPi, white-light LED illumination (100 mW cm−2) | 80% CH3COO−, +0.2 V bias potential, photocurrent density 0.15 mA cm−2 | 2017 (ref. 220) |
Cu (cathode)‖WO3 (photoanode) | 0.5 M KHCO3(pH = 7.5), visible light (>420 nm) using a 500 W Hg lamp | 71.6% carbonic product (65% CH4), +0.65 V bias potential, photocurrent density 0.25 mA cm−2 | 2014 (ref. 222) |
Cu2O (cathode)‖TiO2 (photoanode) | 0.1 M KHCO3(pH = 6.9), AM 1.5G illumination (100 mW cm−2) | 92.6% carbonic products (54% CH4, 30% CO, 3% CH3OH), +0.75 V bias potential, photocurrent density 1.34 mA cm−2 | 2016 (ref. 225) |
Cu–Zn alloy (cathode)‖SrTiO3(photoanode) | 0.1 M KHCO3 (pH = 6.8), 0.1 M KCl + 0.01 M NaOH (anolyte), UV light Hg–Xe lamp (240–300 nm) | 79% HCOO− without bias potential | 2017 (ref. 223) |
Cu (cathode)‖BiVO4 (photoanode) | 0.5 M NaCl, 300 W Xe arc lamp with AM 1.5G filter (100 mW cm−2) | 65.4% HCOO− at 0.75 V bias potential, photocurrent density 0.1 mA cm2; 85.1% HCHO at 0.9 V bias potential, photocurrent density 0.35 mA cm−2; 6.89% CH3OH and 4.4% C2H5OH at 1.35 V bias potential, photocurrent density 0.75 mA cm−2 | 2018 (ref. 224) |
PV cell (GaInP/GaInAs/Ge)-CuO@SnO2 (cathode and anode) | 0.1 M CsHCO3 (pH = 6.75, catholyte)‖0.25 M CsOH (pH = 13.3, anolyte), AM 1.5G illumination (100 mW cm−2) and a voltage of 2.38 V | 13.4% solar to CO efficiency at +0.55 V, photocurrent density 12 mA cm−2 | 2017 (ref. 61) |
In summary, the design of efficient and selective Cu-based electrocatalysts is inspiring but still challenging. Several considerations may also be helpful for engineering efficient systems:
(1) PEC CO2 reduction with Cu-based materials. Here, we could manipulate two aspects, one of which is energy supply, such as solar energy and other forms of renewable and clean energy for lower applied overpotentials. Optimization of this technique may develop commercially feasible CO2 reduction systems. Another aspect to consider is the Cu (photo)cathode. The photoanode corrosion of copper oxide species is known by many researchers and studies have been done to prolong its stability under irradiation. However, the instability should not be considered a disadvantage when using the copper oxide species as (photo)cathodes for CO2 reduction. This is because oxide-derived Cu exhibited better performance than its parent counterpart. Moreover, Cu nanostructures possess localized surface plasmonic (LSPR) effects, which have been used for photocatalytic organic synthesis.237,238 To the best of our knowledge, there is no report yet for CO2 reduction utilizing Cu LSPR effects. Future efforts should be made towards this.
(2) Complete systems coupling the reduction of CO2 and H2O oxidation, as well as other significant oxidation reactions. Studying only the half reaction (CO2 reduction) is not sufficient to achieve a commercially feasible CO2 reduction system. The performance of anodic reactions also needs to be investigated. It is necessary to explore Cu-based materials as both cathodes and anodes.
(3) Hybrid system of Cu-based materials and metal complexes. Metal molecular catalysts or metal complexes could coordinate CO2/intermediates with the metal center and ligands. Covalent-attached metal complexes on Cu-based materials will combine the key features of Cu-based materials and allow molecular-level tunability. This may also address the large overpotential required for the Cu electrode and the instability issue of the molecular catalyst.
(4) 3D Cu porous networks. The pore length and pore size of 3D Cu porous structures will influence the diffusion of reactants and intermediates. Tailoring the pore length as well as the pore size from the macropore to mesopore, micropore, and nanopore region will tune the diffusion and, in turn, the performance of the catalyst. In addition, through designing abundant active sites in this porous network to prolong the retention time of specific intermediates, we could expect high selectivity for specific products.
(5) Cu nanoclusters. Although the size effects of Cu nanoparticles have been discussed widely, ultrasmall, atomically precisely controlled Cu nanoclusters are rarely reported. Similar to Au and Ag nanoclusters,239 investigating the reaction mechanism to get a fundamental understanding via the atomic precise control of Cu nanoclusters is critical to design highly efficient and stable Cu nanocluster electrocatalysts.
(6) Design of the flow cell. The fabrication of an efficient flow cell with a designed gas diffusion electrode (GDE) could dramatically promote the activity and stability of CO2 electrocatalysts. More and more attention has been paid to this area over the past two years,103,110,185,208,240 and future efforts could be also made towards this aspect.
This journal is © The Royal Society of Chemistry 2020 |