Abstract
Direct converting low concentration CO2 in industrial exhaust gases to high-value multi-carbon products via renewable-energy-powered electrochemical catalysis provides a sustainable strategy for CO2 utilization with minimized CO2 separation and purification capital and energy cost. Nonetheless, the electrocatalytic conversion of dilute CO2 into value-added chemicals (C2+ products, e.g., ethylene) is frequently impeded by low CO2 conversion rate and weak carbon intermediates’ surface adsorption strength. Here, we fabricate a range of Cu catalysts comprising fine-tuned Cu(111)/Cu2O(111) interface boundary density crystal structures aimed at optimizing rate-determining step and decreasing the thermodynamic barriers of intermediates’ adsorption. Utilizing interface boundary engineering, we attain a Faradaic efficiency of (51.9 ± 2.8) % and a partial current density of (34.5 ± 6.4) mA·cm−2 for C2+ products at a dilute CO2 feed condition (5% CO2 v/v), comparing to the state-of-art low concentration CO2 electrolysis. In contrast to the prevailing belief that the CO2 activation step () governs the reaction rate, we discover that, under dilute CO2 feed conditions, the rate-determining step shifts to the generation of *COOH () at the Cu0/Cu1+ interface boundary, resulting in a better C2+ production performance.
Subject terms: Environmental chemistry, Energy efficiency, Electrocatalysis, Electrocatalysis
The development of catalysts that operate under low concentration CO2 resembling industrial waste gases holds promise for CO2 reduction. Here, the authors report a vacuum calcination approach for regulating the Cu0/Cu1+ density on Cu-based catalysts that can electro-catalyze low-concentration CO2.
Introduction
Electrochemical reduction of carbon dioxide (CO2RR) offers a promising avenue for advancing carbon neutrality in an environmentally sustainable manner1–5. Presently, only metallic Cu-based catalysts exhibit the capability to catalyze CO2 into a diverse range of hydrocarbons, encompassing C2+ products such as C2H4 and C2H5OH which are in high demand in basic production6–12. In pursuit of high selectivity for C2+ products, previous studies on CO2 reduction reaction (CO2RR) have predominantly been conducted using pure CO213–18. It usually comes from large-scale CO2 capture and sequestration facilities. However, the meticulous enrichment and purification of feed gases still incur substantial capital and energy costs even with latest technologies. For instance, the estimated cost of capturing CO2 from biomass-based combustion power plant ranges from $150 to $400 per ton of CO2, with additional purification costs reported to be $70 to $275 per ton of CO2 depending on the emission source19–28. The practical CO2 purity obtainable from military facilities and industrial manufacturing exhausts, such as those from coal power plants and the steel/petrochemical industry, is relatively low, typically ranging from 5 to 33% (CO2 v/v)29–38. These exhaust gases are far from being fully utilized resulting in a considerable waste of available carbon source. Therefore, employing electrocatalytic CO2 reduction with industrially diluted CO2 as the feed gas, not only avoids the high costs associated with feedstock purification, but also facilitates the sustainable utilization of natural resources and the attainment of the United Nations Sustainable Development Goals (SDGs) 12 & 13, which underscores its considerable significance39–44.
Recent developments in the electroreduction of diluted CO2 have centered on two main aspects: (1) controlling the availability of CO2 (its concentration) to achieve enhanced catalytic performance or modulate the reaction pathway40,45–50; and (2) pioneering the development of efficient catalysts for the electroreduction of dilute CO2 while examining their high catalytic activity from a thermodynamic standpoint51–54. Further studies have utilized simulated flue gas as the feed source for electrocatalytic CO2 reduction35,48–50,55–66. These innovative catalysts are progressively evolving towards functionalized microporous metal-organic frameworks (MOFs) or active molecular enzyme analogues, which are designed for targeted CO2 capture or activation55,56,58,60–62,67–69. However, under low CO2 concentrations, the electroreduction products predominantly consist of C1 products (e.g., CO, CH4), which involve relatively simple reduction processes23,37,50,55,57,67,70,71. Most studies have overlooked the relatively low CO2 feed conditions (≤15% CO2 v/v) in efforts to mitigate the intense competition with hydrogen evolution72–76. Some studies have even used CO, an intermediate product of CO2RR, as the feed source in low-concentration systems to enhance the selectivity for C2+ products77. Consequently, the characterization of CO2 reduction reaction (CO2RR) under such extreme CO2 feed condition intervals close to realistic situations remains poorly understood. Moreover, the rate of CO2 reduction reaction (CO2RR) is closely linked to CO2 mass transport, which is highly influenced by the purity of the feedstock. However, the impact of dilute CO2 feedstock on the kinetics of CO2RR remains insufficiently studied.
Achieving superior selectivity for C2+ products under dilute feedstock conditions (≤15% CO2 v/v) necessitates the construction of catalytically active sites that can promote CO2 conversion rate. Defect engineering is a prevalent strategy on Cu-based catalysts to manipulate reaction intermediates owing to its unique atomic arrangement78–82. Among these defects, the adjacent Cu0/Cu1+ interface boundary has been recognized as highly active catalytic sites for promoting C2+ electroproduction in conventional CO2RR83–93 (i.e., CO2RR in pure CO2 feed). The synergistic combination of Cu0 and Cu1+ active species enhances the electron transfer of CO2 and promotes *CO dimerization79,86,94–103. Furthermore, several experiments have demonstrated that Cu1+ species enhance the absorption strength of crucial carbon intermediates on the catalytic surface, thereby stabilizing the Cu1+ species104. Building on these research endeavors, we foresee that the Cu0/Cu1+ interface holds promise for achieving outstanding selectivity towards C2+ products under dilute-purity feed conditions. Nonetheless, the precise construction of quantifiable Cu0/Cu1+ interface boundaries and the exploration of their distinctive kinetic properties in low-purity CO2 feeds pose significant challenges.
We demonstrate efficient production of C2+ compounds under low CO2 feed conditions (5% v/v) using Cu-based catalysts with abundant Cu0/Cu1+ interface boundaries. We utilize a straightforward and stable technique to regulate the density of the Cu0/Cu1+ interface between Cu(111) and Cu2O(111) facets through adjustments in vacuum calcination duration. We discovered that catalytic sites at the Cu0/Cu1+ interface boundary display higher catalytic responsiveness with dilute-purity CO2 feeds than with high-purity CO2 feeds. Specifically, the selectivity of Cu-based catalysts for C2+ compounds correlates positively with the density of Cu0/Cu1+ interface boundaries in low CO2 feed conditions, but this correlation diminishes in high CO2 feed conditions. Consequently, we attained a faradaic efficiency of (51.9 ± 2.8) % for C2+ products at a low CO2 concentration (5% v/v) using the catalyst featuring the most abundant Cu0/Cu1+ interface. Due to the unique catalytic sites at the Cu0/Cu1+ interface, the CO2RR in the alkaline electrolyte microenvironment with dilute feed purity exhibits a distinct reaction mechanism compared to the widely accepted mechanism under high-purity feeds: the rate-determining step shifts backwards to the *COOH formation () instead of maintaining its prior CO2 activation (). Furthermore, we investigate the thermodynamic impact of catalytic sites at the Cu0/Cu1+ interface on dilute CO2 electrolysis. Our research offers fresh kinetic insights into the catalytic mechanism of CO2RR under dilute feed conditions and underscores the significance of directly utilizing dilute industrial CO2 emissions, which is often overlooked in current studies on heterogeneous electrocatalysis.
Results
Preparation and characterization of catalysts
Our objective was to create Cu-based catalysts with a unique fragmentation structure for CO2RR experiments. We selected commercial σ-Cu enriched in the Cu(111) and Cu2O(111) facet as the catalyst precursor. High-level vacuum calcination can delicately modify the surface crystalline structure of the precursor while simultaneously stabilizing its chemical composition. Using this approach, we produced highly fragmented copper with abundant Cu0/Cu1+ interface boundaries, termed high-boundary-density copper (HB Cu), through vacuum calcination of the precursor at 250 °C for 30 min (see Methods). To examine the relationship between the Cu0/Cu1+ interface structure and C2+ selectivity, we crafted control samples with nearly identical chemical compositions and adjustable Cu0/Cu1+ interface density. Medium- and low-boundary-density copper samples (MB Cu and LB Cu, respectively. Figure 1a were prepared as controls by vacuum calcination of commercial σ-Cu at 250 °C for 90 and 150 min, respectively. Furthermore, to eliminate the influence of Cu valence, we created a pure Cu catalyst by annealing commercial σ-Cu in an H2 atmosphere at 250 °C for 2 h. Scanning electron microscopy (SEM) images confirm the dispersed nanoparticles morphology of all Cu-based catalysts (Fig. 1b–d and Supplementary Figs. 1, 3, 5, 7, 9). Analysis of transmission electron microscopy (TEM) images reveals that the particle sizes of these catalysts are quite similar, ranging from approximately 70 to 130 nm, with a slight increase observed with longer annealing times (Fig. 1e–g and Supplementary Figs. 2, 4, 6, 8, 10). X-ray diffraction (XRD) analysis revealed that the chemical compositions of the Cu-based catalysts included Cu(111) and Cu2O(111), while pure Cu exclusively exhibited the Cu(111) facet.
Fig. 1. Characterization of the copper electrocatalysts.
a Schematic illustration for the preparation of the HB Cu, MB Cu and LB Cu catalysts. TEM of the HB Cu (b), MB Cu (c), and LB Cu (d) catalysts before CO2RR, respectively. Scale bars, 100 nm. Insets show the particle size distribution of the corresponding catalysts. Source data are provided as a Source Data file. HRTEM images showing crystallographic information of HB Cu (e), MB Cu (f), and LB Cu (g) catalysts before CO2RR, respectively. Scale bars, 5 nm. Locally magnified images showing the magnified grain boundaries. Scale bars, 2 nm. HRTEM images showing crystallographic information of HB Cu (h), MB Cu (i), and LB Cu (j) catalysts after 30 min of CO2RR, respectively. Scale bars, 10 nm. Locally magnified images showing the magnified grain boundaries. Scale bars, 2 nm. The yellow and orange dashed lines circle the facet fragments of the Cu(111) and Cu2O(111), respectively, and the white dashed line highlights the interface boundary between the Cu(111) and Cu2O(111) facets.
As depicted in Supplementary Fig.11, the X-ray diffraction (XRD) analysis revealed that all samples primarily consisted of Cu(111) with a minor presence of Cu2O(111), whereas pure Cu exhibited solely the Cu(111) facet. The diffraction peaks for Cu(111) and Cu2O(111) narrow and sharpen with longer calcination times, suggesting a gradual shift in the crystalline phase from concentrated fine fragments to dispersed bulk structures105. The average sizes of the crystalline phase fragments in the three samples, determined using the Scherrer equation, increase with longer calcination times (Supplementary Table 1). This trend corresponds to the changes observed in XRD diffraction peaks. X-ray photoelectron spectroscopy (XPS) and Cu LMM Auger spectra (Supplementary Figs. 12, 13) provide further confirmation of the distinctions and occupancy between Cu1+ and Cu0. The comparable valence ratios observed in all three samples dismiss Cu1+: Cu0 content as a significant factor influencing CO2RR performance79 (Supplementary Table 2).
Characterization of crystalline structure
We accurately determined the crystalline structure of HB Cu and the control groups using High-resolution Transmission Electron Microscopy (HRTEM) imaging. We then employed a random sampling method for normalized statistical analysis of the HRTEM images. Cu0 is mainly present in Cu(111) planes (outlined in yellow), exhibiting a lattice spacing of 2.1 Å, and in trace amounts of Cu(200) (outlined in yellow) with a lattice spacing of 1.8 Å. The interplanar spacing of 2.5 Å corresponds to Cu2O(111) planes (outlined in orange) interplanar spacing. HB Cu comprises a considerable amount of fragmented Cu(111) and Cu2O(111) closely spaced, leading to abundant Cu0/Cu1+ interface boundaries (outlined in white in Fig. 1e–g). These boundaries are anticipated to enhance the absorption of vital carbonate intermediates. MB Cu exhibits enlarged and dispersed facet fragments, resulting in a moderately abundant presence of Cu0/Cu1+ interface boundaries. Conversely, LB Cu primarily comprises uniform Cu(111) facet fragments, with the interfaces being nearly undetectable (Supplementary Figs. 14–16 present additional HRTEM images of HB Cu, MB Cu and LB Cu). The dimensional characteristics of the grain fragments in the HRTEM images closely correspond to the results of the Scherrer equation. To accurately grasp the structural characteristics of the three catalysts, we quantified the density of Cu0/Cu1+ interface boundaries in a large number of random HRTEM images, normalizing their length to the sample’s areas77. Supplementary Fig. 17 illustrates that HB Cu, possessing abundant fragmented crystalline structure, has relatively high density of Cu0/Cu1+ interface boundaries (22.67 μm−1), while MB Cu (14.92 μm−1) and LB Cu (0.94 μm−1) exhibit lower interface boundaries density. This characterization evidence indicates that the interface boundary density follows the sequence HB Cu>MB Cu>LB Cu. To further verify the interface boundary density of these Cu-based catalysts, CO stripping voltammetry was conducted for HB Cu, MB Cu and LB Cu (Supplementary Fig. 18). The results showed that the anodic peaks associated with CO electrooxidation of HB Cu (~1.030 V vs. RHE) exhibit a significant negative shift relative to MB Cu (~1.055 V vs. RHE) and LB Cu (~1.065 V vs. RHE). This suggests that grain boundaries serve as active sites for CO stripping and that increasing the Cu0/Cu1+ interface boundary density lowers the overpotential required for CO oxidation. The order of exposed active site density among the three catalysts is HB Cu>MB Cu>LB Cu, consistent with the random sampling observations from HRTEM. Supplementary Figs. 19–24 offer insights into the morphology and physical composition of all Cu catalysts after 30 min of CO2RR. The surface structural features exhibit minimal changes, with the gradient trends in crystalline interface boundary density between Cu(111) and Cu2O(111) remaining largely unchanged.
Electrocatalytic performance
We subsequently assessed the CO2RR performance of all Cu-based catalysts across various CO2 purities by adjusting the volume ratio of CO2 to N2 in the feed gas stream within the flow cell (Supplementary Fig. 25). Figure 2a, d illustrate the faradaic efficiencies of C2+ products generated by HB Cu across different CO2 purity feedstocks, with the corresponding partial current densities shown in Supplementary Fig. 26. When decreasing the incoming gas purity from pure to 2.5%, we observed a general decrease in the current densities of the C2+ products, particularly noticeable for feed gas purities below 30%, although not precisely mirrored in their faradaic efficiency. In the less negative potential range (−0.51 V ~ −0.71 V vs. RHE), the improvement in C2+ faradaic efficiencies shows an exponential rise at CO2 concentrations of 2.5%, 5%, and 10%. This upward trend continues down to −0.91 V vs. RHE and extends into more negative potential ranges for CO2 concentrations above 30%. The performance across these potential intervals suggests that there is ample *CO2 availability on the catalyst surface for CO2 electroreduction. Increasing the applied potential and decreasing feed purity result in decreased CO2RR efficiency due to insufficient *CO2 supply for electroreduction. This leads to mass transfer limitations and faint hydrogen evolution, consistent with prior research findings. At 5% CO2, a peak FEC2+ of (51.9 ± 2.8) % is attained at −0.71 V vs. RHE (Fig. 2h). The corresponding partial current density (normalized to the geometric area of the GDE, JC2+) was (34.5 ± 6.4) mA·cm−2, ranking relatively high among recent studies investigating electrolysis under dilute CO2 feeds (Supplementary Tables 9-10). Samples with fewer Cu0/Cu1+ interface boundaries (MB Cu and LB Cu) also exhibit catalytic performances towards C2+ products with volcano-shaped curves, influenced by mass transfer limitations (Supplementary Figs. 27, 28). At 5% CO2, MB Cu and LB Cu only reach maximum FEC2+ values of (32.6 ± 4.2) % and (25.2 ± 3.5) %, respectively (Fig. 2h). The C2+ selectivity of the samples exhibits a positive correlation with Cu0/Cu1+ interface density (Fig. 2c, f). Furthermore, in achieving 0.1 A·cm−2 in electrolysis, HB Cu demonstrates a lower onset potential compared to MB Cu and LB Cu, indicating its higher intrinsic activity (Fig. 2i). In comparison to the plot in 5% v/v CO2 (Fig. 2b), the C2+ selectivity of samples remains comparable in pure CO2 (Fig. 2e), indicating that the dominance of Cu0/Cu1+ interface is not evident under conditions abundant in mass transfer. To exclude the possibility of N2 electroreduction reaction(N2RR) on the Cu0/Cu1+ interface, we also examined the main products of N2RR on HB Cu: NH3 and N2H4 (Supplementary Figs. 29, 30). In addition, several control experiments were conducted: (1) commercial σ-Cu nanoparticles without calcination, (2) pure Cu nanoparticles after H2 reduction and (3) Cu-based catalysts with varying Cu/Cu2O ratios obtained through air calcination (the synthesis is shown in Methods). All comparison catalysts were tested under a 5% v/v CO2 feed and exhibited different C2+ selectivities, particularly the pure Cu, which lacked a detectable Cu0/Cu1+ interface. Moreover, the selectivity for C2+ products in Cu-based catalysts with varying Cu/Cu2O ratios was strongly and positively correlated with the Cu0/Cu1+ interface boundary density (Fig. 2f and Supplementary Figs. 31–38). These experimental observations indicate that the Cu0/Cu1+ interface boundary significantly reduces mass transfer limitations, thereby enhancing C2+ selectivity under dilute CO2 feed conditions.
Fig. 2. CO2RR electroreduction performance.
a Faradaic efficiencies (FE) for C2+ products produced by HB Cu at 2.5%, 5%, 7.5%, 10% and 15% CO2 feed concentration over a range of potentials. Curves connecting data points serve as visual guides. b Faradaic efficiencies (FE) for C1 and C2+ products as well as current densities (j) for CO2 electroreduction produced by HB Cu, MB Cu, LB Cu and pure Cu at 5% CO2 feed concentration over a range of potentials. c The FEC2+ of HB Cu, MB Cu and LB Cu during CO2 electroreduction process at 5% CO2 feed concentration. d Faradaic efficiencies (FE) for C2+ products produced by HB Cu at 30%, 60% and pure CO2 feed concentration over a range of potentials. Curves connecting data points serve as visual guides. e Faradaic efficiencies (FE) for C1 and C2+ products as well as current densities (j) for CO2 electroreduction produced by HB Cu, MB Cu, LB Cu and pure Cu at pure CO2 feed concentration over a range of potentials. f The Cu0/Cu1+ interface boundary length per area, measured from the HRTEM images of each sample, plotted against the C2+ products selectivity. g Comparison of FEC2+/FEC1 ratios on HB Cu, MB Cu and LB Cu across various CO2 concentrations. h Comparison of Faradaic efficiencies (FE) of different products produced by the three catalysts under potentials at which the maximum FE for C2+ products is reached. i Partial current densities for CO2 electroreduction under pure N2 atmosphere (dashed line) and 5% CO2 feed concentration (solid line) conditions on HB Cu (red line), MB Cu (blue line), LB Cu (gray line). j Horizontal comparison of several factors for HB Cu, MB Cu and LB Cu. There is no iR correction for voltages. Error bars represent the standard deviation based on at least three separate measurements. Source data are provided as a Source Data file.
Figure 2g further demonstrates the influence of Cu0/Cu1+ interface boundary density on product distribution, denoted as FEC2+/FEC1. In CO2 purities below 30%, FEC2+/FEC1 increases with decreasing CO2 purity, suggesting that dilute-purity CO2 has a greater impact on product distribution compared to the Cu0/Cu1+ interface boundary. Notably, we hypothesize that catalyst surface roughness may also impact CO2 electroreduction properties. Therefore, we determined the electrochemically active surface area of the three samples by conducting cyclic voltammetry (CV) scans at various scan rates (Methods and Supplementary Fig. 39). The three samples, each with varying interface boundary densities, exhibited only minor differences in electrochemically active area. Interestingly, we found no correlation between the electrochemically active area and C2+ selectivity performance under the 5% CO2 v/v feed condition. This further confirms the strong correlation between Cu0/Cu1+ interface boundary density and C2+ selectivity.
Insights into CO2RR under dilute purity feeds
Given that CO2RR performance under dilute purity feeds is often hindered by mass transport limitations, which significantly impede electrocatalytic reaction rates, it is imperative to thoroughly evaluate electrocatalytic kinetics under such conditions. In this study, we analyze the reaction kinetics of the HB Cu catalyst by examining plots of log(j) versus potential (V vs. RHE). Figure 3a–c present the kinetics data across the entire range of CO2 purity feeds for the primary electrocatalytic products—hydrogen, carbon monoxide, and ethylene. The calculated transfer coefficient (α = 2.303RT/F dlog(j)/dE) provides insight into the number of electrons transferred in the rate-determining and preceding steps, while the Tafel slope (dE/dlog(j)) identifies the rate-determining step (RDS) in the reaction process106–108 (the raw data is presented in the Supplementary Figs. 40–43 and Supplementary Tables 4–7).
Fig. 3. Insights into CO2RR mechanism from the reaction kinetics perspective and corresponding Operando Raman spectroscopy distinction.
Tafel plots of the HB Cu’s partial current density for H2 production (a), CO evolution (b), and C2H4 production (c) under varying CO2 purities in 1 M KOH at 25 °C. There is no iR correction for voltages. Operando Raman spectroscopy of HB Cu under pure CO2 condition (d) and 5% CO2 feedstock condition (e) over a range of potentials. f Peak *CO and *CO2− intensity during CO2RR on HB Cu under 5% CO2 feedstock condition (orange lines) and pure CO2 condition (blue lines). g Proposed rate-determining reaction mechanisms for CO2RR on rich Cu0/Cu1+ interface surface(left) and poor Cu0/Cu1+ interface surface(right). Source data are provided as a Source Data file.
Initially, we examine the kinetic characteristics of hydrogen production, which are particularly sensitive to disruptions under dilute CO2 feeds. As depicted in Fig. 3a, the Tafel curves flatten as the CO2 feed purity decreases, with the slopes decreasing from 251 mV dec−1 to 93 mV dec−1. The transfer coefficients calculated from these curves are less than 0.5, suggesting that the formation of surface-bound H* species is the rate determining step in H2 evolution109, regardless of the CO2 feed purity (Detailed derivation is illustrated in the Supporting information). Notably, MB Cu and LB Cu,
which have sparser Cu0/Cu1+ interface boundary densities, exhibit similar kinetic characteristics in H2 evolution. Specifically, the transfer coefficients are less than 0.5 regardless of the CO2 feed conditions, with H* formation dominating H2 evolution (Supplementary Fig. 40 and Supplementary Table 7). This observation suggests that catalytic sites at the Cu0/Cu1+ interface boundaries contribute to regulating H* formation, and these interface boundaries facilitate the rate of H2 evolution as CO2 feed concentration decreases.
In our experimental setup, the kinetics of CO production demonstrates distinctive behavior in contrast to H2 production (Fig. 3b). This is analyzed by examining a mechanistic sequence that entails the rate-determining formation of surface-bound *CO2− species, which serves as the initial intermediate in CO production. As the CO2 feed purity reduced from pure to 2.5%, the Tafel slopes of HB Cu dramatically decrease from 112 mV dec−1 to 39 mV dec−1, corresponding to an increase in the transfer coefficient from 0.5 to 1. These kinetics data reveal an overall increase in the rate of CO production as the CO2 feed concentration decreases. Specifically, the Tafel slopes all fall below 57 mV dec−1 at CO2 feed concentrations below 7.5%, indicating a shift from the initial rate-determining CO2 activation step to the rate-limiting *CO2− hydrogenation. Upon reducing the CO2 feed purity from pure to 5%, both MB Cu and LB Cu samples showed Tafel slope changes from 133 mV dec−1 to 107 mV dec−1 and 192 mV dec−1 to 110 mV dec−1, respectively (Supplementary Fig. 40). These values all correspond to a transfer coefficient close to 0.5, suggesting that the initial CO2 activation step constantly remains RDS during CO formation on Cu catalysts with sparser Cu0/Cu1+ interface. These observations indicate that the Cu0/Cu1+ interface can notably facilitate CO2 activation, leading to a tendency to revert the rate-determining step, particularly evident in the HB Cu catalyst with a high proportion of Cu0/Cu1+ interface. This, in turn, improves the conversion rate of essential carbon intermediates (notably *CO2− and *COOH), thereby establishing a robust basis for subsequent C-C coupling reactions.
Although ethylene evolution shares a common *CO2− intermediate with CO production in the early stages of electroreduction. To differentiate it from the CO formation pathway and ensure the independence of C2+ product evolution route, we define the formation pathway of C2+ product as originating from *CO. As illustrated in Fig. 3c, the Tafel slopes of C2H4 catalyzed by HB Cu exhibit two distinct regions. Under fully saturated CO2 conditions, C2H4 evolution displays multiple parallel Tafel slopes of approximately 192 mV dec−1 with a transfer coefficient of less than 0.5. This indicates that the chemical coupling of two surface-bound carbon-based intermediates predominantly governs ethylene formation, consistent with earlier findings45,110–115. When CO2 purity decreases, we note a series of noticeably flattened Tafel slopes at 85 mV dec−1, along with a corresponding transfer coefficient nearing 0.5. This suggests an increased reaction rate while maintaining the same rate-determining process. Interestingly, as the CO2 feed purity decreased from pure to 5%, we observed a decrease in the Tafel slopes of both MB Cu and LB Cu, from 146 mV dec−1 to 113 mV dec−1 and from 158 mV dec−1 to 142 mV dec−1, respectively, showing comparable downward trends (Supplementary Fig. 40). The results suggest that the Cu0/Cu1+ interface boundary marginally enhances the kinetics of C-C coupling under low CO2 concentrations. However, the enhancement is not significant.
In order to identify essential reaction intermediates and validate the kinetic evolution mechanism in CO2 electroreduction under low CO2 concentration conditions, we performed operando Raman experiments on the three Cu-based catalysts, applying potentials ranging from −0.11 to −1.51 V vs. RHE. Figure 3d, e depict the operando Raman spectra of CO2RR on HB Cu under both pure and 5% CO2 feeds. A series of bands ranging from 140 to 290 cm−1 and a peak at 524 cm−1, identified as characteristic peaks of Cu2O, were consistently observed over HB Cu, indicating the persistent presence of the Cu2O(111) facet and interface boundary during catalysis (in contrast, the Cu2O peaks for MB Cu and LB Cu were less persistent. Refer to Supplementary Figs. 44, 45 for operando Raman spectra of MB Cu and LB Cu). In pure CO2 feed conditions, the peak at approximately 2090 cm−1, attributed to the stretching vibration of *C ≡ O (noted as ν*C≡O), consistently maintains prominence across the entire cathodic potential range. This suggests the buildup of *CO, aligning with a product distribution featuring increased selectivity for CO under high-concentration CO2 feed conditions. Nonetheless, the decrease in the ν*C≡O peak below −0.31 V vs. RHE under a 5% v/v CO2 feed condition suggests the swift depletion of *CO, aligning with an augmented selectivity towards C2+ products (Upper section of Fig. 3f). These findings corroborate the experimental results indicating an improved C2+ selectivity under dilute CO2 feedstock conditions. Moreover, the positive correlation between CO peak intensity and feed gas concentration indicates a strong dependence of *CO surface coverage on CO2 concentration. This insight can guide subsequent simulations of carbonate intermediate coverage on the catalyst surface. Additionally, the peaks at 1292 cm−1 and 1590 cm−1, attributed to the asymmetric stretching vibrations of *CO2− (νas*CO2-), exhibit greater prominence with 5% CO2 compared to pure CO2 conditions at low potentials (Lower section of Fig. 3f), confirming a shift of RDS to *COOH generation under dilute feeds, facilitating the accumulation of *CO2−. Conversely, under pure CO2 conditions, the formation of *CO2− is limited by the rate and challenging to preserve, aligning with the kinetic trends inferred from the Tafel slope. The pronounced weakening of the νas*CO2- peak with increasing potential can be ascribed to limitations in reactant transfer (Fig. 3g, Detailed derivation of the rate-determining step evolution is provided in the Supporting Information).
We aimed to understand the thermodynamic impact of the Cu0/Cu1+ interface boundary under CO2 feed conditions with varying purities through DFT simulations. According to the Henry’s law: local CO2 concentration is proportional to local CO2 partial pressure; local CO2 partial pressure is proportional to CO2 surface coverage. It is reasonable to express local CO2 concentration in terms of the coverage of key carbonate intermediates, such as *CO2− and *CO54,110,116–119. Based on the unstable adsorption of *CO2− species observed in simulation and the strong correlation between *CO peak intensity and CO2 feed concentration in Operando Raman measurements. *CO coverage was ultimately selected to represent CO2 feed concentration in this experimental system. Furthermore, since the *CO peak position in Operando Raman spectra changes gradually with applied potential, we infer that the active sites for *CO adsorption are in a dynamic adsorption state, and CO poisoning does not occur on the catalytic surfaces. We then investigated the coverage effect by computing the Gibbs free energies of reaction at low coverage (1/30 monolayer, ML) and high coverage (14/30 ML). Here, 1/30 ML and 14/30 ML represent the surface coverage of the crucial intermediate *CO under 5% and pure CO2 feed conditions, respectively120–122 (Fig. 4a and Supplementary Figs. 46–49). Thus, stable models of Cu(111), Cu2O(111) surfaces, and symmetric low-angle tilted Cu0/Cu1+ interface boundaries at both high and low coverage were constructed based on HRTEM images. Initially, we examined the adsorption of *CO on the four models. In the Cu0/Cu1+ interface model at low coverage, *CO predominantly binds to the boundary sites. We determined that the adsorption free energy of *CO on the Cu0/Cu1+ interface is −0.052 eV, which is stronger than that on the Cu(111) surface (0.191 eV) but weaker than that on the Cu2O(111) surface (−0.469 eV). At high *CO surface coverage, the adsorption free energy of *CO at the Cu0/Cu1+ interface is 0.113 eV, which is thermodynamically more favorable than on the pristine Cu(111) surface, where it is 0.191 eV. Consequently, the Cu0/Cu1+ interface exhibits stronger *CO binding than pure Cu, promoting the accumulation of the crucial intermediate *CO. To further validate the higher reactivity of the Cu0/Cu1+ interface towards *CO, CO stripping experiments were conducted on three catalysts (Supplementary Fig. 18). The results demonstrated that increasing in Cu0/Cu1+ interface boundary density on the Cu-based catalysts effectively reduces the overpotential required for *CO oxidation.
Fig. 4. DFT calculations for different surfaces regarding the effect of *CO coverage.
a Adsorption configurations and adsorption energies (ΔGads) of *CO at Cu(111), Cu2O(111), Cu0/Cu1+ interface boundary at low *CO coverage and Cu0/Cu1+ interface boundary at high *CO coverage. b The energy profiles of CO2 reduced to *CO and the subsequent C-C coupling at the Cu(111), Cu2O(111) and Cu0/Cu1+ interface boundary. c Reaction free energy difference between CO2 activation, *CO2− hydrogenation and *COOH dehydrogenation with consequential C-C coupling on Cu0/Cu1+ interface boundary at low (orange) and high *CO coverage (gray). The dark brick-red, grey and red spheres represent Cu, C and O atoms respectively. Raw computational and simulation data are provided as Source Data and Supplementary Data 1 files.
Subsequently, we investigated the CO2 electroreduction pathway to OC*COH on different surfaces, a recognized route for generating C2+ products. Using this pathway, we calculated the Gibbs free energy change (ΔG) for each intermediate formation during CO2 electrocatalysis on Cu, Cu2O, and Cu-Cu2O interface surfaces at 0 V (vs. RHE) and under low *CO coverage conditions (Fig. 4b). The ΔG value for *COOH formation on the Cu-Cu2O interface (0.20 eV, compared to the 0 eV state for CO2) was lower than that on the pure Cu surface (1.12 eV). Additionally, the ΔG value for C-C coupling on the Cu-Cu2O interface is 0.59 eV, significantly lower than those on the Cu and Cu2O surfaces. These observations demonstrate that C-C coupling predominantly takes place at the Cu0/Cu1+ interface boundary, particularly under dilute CO2 feed conditions. This finding aligns with our experimental results, indicating that the HB Cu catalyst, which has a more abundant Cu0/Cu1+ interface, exhibits higher selectivity towards C2+ products. To illustrate the impact of *CO coverage on the Cu0/Cu1+ interface boundary, we examined the energy changes along the CO2 reaction pathway using the Cu-Cu2O interface model. We observed a notable reduction in the activation energy for CO2 to *CO2− conversion as the *CO coverage decreased from 14/30 ML to 1/30 ML. This trend mirrors the subsequent hydrogenation of *CO2− to *COOH (Fig. 4c and Supplementary Table 8), indicating that the Cu0/Cu1+ interface boundary can thermodynamically promote the activation and electroreduction of CO2 under conditions of low *CO coverage. Specifically, at high *CO coverage, the activation of CO2 to *CO2− is the dominant factor controlling the overall reaction potential in the CO formation pathway, consistent with the kinetics inferred from Tafel slopes, which suggest that the activation of CO2 to *CO2− is the rate-determining step. As *CO coverage decreases, the dimerization of carbon-containing intermediates becomes thermodynamically favorable, aligning with the kinetic behavior indicated by the reduction of the Tafel slope observed during ethylene formation. The thermodynamic promotion of CO2 activation and reduction in CO2RR becomes more evident with increasing Cu0/Cu1+ interface boundary density. This is consistent with experimental findings demonstrating a positive correlation between the density of Cu0/Cu1+ interface boundary and the performance of C2+ products under dilute CO2 feed conditions.
To further investigate the electronic interaction between the Cu0/Cu1+ interface and the crucial intermediate *CO absorbed on the surface, we performed differential charge analysis (DCA) on the Cu0/Cu1+ interface and pure Cu (Cu(111) surface in this case) with adsorbed *CO. As illustrated in Fig. 5a–d, electrons are transferred from the two Cu surfaces to *CO intermediates. The calculated Bader charge reveals stronger electronic interactions for *CO adsorbed on the Cu0/Cu1+ interface (gaining 0.49 electrons) compared to those on Cu(111) surface (gaining 0.47 electrons). This enhanced *CO adsorption facilitates C-C coupling, which aligns with our experimental results. Raw simulation data are provided as Supplementary Data 1 file. Furthermore, to assess the catalytic stability of grain boundary-rich catalysts under low-concentration CO2 feeds, we conducted stability tests of CO2RR in a 1 M KOH electrolyte with a 5% v/v CO2 feed, observing stable CO2 reduction by HB Cu for 28.5 hours (Fig. 5e, f). During electrolysis, we maintained the test potential at −0.71 V vs. RHE, the optimal potential for C2+ products. The partial current density of C2+ products remained stable at (35 ± 5) mA cm−2, with a corresponding FE stable at (49 ± 3) % over the 28.5 h period. Based on our empirical observations, although a low-concentration CO2 feed is more representative of real production conditions and significantly reduces feed gas purification costs. Optimization of the experimental set-up and the electrocatalytic process is necessary to achieve low energy consumption and high stability for industrial-scale CO2RR application123–125.
Fig. 5. DCA models for Cu(111) and Cu0/Cu1+ interface and experiment setup & stability measurement of CO2RR at 5% CO2 feed concentration.
Plots of differential charge analysis for *CO absorbed on Cu(111) surface (front view: a top view: b) and Cu0/Cu1+ interface (front view: c, top view: d). Blue, red and gray balls stand for copper, oxygen and carbon atoms, respectively. The yellow and blue isosurfaces correspond to the electron accumulation and depletion regions, respectively. e Schematic illustration of electrocatalytic equipment with the gas diffusion electrode for CO2RR. f Stability measurement of CO2RR for 28.5 h of electrolysis at an optimum potential of −0.71 V vs. RHE at 5% CO2 feed concentration and current densities for each product. There is no iR correction for voltages. Source data are provided as a Source Data file. Raw simulation data are provided as Supplementary Data 1 file.
Discussion
Previous techno-economic studies have shown that electrolysis of low-concentration CO2 to produce sustainable chemicals is economically advantageous, as it avoids the high energy demands of feed gas enrichment. This process holds significant potential for environmental remediation, chemical production and other applicative scenarios38. However, poor product selectivity and low catalytic current densities (below the industrial benchmark of 200 mA·cm–2) are the major challenges for industrial-scale low concentrations CO2 electrolysis123–125. To address these challenges: firstly, improving the catalysts’ ability to effectively capture low-concentration CO2 and increase the local concentration of CO2 at catalytic active sites is critical to improve catalytic performance126–130. Secondly, enhancing the catalyst’s capacity to directly convert low-concentration CO2 would significantly improve product selectivity. However, it is worth noting that although some encouraging results have been achieved in low concentrations CO2 electrolysis69,131, the field is still in its early stages with many challenges to be addressed. These include unclear thermodynamics and kinetics mechanisms and the instability of catalytic systems under low-concentration conditions23,48,63.
In summary, we have devised a simple vacuum calcination approach to regulate the density of Cu0/Cu1+ interface boundaries on Cu-based catalysts. By manipulating the crystalline interface, we successfully crafted the Cu-based HB Cu catalyst, which displayed a high interface boundaries density of 22.67 μm−1. At a low flue gas concentration of 5% CO2 feed condition, the catalyst HB Cu exhibited a FEC2+ of (51.9 ± 2.8) % and robust stability during the CO2RR. Kinetic mechanism analysis based on the Tafel slopes engineering reveals that, only at low concentration CO2 feed conditions, the catalytic sites at the Cu0/Cu1+ interface facilitate a backward migration of the rate-determining step (from to ), facilitating the easy formation of carbon-containing intermediates, but intriguingly, this phenomenon does not occur with high purity CO2 feed conditions. Moreover, based on thermodynamic calculations, the energy barriers for CO2 activation and subsequent reduction at Cu0/Cu1+ interface boundary sites are significantly reduced at low CO2 purity comparing to high purity feed conditions. This kinetically and thermodynamically synergistic enhancement endowed Cu catalysts with abundant interface boundary sites with strong capability to directly catalyze low CO2 gas concentrations. HRTEM engineering and CO2RR performance investigated the correlation between crystalline interface boundary density and C2+ selectivity, while ECSA measurements and XRD analysis negated the impact of electrochemically active area and oxidation state content on the electrolysis performance. This study adjusts the density of the Cu0/Cu1+ interface boundary using a straightforward method and reveals the distinctive activity of interface boundary sites under nontraditional catalytic conditions. It not only establishing a green and sustainable electrocatalysis strategy for C2+ hydrocarbons production with high selectivity at near-minimum concentrations of industrial exhaust CO2 but also encourages the investigation of catalytic mechanisms in conditions more representative of real-world production scenarios, such as using dilute feedstocks.
Methods
Reagents
Commercial Cu nanoparticles (99.9%, 25 nm), Nafion solution (5 wt% in water), dimethyl sulfoxide (DMSO, 99.99%) and deuterated water (D2O, 99.9 atom % deuterium) were purchased from Sigma-Aldrich. Cu(CH3COO)2·H2O (purity ≥99.0%), H4N2·H2O (85.0% of volume fraction) and KOH was purchased from Aladdin Ltd. Electrolyte solutions were prepared with deionized water (DIW), which was obtained from a Millipore Auto pure system. All chemicals were of analytical grade and used without further purification. Nitrogen (99.99%), hydrogen (99.99%) and carbon dioxide (99.99%) were purchased from Tianze gas, Inc.
Characterizations
The high-resolution transmission electron microscopy (HRTEM) was observed on a JEOL JEM-2800 at 200 kV. SEM measurements were performed with a JEOL JSM-7800F. Raman spectra were obtained by a confocal Raman spectrometer (Renishaw inVia-Reflex) with a laser wavelength of 633 nm. Electrochemical measurements were performed with the Corrtest instrument (CS350H) electrochemical measurement using a three-electrode system with Ag/AgCl (filled with 3 M KCl) as a reference and Pt plate as a counter electrode. Agilent 7890B gas chromatography (GC) was used to analyzing the gaseous products. 1HNMR spectra were recorded on Bruker DRX 400 Avance MHz spectrometer.
Synthesis of HB Cu, MB Cu and LB Cu
The HB Cu catalyst was synthesized from 99.9% pure Cu nanoparticles with an average size of 25 nm using the vacuum calcination method. In brief, 0.2 g of commercial σ-Cu nanoparticles was placed into a porcelain boat and inserted into a tubular furnace. The sample was evacuated to a pressure below 10 mmHg and heated to 200 °C at a heating rate of 5 °C/min under a vacuum for 30 minutes to produce HB Cu. The MB Cu and LB Cu catalysts were synthesized using a procedure similar to that for HB Cu, with the exception that the vacuum calcination times were extended to 90 min and 150 min, respectively.
Synthesis of the control catalysts
A series of additional controls were obtained by calcination as well as conventional hydrothermal synthesis: (1) Original commercial σ-Cu nanoparticles, without any calcination; (2) Pure Cu nanoparticles after H2 reduction and (3) Cu-based catalysts with different Cu/Cu2O ratios obtained after air calcination and typical synthesis. The pure Cu catalyst was synthesized from 99.9% σ-Cu nanoparticles with an average size of 25 nm using hydrogen calcination method. In brief, 0.2 g of commercial σ-Cu nanoparticles was placed into a porcelain boat and inserted into a tubular furnace. The sample was evacuated to a pressure below 10 mmHg, then pure hydrogen was introduced to atmospheric pressure, repeated 3-4 times, and heated to 200 °C at a heating rate of 5 °C/min under hydrogen atmosphere for 30 min to produce pure Cu. The Cu-based catalysts with different Cu/Cu2O ratios: 1 Cu, 2 Cu and 3 Cu were synthesized using a procedure similar to that for pure Cu, except that the calcination atmosphere was air, and the calcination times were 10 min, 20 min and 30 min, respectively. As for the control catalyst 4 Cu (Cu2O) obtained by the typical synthesis, in brief, 0.2 g of Cu(CH3COO)2·H2O was dissolved in 40 mL of deionized water (DIW), and then 200 μL of H4N2·H2O was added dropwise successively. The mixture was stirred rigorously for 30 min to prepare Cu2O. The achieved yellow precipitate was washed with excess DIW and absolute ethanol for at least three times, which was then dried at 60 °C in an oven under vacuum for next use.
Preparation of the working electrodes
To prepare catalyst inks for the working electrodes, 10 mg of the as-prepared catalyst powder and 30 µL of a 5 wt% Nafion solution were dispersed in 1.0 mL of isopropyl alcohol solution and subjected to ultrasonication for 10 min. The resulting catalyst inks were subsequently applied to carbon paper (AvCarb GDS3250, obtained from Fuel Cell Store) with an area of 0.8 × 1.6 cm2 (catalyst loading ~1.0 mg/cm2) using an airbrush and dried overnight at 60 °C in a vacuum oven.
CO stripping
CO stripping experiment was performed in a three-electrode system by employing a DONGHUA electrochemical workstation (DH7002A). The electrolyte was first bubbled with Ar2 for 30 min, followed by 20 min of pure CO, while the potential of catalyst was held constant at ~0 V (vs. RHE), during which CO monolayer adsorption was developed on the catalyst surface. Extra CO was then expelled from the electrolyte by bubbling the solution with Ar2 for 20 min. Immediately afterwards, CV sweep setup at 20 mV/s was carried out.
CO2RR measurements
Electrochemical measurement was performed in a flow cell electrolyzer by employing a Corrtest instrument (CS350H). The flow cell electrolyzer comprises of two electrolyte chambers separated by an anion exchange membrane and a gas chamber. GDE coated with the prepared catalyst was used as the cathode (working electrode), while the Ag/AgCl (filled with 3 M KCl) electrode and the nickel foam were used as reference and anode (counter electrode), respectively. For this three-electrode experiment, the cathode-anode spacing was small enough that the solution contact resistance was negligible. Therefore, the potential (vs. Ag/AgCl) were converted to RHE values without iR correction:
During electrocatalysis, both gas and electrolyte keep flowing. The flow rates of CO2 (1, 2, 4, 12, 24, 40 sccm) and N2 gases (39, 38, 36, 28, 16, 0 sccm) are regulated by two mass flow controllers according to the gas ratios (2.5%, 5%, 10%, 30%, 60% 100% v/v), and the overall gas flow rate is controlled by another mass flow controller (40 sccm, Alicat), while the electrolyte flow rate is controlled by peristaltic pumps (1.1 mL min−1, Chuangrui, BT100M/YZ1515X). The gaseous phase composition was analyzed online by gas chromatograph (Agilent 7890B). Meanwhile, the cathode electrolyte was collected and analyzed for the liquid product with a Bruker Avance DRX 400 MHz nuclear magnetic resonance spectrometer (1H NMR). Essentially, the diluted cathode electrolyte (500 µL) after electrolysis was mixed with 100 µL of the internal standard solution (25 ppm (v/v) dimethyl sulfoxide, D2O) followed by vortex to be uniform. The 1D 1H spectrum is measured with a water suppression method. The FEs of the products is calculated by using the following formula:
where F is the Faradaic constant, n is the number of electrons transferred, x refers to the mole number of products, and Q is the total charge.
Theoretical methods
In our computational study, we utilized the density functional theory (DFT) as implemented in the Vienna Ab initio Simulation Package (VASP) to carry out all the calculations. The generalized gradient approximation (GGA) with the Revised Perdew-Burke-Ernzerhof (r-PBE) functional was selected to describe the exchange-correlation energy. The ionic cores were represented using the projected augmented wave (PAW) method, and the valence electrons were accounted for within a plane wave basis set, setting the kinetic energy cutoff to 450 eV. To address solvent effects, we employed the implicit solvation model available in VASP, allowing for the explicit calculation of the solvation energy Esol. Additionally, the DFT-D3 method was incorporated to account for van der Waals forces, which are typically underestimated in standard DFT calculations.
The optimization of geometries was conducted with a stringent convergence criterion of 0.05 eV/Å for the forces. For the Brillouin zone integration, we used a 2 x 2 x 1 Monkhorst-Pack grid for all calculations, and the bottom half of the atoms were kept fixed. The changes in free energy (ΔG) for each step of the CO2 reduction reaction (CO2RR) were determined using the computational hydrogen electrode (CHE) model. This model assumes that the chemical potential corresponds to the energy of half a molecule of gas-phase H2 at the standard hydrogen electrode (SHE) potential of 0 V.
To consider the effect of the electrode potential U, () we adjusted the free energy by adding -eU for electron transfer steps, where e is the elementary charge, n is the number of transferred proton-electron pairs, and U is the applied potential. The Gibbs free energy was computed using the following formula:
Here, represents the change in DFT energy, is the zero-point energy, is the entropy at 298.15 K, and T is the absolute temperature (in Kelvin).
For the Cu0/Cu1+ interface boundary model, we constructed a surface consisting of Cu(111) (44) and Cu2O(111) (12) units, forming a (6 × 5) Cu surface structure with specific lattice parameters: a = 16.396, b = 12.396, c = 23.581; α = β = 90°, γ = 120°. The lattice mismatch ratios were 8.55% along the a-axis and 2.58% along the b-axis. The surface coverage of *CO was set to 1/30 ML and 14/30 ML to simulate dilute (5% CO2 v/v) and pure reactant feed conditions, respectively. The reaction free energy for converting CO2 to *CO on the catalyst surface under varying *CO coverages was simulated based on the following reaction scheme:
This approach allowed us to investigate the reaction energetics under different conditions and provide insights into the catalytic performance.
Operando Raman experiments
Operando Raman measurements were conducted in a commercial spectro-electrochemical cell with a three-electrode configuration. To prepare the customized working electrodes, the obtained catalyst inks (10 mg/mL of each of the three catalysts: HB Cu, MB Cu and LB Cu) was uniformly dropped onto the carbon paper (AvCarb GDS 3250, purchased from Fuel Cell Store) with a 1.6 × 1.6 cm2 area (catalyst loading ~1.0 mg/cm2) and dried overnight at 60 °C in a vacuum oven. The obtained carbon paper was then used as working electrodes, with a nickel tape and a saturated Ag/AgCl (filled with 3 M KCl) electrode as the counter electrode and reference electrode, respectively. Operando Raman experiments were conducted on a confocal Raman spectrometer (Renishaw inVia-Reflex) with a laser wavelength of 633 nm. The spectrochemical cell was coupled to a CHI 660E electrochemical analyzer for the electrochemical measurements. Before operando experiments, background spectra were collected without any applied potential. Subsequently, operando spectra were recorded at stepped potentials from −0.11 to −1.51 V vs. RHE in 1 M KOH. All spectra were recorded with a 4 cm−1 resolution and 10 scans.
Supplementary information
Description of Additional Supplementary Files
Source data
Acknowledgements
W. Z. would like to acknowledge the support from National Natural Science Foundation of China (22176086), Natural Science Foundation of Jiangsu Province (BK20210189), the Fundamental Research Funds for the Central Universities (021114380217, 021114380222), the Research Funds from Frontiers Science Center for Critical Earth Material Cycling of Nanjing University, State Key Laboratory of Pollution Control and Resource Reuse, Research Funds for Jiangsu Distinguished Professor, and Carbon Peaking and Carbon Neutrality Technological Innovation Foundation of Jiangsu Province (BE2022861).
Author contributions
L.X. and Y.C. conceived the project. L.X. performed the initial preparation experiments and catalytic performance evaluations. Y.C. and W.Z. supervised research. L.X. wrote the manuscript. Y.J., M.S., J.L., J.Z. and W.Z. supervised the whole project and involved in manuscript preparation and revision.
Peer review
Peer review information
Nature Communications thanks Jing Li, and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data that support the findings of this study are available within the paper and its Supplementary Information file or are available from the corresponding authors upon request. Source Data and Supplementary Data files are provided with this paper. Source data are provided with this paper.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-54590-7.
References
- 1.He, M., Sun, Y. & Han, B. Green Carbon Science: Efficient Carbon Resource Processing, Utilization, and Recycling towards Carbon Neutrality. Angew. Chem. Int. Ed.61, e202112835 (2022). [DOI] [PubMed] [Google Scholar]
- 2.Seh, Z. W. et al. Combining theory and experiment in electrocatalysis: Insights into materials design. Science355, eaad4998 (2017). [DOI] [PubMed] [Google Scholar]
- 3.Zhang, L., Zhao, Z.-J. & Gong, J. Nanostructured Materials for Heterogeneous Electrocatalytic CO2 Reduction and their Related Reaction Mechanisms. Angew. Chem. Int. Ed.56, 11326–11353 (2017). [DOI] [PubMed] [Google Scholar]
- 4.Zimmerman, J. B., Anastas, P. T., Erythropel, H. C. & Leitner, W. Designing for a green chemistry future. Science367, 397–400 (2020). [DOI] [PubMed] [Google Scholar]
- 5.Fu, J. et al. Fight for carbon neutrality with state-of-the-art negative carbon emission technologies. Eco-Environ. Health1, 259–279 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Loiudice, A. et al. Tailoring Copper Nanocrystals towards C2 Products in Electrochemical CO2 Reduction. Angew. Chem. Int. Ed.55, 5789–5792 (2016). [DOI] [PubMed] [Google Scholar]
- 7.Zhang, T. et al. Highly selective and productive reduction of carbon dioxide to multicarbon products via in situ CO management using segmented tandem electrodes. Nat. Catal.5, 202–211 (2022). [Google Scholar]
- 8.Arán-Ais, R. M., Scholten, F., Kunze, S., Rizo, R. & Roldan Cuenya, B. The role of in situ generated morphological motifs and Cu(i) species in C2+ product selectivity during CO2 pulsed electroreduction. Nat. Energy5, 317-325 (2020). [Google Scholar]
- 9.Zhu, P. & Wang, H. High-purity and high-concentration liquid fuels through CO2 electroreduction. Nat. Catal.4, 943–951 (2021). [Google Scholar]
- 10.Dinh, C.-T. et al. CO2 electroreduction to ethylene via hydroxide-mediated copper catalysis at an abrupt interface. Science360, 783–787 (2018). [DOI] [PubMed] [Google Scholar]
- 11.Zhou, Y. et al. Dopant-induced electron localization drives CO2 reduction to C2 hydrocarbons. Nat. Chem.10, 974–980 (2018). [DOI] [PubMed] [Google Scholar]
- 12.Ren, S. et al. Molecular electrocatalysts can mediate fast, selective CO2 reduction in a flow cell. Science365, 367–369 (2019). [DOI] [PubMed] [Google Scholar]
- 13.Kim, D., Kley, C. S., Li, Y. & Yang, P. Copper nanoparticle ensembles for selective electroreduction of CO2 to C2-C3 products. Proc. Natl Acad. Sci.114, 10560-10565 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 14.Yang, D. et al. Selective electroreduction of carbon dioxide to methanol on copper selenide nanocatalysts. Nat. Commun.10, 677 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Fang, M. et al. Hydrophobic, Ultrastable Cuδ+ for Robust CO2 Electroreduction to C2 Products at Ampere-Current Levels. J. Am. Chem. Soc.145, 11323–11332 (2023). [DOI] [PubMed] [Google Scholar]
- 16.Mistry, H. et al. Highly selective plasma-activated copper catalysts for carbon dioxide reduction to ethylene. Nat. Commun.7, 12123 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Ren, D., Ang, B. S.-H. & Yeo, B. S. Tuning the Selectivity of Carbon Dioxide Electroreduction toward Ethanol on Oxide-Derived CuxZn Catalysts. ACS Catal.6, 8239–8247 (2016). [Google Scholar]
- 18.Sullivan, I. et al. Coupling electrochemical CO2 conversion with CO2 capture. Nat. Catal.4, 952–958 (2021). [Google Scholar]
- 19.Kim, D. et al. Electrocatalytic Reduction of Low Concentrations of CO2 Gas in a Membrane Electrode Assembly Electrolyzer. ACS Energy Lett.6, 3488–3495 (2021). [Google Scholar]
- 20.Chu, S. et al. The impact of flue gas impurities and concentrations on the photoelectrochemical CO2 reduction. J. CO2 Utilization60, 101993 (2022). [Google Scholar]
- 21.Van Daele, S. et al. How flue gas impurities affect the electrochemical reduction of CO2 to CO and formate. Appl. Catal. B: Environ.341, 123345 (2024). [Google Scholar]
- 22.Li, X. et al. Low energy-consuming CO2 capture by phase change absorbents of amine/alcohol/H2O. Separation Purification Technol.275, 1383–5866 (2021).
- 23.Al-Attas, T. et al. Permselective MOF-Based Gas Diffusion Electrode for Direct Conversion of CO2 from Quasi Flue Gas. ACS Energy Lett.8, 107–115 (2023). [Google Scholar]
- 24.Hepburn, C. et al. The technological and economic prospects for CO2 utilization and removal. Nature575, 87–97 (2019). [DOI] [PubMed] [Google Scholar]
- 25.Service, R. F. Cost of carbon capture drops, but does anyone want it? Science354, 1362–1363 (2016). [DOI] [PubMed] [Google Scholar]
- 26.Li, M., Irtem, E., Iglesias van Montfort, H.-P., Abdinejad, M. & Burdyny, T. Energy comparison of sequential and integrated CO2 capture and electrochemical conversion. Nat. Commun.13, 5398 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Dongare, S. et al. Reactive capture and electrochemical conversion of CO2 with ionic liquids and deep eutectic solvents. Chem. Society Rev.53, 8563–8631 (2024). [DOI] [PubMed]
- 28.Ho, M. T., Allinson, G. W. & Wiley, D. E. Factors affecting the cost of capture for Australian lignite coal fired power plants. Energy Procedia1, 763–770 (2009). [Google Scholar]
- 29.Boot-Handford, M. E. et al. Carbon capture and storage update. Energy Environ. Sci.7, 130–189 (2014). [Google Scholar]
- 30.D’Alessandro, D. M., Smit, B. & Long, J. R. Carbon Dioxide Capture: Prospects for New Materials. Angew. Chem. Int. Ed.49, 6058–6082 (2010). [DOI] [PubMed] [Google Scholar]
- 31.Bushuyev, O. S. et al. What Should We Make with CO2 and How Can We Make It? Joule2, 825–832 (2018). [Google Scholar]
- 32.Jouny, M., Luc, W. & Jiao, F. General Techno-Economic Analysis of CO2 Electrolysis Systems. Ind. Eng. Chem. Res.57, 2165–2177 (2018). [Google Scholar]
- 33.Kibria, M. G. et al. Electrochemical CO2 Reduction into Chemical Feedstocks: From Mechanistic Electrocatalysis Models to System Design. Adv. Mater.31, 1807166 (2019). [DOI] [PubMed] [Google Scholar]
- 34.Last, G. V. & Schmick, M. T. A review of major non-power-related carbon dioxide stream compositions. Environ. Earth Sci.74, 1189–1198 (2015). [Google Scholar]
- 35.Hou, P., Song, W., Wang, X., Hu, Z. & Kang, P. Well-Defined Single-Atom Cobalt Catalyst for Electrocatalytic Flue Gas CO2 Reduction. Small16, 2001896 (2020). [DOI] [PubMed] [Google Scholar]
- 36.Kwon, Y. et al. Recent advances in integrated capture and electrochemical conversion of CO2. MRS Commun.14, 728–740 (2024).
- 37.Wang, J. et al. Enzymatic Activation and Continuous Electrochemical Production of Methane from Dilute CO2 Sources with a Self-Healing Capsule. J. Am. Chem. Soc.146, 19951–19961 (2024). [DOI] [PubMed] [Google Scholar]
- 38.Peter, S. C. Reduction of CO2 to Chemicals and Fuels: A Solution to Global Warming and Energy Crisis. ACS Energy Lett.3, 1557–1561 (2018). [Google Scholar]
- 39.Sun, Q., Zhao, Y., Ren, W. & Zhao, C. Electroreduction of low concentration CO2 at atomically dispersed Ni-N-C catalysts with nanoconfined ionic liquids. Appl. Catal. B: Environ.304, 120963 (2022). [Google Scholar]
- 40.Wang, X. et al. Efficient Methane Electrosynthesis Enabled by Tuning Local CO2 Availability. J. Am. Chem. Soc.142, 3525–3531 (2020). [DOI] [PubMed] [Google Scholar]
- 41.Luc, W. et al. SO2-Induced Selectivity Change in CO2 Electroreduction. J. Am. Chem. Soc.141, 9902–9909 (2019). [DOI] [PubMed] [Google Scholar]
- 42.Bains, P., Psarras, P. & Wilcox, J. CO2 capture from the industry sector. Prog. Energy Combust. Sci.63, 146–172 (2017). [Google Scholar]
- 43.Capurso, T., Stefanizzi, M., Torresi, M. & Camporeale, S. M. Perspective of the role of hydrogen in the 21st century energy transition. Energy Convers. Manag.251, 114898 (2022). [Google Scholar]
- 44.Graves, C., Ebbesen, S. D., Mogensen, M. & Lackner, K. S. Sustainable hydrocarbon fuels by recycling CO2 and H2O with renewable or nuclear energy. Renew. Sustain. Energy Rev.15, 1–23 (2011). [Google Scholar]
- 45.Song, H., Song, J. T., Kim, B., Tan, Y. C. & Oh, J. Activation of C2H4 reaction pathways in electrochemical CO2 reduction under low CO2 partial pressure. Appl. Catal. B: Environ.272, 119049 (2020). [Google Scholar]
- 46.Kim, B., Ma, S., Molly Jhong, H.-R. & Kenis, P. J. A. Influence of dilute feed and pH on electrochemical reduction of CO2 to CO on Ag in a continuous flow electrolyzer. Electrochim. Acta166, 271–276 (2015). [Google Scholar]
- 47.Sun, M., Cheng, J. & Yamauchi, M. Gas diffusion enhanced electrode with ultrathin superhydrophobic macropore structure for acidic CO2 electroreduction. Nat. Commun.15, 491 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Xu, Y. et al. Oxygen-tolerant electroproduction of C2 products from simulated flue gas. Energy Environ. Sci.13, 554–561 (2020). [Google Scholar]
- 49.Wang, M. et al. Acidic media enables oxygen-tolerant electrosynthesis of multicarbon products from simulated flue gas. Nat. Commun.15, 1218 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Takeda, Y., Mizuno, S., Iwata, R., Morikawa, T. & Kato, N. Gas-fed liquid-covered electrodes used for electrochemical reduction of dilute CO2 in a flue gas. J. CO2 Utilization71, 102472 (2023). [Google Scholar]
- 51.Nabil, S. K. et al. Bifunctional Gas Diffusion Electrode Enables In Situ Separation and Conversion of CO2 to Ethylene from Dilute Stream. Adv. Mater.35, 2300389 (2023). [DOI] [PubMed] [Google Scholar]
- 52.Jiao, L. et al. Single-Atom Electrocatalysts from Multivariate Metal-Organic Frameworks for Highly Selective Reduction of CO2 at Low Pressures. Angew. Chem. Int. Ed.59, 20589–20595 (2020). [DOI] [PubMed] [Google Scholar]
- 53.Kim, B. et al. Over a 15.9% Solar-to-CO Conversion from Dilute CO2 Streams Catalyzed by Gold Nanoclusters Exhibiting a High CO2 Binding Affinity. ACS Energy Lett.5, 749–757 (2020). [Google Scholar]
- 54.Wang, X. et al. Gold-in-copper at low *CO coverage enables efficient electromethanation of CO2. Nat. Commun.12, 3387 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Liu, Z., Yan, T., Shi, H., Pan, H. & Kang, P. Grafting amine-functionalized ligand layer on catalyst for electrochemical CO2 capture and utilization. Appl. Catal. B: Environ.343, 123456 (2024). [Google Scholar]
- 56.Cheng, Y., Hou, J. & Kang, P. Integrated Capture and Electroreduction of Flue Gas CO2 to Formate Using Amine Functionalized SnOx Nanoparticles. ACS Energy Lett.6, 3352–3358 (2021). [Google Scholar]
- 57.Yang, F. et al. Oxide-Derived Bismuth as an Efficient Catalyst for Electrochemical Reduction of Flue Gas. Small19, 2300417 (2023). [DOI] [PubMed] [Google Scholar]
- 58.Liu, Y.-Y., Huang, J.-R., Zhu, H.-L., Liao, P.-Q. & Chen, X.-M. Simultaneous Capture of CO2 Boosting Its Electroreduction in the Micropores of a Metal–organic Framework. Angew. Chem. Int. Ed.62, e202311265 (2023). [DOI] [PubMed] [Google Scholar]
- 59.Shi, H., Pan, H., Cheng, Y., Lu, S. & Kang, P. Imine-Nitrogen-Doped Carbon Nanotubes for the Electrocatalytic Reduction of Flue Gas CO2. ChemElectroChem8, 1792–1797 (2021). [Google Scholar]
- 60.Zhao, H. et al. CO2 selective adsorption over O2 on N−doped activated carbon: Experiment and quantum chemistry study. Appl. Surf. Sci.656, 159727 (2024). [Google Scholar]
- 61.Huang, Y. et al. Impaired conjugation boosts CO2 electroreduction by Ni(ii) macrocyclic catalysts immobilized on carbon nanotubes. J. Mater. Chem. A11, 2969–2978 (2023). [Google Scholar]
- 62.Wang, M. & Luo, J. A coupled electrochemical system for CO2 capture, conversion and product purification. eScience3, 100155 (2023). [Google Scholar]
- 63.Choi, B.-U., Tan, Y. C., Song, H., Lee, K. B. & Oh, J. System Design Considerations for Enhancing Electroproduction of Formate from Simulated Flue Gas. ACS Sustain. Chem. Eng.9, 2348–2357 (2021). [Google Scholar]
- 64.Yao, X. et al. Boosting urea synthesis in simulated flue gas electroreduction by adjusting W–W electronic properties. Green. Chem.26, 8010–8019 (2024). [Google Scholar]
- 65.Yan, T., Liu, S., Liu, Z., Sun, J. & Kang, P. Integrated Electrochemical Carbon Capture and Utilization by Redox-Active Amine Grafted Gold Nanoparticles. Adv. Funct. Mater.34, 2311733 (2024). [Google Scholar]
- 66.Zhang, P., Song, X., Yu, C., Gui, J. & Qiu, J. Biomass-Derived Carbon Nanospheres with Turbostratic Structure as Metal-Free Catalysts for Selective Hydrogenation of o-Chloronitrobenzene. ACS Sustain. Chem. Eng.5, 7481–7485 (2017). [Google Scholar]
- 67.Zhao, Z.-H. et al. Efficient Capture and Electroreduction of Dilute CO2 into Highly Pure and Concentrated Formic Acid Aqueous Solution. J. Am. Chem. Soc.146, 14349–14356 (2024). [DOI] [PubMed] [Google Scholar]
- 68.Chen, B. et al. Molecular Enhancement of Direct Electrolysis of Dilute CO2. ACS Energy Lett.9, 911–918 (2024). [Google Scholar]
- 69.Cobb, S. J., Dharani, A. M., Oliveira, A. R., Pereira, I. A. C. & Reisner, E. Carboxysome-Inspired Electrocatalysis using Enzymes for the Reduction of CO2 at Low Concentrations. Angew. Chem. Int. Ed.62, e202218782 (2023). [DOI] [PubMed] [Google Scholar]
- 70.Nie, B. et al. Tailoring the hierarchical porous nitrogen-doped carbon structures loaded single-atomic Ni catalyst for promoting CO2 electrochemical reduction at low CO2 concentration. Inorg. Chem. Commun.153, 110876 (2023). [Google Scholar]
- 71.Wang, Q. et al. Asymmetric Coordination Induces Electron Localization at Ca Sites for Robust CO2 Electroreduction to CO. Adv. Mater.35, 2300695 (2023). [DOI] [PubMed] [Google Scholar]
- 72.McCoy, D. E., Feo, T., Harvey, T. A. & Prum, R. O. Structural absorption by barbule microstructures of super black bird of paradise feathers. Nat. Commun.9, 1 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Qiao, J., Liu, Y., Hong, F. & Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev.43, 631–675 (2014). [DOI] [PubMed] [Google Scholar]
- 74.Zhang, Y.-J., Sethuraman, V., Michalsky, R. & Peterson, A. A. Competition between CO2 Reduction and H2 Evolution on Transition-Metal Electrocatalysts. ACS Catal.4, 3742–3748 (2014). [Google Scholar]
- 75.Zhang, T., Yuan, B., Wang, W., He, J. & Xiang, X. Tailoring *H Intermediate Coverage on the CuAl2O4/CuO Catalyst for Enhanced Electrocatalytic CO2 Reduction to Ethanol. Angew. Chem. Int. Ed.62, e202302096 (2023). [DOI] [PubMed] [Google Scholar]
- 76.Zhang, Y. et al. Direct reduction of diluted CO gas to C products by copper hydroxyphosphate microrods. AIChE J.69, e18233 (2023). [Google Scholar]
- 77.Pang, Y. et al. Efficient electrocatalytic conversion of carbon monoxide to propanol using fragmented copper. Nat. Catal.2, 251–258 (2019). [Google Scholar]
- 78.Zhang, J. et al. Grain Boundary-Derived Cu+/Cu0 Interfaces in CuO Nanosheets for Low Overpotential Carbon Dioxide Electroreduction to Ethylene. Adv. Sci.9, 2200454 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Wu, Q. et al. Nanograin-Boundary-Abundant Cu2O-Cu Nanocubes with High C2+ Selectivity and Good Stability during Electrochemical CO2 Reduction at a Current Density of 500 mA/cm2. ACS Nano17, 12884–12894 (2023). [DOI] [PubMed] [Google Scholar]
- 80.Sultan, S. et al. Interface rich CuO/Al2CuO4 surface for selective ethylene production from electrochemical CO2 conversion. Energy Environ. Sci.15, 2397–2409 (2022). [Google Scholar]
- 81.Wang, X. et al. Efficient upgrading of CO to C3 fuel using asymmetric C-C coupling active sites. Nat. Commun.10, 5186 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Tang, C., Gong, P., Xiao, T. & Sun, Z. Direct electrosynthesis of 52% concentrated CO on silver’s twin boundary. Nat. Commun.12, 2139 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Gao, D. et al. Selective CO2 Electroreduction to Ethylene and Multicarbon Alcohols via Electrolyte-Driven Nanostructuring. Angew. Chem. Int. Ed.58, 17047–17053 (2019). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Chou, T.-C. et al. Controlling the Oxidation State of the Cu Electrode and Reaction Intermediates for Electrochemical CO2 Reduction to Ethylene. J. Am. Chem. Soc.142, 2857–2867 (2020). [DOI] [PubMed] [Google Scholar]
- 85.Wang, J., Tan, H.-Y., Zhu, Y., Chu, H. & Chen, H. M. Linking the Dynamic Chemical State of Catalysts with the Product Profile of Electrocatalytic CO2 Reduction. Angew. Chem. Int. Ed.60, 17254–17267 (2021). [DOI] [PubMed] [Google Scholar]
- 86.Zhang, X. Y. et al. Direct OC-CHO coupling towards highly C2+ products selective electroreduction over stable Cu0/Cu2+ interface. Nat. Commun.14, 7681 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Chen, Z. et al. Grain-Boundary-Rich Copper for Efficient Solar-Driven Electrochemical CO2 Reduction to Ethylene and Ethanol. J. Am. Chem. Soc.142, 6878–6883 (2020). [DOI] [PubMed] [Google Scholar]
- 88.Kim, C. et al. Cu/Cu2O Interconnected Porous Aerogel Catalyst for Highly Productive Electrosynthesis of Ethanol from CO2. Adv. Funct. Mater.31, 2102142 (2021). [Google Scholar]
- 89.Yang, R. et al. In Situ Halogen-Ion Leaching Regulates Multiple Sites on Tandem Catalysts for Efficient CO2 Electroreduction to C2+ Products. Angew. Chem. Int. Ed.61, e202116706 (2022). [DOI] [PubMed] [Google Scholar]
- 90.Bai, H. et al. Controllable CO adsorption determines ethylene and methane productions from CO2 electroreduction. Sci. Bull.66, 62–68 (2021). [DOI] [PubMed] [Google Scholar]
- 91.Yin, L. et al. Facile and Stable CuInO2 Nanoparticles for Efficient Electrochemical CO2 Reduction. ACS Appl. Mater. Interfaces15, 47135–47144 (2023). [DOI] [PubMed] [Google Scholar]
- 92.Jiang, Y., Li, H., Chen, C., Zheng, Y. & Qiao, S.-Z. Dynamic Cu0/Cu+ Interface Promotes Acidic CO2 Electroreduction. ACS Catal.14, 8310–8316 (2024). [Google Scholar]
- 93.Cao, X. et al. Identification of Cu0/Cu+/Cu0 interface as superior active sites for CO2 electroreduction to C2+ in neutral condition. Chem10, 2089–2102 (2024). [Google Scholar]
- 94.Yuan, X. et al. Controllable Cu0-Cu+ Sites for Electrocatalytic Reduction of Carbon Dioxide. Angew. Chem. Int. Ed.60, 15344–15347 (2021). [DOI] [PubMed] [Google Scholar]
- 95.Li, H. et al. High-Rate CO2 Electroreduction to C2+ Products over a Copper-Copper Iodide Catalyst. Angew. Chem. Int. Ed.60, 14329–14333 (2021). [DOI] [PubMed] [Google Scholar]
- 96.Liu, H. et al. Bottom-up Growth of Convex Sphere with Adjustable Cu(0)/Cu(I) Interfaces for Effective C2 Production from CO2 Electroreduction. Angew. Chem. Int. Ed.63, e202404123 (2024). [DOI] [PubMed] [Google Scholar]
- 97.Zhang, R. et al. Synthesis of n-Propanol from CO2 Electroreduction on Bicontinuous Cu2O/Cu Nanodomains. Angew. Chem. Int. Ed.63, e202405733 (2024). [DOI] [PubMed] [Google Scholar]
- 98.Du, R. et al. Cu▢C(O) Interfaces Deliver Remarkable Selectivity and Stability for CO2 Reduction to C2+ Products at Industrial Current Density of 500 mA cm−2. Small19, 2301289 (2023). [DOI] [PubMed] [Google Scholar]
- 99.Zhang, Z.-Y. et al. SiO2 assisted Cu0–Cu+–NH2 composite interfaces for efficient CO2 electroreduction to C2+ products. J. Mater. Chem. A12, 1218–1232 (2024). [Google Scholar]
- 100.Zhang, Q. et al. Nitrogen cold plasma treatment stabilizes Cu0/Cu+ electrocatalysts to enhance CO2 to C2 conversion. J. Energy Chem.84, 321–328 (2023). [Google Scholar]
- 101.Ma, X. et al. Stabilizing Cu0–Cu+ sites by Pb-doping for highly efficient CO2 electroreduction to C2 products. Green. Chem.25, 7635–7641 (2023). [Google Scholar]
- 102.Dou, T. et al. Sulfurization-derived Cu0–Cu+ sites for electrochemical CO2 reduction to ethanol. J. Power Sources533, 231393 (2022). [Google Scholar]
- 103.Zhang, Z. et al. Highly dispersed Cu-Cu2O-CeOx interfaces on reduced graphene oxide for CO2 electroreduction to C2+ products. J. Colloid Interface Sci.661, 966–976 (2024). [DOI] [PubMed] [Google Scholar]
- 104.Yang, P.-P. et al. Protecting Copper Oxidation State via Intermediate Confinement for Selective CO2 Electroreduction to C2+ Fuels. J. Am. Chem. Soc.142, 6400–6408 (2020). [DOI] [PubMed] [Google Scholar]
- 105.Du, H. et al. Enriching Reaction Intermediates in Multishell Structured Copper Catalysts for Boosted Propanol Electrosynthesis from Carbon Monoxide. ACS Nano17, 8663–8670 (2023). [DOI] [PubMed] [Google Scholar]
- 106.Zhao, H., Cao, H., Zhang, Z. & Wang, Y.-G. Modeling the Potential-Dependent Kinetics of CO2 Electroreduction on Single-Nickel Atom Catalysts with Explicit Solvation. ACS Catal.12, 11380–11390 (2022). [Google Scholar]
- 107.Liu, X. et al. Understanding trends in electrochemical carbon dioxide reduction rates. Nat. Commun.8, 15438 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Bockris, J. O. M. & Nagy, Z. Symmetry factor and transfer coefficient. A source of confusion in electrode kinetics. J. Chem. Educ.50, 839 (1973). [Google Scholar]
- 109.Wuttig, A., Yaguchi, M., Motobayashi, K., Osawa, M. & Surendranath, Y. Inhibited proton transfer enhances Au-catalyzed CO2-to-fuels selectivity. Proc. Natl Acad. Sci.113, E4585–E4593 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Zhan, C. et al. Revealing the CO Coverage-Driven C–C Coupling Mechanism for Electrochemical CO2 Reduction on Cu2O Nanocubes via Operando Raman Spectroscopy. ACS Catal.11, 7694–7701 (2021). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Li, J. et al. Constraining CO coverage on copper promotes high-efficiency ethylene electroproduction. Nat. Catal.2, 1124–1131 (2019). [Google Scholar]
- 112.Cheng, T., Xiao, H. & Goddard, W. A. III Reaction Mechanisms for the Electrochemical Reduction of CO2 to CO and Formate on the Cu(100) Surface at 298 K from Quantum Mechanics Free Energy Calculations with Explicit Water. J. Am. Chem. Soc.138, 13802–13805 (2016). [DOI] [PubMed] [Google Scholar]
- 113.Xiao, H., Goddard, W. A., Cheng, T. & Liu, Y. Cu metal embedded in oxidized matrix catalyst to promote CO2 activation and CO dimerization for electrochemical reduction of CO2. Proc. Natl Acad. Sci.114, 6685–6688 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 114.Xiao, H., Cheng, T., Goddard, W. A. III & Sundararaman, R. Mechanistic Explanation of the pH Dependence and Onset Potentials for Hydrocarbon Products from Electrochemical Reduction of CO on Cu (111). J. Am. Chem. Soc.138, 483–486 (2016). [DOI] [PubMed] [Google Scholar]
- 115.Xiao, H., Cheng, T. & Goddard, W. A. III Atomistic Mechanisms Underlying Selectivities in C1 and C2 Products from Electrochemical Reduction of CO on Cu(111). J. Am. Chem. Soc.139, 130–136 (2017). [DOI] [PubMed] [Google Scholar]
- 116.Montoya, J. H., Shi, C., Chan, K. & Nørskov, J. K. Theoretical Insights into a CO Dimerization Mechanism in CO2 Electroreduction. J. Phys. Chem. Lett.6, 2032–2037 (2015). [DOI] [PubMed] [Google Scholar]
- 117.Wei, P. et al. Coverage-driven selectivity switch from ethylene to acetate in high-rate CO2/CO electrolysis. Nat. Nanotechnol.18, 299–306 (2023). [DOI] [PubMed] [Google Scholar]
- 118.Zhou, Y. et al. Vertical Cu Nanoneedle Arrays Enhance the Local Electric Field Promoting C2 Hydrocarbons in the CO2 Electroreduction. Nano Lett.22, 1963–1970 (2022). [DOI] [PubMed] [Google Scholar]
- 119.Xiong, L. et al. Geometric Modulation of Local CO Flux in Ag@Cu2O Nanoreactors for Steering the CO2RR Pathway toward High-Efficacy Methane Production. Adv. Mater.33, 2101741 (2021). [DOI] [PubMed] [Google Scholar]
- 120.Ji, Y. et al. Selective CO-to-acetate electroreduction via intermediate adsorption tuning on ordered Cu–Pd sites. Nat. Catal.5, 251–258 (2022). [Google Scholar]
- 121.Beenakker, J. J. M., Borman, V. D. & Krylov, S. Y. Molecular transport in subnanometer pores: zero-point energy, reduced dimensionality and quantum sieving. Chem. Phys. Lett.232, 379–382 (1995). [Google Scholar]
- 122.Bartels, L., Meyer, G. & Rieder, K. H. Basic steps involved in the lateral manipulation of single CO molecules and rows of CO molecules. Chem. Phys. Lett.273, 371–375 (1997). [Google Scholar]
- 123.Li, P. et al. Nanoscale Engineering of P-Block Metal-Based Catalysts Toward Industrial-Scale Electrochemical Reduction of CO2. Adv. Energy Mater.13, 2301597 (2023). [Google Scholar]
- 124.Tang, T., Wang, Z. & Guan, J. Achievements and challenges of copper-based single-atom catalysts for the reduction of carbon dioxide to C2+ products. Exploration3, 20230011 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Lv, X., Liu, Z., Yang, C., Ji, Y. & Zheng, G. Tuning Structures and Microenvironments of Cu-Based Catalysts for Sustainable CO2 and CO Electroreduction. Acc. Mater. Res.4, 264–274 (2023). [Google Scholar]
- 126.Leung, D. Y. C., Caramanna, G. & Maroto-Valer, M. M. An overview of current status of carbon dioxide capture and storage technologies. Renew. Sustain. Energy Rev.39, 426–443 (2014). [Google Scholar]
- 127.Sedjo, R. & Sohngen, B. Carbon Sequestration in Forests and Soils. Annu. Rev. Resour. Econ.4, 127–144 (2012). [Google Scholar]
- 128.Li, W.-L., Shuai, Q. & Yu, J. Recent Advances of Carbon Capture in Metal-Organic Frameworks: A Comprehensive Review. Small.20, 2402783 (2024). [DOI] [PubMed]
- 129.Hamilton, S. T., Kelly, M., Smith, W. A. & Park, A.-H. A. Electrolyte–Electrocatalyst Interfacial Effects of Polymeric Materials for Tandem CO2 Capture and Conversion Elucidated Using In Situ Electrochemical AFM. ACS Appl. Mater. Interfaces16, 42021–42033 (2024). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.Monteagudo, J. M., Durán, A., Alonso, M. & Stoica, A.-I. Investigation of effectiveness of KOH-activated olive pomace biochar for efficient direct air capture of CO2. Sep. Purif. Technol.352, 127997 (2025). [Google Scholar]
- 131.Li, Y. et al. Enhancing Local CO2 Adsorption by L-histidine Incorporation for Selective Formate Production Over the Wide Potential Window. Angew. Chem. Int. Ed.62, e202313522 (2023). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Description of Additional Supplementary Files
Data Availability Statement
The data that support the findings of this study are available within the paper and its Supplementary Information file or are available from the corresponding authors upon request. Source Data and Supplementary Data files are provided with this paper. Source data are provided with this paper.