J. Electrochem. Sci. Technol Search

CLOSE


J. Electrochem. Sci. Technol > Volume 16(2); 2025 > Article
Heo, Seong, Bae, Burungale, Mane, Kang, and Ha: Suppressing the Catalytic Barriers in Cu2O Electrocatalyst via Cu Electron Transport Interlayer for Enhanced CO2 Reduction Efficiency

Abstract

The electrochemical carbon dioxide (CO2) reduction is a very promising method for sustainable development, and for this, it is essential to develop a catalyst with high efficiency. In this regard, this study investigated how the presence of the copper (Cu) interlayer affects the CO2 reduction performance of the copper oxide (Cu2O) catalyst. As a result, it was confirmed that when the Cu interlayer was present, the selectivity of the Cu2O catalyst to the formic acid was significantly increased and stability was also improved. In addition, the plausible mechanism involved for enhanced CO2 reduction performance is presented via a wide range of physicochemical and electrochemical characterizations. The findings from our work suggest that the Cu2O is a promising electrocatalyst and the Cu interlayer significantly boosts its CO2 reduction efficiency.

INTRODUCTION

The electrochemical CO2 reduction is a promising strategy to reduce atmospheric CO2 concentrations and enable the transition to sustainable energy sources [13]. In addition, by converting carbon dioxide into useful carbon compounds, dependence on fossil fuels can be reduced [46]. However, stability and efficiency of the electrocatalysts still remain a challenge to commercialize this technology.
Among the various materials used as electrochemical carbon dioxide reduction electrodes, Cu-based catalysts are receiving significant interest. Cu is the only metal catalyst capable of producing C1 compounds such as methane [7,8], as well as C2, C3 compounds such as ethylene [9,10], ethanol [11,12], and propanol [1315]. This remarkable performance stems from the binding energy of Cu with the main reaction intermediate [16,17]. Nevertheless, the full potential of Cu and Cu2O catalysts remains largely untapped, as evidenced by their limited selectivity and stability. The performance of the catalyst is highly dependent on its crystallinity, shape, and surface composition, all of which can profoundly impact its catalytic activity and product selectivity.
An important component in the design of efficient CO2 reduction electrodes is the gas diffusion layer (GDL). The GDL plays an important role in ensuring the effective transport of gas (CO2 and produced products) and liquid (electrolyte) within the electrode structure [1820]. In general, GDL, which is composed of a porous carbon-based material, maintains high conductivity and mechanical stability while supporting the catalyst layer. The porosity of GDL is essential for maintaining a high reaction rate and evenly distributing CO2 to the catalyst site [21,22]. In addition, the hydrophobicity of GDL is important in maintaining the catalytic efficiency without blocking the gas transport path by preventing the electrode from flooding due to the electrolyte [23,24]. However, Gas diffusion electrode(GDE)-based CO2 reduction systems are suffering from problems of catalyst utilization, reduction of reaction rate, and flooding of electrodes or gas accumulation due to inefficient material transport. This is a factor that limits the overall performance and the scalability of the process.
Recent studies have emphasized the importance of optimizing deposition parameters to improve the catalytic performance of Cu2O-based electrodes [2527]. Changes in deposition time can affect the shape and active site of the catalyst, resulting in a change in the CO2 reduction efficiency. In addition, the inclusion of Cu interlayers has shown promising results in improving the overall performance by improving the electron transfer path and stabilizing the active catalytic site [25,28].
This study aims to systematically explore the effect of Cu2O deposition time on the electro chemical reduction of CO2 and to evaluate the effect of Cu interlayers on the performance in GDE-based CO2 reduction system. We integrated electrochemical and surface analysis techniques to optimize the deposition conditions for Cu2O. Additionally, we provide a detailed explanation of the role of the Cu interlayer in enhancing CO2 reduction efficiency.

EXPERIMENTAL

Preparation and characterization of the Electrodeposited Cu Catalysts on the gas-diffusion electrodes

Materials and Chemicals:
All chemicals used in this study were of analytical grade and used without further purification. Copper (II) sulfate pentahydrate (CuSO4·5H2O) and sodium hydroxide (NaOH) were purchased from DUKSAN. Lactic acid, Copper (I) sulfate (Cu2SO4), and sulfuric acid (H2SO4) were obtained from Sigma-Aldrich. Deionized water (18.2 MΩ cm) was used in all experiments.
Preparation of Gas-Diffusion Electrodes (GDEs):
GDEs were prepared using carbon paper as the substrate. The carbon paper was cut into 16 cm2 pieces and cleaned by sequential ultrasonication in acetone, isopropyl alcohol (IPA), and deionized water for 5 minutes each to remove any impurities. The cleaned carbon paper was then dried at 70°C for 1 hour.
Electrodeposition of Cu:
For the preparation of GDEs with a Cu interlayer, Cu was electrodeposited prior to the Cu2O deposition. The electrolyte solution for Cu deposition contained 0.01 M Cu2SO4 and 0.1 M H2SO4. The electrodeposition was performed at a constant current density of –2.5 mA/cm2 for 400 seconds. The electrodes were then rinsed with deionized water and dried under nitrogen gas before proceeding with Cu2O deposition under the same conditions described above.
Electrodeposition of Cu2O:
Cu2O was electrodeposited onto the GDEs from an electrolyte solution containing 0.2 M CuSO4·5H2O and 3 M lactic acid. The pH of the solution was adjusted to 12 using NaOH. The electrodeposition was carried out in a three-electrode system with the GDE as the working electrode, an Ag/AgCl electrode as the reference electrode, and a platinum wire as the counter electrode. A constant voltage of –0.6 V vs. Ag/AgCl was applied for 3 hours. After deposition, the electrodes were rinsed with deionized water and dried at 70°C for 1 hour.

Materials Characterizations

The surface morphology of the electrodeposited catalysts was examined using Field Emission Scanning Electron Microscopy (FE-SEM, SU5000/Hitachi). Energy Dispersive X-ray Spectroscopy (EDS, SU5000/Hitachi) was utilized to analyze the elemental composition and distribution of the deposited layers. X-ray Diffraction (XRD, EMPYREAN/Malvern Panalytical) was employed to determine the crystalline structure of the catalysts. X-ray Photoelectron Spectroscopy (XPS, K-ALPHA+) was conducted to analyze the surface chemical states and composition of the catalysts. FE-SEM, EDS, and XRD analyses were conducted using equipment at Energy Convergence Core Facility in Chonnam National University.

Electrochemical cell and experimental conditions

The electrochemical reduction of CO2 was carried out in a custom-made gas-tight electrochemical cell. The cell was equipped with separate compartments for the working electrode (WE), reference electrode (RE), and counter electrode (CE), each compartment separated by Nafion 117 membranes to prevent cross-contamination. The operation electrode partition was stably supplied to CO2 gas continuously flowing through CO2 gas continuously.
The prepared GDE with electrodeposited Cu2O and Cu/Cu2O catalysts served as the working electrodes. An Ag/AgCl (saturated KCl) electrode was used as the reference electrode, and a bare GDE was used as the counter electrode. The electrolyte used for CO2 reduction was 1 N KOH solution, prepared using deionized water and purged with CO2 gas for 30 minutes prior to the experiments to ensure saturation and a stable pH of approximately 14.
CO2 gas (99.9999% purity) was continuously bubbled through the electrolyte at a flow rate of 20 sccm during the electrochemical measurements. Electrochemical experiments were conducted using a CH Instruments potentiostat (WPG100e/WonATech) in a three-electrode configuration.
Linear sweep voltammetry (LSV) was performed from –1.5 V to 1.0 V vs. RHE at a scan rate of 20 mV/s to evaluate the onset potential and current density for CO2 reduction. And controlled-potential electrolysis was carried out at selected potentials (based on LSV results) for 1 hour. During electrolysis, the gas and liquid products were periodically sampled for analysis.
Gas products were analyzed using gas chromatography (GC, CN/7890B/Agilent Technologies) equipped with a thermal conductivity detector (TCD) and a flame ionization detector (FID). Liquid products were analyzed using high-performance liquid chromatography (HPLC, 1260 infinity II/Agilent Technologies) and gas chromatography with a headspace sampler (Headspace-GC, 8890B/Agilent Technologies) to quantify the concentration of various reduction products. Faradaic efficiency for each product was calculated based on the charge passed and the amount of product formed.
EIS measurements were performed at open circuit potential and selected reduction potentials to study the charge transfer resistance and electrode kinetics. The frequency range for EIS was from 100 kHz to 0.1 Hz with an AC amplitude of 10 mV.
CV measurements were conducted in a non-faradaic region to estimate the electrochemical surface area (ECSA) of the catalysts. The scan rate was varied from 20 mV/s to 200 mV/s to establish the double-layer capacitance.
The measured potential (EAg/AgCl) was converted to the reversible hydrogen electrode (RHE, ERHE) scale using Eq. 1.
(1)
ERHE=EAg/AgCl+0.059 V×pH+0.1976 V

RESULT AND DISCUSSION

Catalyst Characterization

Fig. 1a is a schematic diagram showing the microstructure of the GDE@Cu@Cu2O electrode. GDE has a porous support structure, allowing CO2 to be effectively transported to the catalyst surface. The Cu and Cu2O layers deposited on GDE act as active sites for electrochemical CO2 reduction reactions and are selectively converted to products such as carbon monoxide (CO), formic acid (HCOOH), and ethanol (C2H5OH). Fig 1b shows an X-ray diffraction (XRD) pattern. It provides information about the crystal structure of GDE@Cu2O and GDE@Cu@Cu2O electrodes. The XRD pattern of the GDE@Cu2O electrode showed distinct peaks at 2θ values of about 29.6°, 36.4°, 42.3°, 61.4°, 73.6°, and 77.5° corresponding to the crystal phase of Cu2O, indicating that (110), (111), (200), (220), (311), and (222) well-crystallized Cu2O phase (JCPDS No. 05-0667) was formed in the plane [29,30]. In the case of the XRD pattern of the GDE@Cu@Cu2O electrode, a pattern similar to that of the GDE@Cu2O electrode was shown, but in addition, there were additional peaks suggesting the presence of the Cu interlayer. The Cu2O peaks were present at the same 2θ values (29.6°, 36.4°, 42.3°, 61.4°, 73.6° and 77.5°), confirming the formation of the cuprous oxide phase on top of the Cu interlayer. The XRD pattern also exhibited peak at approximately 43.3o, corresponding to the (111) plane of metallic Cu, confirming the presence of the Cu interlayer (JCPDS No. 04-0836) [31,32].
The surface morphology of the GDE@Cu2O electrode was investigated using FE-SEM at different magnifications, as shown in Fig. 1c and 1d. Fig. 1c presents the surface at ×2.00k magnification, while Fig. 1d shows a more detailed view at ×10.0k magnification. In comparison with the surface morphology of GDE@Cu@Cu2O displayed in Fig. 1f and 1g, distinct differences can be observed. The GDE@Cu2O surface exhibits tetrahedral Cu2O crystals formed directly on the GDE substrate. In contrast, the GDE@Cu@Cu2O surface features a leaf-like structure with smaller particles forming on these leaf-like surfaces. The morphological difference between the two electrodes is due to the presence of the Cu interlayer. The Cu interlayer affects nucleation and growth during the electrodeposition process of Cu2O, preventing the formation of tetrahedral crystals and allowing them to exist in the form of spherical particles. Compared to the surface morphology of GDE@Cu, as shown in Supporting Information Fig. S1a and S1b, the structure of GDE@Cu@Cu2O is extremely similar, suggesting that Cu2O formed on the Cu interlayer.
The elemental composition and distribution of the GDE@Cu@Cu2O electrode surface were analyzed using EDS mapping as shown in Fig. 1e. As can be seen from the mapping results, it shows a uniform distribution of Cu and O throughout the surface, confirming the presence of Cu2O. This supports the structural information obtained from XRD analysis. The presence of carbon (C) is due to the electroplated Cu and the GDE substrate under Cu2O.
XPS analysis was conducted to investigate the surface chemical states and composition of the GDE@Cu2O and GDE@Cu@Cu2O electrodes. The survey spectra, as well as core level spectra C 1s, O 1s, and Cu 2p spectra, were obtained for both samples this is shown in Fig 2. Fig. 2a is the XPS survey spectra of both GDE@Cu2O and GDE@Cu@Cu2O electrodes displayed characteristic peaks corresponding to carbon (C 1s), oxygen (O 1s), and copper (Cu 2p). The survey spectra for GDE@Cu@Cu2O also showed a higher intensity for the Cu 2p peaks compared to GDE@Cu2O, indicating the presence of the additional Cu interlayer [33]. Fig. 2b is the C 1s spectra for both electrodes showed a main peak at ~284.1 eV (C=C) with additional peaks at ~285.0 eV (C–OH) and ~278.9 eV (C=O), resulting from the presence of GDE [34]. Fig 2c is the O 1s spectrum displayed a dominant peak at ~530 eV, corresponding to metal oxides (Cu2O). Fig. 2c shows a peak at 530 eV due to Cu2O as an O1s spectrum. A peak at 531 eV, which is a higher binding energy, indicates the presence of adsorbed water or hydroxyl groups [35]. Fig. 2d is a Cu 2p spectrum characterized by two major peaks at 931.8 eV (Cu 2p3/2) and 951.6 eV (Cu 2p1/2) originating from Cu(I) species of Cu2O [3537]. The Cu 2p3/2 peak at 931.8 eV was stronger in the GDE@Cu@Cu2O sample compared to the GDE@Cu2O sample. This is due to the presence of the Cu interlayer present in the GDE@Cu@Cu2O sample. The Cu interlayer contributes to the additional metallic Cu, which is reflected by a stronger Cu 2p3/2 peak at this binding energy [37,38]. A higher intensity at 931.8 eV means a higher concentration of metal Cu species in the GDE@Cu@Cu2O sample. This indicates the presence of the Cu interlayer increasing the total metal Cu content [38,39]. In addition, the absence of satellite peaks suggests the presence of minimal Cu(II) species in the sample. However, for GDE@Cu@Cu2O, the Cu 2p3/2 and Cu 2p1/2 peaks appear at similar binding energies, whereas the overall intensity is higher and there is a slight presence of satellite peaks, indicating a small amount of Cu(II) species due to slight oxidation between the Cu layers or formation of Cu(OH)2 on the surface [40].

Electrocatalytic Performance

Fig. 3 shows a comparative analysis of the electrochemical performance of the GDE@Cu2O and GDE@Cu@Cu2O electrodes in a CO2-saturated 1 M KOH aqueous solution, showing the effect of the Cu interlayer on the catalytic activity and the overall efficiency of the CO2 reduction reaction. Fig. 3a shows linear sweep voltammetry (LSV) of the GDE@Cu2O and GDE@Cu2O electrodes. The GDE@Cu@Cu2O electrode has a higher current density in a potential range compared to the GDE@Cu2O electrode. This may have been due to the structural deformation induced by the presence of the Cu interlayer, which may more efficiently promote the interaction between the CO2 molecule and the catalyst surface [4143]. Fig. 3b shows the current density over time at a constant potential of –0.8 V (vs. RHE). For about 1 hour, the GDE@Cu@Cu2O electrode maintained a very constant high current density, while the GDE@Cu2O electrode showed a lower current density than the GDE@Cu@Cu2O electrode, and after about 2500 seconds, the current density decreased sharply to less than half. Fig. 3c and 3d show the Faradaic efficiency at various voltages of the GDE@Cu2O and GDE@Cu@Cu2O electrodes, respectively. For GDE@Cu@Cu2O electrodes, HCOOH is the main product, showing a significantly higher Faradaic efficiency compared to GDE@Cu2O. This contrasts with the GDE@Cu2O electrode, where H2 is the main product, which may mean that in the absence of the Cu interlayer, the electrode's ability to effectively reduce CO2 is limited, resulting in higher hydrogen evolution. Therefore, the Cu interlayer is likely to stabilize intermediates such as CO* and HCOO* to promote subsequent reduction to CO and HCOOH, while inhibiting competitive hydrogen evolution reaction (HER) [4447]. Stabilization of CO and HCOO intermediate at the Cu/Cu2O interface is very important for improving the selectivity and efficiency of the CO2 reduction reaction [48,49]. Similar Cu-based electrical catalysts have been extensively studied, and customized catalysts and surface composition have been found to optimize intermediate binding energy and reaction paths [10,50]. Fig. 3e is a Nyquist plot obtained by electrochemical impedance spectroscopy (EIS) at –0.8 V (vs. RHE), through which information on the charge transfer characteristics of the two electrodes can be obtained. The calculated Rs and Rct values are specified in Table S1 of the support information. Considering that the semicircle of GDE@Cu@Cu2O in the Nyquist plot is significantly smaller than that of GDE@Cu2O, the electron transfer resistance (Rct) of GDE@Cu@Cu2O is smaller than that of GDE@Cu2O. The decrease in Rct means that the Cu interlayer improves the electron transfer kinetics between the electrode and the electrolyte, which allows faster electron transfer to the CO2 reduction reaction. This improvement in charge transfer allows electrons to reach the active site more easily, so CO2 reduction can be achieved more efficiently [41,42,51]. Fig. 3f shows the Tafel plots of GDE@Cu2O and GDE@Cu@Cu2O. The Tafel slopes of GDE@Cu2O and GDE@Cu@Cu2O were measured to be –0.23 V/dec and –0.14 V/dec, respectively, indicating a distinct difference in catalytic performance between the two materials in the electrochemical reduction of CO2 to formic acid. The low Tafel slopes of GDE@Cu@Cu2O (–0.14 V/dec) compared to GDE@Cu2O (–0.23 V/dec) suggest that GDE@Cu@Cu2O will have much less overpotential in promoting a more efficient reaction pathway to achieve the same current density increase [52,53]. On the other hand, the high Tafel gradient of Cu2O (–0.23 V/dec) indicates that the electron transfer process is less efficient and requires higher overpotential for the same level of catalytic activity [54,55]. This may be due to the intrinsic limitation of Cu2O or to the absence of a metal Cu interlayer that could facilitate faster electron transport.
To calculate the electrochemically active surface area (ECSA), cyclic voltammetry (CV) measurements were performed on GDE@Cu2O and GDE@Cu@Cu2O electrodes at various scan rates increasing by 20 mV/s from 20 mV/s to 200 mV/s, and these are shown in Fig. 4a and 4b. The potential was converted to RHE contrast value using the Nernst equation (Eq. 1) [56,57].
As can be seen in Fig. 4a, the GDE@Cu2O electrode exhibits typical capacitive behavior within the selected potential window (0.825 V to 1.25 V vs. RHE), with the current density increasing linearly with the scan rate. This behavior is indicative of the charging and discharging of the double layer, confirming that the measurements are conducted within the non-faradaic region [58,59]. Similarly, the CV curves for GDE@Cu@Cu2O, displayed in Fig. 4b, also show linear increases in current density with scan rate, suggesting that the presence of the Cu interlayer does not significantly alter the capacitive behavior within the selected potential window. To further analyze the data, the current densities at the midpoint of the potential window were plotted against the scan rate for both electrodes. The difference between the anodic and cathodic current densities (ΔJ) was halved and plotted against the scan rate, as shown in Fig. 4c. This linear relationship between ΔJ/2 and the scan rate indicates that the measured current is primarily due to the charging of the double layer, and thus, the slope of this line can be used to calculate the double-layer capacitance (Cdl) for each electrode. Fig. 4d depicts the calculated ECSA derived from the CV data using Eq. 2 and 3 provided below [60].
(2)
Cs=V1V21(V)dVmS
(3)
ECSA=CdlCs
The two main parameters, CS and Cdl, are used to calculate ECSA. CS represents the non-capacitance of an ideal flat surface of an electrocatalyst, and for Cu2O-based electrodes, the value of 0.035 mF/cm2, which is commonly used in calculations, was used as reported in similar studies [61]. The double layer capacitance, Cdl, was measured experimentally and represents the charge storage capacity at the interface between the electrode surface and the electrolyte in the non-faradaic region. The higher the Cdl value, the larger the active surface area [62,63]. CS refers to the standard electrical capacity per unit area and is the reference for the expected double layer electrical capacity of the known surface area under similar conditions. The Cdl to CS ratio can be used to estimate the active surface area of the electrode, which provides information on the electrochemical performance of the produced electrode. The calculated results show that the GDE@Cu@Cu2O electrode exhibits a larger ECSA than the GDE@Cu2O, which is thought to be due to the increased surface roughness and improved active surface area due to the presence of the Cu interlayer as shown in the SEM image of Fig. 1. These results indicate that increasing the ECSA provides more active sites, thereby significantly enhancing the electrochemical performance of the electrode [64,65]. In summary, in GDE@Cu@Cu2O catalysts, the Cu interlayer plays a key role in improving the CO2 reduction performance. First, the Cu interlayer promotes electron transfer and increases reaction efficiency by reducing charge transfer resistance (Fig. 3e). Second, by changing the shape of the Cu2O layer to form small spherical particles and increase the surface roughness, the ECSA is expanded (Fig. 1). Third, the Cu/Cu2O interface stabilizes the CO and HCOO intermediates, increasing the selectivity of forming formic acid (HCOOH) and suppressing the hydrogen generation reaction (HER). Unlike alloying or doping, the Cu interlayer simultaneously enhances selectivity and efficiency through electron transfer and structural optimization while maintaining Cu-specific catalytic properties.

CONCLUSIONS

This study is to improve the electrochemical CO2 reduction performance by introducing a Cu interlayer to a Cu2O-based electrode manufactured using an electrodeposition method on a GDE substrate. It was found that the GDE@Cu@Cu2O electrode introduced with the Cu interlayer showed a higher current density than the GDE@Cu2O electrode without the Cu interlayer. Furthermore, the Faradaic efficiency for CO2 reduction to formic acid was enhanced, while the charge transfer resistance was simultaneously decreased. In addition, unlike the GDE@Cu2O electrode which forms a tetrahedral crystal structure, the GDE@Cu@Cu2O electrode was formed in the form of small spherical Cu2O crystals on the leaf-shaped Cu crystals, indicating that ECSA was also significantly increased. This indicates that the Cu interlayer not only enhances electron transfer efficiency but also introduces additional active sites, thereby boosting the electrochemical activity of the Cu2O-based catalyst and enabling more efficient CO2 reduction.

Notes

ACKNOWLEDGMENTS

This research was supported by "Regional Innovation Strategy (RIS)" through the National Research Foundation of Korea (NRF) funded by the Ministry of Education (MOE) (2021RIS-002) and Basic Science Research Capacity Enhancement Project through Korea Basic Science Institute (National research Facilities and Equipment Center) grant funded by the Ministry of Education. (grant No. 2019R1A6C1010024)

Fig. 1.
(a) Schematic illustration of microstructure of GDE@Cu@Cu2O, (b) XRD patterns of GDE@Cu@Cu2O and GDE@Cu2O electrodes, (c, d) FE-SEM images of the GDE@Cu2O surface, and (e) EDS mapping results of the GDE@Cu2O surface, (f, g) FE-SEM images of the GDE@Cu@Cu2O surface, and (h) EDS mapping results of the GDE@Cu@Cu2O surface.
jecst-2024-01137f1.jpg
Fig. 2.
(a) XPS survey spectra, and the core level spectra of (b) C 1s, (c) O 1s, (d) Cu 2p GDE@Cu2O and GDE@Cu@Cu2O
jecst-2024-01137f2.jpg
Fig. 3.
Electrochemical performance of GDE@Cu2O and GDE@Cu@Cu2O electrodes. (a) Linear sweep voltammetry (LSV) curves of GDE@Cu2O and GDE@Cu@Cu2O in CO2-saturated 1N KOH solution, (b) current density over time at–0.8 V vs. RHE for both electrodes, (c, d) Faradaic efficiency of the CO2 reduction reaction at various potential for each product on GDE@Cu2O and GDE@Cu@Cu2O electrodes, respectively, (e) Nyquist plot from Electrochemical Impedance Spectroscopy (EIS) measured at –0.8 V vs. RHE, and (f) Tafel plots comparing the reaction kinetics of the two electrodes.
jecst-2024-01137f3.jpg
Fig. 4.
CVs at various scan rates from 20 mV/s to 200 mV/s of (a) GDE@Cu2O and (b) GDE@Cu@Cu2O. (c) Half of the difference in current density (J) at the center of the CV potential window vs. scan rates and (d) calculated ECSA values.
jecst-2024-01137f4.jpg

REFERENCES

[1] W. Lai, Y. Qiao, Y. Wang and H. Huang, Adv. Mater., 2023, 35(51), 2306288.

[2] C. Li, Y. Ji, Y. Wang, C. Liu, Z. Chen, J. Tang, Y. Hong, X. Li, T. Zheng, Q. Jiang and C. Xia, Nano-Micro Lett., 2023, 15, 113.

[3] J. Zhang, J. Ding, Y. Liu, C. Su, H. Yang, Y. Huang and B. Liu, Joule, 2023, 7(8), 1700–1744.
crossref
[4] R. G. Anderson, J. G. Canadell, J. T. Randerson, R. B. Jackson, B. A. Hungate, D. D. Baldocchi, G. A. Ban-Weiss, G. B. Bonan, K. Caldeira, L. Cao, N. S. Diffenbaugh, K. R. Gurney, L. M. Kueppers, B. E. Law, S. Luyssaert and T. L. O’Halloran, Front. Ecol. Environ., 2011, 9(3), 174–182.
crossref pdf
[5] R. Snoeckx and A. Bogaerts, Chem. Soc. Rev., 2017, 46, 5805–5863.
crossref
[6] W. A. Smith, T. Burdyny, D. A. Vermaas and H. Geerlings, Joule, 2019, 3(8), 1822–1834.
crossref
[7] P. Zhao, H. Jiang, H. Shen, S. Yang, R. Gao, Y. Guo, Q. Zhang and H. Zhang, Angew. Chem. Int. Ed., 2023, 62(49), e202314121.

[8] K. Kim, P. Wagner, K. Wagner and A. J. Mozer, ChemCatChem, 2023, 15(13), e202300396.

[9] Y. Ouyang, L. Shi, X. Bai, C. Ling, Q. Li and J. Wang, ACS Catal., 2023, 13(23), 15448–15456.
crossref pdf
[10] J. Qu, X. Cao, L. Gao, J. Li, L. Li, Y. Xie, Y. Zhao, J. Zhang, M. Wu and H. Liu, Nano-Micro Lett., 2023, 15, 178.

[11] M. Li, N. Song, W. Luo, J. Chen, W. Jiang and J. Yang, Adv. Sci., 2023, 10(2), 2204579.

[12] W. Mo, X.-Q. Tan and W.-J. Ong, ACS Sustain. Chem. Eng., 2023, 11(14), 5782–5799.
crossref pdf
[13] C. Long, X. Liu, K. Wan, Y. Jiang, P. An, C. Yang, G. Wu, W. Wang, J. Guo, L. Li, K. Pang, Q. Li, C. Cui, S. Liu, T. Tan and Z. Tang, Sci. Adv., 2023, 9, eadi6119.

[14] S. Jeong, C. Huang, Z. Levell, R. X. Skalla, W. Hong, N. J. Escorcia, Y. Losovyj, B. Zhu, A. N. Butrum-Griffith, Y. Liu, C. W. Li, Y. Liu and X. Ye, J. Am. Chem. Soc., 2024, 146(7), 4508–4520.
crossref pdf
[15] R. Zhang, J. Zhang, S. Wang, Z. Tan, Y. Yang, Y. Song, M. Li, Y. Zhao, H. Wang, B. Han and R. Duan, Angew. Chem. Int. Ed., 2024, 63(30), e202405733.

[16] S. Mohan, B. Honnappa, A. Augustin, M. Shanmugam, C. Chuaicham, K. Sasaki, B. Ramasamy and K. Sekar, Catalysts, 2022, 12(4), 445.
crossref
[17] Y. Jiang, X. Wang, D. Duan, C. He, J. Ma, W. Zhang, H. Liu, R. Long, Z. Li, T. Kong, X. J. Loh, L. Song, E. Ye and Y. Xiong, Adv. Sci., 2022, 9(16), 2105292.

[18] B. Endrődi, G. Bencsik, F. Darvas, R. Jones, K. Rajeshwar and C. Janáky, Prog. Energy Combust. Sci., 2017, 62, 133–154.

[19] T. Burdyny and W. A. Smith, Energy Environ. Sci., 2019, 12, 1442–1453.
crossref
[20] K. Yang, R. Kas, W. A. Smith and T. Burdyny, ACS Energy Lett., 2021, 6(1), 33–40.
crossref pdf
[21] T. N. Nguyen and C.-T. Dinh, Chem. Soc. Rev., 2020, 49, 7488–7504.
crossref
[22] K. Liu, W. A. Smith and T. Burdyny, ACS Energy Lett., 2019, 4, 639–643.
crossref pdf
[23] Y. Wu, L. Charlesworth, I. Maglaya, M. N. Idros, M. Li, T. Burdyny, G. Wang and T. E. Rufford, ACS Energy Lett., 2022, 7(9), 2884–2892.
crossref pdf
[24] A. Q. Fenwick, A. J. Welch, X. Li, I. Sullivan, J. S. DuChene, C. Xiang and H. A. Atwater, ACS Energy Lett., 2022, 7(2), 871–879.
crossref
[25] A. R. Araujo, K. Rajeshwar and M. V. B. Zanoni, Chem. Eng. J., 2015, 264, 302.

[26] J. Albo, A. Sáez, J. Solla-Gullón, V. Montiel and A. Irabien, Appl. Catal. B Environ., 2015, 176–177, 709–717.

[27] G. O. Larrazábal, T. Shinagawa, A. J. Martín and J. Pérez-Ramírez, Nat. Commun., 2018, 9, 1477.

[28] S. Budi, M. G. Sutrisno and T. Hadinugrahaningsih, Clean. Mater., 2024, 12, 100250.
crossref
[29] R. M. H. H. Jayarathne, P. K. D. D. P. Pitigala and V. P. S. Perera, Institute of Physics Sri Lanka, 2019, 35, 31–38.

[30] S. M. Pawar, J. Kim, A. I. Inamdar, H. Woo, Y. Jo, B. S. Pawar, S. Cho, H. Kim and H. Im, Sci. Rep., 2016, 6, 21310.

[31] A. Ma, X. Wu, X. Li, D. Liu, H. Li, A. T. Kuvarega, B. B. Mamba and J. Gui, Appl. Surf. Sci., 2024, 653, 159397.
crossref
[32] D. Mardiansyah, T. Badloe, K. Triyana, M. Q. Mehmood, N. Raeis-Hosseini, Y. Lee, H. Sabarman, K. Kim and J. Rho, Sci. Rep., 2018, 8, 10639.

[33] A. Cano, J. Rodríguez-Hernández, A. Shchukarev and E. Reguera, J. Solid State Chem., 2019, 273, 1–10.
crossref
[34] P. Xia, L. Zhao, X. Chen, Z. Ye, Z. Zheng, Q. He and I. Sirés, Appl. Catal. B Environ., 2024, 343, 123467.
crossref
[35] Y. Wang, Y. Lü, W. Zhan, Z. Xie, Q. Kuang and L. Zheng, J. Mater. Chem. A, 2015, 3, 12796–12803.
crossref
[36] F. Du, Q.-Y. Chen and Y.-H. Wang, J. Phys. Chem. Solids, 2017, 104, 139–144.
crossref
[37] P. Jiang, D. Prendergast, F. Borondics, S. Porsgaard, L. Giovanetti, E. Pach, J. Newberg, H. Bluhm, F. Besenbacher and M. Salmeron, J. Chem. Phys., 2013, 138, 024704.

[38] R. P. Lopes, Environ. Sci. Pollut. Res., 2019, 26, 14883–14903.
crossref pdf
[39] M. Swadźba-Kwaśny, L. Chancelier, S. Ng, H. G. Manyar, C. Hardacre and P. Nockemann, Dalton Trans., 2012, 41, 219–227.
crossref
[40] C. Li, S. Liu, X. Jin, Z. Zuo, H. Yang, H. Jing and X. Cao, J. Alloys Compd., 2022, 900, 163455.
crossref
[41] L. Li, J. Yang, L. Li, Y. Huang and J. Zhao, Electrochim. Acta, 2022, 402, 139523.
crossref
[42] W. Ni, Z. Yixiang, Y. Yao, X. Wang, R. Zhao, Z. Yang, X. Li and Y.-M. Yan, ACS Appl. Mater. Interfaces, 2022, 14(11), 13261–13270.
crossref
[43] G. M. Tomboc, S. Choi, T. Kwon, Y. J. Hwang and K. Lee, Adv. Mater., 2020, 32(17), 1908398.

[44] P. Prasannalakshmi, N. Shanmugam and A. S. Kumar, J. Appl. Electrochem., 2017, 47, 889–903.
crossref pdf
[45] Q. Zhang, V. Celorrio, K. Bradley, F. Eisner, D. Cherns, W. Yan and D. J. Fermín, J. Phys. Chem. C, 2014, 118(31), 18207–18213.
crossref
[46] P. Mane, I. V. Bagal, H. Bae, V. V. Burungale, C. Seong, S. W. Ryu and J.-S. Ha, Adv. Sustain. Syst., 2022, 6(6), 2200014.

[47] S. Wang, P. Chen, J. H. Yun, Y. Hu and L. Wang, Angew. Chem. Int. Ed., 2017, 56(29), 8500–8504.
crossref pdf
[48] S. Feng, X. Wang, D. Cheng, Y. Luo, M. Shen, J. Wang, W. Zhao, S. Fang, H. Zheng, L. Ji, X. Zhang, W. Xu, Y. Liang, P. Sautet and J. Zhu, Angew. Chem. Int. Ed., 2024, 63(8), e202317942.

[49] L. Li, A. Ozden, S. Guo, C. Wang, M. Zhang, J. Zhang, H. Jiang, W. Wang, H. Dong, D. Sinton and M. Zhong, Nat. Commun., 2021, 12, 5223.

[50] S. Natesakhawat, J. W. Lekse, J. P. Baltrus, P. R. Ohodnicki, B. H. Howard, X. Deng and C. Matranga, ACS Catal., 2012, 2, 1667–1676.
crossref
[51] M. Zhang, X. Xuan, W. Wang, C. Ma and Z. Lin, Adv. Funct. Mater., 2020, 30(52), 2005983.

[52] M. Ye, T. Shao, J. Liu, C. Li, B. Song and S. Liu, Appl. Surf. Sci., 2023, 622, 156981.
crossref
[53] T. Ahmad, S. Liu, M. Sajid, K. Li, M. Ali, L. Liu and W. Chen, Nano Res. Energy, 2022, 1, 9120021.

[54] T. N. Saada, L. Pang, A. H. B. Dourado, L. D. Germano, E. D. Vicentini, A. P. L. Batista, F. Dumeignil, S. Paul, R. Wojcieszak, S. Melinte, G. Sandu, G. Petretto, G.-M. Rignanese, A. H. Braga, T. F. Rosado, D. Meziane, R. Boukherroub and S. Szunerits, Electrochim. Acta, 2021, 390, 138810.
crossref
[55] A. M. Paul, A. Sajeev, R. Nivetha, K. Gothandapani, P. Bhardwaj, G. K, V. Raghavan, G. Jacob, R. Sellapan, S. K. Jeong and A. N. Grace, Diam. Relat. Mater., 2020, 107, 107899.
crossref
[56] V. V. Burungale, H. Bae, A.-N. Cha, J. Heo, S.-W. Ryu, S.-H. Kang and J.-S. Ha, Int. J. Hydrogen Energy, 2023, 48(11), 4193–4206.
crossref
[57] A. M. Teli, S. A. Beknalkar, V. V. Burungale, R. U. Amte, P. J. Morankar, V. V. Satale, M. A. Yewale, R. P. Dhavale and J. C. Shin, J. Energy Storage, 2024, 86, 111155.
crossref
[58] M. Goswami, C. Nithya, N. Sathish, S. Kumar, N. Singh, A. K. Srivastava and S. Kumar, New J. Chem., 2020, 44, 5278–5284.
crossref
[59] N. S. Tantawy, F. E. Heakal and S. Y. Ahmed, J. Energy Storage, 2020, 31, 101625.
crossref
[60] P. Connor, J. Schuch, B. Kaiser and W. Jaegermann, Z. Phys. Chem., 2020, 234(5), 979–994.
crossref
[61] Z. Shen, J. Yan, M. Wang, L. Xing, B. Huang, H. Zhou, W. Li, L. Chen and J. Shi, ACS Sustain. Chem. Eng., 2023, 11(25), 9433–9441.
crossref pdf
[62] P. Liu, J.-G. Zhang, C. E. Tracy and J. A. Turner, Electrochem. Solid-State Lett., 2000, 3, 163.

[63] D. S. Dmitriev and M. I. Tenevich, Nanosystems: Phys. Chem. Math., 2023, 14(5), 590–600.
crossref
[64] M. H. Elbakkay, S. I. El-Dek and A. A. Farghali, Int. J. Hydrogen Energy, 2022, 47(17), 9876–9894.
crossref
[65] M. Asif, W. Haitao, D. Shuang, A. Aziz, G. Zhang, F. Xiao and H. Liu, Sens. Actuators B: Chem., 2017, 239, 243–252.
crossref
TOOLS
Share :
Facebook Twitter Linked In Google+ Line it
METRICS Graph View
  • 0 Crossref
  •   Scopus
  • 463 View
  • 33 Download
Related articles in J. Electrochem. Sci. Technol


ABOUT
ARTICLE CATEGORY

Browse all articles >

BROWSE ARTICLES
AUTHOR INFORMATION
Editorial Office
E-mail: journal@kecs.or.kr    Tel: +82-2-568-9392    Fax: +82-2-568-5931                   

Copyright © 2025 by The Korean Electrochemical Society.

Developed in M2PI

Close layer
prev next