Next Article in Journal
Thermal Conductivity Analysis of Chitin and Deacetylated-Chitin Nanofiber Films under Dry Conditions
Next Article in Special Issue
Structural and Electrochemical Properties of Physically and Chemically Activated Carbon Nanoparticles for Supercapacitors
Previous Article in Journal
Electrochemical Investigation of Phenethylammonium Bismuth Iodide as Anode in Aqueous Zn2+ Electrolytes
Previous Article in Special Issue
Enhanced Desalination Performance of Capacitive Deionization Using Nanoporous Carbon Derived from ZIF-67 Metal Organic Frameworks and CNTs
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effects of Annealing Temperature on the Oxygen Evolution Reaction Activity of Copper–Cobalt Oxide Nanosheets

1
Department of Energy & Electronic Materials, Surface Materials Division, Korea Institute of Materials Science, Changwon 642831, Korea
2
Department of Materials Science and Engineering, Pusan National University, Busan 46241, Korea
3
IT&New Application Battery Development, LG Chem, Daejeon 34112, Korea
*
Authors to whom correspondence should be addressed.
These authors equally contributed to this work.
Submission received: 25 January 2021 / Revised: 2 March 2021 / Accepted: 4 March 2021 / Published: 8 March 2021
(This article belongs to the Special Issue Nano-Materials in Electrocatalyst)

Abstract

:
Developing high performance, highly stable, and low-cost electrodes for the oxygen evolution reaction (OER) is challenging in water electrolysis technology. However, Ir- and Ru-based OER catalysts with high OER efficiency are difficult to commercialize as precious metal-based catalysts. Therefore, the study of OER catalysts, which are replaced by non-precious metals and have high activity and stability, are necessary. In this study, a copper–cobalt oxide nanosheet (CCO) electrode was synthesized by the electrodeposition of copper–cobalt hydroxide (CCOH) on Ni foam followed by annealing. The CCOH was annealed at various temperatures, and the structure changed to that of CCO at temperatures above 250 °C. In addition, it was observed that the nanosheets agglomerated when annealed at 300 °C. The CCO electrode annealed at 250 °C had a high surface area and efficient electron conduction pathways as a result of the direct growth on the Ni foam. Thus, the prepared CCO electrode exhibited enhanced OER activity (1.6 V at 261 mA/cm2) compared to those of CCOH (1.6 V at 144 mA/cm2), Co3O4 (1.6 V at 39 mA/cm2), and commercial IrO2 (1.6 V at 14 mA/cm2) electrodes. The optimized catalyst also showed high activity and stability under high pH conditions, demonstrating its potential as a low cost, highly efficient OER electrode material.

1. Introduction

Electrochemical water splitting is an effective method of hydrogen production [1,2]. In addition, it can be ecofriendly because of the possibility of using renewable energy sources such as photovoltaic, wind, and hydroelectric power. The theoretical potential required for water splitting is 1.23 V [3,4]. However, an additional overpotential is required in both the oxygen and hydrogen evolution reactions (OER and HER, respectively). Specifically, the kinetics of OER is slow and require a higher overpotential than HER at actual operating current density of electrochemical water splitting [5]. In this regard, developing an electrocatalyst that can expedite the OER has drawn much interest and intense efforts [6,7]. The commonly used Ir- or Ru-based electrodes have high OER activities but are costly because of the use of noble metals, which have limited reserves [8,9]. Therefore, there has been ongoing research into earth-abundant transition metal-based electrodes that can replace noble metal electrodes. For example, OER catalysts comprising inexpensive and abundant resources such as transition metal (such as Co, Ni, Fe, and Cu) oxides [10,11], phosphides [12,13], and borates are drawing attention [14,15]. In particular, Co3O4 has emerged as an attractive candidate because of its efficient electrode performance resulting from its nanostructure and low cost [16,17]. However, the electrical conductivity of Co is low; thus, investigation into methods to improve the conductivity is required [18,19]. Methods to improve electrical conductivity and activity for electrocatalysts include composite transition metal catalyst synthesis, structure optimization and surface area enhancement using highly conductive materials [20,21]. Previous studies were conducted on CuxCo2-xO4 (CCO) catalysts, which have higher conductivities as a result of the Cu doping of Co3O4. In addition, various approaches such as chemical deposition, hydrothermal methods, and electrodeposition have been used to improve the surface morphology and electrochemical and crystallographic properties of the CCO catalyst [22,23].
Alkaline water electrolysis (AWE) is a mature technology for H2 production and is the most used water electrolysis method [24,25]. This AWE cells are powered by a highly concentrated aqueous alkaline electrolyte over 20 wt.% KOH of NaOH (3.6 M KOH or NaOH). [26]. However, it is difficult to secure electrode stability with high concentration alkaline electrolyte. Therefore, many AWE cells currently use steel or nickel alloy plated steel. Despite the recent results of many OER catalyst studies, there are not many reported studies on OER evaluation and stable OER catalysts above 20 wt.% KOH other than a few papers [27,28,29]. Therefore, development of stable electrodes even above 20 wt.% KOH is strongly needed. Thus, research is needed to develop a practical and convenient method for optimizing catalysts with a variety of structures and excellent electrochemical performance. Catalyst optimization can improve stability and performance, as well as electrical conductivity. For optimization, the temperature, pressure, and reactants can be controlled as parameters. However, optimization via temperature-controlled heat treatment is the easiest method to improve catalyst stability, as well as change the catalyst structure; furthermore, it is an established process.
In this study, the CCO electrode was synthesized by electrodepositing copper-cobalt hydroxide (CCOH) on nickel foam and then heat treatment, and catalytic activity and durability of the electrode were optimized. In particular, we found that the nanosheets turned into particles on the surface after heat treatment at 300 °C or higher (Figure 1). When comparing the electrochemical properties of the electrodes at a current density of 20 mA/cm2, the lowest overpotential of 276 mV was achieved for the electrode annealed at 250 °C. Furthermore, the catalytic performance was tested at a high pH, and it remained stable for 100 h.

2. Materials and Methods

2.1. Preparation of Copper–Cobalt Oxide Electrodes by Cathodic Electrodeposition

To surface oxidation (NiO) of the Ni foam (0.7 cm × 0.7 cm) substrate, etching with HCl (5 M) for 30 min was carried out, and the residual acid was removed from the substrate by washing with deionized (DI) water [30,31]. To prepare the electrodeposition solution, copper (II) nitrate hemi(pentahydrate) (10 mM, Cu(NO3)2·2.5H2O, ≥98%, Sigma-Aldrich Co., Ltd., St. Louis, MO, USA) and cobalt (II) nitrate hexahydrate (50 mM, Co(NO3)2·6H2O, ≥98%, Sigma-Aldrich Co., Ltd., St. Louis, MO, USA) were prepared in 500 mL of DI water and stirred for 30 min. The Ni foam substrate with surface oxide removed, a Pt mesh (3 cm × 4 cm), and a saturated calomel electrode (SCE) were used as the working (cathode), counter (anode), and reference electrodes, respectively (the distance of each electrode is 1 cm). Then, a constant potential of −1 V vs. SCE was applied for 5 min at 30 °C. After this process, the color of the nickel foam turned dark green. The electrode made by electrodeposition was washed with prepared DI water, and the dried at room temperature. Subsequently, heat treatment (annealing) was performed on the electrodeposited catalysts by heating at a rate of 1 °C/min to 150, 200, 250, and 300 °C in air for 3 h. The catalysts are labeled as CCO-x °C.

2.2. Physicochemical Characterization

Field-emission scanning electron microscopy (FE-SEM, JEOL, JSM-7001F, Tokyo, Japan) were used to observe the surface morphologies of the CCO electrodes. Transmission electron microscopy (TEM, Thermo Fisher Scientific, Talos F200X, Waltham, MA, USA) were used to observe the morphology and structure of CCO-250 °C nanosheets. Crystallinity analysis of copper-cobalt oxide was performed using X-ray diffractometry (XRD, RIGAKU, D/MAX-2500V, Rigaku Corporation, Akishima, Japan), which was performed using copper radiation (1.54 Å, Cu Kα) generated at a voltage and current of 40 kV and 250 mA, respectively. Measurements were carried out at a scan rate of 1°/min from 10° to 70° in 2θ. In addition, the chemical state of the electrode before and after testing the stability in 1 M KOH was obtained using X-ray photoelectron spectroscopy (XPS, Thermo Fisher, ESCALAB 250, Waltham, MA, USA) using an Al Kα (1486.6 eV) radiation source.

2.3. Electrochemical Characterization

Linear sweep voltammetry (LSV) and cyclic voltammetry (CV) were performed using a potentiostat (VMP3, Bio-Logic, Seyssinet-Pariset, France) for OER measurements. The measurements were conducted under alkaline conditions in a three-electrode system, and Pt mesh and Hg/HgO (1 M KOH) were used as the counter and reference electrodes, respectively. The OER tests were conducted in 1 M KOH solution at room temperature. Ni foam with an area of 0.49 cm2 was used as the working electrode. All potentials were corrected according to VRHE = VHg/HgO + 0.059pH + 0.098 V, where VRHE is the potential versus that of the reversible hydrogen potential, VHg/HgO is the potential versus that of the Hg/HgO electrode, and pH is the electrolyte pH. All electrochemical results were 85% IR-corrected [32]. A stability test was performed by applying a constant current of 20 mA/cm2 for about 100 h. The Tafel slope was calculated using Equation (1).
η = A × log 10 ( I I 0 )
here, η represents the overpotential (mV), I is the current density (mA/cm2), I0 is the exchange current density (mA/cm2), and A is the Tafel slope (mV/dec). Thus, the OER mechanism of the electrode was compared. For electrochemical impedance spectroscopy (EIS), measurements were conducted in 1 M KOH in a frequency range of 100 mHz to 200 kHz at a potential of 1.34 V. The electrochemical double layer capacitance (Cdl) was obtained using CV measurements in the non-Faradaic region at a scan rate of 20 to 100 mV/s).

3. Results

CuCo precursors were synthesized on Ni foam substrates from which surface Ni oxide was removed using electrodeposition using previously reported methods. FE-SEM analysis confirmed that the CuCo precursor had been uniformly deposited on the surface of the Ni foam in a nanosheet structure (Figure 2a,b). The XRD patterns (Figure 2c) of the CuCo precursor electrodes treated at all temperatures contain diffractions at 2θ = 12.9°, 25.8°, 33.6° and 38.1° which correspond to the (001), (002), (120) and (121) planes of Cu1xCox (OH)2(NO3) (CCOH), respectively [33,34]. After heat treatment, changes in the morphology and structure were observed via SEM and XRD analyses, respectively. The SEM images of the electrodes annealed at temperatures below 250 °C show that the surface morphology retained a nanosheet structure (Figure 3a–c). However, after annealing at 300 °C or above, the surface morphology became particle-like because of the structural breaking and aggregation of the nanosheets (Figure 3d). The XRD patterns (Figure 4) of the electrodes treated at all temperatures contain diffractions at 2θ = 31.2°, 36.8°, and 65.1°, which correspond to the (022), (113), and (044) planes, respectively, of Cu0.92Co2.08O4 (ICSD card No. 98-003-6356) with an inverse spinel structure. The minor peaks at 35.5° and 38.8° were indexed to the (002) and (111) planes, respectively, which were identified as CuO (ICSD card No. 98-009-2365). In addition, for the electrodes treated at 150 and 200 °C, diffractions at 2θ = 13.0° and 25.7° were observed, and these correspond to CCOH. With increase in the heat treatment temperature, the nitrate (NO3) contained in CCOH disappeared, the intensity of the CCO peak increased, and only diffractions corresponding to oxide species were observed.
The Co oxide structure is more stable than Co hydroxide for OER [35,36]. Thus, TEM and XPS analyses were performed to enable the precise structural analysis and chemical states of oxide species in the electrodes annealed at 250 °C, respectively. These nanosheets of CCO-250 °C were detached from the electrode by sonication, and then characterized by TEM. We confirmed that CuO and CCO were alloyed on the Ni foam substrate (Figure 5a). In addition, the components of the CCO-250 °C electrode were analyzed using energy-dispersive X-ray spectroscopy (EDS, Figure 5b). As a result of the analysis, the uniform distribution of Cu, Co, and O atoms in the CCO-250 C electrode was confirmed. When examining the Co 2p spectrum, Co2+ peaks were observed at 797.4 and 781.9 eV, and Co3+ peaks were observed at 795.4 and 779.9 eV (Figure 5c). The Cu 2p spectrum also contained Cu+ ion peaks at 934.0 and 953.8 eV and Cu2+ ion peaks at 935.2 and 955.2 eV (Figure 5d). The peak quantifications for the elements Co2+, Co3+, Cu+ and Cu2+ presented in CCO-250 °C were 34, 17, 23, and 26%, respectively. In general, in the spinel structure of CuxCo3-xO4, when x is 0.7 or more, Co3+ enters the tetrahedral site of spinel structure and forms an inverse spinel structure. These results indicated the formation of Cu0.92Co2.08O4, CuO, and Cu2O through the electrodeposition and heat treatment [37].
In addition, we investigated the electrocatalytic properties of the samples for the OER with respect to the heat treatment temperature. Figure 6a shows the polarization curves on the reversible hydrogen electrode (RHE) scale recorded at a scan rate of 2 mV/s in 1 M KOH solution. As the heat treatment temperature increased from 150 to 200 to 250 °C, at a low current density of 20 mA/cm2, the overpotentials decreased from 291 to 280 to 276 mV, respectively. In contrast, that of the electrode annealed at 300 °C increased to 281 mV. The same trend was observed at a high current density of 200 mA/cm2, and the lowest overpotential of 359 mV was obtained after heat treatment at 250 °C (Figure 6b). The performance of the proposed OER electrodes was compared at a current density of 10 mA/cm2 (Figure 6c). The measured overpotential of the CCO-250 °C electrode was lower than that of the reported OER catalysts under the same or similar conditions (Table 1) [38,39,40,41,42,43,44,45,46]. This result indicates that the CCO-250 °C electrode has higher OER catalytic activity compared to other catalysts. The Tafel slope relates the rate of the electrochemical reaction and the overpotential.
η = a + b log j
here, η is the overpotential, b is the Tafel slope, and j is the current density [47,48]. The Tafel slopes of the CCO electrode were 76, 77, 77, and 75 mV/dec for the electrodes treated at 150, 200, 250, and 300 °C, respectively, at an onset potential of 1.55 V (Figure 6d). These Tafel slope values were not affected by the temperature change and show similar trends, suggesting that even when the heat treatment temperature changes, the same OER mechanism applies and the catalyst maintains the same reaction rate. Electrochemical impedance spectroscopy (EIS) measurements of the CCO electrode were conducted for the samples annealed at 150, 200, 250, and 300 °C at 1.5 V (Figure 6e). The data obtained by impedance analysis is represented as a Nyquist plot, which shows the electrolyte resistance and charge transfer resistance of the electrode. As shown in Figure 6e, two semicircles of the Nyquist plot are observed. The first semicircle in the high frequency region and the second semicircle in the low frequency region represent the solid oxide film resistance (R1) and the charge transfer resistance (Rct) from the electrolyte to the catalyst surface, respectively [49,50]. The series resistance (Rs) was the same because of the use of the same electrolyte temperature, distance between the electrodes, electrode area, and experimental settings of the cell. The transfer resistance for charge transfer was the lowest for the electrode annealed at 250 °C. In addition, the electrochemically active surface area (ECSA)was determined using the Cdl to evaluate the catalytic activity of the catalyst. Cdl was calculated from the CV measurements between −0.05 and 0.05 V versus the open-circuit potential (OCV) at different scan rates. (Figure 6f). The electrodes annealed at 150, 200, 250, and 300 °C have Cdl values of 21, 52, 60, and 25 mF, respectively, indicating that the highest electrode activity was obtained at 250 °C. At temperatures above 300 °C, the surface area decreased because of particle aggregation, resulting in a decrease in electrode activity. Thus, based on catalyst performance, the electrochemical analysis results confirm that the catalyst performance was optimal at an annealing temperature of 250 °C.
Next, stability tests of the CCO-250 °C electrode, which showed the best electrode performance of the prepared electrodes, were conducted. When a constant current at a current density of 20 mA/cm2 was applied in a 1 M KOH aqueous solution, the electrode performance was maintained 100 h testing (Figure 7a). An SEM image of the CCO-250 °C catalyst after stability testing is exhibited in Figure 7b, showing that the nanosheet structure of the surface remained intact. After stability testing, XPS analysis was performed. As shown in Figure 7c,d, the Co2+ (797.1 and 781.8 eV), Co3+ (795.0 and 779.9 eV), Cu+ (934.1 and 953.8 eV), and Cu2+ (935.2 and 954.9 eV) peak positions remained almost unchanged. The Cu 2p spectrum showed the same Cu2+:Cu+ ratio as the XPS spectrum before OER. However, the Co2+:Co3+ ratio in the Co 2p spectrum changed from 67:33 to 81:19. The reason for this difference is that as the stability test proceeded, the Co2+ ions were oxidized to Co3+ ions, resulting in a change in peak intensity. During alkaline water electrolysis (AWE), the electrode performance and stability must be maintained under extreme conditions. Because AWE systems operate under highly alkaline conditions, the OER activity of the CCO-250 °C electrode was tested in 5 M KOH. By comparing the LSV curves obtained at the same potential but in KOH solutions of different concentrations, we found that the OER catalyst activity was higher in 5 M KOH (Figure 8a), probably because of the higher concentration of hydroxide anions in the 5 M KOH solution [51,52,53]. In addition, a stability test was conducted for 100 h with the same electrode, and it maintained its performance and stability even in the extreme environment (Figure 8b). Furthermore, SEM analysis of the electrode surface after the stability testing revealed that the nanosheet structure was maintained (Figure 8c).

4. Conclusions

In this study, CCO electrodes were synthesized by a practical and convenient method: electrodeposition and heat treatment. The low electrical conductivity of Co3O4, a non-noble metal electrode, was improved by the addition of copper. The effect of the changes in morphology and structure with respect to heat treatment temperature on the CCO electrode activities was investigated, allowing the optimization of the catalysts. The structure, chemical state, and morphology were analyzed using various techniques such as SEM, TEM, XRD and XPS. We confirmed that 250 °C annealed CCOH had the lowest overpotential of 276 mV at a current density of 20 mA/cm2 because all hydroxides are converted to oxides. In terms of the morphology, SEM analysis revealed that the surface maintained the nanosheet structure. Stability tests of the electrodes were conducted at 1 KOH. The synthesized electrodes showed high stability for 100 h at a current density of 20 mA/cm2. After the stability tests, the OER activity of the electrode was maintained at the same potential. The oxidation state of Co near surface through the OER increased, but the morphology and structural changes of electrode did not occur. The results indicate that the CCO electrodes are suitable and stable OER electrodes for use under extreme environments. Specifically, when stability tests were conducted at a high pH of 5 M KOH, high performance and stability were maintained and the nanosheet structure remained intact. Thus, the optimum CCO electrode was obtained by electrodeposition followed by heat treatment at 250 °C.

Author Contributions

Conceptualization, Y.S.P., S.M.C. and J.L.; Data curation, G.H.K., J.J., J.-H.L. and J.L.; Formal analysis, G.H.K., Y.S.P., J.Y. and M.J.J.; Funding acquisition, S.M.C.; Investigation, G.H.K. and H.-S.P.; Project administration, Y.H.P., S.M.C. and J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Fundamental Research Program of the Korean Institute of Materials Science (PNK7550) & the Hydrogen Energy Innovation Technology Development Program of the NRF (No. 2019M3E6A106367512) & the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT) (No. 2021R1A2C2014348).

Data Availability Statement

Data is available on the request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pu, Z.; Luo, Y.; Asiri, A.M.; Sun, X. Efficient Electrochemical Water Splitting Catalyzed by Electrodeposited Nickel Diselenide Nanoparticles Based Film. ACS Appl. Mater. Interfaces 2016, 8, 4718–4723. [Google Scholar] [CrossRef] [PubMed]
  2. Kuai, L.; Geng, J.; Chen, C.; Kan, E.; Liu, Y.; Wang, Q.; Geng, B. A Reliable Aerosol-Spray-Assisted Approach to Produce and Optimize Amorphous Metal Oxide Catalysts for Electrochemical Water Splitting. Angew. Chem. Int. Ed. 2014, 53, 7547–7551. [Google Scholar] [CrossRef]
  3. Liu, G.; Ye, S.; Yan, P.; Xiong, F.; Fu, P.; Wang, Z.; Chen, Z.; Shi, J.; Li, C. Enabling an integrated tantalum nitride photoanode to approach the theoretical photocurrent limit for solar water splitting. Energy Environ. Sci. 2016, 9, 1327–1334. [Google Scholar] [CrossRef]
  4. Bhatt, M.D.; Lee, J.S. Recent theoretical ptrogress in the development of photoanode materials for solar water splitting photoelectrochemical cells. J. Mater. Chem. A 2015, 3, 10632–10659. [Google Scholar] [CrossRef]
  5. Wang, J.; Yue, X.; Yang, Y.; Sirisomboonchai, S.; Wang, P.; Ma, X.; Abudula, A.; Guan, G. Earth-abundant transition-metal-based bifunctional catalysts for overall electrochemical water splitting: A review. J. Alloys Compd. 2020, 819, 153346. [Google Scholar] [CrossRef]
  6. Wu, X.; Zhu, C.; Wang, L.; Guo, S.; Zhang, Y.; Li, H.; Huang, H.; Liu, Y.; Tang, J.; Kang, Z. Control Strategy on Two-/Four-Electron Pathway of Water Splitting by Multidoped Carbon Based Catalysts. ACS Catal. 2017, 7, 1637–1645. [Google Scholar] [CrossRef]
  7. Liu, J.; Liu, Y.; Liu, N.; Han, Y.; Zhang, X.; Huang, H.; Lifshitz, Y.; Lee, S.T.; Zhong, J.; Kang, Z. Metal-free efficient photocatalyst for stable visible water splitting via a two-electron pathway. Science 2015, 347, 6225. [Google Scholar] [CrossRef]
  8. Cherevko, S.; Geiger, S.; Kasian, O.; Kulyk, N.; Grote, J.-P.; Savan, A.; Shrestha, B.R.; Merzlikin, S.; Breitbach, B.; Ludwig, A.; et al. Oxygen and hydrogen evolution reactions on Ru, RuO2, Ir, IrO2 thin film electrodes in acidic and alkaline electrolytes: A comparative study on activity and stability. Catal. Today 2016, 262, 170–180. [Google Scholar] [CrossRef]
  9. Audichon, T.; Mayousse, E.; Morisset, S.; Morais, C.; Comminges, C.; Napporn, T.W.; Kokoh, K.B. Electroactivity of RuO2-IrO2 mixed nanocatalysts toward the oxygen evolution reaction in a water electrolyzer supplied by a solar profile. Int. J. Hydrogen Energy 2014, 39, 16785–16796. [Google Scholar] [CrossRef]
  10. Sahoo, S.; Rout, C.S. Facile Electrochemical Synthesis of Porous Manganese-Cobalt-Sulfide Based Ternary Transition Metal Sulfide Nanosheets Architectures for High Performance Energy Storage Applications. Electrochim. Acta 2016, 220, 57–66. [Google Scholar] [CrossRef]
  11. Purcar, V.; Rădiţoiu, V.; Rădiţoiu, A.; Raduly, F.M.; Manea, R.; Frone, A.; Anastasescu, M.; Ispas, G.C.; Căprărescu, S. Bilayer coationgs based on silica materials and iron (III) phthalocyanine—Sensitized TiO2 photocatalyst. Mater. Res. Bull. 2021, 138, 111222. [Google Scholar] [CrossRef]
  12. Ma, B.; Yang, Z.; Chen, Y.; Yuan, Z. Nikel cobalt phosphide with three-dimensional nanostructure as a highly efficient electrocatalyst for hydrogen evolution reaction in both acidic and alkaline electrolytes. Nano Res. 2019, 12, 375–380. [Google Scholar] [CrossRef]
  13. Khodabakhshi, M.; Chen, S.; Ye, T.; Wu, H.; Yang, L.; Zhang, W.; Chang, H. Hierarchical High Wrinkled Trimetallic NiFeCu Phosphide Nanosheets on Nanodendrite Ni3S2/Ni Foam as an Efficient Electrocatalyst for the Oxygen Evolution Reaction. ACS Appl. Mater. Interfaces 2020, 12, 36268–36276. [Google Scholar] [CrossRef]
  14. Sabatini, M.T.; Boulton, L.T.; Sheppard, T.D. Borate ester: Simple catalysts for the sustainable synthesis of complex amides. Sci. Adv. 2017, 3, e1701028. [Google Scholar] [CrossRef] [Green Version]
  15. Jiang, W.J.; Niu, S.; Tang, T.; Zhang, Q.H.; Liu, X.Z.; Zhang, Y.; Chen, Y.Y.; Li, J.H.; Gu, L.; Wan, L.J.; et al. Crystallinity-Modulated Electrocatalytic Activity of a Nickell(II) Borate Thin Layer on Ni3B for Efficient Water Oxidation. Angew. Chem. Int. Ed. 2017, 56, 6572–6577. [Google Scholar] [CrossRef] [PubMed]
  16. Bai, B.; Arandiyan, H.; Li, J. Comparison of the performance for oxidation of formaldehyde on nano-Co3O4, 2D-Co3O4, and 3D-Co3O4 catalysts. Appl. Catal. B 2013, 142–143, 677–683. [Google Scholar] [CrossRef]
  17. Ma, C.; Wang, D.; Xue, W.; Dou, B.; Wang, H.; Hao, Z. Investigation of Formaldehyde Oxidation over Co3O4-CeO2 and Au/Co3O4-CeO2 Catalysts at Room Temperature: Effective Removal and Determination of Reaction Mechanism. Environ. Sci. Technol. 2011, 45, 3628–3634. [Google Scholar] [CrossRef]
  18. Dăncilă, A.M.; Căprărescu, S.; Bobirică, C.; Purcar, V.; Gărleanu, G.; Vasile, E.; Modrogan, C.; Borda, C.; Dobrotă, D. Optimization of the Technological Parameters for Obtaining Zn-Ti Based Composites to Increase the Performance of H2S Removal from Syngas. Processes 2020, 8, 562. [Google Scholar] [CrossRef]
  19. Purcar, V.; Răditoiu, V.; Nicolae, C.A.; Frone, A.N.; Răditoiu, A.; Raduly, F.M.; Somoghi, R.; Gabor, R.A.; Căprărescu, S. Synthesis and morpho-structural properties of TiO2-based materials. J. Optoelectron. Adv. Mater. 2019, 21, 281–286. [Google Scholar]
  20. Purcar, V.; Şomoghi, R.; Niţu, S.G.; Nicolae, C.A.; Alexandrescu, E.; Gîfu, I.C.; Gabor, A.R.; Stroescu, H.; Ianchiş, R.; Căprărescu, S.; et al. The Effect of Different Coupling Agents on Nano-ZnO Materials Obtained via Sol-Gel Process. Nanomaterials 2017, 7, 439. [Google Scholar] [CrossRef] [Green Version]
  21. Purcar, V.; Cinteza, O.; Donescu, D.; Bala, D.; Ghiurea, M.; Petcu, C.; Caprarescu, S. Surface modification of silica particles assisted by CO2. J. Supercrit. Fluids 2014, 87, 34–39. [Google Scholar] [CrossRef]
  22. Choi, W.S.; Jang, M.J.; Park, Y.S.; Lee, K.H.; Lee, J.Y.; Seo, M.H.; Choi, S.M. Three-Dimensional Honeycomb-Like Cu0.81Co2.19O4 Nanosheet Arrays Supported by Ni Foam and Their High Efficiency as Oxygen Evolution Electrodes. ACS Appl. Mater. Interfaces 2018, 10, 38663–38668. [Google Scholar] [CrossRef]
  23. Park, H.; Park, B.H.; Choi, J.; Kim, S.; Kim, T.; Youn, Y.S.; Son, N.; Kim, J.H.; Kang, M. Enhanced Electrochemical Properties and OER Performances by Cu Substitution in NiCo2O4 Spinel Structure. Nanomaterials 2020, 10, 1727. [Google Scholar] [CrossRef]
  24. Ursúa, A.; Barrios, E.L.; Pascual, J.; Martin, I.S.; Sanchis, P. Integration of commercial alkaline water electrolysers with renewable energies: Limitations and improvements. Int. J. Hydrogen Energy 2016, 41, 12852–12861. [Google Scholar] [CrossRef]
  25. Schalenbach, M.; Kasian, O.; Mayrhofer, K.J.J. An alkaline water electrolyzer with nickel electrodes enables efficient high current density operation. Int. J. Hydrogen Energy 2018, 43, 11932–11938. [Google Scholar] [CrossRef]
  26. Miller, H.A.; Bouzek, K.; Hnat, J.; Loos, S.; Bernäcker, C.I.; Weißgärber, T.; Röntzsch, L.; Meier-Haack, J. Green hydrogen from anion exchange membrane water electrolysis: A review of recent developments in critical materials and operating conditions. Sustain. Energy Fuels 2020, 4, 2114–2133. [Google Scholar] [CrossRef]
  27. Andronescu, C.; Seisel, S.; Wilde, P.; Barwe, S.; Masa, J.; Chen, Y.T.; Ventosa, E.; Schuhmann, W. Influence of Temperature and Electrolyte Concentration on the Structure and Catalytic Oxygen Evolution Activity of Nickel-Iron Layered Double Hydroxide. Chem. Eur. J. 2018, 24, 13773–13777. [Google Scholar] [CrossRef] [PubMed]
  28. Liu, Y.; Liang, X.; Gu, L.; Zhang, Y.; Li, G.D.; Zou, X.; Chen, J.S. Corrosion engineering towards efficient oxygen evolution electrodes with stable catalytic activity for over 6000 hours. Nat. Commun. 2018, 9, 2609. [Google Scholar] [CrossRef] [PubMed]
  29. Lee, J.; Lee, H.; Lim, B. Chemical transformation of iron alkoxide nanosheets to FeOOH nanoparticles for highly active and stable oxygen evolution electrocatalysts. J. Ind. Eng. Chem. 2018, 58, 100–104. [Google Scholar] [CrossRef]
  30. Guo, C.X.; Li, C.M. Room temperature-formed iron-doped nickel hydroxide on nickel foam as a 3D electrode for low polarized and high-current-density oxygen evolution. Chem. Commun. 2018, 54, 3262–3265. [Google Scholar] [CrossRef]
  31. Maiti, U.N.; Lim, J.; Lee, K.E.; Lee, W.J.; Kim, S.O. Three-Dimensional Shape Engineered, Interfacial Gelation of Reduced Graphene Oxide for High Rate, Large Capacity Supercapacitors. Adv. Mater. 2014, 26, 615–619. [Google Scholar] [CrossRef]
  32. Yu, L.; Ren, Z. Systematic study of the influence of iR compensation on water electrolysis. Mater. Today Phys. 2020, 14, 100253. [Google Scholar] [CrossRef]
  33. Jalal, M.R.; Hojjati, H.; Jalal, J.R.; Ebrahimi, S.; Bighashi, M.R.Z. Synthesis of Copper Hydroxide Nitrate (Cu2(OH)3NO3) micro-sheets by plasma electrolysis of Cu(NO3)2 aqueous solution in atmospheric air. J. Interfaces Thin Films Low Dimens. Syst. JITL 2018, 1, 109–112. [Google Scholar]
  34. Wang, X.; Huang, L. A novel one-step method to synthesize copper nitrate hydroxide nanorings. Trans. Nonferrous Met. Soc. China 2009, 19, s480–s484. [Google Scholar] [CrossRef]
  35. Kalyani, M.; Emerson, R.N. Electgrodeposition of nano crystalline cobalt oxide on porous copper electrode for supercapacitor. J. Mater. Sci. Mater. Electron. 2019, 30, 1214–1226. [Google Scholar] [CrossRef]
  36. Xia, H.; Peng, Z.; Cuncail, V.; Zhao, Y.; Hao, J.; Huang, Z. Self-supported porous Cobalt Oxide Nanowires with enhanced Electrocatalytic performance toward Oxygen evolution reaction. J. Chem. Sci. 2016, 128, 1879–1885. [Google Scholar] [CrossRef] [Green Version]
  37. Chi, B.; Lin, H.; Li, J. Cations distribution of CuxCo3−xO4 and its electrocatalytic activities for oxygen evolution reaction. Int. J. Hydrog. Energy 2008, 33, 4763–4768. [Google Scholar] [CrossRef]
  38. Zhang, B.; Jiang, K.; Wang, H.; Hu, S. Fluoride-Induced Dynamic Surface Self-Reconstruction Produces Unexpectedly Efficient Oxygen-Evolution Catalyst. Nano Lett. 2019, 19, 530–537. [Google Scholar] [CrossRef] [PubMed]
  39. Liu, K.; Zhang, C.; Sun, Y.; Zhang, G.; Shen, X.; Zou, F.; Zhang, H.; Wu, Z.; Wegener, E.C.; Taubert, C.J.; et al. High-Performance Transition Metal Phosphide Alloy Catalyst for Oxygen Evolution Reaction. ACS Nano 2018, 12, 158–167. [Google Scholar] [CrossRef]
  40. Hai, G.; Jia, X.; Zhang, K.; Liu, X.; Wu, Z.; Wang, G. High-performance oxygen evolution catalyst using two-dimensional ultrathin metal-organic frameworks nanosheets. Nano Energy 2018, 44, 345–352. [Google Scholar] [CrossRef]
  41. Xu, H.; Wang, B.; Shan, C.; Xi, P.; Liu, W.; Tang, Y. Ce-Doped NiFe-Layered Double Hydroxide Ultrathin Nanosheets/Nanocarbon Hierarchical Nanocomposite as an Efficient Oxygen Evolution Catalyst. ACS Appl. Mater. Interfaces 2018, 10, 6336–6345. [Google Scholar] [CrossRef]
  42. Wang, Y.; Wang, S.; Liu, D.; Zhou, L.; Du, R.; Li, T.T.; Miao, T.; Qian, J.; Hu, Y.; Huang, S. Normal-pulse-voltage-assisted in situ fabrication of graphene-wrapped MOF-derived CuO nanoflowers for water oxidation. Chem. Commun. 2020, 56, 8750–8753. [Google Scholar] [CrossRef]
  43. Bai, L.; Hsu, C.S.; Alexander, D.T.L.; Chen, H.M.; Hu, X. A Cobalt-Iron Double-Atom Catalyst for the Oxygen Evolution Reaction. J. Am. Chem. Soc. 2019, 141, 14190–14199. [Google Scholar] [CrossRef]
  44. Stern, L.A.; Feng, L.; Song, F.; Hu, X. Ni2P as a Janus catalyst for water splitting: The oxygen evolution activity of Ni2P nanoparticles. Energy. Environ. Sci. 2015, 8, 2347–2351. [Google Scholar] [CrossRef]
  45. Babar, P.T.; Lokhande, A.C.; Pawar, B.S.; Gang, M.G.; Jo, E.; Go, C.; Suryawanshi, M.P.; Pawar, S.M.; Kim, J.H. Electrocatalytic performance evaluation of cobalt hydroxide and cobalt oxide thin films for oxygen evolution reaction. Appl. Surf. Sci. 2018, 427, 253–259. [Google Scholar] [CrossRef]
  46. Dang, W.J.; Shen, Y.Q.; Lin, M.; Jiao, H.; Xu, L.; Wang, Z.L. Noble-metal-free electrocatalyst based on a mixed CoNi metal-organic framework for oxygen evolution reaction. J. Alloys Compd. 2019, 792, 69–76. [Google Scholar] [CrossRef]
  47. Hu, Q.; Li, G.; Liu, X.; Zhu, B.; Li, G.; Fan, L.; Chai, X.; Zhang, Q.; Liu, J.; He, C. Coupling pentlandite nanoparticles and dual-doped carbon networks to yield efficient and stable electrocatalysts for acid water oxidation. J. Mater. Chem. A 2019, 7, 461–468. [Google Scholar] [CrossRef]
  48. Wang, F.; Xue, H.; Tian, Z.; Xing, W.; Feng, L. Fe2P as a novel efficient catalyst promoter in Pd/C system for formic acid electro-oxidation in fuel cells reaction. J. Power Source 2018, 375, 37–42. [Google Scholar] [CrossRef]
  49. Zhang, W.; Qi, J.; Liu, K.; Cao, R. A Nickel-Based Integrated Electrode from an Autologous Growth Strategy for Highly Efficient Water Oxidation. Adv. Energy Mater. 2016, 6, 1502489. [Google Scholar] [CrossRef]
  50. Gao, M.; Sheng, W.; Zhuang, Z.; Fang, Q.; Gu, S.; Jing, J.; Yan, Y. Efficient Water Oxidation Using Nanostructured α–Nickel-Hydroxide as an Electrocatalyst. J. Am. Chem. Soc. 2014, 136, 7077–7084. [Google Scholar] [CrossRef]
  51. Andronescu, C.; Barwe, S.; Ventosa, E.; Masa, J.; Vasile, E.; Konkena, B.; Möller, S.; Schuhmann, W. Powder Catalyst Fixation for Post-Electrolysis Structural Characterization of NiFe Layered Double Hydroxide Based Oxygen Evolution Reaction Electrocatalysts. Angew. Chem. Int. Ed. 2017, 56, 11258–11262. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Yang, C.; Rousse, G.; Svane, K.L.; Pearce, P.E.; Abakumov, A.M.; Deschamps, M.; Cibin, G.; Chadwich, A.V.; Corte, D.A.; Hansen, H.A.; et al. Cation insertion to break the activity/stability relationship for highly active oxygen evolution reaction catalyst. Nat. Commun. 2020, 11, 1378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Lu, X.; Zhao, C. Electrodeposition of hierarchically structured three-dimensional nickel-iron electrodes for efficient oxygen evolution at high current densities. Nat. Commun. 2015, 6, 6616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Schematic illustration of CCO catalyst preparation using electrodeposition and heat treatment at 200, 250, and 300 °C.
Figure 1. Schematic illustration of CCO catalyst preparation using electrodeposition and heat treatment at 200, 250, and 300 °C.
Nanomaterials 11 00657 g001
Figure 2. (a,b) SEM images and (c) XRD pattern of electrodeposited CCOH. (d) Proposed structure of Cu1-xCox(OH)2(NO3).
Figure 2. (a,b) SEM images and (c) XRD pattern of electrodeposited CCOH. (d) Proposed structure of Cu1-xCox(OH)2(NO3).
Nanomaterials 11 00657 g002
Figure 3. SEM images showing the changes in CCO morphology with temperature: (a) 150, (b) 200, (c) 250, and (d) 300 °C, respectively.
Figure 3. SEM images showing the changes in CCO morphology with temperature: (a) 150, (b) 200, (c) 250, and (d) 300 °C, respectively.
Nanomaterials 11 00657 g003
Figure 4. XRD patterns of CCO after treatment at different temperatures: (from top to bottom) 150, 200, 250, and 300 °C.
Figure 4. XRD patterns of CCO after treatment at different temperatures: (from top to bottom) 150, 200, 250, and 300 °C.
Nanomaterials 11 00657 g004
Figure 5. (a) High-resolution TEM image (blue, CuO d(200) = 0.229 nm; red, CCO d(133) = 0.187 nm and d(224) = 0.163 nm), (b) EDS elemental maps of (clockwise) O, Co, and Cu, and (c) Co 2p and (d) Cu 2p XPS core level spectrum of CCO-250 °C.
Figure 5. (a) High-resolution TEM image (blue, CuO d(200) = 0.229 nm; red, CCO d(133) = 0.187 nm and d(224) = 0.163 nm), (b) EDS elemental maps of (clockwise) O, Co, and Cu, and (c) Co 2p and (d) Cu 2p XPS core level spectrum of CCO-250 °C.
Nanomaterials 11 00657 g005
Figure 6. (a,b) LSV curves (85% IR-compensated) and (c) overpotentials at various current densities for CCO catalysts (150, 200, 250, and 300 °C) in 1 M KOH at room temperature. (d) Tafel plots, (e) EIS results, and (f) Cdl of CCO catalysts (150, 200, 250, and 300 °C).
Figure 6. (a,b) LSV curves (85% IR-compensated) and (c) overpotentials at various current densities for CCO catalysts (150, 200, 250, and 300 °C) in 1 M KOH at room temperature. (d) Tafel plots, (e) EIS results, and (f) Cdl of CCO catalysts (150, 200, 250, and 300 °C).
Nanomaterials 11 00657 g006
Figure 7. (a) Chronopotentiometry curve of CCO-250 °C at a current density of 20 mA cm−2 in 1 M KOH for 100 h. Inset show the SEM image of surface morphology for CCO after durability test. (b) XRD patterns, (c) Co 2p and (d) Cu 2p of CCO-250 °C after durability testing.
Figure 7. (a) Chronopotentiometry curve of CCO-250 °C at a current density of 20 mA cm−2 in 1 M KOH for 100 h. Inset show the SEM image of surface morphology for CCO after durability test. (b) XRD patterns, (c) Co 2p and (d) Cu 2p of CCO-250 °C after durability testing.
Nanomaterials 11 00657 g007
Figure 8. (a) LSV curves (85% IR-compensated) for CCO-250 °C in 1 and 5 M KOH at room temperature. (b) Chronopotentiometry curve of CCO (250 °C) at a current density of 20 mA cm−2 in 5 M KOH for 100 h. (c) SEM image showing the surface morphology of CCO (250 °C) after durability testing.
Figure 8. (a) LSV curves (85% IR-compensated) for CCO-250 °C in 1 and 5 M KOH at room temperature. (b) Chronopotentiometry curve of CCO (250 °C) at a current density of 20 mA cm−2 in 5 M KOH for 100 h. (c) SEM image showing the surface morphology of CCO (250 °C) after durability testing.
Nanomaterials 11 00657 g008
Table 1. Comparison of the reported OER activity.
Table 1. Comparison of the reported OER activity.
SampleTemperatureElectrolyteCurrent Density
(mA/cm2)
Overpotential
(mV)
Ref.
CCORoom temp.1 M KOH10240Our work
NiFe-OH-F243[38]
Fe-Co-P252[39]
NiFe-UMNs260[40]
1% Ce-NiFe-LDH/CNT270[41]
NiFe-LDH/CNT299[41]
CuO NF@G/CF320[42]
Co-N-C321[43]
Ni(OH)2330[44]
Co3O4340[45]
Ni-NDC/PANI-NF361[46]
Ni365[44]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kim, G.H.; Park, Y.S.; Yang, J.; Jang, M.J.; Jeong, J.; Lee, J.-H.; Park, H.-S.; Park, Y.H.; Choi, S.M.; Lee, J. Effects of Annealing Temperature on the Oxygen Evolution Reaction Activity of Copper–Cobalt Oxide Nanosheets. Nanomaterials 2021, 11, 657. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11030657

AMA Style

Kim GH, Park YS, Yang J, Jang MJ, Jeong J, Lee J-H, Park H-S, Park YH, Choi SM, Lee J. Effects of Annealing Temperature on the Oxygen Evolution Reaction Activity of Copper–Cobalt Oxide Nanosheets. Nanomaterials. 2021; 11(3):657. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11030657

Chicago/Turabian Style

Kim, Geul Han, Yoo Sei Park, Juchan Yang, Myeong Je Jang, Jaehoon Jeong, Ji-Hoon Lee, Han-Saem Park, Yong Ho Park, Sung Mook Choi, and Jooyoung Lee. 2021. "Effects of Annealing Temperature on the Oxygen Evolution Reaction Activity of Copper–Cobalt Oxide Nanosheets" Nanomaterials 11, no. 3: 657. https://0-doi-org.brum.beds.ac.uk/10.3390/nano11030657

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

Article Metrics

Back to TopTop