Next Article in Journal
Biomass-Derived Carbon Anode for High-Performance Microbial Fuel Cells
Next Article in Special Issue
Photocatalytic Oxidative Desulfurization of Thiophene by Exploiting a Mesoporous V2O5-ZnO Nanocomposite as an Effective Photocatalyst
Previous Article in Journal
3D-Structured Au(NiMo)/Ti Catalysts for the Electrooxidation of Glucose
Previous Article in Special Issue
Explorative Sonophotocatalytic Study of C-H Arylation Reaction of Pyrazoles Utilizing a Novel Sonophotoreactor for Green and Sustainable Organic Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrochemical Characteristics of Nanosized Cu, Ni, and Zn Cobaltite Spinel Materials

by
Mohamed Mokhtar M. Mostafa
1,*,
Wejdan Bajafar
1,
Lin Gu
2,
Katabathini Narasimharao
1,
Mohamed Abdel Salam
1,
Abdulmohsen Alshehri
1,
Nezar H. Khdary
3,
Sulaiman Al-Faifi
1 and
Abhishek Dutta Chowdhury
2,*
1
Chemistry Department, Faculty of Science, King Abdulaziz University, Jeddah 21589, Saudi Arabia
2
College of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, China
3
King Abdulaziz City for Science and Technology, Riyadh 11442, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Submission received: 17 June 2022 / Revised: 1 August 2022 / Accepted: 11 August 2022 / Published: 13 August 2022
(This article belongs to the Special Issue The Role of Catalysts in Functionalization of C-H and C-C Bonds)

Abstract

:
For a long time, transition metal oxide systems have been considered well explored materials in heterogeneous catalysis. Amongst, the spinel-type oxides, materials such as cobaltites (Co3O4) received significant attention, owing to their use in many industrial applications. In the present study, nanosized Cu, Ni, and Zn cobaltite spinel oxides were synthesized by a simple hydrothermal method. Physicochemical characterization of the synthesized materials was performed utilizing XRD, HRTEM, CO2-TPD, and XPS techniques. The textural characteristics (BET-surface area, pore size, etc.) of samples were determined from N2 physisorption measurements at −196 °C. The CO2-electrocatalytic reduction was selected as a model reaction to evaluate the electrochemical performance of the synthesized spinel cobaltites. For Ni, Cu, and Zn spinel materials, hydrogen was produced as the main product at the whole potential, along with other products, such as CO and HCOOH. Despite the advantages, the catalytic electrochemical CO2 reduction performance of spinel cobaltite catalysts is still far from adequate, which is principally ascribed to the low number of active sites combined with poor electrical conductivity.

1. Introduction

In the past decades, researchers utilized noble metal (Au, Ag, and Pd) based electrocatalysts to obtain superior performances for electrochemical reduction of CO2 [1]. However, these catalysts are less selective, expensive, have limited availability, and their catalytic properties decline rapidly. Therefore, it is necessary to find an inexpensive alternative catalyst to obtain enhanced catalytic performance in CO2 electrochemical reduction [1,2]. Recently, transition metal oxide-based electrocatalysts have provoked increasing attention [2]. These materials possessed additional advantages, such as high abundance, low cost, and easy preparation [3]. The widest class of materials examined in electrocatalysis is the transition metal oxide systems, which have been employed for a long time in heterogeneous catalysis. The research was focused on dioxides, perovskites, pyrochlores, and spinel-type transition metal oxides [4]. Among them, spinel type oxides have received great interest due to their unique crystal structure and chemical and thermal stability. Spinel type oxides typically expressed with the general formula of AB2O4, where AII is a divalent cation, e.g., Mn, Co, Mg, Fe, Cr, Zn, Ni, and Cu, and BIII is a trivalent cation, e.g., Al, Ti, Ga, V, Cr, Mn, Fe, Co, and Ni [5,6]. Many of spinel oxides have intriguing electronic, optical, and magnetic properties, making them suitable for a wide range of technological applications [5].
Numerous metal oxides have been studied extensively as electrocatalysts, including Co3O4-based catalysts [7,8,9,10] for oxidation of water, oxygen reduction reactions, and hydrogen evolution processes. Nanostructured spinel Co3O4’s electrochemical activity is mainly due to the presence of mixed oxidation states (Co2+ in tetrahedral and Co3+ in octahedral sites) and highly active surface sites [11]. According to Gao et al., ultrathin layers of Co3O4 exhibit strong CO2 reduction activity and are particularly selective for a formic acid generation [12]. Because of the irregular electrical structure of ultrathin films and the ease with which electronic transfer, carrier transport, and low corrosion rate are facilitated by the films, this activity was linked not only to an increased number of surface sites typical of low-dimensional systems. For effective CO2 adsorption and enhanced H+ transfer, nanosized Co-based spinel oxides have been shown to have significant CO2RR activity [5,12] due to the presence of both metal (for an electron transfer into CO2) and crystalline metal oxide phases. Bulk and transition metal containing cobaltites (Co3O4, NiCo2O4, CuCo2O4 and ZnCo2O4) are among the spinel-type oxides that have attracted a great deal of attention over the past several decades due to their wide range of industrial applications. Anode and cathode compounds for oxygen evolution reactions have been extensively studied due to their promising activity and stability in alkaline media. These oxide materials are not stable in acidic environments. The benefits of using these oxides as electrode materials include their activity, availability, low cost, thermodynamic stability, low electrical resistance, and friendliness to the environment. There are two types of cubic oxide lattices: normal Co3O4 and MIICo2O4 (inverted spinel). For decades, Co3O4 and NiCo2O4 have been known to be active and stable catalysts, not only for oxygen evolution and reduction but also for organic electro-oxidation in alkaline environments [13,14,15,16,17,18].
There are different methods that can be utilized to prepare spinels, including solid-state, hydrothermal, sol-gel, co-precipitation, solvothermal, and sonochemical techniques [19]. Table S1 summarizes some spinels utilized for the CO2 reduction reaction, where cobalt containing spinel catalysts resulted in the formation of formic acid. In this context, the present study evaluates the electrochemical characteristics and catalytic activity of metal (Zn, Ni and Cu) spinel cobaltites catalysts for CO2 reduction.

2. Results and Discussion

The XRD analysis was used to investigate the crystal structure and phase detection of synthesized spinels, and the obtained patterns are shown in Figure 1. Reflections at 19°, 31.6°, 37.4°, and 44.6° correspond to the (110), (220), (310), (400), (511), and (440) planes for the cubic structure of the spinel [20]. The major reflections in the XRD patterns of the spinels can be indexed to NiCo2O4 (JCPDS file no. 73-1702), CuCo2O4 (JCPDS file no. 1-1155) and ZnCo2O4 (JCPDS file no. 23-1390) [21,22,23]. There were no reflections observed for any other crystal phase, indicating the purity of the synthesized materials. It is interesting to note that the NiCo2O4 sample exhibited relatively broad reflections compared to other two samples, indicating the relatively small crystallite size in the case of the NiCo2O4 sample. The average crystallite size of the synthesized spinels was calculated by using the Scherrer equation. The calculated average crystallite sizes were 14 nm, 30 nm, and 31 nm for NiCo2O4, ZnCo2O4, and CuCo2O4 samples respectively.
The TEM images (Figure 2) show that Ni spinel has a typical three-dimensional porous architecture with interconnected nanowires. High-resolution (HR) TEM images (Figure 2a,b) show that Ni spinel sample contained uniform nanowires with an average diameter of approximately 12 nm. From Figure 2c–f, it can be observed that the Cu spinel and the Zn spinel consist of nonuniform nanoparticles with an average diameter of approximately 30 nm, which is in concurrence with the XRD results.
Figure 3 illustrates the nitrogen adsorption-desorption isotherms of synthesized spinel catalysts. All the isotherms can be classified as type IV according to IUPAC classification [24,25]. The hysteresis loops in the isotherms indicate the mesoporous nature of all spinels, especially the NiCo2O4. The H3-type hysteresis loop of spinel catalysts indicates that the type of holes is slit pores formed by the aggregation of sheet particles [23,26]. Table 1 displays the BET surface area (SBET), micropore surface area (Smicro), mesopore surface area (Smeso), total pore volume (Vtotal), micropore volume (Vmicro), and mesopore volume (Vmeso) of all the investigated samples. The synthesized nanosized NiCo2O4, ZnCo2O4 and CuCo2O4 spinel catalysts possessed a BET surface area of 41, 32, and 28 m2/g, respectively [22,23]. The pore size distribution patterns were obtained by applying the Non-Local Density Functional Theory (NLDFT) method. The obtained patterns indicated that the synthesized samples have multimodal distribution curves in the microporous, mesoporous, and macroporous ranges [21]. Smeso and Vmeso are higher in NiCo2O4 spinel than other spinels, indicating that this particular spinel exhibited a larger surface area and mesoporous feature. The hierarchy factor (HF) increases in CuCo2O4 and ZnCo2O4 spinels, which indicates that the BET surface area was significantly reduced as a function of the pronounced suppression of the mesopore volume of these samples. NLDFT calculations show a high number of pores with a diameter of approximately 2 nm, which is on the edge between micro and mesopores. However, the samples possessed specific surface areas of 41, 28, and 32 m2/g, which would be characteristic for materials with much wider pores. The average pore size was calculated as 3 nm, 3 nm, and 13 nm for Cu, Zn, and Ni cobaltite samples, respectively. The presence of a H3 hysteresis loop indicates that the samples have slit-type pores due to the aggregation of spinel particles [23,26]. Therefore, the meso and macro porosity observed in the samples is due to the arrangement of particles, which explains the low surface area of the samples. In general, the presence of micropores is responsible for a high surface area. These observed results indicate that the synthesized spinel materials possessed a relatively high surface area and porosity, which enhance the diffusion of molecules. In addition, the nanowire or nanoparticle morphology could increase the contact area for reacting molecules, providing enough active sites for reactions.
The surface structure and oxidation state of the metals presented in the synthesized spinels were analyzed using XPS analysis. The full survey and deconvoluted XPS spectra for the three samples are shown in Figures S1 and S2, respectively. For the sake of clarity, only 2p3/2 components for the metal (Cu, Ni, Zn, and Co) species are presented in the figure. The Gaussian fitted XPS spectrum of the CuCo2O4 sample shows the presence of Cu, Co, and O elements. It is known that chemical state differentiation is not straight forward with XPS, as the Cu2p3/2 peak for Cu°, Cu+, and Cu2+ species appears in the range of 933–933.5 eV [24,27]. However, it is possible to distinguish the Cu oxidation states; using the satellite peak features of Cu2p as the Cu (II) species exhibits strong observable satellite features around 943 eV [28]. The presence of strong satellite peaks confirm the characteristic of Cu2+ species in the synthesized CuCo2O4 sample; therefore, the Cu 2p3/2 peaks observed at 932.9 eV and 934.7 eV could be attributed to Cu (II) oxide and Cu (II) hydroxide species. The Co2p region present in the synthesized three spinel samples (CuCo2O4, NiCo2O4, and ZnCo2O4) and the Co 2p3/2 can be fitted into two peaks. The fitted peaks at 780.0 and 781.6 eV correspond to Co3+ and Co2+ species. It was reported that bulk cobalt oxide, such as Co3O4, possesses mixed oxidation state of Co2+ and Co3+, and it is expected to show satellite features due to both Co2+ and Co3+ states. Co2+ has an observable satellite feature around 786 eV [29]. Moreover, the integral area of the Co3+ peak is much larger than that of the Co2+ peak, which reveals that the main oxidation state of Co element in the three spinel samples is Co3+ [24,27]. The presence of mixed oxidation states of the same cation in spinel systems (Co2+ and Co3+) and the presence of highly active surface sites are the primary sources of electrochemical activity in cobaltite spinels. In this case, the ultrahigh part of low-coordinated surface Co atoms may act as the potential active sites for efficiently adsorbing CO2 [30,31].
It is well reported that Ni contained spinels can show complex multiplet-split peaks [32]. The Ni 2p3/2 can be fitted into two peaks at 855.7 eV and 858.3 eV. The first peak could be assigned to Ni3+ and the other peak at 858.3 eV could be related to Ni2+ cations in octahedral or tetrahedral sites of the NiCo2O4 spinel structure. This binding energy could be assigned to Ni cations or surface Ni(OH)2 species [33]. The Zn 2p3/2 component showed a strong peak at 1021.0 eV, corresponding to the Zn2+ in the ZnCo2O4 structure. It was also observed that the binding energy different between the Zn 2p3/2 and Zn 2p1/2 peaks is approximately 23 eV, which is in accordance with Hu et al. [34].
The O 1s spectra of three samples can be fitted into three peaks. The peak at 529.5 eV matches to the oxygen in the crystal lattice that bonded with metal ions, the peak at 531.5 eV corresponds to the oxygen of OH- groups, and the peak at 533.3 eV corresponds to the oxygen of physically adsorbed H2O molecules [20,24]. These observations clearly indicating that the surfaces of synthesized spinels are hydroxylated, which could be useful functional groups for catalyzing different types of reactions.
CO2-TPD experiments were performed to evaluate the number of active sites presented in the synthesized spinel catalysts to adsorb the CO2 molecules. The CO2-TPD patterns obtained for the three samples are shown in Figure 4. The deconvolution of the desorption peaks for TPD pattern of the NiCo2O4 sample indicates the presence of two desorption peaks in the temperature range of 430−730 °C, while the CuCo2O4 sample exhibited between 500 and 770 °C. However, the ZnCo2O4 sample yielded peaks between 625 and 730 °C. In addition, the samples showed intense and broad CO2 desorption peaks at high temperatures, and these observed results indicated that the samples possessed strong CO2 adsorption ability. Furthermore, it was found that the NiCo2O4 catalyst can adsorb a large number of CO2 molecules (1089.1 μmol/g compared to the other two samples (Table 2)). Similar results were previously observed that NiCo2O4 catalyst has a higher CO2 adsorption ability than both copper and zinc cobaltites [35].

3. Catalytic Activity

Using an H-cell separated by a Nafion proton exchange membrane, the catalytic activity of Ni spinel, Cu spinel, and Zn spinel for CO2 reduction was measured. There was a sealed compartment for the working electrode and the reference electrode (Ag/AgCl electrode) and a separate compartment for the counter electrode (Pt electrode). Three electrodes were used to measure the linear sweep voltammetry (LSV) of the catalysts in CO2 saturated 0.1 M KHCO3 aqueous solution. As a result, the current density was normalized to the area of the carbon paper electrode, approximately 1 cm2. Figure 5a shows the LSV scans of the spinel catalysts. In terms of potential range, the current density of Ni spinel is much higher than that of Cu and Zn spinel. It suggests that Cu and Zn spinel has a lower overpotential to reduce CO2 than other spinel materials. As shown in Figures S3–S5, the Ni, Zn, and Cu materials were also evaluated at different times in CO2-saturated 0.1 M KHCO3 and reproduced independently by two separate systems. This observation supports its practical applicability in terms of efficiency, flexibility, and durability. To further gain insights into the intrinsic activity of the Ni, Cu, and Zn spinel, measurement of double layer capacitance (Cdl) was herein tested, which is proportional to the electrocatalytic active surface areas (ECSA) in the electrocatalysts [36] (Figures S6–S8). As shown in Figure 5b, the Cdl of Cu spinel is ≈ 1.85 and 1.12 times higher than that of Zn and Ni spinel, suggesting that the Cu spinel has relatively more exposed active electrocatalytic sites. When the LSV curves are normalized by the ECSA to highlight the intrinsic activity, as shown in Figure 5c, Zn spinel showed relatively better CO2RR activity than Ni and Cu spinel. Figure S9 shows the Nyquist plots of Ni, Cu, and Zn spinel for CO2RR, the contact and solution resistance (14.07 Ω) of the Zn spinel is also the lowest among Ni spinel (15.53 Ω) and Cu spinel (16.34 Ω).
To analyze the reduction product of CO2 and gain insight into the selectivity of the spinel catalysts, steady-state electrolysis was conducted at different potentials from −0.9 to −1.1 V vs. RHE. At each potential, CO2 reduction reaction was conducted to reach a total electric charge of 50 C. For Ni, Cu, and Zn spinel, hydrogen was produced as the main product at the whole potential, along with other products, such as CO and HCOOH (Figure 5d–f). The FE of Ni spinel to H2 is higher than 90% at −1.1 V vs. RHE (Figure 5d). Similarly, for the Cu and Zn spinel, the formation of H2 is still the major product; and the selectivity toward H2, CO, and HCOOH is less potential-dependent (Figure 5e,f). For Ni and Zn spinel, the FE of H2 is slightly lower than that of Cu spinel, e.g., 82 and 90% at −0.9 V vs. RHE and the C1 product FE of less than 20% was also observed (Figures S10 and S11). The comparison of partial current density for H2, CO, and formate production with different catalysts is shown in Figure 5g–i. It can be seen that the partial current density for H2 species production would remarkably increase along with the raise of polarization potential, approximate 12.1 mA cm−2 with Ni spinel, 10.6 mA cm−2 with Cu spinel and 14.6 mA cm−2 with Zn spinel at −1.1 V vs. RHE, indicative of H2 as the main product from the water.
The stability of Ni, Cu, and Zn spinel was evaluated using i-t measurement at −0.9 V vs. RHE under a continuous flow of CO2 (Figure 5j–l). Figure 5j,l showed that the Ni and Zn spinel were relatively more stable than Cu spinel and maintained their excellent activity after 6 h at −0.9 V vs. RHE. Meanwhile, for Ni, Zn, and Cu spinel, the FEs of CO and HCOOH are stably maintained throughout the whole CO2RR process, respectively.
Besides, we also discuss the chemical composition and valence state of Zn, Ni, and Cu spinel coated on carbon paper before and after the CO2RR. As shown in Figure 6a, the M 2p3/2 signals (M = Co, Zn) were negatively shifted by 0.7 and 0.5 eV, compared to those for fresh Zn spinel. The negative shift confirms the electronic charge redistribution on their atomic interface between Zn–Co. In the Ni spectra (Figure 6b), the valence states of Ni2+ and Ni3+ were relatively stable without significant changes, while the Co3+ 2p3/2 and Co2+ 2p3/2 shifted to the lower binding energy after the CO2RR for Ni spinel, indicating a strong charge transfer from Co to O atom. For Cu spinel (Figure 6c), the significant binding energy shifted in XPS spectra of Cu and Co also suggests the strong charge transfer from Cu to Co atom through the Cu—Co atomic interface, leading to the decreased electron density of Cu.

4. Material and Methods

4.1. Materials

Zinc nitrate hexahydrate, cobalt nitrate hexahydrate, copper nitrate trihydrate, and nickel nitrate hexahydrate were purchased from Sigma-Aldrich, Darmstadt, Germany. Pure urea ethanol was obtained from Fisher Scientific, Waltham, MA USA.

4.2. Synthesis of Different Spinel Catalysts

The ZnCo2O4, CuCo2O4, and NiCo2O4 spinel catalysts were prepared by a co-precipitation followed by hydrothermal treatment under autogenous pressure. One mmol of Zn (NO3)2.6H2O, Cu (NO3)2.3H2O, and Ni (NO3)2.6H2O were mixed individually with two mmol of Co (NO3)2.6H2O in 50 mL of deionized water (DI) under vigorous stirring. A total of 60 mmol of urea dissolved in 100 mL of deionized water was then added drop by drop into the above solutions as a precipitating agent at pH 11–12 and thereafter stirred for 30 min. Then, the total contents were poured into a 200 mL Teflon-lined stainless-steel autoclave and subjected to hydrothermal treatment at 180 °C for 6h. After hydrothermal treatment, the obtained precipitate was filtered and washed with t deionized water and ethanol. Finally, the washed cake was dried in an electric oven for 6 h at 60 °C and calcined at 400 °C for 3 h in air at a heating rate of 3 °C/min.

4.3. Characterization of Catalysts

The crystallographic structure of all prepared solid samples was characterized by X-ray diffraction using Bruker diffractometer (Brucker D8 Advance, Karlsruhe, Germany) equipped with copper Kα (λ = 1.5405 Å) and a monochromator at 40 kV and 40 mA, in the 2 θ range of 5–80°. The HRTEM images of the samples were obtained using a FEI Tecnai G2 F30 TEM microscope, in Bright Field (BF) imaging mode. The samples were prepared for TEM image analysis by dispersing catalyst powder in alcohol by ultra-sonication, and a droplet of solution was dropped onto ultra-thin carbon film. The BET surface area, total pore volume, and pore radius of different spinel catalysts were determined from results of N2 adsorption/desorption experiments performed at −196 °C using a Quantachrome Autosorb-iQ instrument after degassing the samples at 120 °C for 5 h. The surface elemental compositions of the samples were analyzed by X-ray photoelectron spectroscopy utilizing a Thermo Fisher scientific with monochromatic X-rays of Al Kα ~ −10 to 1350 eV radiation at a size of 400 μm at a pressure of 10−9 mbar. A full-pass energy spectrum of 200 eV and in a narrow 50 eV band was applied. The binding energy of the transverse carbon line (C 1 s) was used for calibration, and the positions of the peaks were corrected relative to the C 1 s signal position.

4.4. Electrochemical Measurements

Electrochemical CO2 reduction was performed using CH instruments 760E potentiostat equipped with a three-electrode system at room temperature in CO2-saturated 0.1 M KHCO3 (pH = 6.8). Before conducting the test, CO2 was continuously fed into the cathode compartment (20 sccm, 30 min) to form CO2-saturated 0.1 M KHCO3. The Pt electrode was used as a counter electrode and Ag/AgCl electrode as the reference electrode. An H-type device made up of two compartment cells, separated by a Nafion−117 proton-exchange membrane was used to separate the cathode and anode, as well as prevent diffusion of products. Linear sweep voltammetry (LSV) scans were carried out in CO2 saturated 0.1 M KHCO3 solution at a scan rate of 50 mV s−1. A total of 5 mg of Zn, Ni, and Cu spinel power were dispersed into a solution containing 500 μL isopropanol and 500 μL deionized water and 40 μL of Nafion solution (5%); subsequently, the ink was dripped onto carbon paper (5 cm × 1 cm). The electrolyte solution was saturated with CO2 before the tests and the potential was converted to RHE according to following equation.
E (V vs. RHE) = E (V vs. Ag/AgCl) + 0.197 V + 0.0591 × pH
The electrochemical active surface area (ECSA)-normalized current density for as-prepared catalysts was calculated by:
ECSA-normalized current density = current density × Cs/Cdl
where Cs is the specific capacitance. In this work, 0.04 mF cm−2 was adopted as the value of Cs based on previously reported electrocatalysts [36,37].
The i-t curve was measured at different potentials and the products were analyzed using gas chromatography (GC) and NMR. The quantification of gas products (H2 and CO) was obtained from the peak areas using the standard calibration curves. The gas samples were analyzed, using ruimin gas chromatographic instrument (GC) equipped with a thermal conductivity detector. The liquid products were analyzed by using proton nuclear magnetic resonance (Bruker Avance 600 MHz spectrometer) with dimethyl sulfoxide (DMSO) as an internal standard. In each of the NMR measurements, 0.6 mL of product solution (30 mL of the electrolyte was mixed with 2 μL dimethyl sulfoxide (DMSO) as an internal standard) and 0.06 mL deuterated water (D2O, to lock the magnetic field) was added. The one dimensional 1H spectrum was measured with water suppression.

5. Conclusions

In this study, nanosized Cu, Ni, and Zn cobaltites were synthesized by a simple hydrothermal method under autogenous pressure and subsequent calcination at 400 °C in air. The results have shown that the synthesis of spinels was successful confirmed by XRD results. The morphology and particle sizes of the spinels were investigated by using HRTEM analysis. The current density of the mesoporous NiCo2O4 is significantly higher than that of CuCo2O4 and ZnCo2O4 spinels over the whole potential range. It is indicative of Cu, and Zn spinel is slightly less active in terms of its overpotential to reduce CO2. Hydrogen was produced as the main product at the whole potential, along with other products, such as CO and HCOOH. The synthesized Ni and Zn spinel were relatively more stable than the Cu spinel, maintained their excellent activity, and delivered a durable catalyst lifetime. Information related to the structure-reactivity relationship has also been provided based on the available characterization and electrochemical data. The proper understanding of higher selectivity toward hydrogen and/or poorer selectivity of carbon-based products will allow us to develop a future generation of effective spinel-derived catalysts, which could have immense application in the distant future.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/catal12080893/s1, Figure S1: XPS survey spectrum of (a) NiCo2O4, (b) CuCo2O4, (c) ZnCo2O4; Figure S2: Deconvoluted X-ray photoelectron spectra for CuCo2O4, NiCo2O4 and ZnCo2O4 before reaction; Figure S3: Ni spinel was evaluated at different times in CO2-saturated 0.1 M KHCO3 and reproduced independently by two different systems; Figure S4: Cu spinel was evaluated at different times in CO2-saturated 0.1 M KHCO3 and reproduced independently by two different systems; Figure S5: Zn spinel was evaluated at different times in CO2-saturated 0.1 M KHCO3 and reproduced independently by two different systems; Figure S6: Cyclic voltammogram (CV) curves of Ni spinel materials in the non-Faradaic potential region recorded at different scan rates (10−50 mV s−1); Figure S7: Cyclic voltammogram (CV) curves of Cu spinel materials in the non-Faradaic potential region recorded at different scan rates (10−50 mV s−1); Figure S8: Cyclic voltammogram (CV) curves of Zn spinel materials in the non-Faradaic potential region recorded at different scan rates (10−50 mV s−1); Figure S9: Nyquist plots of Ni, Zn and Cu spinel in CO2-saturated 0.1 M KHCO3; Figure S10: The FE of H2 for the different catalysts in the potential range of −0.9 to −1.1 V; Figure S11: The FE of C1 product for the different catalysts in the potential range of −0.9 to −1.1 V; Table S1: Summary of the electrocatalysts toward the CO2 reduction reaction. References [30,38,39,40,41,42,43] are cited in the Supplementary Materials.

Author Contributions

Conceptualization, M.M.M.M. and A.D.C.; methodology, W.B. and L.G.; software, L.G.; validation, K.N., M.A.S. and A.A.; formal analysis, N.H.K.; investigation, S.A.-F.; resources, A.A.; data curation, M.A.S.; writing—original draft preparation, K.N.; writing—review and editing, M.M.M.M. and A.D.C.; visualization, L.G.; supervision, A.D.C.; project administration, M.M.M.M.; funding acquisition, M.M.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia, Project number IFPRC-108-130-2020.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Acknowledgments

The authors extend their appreciation to the Deputyship for Research & Innovation, Ministry of Education in Saudi Arabia for funding this research work through project number IFPRC-108-130-2020 and the Deanship of Scientific Research (DSR), King Abdulaziz University, Jeddah, Saudi Arabia. Support from a start-up research grant from Wuhan University (China) and the National Natural Science Foundation of China (NSFC) (Grant No. 22050410276 to A.D.C.) is also acknowledged.

Conflicts of Interest

The authors claim that they do not have any known conflicting financial interests or personal affiliations that may seem to have impacted the work presented in this study.

References

  1. Verma, S.; Lu, X.; Ma, S.; Masel, R.I.; Kenis, P.J.A. The Effect of Electrolyte Composition on the Electroreduction of CO2 to CO on Ag Based Gas Diffusion Electrodes. Phys. Chem. Chem. Phys. 2016, 18, 7075–7084. [Google Scholar] [CrossRef] [PubMed]
  2. Kim, B.; Hillman, F.; Ariyoshi, M.; Fujikawa, S.; Kenis, P.J.A. Effects of Composition of the Micro Porous Layer and the Substrate on Performance in the Electrochemical Reduction of CO2 to CO. J. Power Sources 2016, 312, 192–198. [Google Scholar] [CrossRef]
  3. Xiong, P.; Yang, F.; Ding, Z.; Jia, Y.; Liu, J.; Yan, X.; Chen, X.; Yang, C. Preparation and Electrocatalytic Properties of Spinel CoxFe3-XO4 Nanoparticles. Int. J. Hydrogen Energy 2020, 45, 13841–13847. [Google Scholar] [CrossRef]
  4. Khdary, N.H.; Alayyar, A.S.; Alsarhan, L.M.; Alshihri, S.; Mokhtar, M. Metal Oxides as Catalyst/Supporter for CO2 Capture and Conversion, Review. Catalysts 2022, 12, 300. [Google Scholar] [CrossRef]
  5. Glazer, A.M. Crystallography: A Very Short Introduction; Oxford University Press: Oxford, UK, 2016; ISBN 9780198717591. [Google Scholar]
  6. Singh, S.K.; Dhavale, V.M.; Kurungot, S. Surface-Tuned Co3O4 Nanoparticles Dispersed on Nitrogen-Doped Graphene as an Efficient Cathode Electrocatalyst for Mechanical Rechargeable Zinc-Air Battery Application. ACS Appl. Mater. Interfaces 2015, 7, 21138–21149. [Google Scholar] [CrossRef] [PubMed]
  7. Ramsundar, R.M.; Debgupta, J.; Pillai, V.K.; Joy, P.A. Co3O4 Nanorods—Efficient Non-Noble Metal Electrocatalyst for Oxygen Evolution at Neutral PH. Electrocatalysis 2015, 6, 331–340. [Google Scholar] [CrossRef]
  8. Jin, H.; Wang, J.; Su, D.; Wei, Z.; Pang, Z.; Wang, Y. In-Situ Cobalt-Cobalt Oxide/N-Doped Carbon Hybrids as Superior Bi-Functional Electrocatalysts for Hydrogen and Oxygen Evolution. J. Am. Chem. Soc. 2015, 137, 2688–2694. [Google Scholar] [CrossRef] [PubMed]
  9. Hamdani, M.; Singh, R.N.; Chartier, P. Co3O4 and Co-Based Spinel Oxides Bifunctional Oxygen Electrodes. Int. J. Electrochem. Sci. 2010, 5, 556–577. [Google Scholar]
  10. Wang, J.; Wang, Q.; She, W.; Xie, C.; Zhang, X.; Sun, M.; Xiao, J.; Wang, S. Tuning the Electron Density Distribution of the Co-N-C Catalysts through Guest Molecules and Heteroatom Doping to Boost Oxygen Reduction Activity. J. Power Sources 2019, 418, 50–60. [Google Scholar] [CrossRef]
  11. Raveau, B.; Seikh, M.M. Magnetic and Physical Properties of Cobalt Perovskites. In Handbook of Magnetic Materials; Elsevier B.V.: Amsterdam, The Netherlands, 2015; Volume 23, pp. 161–289. [Google Scholar]
  12. Gao, S.; Lin, Y.; Jiao, X.; Sun, Y.; Luo, Q.; Zhang, W.; Li, D.; Yang, J.; Xie, Y. Partially Oxidized Atomic Cobalt Layers for Carbon Dioxide Electroreduction to Liquid Fuel. Nature 2016, 529, 68–71. [Google Scholar] [CrossRef]
  13. Harris, V.G. Microwave magnetic materials. In Handbook of Magnetic Materials; Elsevier: Amsterdam, The Netherlands, 2012; pp. 1–63. [Google Scholar]
  14. Basahel, S.N.; Abd El-Maksod, I.H.; Abu-Zied, B.M.; Mokhtar, M. Effect of Zr4+ doping on the stabilization of ZnCo-mixed oxide spinel system and its catalytic activity towards N2O decomposition. J. Alloys Compd. 2010, 493, 630–635. [Google Scholar] [CrossRef]
  15. Bruce, P.G. Solid State Electrochemistry; Cambridge University Press: Cambridge, UK, 1995; ISBN 0521400074. [Google Scholar]
  16. Singh, R.-N.; Hamdani, M.; Koenig, J.-F.; Poillerat, G.; Gautier, J.L.; Chartier, P. Thin Films of CO3O4 and NiCo2O4 Obtained by the Method of Chemical Spray Pyrolysis for Electrocatalysis III. The Electrocatalysis of Oxygen Evolution. J. Appl. Electrochem. 1990, 20, 442–446. [Google Scholar] [CrossRef]
  17. Singh, R.N.; Koenig, J.-F.; Poillerat, G.; Chartier, P. Electrochemical Studies on Protective Thin CO3O4 and NiCo2O4 Films Prepared on Titanium by Spray Pyrolysis for Oxygen Evolution. J. Electrochem. Soc. 1990, 137, 1408–1413. [Google Scholar] [CrossRef]
  18. Rashkova, V.; Kitova, S.; Konstantinov, I.; Vitanov, T. Vacuum Evaporated Thin Films of Mixed Cobalt and Nickel Oxides as Electrocatalyst for Oxygen Evolution and Reduction. Electrochim. Acta 2002, 47, 1555–1560. [Google Scholar] [CrossRef]
  19. Zhao, Q.; Yan, Z.; Chen, C.; Chen, J. Spinels: Controlled Preparation, Oxygen Reduction/Evolution Reaction Application, and Beyond. Chem. Rev. 2017, 117, 10121–10211. [Google Scholar] [CrossRef] [PubMed]
  20. Shi, R.; Zhang, Y.; Wang, Z. Facile Synthesis of a ZnCo2O4 Electrocatalyst with Three-Dimensional Architecture for Methanol Oxidation. J. Alloys Compd. 2019, 810, 151879. [Google Scholar] [CrossRef]
  21. Anu Prathap, M.U.; Srivastava, R. Synthesis of NiCo2O4 and Its Application in the Electrocatalytic Oxidation of Methanol. Nano Energy 2013, 2, 1046–1053. [Google Scholar] [CrossRef]
  22. Silambarasan, M.; Padmanathan, N.; Ramesh, P.S.; Geetha, D. Spinel CuCo2O4 Nanoparticles: Facile One-Step Synthesis, Optical, and Electrochemical Properties. Mater. Res. Express 2016, 3, 095021. [Google Scholar] [CrossRef]
  23. Hao, S.; Zhang, B.; Ball, S.; Copley, M.; Xu, Z.; Srinivasan, M.; Zhou, K.; Mhaisalkar, S.; Huang, Y. Synthesis of Multimodal Porous ZnCo2O4 and Its Electrochemical Properties as an Anode Material for Lithium Ion Batteries. J. Power Sources 2015, 294, 112–119. [Google Scholar] [CrossRef]
  24. Li, X.; Jiang, L.; Zhou, C.; Liu, J.; Zeng, H. Integrating Large Specific Surface Area and High Conductivity in Hydrogenated NiCo2O4 Double-Shell Hollow Spheres to Improve Supercapacitors. NPG Asia Mater. 2015, 7, e165. [Google Scholar] [CrossRef]
  25. Lu, L.; Min, F.; Luo, Z.; Wang, S.; Teng, F.; Li, G.; Feng, C. Synthesis and Electrochemical Properties of CuCo2O4 as Anode Material for Lithium-Ion Battery. J. Nanosci. Nanotechnol. 2017, 17, 4763–4771. [Google Scholar] [CrossRef]
  26. Xiao, X.; Wang, G.; Zhang, M.; Wang, Z.; Zhao, R.; Wang, Y. Electrochemical Performance of Mesoporous ZnCo2O4 Nanosheets as an Electrode Material for Supercapacitor. Ionics 2018, 24, 2435–2443. [Google Scholar] [CrossRef]
  27. Wu, H.; Qin, M.; Zhang, L. NiCo2O4 Constructed by Different Dimensions of Building Blocks with Superior Electromagnetic Wave Absorption Performance. Compos. Part B Eng. 2020, 182, 107620. [Google Scholar] [CrossRef]
  28. Vijayakumar, S.; Lee, S.H.; Ryu, K.S. Hierarchical CuCo2O4 Nanobelts as a Supercapacitor Electrode with High Areal and Specific Capacitance. Electrochim. Acta 2015, 182, 979–986. [Google Scholar] [CrossRef]
  29. Zhu, J.; Gao, Q. Mesoporous MCo2O4 (M = Cu, Mn and Ni) Spinels: Structural Replication, Characterization and Catalytic Application in CO Oxidation. Microporous Mesoporous Mater. 2009, 124, 144–152. [Google Scholar] [CrossRef]
  30. Sekar, P.; Calvillo, L.; Tubaro, C.; Baron, M.; Pokle, A.; Carraro, F.; Martucci, A.; Agnoli, S. Cobalt Spinel Nanocubes on N-Doped Graphene: A Synergistic Hybrid Electrocatalyst for the Highly Selective Reduction of Carbon Dioxide to Formic Acid. ACS Catal. 2017, 7, 7695–7703. [Google Scholar] [CrossRef]
  31. Parveen, S.; Nguyen, H.H.; Premkumar, T.; Puschmann, H.; Govindarajan, S. Nano Spinel Cobaltites and Their Catalytic and Electrochemical Properties: Facile Synthesis of Metal (Co, Ni, and Zn) and Mixed Metal (Co–Ni and Co–Zn) Complexes of Schiff Bases Prepared from α-Ketoglutaric Acid and Ethyl Carbazate. New J. Chem. 2020, 44, 12729–12740. [Google Scholar] [CrossRef]
  32. Lu, X.F.; Wu, D.J.; Li, R.Z.; Li, Q.; Ye, S.H.; Tong, Y.X.; Li, G.R. Hierarchical NiCo2O4 Nanosheets@hollow Microrod Arrays for High-Performance Asymmetric Supercapacitors. J. Mater. Chem. A 2014, 2, 4706–4713. [Google Scholar] [CrossRef]
  33. Moulder, J.F.; Stickle, W.F.; Sobol, P.E.; Bomben, K.D.; Chastain, J. Handbook of X-ray Photoelectron Spectroscopy AReference Book of Standard Spectra for Identification and Interpretation of XPS Data; Perkin-Elmer: Waltham, MA, USA, 1979. [Google Scholar]
  34. Hu, L.; Qu, B.; Li, C.; Chen, Y.; Mei, L.; Lei, D.; Chen, L.; Li, Q.; Wang, T. Facile Synthesis of Uniform Mesoporous ZnCo2O4 Microspheres as a High-Performance Anode Material for Li-Ion Batteries. J. Mater. Chem. A 2013, 1, 5596–5602. [Google Scholar] [CrossRef]
  35. Fang, Y.Y.; Wang, X.Z.; Chen, Y.Q.; Dai, L.Y. NiCo2O4 Nanoparticles: An Efficient and Magnetic Catalyst for Knoevenagel Condensation. J. Zhejiang Univ. Sci. A 2020, 21, 74–84. [Google Scholar] [CrossRef]
  36. Zheng, W.; Liu, M.; Lee, L.Y.S. Best Practices in Using Foam-Type Electrodes for Electrocatalytic Performance Benchmark. ACS Energy Lett. 2020, 5, 3260–3264. [Google Scholar] [CrossRef]
  37. Huang, Y.; Jiang, L.W.; Shi, B.Y.; Ryan, K.M.; Wang, J.J. Highly Efficient Oxygen Evolution Reaction Enabled by Phosphorus Doping of the Fe Electronic Structure in Iron–Nickel Selenide Nanosheets. Adv. Sci. 2021, 8, 2101775. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, Q.; He, A.; Dong, W.; Du, J.; Liu, Z.; Tao, C. Co, N Co-Doped Porous Carbon Supported Spinel Co3O4 for Highly Selective Electroreduction of CO2 to Formate. Vacuum 2022, 197, 110803. [Google Scholar] [CrossRef]
  39. Gao, S.; Jiao, X.; Sun, Z.; Zhang, W.; Sun, Y.; Wang, C.; Hu, Q.; Zu, X.; Yang, F.; Yang, S.; et al. Ultrathin Co3O4 Layers Realizing Optimized CO2 Electroreduction to Formate. Angew. Chem. 2016, 128, 708–712. [Google Scholar] [CrossRef]
  40. Yang, Y.; He, A.; Yang, M.; Zou, Q.; Li, H.; Liu, Z.; Du, J. Selective electroreduction of CO2 to ethanol over a highly stable catalyst derived from polyaniline/CuBi2O4. Catal. Sci. Technol. 2021, 11, 5908–5916. [Google Scholar] [CrossRef]
  41. Karim, K.M.R.; Tarek, M.; Sarkar, S.M.; Mouras, R.; Ong, H.R.; Abdullah, H.; Cheng, C.K.; Khan, M.M.R. Photoelectrocatalytic Reduction of CO2 to Methanol over CuFe2O4@PANI Photocathode. Int. J. Hydrogen Energy 2021, 46, 24709–24720. [Google Scholar] [CrossRef]
  42. Tarek, M.; Rezaul Karim, K.M.; Sarkar, S.M.; Deb, A.; Ong, H.R.; Abdullah, H.; Cheng, C.K.; Rahman Khan, M.M. Hetero-Structure CdS–CuFe2O4 as an Efficient Visible Light Active Photocatalyst for Photoelectrochemical Reduction of CO2 to Methanol. Int. J. Hydrogen Energy 2019, 44, 26271–26284. [Google Scholar] [CrossRef]
  43. Simon, C.; Zander, J.; Kottakkat, T.; Weiss, M.; Timm, J.; Roth, C.; Marschall, R. Fast Microwave Synthesis of Phase-Pure Ni2FeS4 Thiospinel Nanosheets for Application in Electrochemical CO2 Reduction. ACS Appl. Energy Mater. 2021, 4, 8702–8708. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of synthesized NiCo2O4, CuCo2O4, and ZnCo2O4.
Figure 1. XRD patterns of synthesized NiCo2O4, CuCo2O4, and ZnCo2O4.
Catalysts 12 00893 g001
Figure 2. HRTEM images with different magnification for NiCo2O4 (a,b), CuCo2O4 (c,d), ZnCo2O4 (e,f).
Figure 2. HRTEM images with different magnification for NiCo2O4 (a,b), CuCo2O4 (c,d), ZnCo2O4 (e,f).
Catalysts 12 00893 g002
Figure 3. (a) N2 adsorption-desorption isotherms for NiCo2O4, CuCo2O4, and ZnCo2O4. (b) Pore size distribution curves for NiCo2O4, CuCo2O4, and ZnCo2O4.
Figure 3. (a) N2 adsorption-desorption isotherms for NiCo2O4, CuCo2O4, and ZnCo2O4. (b) Pore size distribution curves for NiCo2O4, CuCo2O4, and ZnCo2O4.
Catalysts 12 00893 g003
Figure 4. CO2-TPD patterns for NiCo2O4, CuCo2O4 and ZnCo2O4 samples.
Figure 4. CO2-TPD patterns for NiCo2O4, CuCo2O4 and ZnCo2O4 samples.
Catalysts 12 00893 g004
Figure 5. CO2RR performance of Ni, Cu, and Zn spinel: (a) LSV curves, (b) plots showing the extraction of the Cdl for the estimation of the ECSA, (c) ECSA-normalized LSV curves, (df) FE of major products vs. potential for the Ni, Cu, and Zn spinel catalyst at a potential range of −0.9 V to −1.1 V in 0.1 M KHCO3, (g) H2, (h) CO, and (i) formate current density plots as a function of potential (vs. RHE) during CO2 flow rate of 20 sccm, (jl) stability test of Ni, Cu, and Zn spinel, Left y-axis: the current density (j) of Ni, Cu, and Zn spinel at −0.9 V vs. RHE. Right y-axis: the faradaic efficiency of H2, HCOOH, and CO.
Figure 5. CO2RR performance of Ni, Cu, and Zn spinel: (a) LSV curves, (b) plots showing the extraction of the Cdl for the estimation of the ECSA, (c) ECSA-normalized LSV curves, (df) FE of major products vs. potential for the Ni, Cu, and Zn spinel catalyst at a potential range of −0.9 V to −1.1 V in 0.1 M KHCO3, (g) H2, (h) CO, and (i) formate current density plots as a function of potential (vs. RHE) during CO2 flow rate of 20 sccm, (jl) stability test of Ni, Cu, and Zn spinel, Left y-axis: the current density (j) of Ni, Cu, and Zn spinel at −0.9 V vs. RHE. Right y-axis: the faradaic efficiency of H2, HCOOH, and CO.
Catalysts 12 00893 g005
Figure 6. XPS spectra of spinels (cf. both before and after the reaction): (a) Zn and Co 2p spectra of Zn spinel; (b) Ni 2p and Co 2p spectra of Ni spinel; (c) Cu 2p and Co 2p spectra of Cu spinel.
Figure 6. XPS spectra of spinels (cf. both before and after the reaction): (a) Zn and Co 2p spectra of Zn spinel; (b) Ni 2p and Co 2p spectra of Ni spinel; (c) Cu 2p and Co 2p spectra of Cu spinel.
Catalysts 12 00893 g006
Table 1. Textural properties of different spinels from N2- physisorption at −196 °C.
Table 1. Textural properties of different spinels from N2- physisorption at −196 °C.
SampleSBET
(m2/g)
Smicro
(m2/g)
Smeso
(m2/g)
Vtotal
(cc/g)
Vmicro
(cc/g)
Vmeso
(cc/g)
Average Pore Size (nm)Pore Radius (nm)C-ConstantHF
NiCo2O4410350.27260.0000.26138.745.70.000
CuCo2O42811120.03520.0070.02331.7121.60.088
ZnCo2O43212160.04070.0070.02731.7100.50.086
Table 2. Data obtained from CO2-TPD measurements for the investigated samples.
Table 2. Data obtained from CO2-TPD measurements for the investigated samples.
CatalystTemp. (°C)CO2 Uptake (μmol/g)Total CO2 Uptake (μmol/g)
NiCo2O4585.5646.61089.1
652442.5
CuCo2O4686.2626.2626.4
ZnCo2O4620.8152697.7
720545.8
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mostafa, M.M.M.; Bajafar, W.; Gu, L.; Narasimharao, K.; Abdel Salam, M.; Alshehri, A.; Khdary, N.H.; Al-Faifi, S.; Chowdhury, A.D. Electrochemical Characteristics of Nanosized Cu, Ni, and Zn Cobaltite Spinel Materials. Catalysts 2022, 12, 893. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12080893

AMA Style

Mostafa MMM, Bajafar W, Gu L, Narasimharao K, Abdel Salam M, Alshehri A, Khdary NH, Al-Faifi S, Chowdhury AD. Electrochemical Characteristics of Nanosized Cu, Ni, and Zn Cobaltite Spinel Materials. Catalysts. 2022; 12(8):893. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12080893

Chicago/Turabian Style

Mostafa, Mohamed Mokhtar M., Wejdan Bajafar, Lin Gu, Katabathini Narasimharao, Mohamed Abdel Salam, Abdulmohsen Alshehri, Nezar H. Khdary, Sulaiman Al-Faifi, and Abhishek Dutta Chowdhury. 2022. "Electrochemical Characteristics of Nanosized Cu, Ni, and Zn Cobaltite Spinel Materials" Catalysts 12, no. 8: 893. https://0-doi-org.brum.beds.ac.uk/10.3390/catal12080893

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