Next Article in Journal
Light-Assisted Catalysis in Water and Indoor Air Cleaning: Challenges and Perspectives
Next Article in Special Issue
Mechanism of Methanol Synthesis from CO2 Hydrogenation over Cu/γ-Al2O3 Interface: Influences of Surface Hydroxylation
Previous Article in Journal
Conversion of CO2 Hydrogenation to Methanol over K/Ni Promoted MoS2/MgO Catalyst
Previous Article in Special Issue
Tailoring of Three-Atom Metal Cluster Catalysts for Ammonia Synthesis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimized Ni, Co, Mn Oxides Anchored on Graphite Plates for Highly Efficient Overall Water Splitting

1
Key Laboratory of Carbon Materials of Zhejiang Province, Wenzhou Key Lab of Advanced Energy Storage and Conversion, Zhejiang Province Key Lab of Leather Engineering, College of Chemistry and Materials Engineering, Wenzhou University, Wenzhou 325035, China
2
Printable Electronics Research Center & i-Lab, Suzhou Institute of Nano-Tech and Nano-Bionics, Chinese Academy of Sciences, Suzhou 215123, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(7), 1031; https://doi.org/10.3390/catal13071031
Submission received: 22 May 2023 / Revised: 15 June 2023 / Accepted: 19 June 2023 / Published: 23 June 2023
(This article belongs to the Special Issue Theory-Guided Electrocatalysis and Photocatalysis)

Abstract

:
Nickel, cobalt, and manganese oxides are easily obtainable non-noble metal catalysts for water splitting. However, the relationship between composition and catalysts’ performance still needs systematic studies. Herein, guided by theoretical calculations, a low overpotential, easily prepared Mn-doped Co3O4 was deposited on graphite plates for water splitting. The 30% Mn-doped Co3O4 (Co2.1Mn0.9O4) required the lowest overpotential for oxygen evolution reaction (OER), in which the Co2.1Mn0.9O4 reached 20, 30, and 50 mA cm−2 in the overpotentials of 425, 451, and 487 mV, respectively, with 90% IR compensation. Under overall water-splitting conditions, the current density reached 30 mA cm−2 at an overpotential of 0.78 V without IR compensation. Charge density difference analysis illustrates that doped Mn provides electrons for O atoms, and that Mn doping also promotes the electron fluctuation of Co atoms. XPS analysis reveals that Mn-doping increases the chemical valence of the Co atom, and that the doped Mn atom also exhibits higher chemical valence than the Mn of Mn3O4, which is advantageous to boost the form of based-OOH* radical, then decrease the overpotential. Considering the particular simplicity of growing the Co2.1Mn0.9O4 on graphite plates, this work is expected to provide a feasible way to develop the high-performance Co-Mn bimetallic oxide for water splitting.

Graphical Abstract

1. Introduction

In contemporary study giving equal importance to energy and the environment, electrolytic water has emerged as a critical avenue for the effective utilization of clean energy resources [1,2]. Noble metal catalysts containing Ru, Rh, Pd, Os, Ir, and Pt exhibit excellent catalyst performance. However, due to resource scarcity and high price, their large-scale applications are limited; therefore, it is of great importance to develop non-noble metal catalysts [3,4]. Numerous materials are nascent non-noble catalysts for water splitting, i.e., high-entropy materials, transition metal phosphides/oxides/sulfides/selenides/tellurides, perovskite composites such as LaSr3−y, LaSr2.7Co1.5Fe1.5O10, and their composites [5,6,7,8,9,10,11]. Transition metal oxides have been extensively studied in recent years due to their easy preparation and low cost, especially Co, Ni, and Mn oxides [12,13,14,15,16]. For instance, Zhang et al. deposited Co/CoO on N-doped graphene, and as-prepared Co/CoO/N-G achieved 10 mA cm−2 at an overpotential of 315 mV with 80% IR compensation [17]. Yang et al. deposited graphene sheets on carbon cloth and then deposited NiO on the graphene sheets to obtain the catalytic electrode. The prepared electrode achieved 10 mA cm−2 at an overpotential of 425 mV (did not clarify whether IR compensation was utilized) [18]. Zhang et al. introduced oxygen vacancies into the surface of MnO2 nanorods, and the prepared MnO2 achieved a current density of 5 mA cm−2 at an overpotential of 410 mV with 90% IR compensation [19]. However, monometallic oxides are still susceptible to corrosion by alkaline electrolytes and insufficient catalyst activity [20,21]. Doping by other elements is a commonly used method to enhance material stability and performance, which has been confirmed in many research fields. For example, in the field of energy storage, Poudel et al. reported that 3.0 at.% Mn-doped WS2 delivered 107.79 mAh g−1 for an asymmetric supercapacitor at the current density of 1.0 A g−1, while the WS2 only delivered 37.52 mAh g−1 in the same condition [22]. Furthermore, the 2.0 at.% Gd-doped MnO2 improved the capacity from 49.5 mAh g−1 to 82.1 mAh g−1, and the improved amplitude reached 65.8% [23]. It is believed that doping by other elements has introduced new active centers and regulated the electronic structure of the original catalysts, thus soundly improving the catalyst activity. However, for a surface-doped catalyst, the catalyst activity will decrease dramatically after the surface layer is corroded.
The bimetallic material is equivalent to a uniform solid solution, in which the activity of the catalyst does not decrease significantly after the surface layer is corroded. Recent studies have suggested that bimetallic compounds/composites with appropriate elemental ratios exhibit similar properties to doped materials. For example, Tang et al. discovered the OER performance of CoδMn2−δCH (CH: carbonate hydroxide) following the order of Co1Mn1CH > Co1.5Mn0.5CH > CoCH > Co0.5Mn1.5CH in 1 mol/L KOH electrolyte. This indicates that a small amount of Mn-doping CoCH can enhance the OER performance, but excessive doping will lead to inferior OER performance [24]. Similarly, Nai et al. investigated the OER performance of Ni-Co bimetallic hydroxides and found that the polarization voltage followed the order of Ni(OH)2 > Ni2Co(OH)x > NiCo(OH)x > Co(OH)2 > NiCo8.3(OH)x > NiCo5.7(OH)x > NiCo2.7(OH)x. This suggests that a small amount of Ni substituting Co(OH)2 can reduce the overpotential, and the optimal performance is achieved when approximately 1/4 Co atoms are substituted by Ni. However, substituting more than 50% of the Co atoms will increase the overpotential [25]. Based on the above studies, it is reasonable to assume that the utilization of Ni, Co, and Mn bimetallic oxides could be a promising approach for enhancing water electrolysis performance. However, to the best of our knowledge, a systematic, comprehensive, and in-depth investigation of the Ni, Co, and Mn bimetallic oxides for water splitting is still lacking.
In this work, theoretical calculations were conducted on Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, and Co1−xMnxO (x = 0, 0.2, 0.4, 0.6, 0.8) to predict the catalytic activity variation as the elements proportion change. Calculations results predicted that the Ni-doping MnO, Co-doping MnO, Ni-doping Co3O4, and Mn-doping Co3O4 may exhibit lower overpotential. Usually, catalysts need to deposit on substrates for application. Recent reports indicate that graphite is a practical catalyst substrate due to its high electronic conductivity and low cost, and its low catalytic activity does not obscure the catalyst’s performance [26,27,28,29]. Herein, the graphite plates were applied as the support to investigate the catalytic activity of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, and Co1−xMnxO. The experiment results illustrated that Co2.1Mn0.9O4 is most favorable for water splitting, which was consistent with the theoretical calculation predicted results.

2. Results and Discussion

2.1. DFT Calculation Results

Before conducting the experiments, band structure and density of states (DOS) of the Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, and Co1−xMnxO were calculated (the results are shown in Figures S1 and S2, and Table S1). A smaller band gap value indicates better conductivity of the material and lower overpotential when the material is employed as a catalyst for water splitting [30,31,32]. It is worth noting that the band gap values obtained from theoretical calculations are significantly lower than those obtained from experiments, primarily due to the perfect crystal structures used for theoretical calculations, whereas materials tested in experiments contain crystallographic defects, such as impurities, atomic vacancies, and surface contamination. Despite the significant numerical discrepancy between the theoretical calculations and experiments, it does not impede the analysis of the variation of the materials’ properties. According to Figure 1a and Table S1, the band gap values of MnO, CoO, and NiO are 0, 0.157, and 0.236 eV, respectively. This suggests that the electronic conductivity of MnO is higher than that of CoO, and CoO also exhibits higher electronic conductivity than NiO. Furthermore, the band gap of Ni0.2Mn0.8O and Co0.2Mn0.8O is 0 eV, which proves that 20% Ni or Co substituted MnO maintains its original conductivity. Moreover, the electronic conductivity of Ni0.8Co0.2O is lower than that of NiO, which also illustrates that a small amount of Co doping NiO improved the electronic conductivity. However, if the doping amount exceeds a certain limit, the band gap will increase, i.e., Ni0.4Mn0.6O and Co0.4Mn0.6O show a higher band gap than MnO, and Ni0.6Co0.4O shows a higher band gap than NiO.
The d-band center energy calculated from DOS was listed in Table S2 and plotted in Figure 1b. According to the d-band center theory, the closer the d-band center value to the Fermi level, the stronger adsorption of adsorbate on the exposed surface of the adsorbent [33,34]. In the case of water electrolysis, the high adsorption energy resulted in higher overpotential, which originated from the challenge of electronically insulated O2/H2 desorption from the adsorbent. For monometallic oxides, the d-band center of MnO, Co3O4, NiO, Mn3O4, and CoO are 5.0143, 4.9905, 3.7307, 3.2971, 2.0131 eV, respectively, which implies that the catalysts’ activity of those oxides abides the rule of MnO ≈ Co3O4 > NiO > Mn3O4 > CoO. The d-band center of Mn-doped CoO moves away from the Fermi level with the Mn concentration increasing, suggesting a gradual decrease in the overpotential with the increase of Mn concentration. For Co-doped NiO, the d-band center moves closer to the Fermi level upon increasing Co concentration, indicating an increase in the overpotential with the increase of Co concentration. Similarly, for Ni-doped MnO, the d-band center moves closer to the Fermi level with increasing Ni concentration, indicating an increase in the overpotential with an increase in the Ni concentration. Notice that the d-band center value of MnO and Co3O4 is far away from the Fermi level and the difference is negligible. Combining the high electronic conductive property of MnO and Co3O4, we can conclude that MnO-based and Co3O4-based oxides will exhibit higher catalyst activity than other oxides; in other words, a small amount of Ni-doped MnO, Co-doped MnO, Ni-doped Co3O4, and Mn-doped Co3O4 will show lower overpotential for water splitting.

2.2. Electrode Preparation and Characterization

Guided by the theoretical calculations, graphite plates were wetted by the solutions containing Co2+/Ni2+, Co2+/Mn2+, or Ni2+/Mn2+, then dried and heated at 300 °C for 2 h to deposit the catalyst on the surface. Figure 2a schematizes the experimental process. To simplify the discussion, the samples were named in the form of M1M2-n, where M1 and M2 represent the metal cations used in the experiment, and n represents the proportion of M2 in the M1/M2 solution. For example, CoNi-1 represents the electrode obtained using a Co2+/Ni2+ solution where Co accounts for 90% and Ni accounts for 10%. Similarly, CoNi-2 represents the electrode obtained using a Co2+/Ni2+ solution where Co accounts for 80% and Ni accounts for 20%, and so on. The solution was dried, then the obtained powder was heated at 300 °C for 2 h to test the XRD, considering that, if the electrodes were tested directly, the catalyst’s peak may be obscured by the diffraction peaks of graphite, resulting in difficult identification.
In the case of CoNi-n (Figure 2b), when the Ni content was between 10% and 40%, the diffraction peaks matched Co3O4 (PDF#74-1657), implying that the powder obtained was Ni-doped Co3O4. At a Ni content ranging from 50% to 60%, the XRD patterns matched NiO (PDF #78-0423) and Co3O4. The actual powder may be a mixture of Co-doped NiO and Ni-doped Co3O4. At a Ni content of 70% to 90%, the XRD peaks matched NiO and metallic Ni. The actual yield powder may be the mixture of Ni and Co-doped NiO. For CoMn-n (Figure 2c), when the Mn content was 90%, the XRD peaks matched Mn3O4 (PDF#80-0382), Co3O4, and MnO (PDF#75-0626); thus, the actual yield powder may be the mixture of Co-doped Mn3O4, Mn-doped Co3O4, and Co-doped MnO. When the Mn decreased to 50~80%, the XRD peaks matched Mn3O4 and Co3O4, which implies that the MnO or Co-doped MnO was erased. When the Mn content decreased to 10~40%, the XRD peaks matched Co3O4; thus, the yield powder was Mn-doped Co3O4. In the case of NiMn-n (Figure 2d), both NiO and metallic Ni were detected when the Ni content was between 70% and 90%; thus, the yield powder should be the mixture of Mn-doped NiO and Ni. When the Ni content decreased to 20~60%, the XRD peaks matched NiO and Mn3O4; therefore, the yield powder may be the mixture of Mn-doped NiO and Ni-doped Mn3O4. At a Ni content of 10%, the XRD peaks matched Mn3O4 and MnO; thus, the yield powder may be a mixture of Ni-doped Mn3O4 and Ni-doped MnO.
Notice that Ni/NiO, Co3O4, and Mn3O4 were obtained by heating Ni(CH3COO)2·4H2O, Co(CH3COO)2·4H2O, and Mn(CH3COO)2·4H2O, respectively (Figure S3). In this case, the MnO cannot be synthesized by the experimental method of this work. According to Figure 2c,d, the single phase of Ni-doped MnO and Co-doped MnO failed to synthesize, while the Ni-doped Co3O4 and Mn-doped Co3O4 were successfully synthesized. According to Figure 2b,c, to successfully fabricate the Ni-doped Co3O4 and Mn-doped Co3O4, the concentration of Ni or Mn should be in the range of 0~40%.
Subsequently, electrolysis water tests were performed, and the linear sweep voltammetry (LSV) curves without IR compensation for CoNi-n, CoMn-n, and NiMn-n are shown in Figure 3a, Figure 3b and Figure 3c, respectively. In CoNi-n, CoNi-6 exhibited the lowest overpotential compared to others. In CoMn-n, CoMn-3 showed the lowest overpotential, and in NiMn-n, NiMn-3 showed the lowest overpotential. Due to the smaller difference in polarization voltage of those electrodes, more in-depth experimental research is still needed, especially for CoNi-n and CoMn-n. Based on the XRD analysis presented in Figure 2, CoNi-6 may be the mixture of Co-doped NiO and Ni-doped Co3O4, CoMn-3 was Co2.1Mn0.9O4, and NiMn-3 may be the mixture of Mn-doped NiO and Ni-doped Mn3O4. To further investigate the catalytic activity of CoNi-6, CoMn-3, and NiMn-3, the LSV of OER with 90% IR compensation was tested and the results are shown in Figure 3d. We can see that CoMn-3 showed the lowest overpotential compared to CoNi-6 and NiMn-3. The LSV curve of CoMn-3 was also compared with Ni/NiO, Co3O4, and Mn3O4, and CoMn-3 still showed the lowest overpotential (Figure S4). It should be mentioned that the electronegativity of Co, Ni, and Mn are 1.91, 1.88, and 1.55, respectively. The higher the electronegativity, the stronger the ability to pull electrons. Recent literature proves that the HER is a two-step process, and OER is a four-step process, the overpotential of OER is higher than that of HER, and the forming of based-OOH* radical from based-O* radical is the rate-determining step (reaction equation: based-O* + OH → based-OOH*) [3,6,21]. The higher electronegativity of the reactive site, the easier it is to capture the electron from OH, which makes easier the formation of the based-OOH* radical. It implies that Co3O4 will exhibit lower OER overpotential than NiO, and NiO will exhibit lower OER overpotential than Mn3O4. This conclusion has also been verified by the experiment (Figure S4). Herein, the Gibbs free energy of the four elementary reactions of the OER on Co3O4, Mn3O4, and CoMn-3 was calculated, and results are shown in Figures S5–S7. It can be seen that the calculated overpotential of OER on Co3O4, Mn3O4, Co-site of CoMn-3, and Mn-site of CoMn-3 are 0.753, 0.806, 0.594, and 0.619 V, respectively, which further proves that the OER on Co3O4 exhibits lower overpotential than on Mn3O4, and that the Co-site of CoMn-3 has more activity than the Mn-site. As for the Ni-doped Co3O4 or Mn-doped Co3O4, due to their similar molecular formula (Co3−γMγO4, M = Ni, Mn, 0 ≤ γ ≤ 0.40), Co3−γMnγO4 has more empty electron orbitals to accept external electrons; therefore, it is easier to transform based-O* to based-OOH*, and lower overpotential is required.
According to Figure 3d, the current densities of CoNi-6, CoMn-3, and NiMn-3 reached 10 mA cm−2 at an overpotential of 378, 374, and 388 mV, respectively, and reached 100 mA cm−2 at an overpotential of 477, 472, and 550 mV, respectively. The polarization voltage difference between CoMn-3 and CoNi-6 is within the measurement error range, so we are not entirely certain that CoMn-3 is better than CoNi-6. After this work, more research still needs to be done to distinguish the catalyst activity of those materials. In this work, we discuss and analyze the catalytic activity of those materials based on the measured data, as the CoMn-n and CoNi-n still need more in-depth experimental research. Figure 3e illustrates the turnover frequency (TOF) curves of CoNi-6, CoMn-3, and NiMn-3; the higher the TOF values, the higher the OER activity of the material. At an overpotential of 300, 350, 400, 450, and 500 mV, the TOF values of CoMn-3 were 0.09896 × 10−4, 0.20659 × 10−4, 0.6773 × 10−4, 2.25077 × 10−4, and 5.49996 × 10−4 s−1, respectively, and the corresponding TOF values for CoNi-6 were 0.05932 × 10−4, 0.13946 × 10−4, 0.51648 × 10−4, 1.63669 × 10−4, and 4.19032 × 10−4 s−1, respectively. Accordingly, CoMn-3 exhibited superior catalytic activity to CoNi-6, especially at high current densities. In contrast, NiMn-3 exhibited the lowest catalytic activity, demonstrating the inferior OER activity of NiMn-3. To assess the practical application property of the electrodes, current–voltage (i-t) tests without IR compensation were conducted at an initial current density of 10 mA cm−2. Based on results shown in Figure 3f, CoMn-3 showed better stability than CoNi-6 and NiMn-3 according to the i-t curves. Furthermore, LSV curves without IR compensation were carried out on the electrodes that had undergone a 55 h i-t test and compared with the LSV curve before going through the i-t test. As shown in Figure 3g, after the i-t test, the LSV curves shifted to the higher voltage. The LSV curve of CoMn-3 only slightly shifted, NiMn-3 showed a relatively larger shift amplitude, and CoNi-6 showed the largest shift amplitude. These results demonstrate that CoMn-3 is more stable than CoNi-6 and NiMn-3 for OER.
To further confirm the difference between CoNi-6, CoMn-3, and NiMn-3, the HER performance was evaluated with 90% IR compensation, and the LSV curves are presented in Figure 3h. CoMn-3 manifested a significantly lower polarization than CoNi-6, while NiMn-3 displayed the highest voltage polarization. CoMn-3 achieved a current density of 100 mA cm−2 with a 437 mV overpotential, whereas CoNi-6 required an overpotential of 471 mV, and NiMn-3 required an overpotential of 529 mV to reach the same current density. After the electrodes underwent a 55 h i-t test of OER, the LSV curve of HER without IR compensation was tested and is shown in Figure 3i. The LSV curve without going through the i-t test is also shown for comparison. It can be seen that CoMn-3 exhibited a high degree of consistency with the initial LSV curve, implying excellent stability of CoMn-3 in alkaline electrolytes, while CoNi-6 and NiMn-3 showed a remarkable voltage shift to higher overpotential.
To reveal the discrepancy between CoNi-n, CoMn-n, and NiMn-n, scanning electron microscopy (SEM) was adopted to reveal the morphology of those electrodes. As shown in Figure S8, CoNi-n were composed of flake-like sheets; the scale of the sheets was 0.1~2 μm. Figure S9 shows that CoMn-1, CoMn-2, CoMn-5, CoMn-6, CoMn-8, and CoMn-9 were composed of flake-like sheets, while CoMn-3 was composed of nanodots, CoMn-4 was composed of nanodots deposed on sheets, and CoMn-7 was composed of flocculent-like nanodots. In the condition of NiMn-n (Figure S10), except for CoMn-4, the others were composed of micrometer fiber. The higher-resolution images of CoNi-6, CoMn-3, and NiMn-3 are shown in Figure 4a–c. CoNi-6 exhibited a flake-like structure of less than 1 μm with a relatively uniform size. CoMn-3 exhibited spherical particles with a diameter of 100~200 nm, and NiMn-3 exhibited fibers with a diameter of 100~500 nm. Ni/NiO, Co3O4, and Mn3O4 deposed on graphite plates were also investigated the morphology (Figure S11), whereby the Ni/NiO was composed of sheets, Co3O4 was nanodots, and Mn3O4 was micrometer fiber. The CoMn-3 and Co3O4 exhibited similar structure and morphology, and the NiMn-3 and Mn3O4 exhibited similar structure and morphology, which proves that the small amount of Mn-doping Co3O4 or Ni-doping Mn3O4 would not devastate the structure of Co3O4 and Mn3O4.
The X-ray photoelectron spectroscopy (XPS) survey of CoNi-6, CoMn-3, and NiMn-3 is shown in Figure 4d. Apart from C and O elements, Ni and Co were detected in CoNi-6, Ni and Mn were detected in NiMn-3, and Co and Mn were detected in CoMn-3. The C 1s peak can be fitted to the C-C peak at 284.80 eV and the C-O/C=O peak at 285.1–285.3 eV (Figure 4e) [35]. Likewise, the O 1s peak can be deconvoluted to the O-Co/O-Ni, O-Co/O-Mn, O-Ni/O-Mn peaks at 529.3–529.8 eV and H-O-H bond for the residual water at 531.4–531.7 eV (Figure 4f) [12,36,37]. The Co 2p peak of CoNi-6 can be resolved into two peaks located at 779.85 eV and 795.02 eV for Co-Ni 2p3/2 and Co-Ni 2p1/2, respectively, and two other peaks located at 781.16 eV and 796.67 eV for Co2+ 2p3/2 and Co2+ 2p1/2, respectively. Three satellite peaks were observed at 785.40 eV, 789.46 eV, and 803.98 eV (Figure 4g) [38]. In addition, in 773.15 eV, a weak, blurry peak was also observed, which may have originated from the interaction of Co3O4 and NiO. Similarly, the Co 2p peak of CoMn-3 can be deconvoluted into two peaks located at 779.96 eV and 794.96 eV for Co-Mn 2p3/2 and Co-Mn 2p1/2, and two other peaks located at 781.07 eV and 796.16 eV for Co2+ 2p3/2 and Co2+ 2p1/2, respectively. Three satellite peaks were observed at 786.22 eV, 790.02 eV, and 803.12 eV for CoMn-3 (Figure 4g) [39]. The Ni 2p peak of CoNi-6 can be resolved into peaks of Ni-Co 2p3/2 at 856.07 eV and Ni-Co 2p1/2 at 873.26 eV, as well as Ni2+ 2p3/2 at 861.67 eV and Ni2+ 2p1/2 at 880.08 eV. Additionally, three peaks at 854.35 eV, 864.63 eV, and 871.47 eV are satellite peaks (Figure 4h) [38,40]. The Ni 2p peak of NiMn-3 can be deconvoluted into Ni-Mn 2p3/2 at 855.76 eV and Ni-Mn 2p1/2 at 872.61 eV, as well as Ni2+ 2p3/2 at 861.34 eV and Ni2+ 2p1/2 at 879.83 eV. The peaks located at 854.44 eV and 866.11 eV are two satellite peaks (Figure 4h) [41]. Finally, the Mn 2p peak of CoMn-3 was analyzed and fitted to the Mn-Co 2p3/2 peak of Mn2+, Mn3+, and Mn4+ located at 641.35 eV, 642.73 eV, 645.26 eV, respectively, and Mn-Co bond of Mn 2p1/2 peak of Mn2+, Mn3+, Mn4+ located at 652.92, 654.00, 655.81 eV, respectively [42]. Similarly, the Mn 2p peak of NiMn-3 was analyzed and fitted to the Mn-Ni bond of Mn 2p3/2 peak of Mn2+, Mn3+, Mn4+ located at 641.24 eV, 642.59 eV, 644.71 eV, respectively, and the Mn-Co bond of Mn 2p1/2 peak of Mn2+, Mn3+, Mn4+ located at 652.73 eV, 653.90 eV, 655.92 eV, respectively [43]. In addition, a weak peak located at 638.16 eV was also observed, which may have originated from the interaction of Mn3O4 and NiO. Comparing the Co 2p spectra of CoNi-6 and CoMn-3, it can be inferred that CoNi-6 has more components than CoMn-3. Similarly, NiMn-3 has more components than CoMn-3 by comparing the Mn 2p spectra. Combining previous XRD data, it can be confirmed that CoMn-3 is Co2.1Mn0.9O4.
To further analyze the charge redistribution of the CoMn-3, the charge density difference of CoMn-3 was calculated based on DFT, and the results are shown in Figure S12. The result demonstrates that Mn atoms provided electrons to O atoms, and the Mn-doping promoted some Co atoms to gain electrons and some Co atoms to lose electrons. The chemical valence state of Mn and Co atoms was further investigated by XPS (Figure S13). Comparing the Co 2p spectra of Co3O4 and CoMn-3, it can be seen that the binding energy of Co 2p of CoMn-3 shifted to higher energy, proving that the Co site of CoMn-3 lost the electrons and increased the chemical valence, which is an advantage to capture the electron from OH to boost the based-O* transfer to based-OOH* radical. The Mn 2p spectra of CoMn-3 also showed higher binding energy than that of Mn3O4, which is also an advantage to boost the based-O* transfer to based-OOH* radical. Thus, the Mn-doping Co3O4 simultaneously improved the catalytic activity of Co and Mn sites compared to monometallic Co3O4 and Mn3O4.
Due to the best HER/OER performance, the overall water-splitting properties of CoMn-3 without IR compensation were tested in a two-electrode system to evaluate its potential practicality. Figure 5a shows the LSV curve. The current density reached 10 mA cm−2 at 1.78 V and reached 30 mA cm−2 at 2.01 V. At 1.80 V, the current density reached 11.19 mA cm−2, and the graphite plate only contributed the current density of 0.86 mA cm−2; thus, the tested data reflect the catalytic activity of CoMn-3. At 2.0 V, the current density reached 29.30 mA cm−2, and the current density contributed by the graphite plate is still less than 15% (Figure S14). The rate CV curves are shown in Figure S15, accordingly, and the double-layer capacitance was calculated to be 15.86 mF cm−2 (Figure 5b). Considering that the theoretical value of double-layer capacitance for a perfect, defect-free planar electrode is 0.040 mF cm−2, the active surface area of CoMn-3 was estimated to be 396 cm2 (15.86/0.040) in a 1.0 cm2 electrode, which indicates a remarkably high active surface area [44]. Additionally, Figure 5c presents the i-t curve obtained at 1.78 V, revealing that, even after 55 h of testing, CoMn-3 shows acceptable stability. Moreover, Figure 5d depicts the comparison of the LSV curve before and after the i-t test, with the curves almost overlapping, which suggests that CoMn-3 may be a potentially applicable water-splitting catalyst.

3. Materials and Methods

3.1. Theoretical Calculations

The crystal structures of NiO, CoO, MnO, Co3O4, and Mn3O4 were downloaded from the Materials Project website, and all downloaded crystal structures are the lowest energy structures. The oxides of CoxNi1−xO, NixMn1−xO, and Co1−xMnxO, (x = 0, 0.2, 0.4, 0.6, 0.8) were obtained by atomic substitution. By replacing atoms, Co0.8Ni0.2O, Co0.6Ni0.4O, Co0.4Ni0.6O, Co0.2Ni0.8O, Ni0.8Mn0.2O, Ni0.6Mn0.4O, Ni0.4Mn0.6O, Ni0.2Mn0.8O, Co0.8Mn0.2O, Co0.6Mn0.4O, Co0.4Mn0.6O, and Co0.2Mn0.8O were obtained. The academic version of CASTEP was used to calculate the band structure and density of states, and all crystal structures were read and visualized using the VESTA software. The structures were optimized before band structure and density-of-states calculations. The main parameters of theoretical calculations are as follows: GGA-PBE function was used, the energy convergence criterion was set to 1 × 10−5 eV/atom, the force convergence criterion was set to 0.03 eV/Å, the cutoff energy was set to 550 eV, and the band structure and DOS were calculated using the primitive cell. Spin polarization was considered in the calculation of band structure and density of states, and the LDA + U function was used to correct the strong correlation of Ni, Co, and Mn. The center values of the d-bands for both the up and down spins of each system were calculated using the following formula [45,46]:
ε , = l o w h i g h E × P D O S ( E ) d E l o w h i g h P D O S ( E ) d E
In the equation, E is the energy of electrons, and projected DOS (PDOS) is the electronic density of states. Considering the degeneracy of the up and down spin, the average of ε and ε is taken as the evaluation of the catalyst performance of the materials.
The Gibbs energy of the OER and charge density difference of CoMn-3 were calculated on the VASP code. The electron interaction function, K-point set, energy convergence, and force convergence criterion are the same as the band structure calculation. The oxygen evolution energy was calculated based on the following equation:
* + OH → OH* + e, ΔG1 = G(OH*) − G(*) − G(OH)
OH* + OH → O* + H2O + e, ΔG2 = G(O*) + G(H2O) − G(OH*) − G(OH)
O* + OH → OOH* + e, ΔG3 = ΔG(OOH*) − G(O*) − G(OH)
OOH* + OH → O2 + * + H2O + e, ΔG4 = 4.92 − ΔG1 − ΔG2 − ΔG3
where * represents the based, OH* represents the based OH free radical, and so on. The highest energy is the speed-determining step.

3.2. Preparation and Testing of Electrodes

Ni(CH3COO)2·4H2O, Co(CH3COO)2·4H2O, and Mn(CH3COO)2·4H2O were purchased from Aladdin Chemistry Co., Ltd., Shanghai, China, and used directly without further treatment. A 2-millimeter-thick graphite plate was purchased from Dongguan Yongyao Graphite Material Co., Ltd., Dongguan, China. The experimental process was as follows: Prepare 1.0 mol/L aqueous solutions of Ni(CH3COO)2·4H2O, Co(CH3COO)2·4H2O, and Mn(CH3COO)2·4H2O, then immerse the 1.0 × 2.0 cm2 in size of graphite plates into the solution for 30 min. Remove the sucked graphite plates from the solution and dry, and then place them in a tubular furnace. Heat the furnace to 300 °C at a rate of 5 °C/min and hold at 300 °C for 2 h. After the heating is finished and cooled down to room temperature spontaneously, obtain the monometallic oxide growth graphite plates. Next, replace the solution with a mixed solution containing Co2+/Ni2+ to prepare CoNi-n grown graphite plates, the molar ratio of Co2+, Ni2+ changing from 0.8/0.2, 0.6/0.4, 0.4/0.6, and 0.2/0.8, while maintaining the total metal ion concentration at 1.0 mol/L. Prepare the CoMn-n grown graphite plates by using the Co2+/Mn2+ solution and repeat the same protocol, and prepare the NiMn-n grown graphite plates by using the Ni2+/Mn2+ solution by repeating the same protocol. Then use the oxides-grown graphite plates as the working electrode of the electrolyte water.
The Shimadzu-XRD-7000 XRD (Shimadzu, Japan) tester was used to determine the phase structure and composition of the metal oxides. The structure of the electrodes was observed by Nova 200 NanoSEM (FEI Company, Houston, TX, USA) scanning electron microscopes, and the atomic valence state of the metal oxide on the electrode’s surface was investigated using the Axis Ultra DLD X-ray photoelectron spectrometer (Kratos, Japan). The 1.0 mol/L NaOH aqueous solution was used as the electrolyte, and the EC-100B electrochemical workstation (Mesobiosystems Co., Ltd., Wuhan, China) was used to test the linear voltammetry curve (LSV) without IR compensation, current–time curve, and CV, while the CHI660E electrochemical workstation (Shanghai Chenhua Instrument Co., Ltd., Shanghai, China) was used to test the LSV with 90% IR compensation. For the OER and HER tests, the three-electrode system was adopted, wherein a Pt electrode was used as the auxiliary electrode, a saturated Ag/AgCl electrode was used as the reference electrode, and the scan rate for LSV was 10 mV/s. The working electrode voltage was normalized to the hydrogen standard electrode (RHE) voltage, and the normalization formula was [47]:
E(RHE) = E(Ag/AgCl) + 0.05916 × pH +0.1976 (V)
Herein, the pH value was 14; thus,
E(RHE) = E(Ag/AgCl) + 1.0236 (V)
The overpotential calculation formula was:
η = E(RHE) − 1.23 (V)
For the overall water splitting, only a two-electrode system was adopted, in which the as-prepared electrode was used as both the cathode and anode.

4. Conclusions

To summarize, this study employed theoretical calculations to predict the overpotential of CoxNi1−xO, NixMn1−xO, Co1−xMnxO, Co3O4, and Mn3O4 as catalysts for water splitting, indicating that doped MnO or Co3O4 would exhibit lower overpotential. Subsequently, a series of nickel, cobalt, and manganese oxides were deposited on graphite plates and the OER/HER performances were investigated, whereby Co2.1Mn0.9O4 exhibited the highest catalytic activity. In overall water splitting, Co2.1Mn0.9O4 reached 10 mA cm−2 at 1.78 V and reached 30 mA cm−2 at 2.01 V. XPS revealed the chemical valence of the Mn site, and the Co site was increased compared to Mn3O4 and Co3O4, which is an advantage to boost the form of based-OOH* radical. This work once again demonstrates the advantages of theoretical calculations in guiding experimental investigations.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13071031/s1, Figure S1: Band structure Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0, 0.2, 0.4, 0.6, 0.8); Figure S2: DOS of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0, 0.2, 0.4, 0.6, 0.8); Figure S3: (a) XRD pattern of heated Mn(CH3COO)2·4H2O, (b) XRD pattern of heated Ni(CH3COO)2·4H2O, (c) XRD pattern of heated Co(CH3COO)2·4H2O; Figure S4: LSV of CoMn-3, Ni/NiO, Co3O4, and Mn3O4 without IR compensation. Notice that Ni/NiO, Co3O4 and Mn3O4 were deposed on graphite plates by immersing the graphite plates in 1.0 mol/L Ni(CH3COO)2·4H2O, Co(CH3COO)2·4H2O, and Mn(CH3COO)2·4H2O aqueous solution, than dried, and heated at 300 °C for 2 h; Figure S5: the OER process of Co3O4 and Mn3O4 for Gibbs energy calculation; Figure S6: the OER process of CoMn-3 for Gibbs energy calculation; Figure S7: the calculated Gibbs free energy of OER on Co3O4, Mn3O4 and CoMn-3; Figure S8: SEM images of CoNi-n, n = 1~9; Figure S9: SEM images of CoMn-n, n = 1~9; Figure S10: SEM images of NiMn-n, n = 1~9; Figure S11: (a,b) SEM images Ni/NiO, (c,d) SEM images Co3O4, (e,f) SEM images Mn3O4; Figure S12: the charge density difference of CoMn-3; Figure S13: the XPS analysis of Co3O4, Mn3O4 and CoMn-3; Figure S14: Compare the LSV curves of CoMn-3 and graphite plate tested in a two-electrodes system; Figure S15: Rate CV curves of CoMn-3||CoMn-3 symmetric cell; Table S1: Band gap of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0.2, 0.4, 0.6, 0.8); Table S2: d-band center values of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0.2, 0.4, 0.6, 0.8).

Author Contributions

J.L.: Methodology, Investigation, Formal Analysis. Y.D.: Conceptualization, Funding acquisition, Review and editing. H.J.: Funding acquisition. T.Z.: DFT calculation, Analysis, Writing. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Post Doctoral Funding Project of Zhejiang Province (ZJ2022023), National Natural Science Foundation of China (22073069 and 21773082).

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors would like to acknowledge the public testing platform of the Analytical & Testing Center, College of Chemistry and Materials Engineering of Wenzhou University for their help with characterizations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chen, Z.; Yun, S.; Wu, L.; Zhang, J.; Shi, X.; Wei, W.; Liu, Y.; Zheng, R.; Han, N.; Ni, B.-J. Waste-Derived Catalysts for Water Electrolysis: Circular Economy-Driven Sustainable Green Hydrogen Energy. Nano-Micro Lett. 2023, 15, 4. [Google Scholar]
  2. Chatenet, M.; Pollet, B.G.; Dekel, D.R.; Dionigi, F.; Deseure, J.; Millet, P.; Braatz, R.D.; Bazant, M.Z.; Eikerling, M.; Staffell, I.; et al. Water electrolysis: From textbook knowledge to the latest scientific strategies and industrial developments. Chem. Soc. Rev. 2022, 51, 4583–4762. [Google Scholar] [PubMed]
  3. Li, Y.; Sun, Y.; Qin, Y.; Zhang, W.; Wang, L.; Luo, M.; Yang, H.; Guo, S. Recent Advances on Water-Splitting Electrocatalysis Mediated by Noble-Metal-Based Nanostructured Materials. Adv. Energy Mater. 2020, 10, 1903120. [Google Scholar] [CrossRef]
  4. Dinh, T.; Dobo, Z.; Kovacs, H. Phytomining of noble metals—A review. Chemosphere 2022, 286, 131805. [Google Scholar] [PubMed]
  5. Wang, J.; Liao, T.; Wei, Z.; Sun, J.; Guo, J.; Sun, Z. Heteroatom-Doping of Non-Noble Metal-Based Catalysts for Electrocatalytic Hydrogen Evolution: An Electronic Structure Tuning Strategy. Small Methods 2021, 5, 2000988. [Google Scholar]
  6. Wang, C.; Shang, H.; Xu, H.; Du, Y. Nanoboxes endow non-noble-metal-based electrocatalysts with high efficiency for overall water splitting. J. Mater. Chem. A 2021, 9, 857–874. [Google Scholar]
  7. Zhou, Q.; Liao, L.; Zhou, H.; Li, D.; Tang, D.; Yu, F. Innovative strategies in design of transition metal-based catalysts for large-current-density alkaline water/seawater electrolysis. Mater. Today Phys. 2022, 26, 100727. [Google Scholar]
  8. Xu, X.M.; Shao, Z.P.; Jiang, S.P. High-Entropy Materials for Water Electrolysis. Energy Technol. 2022, 10, 2200573. [Google Scholar] [CrossRef]
  9. Xu, X.; Pan, Y.; Ge, L.; Chen, Y.; Mao, X.; Guan, D.; Li, M.; Zhong, Y.; Hu, Z.; Peterson, V.K.; et al. High-Performance Perovskite Composite Electrocatalysts Enabled by Controllable Interface Engineering. Small 2021, 17, 2101573. [Google Scholar]
  10. Tang, J.; Xu, X.; Tang, T.; Zhong, Y.; Shao, Z. Perovskite-Based Electrocatalysts for Cost-Effective Ultrahigh-Current-Density Water Splitting in Anion Exchange Membrane Electrolyzer Cell. Small Methods 2022, 6, 2201099. [Google Scholar] [CrossRef]
  11. Poudel, M.B.; Logeshwaran, N.; Kim, A.R.; Karthikeyan, S.C.; Vijayapradeep, S.; Yoo, D.J. Integrated core-shell assembly of Ni3S2 nanowires and CoMoP nanosheets as highly efficient bifunctional electrocatalysts for overall water splitting. J. Alloys Compd. 2023, 960, 170678. [Google Scholar] [CrossRef]
  12. Dong, Y.; Deng, Z.; Xu, Z.; Liu, G.; Wang, X. Synergistic Tuning of CoO/CoP Heterojunction Nanowire Arrays as Efficient Bifunctional Catalysts for Alkaline Overall Water Splitting. Small Methods 2023, 2300071. [Google Scholar] [CrossRef]
  13. Shah, A.A.; Kumar, M.; Aftab, U.; Abro, M.I. Seawater-Extracted MgO-Doped Co3O4 Composite for Electrochemical Water Splitting. Chem. Eng. Technol. 2023, 46, 459–465. [Google Scholar] [CrossRef]
  14. Sai, K.N.S.; Tang, Y.; Dong, L.; Yu, X.-Y.; Hong, Z. N2 plasma-activated NiO nanosheet arrays with enhanced water splitting performance. Nanotechnology 2020, 31, 455709. [Google Scholar] [CrossRef]
  15. Hu, Q.; Liu, X.; Zhu, B.; Li, G.; Fan, L.; Chai, X.; Zhang, Q.; Liu, J.; He, C. Redox route to ultrathin metal sulfides nanosheet arrays-anchored MnO2 nanoparticles as self-supported electrocatalysts for efficient water splitting. J. Power Sources 2018, 398, 159–166. [Google Scholar] [CrossRef]
  16. Li, X.; Liu, P.F.; Zhang, L.; Zu, M.Y.; Yang, Y.X.; Yang, H.G. Enhancing alkaline hydrogen evolution reaction activity through Ni-Mn3O4 nanocomposites. Chem. Commun. 2016, 52, 10566. [Google Scholar] [CrossRef]
  17. Zhang, W.; Li, C.; Ji, J.-Y.; Niu, Z.; Gu, H.; Abrahams, B.F.; Lang, J.-P. Tailorable carbon cloth electrodes covered with heterostructured Co/CoO/CoN interfaces for scalable electrocatalytic overall water splitting. Chem. Eng. J. 2023, 461, 141937. [Google Scholar]
  18. Yang, W.; Wei, A.; Liu, J.; Xiao, Z.; Zhao, Y.; Zhang, Y. Rational construction of vertical few layer graphene/NiO core-shell nanoflake arrays for efficient oxygen evolution reaction. Mater. Res. Bull. 2021, 139, 111260. [Google Scholar] [CrossRef]
  19. Zhuang, Q.; Ma, N.; Yin, Z.; Yang, X.; Yin, Z.; Gao, J.; Xu, Y.; Gao, Z.; Wang, H.; Kang, J.; et al. Rich Surface Oxygen Vacancies of MnO2 for Enhancing Electrocatalytic Oxygen Reduction and Oxygen Evolution Reactions. Adv. Energy Sustain. Res. 2021, 2, 2100030. [Google Scholar] [CrossRef]
  20. Wang, Q.; Dastafkan, K.; Zhao, C. Design strategies for non-precious metal oxide electrocatalysts for oxygen evolution reactions. Curr. Opin. Electrochem. 2018, 10, 16–23. [Google Scholar] [CrossRef]
  21. Boppella, R.; Tan, J.; Yun, J.; Manorama, S.V.; Moon, J. Anion-mediated transition metal electrocatalysts for efficient water electrolysis: Recent advances and future perspectives. Coord. Chem. Rev. 2021, 427, 213552. [Google Scholar]
  22. Poudel, M.B.; Ojha, G.P.; Kim, A.A.; Kim, H.J. Manganese-doped tungsten disulfide microcones as binder-free electrode for high performance asymmetric supercapacitor. J. Energy Storage 2022, 47, 103674. [Google Scholar] [CrossRef]
  23. Poudel, M.B.; Kim, H.J. Synthesis of high-performance nickel hydroxide nanosheets/gadolinium doped-α-MnO2 composite nanorods as cathode and Fe3O4/GO nanospheres as anode for an all-solid-state asymmetric supercapacitor. J. Energy Chem. 2022, 64, 475–484. [Google Scholar] [CrossRef]
  24. Tang, T.; Jiang, W.-J.; Niu, S.; Liu, N.; Luo, H.; Chen, Y.-Y.; Jin, S.-F.; Gao, F.; Wan, L.-J.; Hu, J.-S. Electronic and Morphological Dual Modulation of Cobalt Carbonate Hydroxides by Mn Doping toward Highly Efficient and Stable Bifunctional Electrocatalysts for Overall Water Splitting. J. Am. Chem. Soc. 2017, 139, 8320–8328. [Google Scholar] [CrossRef] [PubMed]
  25. Nai, J.; Yin, H.; You, T.; Zheng, L.; Zhang, J.; Wang, P.; Jin, Z.; Tian, Y.; Liu, J.; Tang, Z.; et al. Efficient Electrocatalytic Water Oxidation by Using Amorphous Ni-Co Double Hydroxides Nanocages. Adv. Energy Mater. 2015, 5, 1401880. [Google Scholar] [CrossRef]
  26. Liu, X.; Wang, Y.; Tian, P.; Zhou, S.; Gao, H.; Su, J.; Tian, X.; Zang, J. Amorphous NiFe-OH/Ni-Cu-P supported on selfsupporting expanded graphite sheet as efficient bifunctional electrocatalysts for overall water splitting. Int. J. Hydrogen Energ. 2020, 45, 30387–30395. [Google Scholar] [CrossRef]
  27. Gugtapeh, H.S.; Rezaei, M. Facile electrochemical synthesis of Ni-Sb nanostructure supported on graphite as an affordable bifunctional electrocatalyst for hydrogen and oxygen evolution reactions. J. Electroanal. Chem. 2022, 922, 116726. [Google Scholar] [CrossRef]
  28. Sun, H.; Tian, C.; Li, Y.; Wu, J.; Wang, Q.; Yan, Z.; Li, C.-P.; Cheng, F.; Du, M. Coupling NiCo Alloy and CeO2 to Enhance Electrocatalytic Hydrogen Evolution in Alkaline Solution. Adv. Sustain. Syst. 2020, 4, 2000122. [Google Scholar] [CrossRef]
  29. Sun, H.; Tian, C.; Fan, G.; Qi, J.; Liu, Z.; Yan, Z.; Cheng, F.; Chen, J.; Li, C.-P.; Du, M. Boosting Activity on Co4N Porous Nanosheet by Coupling CeO2 for Efficient Electrochemical Overall Water Splitting at High Current Densities. Adv. Funct. Mater. 2020, 30, 1910596. [Google Scholar] [CrossRef]
  30. Zhang, R.; Luo, X.; Zhao, F.; Xu, Q.; Xu, Y.; Xu, Y.; Chen, S.; Fan, X.; Qiao, X. Mono-crystalline Ge1−xSnxSe micro-sheets with hexagonal morphologies for Visible-NIR photodetectors: Increased carrier concentration, narrowed band gap and improved performances. J. Solid State Chem. 2022, 310, 123068. [Google Scholar] [CrossRef]
  31. Saad, A.A.; Lattief, F.A. Test of conductor and semiconductor electrocatalysts in high voltage alkaline electrolyzer as production media for green hydrogen. Int. J. Hydrogen Energ. 2023, 48, 1717–1732. [Google Scholar]
  32. Xuan, C.; Xu, Q.; Han, L.; Hou, B. Electronic structure exquisite modulation of NiSe2 interface via rationally controlling Fe doping for boosting electrochemical oxygen evolution activity. Chem. Eng. J. 2023, 464, 142620. [Google Scholar]
  33. Jiao, S.; Fu, X.; Huang, H. Descriptors for the Evaluation of Electrocatalytic Reactions: D-Band Theory and Beyond. Adv. Funct. Mater. 2022, 32, 2107651. [Google Scholar]
  34. Wang, J.; Xin, S.; Xiao, Y.; Zhang, Z.; Li, Z.; Zhang, W.; Li, C.; Bao, R.; Peng, J.; Yi, J.; et al. Manipulating the Water Dissociation Electrocatalytic Sites of Bimetallic Nickel-Based Alloys for Highly Efficient Alkaline Hydrogen Evolution. Angew. Chem. Int. Ed. 2022, 61, e202202518. [Google Scholar]
  35. Hsiao, M.-C.; Liao, S.-H.; Yen, M.-Y.; Liu, P.-I.; Pu, N.-W.; Wang, C.-A.; Ma, C.-C.M. Preparation of Covalently Functionalized Graphene Using Residual Oxygen-Containing Functional Groups. ACS Appl. Mater. Interfaces 2010, 2, 3092–3099. [Google Scholar] [CrossRef]
  36. Aswathy, N.R.; Varghese, J.; Vinodkumar, R. Effect of annealing temperature on the structural, optical, magnetic and electrochemical properties of NiO thin films prepared by sol-gel spin coating. J. Mater. Sci.-Mater. Electron. 2020, 31, 16634–16648. [Google Scholar] [CrossRef]
  37. Bigiani, L.; Maccato, C.; Barreca, D.; Gasparotto, A. MnO2 nanomaterials functionalized with Ag and SnO2: An XPS study. Surf. Sci. Spectra 2020, 27, 024005. [Google Scholar] [CrossRef]
  38. Lim, T.H.; Cho, S.J.; Yang, H.S.; Engelhard, M.H.; Kim, D.H. Effect of Co/Ni ratios in cobalt nickel mixed oxide catalysts on methane combustion. Appl. Catal. A-Gen. 2015, 505, 62–69. [Google Scholar] [CrossRef] [Green Version]
  39. Lu, Z.; Chen, X.; Liu, P.; Huang, X.; Wei, J.; Ren, Z.; Yao, S.; Fang, Z.; Wang, T.; Masa, J. Co-Mn Hybrid Oxides Supported on N-Doped Graphene as Efficient Electrocatalysts for Reversible Oxygen Electrodes. J. Electrochem. Soc. 2018, 165, H580–H589. [Google Scholar] [CrossRef]
  40. Nassar, M.A.; El-dek, S.I.; Rouby, W.M.A.E.; El-Deen, A.G. Highly efficient asymmetric supercapacitor-based on Ni-Co oxides intercalated graphene as positive and Fe2O3 doped graphene as negative electrodes. J. Energy Storage 2021, 44, 103305. [Google Scholar]
  41. Shan, X.; Guo, F.; Page, K.; Neuefeind, J.C.; Ravel, B.; Abeykoon, A.M.M.; Kwon, G.; Olds, D.; Su, D.; Teng, X. Framework Doping of Ni Enhances Pseudocapacitive Na-Ion Storage of (Ni)MnO2 Layered Birnessite. Chem. Mater. 2019, 31, 8774–8786. [Google Scholar] [CrossRef]
  42. Gao, F.; Tang, X.; Yi, H.; Li, J.; Zhao, S.; Wang, J.; Chu, C.; Li, C. Promotional mechanisms of activity and SO2 tolerance of Co- or Ni-doped MnOx-CeO2 catalysts for SCR of NOx with NH3 at low temperature. Chem. Eng. J. 2017, 317, 20–31. [Google Scholar] [CrossRef]
  43. Dong, Y.; Zhao, J.; Zhang, J.-Y.; Chen, Y.; Yang, X.; Song, W.; Wei, L.; Li, W. Synergy of Mn and Ni enhanced catalytic performance for toluene combustion over Ni-doped α-MnO2 catalysts. Chem. Eng. J. 2020, 388, 124244. [Google Scholar] [CrossRef]
  44. Huang, X.F.; Chang, S.; Wee, S.V.; Ding, J.; Xue, J.M. Three-dimensional printed cellular stainless steel as a high-activity catalytic electrode for oxygen evolution. J. Mater. Chem. A 2017, 5, 18176–18182. [Google Scholar] [CrossRef]
  45. Nørskov, J.K. Chemisorption on metal surfaces. Rep. Prog. Phys. 1990, 53, 1253–1295. [Google Scholar] [CrossRef]
  46. Nørskov, J.K. Electronic factors in catalysis. Prog. Surf. Sci. 1991, 38, 103–144. [Google Scholar] [CrossRef]
  47. Liang, C.; Zou, P.; Nairan, A.; Zhang, Y.; Liu, J.; Liu, K.; Hu, S.; Kang, F.; Fan, H.J.; Yang, C. Exceptional performance of hierarchical Ni-Fe oxyhydroxide@NiFe alloy nanowire array electrocatalysts for large current density water splitting. Energy Environ. Sci. 2020, 13, 86–95. [Google Scholar] [CrossRef]
Figure 1. (a) Band gap values of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0, 0.2, 0.4, 0.6, 0.8). (b) d-band center values of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO.
Figure 1. (a) Band gap values of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO (x = 0, 0.2, 0.4, 0.6, 0.8). (b) d-band center values of Co3O4, Mn3O4, CoxNi1−xO, NixMn1−xO, Co1−xMnxO.
Catalysts 13 01031 g001
Figure 2. (a) Schematic of the experiment process, (b) XRD of CoNi-n, (c) XRD of CoMn-n, (d) XRD of NiMn-n.
Figure 2. (a) Schematic of the experiment process, (b) XRD of CoNi-n, (c) XRD of CoMn-n, (d) XRD of NiMn-n.
Catalysts 13 01031 g002
Figure 3. (ac) LSV curves without IR compensation with arrows enlarged details. (a) LSV of CoNi-n, (b) LSV of CoMn-n, and (c) LSV of NiMn-n. (dg) OER test of CoNi-6, CoMn-3, and NiMn-3. (d) LSV with 90% IR compensation, I TOF curve, (f) i-t curves without IR compensation, (g) LSV without IR compensation. (h,i) HER test of CoNi-6, CoMn-3, and NiMn-3. (h) LSV with 90% IR compensation, (i) LSV without IR compensation.
Figure 3. (ac) LSV curves without IR compensation with arrows enlarged details. (a) LSV of CoNi-n, (b) LSV of CoMn-n, and (c) LSV of NiMn-n. (dg) OER test of CoNi-6, CoMn-3, and NiMn-3. (d) LSV with 90% IR compensation, I TOF curve, (f) i-t curves without IR compensation, (g) LSV without IR compensation. (h,i) HER test of CoNi-6, CoMn-3, and NiMn-3. (h) LSV with 90% IR compensation, (i) LSV without IR compensation.
Catalysts 13 01031 g003
Figure 4. (a) SEM image of CoNi-6, (b) SEM image of CoMn-3, (c) SEM image of NiMn-3, (di) XPS spectra of CoNi-6, CoMn-3, and NiMn-3.
Figure 4. (a) SEM image of CoNi-6, (b) SEM image of CoMn-3, (c) SEM image of NiMn-3, (di) XPS spectra of CoNi-6, CoMn-3, and NiMn-3.
Catalysts 13 01031 g004
Figure 5. Overall water-splitting performance of CoMn-3. (a) LSV curve, (b) double-layer capacitance analysis curve, (c) i-t curve measured at 1.78 V, and (d) comparison of the LSV curves before and after i-t testing.
Figure 5. Overall water-splitting performance of CoMn-3. (a) LSV curve, (b) double-layer capacitance analysis curve, (c) i-t curve measured at 1.78 V, and (d) comparison of the LSV curves before and after i-t testing.
Catalysts 13 01031 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lin, J.; Ding, Y.; Jin, H.; Zeng, T. Optimized Ni, Co, Mn Oxides Anchored on Graphite Plates for Highly Efficient Overall Water Splitting. Catalysts 2023, 13, 1031. https://doi.org/10.3390/catal13071031

AMA Style

Lin J, Ding Y, Jin H, Zeng T. Optimized Ni, Co, Mn Oxides Anchored on Graphite Plates for Highly Efficient Overall Water Splitting. Catalysts. 2023; 13(7):1031. https://doi.org/10.3390/catal13071031

Chicago/Turabian Style

Lin, Jie, Yihong Ding, Huile Jin, and Tianbiao Zeng. 2023. "Optimized Ni, Co, Mn Oxides Anchored on Graphite Plates for Highly Efficient Overall Water Splitting" Catalysts 13, no. 7: 1031. https://doi.org/10.3390/catal13071031

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