Jie
Dong
a,
Saiyi
Chen
a,
Cuncai
Lv
b,
Mark G.
Humphrey
c,
Chi
Zhang
*a and
Zhipeng
Huang
*a
aChina-Australia Joint Research Center for Functional Molecular Materials, School of Chemical Science and Engineering, Tongji University, Shanghai, 200092, China. E-mail: chizhang@tongji.edu.cn; zphuang@tongji.edu.cn
bKey Laboratory of High-precision Computation and Application of Quantum Field Theory of Hebei Province, Hebei Key Lab of Optic-electronic Information and Materials, The College of Physics Science and Technology, Hebei University, Baoding 071002, China
cResearch School of Chemistry, Australian National University, Canberra, ACT 2601, Australia
First published on 9th January 2024
Developing an efficient bi-functional water-splitting catalyst is crucial for advancing sustainable hydrogen energy applications. A novel sulfur-doped cobalt molybdenum oxide (CoMoO) catalyst with a hydrangea-like structure was synthesized in situ on a carbon fiber paper substrate using a simple one-step hydrothermal process. The optimized sample exhibits excellent hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) catalytic activity with an overpotential of 105 and 205 mV at a current density of 10 mA cm−2 under alkaline conditions, respectively. Experimental characterization shows that the introduction of sulfur efficiently modifies the composition and morphology of the CoMoO electrocatalyst, which increases the electrochemically active surface area and improves the electrocatalytic kinetics. The improved electrocatalytic performance can be attributed to the synergistic effects of the various metal ion constituents. Mo4+ enhances water adsorption and accelerates the reaction kinetics of H2 generation from Co2+ or Sn2−. Meanwhile, Sn2− can accelerate the charge transfer, facilitating the OER process. Additionally, the presence of Mo4+ contributes to the stabilization of OER intermediates, thereby enhancing the efficiency of the OER process on CoOOH active sites. The hydrangea-like structure also offers plentiful active sites and efficient mass transfer. Besides, this electrolyzer obtains a current density of 10 mA cm−2 at a cell voltage of 1.61 V.
Transition metal-based catalysts have been extensively studied for efficient electrolysis technology due to their low cost, abundance, low toxicity, and high activity.18–22 In particular, bimetallic oxide catalysts have shown promise due to their multiple oxidation states, structural stability, unique electronic structures, and abundant defects, making them highly suitable for bi-functional HER and OER electrocatalysis.23–25 Among them, the bimetallic oxide CoMoO4 is a promising material with good intrinsic activity. Introducing the high-valent metal Mo into CoMoO4 can modulate the electronic structure of Co, improve electrical conductivity, and optimize water-splitting intermediate's adsorption energy, thereby enhancing catalytic performance.26–28 Therefore, developing bimetallic oxides with multifunctional catalytic properties is currently a hot topic in energy-related fields.
Non-metallic element doping effectively optimizes catalytic performance by modulating the electronic structure, enhancing wettability, lowering the kinetic energy barrier, and introducing additional active sites.29–32 Pang's work involved the introduction of nitrogen doping to enhance the electronic conductivity and OH adsorption strength of Co3O4, resulting in accelerated reaction kinetics and improved catalytic activity for the OER.33 Liu's research group has shown that phosphorus doping optimizes the electronic configuration and enhances the valence state of iron ions.34 This results in a nearly twofold increase in the mass and specific activity of LaFeO3−δ.34 Similar to doping with P or N, sulfur doping can also improve the electrochemical performance of metal oxides.32 Sulfur doping can replace oxygen atoms to improve intrinsic conductivity and introduce abundant oxygen vacancies to provide enough catalytically active sites.32 Inspired by these findings, we aim to explore the potential of introducing S into cobalt molybdate to enhance the performance of the HER and OER further. However, the direct synthesis of metal oxides doped with non-metallic elements through these strategies still poses challenges due to the requirement of additional reactions with compounds containing heteroatoms. The synthesis of heteroatoms also suffers from drawbacks such as using toxic precursors, complex processes, and releasing toxic gases (H2S, PH3, etc.). While extensive studies have shown the benefits of doping non-metallic elements in metal oxides to improve electrochemical and electrocatalytic properties, there is limited literature regarding sulfur doping.
Based on the above discussion, we synthesised a bi-functional S-doped cobalt molybdate (S-CoMoO-12.4) electrocatalyst on carbon fiber paper (CFP) by a simple one-step hydrothermal method. Such a catalyst possesses a hydrangea-like structure with a large specific surface area and rapid mass transfer capabilities. As expected, S-CoMoO-12.4 demonstrates ultralow overpotentials of 105 and 205 mV at a current density of 10 mA cm−2 in alkaline HER and OER processes, respectively. The experimental results show that introducing sulfur could effectively tune the electrocatalyst composition and morphology, increase the electrochemically active surface area and improve the electrocatalytic kinetics. The improved electrocatalytic performance can be attributed to the synergistic effects of the various ion constituents. Mo4+ as a reactant adsorption site enhances the adsorption of H2O molecules, thereby promoting the occurrence of the HER in the Co2+ or Sn2− site. Moreover, Mo4+ facilitates the attraction of electrons from the OER intermediate, stabilizing the intermediate and enhancing OER activity. Sn2− species also promote the OER processes, attributed to their ability to accelerate charge transfer. A two-electrode electrolyzer achieves 10 mA cm−2 with a cell voltage of 1.61 V for overall water splitting in 1.0 M KOH. This work presents a promising strategy for rationalizing and synthesizing efficient bimetallic electrocatalysts.
The synthesis of S-CoMoO-12.4 was carried out using a one-step hydrothermal method. In a beaker, 3.0 mmol of Co(NO3)2·6H2O, 15.0 mmol of urea, 3.6 mmol of (NH4)6Mo7O24·4H2O, and 37.3 mmol of thiourea are initially dissolved in 80 mL of deionized water, along with 6.0 mmol of NH4F. Magnetic stirring is applied to the mixture for 30 minutes before transferring it to a 150 mL Teflon-lined autoclave containing a CFP (3 cm × 6 cm). The autoclave is sealed, and the temperature is gradually raised at a rate of 1.5 °C min−1 to reach 120 °C, which is maintained for 7 h. Subsequently, the temperature is presented at a rate of 1.3 °C min−1 to 200 °C and held there for 8 h. Finally, the autoclave is allowed to cool to room temperature, and the synthesized S-CoMoO is washed with deionized H2O and C2H5OH. S-CoMoO-R (R = 3.1, 6.2, 9.3, 15.5, 18.6, representing the feeding atomic ratio of S/Co) is utilized as a control sample. The synthesis of CoMoO follows a similar procedure to that of S-CoMoO, except that thiourea is not included in the initial feedstock.
The typical morphology of S-CoMoO samples is shown in Fig. S1.† At low R, the nanosheets (thickness approximately 37 nm) were randomly arranged on the CFP surface (Fig. S1a†). As the R increases, the nanosheet thickness in S-CoMoO-6.2 (Fig. S1b†) becomes nearly twice that in S-CoMoO-3.1. In addition, these nanosheets self-assemble into microspheres with an average diameter of about 2.4 μm. However, as the R increases, the nanosheets transition from self-assembled microspheres to varying-sized cubes, as observed in Fig. S1c.† As R continues to grow, the S-CoMoO-12.4 catalyst exhibited a morphology resembling hydrangea-like microspheres with an average diameter of approximately 1.9 μm (Fig. 2b and c). Nevertheless, a high R leads to a blocky cobalt molybdate morphology (Fig. S1e and f†), significantly reducing the surface area. Also, the Brunauer–Emmett–Teller (BET) surface area (Fig. S2†) of CoMoO-15.5 (12.5 m2 g−1) and CoMoO-18.6 (6.7 m2 g−1) is smaller than that of CoMoO-12.4 (38.5 m2 g−1) (Fig. S3a†). The SEM images demonstrate the significant influence of R on the morphology of cobalt molybdate. Varied S/Co feeding atomic ratios can influence the growth and self-assembly of CoMoO nanosheets, thereby offering a means to adjust the structure and morphology of CoMoO and enhance its catalytic performance. To better understand how the actual sulfur levels affect their morphology, we used energy dispersive X-ray (EDX) spectroscopy to determine their elemental contents (Fig. S3b, S4 and Table S1†), the results of which are shown in Fig. S3b, S4 and Table S1.† The TEM image (Fig. 2d) of S-CoMoO-12.4 shows the structure to be an embroidered spherical structure with an average diameter of 2.0 μm, consistent with the SEM results. In addition, the high-resolution TEM (HRTEM) images (Fig. 2e–i) of S-CoMoO-12.4 reveal two distinct lattice stripes with lattice spacings of 0.32 and 0.20 nm, corresponding to the XRD results at 29.2° and 46.8° of S-CoMoO-12.4. The high-angle annular dark field scanning transmission electron microscopy (HAADF-STEM) images and synchronous EDX mapping (Fig. 2j) show that Co, S, Mo, and O elements are uniformly distributed throughout the entire sample, further confirming successful sulfur doping.
XPS was employed to reveal further the effect of S doping on the sample's electronic structure. The XPS analysis investigated the chemical composition and element states in Fig. 3 and S5.† The survey spectrum (Fig. S5†) indicates the presence of Co, S, Mo, and O elements in the S-CoMoO-12.4 sample. The C 1s peak at 284.8 eV was used as the XPS correction peak. The Co 2p XPS spectrum of S-CoMoO-12.4 (Fig. 3a) exhibited peaks at binding energies near 781.8 and 797.6 eV, corresponding to the Co 2p3/2 and Co 2p1/2 energy levels. In addition, the binding energy difference between Co 2p3/2 and Co 2p1/2 is 15.8 eV, a sign of the Co2+ oxidation state.35,36 Satellite peaks were observed at 786.4 and 802.8 eV. With the incorporation of S, a group of new peaks at 779.5 (Co 2p3/2) and 794.3 eV (Co 2p1/2) appeared, indicating the presence of trivalent cobalt in S-CoMoO-12.4.37,38 This observation indicates that the highly electronegative O atom is partly substituted by the less electronegative S atom, leading to the non-metallic atom being less attractive to the electrons of the metal Co atom, causing the Co atom to be more prone to losing electrons and thus increasing its oxidation state.39 In the Mo 3d spectrum of S-CoMoO-12.4 (Fig. 3b), two peaks at 233.0 and 236.1 eV can be deconvoluted, attributed to Mo 3d5/2 and Mo 3d3/2, respectively. The width between these spin–orbit split peaks is 3.1 eV, suggesting that Mo 3d is the same as Mo6+.40,41 Additionally, the appearance of two new peaks located at 231.4 eV and 228.3 eV indicates the generation of new Mo4+ species, providing further evidence for the successful introduction of S.42 The presence of Mo4+ and the formation of oxygen vacancies (OV) may be attributed to the capture of electrons by Mo6+ from the nearby lattice oxygen (OLat).43 Specific reactions may be OLat − 2e− → 1/2O2 + OV2+ and Mo6+ + 2e− → Mo4+. The S 2p spectrum (Fig. 3c) exhibited a peak at 163.2 eV attributed to the binding of Sn2− (polysulfide),44,45 and other peaks at 168.9 eV can be attributed to surface oxidation in air.46 Notably, forming Sn2− species could induce changes in the metal site's electronic structure and d-band center, thereby promoting the catalytic process.47
Furthermore, the O 1s spectrum (Fig. 3d) can be deconvoluted into three peaks. The peak at 530.9 eV can be attributed to an oxygen atom bound to the metal. The peak at 532.4 eV is associated with a hydroxyl species, accompanied by adsorbed water molecules on or near the surface. Finally, the peak observed at 531.7 eV indicates an oxygen vacancy.48 In addition, the OV/OLat on the surface of CoMoO after sulfur doping increased from 0.79 to 1.14. These oxygen vacancies significantly improve the conductivity of the catalyst and accelerate the kinetics of surface redox reactions, thereby enhancing its catalytic performance.49,50
To investigate the kinetics of the HER, the Tafel plot was employed, with a smaller Tafel slope indicating more favorable reaction kinetics. In agreement with the LSV results, the S-CoMoO-12.4 (Fig. 4c) catalyst exhibited the lowest Tafel slope value of 108.7 mV dec−1, indicating the fastest HER kinetics due to its fast electron transport. Notably, the S-CoMoO-12.4 catalyst exhibited a remarkably low Tafel slope, suggesting that the HER process of this catalyst follows a Volmer–Heyrovsky mechanism.51 To gain a more comprehensive understanding of the inherent activity of the HER, we conducted a normalization of the TOF of the HER in relation to the number of active sites, as depicted in Fig. S6.† At an overpotential of 300 mV, the TOF of the S-CoMoO-12.4 catalyst was measured to be 2.8×10−2 s−1, surpassing the TOF values of CoMoO (1.8 × 10−4 s−1). This observation aligns with the superior HER activity exhibited by the S-CoMoO-12.4 catalyst.
Apart from activity, the stability of electrocatalysts is of utmost importance for practical applications. As shown in Fig. 4d, S-CoMoO-12.4 showed negligible degradation after 40 hours of operating at 10 mA cm−2, highlighting its remarkable stability in the HER process. The LSV curve inset of Fig. 4d indicates only a slight change of 23 mV compared to the initial curve after 40 hours of the HER, further supporting its exceptional stability. Additionally, SEM images (Fig. S7†) reveal that the morphology of S-CoMoO-12.4 remains virtually unchanged following 40 hours of continuous operation. To elucidate the potential chemical structure of CoMoO-12.4 after the HER stability test, the XPS spectra of CoMoO-12.4 before and after the 40 hours stability test were compared. Fig. S8 and S9† show that CoMoO-12.4 after stability testing still shows characteristic peaks corresponding to Co 2p, Mo 3d, S 2p, and O 1s, indicating no significant alteration in chemical composition. However, the absence of the Co3+ 2p peak after stability testing can be attributed to the reduction of Co3+ to Co2+ (Fig. S8a†).
EIS is employed to investigate the kinetic properties of electrodes further. The obtained EIS data (Fig. 5a, S10 and Table S2†) were fitted with an equivalent circuit model. Additionally, the charge transfer resistance (Rct) of CoMoO (10.03 Ω) was found to be significantly higher than that of S-CoMoO-12.4 (2.66 Ω), indicating faster HER kinetics and an expedited faradaic process on the S-CoMoO-12.4 electrode. This can be attributed to the reduced energy barrier for the HER process resulting from the sulfur doping.47 In addition, the results obtained from EIS fitting (Table S2†) show that the resistances of the equivalent series resistance (Rs) of S-CoMoO-12.4 (1.38 Ω) is smaller than that of CoMoO (1.76 Ω), confirming that sulfur doping can improve the conductivity of CoMoO.
Generally speaking, sulfur-doped electrocatalysts possess a distinct advantage in their high ratio of active catalytic surface sites per unit mass of catalyst material. To demonstrate this advantage, the effective electrochemically active surface area (ECSA) of the S-CoMoO-12.4 electrocatalyst was assessed by cyclic voltammetry (CV) measurements to determine double-layer capacitance (Cdl). The ECSA was calculated via the equation ECSA = Cdl/Cs, where Cs represents the specific capacitance of the sample. This study employed a Cs value of 0.04 mF cm−2 based on previously reported data for metal oxides/hydroxides in alkaline solutions.36 CV curves were acquired at different scan rates (ranging from 10 to 50 mV s−1) within the non-Faraday range (0.281–0.381 V vs. RHE) (Fig. S11a and S10b†). The current density (Δj) difference at the intermediate potential (0.33 V) was calculated for each scan rate. The Cdl value was determined by calculating the slope of the Δj/2 versus the scan rate curve. Analysis of Fig. S11c† yielded a Cdl value of approximately 0.10 mF cm−2 for S-CoMoO-12.4, which is higher than that of the CoMoO sample (0.05 mF cm−2), suggesting that S doping could be favorable for exposing more electrochemically active sites in the HER process. Furthermore, it benefits from the hydrangea-like structure with large surface areas and increasingly plentiful exposed edges that afford a mass of active sites for the electrochemical process.52 The sample's normalized LSV curves (Fig. 5c), obtained from their ECSA measurements, further corroborate that S-CoMoO-12.4 exhibits superior intrinsic HER activity compared to CoMoO, thus confirming that sulfur doping enhances the properties of the active site. Detailed analysis of the actual HER active site will be elaborated in the following sections.
Following Elbert et al.'s procedure, we conducted a kinetic analysis to assess the standard activation free energies for the three elementary reaction steps of the HER (for details see the ESI†).53,54 We utilized the dual-pathway kinetic model to fit the experimental data (from Fig. 5b) of kinetic current density in order to obtain the mechanism and kinetic parameters of the HER.53,54 The advantage of this method is that it can consider various complex factors and uncertainties in the actual system and more accurately describe the behavior of the real HER. In addition, without the crystallographic information file for CoMoO, we can't do density functional theoretical calculations on it. We constructed a free energy diagram based on the fitted parameters of standard activation free energy. This diagram (Fig. 5c) visually represents the reaction barriers involved in the HER on CoMoO and S-CoMoO-12.4 catalysts. The parameters used in the diagram include for the Tafel step,
for the Volmer step (water dissociation), ΔG0ad for the standard free energy of adsorption H, and
for the Heyrovsky step (electrochemical recombination).53,54 The free energy is lower for the Heyrovsky step than for the Tafel step at 0 V
(Fig. 5b), indicating that the Volmer–Heyrovsky pathway dominates the HER for both CoMoO and S-CoMoO-12.4.53,54 The result is in good agreement with the Tafel slope analysis. As shown in Fig. 5c and d, the activation energies on S-CoMoO-12.4 are lower for both the Volmer step
and the Heyrovsky step
compared to those on bare CoMoO. Specifically, ΔG1 (S-CoMoO-12.4, 105 meV) < ΔG1 (CoMoO, 118 meV) and ΔG2 (S-CoMoO-12.4, 312 meV) < ΔG2 (CoMoO, 356 meV). This suggests that sulfur doping accelerates the water dissociation and electrochemical recombination steps of the HER in the alkaline environment.
Fig. 6e shows the long-term durability test of S-CoMoO-12.4 conducted for 40 hours at an applied current density of 10 mA cm−2. Impressively, S-CoMoO-12.4 demonstrates good electrochemical OER stability during this extended testing period. The inset of Fig. 6e clearly shows a minimal shift of 62 mV in the polarization curve of S-CoMoO-12.4 before and after the 40 hours OER test, suggesting remarkable stability. In addition, the high-resolution Co 2p spectra of S-CoMoO-12.4 obtained before and after the OER test (Fig. S16†) confirm significant changes in the oxidation state of cobalt. Notably, the main peaks of Co 2p are positively shifted by about 0.9 eV to 780.4 and 779.5 eV. These energy shifts can be attributed to CoOOH species, which are identified as the active centers for the OER process.55 In addition, the decay of the characteristic satellite peak of Co2+ proves the conversion of Co2+ to the CoOOH active center during OER catalysis. Furthermore, the SEM image in Fig. S17† indicates that the morphology of S-CoMoO-12.4 exhibits minimal alteration even after 40 hours of uninterrupted operation.
The S-CoMoO-12.4 catalyst demonstrates favorable electrocatalytic performance in both the HER and OER. To showcase the potential application of S-CoMoO-12.4 in overall water splitting, a two-electrode electrolyzer employs S-CoMoO-12.4 as both the anode and cathode materials. For comparative purposes, a commercial cell consisting of RuO2 (anode) and 20 wt% Pt/C (cathode) is also employed. The S-CoMoO-12.4‖S-CoMoO-12.4 system exhibits a remarkable current density of 10 mA cm−2 at a cell voltage of only 1.61 V (Fig. 6d). Surprisingly, as the current density is 100 mA cm−2, the required overpotential of our S-CoMoO-12.4‖S-CoMoO-12.4 system (1.80 V) is even lower than that of the commercial 20 wt% Pt/C‖RuO2 cell (1.89 V). Furthermore, the long-term durability of the S-CoMoO-12.4 catalyst for overall water splitting is confirmed by a negligible degradation of only 23 mV after 1000 cycles (Fig. S18†). These findings unequivocally establish the S-CoMoO-12.4 composite as a highly efficient electrocatalyst for water splitting.
To investigate the underlying mechanism of activity enhancement, we conducted XPS characterization of CoMoO and S-CoMoO-12.4 samples after HER or OER testing. Firstly, we analyzed the potential mechanism of HER activity enhancement. The Co 2p spectra of all samples after the HER showed two sets of peaks, where 781.8 and 797.6 eV corresponded to Co2+ and 786.4 and 802.8 eV corresponded to satellite peaks (Fig. S19 and S20†). The disappearance of the Co3+ peak in the initial S-CoMoO-12.4 (Fig. S19†) indicated that Co3+ in S-CoMoO-12.4 may have been reduced to Co2+ during the HER. The Mo 3d spectra of both samples after the HER (Fig. S19 and S20†) were similar to those before testing, but only the S-CoMoO-12.4 sample contained a Mo4+ peak (228.3 and 231.4 eV). After HER testing, the Mo4+ content in S-CoMoO-12.4 increased from 49.73% to 62.75%, suggesting that some of the Mo6+ in S-CoMoO-12.4 was reduced (Fig. 7c and Table S5†). The unsaturated Mo4+ sites can achieve electron focusing on the Mo site, thereby enhancing the adsorption of H2O.56,57
In addition, the O 1s spectra of S-CoMoO-12.4 after HER testing (Fig. S19 and S20†) showed a higher H2O peak content (25.77%) compared to CoMoO (19.73%), further confirming this observation. Furthermore, the Ov/OLat value of S-CoMoO-12.4 after HER testing (1.43) was higher than that of CoMoO (0.75), indicating that S-CoMoO-12.4 had more oxygen vacancies after HER testing (Fig. 7c). In transition metal oxides, abundant Ov can enhance electronic conductivity and facilitate water dissociation, significantly promoting the alkaline HER reaction kinetics.58–60 As mentioned above, a large number of Ov in the catalyst are highly beneficial for promoting alkaline HER. Moreover, bridging Sn2− species is believed to serve as active sites for the HER, owing to their moderate Gibbs free energy for atomic hydrogen adsorption.61–63 Based on the above results, we propose the potential mechanism of the HER on S-CoMoO-12.4, as illustrated in Fig. 7e. Specifically, the process begins with the initial adsorption of H2O molecules on Mo4+ sites, which is followed by the subsequent splitting of water into hydroxide ions (OH−) and hydrogen (H) atoms (hydrogen absorption). The H atoms then migrate to adjacent Co2+ or Sn2− sites and ultimately recombine to form H2, while the OH− desorbs from Mo4+ or Sn2− sites (electrochemical hydrogen desorption).
The Co 2p spectra of all samples were obtained after OER testing to investigate the potential mechanism of OER activity enhancement (Fig. S21 and S22†). The spectra showed three sets of different peaks, with peaks at 781.8 and 797.6 eV corresponding to Co2+, peaks at 780.4 and 795.6 eV corresponding to CoOOH, and peaks at 786.4 and 802.8 eV corresponding to satellite peaks (Fig. S21 and S22†). The appearance of the CoOOH peak in the OER process was attributed to the oxidation of Co2+. Furthermore, the CoOOH/(Co2+ + CoOOH) value of S-CoMoO-12.4 (54.75%) after OER testing was higher than that of CoMoO (39.66%), indicating that S-CoMoO-12.4 contained more OER active species (CoOOH) after OER testing. The Mo 3d spectra of both samples after OER testing (Fig. S21, S22 and Table S5†) were similar to those before testing, but only the S-CoMoO-12.4 sample contained a Mo4+ peak (228.3 and 231.4 eV). Compared to before OER testing, the Mo4+ content in S-CoMoO-12.4 decreased from 49.73% to 21.96%, indicating that Mo4+ in S-CoMoO-12.4 was partially oxidized during the OER. Mo4+ can attract electrons from oxygen evolution intermediates (OOH*), promoting their formation and stabilization, and increasing oxygen release.64
Furthermore, the Ov/OLat value of S-CoMoO-12.4 (1.40) after OER testing (Fig. 7d, S21 and S22†) was higher than that of CoMoO (0.59), indicating that S-CoMoO-12.4 had more Ov after OER testing. Ov can facilitate the surface reconstruction of transition metal oxides into metal hydroxides, thereby promoting the OER.65 The Sn2− can accelerate the charge transfer, thereby promoting the OER catalytic process.66 Additionally, the strong chemical bond of Sn2− in S-CoMoO-12.4 can improve electrochemical stability and accelerate the charge transfer process of S-CoMoO-12.4. Previous studies have demonstrated that transferring electrons from metals to heteroatoms can weaken metal bonds, thereby reducing the energy barrier for forming intermediates and enhancing OER activity.67–69 Based on the results above, the OER mechanism of S-CoMoO-12.4 may involve the adsorbate evolution mechanism, which encompasses a four-step synergistic electron transfer cycle for oxygen generation, as illustrated in Fig. 7e. The superior OER performance of S-CoMoO-12.4 can be ascribed to the synergistic effect of its CoOOH, Mo4+, and Sn2− constituents. The presence of Mo4+ facilitates the attraction of electrons from the OER intermediates, thereby contributing to the stabilization of the OER intermediates and consequently promoting oxygen generation on CoOOH active sites. Additionally, the inclusion of Sn2− is found to accelerate the charge transfer in the OER process. These combined effects collectively contribute to the remarkable OER performance observed in S-CoMoO-12.4.
Overall, the exceptional bi-functional activity and stability of S-CoMoO-12.4 can be attributed to several key factors. Firstly, the unique hydrangea structure exhibits a larger surface area and an increased number of exposed edges, which provide abundant active sites for electrochemical processes. Secondly, the introduction of S-doping alters the electronic structure of CoMoO, leading to improved conductivity and facilitating electron transfer. Thirdly, the presence of unsaturated Mo4+ sites enhances the adsorption of H2O molecules, thereby facilitating the HER. The Mo4+ species in the catalyst can also attract electrons from the OER intermediates, significantly enhancing the OER activity. Fourth, the existence of Sn2− species promotes both the HER and OER processes. It serves as the active site of the HER and also accelerates the charge transfer in the OER process. Lastly, Ov plays a critical role in enhancing the catalytic performance of transition metal oxides. Specifically, Ov promotes the restructuring of transition metal oxide surfaces into metal hydroxides, thereby facilitating the OER. Moreover, the abundance of Ov enhances electronic conductivity and facilitates water dissociation, thereby significantly influencing the kinetics of alkaline HER. Collectively, these factors synergistically influence the rate and efficiency of electrochemical reactions, playing a crucial role in improving the activity and efficiency of the catalyst.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3ta07380g |
This journal is © The Royal Society of Chemistry 2024 |