Ru and S co-modification-induced synergistic morphology and electronic engineering of nickel-iron hydroxide with efficient oxygen evolution

Shengxia Yang Yukang Pan Tianyu Kong Chaoran Jia Yueyang Cui Xuehua Li Yannan Zhou Haijun Liu Xinyu Zhang Bin Dong Qunwei Tang

Citation:  Shengxia Yang, Yukang Pan, Tianyu Kong, Chaoran Jia, Yueyang Cui, Xuehua Li, Yannan Zhou, Haijun Liu, Xinyu Zhang, Bin Dong, Qunwei Tang. Ru and S co-modification-induced synergistic morphology and electronic engineering of nickel-iron hydroxide with efficient oxygen evolution[J]. Chinese Chemical Letters, 2025, 36(12): 111540. doi: 10.1016/j.cclet.2025.111540 shu

Ru and S co-modification-induced synergistic morphology and electronic engineering of nickel-iron hydroxide with efficient oxygen evolution

English

  • Hydrogen features high energy density, high fuel efficiency, and being free of pollution, which is regarded as an ideal substitute for fossil fuels. Electrochemical water splitting is a promising method for producing clean hydrogen fuel by using renewable energy [13]. The oxygen evolution reaction (OER) and the hydrogen evolution reaction (HER) are the main half-reactions that occur at the anode and cathode in electrochemical water splitting respectively. Among them, the OER is an important basis for supporting a variety of cathode reactions such as the HER, carbon dioxide reduction, and nitrogen reduction reaction to achieve the conversion and storage of renewable energy [47]. However, the OER involves multiple electron-protons coupling steps, which limits the efficiency of the overall water electrolysis [811]. Therefore, developing OER catalysts with sufficient catalytic activity and stability is crucial to the overall efficiency of the water splitting reaction.

    It is reported that the OER catalytic activity of some non-precious metal materials is close to or sometimes even better than that of noble metal IrO2 or RuO2, which include crystalline/amorphous metal oxides [12,13], layered double hydroxides [14,15], spinel-type oxides [16,17], perovskite-type oxides [18,19], transition metal sulfide (TMS) [2022] and so on. Among them, nickel-iron-based catalysts stand out due to their great balanced oxygen adsorption/desorption abilities and are long considered as transition metals with relatively high activity [2325]. Yet the evaluations of nickel-iron-based catalysts have always been carried out under low current densities. In practical applications, the catalytic current densities that catalysts are forced to provide often exceed 500 mA/cm2 [26,27]. However, in the OER process at high current densities, it is often accompanied by various unfavorable microstructure evolutions of the catalytically active species (such as oxidative decomposition, structural changes, metal leaching and irregular aggregation) and the detachment of the catalytically active species from the working electrode, which is very likely to lead to the deactivation of the catalyst. High current density means rapid charge/mass transfer [2830]. Therefore, a large active surface area, high electrical conductivity and excellent hydrophilicity/hydrophobicity must be integrated into electrocatalysts to provide high-speed kinetics. Although it is crucial to develop nickel-iron catalysts that can operate stably at high current densities, it is still challenging.

    It is also worth noting that iron, usually associated with negative emotions and considered undesirable, is a positive and necessary metal-based catalyst for the production of OER electrodes [31,32]. Since nickel-iron catalysts exist in the form of flake-like structures, researchers have long paid more attention to using the strategy of electronic regulation to enhance the intrinsic activity of the catalysts, while ignoring the impact on the morphological structure of nickel-iron hydroxides during the process of regulating the intrinsic activity and this point has not been taken seriously for a long time [33,34]. If the morphology of nickel-iron catalytic materials is optimized simultaneously with the regulation of their intrinsic activity, by reducing the nanoscale size of the catalysts to increase their electrochemically active area, and consequently increasing the number of surface-active sites, then the OER performance of the catalysts will be further improved. However, reports on the synchronous regulation of the electronic structure and morphology of nickel-iron catalysts are relatively lacking at present.

    In light of the above analysis, herein, we have synthesized a nickel-iron hydroxide co-regulated by non-metal sulfur and metal ruthenium (SARuT-FeNiOHx-5h) through a simple multi-step room-temperature impregnation strategy. Physical characterizations and theoretical analyses show that the regulation by ruthenium and sulfur can not only modulate the electronic structure of the nickel-iron active centers and optimize the OER reaction pathway, but also adjust the growth process of the nickel-iron hydroxide catalyst, enabling it a larger specific surface area (enhanced 1.25 times). This effectively promotes the substantial contact between the catalyst and the electrolyte solution, thus enabling SARuT-FeNiOHx-5h to require a low overpotential of 253 mV when reaching a high current density of 1000 mA/cm2 with long-term stability (500 h), and these features endow it with a large prospect in practical industrial-level applications.

    The SARuT-NiFeOHx-5h was synthesized via a stepwise room temperature soaking strategy as described in Fig. 1a. First, Ru modified NiFeOHx supported on iron foam (RuT-NiFeOHx-5h) was achieved through a 5-h impregnation process, which was then used as template to prepare the SARuT-NiFeOHx-5h by introducing S species in a sodium sulfide solution. The mechanism of the above synthetic process was shown in Fig. 1b. Specifically, the surface of the IF first underwent an oxidation–reduction reaction with oxygen, generating a small amount of Fe3+ and OH⁻. When they encountered the added Ru3+ and Ni2+, the four species precipitated and grew attached to the surface of the IF, forming RuT-NiFeOHx-5h sample. During this process, CO2 from the air spontaneously dissolved in the solution to form CO32- species, which acted as a charge-balancing anionic regulator. Additionally, when the RuT-NiFeOHx-5h was soaked in a solution containing sodium sulfide, S2- would combine with Fe3+, Ru3+, and Ni2+ on the surface of RuT-NiFeOHx-5h to form local sulfides, thereby achieving the regulation of the catalytic performance of RuT-NiFeOHx-5h. Throughout the process of synthesizing SARuT-NiFeOHx-5h, the following chemical reactions (Eqs. 1–5) may be involved.

    $ \mathrm{Fe} \rightarrow \mathrm{Fe}^{2+}+2 \mathrm{e}^{-} $

    (1)

    $ \mathrm{Fe}^{2+} \rightarrow \mathrm{Fe}^{3+}+\mathrm{e}^{-} $

    (2)

    $ \mathrm{O}_2+2 \mathrm{H}_2 \mathrm{O}+4 \mathrm{e}^{-} \rightarrow 4 \mathrm{OH}^{-} $

    (3)

    $ \mathrm{Fe}^{3+}+\mathrm{Ni}^{2+}+\mathrm{Ru}^{3+}+\mathrm{OH}^{-} \rightarrow \mathrm{Ru}_{\mathrm{T}}-\mathrm{FeNiOH}_{\mathrm{x}} $

    (4)

    $ \mathrm{Ru}_{\mathrm{T}}-\mathrm{FeNiOH}_{\mathrm{x}}+\mathrm{S}^{2-} \rightarrow \mathrm{S}_{\mathrm{A}} \mathrm{Ru}_{\mathrm{T}}-\mathrm{FeNiOH}_{\mathrm{x}} $

    (5)

    Figure 1

    Figure 1.  (a) Schematic illustration of the synthesis process of SARuT-NiFeOHx-5h catalyst. (b) Synthesis mechanism of SARuT-NiFeOHx-5h. (c) XRD patterns of the prepared NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. Comparison of the SEM images of (d) NiFeOHx-5h, (e) RuT-NiFeOHx-5h, and (f) SARuT-NiFeOHx-5h. (g) TEM and (h) HRTEM images of SARuT-NiFeOHx-5h.

    Fig. S1 (Supporting information) was the optical photograph of IF, NiFeOHx-5h, RuT-NiFeOHx-5h and SARuT-NiFeOHx-5h. At each stage, the corresponding color has changed significantly. Fig. 1c showed the X-ray diffraction (XRD) of three substances including NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. It was found that among the three samples, only three obvious characteristic peaks of the IF appeared at 45.4°, 65.2°, and 82.4°, which corresponded to the (110), (200), and (220) crystal faces of the iron (PDF No. 00–001–1267).

    The morphology and structure of the prepared samples were further characterized by scanning electron microscopy (SEM). As shown in Fig. S2 (Supporting information), the initial IF had a relatively smooth and flat surface, which provided an idea foundation for the subsequent growth of the nickel-iron hydroxide precursor. After a 5-h room temperature soaking, the NiFeOHx-5h grew on the surface of the IF in the form of interwoven nanosheets and the size of the nanosheets being approximately 4–6 µm in length (Fig. 1d). When ruthenium ions were involved in the growth of NiFeOHx-5h, it was observed that the size of the nickel-iron hydroxide nanosheets noticeably decreased (about 500 nm), which might be related to the rapid nucleation of Ru(OH)3 to promote catalyst growth (Fig. 1e). After further S2- regulation, the smaller size of the nickel-iron hydroxide nanosheets was remained, indicating that the sulfur modification has not destroyed the structure of the RuT-NiFeOHx-5h (Fig. 1f). Figs. 1g and h were scanning transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images of SARuT-NiFeOHx-5h. As revealed, SARuT-NiFeOHx-5h showed a relatively thin petal-like nanosheet structure, and the size of the nanosheets was about 200 nm, which was corresponding to the SEM results in Fig. 1f. It is noteworthy that in the HRTEM images of SARuT-NiFeOHx-5h hardly any long-range ordered lattice fringes were found, confirming the amorphous state of the catalyst. SEM mapping and energy dispersive X-ray spectroscopy (EDS) revealed, as expected, the presence of Fe, Ni, O, S, and Ru (Fig. S3 in Supporting information), with the atomic proportions of Fe, Ni, O, S, and Ru being 51.23%, 8.44%, 35.92%, 4.13%, and 0.28%, respectively (Table S1 in Supporting information). Combined with the above XRD, SEM, SEM Mapping, and HRTEM analyses, it was believed that the nanosheets born on the iron foam were mainly amorphous nickel-iron hydroxide.

    In the comparative experiments mentioned above, it was found that the regulation of Ru3+ could effectively reduce the size of nickel-iron hydroxide nanosheets. Therefore, by comparing the morphology of NiFeOHx-5h under different Ru3+ additions (0, 0.00001, 0.001, 0.005, and 0.01 mol/L), the regulatory mechanism of Ru3+ on the growth and performance of nickel-iron hydroxide was further explored. As shown in Figs. 2a-e, the size of nickel-iron hydroxide nanosheets regulated by ruthenium gradually decreased with the increase of Ru3+ concentration. It was also found that when the concentration of Ru3+ in the soaking solution was 0.005 mol/L, the RuT-NiFeOHx-5h-0.005 exhibited the best OER performance, indicating that Ru3+ had the best regulatory effect on nickel-iron hydroxide under this condition (Fig. 2f). Based on SEM and LSV results, it was speculated that the regulation of ruthenium on NiFeOHx was related to the solubility product constants (KSP) of Ru(OH)3, Ni(OH)2, and Fe(OH)3. Specifically, the KSP value of Ru(OH)3 was 1.0 × 10–36, which was much smaller than that of Ni(OH)2 (2.0 × 10–15). Considering the trace amounts of Fe3+ in the solution, Ru3+ would rapidly combine with the small amount of OH- to form a large number of fine nanoparticles, which acted as crystal nuclei to promote the attachment and growth of nickel-iron hydroxide on their surfaces, thereby forming a large number of fine nanosheet structures (Fig. 2g). Compared with the pure NiFeOHx-5h, this abundance of small-sized nanosheets could effectively promote full contact between the catalyst and the electrolyte, thereby enhancing the overall catalytic performance of the NiFeOHx-5h.

    Figure 2

    Figure 2.  The influence of the concentration of Ru3+ in the soaking system on the growth process of the RuT-NiFeOHx-5h. SEM images of the RuT-NiFeOHx-5h at different Ru3+ addition levels: (a) 0 mol/L, (b) 0.0001 mol/L, (c) 0.001 mol/L, (d) 0.005 mol/L, and (e) 0.01 mol/L. (f) The polarization curves of RuT-NiFeOHx-5h at different Ru3+ concentrations. (g) Schematic diagram of the synthesis mechanism of RuT-NiFeOHx-5h catalyst.

    Further investigations of the chemical composition and valence state of the surface elements were probed by the X-ray photoelectron spectroscopy (XPS) technique. As shown in Fig. S4a (Supporting information), the S and Ru signal, together with the Ni, Fe, and O element peaks, were clearly present in the full spectrum of SARuT-NiFeOHx-5h, in agreement with the EDS results above. In the core-level, the Fe 2p spectra of NiFeOHx-5h exhibited two main peaks attributed to Fe 2p1/2 and Fe 2p3/2 respectively. The characteristic peaks with binding energies of 725.0 eV and 711.3 eV were assigned to Fe3+, another couple of characteristic peaks with binding energy of 733.0 and 715.5 eV were assigned to satellite peaks [3537]. Compared with pure NiFeOHx-5h, the peaks of Fe3+ in RuT-NiFeOHx-5h have shifted towards lower binding energy by approximately 0.3 eV. This change indicated an increase in the valence state of the Fe element in RuT-NiFeOHx-5h. On the contrary, the characteristic peaks of Fe3+ in SARuT-NiFeOHx-5h were located at lower binding energies of 724.0 and 710.6 eV (compared to RuT-NiFeOHx-5h), which might be attributed to the weak electronegativity of coordination sulfur relative to oxygen (Fig. S4b in Supporting information). For Ni 2p in Fig. S4c (Supporting information), the peaks at 873.62 and 855.7 eV, together with satellite peaks at 873.6 and 861.9 eV (Sat.), are indicative of the Ni 2p3/2 and Ni 2p1/2 orbitals of the Ni2+ species [3840]. The binding energy of Ni2+ in RuT-NiFeOHx-5h exhibited a positive shift of 0.5 eV compared to NiFeOHx-5h. This might originate from the interfacial electron transfer from Ni to Ru, thereby raising the average oxidation state of nickel. Conversely, the positions of the Ni2+ peaks in SARuT-NiFeOHx-5h were very similar to those in RuT-NiFeOHx-5h, indicating that sulfur primarily enhanced the OER performance of the catalyst by affecting the electronic structure of iron. In addition to the XPS spectra of the metal species, O 1s XPS spectra of three samples were shown in Fig. S4d (Supporting information). As shown, the O 1s spectra of catalysts all contained M-O bond, M-OH bond, and adsorbed water located at 529.8, 531.3, and 532.8 eV, respectively [4143]. It is evident that the regulation by sulfur and ruthenium resulted in a positive shift in the binding energy of oxygen, corresponding to the increase in the valence state of nickel and iron. In Fig. S4e (Supporting information), the Ru 3p core level spectrum of RuT-NiFeOHx-5h showed two distinct characteristic peaks at 486.7 and 463.3 eV, corresponding to the Ru 3p1/2 and Ru 3p3/2 orbitals, respectively. These peaks unequivocally indicated the presence of Ru3+ [4446]. Compared to RuT-NiFeOHx-5h, the Ru characteristic peaks of SARuT-NiFeOHx-5h exhibited a negative shift of approximately 0.4 eV, which implied a decrease in the oxidation state of Ru atoms in SARuT-NiFeOHx-5h. Regarding the S 2p spectrum depicted in Fig. S4f (Supporting information), SARuT-NiFeOHx-5h exhibited two peaks at 168.5 and 162.8 eV, corresponding to the S 2p1/2 and S 2p3/2 orbitals, respectively, unequivocally indicating the presence of S2- [47,48].

    The alkaline OER performances of SARuT-NiFeOHx-5h and reference catalysts were evaluated in a 1.0 mol/L KOH electrolyte. As illustrated in Fig. 3a, a comparison was made of the linear sweep voltammetry (LSV) curves for NiFeOHx under different soaking synthesis time. At a reaction time of 5-h, the NiFeOHx-5h exhibited optimal activity, corresponding to an overpotential of 320 mV at a current density of 1000 mA/cm2. To further investigate the effect of Ru3+ regulation mode on the OER performance of the catalyst, LSV test of RuT-NiFeOHx-5h and RuA-NiFeOHx-5h was conducted. As shown in Fig. 3b, RuT-NiFeOHx-5h required a smaller potential at the same current density. This observation signifies that Ru3+ contributed significantly to the NiFeOHx-5h growth process. Similarly, it was found that SA-NiFeOHx-5h showed better OER performance when modified by S2- after synthesis of NiFeOHx-5h (Fig. 3c). These results suggested that sulfur was more likely to enhance the performance of catalyst by modifying the surface electronic structure of NiFeOHx-5h. Based on the above analysis, the OER properties of different catalysts were further compared under the optimal conditions (5-h soaking synthesis time, adding Ru3+ together with growth of nickel-iron hydroxide, adding S2- after the growth of nickel-iron hydroxide) to reveal the regulatory results of Ru3+ and S2- on nickel-iron hydroxide. As expected, SARuT-NiFeOHx-5h, in Figs. 3d and e, exhibited the optimal OER activity, requiring overpotentials of only 253, 296, and 321 mV to attain the current densities of 100, 500, and 1000 mA/cm2, respectively. While the IF catalyst always afforded the largest η values and needed of 366 and 462 mV to reach the current densities of 10 and 100 mA/cm2, which was worse than those of NiFeOHx-5h (η100 = 275 mV, η500 = 321 mV, and η1000 = 341 mV), SA-NiFeOHx-5h (η100 = 268 mV, η500 = 315 mV, and η1000 = 340 mV), RuT-NiFeOHx-5h (η100 = 262 mV, η500 = 310 mV, and η1000 = 330 mV), and SARuT-NiFeOHx-5h (η100 = 253 mV, η500 = 296 mV, and η1000 = 321 mV). The large differences in overpotential between catalysts demonstrated that the presence of Ru atom and S atom could effectively regulate the intrinsic OER activity of NiFeOHx-5h. Even if the electrolyte was seawater (1.0 mol/L KOH containing 0.5 mol/L NaCl), the above rule remains unchanged (Fig. S7 in Supporting information). Meanwhile, due to its ultra-low overpotential, SARuT-NiFeOHx-5h demonstrated superior OER activity compared to most reported transition-metal-based catalysts and commercial iridium or ruthenium oxides (Table S2 in Supporting information). Next, the LSV polarization curves of the prepared samples were processed according to the Tafel equation to obtain the Tafel slopes for each sample (Fig. 3f). Specifically, the Tafel slopes for IF, NiFeOHx-5h, SA-NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h were 88, 68, 67, 63, and 61 mV/dec, respectively. Notably, SARuT-NiFeOHx-5h exhibited the smallest Tafel slope, indicating its superior catalytic activity and faster reaction kinetics among the samples. In addition, Nyquist plots were applied to further analyze the electronic and mass transport of catalyst (Fig. 3g). It was evident that SARuT-NiFeOHx-5h retained the smallest semicircle, indicative of the lowest electron transfer impedance. This phenomenon could be attributed to the redistribution of the electron configuration of the entire catalytic material, resulting from the modification of Ru3+ and S2- on NiFeOHx-5h. Meanwhile, to reflect the morphology regulation of ruthenium, the NiFeOHx-5h and RuT-NiFeOHx-5h were evaluated by electrochemically active surface area (ECSA). It was obvious that the presence of ruthenium in Fig. S5 (Supporting information) and Table S3 (Supporting information) increased the ECSA value of NiFeOHx-5h by 1.25 times, which could effectively promote full contact between the catalyst and the electrolyte. Finally, the stability of the SARuT-NiFeOHx-5h was also evaluated at a current density of 1000 mA/cm2. As shown in Fig. 3h and Fig. S6 (Supporting information), during the long-term stability test, the potential and structure change of SARuT-NiFeOHx-5h could be negligible, indicating that SARuT-NiFeOHx-5h could maintain idea stability when running under industrial current density.

    Figure 3

    Figure 3.  (a) Polarization curves of NiFeOHx with different immersion time with iR compensation. (b) Comparison of polarization curves of NiFeOHx-5h with diverse Ru3+ doping methods with iR compensation. (c) Comparison of polarization curves tests of NiFeOHx-5h with diverse S2- doping strategies with iR compensation. Comparison of (d) polarization curves and (e) overpotentials of IF, NiFeOHx-5h, RuT-NiFeOHx-5h, SA-NiFeOHx-5h, and SARuT-NiFeOHx-5h catalysts in 1.0 mol/L KOH with iR. (f) Tafel and (g) Nyquist plots of different samples with iR compensation. (h) Chronopotentiometry curve of SARuT-NiFeOHx-5h at a current density of 1000 mA/cm2 for 500h without iR compensation.

    The Gibbs free energy changes of reaction intermediates on the surface of nickel-iron hydroxide were investigated using density functional theory (DFT) calculations. The structures of NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h were shown in Fig. 4a. According to literature report, the Ni site in (001) crystal plane of nickel-iron hydroxide was selected to simulate the effect of sulfur ruthenium atom regulation on the OER performance of the catalyst. Meanwhile, the proposed adsorption evolution mechanism (AEM) of OER on the surface of catalyst was applied in Fig. 4b. Fig. 4c was the density of states (DOS) diagram of the three catalysts. As displayed, the center of the d band of SARuT-NiFeOHx-5h was closest to the Fermi level, indicating that the adsorption effect of reaction intermediates on the catalyst surface was significantly enhanced. Fig. 4d showed the calculated Gibbs free energy and the corresponding values. It was obvious that the rate determining step RDS of the NiFeOHx-5h was *O → *OOH (3.94 eV). On the contrary, for RuT-NiFeOHx-5h and SARuT-NiFeOHx-5h, the above RDS changed to *OOH→O2 and the reaction energy barrier decreased to 3.08 and 2.96 eV, respectively. Therefore, the calculated results of DFT revealed that the synergistic regulation of S and Ru could effectively improve the OER activity nickel-iron hydroxide, which was consistent with the experimental analysis.

    Figure 4

    Figure 4.  (a) The optimized structural model of NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. (b) AEM mechanism of OER on the surface of catalyst. (c) DOS diagram of NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. (d) The calculated Gibbs free energy and the corresponding values.

    In this work, a nickel-iron hydroxide catalyst supported on iron foam modified with nonmetallic sulfur and metallic ruthenium (SARuT-NiFeOHx-5h) was synthesized by a two-step room temperature impregnation strategy. It was found that the presence of ruthenium can effectively regulate the growth process of nickel-iron hydroxide, which significantly reduced the size of nanosheets and increased the contact area between catalyst and electrolyte (1.25 times). The subsequent modification of nickel-iron hydroxide by sulfur could promote the increase of metal valence in the active center of nickel, and thus enhanced the intrinsic activity of the catalyst. Resultantly, the optimal SARuT-NiFeOHx-5h exhibited an excellent OER activity with a lower overpotential 321 mV at 1000 mA/cm2. Meanwhile, the stability test of up to 500 h at such a high current density endowed SARuT-NiFeOHx-5h a large prospect for practical application.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    Shengxia Yang: Writing – review & editing, Writing – original draft, Investigation, Formal analysis, Data curation, Conceptualization. Yukang Pan: Writing – review & editing, Writing – original draft, Formal analysis, Conceptualization. Tianyu Kong: Writing – review & editing, Formal analysis. Chaoran Jia: Conceptualization. Yueyang Cui: Writing – original draft, Formal analysis. Xuehua Li: Writing – original draft, Formal analysis. Yannan Zhou: Writing – original draft, Formal analysis. Haijun Liu: Writing – review & editing, Funding acquisition, Formal analysis. Xinyu Zhang: Writing – review & editing, Resources, Project administration, Investigation, Funding acquisition, Formal analysis. Bin Dong: Writing – review & editing, Supervision, Investigation, Formal analysis, Conceptualization. Qunwei Tang: Writing – review & editing, Writing – original draft, Resources, Project administration, Investigation, Funding acquisition, Formal analysis.

    This work is financially supported by Shandong Provincial Natural Science Foundation (No. ZR2024QB021) and Qingdao Natural Science Foundation (No. 24–4-4-zrjj-21-jch) and National Natural Science Foundation of China (Nos. 62204098, 62304124, 22309107).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2025.111540.


    1. [1]

      M. Chen, N. Kitiphatpiboon, C. Feng, et al., eScience 3 (2023) 100111. doi: 10.1016/j.esci.2023.100111

    2. [2]

      G.A. Gebreslase, M.V. Martínez-Huerta, M.J. Lázaro, J. Energy Chem. 67 (2022) 101–137. doi: 10.1016/j.jechem.2021.10.009

    3. [3]

      K. Ayers, N. Danilovic, R. Ouimet, et al., Annu. Rev. Chem. Biomol. Eng. 10 (2019) 219–239. doi: 10.1146/annurev-chembioeng-060718-030241

    4. [4]

      R. Ram, L. Xia, H. Benzidi, et al., Science 384 (2024) 1373–1380. doi: 10.1126/science.adk9849

    5. [5]

      J. Zhang, X. Fu, S. Kwon, et al., Science 387 (2025) 48–55. doi: 10.1126/science.ado9938

    6. [6]

      Y. Chen, Q. Li, Y. Lin, et al., Nat. Commun. 15 (2024) 7278. doi: 10.1038/s41467-024-51521-4

    7. [7]

      S. Lu, Z. Zhang, C. Cheng, et al., Angew. Chem. Int. Ed. 64 (2025) e202413308. doi: 10.1002/anie.202413308

    8. [8]

      J. Song, C. Wei, Z.F. Huang, et al., Chem. Soc. Rev. 49 (2020) 2196–2214. doi: 10.1039/c9cs00607a

    9. [9]

      J. Zhao, Y. Guo, Z. Zhang, et al., Nat. Nanotechnol. 20 (2024) 57–66.

    10. [10]

      S. Yu, J. Li, Y. Du, et al., Coord. Chem. Rev. 520 (2024) 216144. doi: 10.1016/j.ccr.2024.216144

    11. [11]

      J. Xu, M. Li, B. Dong, L. Feng, Chin. Chem. Lett. 35 (2024) 108798. doi: 10.1016/j.cclet.2023.108798

    12. [12]

      J. Zhang, L. Xu, X. Yang, et al., Angew. Chem. Int. Ed. 63 (2024) e202405641. doi: 10.1002/anie.202405641

    13. [13]

      C.H. Park, H. Lee, J.S. Choi, et al., Adv. Mater. 36 (2024) 2403392. doi: 10.1002/adma.202403392

    14. [14]

      R. Fan, C. Liu, Z. Li, et al., Nat. Sustain. 7 (2024) 158–167. doi: 10.1038/s41893-023-01263-w

    15. [15]

      P. Roy Chowdhury, H. Medhi, K.G. Bhattacharyya, C. Mustansar Hussain, Coord. Chem. Rev. 483 (2023) 215083. doi: 10.1016/j.ccr.2023.215083

    16. [16]

      L. Chong, G. Gao, J. Wen, et al., Science 380 (2023) 609–616. doi: 10.1126/science.ade1499

    17. [17]

      M.G. Ahmed, Y.F. Tay, X. Chi, et al., Angew. Chem. Int. Ed. 64 (2025) e202416757. doi: 10.1002/anie.202416757

    18. [18]

      Y. Shan, J. Pan, Y. Song, et al., Chem. Phys. 582 (2024) 112292. doi: 10.1016/j.chemphys.2024.112292

    19. [19]

      P. Wang, C. Zhang, J. Ding, et al., Small 20 (2024) 2305662. doi: 10.1002/smll.202305662

    20. [20]

      S. Yu, J. Li, J. Yin, et al., Chin. Chem. Lett. 35 (2024) 110068. doi: 10.1016/j.cclet.2024.110068

    21. [21]

      S. Yu, D. Liu, C. Wang, et al., J. Colloid Interface Sci. 653 (2024) 1464–1477. doi: 10.1016/j.jcis.2023.10.006

    22. [22]

      S. Yu, N. Zhang, J. Li, et al., ACS Sustain. Chem. Eng. 12 (2024) 4551–4564. doi: 10.1021/acssuschemeng.3c07704

    23. [23]

      G.L. Li, F. Deng, T.G. Ma, et al., Chem. Eng. J. 505 (2025) 159290. doi: 10.1016/j.cej.2025.159290

    24. [24]

      Z. Li, G. Lin, L. Wang, et al., Nat. Catal. 7 (2024) 944–952. doi: 10.1038/s41929-024-01209-1

    25. [25]

      Y. Liu, Q. Song, T. Xu, et al., Appl. Catal. B Environ. Energy 363 (2025) 124811. doi: 10.1016/j.apcatb.2024.124811

    26. [26]

      D. Li, R. Xiang, F. Yu, et al., Adv. Mater. 36 (2024) 2305685. doi: 10.1002/adma.202305685

    27. [27]

      J. Liu, W. Du, S. Guo, et al., Adv. Sci. 10 (2023) 2300717. doi: 10.1002/advs.202300717

    28. [28]

      N. Yu, F.L. Wang, X.Y. Jiang, et al., J. Energy Chem. 102 (2025) 208–217. doi: 10.1016/j.jechem.2024.10.043

    29. [29]

      R. Wan, T. Yuan, L. Wang, et al., Nat. Catal. 7 (2024) 1288–1304. doi: 10.1038/s41929-024-01266-6

    30. [30]

      Y. Park, H.Y. Jang, T.K. Lee, et al., Nat. Commun. 16 (2025) 579. doi: 10.1038/s41467-025-55910-1

    31. [31]

      X. -Y. Zhang, Y. -R. Zhu, Y. Chen, et al., Chem. Eng. J. 399 (2020) 125831. doi: 10.1016/j.cej.2020.125831

    32. [32]

      X. Zhou, T. Yang, T. Li, et al., Nano Res. Energy 2 (2023) e9120086. doi: 10.26599/nre.2023.9120086

    33. [33]

      C. Yin, J. Li, S. Wang, et al., Carbon Energy 6 (2024) e553. doi: 10.1002/cey2.553

    34. [34]

      J. Wu, Z. Ma, L. Yu, et al., Sci. China Chem. 67 (2024) 2755–2766. doi: 10.1007/s11426-024-2077-2

    35. [35]

      Y. Gao, X. Sun, G.A.M. Mersal, et al., Adv. Compos. Hybrid Mater. 8 (2025) 117. doi: 10.1109/icosse65712.2025.00028

    36. [36]

      H. Liu, X. Lu, P. Shen, et al., ACS Appl. Nano Mater. 7 (2024) 7836–7844. doi: 10.1021/acsanm.4c00444

    37. [37]

      X. Tian, Y. Wang, F. Sun, et al., J. Alloy. Compd. 997 (2024) 174942. doi: 10.1016/j.jallcom.2024.174942

    38. [38]

      L. Fan, P. Zhang, B. Zhang, et al., ACS Energy Lett. 3 (2018) 2865–2874. doi: 10.1021/acsenergylett.8b01897

    39. [39]

      A.A. Abdelraouf, A.M. Abdelrahim, M.G. Abd El-Moghny, M.S. El-Deab, Int. J. Hydrog. Energy 101 (2025) 556–567. doi: 10.1016/j.ijhydene.2024.12.395

    40. [40]

      K. Shen, Y. Tang, Q. Zhou, et al., Chem. Eng. J. 471 (2023) 144827. doi: 10.1016/j.cej.2023.144827

    41. [41]

      Q. Zhou, C. Xu, J. Hou, et al., Nano-Micro Lett. 15 (2023) 95. doi: 10.1007/s40820-023-01066-w

    42. [42]

      Y. Jiang, Z. Song, M. Qu, et al., Small 20 (2024) 2400859. doi: 10.1002/smll.202400859

    43. [43]

      L. Kong, L. Pan, H. Guo, et al., J. Colloid Interface Sci. 664 (2024) 178–185. doi: 10.1016/j.jcis.2024.03.002

    44. [44]

      W.X. Zheng, X.X. Cheng, P.P. Chen, et al., Nat. Commun. 16 (2025) 337. doi: 10.70114/acmsr.2025.2.1.p337

    45. [45]

      L. Zhou, Y. Shao, F. Yin, et al., Nat. Commun. 14 (2023) 7644. doi: 10.1038/s41467-023-43466-x

    46. [46]

      R. Yao, K. Sun, K. Zhang, et al., Nat. Commun. 15 (2024) 2218. doi: 10.1038/s41467-024-46553-9

    47. [47]

      Y. Huang, L.W. Jiang, H. Liu, J.J. Wang, Chem. Eng. J. 441 (2022) 136121. doi: 10.1016/j.cej.2022.136121

    48. [48]

      M. Li, H. Li, H. Fan, et al., Nat. Commun. 15 (2024) 6154. doi: 10.1038/s41467-024-50535-2

  • Figure 1  (a) Schematic illustration of the synthesis process of SARuT-NiFeOHx-5h catalyst. (b) Synthesis mechanism of SARuT-NiFeOHx-5h. (c) XRD patterns of the prepared NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. Comparison of the SEM images of (d) NiFeOHx-5h, (e) RuT-NiFeOHx-5h, and (f) SARuT-NiFeOHx-5h. (g) TEM and (h) HRTEM images of SARuT-NiFeOHx-5h.

    Figure 2  The influence of the concentration of Ru3+ in the soaking system on the growth process of the RuT-NiFeOHx-5h. SEM images of the RuT-NiFeOHx-5h at different Ru3+ addition levels: (a) 0 mol/L, (b) 0.0001 mol/L, (c) 0.001 mol/L, (d) 0.005 mol/L, and (e) 0.01 mol/L. (f) The polarization curves of RuT-NiFeOHx-5h at different Ru3+ concentrations. (g) Schematic diagram of the synthesis mechanism of RuT-NiFeOHx-5h catalyst.

    Figure 3  (a) Polarization curves of NiFeOHx with different immersion time with iR compensation. (b) Comparison of polarization curves of NiFeOHx-5h with diverse Ru3+ doping methods with iR compensation. (c) Comparison of polarization curves tests of NiFeOHx-5h with diverse S2- doping strategies with iR compensation. Comparison of (d) polarization curves and (e) overpotentials of IF, NiFeOHx-5h, RuT-NiFeOHx-5h, SA-NiFeOHx-5h, and SARuT-NiFeOHx-5h catalysts in 1.0 mol/L KOH with iR. (f) Tafel and (g) Nyquist plots of different samples with iR compensation. (h) Chronopotentiometry curve of SARuT-NiFeOHx-5h at a current density of 1000 mA/cm2 for 500h without iR compensation.

    Figure 4  (a) The optimized structural model of NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. (b) AEM mechanism of OER on the surface of catalyst. (c) DOS diagram of NiFeOHx-5h, RuT-NiFeOHx-5h, and SARuT-NiFeOHx-5h. (d) The calculated Gibbs free energy and the corresponding values.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  20
  • HTML全文浏览量:  6
文章相关
  • 发布日期:  2025-12-15
  • 收稿日期:  2025-04-10
  • 接受日期:  2025-07-02
  • 修回日期:  2025-06-28
  • 网络出版日期:  2025-07-03
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章