Next Article in Journal
Intelligent Optimization of Single-Stand Control in Directional Drilling with Single-Bent-Housing Motors
Previous Article in Journal
Optimizing Texture and Drying Behavior of Squid (Todarodes pacificus) for Elder-Friendly Applications Using Alkaline Pretreatment and Intermittent Drying: An Experimental and Numerical Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Rapid Joule-Heating Synthesis of Efficient Low-Crystallinity Ru-Mo Oxide Catalysts for Alkaline Hydrogen Evolution Reaction

1
Key Laboratory of Marine Chemistry Theory and Technology, Ministry of Education, College of Chemistry and Chemical Engineering, Ocean University of China, Qingdao 266100, China
2
Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensification Technology, Guangxi University, Nanning 530004, China
*
Author to whom correspondence should be addressed.
Processes 2025, 13(8), 2594; https://doi.org/10.3390/pr13082594
Submission received: 25 July 2025 / Revised: 11 August 2025 / Accepted: 14 August 2025 / Published: 16 August 2025
(This article belongs to the Section Environmental and Green Processes)

Abstract

Electrocatalytic water splitting has been demonstrated to be a highly efficient and promising technology for green hydrogen production. However, the inefficiency and instability of the cathode hinder its wide application in water electrolysis. Herein, we report a rapid Joule heating method for synthesizing the Ru-Mo oxide catalyst. Comprehensive characterization results confirmed that the as-prepared catalyst featured an internal porous structure with low crystallinity, which weakened the strength of Ru-H bonds through structural and electronic modulation. The enhanced HER performance was attributed to the incorporation of Mo4+ species, which strengthened Ru-O-Mo interactions. As tested, the optimized catalyst exhibited ultralow overpotentials (25.08 mV and 120.52 mV @ 10 and 100 mA cm−2, respectively) and excellent stability (100 h @ 100 mA cm−2) in a 1 M KOH solution. Meanwhile, the as-prepared catalyst was equipped in an anion exchange membrane (AEM) alkaline water electrolyzer, which could deliver 185 mA cm−2 at only 2.16 V with 100% Faradaic efficiency. This study provides a feasible strategy for constructing highly efficient low-crystallinity electrocatalysts.

1. Introduction

Hydrogen (H2), characterized by zero emissions and high energy density (140.4 MJ/kg), has emerged as a sustainable alternative to fossil fuels and an efficient energy carrier [1]. However, hydrogen production primarily relies on fossil fuel reforming and industrial by-products, which result in substantial energy consumption and environmental degradation due to carbon emissions. In contrast, electrocatalytic water splitting is regarded as a promising pathway for industrial-scale green hydrogen generation, featuring both pure H2 production and environmental friendliness [2]. Nevertheless, the high energy consumption (4.2 kWh Nm−3 H2) restricts its widespread applications [3]. Therefore, the development of efficient and robust electrocatalysts is essential for energy conservation and advancing large-scale utilization. Platinum-based catalysts (e.g., commercial Pt/C) exhibit exceptional hydrogen evolution reaction (HER) activity due to nearly optimal hydrogen adsorption energy (ΔG*H ≈ 0 eV). However, the low reserves and high cost of platinum severely impede its large-scale application in water electrolysis [4]. As a result, it is critical to develop cost-effective electrocatalysts that deliver high efficiency and stability, thereby advancing electrocatalytic water splitting toward an industrially viable process.
Among various alternatives, ruthenium (Ru)-based materials, such as RuO2 and unsaturated RuOx clusters, have received much attention in water electrolysis due to their higher reserves and lower cost (approximately half that of Pt) [5,6,7,8]. Specifically, RuO2 facilitates the adsorption and dissociation of water molecules, which makes it an ideal catalyst for the water dissociation reaction under alkaline conditions [9,10]. However, the high adsorption of hydrogen intermediates on Ru (ΔG*H ≪ 0 eV) results in inferior HER performance [11]. Introducing suitable secondary metals (e.g., Mo, Ni, and Co) into Ru-based materials has proven to be an effective strategy to overcome the dynamic constraint [12]. In detail, the interaction among these elements and Ru-based materials has been reported to optimize the electronic structure by lowering the d-band center, thereby weakening Ru-H interactions and tuning the hydrogen adsorption free energy (ΔG*H) closer to the ideal value (0 eV) [13,14]. Accordingly, molybdenum (Mo) has been comprehensively explored due to its abundant reserves and tunable electronic structures [15,16] and exhibits potential as a rational secondary metal. Recent studies have demonstrated that the synergistic effect between Ru and Mo species can accelerate the efficiency of water splitting, in accordance with the Brewer–Engel theory [17]. For instance, Fan et al. synthesized Ru-, RuO2-, and MoO3-embedded carbon nanorods (Ru-RuO2/MoO3 CNRs-350) [18], wherein the strong interaction between Ru and Mo species enabled the catalyst to exhibit ultralow overpotential (9.2 and 65.4 mV, respectively) at a current density of 10 and 100 mA cm−2 in 1.0 M KOH, which is comparable to that of Pt/C (20 wt%).
Despite the advances described above, it is difficult to overcome the inherent activity limitations of Ru-based materials. It is critical to develop effective strategies, such as phase modulation [19], morphology engineering [20], heteroatom doping [21], heterostructure construction [22], etc. Recently, low-crystallinity materials have gained attention as another promising approach for enhancing intrinsic activity [23]. Specifically, the low crystallinity not only provides abundant defects and active sites to facilitate the adsorption of intermediates but also enhances the conductivity of active materials and accelerates electron transfer rates [24,25]. Furthermore, the relatively disordered atomic arrangement of low-crystallinity materials can realize rapid charge transfer, thereby creating an ideal electronic structure for active sites [8,26]. Guo et al. created a rough layer with extremely low crystallinity on the surface of Co-Ni-Mo-O nanosheets using an electrochemical activation strategy [27], which increased the active sites and improved internal conductivity compared to fully polycrystalline or amorphous counterparts. Such optimized Co-Ni-Mo-O/NF-ER catalysts exhibited exceptional HER activity (42 mV @ 10 mA cm−2 and 400 h @ 250 mA cm−2). However, the conventional synthesis method (e.g., chemical vapor deposition) can destroy the low-crystalline matrix during long-term operation with high temperatures (>700 K) [28]. Therefore, the controllable synthesis of low-crystallinity materials remains a bottleneck for the development of scalable applications.
Based on the considerations described above, we report a rapid Joule heating method for the synthesis of low-crystallinity Ru-Mo oxide catalysts (Ru1Mo1-JH). Compared to the catalyst synthesized via conventional tube furnace heating methods (Ru1Mo1-LT), Ru1Mo1-JH features significantly reduced crystallinity and increased Mo4+ content, which strengthens metal-oxygen bonds and weakens Ru-H interactions. Such synergistic electronic modulation endows the catalyst with exceptional HER performance. The catalyst Ru1Mo1-JH exhibits ultralow overpotentials (25.08 mV and 120.52 mV at current densities of 10 and 100 mA cm−2, respectively) and excellent stability (100 h @ 100 mA cm−2) in 1 M KOH, which surpassed commercial Pt/C (42.27 mV and 149.27 mV) under identical conditions. Furthermore, an anion exchange membrane electrolyzer (AEM) assembled with an Ru1Mo1-JH electrode achieves 185 mA cm−2 at only 2.16 V with 100% Faradaic efficiency. These results highlight the potential of Ru1Mo1-JH as a cost-effective alternative to Pt/C for industrial applications.

2. Experimental

2.1. Synthesis of Ru1Mo1-JH Catalyst

This work successfully synthesized low-crystallinity Ru-Mo oxide catalysts through a two-step process that combines hydrothermal treatment and rapid Joule heating. To optimize the strength of the Ru-H bond by tailoring the electronic structure, we first investigated the impact of Mo species on the water splitting performance of Ru-based catalysts. Homogeneous solutions of the precursors, including RuCl3 (Aladdin Ltd., Shanghai, China) and MoCl5 (Aladdin Ltd., Shanghai, China) with different Ru/Mo ratios of 0:4, 1:3, 2:2, 3:1, and 4:0, were transferred to a reactor, in which hydrothermal treatment took place at 160 °C for 5 h. Subsequently, these catalysts were heated to 470 °C for 60 s using a rapid Joule heating device, denoted as Mo-JH, Ru1Mo3-JH, Ru1Mo1-JH, Ru3Mo1-JH, and Ru-JH, respectively. To further elucidate the role of low crystallinity, a parallel catalyst (Ru1Mo1-LT) was synthesized using the homogeneous solution with the RuCl3/MoCl5 ratio (1:1) and treated using a conventional thermal method at the same temperature (470 °C) for 1 h to ensure full crystallization. The following text will focus on the performance comparison of these catalysts and explore the improvement mechanism of the Ru1Mo1-JH.

2.2. Characterizations

X-ray diffraction (XRD) was employed on the Bruker D8 focus (Bruker AXS GmbH, Berlin, Germany) using Cu Kα radiation with a scan rate of 10°min−1. Transmission electron microscopy (TEM) was performed on an FEI Tecnai F20 microscope (FEI, Hillsboro, Oregon, OR, USA) at an acceleration voltage of 200 kV, yielding high-resolution transmission electron microscopy (HRTEM) and energy-dispersive X-ray spectroscopy (EDX) data via a JEOL JEM 2100 microscope (JEOL, Tokyo, Japan) at the same voltage. Scanning electron microscopy (SEM) images were obtained using a Hitachi S-4800 microscope (Hitachi, Nagano, Japan) at an acceleration voltage of 5–10 kV. The X-ray photoelectron spectroscopy (XPS) spectra were obtained using an ESCALAB 250Xi (Thermo Fisher Scientific, Waltham, MA, USA) with a monochromatic Al Kα source, and the spectra were analyzed using Avantage 6.9 software. Brunauer–Emmett–Teller (BET) measurements were performed on an ASAP2460 (Micromeritics Instrument Corporation, Norcross, Georgia, GA, USA) using nitrogen as the adsorption gas, which was degassed at 120 degrees Celsius for 8 h. Fourier-transform infrared spectroscopy (FTIR) analysis was employed on a Nicolet iS 10 (Thermo Fisher Scientific, Waltham, MA, USA) using a silicon crystal to obtain the absorption infrared spectrum of the samples.

2.3. Electrochemical Measurements

The electrochemical HER of catalysts was performed on a CHI 660E workstation (CH Instruments, Shanghai, China) using a typical three-electrode system in 1 M KOH (Sinopharm Chemical Reagent Co. Ltd., Shanghai, China) electrolyte. A graphite rod, Hg/HgO, and a catalyst-loaded glassy carbon electrode (GCE) (Shanghai Jingchong Electronic Technology Development Co. Ltd., Shanghai, China) served as the counter, reference, and working electrodes, respectively. The catalyst ink consisted of 5 mg of catalyst, 245 μL of ethanol (Sinopharm Chemical Reagent Co. Ltd., Shanghai, China), 245 μL of water, and 10 μL of 5 wt% Nafion solution (Shanghai Jingchong Electronic Technology Development Co. Ltd., Shanghai, China). Afterward, the GCE was dropped into 24 μL of as-prepared ink and dried before undergoing electrochemical tests. Linear sweep voltammetry (LSV) measurements were performed at a scan rate of 5 mV s−1 with 95% iR-compensation. The Nernst equation (ERHE = EHg/HgO + 0.0592pH + 0.098 V) was referenced for the conversion of the applied potentials. The Tafel slope was obtained via linear fit to LSV curves using the Tafel equation (η = a + b log(j)). Cyclic voltammetry (CV) tests with different scan rates (ranging from 20 to 100 mV s−1) were conducted to calculate the double-layer capacitance (Cdl) values of the catalysts. Electrochemical impedance spectroscopy (EIS) was performed at frequencies ranging from 105 Hz to 10−2 Hz. The Faradaic efficiency of Ru1Mo1-JH was calculated by measuring the amount of evolved H2 through a water drainage method at 185 mA cm−2, expressed as follows:
Faradaic efficiency = NG/NT
NG = V/Vm
NT = Q/(2 × F)
where NG and NT are the practical and theoretical number of hydrogen molecules, respectively; V is the measured volume of hydrogen; Vm is the molar volume of gas; and Q indicates the number of electrons transferred to generate hydrogen.

3. Results and Discussion

3.1. Structural Characterization of Ru1Mo1-JH

The Ru1Mo1 precursor was synthesized using the hydrothermal method to form powdered nanoparticles. As shown in its SEM image (Figure 1a), the nanoparticles exhibited a 3D-connected structure. After rapid Joule heating treatment at 470 °C for 60 s, the Ru1Mo1-JH catalyst exhibited significantly reduced agglomeration and uniform dispersion of nanoparticles (Figure 1b), facilitating the exposure of abundant active sites and thus enhancing catalytic activity, which benefited from the transient heating and rapid quenching characteristics of Joule heating technology [29]. Joule heating is capable of reaching high temperatures up to 3000 K within hundreds of milliseconds and cooling below 323 K within seconds. Based on rapid energy input and instantaneous quenching, ultrafast Joule heating technology can effectively circumvent the aggregation of nanoparticles and maintain the stability of catalytic particle dispersion [30]. In contrast, Ru1Mo1 was heated for a longer time to obtain Ru1Mo1-LT, which induced sintering, resulting in irregular blocks with smooth surfaces (Figure 1c). Such morphology reduced the number of active sites and diminished the catalytic performance. The TEM image (Figure 1d) revealed that the morphology of Ru1Mo1-JH nanoparticles was maintained after rapid Joule heating. The HRTEM images (Figure 1e,f) revealed that Ru1Mo1-JH had a large area of amorphous parts with interconnected channels and voids between nanoparticles, thus exposing more active sites. Such a structure was also conducive to electrolyte ion transport, intermediate activation, and the escape of gases produced by the reaction [31]. Region I in Figure 1f was selected for fast Fourier transform (FFT), which indicated that Ru1Mo1-JH was composed of randomly oriented nanoparticles (Figure 1g). Furthermore, the elemental mapping of Ru1Mo1-JH showed that the Mo, Ru, and O elements were uniformly distributed over the entire structure of Ru1Mo1-JH (Figure 1h). However, the TEM of Ru1Mo1-LT showed larger particle dimensions and increased agglomeration (Figure 1i), which limited the exposure of active sites. As shown in Figure 1j, Ru1Mo1-LT was distributed with obvious lattice fringes. It was initially determined that the crystallinity of the catalyst was relatively lower after rapid Joule heating, while a crystal material was formed after conventional thermal treatment. It should be noted that low-crystallinity materials have a high specific surface area, which can improve the efficiency of catalysts by providing fully exposed active sites. This finding was also confirmed via SEM in Figure 1b. According to the line intensity distribution map of area III in Figure 1k, the spacing of the fringes was 0.34 nm, which confirmed the complete crystallization of Ru1Mo1-LT. These results further provided strong evidence for the low crystallinity of Ru1Mo1-JH and the complete crystallization of Ru1Mo1-LT.
The phase composition and crystal structure of the samples were further analyzed via XRD. As shown in Figure 2a, XRD of as-prepared catalysts exhibited phase structures of RuO2 (PDF#21-1172), MoO2 (PDF#50-0739), and MoO3 (PDF#85-2405). The gradual enhancement of XRD diffraction peak intensities with prolonged heating time indicated an increase in crystallinity. The half-peak full widths (FWHM) of the strongest diffraction peaks for Ru1Mo1, Ru1Mo1-JH, and Ru1Mo1-LT were calculated as 0.60 Å−1, 0.40 Å−1, and 0.24 Å−1, respectively. According to previous studies, an FWHM range of 0.3–0.5 Å−1 was associated with amorphous structures, which confirmed the low crystallinity of Ru1Mo1-JH [32,33,34]. Quantitative phase analysis (Figure 2b) revealed that Joule heating in air resulted in Ru and Mo species existing in the form of oxides. XPS was employed to further probe the elemental composition and chemical state of Ru1Mo1-JH and Ru1Mo1-LT. The XPS full-spectrum scanning results (Figure 2c) confirmed the existence of Ru, Mo, and O in both samples. In the Mo 3d orbital spectra (Figure 2d), there was no significant difference in the binding energy of Mo6+ between the two catalysts. It should be noted that the peak of Mo4+ in Ru1Mo1-JH exhibited a 0.34 eV shift toward higher binding energy compared to Ru1Mo1-LT. Moreover, the peak integral area ratio of Mo6+/Mo4+ increased from 1.5 to 3.3, suggesting the progressive oxidation of Mo4+ to Mo6+. The Ru 3p spectra (Figure 2e) revealed a 0.20 eV positive shift in the Ru 3p1/2 peak for Ru1Mo1-JH, accompanied by a 3.95% increase in the Ru 3p3/2 peak area. According to the results described above, it could be concluded that MoO2, as a superior electron-accepting support, formed strong Ru-MoO2 interactions in Ru1Mo1-JH, allowing for significant electron transfer from Ru to MoO2 and eventually resulting in the redistribution of electrons among Ru sites [35]. These observations suggest that Mo4+ incorporation strengthened Ru-O-Mo interactions and reduced the d-electron density of Ru, which could lower the d-band center, thereby optimizing the electronic structure [36]. For the O 1s in Figure 2f, Ru1Mo1-JH exhibited an approximately 0.45 eV positive shift compared to Ru1Mo1-LT, with a higher proportion of O2− in the metal oxide, which proved that enhanced metal–oxygen bonding occurred. Notably, the contents of OH and H2O on the surface of Ru1Mo1-LT were 54.11%, whereas the OH and H2O on the surface of Ru1Mo1-JH accounted for only 23.89%. According to previous studies, low crystallinity formed by rapid heating could optimize the electronic structure of Ru/Mo-based catalysts and reduce the d-electron density of Ru, weakening the adsorption strength of Ru for H and ultimately affecting HER performance [37]. As shown in Figure 2g, a band at 1958 cm−1 on the spectra of Ru1Mo1-LT was observed and attributed to Ru-H species [38,39], while a band at 1956 cm−1 on the spectra of Ru1Mo1-JH was much weaker, which further confirmed that low crystallinity contributed to weakening Ru-H interactions. In addition, the BET surface area of Ru1Mo1-JH was found to be 69.01 m2/g, which exceeded that of other reported catalysts, indicating that Ru1Mo1-JH was characterized by abundant active sites.
To investigate the morphological and electronic structures of Ru1Mo1-JH after water splitting, XRD and XPS were carried out. The XRD patterns showed that the characteristic peak intensities of Ru1Mo1-JH had slightly decreased after the reaction, but the phase structures of RuO2, MoO2, and MoO3 still existed (Figure 3a). Typically, in order to collect Ru1Mo1-JH after the reaction more effectively, the powder of Ru1Mo1-JH was dripped onto nickel foam (NF), and then an LSV scan was carried out before XRD was conducted. The three dominant peaks originated from the NF substrate. Moreover, XPS was employed to investigate the elemental composition and electronic structure of the material after the reaction. The XPS full-spectrum scanning results confirmed the existence of Ru, Mo, and O in Ru1Mo1-JH both before and after the reaction (Figure 3b). As shown in Figure 3c, the binding energies of the Mo4+ and Mo6+ peaks in Ru1Mo1-JH revealed positive shifts of 0.17 eV and 0.24 eV, respectively. The Ru 3p spectra revealed negligible binding energy shifts for Ru 3p1/2 and Ru 3p3/2 peaks in Ru1Mo1-JH after the reaction (Figure 3d). The elements Ru and Mo in Ru1Mo1-JH existed in the same valence state as Mo4+, Mo6+, and Ru4+, both before and after the reaction. However, the O 1s spectrum exhibited relatively significant changes after the reaction. Ru1Mo1-JH after the reaction showed a positive shift of approximately 0.40 eV compared to before the reaction (Figure 3e), indicating enhanced metal–oxygen bonding. Additionally, the content of OH and H2O on the surface of Ru1Mo1-JH increased, which might be attributed to the influence of alkaline solutions during the water splitting process. These results confirmed the excellent structural stability of Ru1Mo1-JH.

3.2. HER Electrochemical Performance of Ru1Mo1-JH

A conventional three-electrode cell configuration was used to comprehensively evaluate the electrocatalytic activity of Ru1Mo1-JH in a 1 M KOH solution at 25 °C, with other as-prepared catalysts measured for comparison. To confirm the optimum content of Mo, HER activities of catalysts synthesized with homogeneous solutions of different RuCl3/MoCl5 ratios were explored. As shown in Figure 4a, the overpotential of Ru1Mo1-JH at 10 mA cm−2 was only 25.08 mV, which was much lower than that of the contrasted samples. Typically, the Tafel slope of Ru1Mo1-JH (48.89 mV dec−1) surpassed that of Mo-JH (99.15 mV dec−1), Ru1Mo3-JH (65.42 mV dec−1), Ru3Mo1-JH (242.30 mV dec−1), and Ru-JH (274.97 mV dec−1), implying the superior catalytic kinetics of Ru1Mo1-JH (Figure 4b,c). Furthermore, the Cdl value of Ru1Mo1-JH (83.09 mF cm−2) was significantly higher than that of Mo-JH (0.30 mF cm−2), Ru1Mo3-JH (11.10 mF cm−2), Ru3Mo1-JH (8.89 mF cm−2), and Ru-JH (2.20 mF cm−2), which revealed that Ru1Mo1-JH was characterized by a large electrochemically active area (Figure 4d). To investigate the electron transfer ability and electrical conductivity of catalysts, EIS measurements were carried out with the corresponding equivalent circuit model, wherein Rs and Rct represented the solution resistance and charge transfer resistance, respectively. As shown in Figure 4e,f, Ru1Mo1-JH displayed the lowest Rct of 14.48 Ω compared with other reference catalysts, indicating its rapid electron transport and high electrical conductivity. These results confirmed the successful optimization of the electronic structure of Ru1Mo1-JH.
The precursors of Ru1Mo1, after activation at 470 °C for different amounts of time (0 s, 60 s, and 1 h), were measured to explore the role of low crystallinity, along with the control groups, Mo-JH, Ru-JH, and commercial Pt/C. As shown in Figure 5a, Ru1Mo1-JH displayed overpotentials of 25.08 mV and 120.52 mV to deliver current densities of 10 mA cm−2 and 100 mA cm−2, respectively, outperforming commercial Pt/C (the overpotentials at 10 mA cm−2 and 100 mA cm−2 were 42.27 mV and 149.27 mV, respectively). Moreover, as shown in Figure 5b, the Tafel slope of Ru1Mo1-JH (49.89 mV dec−1) was also lower than that of Mo-JH (99.15 mV dec−1), Ru-JH (274.97 mV dec−1), Ru1Mo1 (99.40 mV dec−1), Ru1Mo1-LT (173.04 mV dec−1), and Pt/C (88.81 mV dec−1), indicating its rapid reaction kinetics following the Volmer–Heyrovsky mechanism (Volmer step: H2O + e + * → *H + OH; Heyrovsky step: *H +H2O + e → H2 + OH + *) [40,41]. Based on the comparison of the catalysts in Figure 5c, it can be seen that the Mo species significantly enhanced catalytic activity, which further optimized HER activity through the lowering of crystallinity achieved via Joule heating. Notably, Ru1Mo1-LT, synthesized through conventional thermal treatment, exhibited a high overpotential of 253.01 mV to deliver 10 mA cm−2, which was attributed to the blockage of active sites. Meanwhile, a large Tafel slope of Ru1Mo1-LT (173.04 mV dec−1) indicated its high kinetic barrier. This might be due to the excessive adsorption capacity of Ru1Mo1-LT for OH and H2O intermediates, which significantly hindered the performance of HER activity at high current densities [42]. These results showcased the critical role of crystallinity modulation in enhancing HER performance.
As shown in Figure 5d, Ru1Mo1-JH showed a much lower Rct value of 2.2 Ω (Figure 4d) compared with Pt/C (24.57 Ω), Ru-JH (547.3 Ω), Mo-JH (4826 Ω), Ru1Mo1 (528.7 Ω), and Ru1Mo1-LT (481.9 Ω), which was consistent with its low overpotential and favorable kinetics. In contrast, Ru1Mo1-LT showed a significantly higher Rct than that of Ru1Mo1-JH, highlighting the improved charge transport enabled by low crystallinity [43]. Unsurprisingly, Ru1Mo1-JH possessed a high Cdl value of 83.09 mF cm−2 (Figure 5e), which indicated that Ru1Mo1-JH had a larger electrochemically active area and could expose a greater number of active sites compared with the crystalline Ru1Mo1-LT. Furthermore, the durability of Ru1Mo1-JH was evaluated via a chronopotentiometry test (100 mA cm−2). In addition, the superior alkaline HER performance of as-prepared Ru1Mo1-JH was retained even after 100 h of operation at 100 mA cm−2 without obvious decay (Figure 5f), indicating its high stability for potential industrial applications. According to XPS analysis, the interactions of Ru-Mo-O in Ru1Mo1-JH were enhanced, making it less likely for active site loss to occur when the catalyst exerts its catalytic effect, resulting in excellent stability. In order to clarify the excellent HER performance of Ru1Mo1-JH, the electrocatalytic performances of as-prepared Ru1Mo1-JH and recently reported state-of-the-art HER catalysts in alkaline solution are summarized in Figure 5g. It should be noted that the performance of the Ru-Mo oxide low-crystallinity catalyst prepared via Joule heating in this study outperformed most other Mo-based and Ru-based electrocatalysts, proving that Ru1Mo1-JH exhibited excellent HER performance.

3.3. Practical Water Splitting

To evaluate the practical HER activity of Ru1Mo1-JH, an AEM electrolyzer coupled with a quantification system for H2 was constructed to simulate industrial hydrogen production. As illustrated in Figure 6a, the AEM electrolyzer was equipped with Ru1Mo1-JH loading on nickel foam (NF) and RuO2 as the cathode and anode, respectively, which were separated by an anion exchange membrane in a 1 M KOH electrolyte for overall water splitting. First, the actual volume of H2 produced by Ru1Mo1-JH//RuO2 under a constant current density of 100 mA cm−2 was measured (Figure 6c). A comparative analysis with the theoretical volume of H2 in Figure 6b confirmed that the Faraday efficiency of the Ru1Mo1-JH catalyst could reach 100%. Remarkably, the AEM electrolyzer composed of Ru1Mo1-JH and RuO2 required only 2.16 V to operate at 185 mA cm−2 (Figure 6d), which was lower than that of Pt/C//RuO2. These results demonstrated the exceptional HER activity and industrial viability of Ru1Mo1-JH in practical electrolysis systems.

4. Conclusions

In summary, this work reports a rapid Joule heating method for synthesizing low-crystallinity Ru-Mo oxide catalysts featuring an internal porous structure that maximizes active site exposure. The optimized Ru1Mo1-JH catalyst exhibits excellent HER performance in 1 M KOH, achieving a low overpotential of 25.08 mV at 10 mA cm−2 and a Tafel slope of 48.89 mV dec−1, which is indicative of its rapid reaction kinetics. The high value of Cdl (83.09 mF cm−2) and minimal Rct (14.48 Ω) further confirm Ru1Mo1-JH has a large electrochemically active surface area and efficient interfacial charge transfer. Compared with the fully crystallized Ru1Mo1-LT, the results indicate that low crystallinity plays a critical role in weakening Ru-H bond strength and optimizing hydrogen intermediate adsorption. Furthermore, Ru1Mo1-JH achieved 100% Faradaic efficiency and required only 2.16 V to deliver 185 mA cm−2 in an AEM electrolyzer. All these results position Ru1Mo1-JH as a highly promising and cost-effective catalyst for industrial hydrogen production.

Author Contributions

T.S.: writing—original draft, investigation. X.H.: investigation. Z.Z.: investigation. Z.L.: investigation. K.H.: investigation. X.M.: writing—review and editing, supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (Grant No: 22408356), the Shandong Provincial Natural Science Foundation (Grant No.: ZR2024MB052), the Qingdao Natural Science Foundation (Grant No.: 24-4-4-zrjj-195-jch), and the Dean/Opening Project of Guangxi Key Laboratory of Petrochemical Resource Processing and Process Intensification Technology (No.: 2023K001, 2024K006).

Data Availability Statement

The authors declare that the data supporting the findings of this study are available in the paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Hosseini, S.E.; Wahid, M.A. Hydrogen production from renewable and sustainable energy resources: Promising green energy carrier for clean development. Renew. Sust. Energ Rev. 2016, 57, 850–866. [Google Scholar] [CrossRef]
  2. Zhang, E.; Song, W. Review-Self-Supporting Electrocatalysts for HER in Alkaline Water Electrolysis. J. Electrochem. Soc. 2024, 171, 052503. [Google Scholar] [CrossRef]
  3. Xiong, H.; Zhuang, R.; Cheng, B.; Liu, D.; Du, Y.; Wang, H.; Liu, Y.; Xu, F.; Wang, H. Self-Supported Metallic Alkaline Hydrogen Evolution Electrocatalysts Tolerant for Ampere-Level Current Densities. Adv. Energy Mater. 2025, 15, 2404077. [Google Scholar] [CrossRef]
  4. Huang, L.; Wei, M.; Jadoon, S.; Ali, A.; Xia, B.Y. Well-connection of micro-platinum and cobalt oxide flower array with optimized water dissociation and hydrogen recombination for efficient overall water splitting. Chem. Eng. J. 2020, 398, 125669. [Google Scholar] [CrossRef]
  5. Liu, Y.; Cheng, L.; Zhou, S.; Niu, C.; Isimjan, T.T.; Yang, X. Electronic regulation of hcp-Ru by d-d orbital coupling for robust electrocatalytic hydrogen oxidation in alkaline electrolytes. J. Colloid Interface Sci. 2025, 677, 997–1004. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, Q.; Lao, M.; Yu, Y.; Ma, M.; Li, M.; Fei, Z.; Dyson, P.J.; Wang, S.; Min, D. Manipulation of the Electronic Structure of Ruthenium Nanoclusters by Ni-N4 Sites Enhances the Alkaline Hydrogen Evolution Reaction. Adv. Funct. Mater. 2025, 35, 2416071. [Google Scholar] [CrossRef]
  7. Wang, L.; Liu, Y.; Chen, Z.; Dai, Q.; Dong, C.-L.; Yang, B.; Li, Z.; Hu, X.; Lei, L.; Hou, Y. Theory-guided design of electron-deficient ruthenium cluster for ampere-level current density electrochemical hydrogen evolution. Nano Energy 2023, 115, 108694. [Google Scholar] [CrossRef]
  8. Wang, L.; Ma, M.; Zhang, C.; Chang, H.-H.; Zhang, Y.; Li, L.; Chen, H.-Y.; Peng, S. Manipulating the Microenvironment of Single Atoms by Switching Support Crystallinity for Industrial Hydrogen Evolution. Angew. Chem. Int. Edit. 2024, 63, e202317220. [Google Scholar] [CrossRef]
  9. Xu, Y.; Yang, C.; Deng, Q.; Zhou, Y.; Mao, C.; Song, Y.; Zhu, M.; Zhang, Y. Bi-Porphyrins MOF with confinement and ion-attracting effects in concert with RuO2-doped CNT as efficient electrocatalysts for HER in acidic and alkaline media. Appl. Surf. Sci. 2023, 612, 155870. [Google Scholar] [CrossRef]
  10. Liu, Z.; Jiang, J.; Liu, Y.; Huang, G.; Yuan, S.; Li, X.; Li, N. Boosted hydrogen evolution reaction based on synergistic effect of RuO2@MoS2 hybrid electrocatalyst. Appl. Surf. Sci. 2021, 538, 148019. [Google Scholar] [CrossRef]
  11. Sun, J.; Zhao, Z.; Li, Z.; Zhang, Z.; Zhang, R.; Meng, X. Ultrafast carbothermal shocking fabrication of cation vacancy-rich Mo doped Ru nanoparticles on carbon nanotubes for high-performance water/seawater electrolysis. J. Mater. Chem. A 2023, 11, 22430–22440. [Google Scholar] [CrossRef]
  12. Kandel, M.R.; Pan, U.N.; Paudel, D.R.; Dhakal, P.P.; Kim, N.H.; Lee, J.H. Hybridized bimetallic phosphides of Ni–Mo, Co–Mo, and Co–Ni in a single ultrathin-3D-nanosheets for efficient HER and OER in alkaline media. Compos. Part B Eng. 2022, 239, 109992. [Google Scholar] [CrossRef]
  13. Li, W.; Liu, R.; Yu, G.; Chen, X.; Yan, S.; Ren, S.; Chen, J.; Chen, W.; Wang, C.; Lu, X. Rationally Construction of Mn-Doped RuO2 Nanofibers for High-Activity and Stable Alkaline Ampere-Level Current Density Overall Water Splitting. Small 2024, 20, 2307164. [Google Scholar] [CrossRef] [PubMed]
  14. Zhang, Y.; Lu, R.; Wang, C.; Zhao, Y.; Qi, L. Electronic and Vacancy Engineering of Mo-RuCoOx Nanoarrays for High-Efficiency Water Splitting. Adv. Funct. Mater. 2023, 33, 2303073. [Google Scholar] [CrossRef]
  15. Zhao, Z.; Sun, J.P.; Li, X.; Li, X.; Zhang, Z.; Meng, X. Joule heating synthesis of NiFe alloy/MoO2 and in-situ transformed (Ni,Fe) OOH/MoO2 heterostructure as effective complementary electrocatalysts for overall splitting in alkaline seawater. Appl. Catal. B-Environ. Energy 2024, 340, 123277. [Google Scholar] [CrossRef]
  16. Du, W.; Shi, Y.M.; Zhou, W.; Yu, Y.; Zhang, B. Unveiling the In Situ Dissolution and Polymerization of Mo in Ni4Mo Alloy for Promoting the Hydrogen Evolution Reaction. Angew. Chem. Int. Edit. 2021, 60, 7051–7055. [Google Scholar] [CrossRef]
  17. Jakšić, M.M. Electrocatalysis of hydrogen evolution in the light of the brewer—Engel theory for bonding in metals and intermetallic phases. Electrochim. Acta 1984, 29, 1539–1550. [Google Scholar] [CrossRef]
  18. Fan, L.; Li, Q.; Wang, D.; Meng, T.; Yan, M.; Xing, Z.; Wang, E.; Yang, X. Electrospun Ru-RuO2/MoO3 carbon nanorods with multi-active components: A Pt-like catalyst for the hydrogen evolution reaction. Chem. Commun. 2020, 56, 739–742. [Google Scholar] [CrossRef]
  19. Zhang, J.; Wang, T.; Liu, P.; Liao, Z.; Liu, S.; Zhuang, X.; Chen, M.; Zschech, E.; Feng, X. Efficient hydrogen production on MoNi4 electrocatalysts with fast water dissociation kinetics. Nat. Commun. 2017, 8, 15437. [Google Scholar] [CrossRef]
  20. Sun, Y.K.; Sun, W.Y.; Li, G.C.; Wang, L.; Huang, J.; Meng, A.; Li, Z. Interface, vacancy, and morphology engineering synergistically improve In2S3@Cu2S electrocatalytic performance for pH-universal HER. J. Mater. Chem. A 2023, 11, 2262–2272. [Google Scholar] [CrossRef]
  21. Yin, J.W.; Lu, T.Y.; Li, J.; Liu, J.; Lin, Y.; Sun, D.; Xu, L.; Zhao, Q.; Pang, H.; Zhang, S.; et al. Synergistic Alkaline Hydrogen Evolution Catalysis over MoC Triggered by Doping Single Ru Atoms. Adv. Funct. Mater. 2025, 35, 2417034. [Google Scholar] [CrossRef]
  22. Xiang, L.J.; Li, N.; Zhao, L.; Wang, K.; Pang, B.; Liu, Z.; Guo, J. Boosting alkaline hydrogen evolution via in-plane heterostructure construction with Ultra-Exposed heterointerfaces. Chem. Eng. J. 2024, 499, 155833. [Google Scholar] [CrossRef]
  23. Peng, C.L.; Zhao, W.P.; Li, Z.X.; Kuang, Z.; Cheng, G.; Miller, J.T.; Sun, S.; Chen, H. Eutectic molten salt assisted synthesis of highly defective and flexible ruthenium oxide for efficient overall water splitting. Chem. Eng. J. 2021, 425, 131707. [Google Scholar] [CrossRef]
  24. Gao, X.Y.; Li, B.; Sun, X.Z.; Wu, B.; Hu, Y.; Ning, Z.; Li, J.; Wang, N. Engineering heterostructure and crystallinity of Ru/RuS2 nanoparticle composited with N-doped graphene as electrocatalysts for alkaline hydrogen evolution. Chin. Chem. Lett. 2021, 32, 3591–3595. [Google Scholar] [CrossRef]
  25. Zai, S.F.; Dong, A.Q.; Li, J.; Wen, Z.; Yang, C.C.; Jiang, Q.Q. Low-crystallinity mesoporous NiGaFe hydroxide nanosheets on macroporous Ni foam for high-efficiency oxygen evolution electrocatalysis. J. Mater. Chem. A 2021, 9, 6223–6231. [Google Scholar] [CrossRef]
  26. Qin, Y.Q.; Du, K.C.; Wang, Y.Q.; Jiang, S.; Miao, L.; Lu, M.; Wang, P. Support Regulation Strategy Synergistic with Ru Single-Atom Catalyst for Energy-Saving Hydrogen Production Assisted by Green Electrosynthesis of Azotetrazolate. Inorg. Chem. 2025, 64, 4544–4554. [Google Scholar] [CrossRef]
  27. Guo, K.L.; Jia, J.Z.; Wu, H.X. Electrochemical activation-induced Co-Ni-Mo-O nanosheets with low crystallinity and abundant active sites for efficient and ultra-stable hydrogen generation. Dalton Trans. 2023, 52, 9138–9147. [Google Scholar] [CrossRef] [PubMed]
  28. Yang, Y.Q.; Pang, D.W.; Wang, C.J.; Fu, Z.; Liu, N.; Liu, J.; Wu, H.; Jia, B.; Guo, Z.; Fan, X.; et al. Vacancy and Dopant Co-Constructed Active Microregion in Ru-MoO3-x/Mo2AlB2 for Enhanced Acidic Hydrogen Evolution. Angew. Chem. Int. Edit. 2025, 64, e202504084. [Google Scholar] [CrossRef]
  29. Yang, L.; Pang, H.; Ren, W.; Wang, X.; Wei, Y.; Lu, P.; Chen, J.; Tian, W.; Huang, M.; Wang, H. Joule Heating Driven Graphitization Regulation and Ni Single-Atom Modification in Hard Carbon for Low-Voltage and High-Rate Potassium-Ion Storage. Adv. Funct. Mater. 2025, e16237. [Google Scholar] [CrossRef]
  30. Zhao, Z.; Wu, T.; Li, X.; Chena, Y.; Meng, X. Progress and perspectives of rapid Joule heating for the preparation of highly efficient catalysts. Mater. Horiz. 2025, 12, 734–759. [Google Scholar] [CrossRef]
  31. Ji, D.X.; Peng, S.J.; Lu, J.; Li, L.; Yang, S.; Yang, G.; Qin, X.; Srinivasand, X.; Ramakrishna, S. Design and synthesis of porous channel-rich carbon nanofibers for self-standing oxygen reduction reaction and hydrogen evolution reaction bifunctional catalysts in alkaline medium. J. Mater. Chem. A 2017, 5, 7507–7515. [Google Scholar] [CrossRef]
  32. Inoue, A.; Kato, A.; Zhang, T.; Kim, S.G.; Masumoto, T. Mg-Cu-Y Amorphous-Alloys with High Mechanical Strengths Produced by a Metallic Mold Casting Method. Mater. Trans. JIM 1991, 32, 609–616. [Google Scholar] [CrossRef]
  33. Senkov, O.N.; Scott, J.M.; Miracle, D.B. Development of low density Ca-Mg-Al-based bulk metallic glasses. Mater. Trans. 2007, 48, 1610–1616. [Google Scholar] [CrossRef]
  34. Ma, D.; Stoica, A.D.; Wang, X.L. Power-law scaling and fractal nature of medium-range order in metallic glasses. Nat. Mater. 2009, 8, 30–34. [Google Scholar] [CrossRef] [PubMed]
  35. Ren, P.Y.; Wang, W.; Hong, Y.L.; Zheng, Y.; Ren, X.; Du, W.; Chen, Y.; Tan, S.; Zhang, W. Ru Nanoparticle-Anchored MoO2 Nanoflowers as Electrocatalysts for the Hydrogen Evolution Reaction. ACS Appl. Nano Mater. 2023, 6, 13926–13934. [Google Scholar] [CrossRef]
  36. Wang, L.; Chen, Y.; Liu, Y.N.; Dai, Q.; Chen, Z.; Yang, X.; Luo, Y.; Li, Z.; Yang, B.; Zheng, M.; et al. Electron Redistribution of Ru Site on MoO2@NiMoO4 Support for Efficient Ampere-Level Current Density Electrolysis of Alkaline Seawater. Small 2024, 20, 2311477. [Google Scholar] [CrossRef]
  37. Chang, Y.N.; Kong, L.Y.; Xu, D.D.; Lu, X.; Wang, S.; Li, Y.; Bao, J.; Wang, Y.; Liu, Y. Mo Migration-Induced Crystalline to Amorphous Conversion and Formation of RuMo/NiMoO4 Heterogeneous Nanoarray for Hydrazine-Assisted Water Splitting at Large Current Density. Angew. Chem. Int. Edit. 2025, 64, e202414234. [Google Scholar] [CrossRef]
  38. Liu, J.; Liu, W.; Li, Z.; Wang, J.; Fan, R.; Liu, C.; Gao, B.; Zou, Z.; Li, Z. Mitigating Hydrogen Poisoning for Robust Ammonia-to-Hydrogen Conversion over Photothermal Catalysts. ACS Catal. 2025, 15, 10470–10479. [Google Scholar] [CrossRef]
  39. Zhu, S.; Qin, X.; Xiao, F.; Yang, S.; Xu, Y.; Tan, Z.; Li, J.; Yan, J.; Chen, Q.; Chen, M.; et al. The role of ruthenium in improving the kinetics of hydrogen oxidation and evolution reactions of platinum. Nat. Catal. 2021, 4, 711–718. [Google Scholar] [CrossRef]
  40. Zhao, G.Q.; Rui, K.; Dou, S.X.; Sun, W. Heterostructures for Electrochemical Hydrogen Evolution Reaction: A Review. Adv. Funct. Mater. 2018, 28, 1803291. [Google Scholar] [CrossRef]
  41. Cao, D.; Xu, H.X.; Li, H.L.; Feng, C.; Zeng, J.; Cheng, D. Volcano-type relationship between oxidation states and catalytic activity of single-atom catalysts towards hydrogen evolution. Nat. Commun. 2022, 13, 5843. [Google Scholar] [CrossRef] [PubMed]
  42. Hou, L.Q.; Li, C.; Jang, H.; Kim, M.K.; Jiang, J.-Z.; Cho, J.; Liu, S.; Liu, X. Rationally Designed Mo/Ru-Based Multi-Site Heterogeneous Electrocatalyst for Accelerated Alkaline Hydrogen Evolution Reaction. Adv. Mater. 2024, 36, 2410039. [Google Scholar] [CrossRef] [PubMed]
  43. Lin, S.Q.; Zhu, L.P.; Tang, Z.; Wang, Z.L. Spin-selected electron transfer in liquid-solid contact electrification. Nat. Commun. 2022, 13, 5230. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Components and structure of the Ru1Mo1-LT counterpart and Ru1Mo1-JH. SEM images of (a) Ru1Mo1; (b) Ru1Mo1-JH; and (c) Ru1Mo1-LT. (df) TEM images of Ru1Mo1-JH. (g) FFT image corresponding to Area I. (h) EDX spectra of Ru1Mo1-JH. (i,j) TEM images of Ru1Mo1-LT. (k) HRTEM image of Ru1Mo1-LT and FFT image corresponding to Area II.
Figure 1. Components and structure of the Ru1Mo1-LT counterpart and Ru1Mo1-JH. SEM images of (a) Ru1Mo1; (b) Ru1Mo1-JH; and (c) Ru1Mo1-LT. (df) TEM images of Ru1Mo1-JH. (g) FFT image corresponding to Area I. (h) EDX spectra of Ru1Mo1-JH. (i,j) TEM images of Ru1Mo1-LT. (k) HRTEM image of Ru1Mo1-LT and FFT image corresponding to Area II.
Processes 13 02594 g001
Figure 2. Surface chemical state and coordination environment analysis of Ru1Mo1-JH and Ru1Mo1-LT. (a) XRD spectrum. (b) Mass fraction of phase. (c) XPS score. XPS spectra of (d) Mo 3d; (e) Ru 3p; (f) O 1s. (g) FITR spectra. (h) Comparison of the BET surface area between Ru1Mo1-JH and other reported catalysts.
Figure 2. Surface chemical state and coordination environment analysis of Ru1Mo1-JH and Ru1Mo1-LT. (a) XRD spectrum. (b) Mass fraction of phase. (c) XPS score. XPS spectra of (d) Mo 3d; (e) Ru 3p; (f) O 1s. (g) FITR spectra. (h) Comparison of the BET surface area between Ru1Mo1-JH and other reported catalysts.
Processes 13 02594 g002
Figure 3. Surface chemical state and coordination environment analysis of Ru1Mo1-JH before and after the water splitting reaction. (a) XRD spectra. (b) XPS score. XPS spectra of (c) Mo 3d, (d) Ru 3p, and (e) O 1 s.
Figure 3. Surface chemical state and coordination environment analysis of Ru1Mo1-JH before and after the water splitting reaction. (a) XRD spectra. (b) XPS score. XPS spectra of (c) Mo 3d, (d) Ru 3p, and (e) O 1 s.
Processes 13 02594 g003
Figure 4. Electrocatalytic HER performance test of catalysts with varying Ru/Mo ratios in 1.0 M KOH solution. (a) LSV curves. (b) Tafel plots. (c) Comparison of overpotentials for current densities of 10 mA cm−2 with Tafel slopes in as-prepared electrocatalysts. (d) Cdl values obtained from the linear fitting of capacitance currents and scan rates. (e) EIS Nyquist plots. (f) Rct values of as-prepared electrocatalysts.
Figure 4. Electrocatalytic HER performance test of catalysts with varying Ru/Mo ratios in 1.0 M KOH solution. (a) LSV curves. (b) Tafel plots. (c) Comparison of overpotentials for current densities of 10 mA cm−2 with Tafel slopes in as-prepared electrocatalysts. (d) Cdl values obtained from the linear fitting of capacitance currents and scan rates. (e) EIS Nyquist plots. (f) Rct values of as-prepared electrocatalysts.
Processes 13 02594 g004
Figure 5. Electrocatalytic HER performance test of catalysts in 1.0 M KOH solution. (a) LSV curves. (b) Tafel plots. (c) Comparison of overpotentials for current density of 10 mA cm−2 with Tafel slopes in as-prepared electrocatalysts. (d) EIS Nyquist plots. (e) Cdl values obtained from the linear fitting of capacitance currents and scan rates. (f) Chronopotentiometry measurement without iR compensation of Ru1Mo1-JH at 100 mA cm−2. (g) Comparison of overpotentials for current densities of 10 mA cm−2 and 100 mA cm−2 in different works.
Figure 5. Electrocatalytic HER performance test of catalysts in 1.0 M KOH solution. (a) LSV curves. (b) Tafel plots. (c) Comparison of overpotentials for current density of 10 mA cm−2 with Tafel slopes in as-prepared electrocatalysts. (d) EIS Nyquist plots. (e) Cdl values obtained from the linear fitting of capacitance currents and scan rates. (f) Chronopotentiometry measurement without iR compensation of Ru1Mo1-JH at 100 mA cm−2. (g) Comparison of overpotentials for current densities of 10 mA cm−2 and 100 mA cm−2 in different works.
Processes 13 02594 g005
Figure 6. AEM device performance based on Ru1Mo1-JH as the cathode. (a) Schematic picture of AEM. (b) Faraday efficiency. (c) Process of the Faraday efficiency test. (d) LSV curves of the electrolyzer.
Figure 6. AEM device performance based on Ru1Mo1-JH as the cathode. (a) Schematic picture of AEM. (b) Faraday efficiency. (c) Process of the Faraday efficiency test. (d) LSV curves of the electrolyzer.
Processes 13 02594 g006
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

Shi, T.; Huang, X.; Zhao, Z.; Li, Z.; Huang, K.; Meng, X. Rapid Joule-Heating Synthesis of Efficient Low-Crystallinity Ru-Mo Oxide Catalysts for Alkaline Hydrogen Evolution Reaction. Processes 2025, 13, 2594. https://doi.org/10.3390/pr13082594

AMA Style

Shi T, Huang X, Zhao Z, Li Z, Huang K, Meng X. Rapid Joule-Heating Synthesis of Efficient Low-Crystallinity Ru-Mo Oxide Catalysts for Alkaline Hydrogen Evolution Reaction. Processes. 2025; 13(8):2594. https://doi.org/10.3390/pr13082594

Chicago/Turabian Style

Shi, Tao, Xiaoling Huang, Zhan Zhao, Zizhen Li, Kelei Huang, and Xiangchao Meng. 2025. "Rapid Joule-Heating Synthesis of Efficient Low-Crystallinity Ru-Mo Oxide Catalysts for Alkaline Hydrogen Evolution Reaction" Processes 13, no. 8: 2594. https://doi.org/10.3390/pr13082594

APA Style

Shi, T., Huang, X., Zhao, Z., Li, Z., Huang, K., & Meng, X. (2025). Rapid Joule-Heating Synthesis of Efficient Low-Crystallinity Ru-Mo Oxide Catalysts for Alkaline Hydrogen Evolution Reaction. Processes, 13(8), 2594. https://doi.org/10.3390/pr13082594

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