Next Article in Journal
A Phosphorylated Dendrimer-Supported Biomass-Derived Magnetic Nanoparticle Adsorbent for Efficient Uranium Removal
Previous Article in Journal
Interfacing Langmuir–Blodgett and Pickering Emulsions for the Synthesis of 2D Nanostructured Films: Applications in Copper Ion Adsorption
Previous Article in Special Issue
Synthesis of ZnFe2O4 Nanospheres with Tunable Morphology for Lithium Storage
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Porous Ruthenium–Tungsten–Zinc Nanocages for Efficient Electrocatalytic Hydrogen Oxidation Reaction in Alkali

Anhui Province Key Laboratory of Value-Added Catalytic Conversion and Reaction Engineering, Anhui Province Engineering Research Center of Flexible and Intelligent Materials, School of Chemistry and Chemical Engineering, Hefei University of Technology, Hefei 230009, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Nanomaterials 2024, 14(9), 808; https://doi.org/10.3390/nano14090808
Submission received: 14 April 2024 / Revised: 29 April 2024 / Accepted: 2 May 2024 / Published: 6 May 2024

Abstract

:
With the rapid development of anion exchange membrane technology and the availability of high-performance non-noble metal cathode catalysts in alkaline media, the commercialization of anion exchange membrane fuel cells has become feasible. Currently, anode materials for alkaline anion-exchange membrane fuel cells still rely on platinum-based catalysts, posing a challenge to the development of efficient low-Pt or Pt-free catalysts. Low-cost ruthenium-based anodes are being considered as alternatives to platinum. However, they still suffer from stability issues and strong oxophilicity. Here, we employ a metal–organic framework compound as a template to construct three-dimensional porous ruthenium–tungsten–zinc nanocages via solvothermal and high-temperature pyrolysis methods. The experimental results demonstrate that this porous ruthenium–tungsten–zinc nanocage with an electrochemical surface area of 116 m2 g−1 exhibits excellent catalytic activity for hydrogen oxidation reaction in alkali, with a kinetic density 1.82 times and a mass activity 8.18 times higher than that of commercial Pt/C, and a good catalytic stability, showing no obvious degradation of the current density after continuous operation for 10,000 s. These findings suggest that the developed catalyst holds promise for use in alkaline anion-exchange membrane fuel cells.

1. Introduction

In recent years, hydrogen energy has received extensive attention as a potential alter-native to fossil fuels due to its high energy density of up to 140 MJ Kg−1, as well as its clean and environmentally friendly characteristics [1]. Hydrogen fuel cells are known for their high energy conversion efficiency and zero-carbon emissions, making them one of the most competitive technologies for hydrogen utilization [2]. Fuel cells are classified into five categories based on the type of electrolyte and the operating conditions, including proton exchange membrane fuel cells (PEMFCs), anion exchange membrane fuel cells (AEMFC), phosphoric acid fuel cells, molten carbonate fuel cells, and solid oxide fuel cells [3]. So far, PEMFCs have been at the forefront of cutting-edge energy conversion technologies used in various applications, such as automobiles, communications, and aerospace. However, PEMFCs require platinum (Pt)-based electrocatalysts, acid-resistant electric pile hardware, and perfluorinated membranes, resulting in high production costs [4]. The AEMFC represents a highly efficient and environmentally friendly power generation technology. It is regarded as a new-generation energy power system due to its high cell efficiency, excellent catalyst stability, and flexible fuel selectivity. Anion-exchange membranes are seen as the next generation of cost-effective fuel cells because they can use efficient, non-precious metal catalysts in the cathodic oxygen reduction reaction, significantly reducing fuel cell costs [5]. However, the kinetics of Pt group metals required for anode hydrogen oxidation reactions (HORs) in an alkaline media are 2–3 orders of magnitude lower than those in an acidic media. Therefore, improving the performance of noble metal catalysts or developing efficient low-cost alkaline HOR catalysts remains a major challenge [6,7].
Ruthenium (Ru) possesses a similar hydrogen binding energy to Pt, but is much less expensive than Pt, making it a potential substitute to Pt for alkaline HORs [8,9]. However, Ru exhibits oxyphilic properties at anodic potentials above 0.1 V compared to a reversible hydrogen electrode (RHE), leading to a decreased electrochemical HOR performance [10]. Several strategies have recently been developed to enhance the performance of Ru-based catalysts, including compositional effects [11,12,13], carrier effects [14,15,16,17], size effects [18,19], and structural effects [20,21]. In the case of loaded catalysts, they are usually overlooked when the particles are loaded or embedded in the carrier, which may result in underutilization of the active sites [22,23].
Metal–organic frameworks (MOFs), three-dimensional ordered porous structure materials, have garnered significant attention due to their high specific surface area and controllable porous structure [24]. Furthermore, MOF structures offer the advantages of being easily designed and modified, enabling the direct synthesis of various functional materials derived from MOFs. MOF-derived catalysts often possess hollow structures, facilitating mass transfer and intermediate diffusion, thus improving catalytic performance [25,26,27,28]. For example, Ma et al. [29] used octahedral zeolitic imidazolate framework-8 (ZIF-8) as a template to construct sub-2 nm RuMo-anchored hollow carbon catalysts, exhibiting a high activity for alkaline HOR, with a mass activity of 3.83 A mgRu−1, which is 25 times higher than that of commercial Ru/C. Qiu et al. [30] reported a general method for synthesizing binary, ternary, and high-entropy nanoparticles using a 2D MOF-assisted pyrolysis-replacement alloying technique. The developed Co0.2Ru0.7Pt0.1/PNC nanosheets exhibited a mass activity as high as 1.84 A mg−1 at an overpotential of 50 mV. Therefore, constructing high surface area catalysts and improving their exposure rates are the key factors for enhancing the alkaline HOR performance.
Here, we synthesize a three-dimensional hollow porous ruthenium–tungsten–zinc (Ru-W-Zn-O) nanocage through a hydrothermal reaction and high-temperature pyrolysis processes. The optimized Ru1W0.14Zn1.47Oy catalyst exhibits an enhanced HOR performance in alkaline conditions compared to commercial Pt/C and Ru/C. Specifically, the kinetic current densities (jk) were 1.82 and 1.04 times higher than those of commercial Pt/C and Ru/C, respectively. After a constant voltage test for 10,000 s, the current density of Ru1W0.14Zn1.47Ox shows no significant decay. These fundings make Ru1W0.14Zn1.47Ox a promising candidate for alkaline AEMFC anode materials.

2. Materials and Methods

2.1. Materials

Zn (NO3)2·6H2O, cetyltrimethylammonium bromide (CTAB), potassium hydroxide (KOH), isopropanol, and methanol were purchased from Sinopharm (Shanghai, China), and 2-methyl imidazole was purchased from Beijing Jin Ming Biotechnology Co., Ltd. (Beijing, China). RuCl3·xH2O, Ru/C (20 wt%), and Na2WO4 were purchased from Macklin (Shanghai, China). Pt/C (20 wt%) was purchased from Suzhou Sinero Technology Co., Ltd. (Suzhou, China). Nafion®117 solution was purchased from Sigma-Aldrich (St. Louis, MO, USA) (~5% in a mixture of lower aliphatic alcohols and water, containing 15~20% water). All reagents were analytically pure and used as received without further purification.

2.2. Material Characterization

Transmission electron microscopy (TEM) and Field-Emission Scanning Electron Microscopy (SEM) images were obtained from a Hitachi H7700 and SU8020 (Tokyo, Japan) at an accelerating voltage of 100 KV. The data of high-revolution TEM images and EDS elemental mapping were obtained from a JEM-2100F (Tokyo, Japan). The catalyst solution was applied onto 300-mesh copper grids coated with formvar/carbon support film (Beijing Zhongjing Key Technology Co., Ltd., Beijing, China). The powder X-ray diffraction (XRD) patterns were obtained on Philips X’Pert Pro Super with a Cu Ka radiation source (λ = 1.541841 Å). X-ray photoelectron spectra (XPS) were measured using an Al Kα radiation source on a Thermo Fisher ESCALAB 250Xi (Waltham, MA, USA). The peak shifts caused by apparent charging were calibrated using the carbon C 1s peak set to 284.8 eV. All spectra were collected in ambient conditions. All electrochemical performances were conducted with an electrochemical workstation (Autolab, Metrohm, Herisau, Switzerland). The noble metal mass content of Ru in Ru1W0.14Zn1.47Ox and RuZnOx nanocages were determined by inductively coupled plasma-Mass Spectrometry (ICP-MS, Agilent Technologies 7500 series, Santa Clara, CA, USA).

2.3. Methods

2.3.1. Preparation of ZIF-8

In a typical preparation, 1.22 mmol of Zn (NO3)2·6H2O was dissolved in 12.5 mL of deionized water and stirred at 350 rpm. Separately, 70.64 mmol of 2-methyl imidazole and 1.22 mmol of CTAB were mixed with 87.5 mL deionized water. The Zn (NO3)2·6H2O solution was then quickly injected into the second solution and stirred continuously for 5 min. The above operations were conducted at room temperature, followed by standing for several hours to obtain a uniformly dispersed ZIF-8 solution.

2.3.2. Synthesis of Ru1W0.14Zn1.47Ox Nanocage

As is typical for the synthesis of nanocages, 8 mL of the white ZIF-8 solution was taken for centrifugation, and the precipitate was rinsed with methanol several times. Then, 3 mL of deionized water was added to the precipitate and dispersed by ultrasound. Meanwhile, 10 mg of RuCl3·xH2O and 2 mg of Na2WO4 were dispersed in a mixed solution of 2 mL deionized water and 1 mL methanol. The resulting solution was then injected into the ZIF-8 suspension and thoroughly mixed by stirring for 30 min. After that, the mixture was transferred into a 20 mL Teflon-lined autoclave and heated at 80 °C for 2 h. Finally, the resulting sample (denoted as RuW-2@ZIF-8) was then placed in a porcelain boat and calcined in a reducing gas, H2/Ar (95%). The temperature was raised to 300 °C at a heating rate of 10 °C/min and maintained for 2 h. After cooling to room temperature, the Ru1W0.14Zn1.47Ox nanocage was finally obtained. Soon afterwards, using 6 mg of W precursor while Ru was unchanged (denoted as RuW-6@ZIF-8) as the precatalyst, the Ru1W0.33Zn1.14Ox nanocage could be obtained. The catalyst RuZnxOy was obtained without using W precursor and Ru was kept unchanged. The atomic ratio between Ru, W, and Zn was obtained based on the XPS analysis.

2.3.3. Electrochemical Measurements

As is generally carried out for electrochemical measurements, 1 mg of the developed catalyst was added to 400 μL isopropanol, while 1 mg of carbon black was added to 390 μL of isopropanol and 10 μL of Nafion (5 wt%) solution. The mixture was continuously ultrasonicated for several hours until a homogeneous ink formed. Then, 15 μL of catalyst ink was evenly pipetted onto the glassy carbon working electrode. The catalyst loading on the working electrode was 0.19 mg/cm2.
Electrochemical tests were performed using a three-electrode system. The working electrode was a rotating disk electrode with a disk area of 0.196 cm2 and a rotation speed ranging from 400 to 2500 rpm. A saturated calomel electrode (SCE) and a graphite carbon electrode were used as the reference electrode and the counter electrode, respectively. Prior to testing, the 0.1 M KOH electrolyte was saturated with H2. The catalysts were first tested using cyclic voltammetry (CV) followed by linear sweep voltammetry (LSV) to ensure a quick attainment of steady state. The voltage range of CV was −0.05~0.2 V vs. RHE, with a scan rate of 10 mV s −1. For LSV, the scan rate was 1 mV s −1, and the rotation speed was 1600 rpm. The stability test of Ru1W0.14Zn1.47Ox was performed using a chronoamperometry test (CA) at a constant potential of 50 mV vs. RHE in H2-saturated 0.1 M KOH. Unless otherwise mentioned, the potentials in this work were converted to the reversible hydrogen electrode by the equation:
E R H E = E S C E + 0.059 × p H + 0.253
Due to the fact that Ru-based materials tend to absorb OH* in the hydrogen underpotential deposition (UPD) area. In this work, the Cu UPD method was employed to determine the electrochemical active surface areas (ECSAs) of Ru-based catalysts. The catalysts were subjected to multiple cycles of scanning, ranging from 0 to 0.7 V vs. RHE in order to obtain a stable CV curve as a background measurement in an Ar-saturated 0.1 M H₂SO₄ solution, with a scan rate of 10 mV s−1. When the deposition of Cu occurs at 0.3 V vs. RHE for a duration of 100 s, the stripping of Cu-UPD commences at potentials ranging from 0.3–0.7 V vs. RHE in an Ar-saturated 0.1M H2SO4 solution containing 2 mM CuSO4, with a scanning rate of 10 mV s−1. The value of ECSA (cm2 g−1) can be calculated using the equation:
E C S A ( c m m e t a l 2 g m e t a l ) = Q C u m m e t a l · 420   μ C   c m 2
where QCu is the measured integral charge, mmetal is the mass loading of metals on the electrode, 420 μC cm−2metal is the surface charge density which is assumed for a monolayer adsorption of Cu-UPD on metal.
After LSV test, electrochemical impedance spectroscopy (EIS) was performed over a frequency range from 200 kHz to 0.1 kHz with a voltage perturbation of 10 mV. The potential after iR correction is calculated using the equation:
E i R = E i R
where E and R correspond to the measured potential and the solution resistance, respectively. The kinetic current density (jk) can be obtained by the Koutecky–Levich equation:
1 j = 1 j k + 1 j d = 1 j k + 1 B c 0 w 1 / 2
where j, jd, B, c0, and ω represent the measured current, the diffusional current, the Levich constant, the solubility of H2 (7.33 × 10−4 mol L−1), and the rotating speed, respectively. The exchange current density (j0) could be obtained by the Butler–Volmer (B-V) equation:
j k = j 0 e α F R T η e 1 α F R T η
where α, η, R, T, and F represent the transfer coefficient, the overpotential, the molar gas constant (8.314 J mol−1 K−1), the operating temperature, and the Faraday constant, respectively. j0 could be also obtained from the approximate B-V equation:
j = j 0 η F R T

3. Results and Discussion

3.1. Catalyst Synthesis and Characterization

We fabricated the Ru1W0.14Zn1.47Ox nanocage through a two-step process involving hydrothermal and calcination treatments. ZIF-8 was first prepared as the template as detailed in the Methods section. The morphology of the pure ZIF-8 was imaged by TEM, showing a cubic structure with an average size of approximately 100 nm. Furthermore, the TEM images demonstrate that ZIF-8 was uniformly distributed on the copper grid without any agglomeration (Figure 1a). To prepare the Ru-W-Zn-O nanocomposites, RuCl3·xH2O and Na2WO4 were used as the Ru and W precursors, respectively. Different ratios of Ru/W were applied during the hydrothermal reaction to control the compositions. After introducing Ru or W ions, ZIF-8 underwent hydrolysis during the hydrothermal treatment, leading to the formation of the hollow nanocage structure, as shown in Figure 1b–d, with Ru and W confined in the substrate. Compared to Ru@ZIF-8 (Figure 1b), the relatively lighter contrast of the hollow structures of RuW@ZIF-8 (Figure 1c,d), indicating the introduction of W might accelerate the replacement of Zn and creating a favorable structure for further electrochemical investigation. To obtain the metallic Ru, which was crucial for electrocatalytic HOR, the precatalysts were subsequently annealed at 300 °C under Ar/H2 (5%) atmosphere and kept at this temperature for 2 h. The atomic ratios of the Ru-W-Zn-O catalysts after annealing were confirmed by XPS analysis. After annealing, ZIF-8 still shows the cubic structure (Figure 1e). In contrast, RuZnxOy, Ru1W0.14Zn1.47Ox, and Ru1W0.33Zn1.14Ox catalysts corresponding to Ru@ZIF-8, RuW-2@ZIF8, and RuW-6@ZIF-8 precatalysts, respectively, maintain the pristine hollow structure without any aggregations or collapses. Moreover, the high-temperature treatment also generated pores on the nanocages, which could facilitate H2 transfer and improve the electrocatalytic performance (Figure 1f,g). The SEM images of the annealed samples at different magnifications also demonstrated that the catalysts retained their cubic structure (Figure 2a–f).
We further conducted XRD characterization to investigate the phase structures of the catalysts. After hydrothermal and annealing treatments, ZIF-8 retained its pristine structure (Figure 3a). The XRD pattern of RuZnxOy reveals the presence of ZnO impurity and the hexagonal close-packed Ru (h-Ru) phase (JCPDS No. 06-0663). In contrast, after the induction of W, the peaks of ZnO impurities are absent, indicating that Ru-W-Zn-O composites tend to exhibit amorphous structures (Figure 3b). The Ru1W0.14Zn1.47Ox catalyst shows a broad peak at approximately 42°, which could be assigned to h-Ru. However, with increasing W content, the characteristic peaks of h-Ru disappear and obvious diffraction peaks of tungsten oxide are observed, which could be attributed to the highly dispersion of W within the nanocomposites.
The high-resolution TEM (HRTEM) images were employed to investigate the atomic lattice of the catalysts. Figure 4a,b show that both RuZnxOy and Ru1W0.14Zn1.47Ox nanocages retain a size of approximately 100 nm. These nanocages are composed of numerous nanoparticles and exhibit porous structures. In the HRTEM image of RuZnxOy (Figure 4c), the lattice distances of 2.12 Å and 2.65 Å are ascribed to h-Ru (0002) and ZnO (0002) facets, respectively. Ru nanoparticles, with a size ranging from 2 to 5 nm, are seen to be well distributed on the substrate. However, continuous lattice fringes of ZnO are absent from the Ru1W0.14Zn1.47Ox catalyst after the introduction of W (Figure 4d), and the size of the Ru nanoparticles decreased to approximately 1~2 nm, which is consistent with the XRD results. Additionally, EDS elemental mapping and line-scanning analysis (Figure 4e–g) demonstrate that Ru, W, and Zn are evenly distributed in the nanocages.
We further conducted XPS analysis to characterize the composition and oxidation state of the catalyst. The XPS survey spectrum of RuZnxOy indicates the presence of Ru, Zn, C, and O elements. Ru1W0.14Zn1.47Ox and Ru1W0.33Zn1.14Ox exhibit an additional characteristic peak of W compared to RuZnxOy, indicating successful W and Ru doping (Figure 5). The Ru 3p was selected to investigate the valence state of Ru since the binding energies of Ru 3d overlap with those C 1s [31]. The high-resolution XPS spectra of Ru 3p of RuZnxOy and Ru1W0.14Zn1.47Ox after annealing in H2/Ar atmosphere (Figure 6a) exhibit two peaks at 461.69 and 461.38 eV, corresponding to metallic Ru [32]. The binding energy of the Ru 3p in the Ru1W0.14Zn1.47Ox catalyst is positively shifted compared to RuZnxOy, indicating a modification of the electronic structure of Ru due to the introduction of W. The electron transfer between the active site of the Ru and other atoms can modulate the d-band structure of Ru and optimize the binding ability between H* and Ru, thus enhancing the HOR performance. The high-resolution W 4f spectrum of Ru1W0.14Zn1.47Ox (Figure 6b) shows peaks at 37.31 eV and 35.27 eV corresponding to W-O, and at 33.94 eV and 31.54 eV corresponding to W-C [32]. However, Ru1W0.33Zn1.14Ox only shows peaks at 37.63 and 35.4 eV corresponding to W-O. As shown in Figure 6c, the peaks at 1021.94 and 1022.15 eV corresponding to Zn 2p3/2, and those at 1044.86 and 1045.19 eV corresponding to Zn 2p1/2, indicating the presence of Zn2+. Figure 6d shows that the peaks located at 531.45 and 531.71 eV corresponding to C=O, and those at 530.02 and 530.40 eV corresponding to W/Zn-O, demonstrate the formation of oxidized W and Zn [33]. These results confirm that Ru remains the metallic state in the developed RuZnxOy, Ru1W0.14Zn1.47Ox, and Ru1W0.33Zn1.14Ox catalysts, and the slight shift in Ru 3p between RuZnxOy and W-doped composites is attributed to the charge transfer effect, which can optimize the electronic structure of the active sites.

3.2. Electrochemical Characterization

The electrochemical HOR performances of the catalysts were evaluated using a rotating disk electrode in a three-electrode system (Figure 7) in H2-saturated 0.1 M KOH.
The HOR polarization curves of the catalysts were obtained through LSV tests conducted at a rotating speed of 1600 rpm with a scan rate of 1 mV s−1. Figure 8a shows that the Ru1W0.14Zn1.47Ox catalyst exhibits a higher anode current density compared to RuZnxOy, commercial Ru/C, and Pt/C. For commercial Ru/C, the catalyst loading was adjusted to 0.5 mg cm−2 to prepare a well-dispersed catalyst membrane on the working electrode. The result suggests that the introduction of W is essential for enhancing HOR activity. However, the electrochemical performance decreases when there is an excessive amount of W, indicating that an excessive usage of W might cover the active sites of Ru and reduce the porosity of the composites, thus hindering H2 transfer and decreasing the electrochemical performance. The geometric current density (jg) of Ru1W0.14Zn1.47Ox increases rapidly at around 0 V vs. RHE compared to RuZnxOy, commercial Pt/C, and Ru/C, as confirmed by the micro-polarization region (−5 to 5 mV) analysis (Figure 8b). The exchange current density (j0) of Ru1W0.14Zn1.47Ox was calculated as 1.54 mA cm−2 in the micro-polarization region using the Butler–Volmer equation, which is 1.32 and 2.03 times higher than that of commercial Pt/C and Ru/C, respectively. The Cu-UPD analysis showed that the ECSA of Ru1W0.14Zn1.47Ox reached 116 m2 g−1 (Figure 8c). After normalization by ECSA, the specific activity (j0,s) of Ru1W0.14Zn1.47Ox was found to be 0.026 mA cmECSA−2, close to that of the commercial Ru/C catalyst [34]. This suggests that the enhanced HOR activity of Ru1W0.14Zn1.47Ox can be attributed to the favorable structure created, which has abundant exposed active sites. The polarization curves of Ru1W0.14Zn1.47Ox at various rotation speeds, ranging from 400 to 2500 rpm, showing that the anodic current densities consistently increase with the higher rotation speeds (Figure 8d). The Koutecký–Levich plot at an overpotential of 50 mV reveals a linear relationship between j−1 and ω−1/2.
The slope of Ru1W0.14Zn1.47Ox is 11.1 cm2 mA−1 rpm−1/2, which is close to the theoretical value of 14.8 cm2 mA−1 rpm−1/2, confirming that the current is mainly derived from the HOR process involving a 2e transfer (inset of Figure 8d). The Tafel plots (Figure 8e) further demonstrate that the introduction of W significantly promotes the HOR process in alkali. The Ru1W0.14Zn1.47Ox catalyst exhibits the highest j0 and jg at 50 mV vs. RHE compared to commercial Pt/C, Ru/C, and RuZnxOy, as shown in Figure 8f. Using the Koutecký–Levich equation, we obtained a geometric kinetic current density (jk) of 7.69 mA cm−2 for the Ru1W0.14Zn1.47Ox catalyst, which is 1.82 times higher than that of commercial Pt/C, and similar to that of Ru/C, despite the Ru/C loading being approximately 2.63 times higher than that of Ru1W0.14Zn1.47Ox. The mass content of Ru in Ru1W0.14Zn1.47Ox and RuZnOx nanocages has been determined to be approximately 4.5 wt% by ICP-MS. The mass activity (jk,m) was obtained by normalizing the kinetic current density with respect to the mass of noble metal. As shown in Figure 8g, the jk,m of Ru1W0.14Zn1.47Ox catalyst is 0.9 mA µg−1 at 50 mV vs. RHE, which is 1.22, 8.18, and 3.04 times higher than that of RuZnOx, commercial Pt/C, and Ru/C, respectively. Furthermore, the value of jk,m for the Ru1W0.14Zn1.47Ox catalyst indicates the enhanced catalytic activity compared to other reported HOR electrocatalysts in alkaline media (Figure 8h).
Besides HOR activity, operating stability is another important factor for application. Therefore, we conducted a stability test using chronoamperometry at a constant potential of 50 mV vs. RHE in H2-saturated 0.1 M KOH. Figure 8i shows that the HOR current density of the Ru1W0.14Zn1.47Ox catalyst remains stable without noticeable decay over 10,000 s. These results demonstrate the excellent HOR activity and stability of the Ru1W0.14Zn1.47Ox catalyst in alkali, suggesting its potential application in future AEMFCs.

4. Conclusions

In summary, we have synthesized a series of Ru-W-Zn-O nanocage composites via hydrothermal and annealing treatments using ZIF-8 as the template. The noble metal loading of the Ru1W0.14Zn1.47Ox catalyst is only 4.5 wt%, which has greatly reduced the cost of the anode material. Meanwhile, the optimized Ru1W0.14Zn1.47Ox nanocage catalyst, with a moderate amount of W, exhibits a high electrochemical surface area of 116 m2 g−1 and excellent HOR activity, with a mass activity of 0.9 mA µg−1 at an overpotential of 50 mV, surpassing that of commercial Pt/C and Ru/C. The high catalytic activity is primarily attributed to the incorporation of W, which modulates the d-band structure of Ru and optimizes the binding ability between active sites and intermediates. Moreover, the three-dimensional hollow porous structure promotes mass transfer and diffusion, accelerating H2 transfer and improving the electrocatalytic performance. Our work presents a high-performance Pt-free catalyst material for alkaline HOR and offers valuable insights for developing highly active electrocatalysts.

Author Contributions

Conceptualization, X.S.; methodology, X.S.; validation, X.S.; investigation, X.S., Z.C., H.L. and S.C.; resources, Y.-R.Z.; data curation, X.S.; writing—original draft preparation, X.S. and Z.C.; writing—review and editing, Y.-R.Z.; supervision, Y.-R.Z.; project administration. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Zhang, X.; Xiao, X.; Chen, J.; Liu, Y.; Pan, H.; Sun, W.; Gao, M. Toward the fast and durable alkaline hydrogen oxidation reaction on ruthenium. Energy Environ. Sci. 2022, 15, 4511–4526. [Google Scholar] [CrossRef]
  2. Xue, Y.; Shi, L.; Liu, X.; Fang, J.; Wang, X.; Setzler, B.P.; Zhu, W.; Yan, Y.; Zhuang, Z. A highly-active, stable and low-cost platinum-free anode catalyst based on RuNi for hydroxide exchange membrane fuel cells. Nat. Commun. 2020, 11, 5651. [Google Scholar] [CrossRef] [PubMed]
  3. Lu, Y.; Cai, Y.; Souamy, L.; Song, X.; Zhang, L.; Wang, J. Solid oxide fuel cell technology for sustainable development in China: An over-view. Int. J. Hydrog. Energy 2018, 43, 12870–12891. [Google Scholar] [CrossRef]
  4. Yang, Z.; Lai, W.; He, B.; Wang, J.; Yu, F.; Liu, Q.; Liu, M.; Zhang, S.; Ding, W.; Lin, Z.; et al. Tailoring Interfacial Chemistry of Defective Carbon-Supported Ru Catalyst Toward Efficient and CO-Tolerant Alkaline Hydrogen Oxidation Reaction. Adv. Energy Mater. 2023, 13, 2300881. [Google Scholar] [CrossRef]
  5. Su, L.; Fan, X.; Jin, Y.; Cong, H.; Luo, W. Hydroxyl-Binding Energy-Induced Kinetic Gap Narrowing between Acidic and Alkaline Hydrogen Oxidation Reaction on Intermetallic Ru3Sn7 Catalyst. Small 2023, 19, 2207603. [Google Scholar] [CrossRef] [PubMed]
  6. Su, L.; Jin, Y.; Gong, D.; Ge, X.; Zhang, W.; Fan, X.; Luo, W. The Role of Discrepant Reactive Intermediates on Ru-Ru2P Heterostructure for pH-Universal Hydrogen Oxidation Reaction. Angew. Chem. Int. Ed. 2022, 62, 2215585. [Google Scholar]
  7. Wang, P.; Wang, C.; Yang, Y.; Chen, S.; Cheng, Z.; Huang, M.; Tong, H.; Chen, Q. RuP Nanoparticles Supported on N, O Codoped Porous Hollow Carbon for Efficient Hydrogen Oxidation Reaction. Adv. Mater. Interfaces 2022, 9, 2102193. [Google Scholar] [CrossRef]
  8. Zhou, Y.; Xie, Z.; Jiang, J.; Wang, J.; Song, X.; He, Q.; Ding, W.; Wei, Z. Lattice-confined Ru clusters with high CO tolerance and activity for the hydrogen oxidation reaction. Nat. Catal. 2020, 3, 454–462. [Google Scholar] [CrossRef]
  9. Mao, J.; He, C.-T.; Pei, J.; Liu, Y.; Li, J.; Chen, W.; He, D.; Wang, D.; Li, Y. Isolated Ni Atoms Dispersed on Ru Nanosheets: High-Performance Electrocatalysts toward Hydrogen Oxidation Reaction. Nano Lett. 2020, 20, 3442–3448. [Google Scholar] [CrossRef]
  10. Cui, Y.; Xu, Z.; Chen, D.; Li, T.; Yang, H.; Mu, X.; Gu, X.; Zhou, H.; Liu, S.; Mu, S. Trace oxophilic metal induced surface reconstruction at buried RuRh cluster interfaces possesses extremely fast hydrogen redox kinetics. Nano Energy 2021, 90, 106579. [Google Scholar] [CrossRef]
  11. Ohyama, J.; Kumada, D.; Satsuma, A. Improved hydrogen oxidation reaction under alkaline conditions by ruthenium–iridium alloyed nanoparticles. J. Mater. Chem. A 2016, 4, 15980–15985. [Google Scholar] [CrossRef]
  12. Scofield, M.E.; Zhou, Y.; Yue, S.; Wang, L.; Su, D.; Tong, X.; Vukmirovic, M.B.; Adzic, R.R.; Wong, S.S. Role of Chemical Composition in the Enhanced Catalytic Activity of Pt-Based Alloyed Ultrathin Nanowires for the Hydrogen Oxidation Reaction under Alkaline Conditions. ACS Catal. 2016, 6, 3895–3908. [Google Scholar] [CrossRef]
  13. Qin, B.; Yu, H.; Gao, X.; Yao, D.; Sun, X.; Song, W.; Yi, B.; Shao, Z. Ultrathin IrRu nanowire networks with high performance and durability for the hydrogen oxidation reaction in alkaline anion exchange membrane fuel cells. J. Mater. Chem. A 2018, 6, 20374–20382. [Google Scholar] [CrossRef]
  14. Yang, X.; Ouyang, B.; Shen, P.; Sun, Y.; Yang, Y.; Gao, Y.; Kan, E.; Li, C.; Xu, K.; Xie, Y. Ru Colloidosome Catalysts for the Hydrogen Oxidation Reaction in Alkaline Media. J. Am. Chem. Soc. 2022, 144, 11138–11147. [Google Scholar] [CrossRef] [PubMed]
  15. Ming, M.; Zhang, Y.; He, C.; Zhao, L.; Niu, S.; Fan, G.; Hu, J.S. Room-Temperature Sustainable Synthesis of Selected Platinum Group Metal (PGM = Ir, Rh, and Ru) Nanocatalysts Well-Dispersed on Porous Carbon for Efficient Hydrogen Evolution and Oxidation. Small 2019, 15, 1903057. [Google Scholar] [CrossRef] [PubMed]
  16. Wang, L.; Xu, Z.; Kuo, C.H.; Peng, J.; Hu, F.; Li, L.; Chen, H.Y.; Wang, J.; Peng, S. Stabilizing Low-Valence Single Atoms by Constructing Metalloid Tungsten Carbide Supports for Efficient Hydrogen Oxidation and Evolution. Angew. Chem. Int. Ed. 2023, 62, 2311937. [Google Scholar]
  17. Jiang, J.; Tao, S.; He, Q.; Wang, J.; Zhou, Y.; Xie, Z.; Ding, W.; Wei, Z. Interphase-oxidized ruthenium metal with half-filled d-orbitals for hydrogen oxidation in an alkaline solution. J. Mater. Chem. A 2020, 8, 10168–10174. [Google Scholar] [CrossRef]
  18. Yang, F.; Wang, Y.; Cui, Y.; Yang, X.; Zhu, Y.; Weiss, C.M.; Li, M.; Chen, G.; Yan, Y.; Gu, M.D.; et al. Sub-3 nm Pt@Ru toward Outstanding Hydrogen Oxidation Reaction Performance in Alkaline Media. J. Am. Chem. Soc. 2023, 145, 27500–27511. [Google Scholar] [CrossRef] [PubMed]
  19. Ohyama, J.; Sato, T.; Yamamoto, Y.; Arai, S.; Satsuma, A. Size Specifically High Activity of Ru Nanoparticles for Hydrogen Oxidation Reaction in Alkaline Electrolyte. J. Am. Chem. Soc. 2013, 135, 8016–8021. [Google Scholar] [CrossRef]
  20. He, S.; Tu, Y.; Zhang, J.; Zhang, L.; Ke, J.; Wang, L.; Du, L.; Cui, Z.; Song, H. Ammonia-Induced FCC Ru Nanocrystals for Efficient Alkaline Hydrogen Electrocatalysis. Small 2023, 20, 08053. [Google Scholar] [CrossRef]
  21. Li, Y.; Yang, C.; Yue, J.; Cong, H.; Luo, W. Polymorphism-Interface-Induced Work Function Regulating on Ru Nanocatalyst for Enhanced Alkaline Hydrogen Oxidation Reaction. Adv. Funct. Mater. 2023, 33, 11586. [Google Scholar] [CrossRef]
  22. Mahmood, J.; Li, F.; Jung, S.-M.; Okyay, M.S.; Ahmad, I.; Kim, S.-J.; Park, N.; Jeong, H.Y.; Baek, J.-B. An efficient and pH-universal ruthenium-based catalyst for the hydrogen evolution reaction. Nat. Nanotechnol. 2017, 12, 441–446. [Google Scholar] [CrossRef] [PubMed]
  23. Zhuang, Z.W.; Wang, Y.; Xu, C.Q.; Liu, S.J.; Chen, C.; Peng, Q.; Zhuang, Z.B.; Xiao, H.; Pan, Y.; Lu, S.Q.; et al. Three-dimensional open nano-netcage electrocatalysts for efficient pH-universal overall water splitting. Nat. Commun. 2019, 10, 4875. [Google Scholar] [CrossRef] [PubMed]
  24. Li, Y.Z.; Abbott, J.; Sun, Y.C.; Sun, J.M.; Du, Y.C.; Han, X.J.; Wu, G.; Xu, P. Ru nanoassembly catalysts for hydrogen evolution and oxidation reactions in electrolytes at various pH values. Appl. Catal. B-Environ. Energy 2019, 258, 117952. [Google Scholar] [CrossRef]
  25. Wu, Y.H.; Li, Y.W.; Gao, J.K.; Zhang, Q.C. Recent advances in vacancy engineering of metal-organic frameworks and their derivatives for electrocatalysis. Sus. Mat. 2021, 1, 66–87. [Google Scholar] [CrossRef]
  26. Li, G.D.; Zhao, S.L.; Zhang, Y.; Tang, Z.Y. Metal-Organic Frameworks Encapsulating Active Nanoparticles as Emerging Composites for Catalysis: Recent Progress and Perspectives. Adv. Mater. 2018, 30, 1800702. [Google Scholar] [CrossRef] [PubMed]
  27. Duan, J.J.; Chen, S.; Li, Y.B.; Zhao, C. Closely Arranged 3D-0D Graphene-Nickel Sulfide Superstructures for Bifunctional Hydrogen Electrocatalysis. ACS Appl. Energy Mater. 2018, 1, 6368–6373. [Google Scholar] [CrossRef]
  28. Zhu, J.; Chen, F.; Zhang, Z.; Li, M.; Yang, Q.; Yang, Y.; Bao, Z.; Ren, Q. M-Gallate (M = Ni, Co) Metal–Organic Framework-Derived Ni/C and Bimetallic Ni–Co/C Catalysts for Lignin Conversion into Monophenols. ACS Sustain. Chem. Eng. 2019, 7, 12955–12963. [Google Scholar] [CrossRef]
  29. Ma, M.; Chen, C.; Zhang, X.; Zhao, H.; Wang, Q.; Du, G.; Xie, Z.; Kuang, Q. Mo-modified electronic effect on sub-2 nm Ru catalyst for enhancing hydrogen oxidation catalysis. J. Mater. Chem. A 2023, 11, 10807–10812. [Google Scholar] [CrossRef]
  30. Qiu, Z.; Li, Y.; Gao, Y.; Meng, Z.; Sun, Y.; Bai, Y.; Suen, N.T.; Chen, H.C.; Pi, Y.; Pang, H. 2D MOF-assisted Pyrolysis-displacement-alloying Synthesis of High-entropy Alloy Nanoparticles Library for Efficient Electrocatalytic Hydrogen Oxidation. Angew. Chem. Int. Ed. 2023, 62, 2306881. [Google Scholar] [CrossRef]
  31. Lao, M.M.; Zhao, G.Q.; Li, P.; Ma, T.Y.; Jiang, Y.Z.; Pan, H.G.; Dou, S.X.; Sun, W.P. Manipulating the Coordination Chemistry of Ru-N(O)-C Moieties for Fast Alkaline Hydrogen Evolution Kinetics. Adv. Funct. Mater. 2021, 31, 2100698. [Google Scholar] [CrossRef]
  32. Han, P.; Yang, X.; Wu, L.; Jia, H.; Chen, J.; Shi, W.; Cheng, G.; Luo, W. A Highly-Efficient Boron Interstitially Inserted Ru Anode Catalyst for Anion Exchange Membrane Fuel Cells. Adv. Mater. 2023, 36, 2304496. [Google Scholar] [CrossRef]
  33. Yang, Y.; Shao, X.; Zhou, S.; Yan, P.; Isimjan, T.T.; Yang, X. Interfacial Electronic Coupling of NC@WO3-W2C Decorated Ru Clusters as a Reversible Catalyst toward Electrocatalytic Hydrogen Oxidation and Evolution Reactions. ChemSusChem 2021, 14, 2992–3000. [Google Scholar] [CrossRef] [PubMed]
  34. Feng, Y.; Lu, S.; Fu, L.; Yang, F.; Feng, L. Alleviating the competitive adsorption of hydrogen and hydroxyl intermediates on Ru by d–p orbital hybridization for hydrogen electrooxidation. Chem. Sci. 2024, 15, 2123–2132. [Google Scholar] [CrossRef] [PubMed]
  35. Zhang, B.X.; Zhang, B.H.; Zhao, G.Q.; Wang, J.M.; Liu, D.Q.; Chen, Y.P.; Xia, L.X.; Gao, M.X.; Liu, Y.F.; Sun, W.P.; et al. Atomically dispersed chromium coordinated with hydroxyl clusters enabling efficient hydrogen oxidation on ruthenium. Nat. Commun. 2022, 13, 5894. [Google Scholar] [CrossRef] [PubMed]
  36. Zhan, C.; Xu, Y.; Bu, L.; Zhu, H.; Feng, Y.; Yang, T.; Zhang, Y.; Yang, Z.; Huang, B.; Shao, Q.; et al. Sub nanometer high-entropy alloy nanowires enable remarkable hydrogen oxidation catalysis. Nat. Commun. 2021, 12, 6261. [Google Scholar] [CrossRef] [PubMed]
  37. Zhang, X.B.; Xia, L.X.; Zhao, G.Q.; Zhang, B.X.; Chen, Y.P.; Chen, J.; Gao, M.X.; Jiang, Y.Z.; Liu, Y.F.; Pan, H.G.; et al. Fast and Durable Alkaline Hydrogen Oxidation Reaction at the Electron-Deficient Ruthenium-Ruthenium Oxide Interface. Adv. Mater. 2023, 35, 2208821. [Google Scholar] [CrossRef] [PubMed]
  38. Hamo, E.R.; Singh, R.K.; Douglin, J.C.; Chen, S.; Hassine, M.B.; Carbo-Argibay, E.; Lu, S.; Wang, H.; Ferreira, P.J.; Rosen, B.A.; et al. Carbide-Supported PtRu Catalysts for Hydrogen Oxidation Reaction in Alkaline Electrolyte. ACS Catal. 2021, 11, 932–947. [Google Scholar] [CrossRef]
  39. Li, Y.B.; Yang, C.Y.; Ge, C.X.; Yao, N.; Yin, J.L.; Jiang, W.Y.; Cong, H.J.; Cheng, G.Z.; Luo, W.; Zhuang, L. Electronic Modulation of Ru Nanosheet by d-d Orbital Coupling for Enhanced Hydrogen Oxidation Reaction in Alkaline Electrolytes. Small 2022, 18, 2202404. [Google Scholar] [CrossRef]
Figure 1. Morphology characterization. (ad) TEM images of as-prepared (a) ZIF-8, precatalysts of (b) Ru@ZIF-8, (c) RuW-2@ZIF-8, and (d) RuW-6@ZIF-8, respectively. (e,f) TEM images of (e) ZIF-8 after annealing, (f) RuZnxOy, (g) Ru1W0.14Zn1.47Ox, and (h) Ru1W0.33Zn1.14Ox, respectively.
Figure 1. Morphology characterization. (ad) TEM images of as-prepared (a) ZIF-8, precatalysts of (b) Ru@ZIF-8, (c) RuW-2@ZIF-8, and (d) RuW-6@ZIF-8, respectively. (e,f) TEM images of (e) ZIF-8 after annealing, (f) RuZnxOy, (g) Ru1W0.14Zn1.47Ox, and (h) Ru1W0.33Zn1.14Ox, respectively.
Nanomaterials 14 00808 g001
Figure 2. SEM images of (a,d) ZIF-8 after annealing, (b,e) Ru1W0.14Zn1.47Ox, and (c,f) Ru1W0.33Zn1.14Ox.
Figure 2. SEM images of (a,d) ZIF-8 after annealing, (b,e) Ru1W0.14Zn1.47Ox, and (c,f) Ru1W0.33Zn1.14Ox.
Nanomaterials 14 00808 g002
Figure 3. XRD patterns of (a) ZIF-8 with different treatment ways, (b) RuZnxOy, Ru1W0.33Zn1.14Ox and Ru1W0.14Zn1.47Ox.
Figure 3. XRD patterns of (a) ZIF-8 with different treatment ways, (b) RuZnxOy, Ru1W0.33Zn1.14Ox and Ru1W0.14Zn1.47Ox.
Nanomaterials 14 00808 g003
Figure 4. TEM images of (a) RuZnxOy and (b) Ru1W0.14Zn1.47Ox, respectively. HRTEM images of (c) RuZnxOy and (d) Ru1W0.14Zn1.47Ox, respectively. (e) The EDS elemental mapping analysis of Ru1W0.14Zn1.47Ox. Scale bar: 50 nm. (f,g) EDX line-scanning profile of Ru1W0.14Zn1.47Ox catalyst. Scale bar: 50 nm.
Figure 4. TEM images of (a) RuZnxOy and (b) Ru1W0.14Zn1.47Ox, respectively. HRTEM images of (c) RuZnxOy and (d) Ru1W0.14Zn1.47Ox, respectively. (e) The EDS elemental mapping analysis of Ru1W0.14Zn1.47Ox. Scale bar: 50 nm. (f,g) EDX line-scanning profile of Ru1W0.14Zn1.47Ox catalyst. Scale bar: 50 nm.
Nanomaterials 14 00808 g004
Figure 5. XPS survey spectra of RuZnxOy, Ru1W0.14Zn1.47Ox, and Ru1W0.33Zn1.14Ox.
Figure 5. XPS survey spectra of RuZnxOy, Ru1W0.14Zn1.47Ox, and Ru1W0.33Zn1.14Ox.
Nanomaterials 14 00808 g005
Figure 6. High-resolution XPS spectra of (a) Ru 3p of RuZnxOy, Ru1W0.14Zn1.47Ox and Ru1W0.33Zn1.14Ox. High-resolution XPS spectra of (b) W 4f, (c) Zn 2p, and (d) O 1s of Ru1W0.14Zn1.47Ox. and Ru1W0.33Zn1.14Ox, respectively.
Figure 6. High-resolution XPS spectra of (a) Ru 3p of RuZnxOy, Ru1W0.14Zn1.47Ox and Ru1W0.33Zn1.14Ox. High-resolution XPS spectra of (b) W 4f, (c) Zn 2p, and (d) O 1s of Ru1W0.14Zn1.47Ox. and Ru1W0.33Zn1.14Ox, respectively.
Nanomaterials 14 00808 g006
Figure 7. Schematic illustration of the three-electrode system.
Figure 7. Schematic illustration of the three-electrode system.
Nanomaterials 14 00808 g007
Figure 8. Electrocatalytic HOR performances. (a) Polarization curves of Ru1W0.14Zn1.47Ox, Pt/C, Ru/C, RuZnxOy, and Ru1W0.33Zn1.14Ox, respectively, in H2-saturated 0.1 M KOH solution. (c) The Cuupd stripping voltammogram of Ru1W0.14Zn1.47Ox. The loading of Ru is about 10 µg. (b) Micropolarization region between −5 to 5 mV vs. RHE of the corresponding catalysts. (d) Polarization curves of Ru1W0.14Zn1.47Ox at different rotation rates. The Koutecky–Levich plot of Ru1W0.14Zn1.47Ox at an overpotential of 50 mV vs. RHE inset of (d). (e) Tafel plots present the kinetic current densities of Ru1W0.14Zn1.47Ox, Pt/C, Ru/C, RuZnxOy, and Ru1W0.33Zn1.14Ox, respectively, derived from the Butler–Volmer equation fitting. (f) Comparison of the jg, jk, and j0 of Ru1W0.14Zn1.47Ox, RuZnxOy, Pt/C, and Ru/C at 50 mV vs. RHE. (g) Comparison of mass normalized current density for Ru1W0.14Zn1.47Ox, RuZnOx, commercial Pt/C, and Ru/C. (h) Comparison of the jk,m at an overpotential of 50 mV vs. RHE for Ru1W0.14Zn1.47Ox in this work with other alkaline HOR catalysts. The column diagrams are duplicated from the literature: Ru-Cr1(OH)x-2.2 ([35]), PtRu/C ([36]), Ru–RuO2 ([37]), Ru colloidosome ([14]), Pt7Ru3 NWs ([12]), PtRu/Mo2C-TaC ([38]), Ru3Sn7/C ([5]), RuFe nanosheet ([39]), Ru–Ir/C-20 ([11]). (i) Chronoamperometry (j-t) response of Ru1W0.14Zn1.47Ox in H2-saturated 0.1 M KOH solution at 50 mV vs. RHE. The loading of Ru is 8.55 µg cm−2 in all electrocatalytic tests, except Cu-UPD testing.
Figure 8. Electrocatalytic HOR performances. (a) Polarization curves of Ru1W0.14Zn1.47Ox, Pt/C, Ru/C, RuZnxOy, and Ru1W0.33Zn1.14Ox, respectively, in H2-saturated 0.1 M KOH solution. (c) The Cuupd stripping voltammogram of Ru1W0.14Zn1.47Ox. The loading of Ru is about 10 µg. (b) Micropolarization region between −5 to 5 mV vs. RHE of the corresponding catalysts. (d) Polarization curves of Ru1W0.14Zn1.47Ox at different rotation rates. The Koutecky–Levich plot of Ru1W0.14Zn1.47Ox at an overpotential of 50 mV vs. RHE inset of (d). (e) Tafel plots present the kinetic current densities of Ru1W0.14Zn1.47Ox, Pt/C, Ru/C, RuZnxOy, and Ru1W0.33Zn1.14Ox, respectively, derived from the Butler–Volmer equation fitting. (f) Comparison of the jg, jk, and j0 of Ru1W0.14Zn1.47Ox, RuZnxOy, Pt/C, and Ru/C at 50 mV vs. RHE. (g) Comparison of mass normalized current density for Ru1W0.14Zn1.47Ox, RuZnOx, commercial Pt/C, and Ru/C. (h) Comparison of the jk,m at an overpotential of 50 mV vs. RHE for Ru1W0.14Zn1.47Ox in this work with other alkaline HOR catalysts. The column diagrams are duplicated from the literature: Ru-Cr1(OH)x-2.2 ([35]), PtRu/C ([36]), Ru–RuO2 ([37]), Ru colloidosome ([14]), Pt7Ru3 NWs ([12]), PtRu/Mo2C-TaC ([38]), Ru3Sn7/C ([5]), RuFe nanosheet ([39]), Ru–Ir/C-20 ([11]). (i) Chronoamperometry (j-t) response of Ru1W0.14Zn1.47Ox in H2-saturated 0.1 M KOH solution at 50 mV vs. RHE. The loading of Ru is 8.55 µg cm−2 in all electrocatalytic tests, except Cu-UPD testing.
Nanomaterials 14 00808 g008
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

Sun, X.; Cheng, Z.; Liu, H.; Chen, S.; Zheng, Y.-R. Porous Ruthenium–Tungsten–Zinc Nanocages for Efficient Electrocatalytic Hydrogen Oxidation Reaction in Alkali. Nanomaterials 2024, 14, 808. https://doi.org/10.3390/nano14090808

AMA Style

Sun X, Cheng Z, Liu H, Chen S, Zheng Y-R. Porous Ruthenium–Tungsten–Zinc Nanocages for Efficient Electrocatalytic Hydrogen Oxidation Reaction in Alkali. Nanomaterials. 2024; 14(9):808. https://doi.org/10.3390/nano14090808

Chicago/Turabian Style

Sun, Xiandi, Zhiyuan Cheng, Hang Liu, Siyu Chen, and Ya-Rong Zheng. 2024. "Porous Ruthenium–Tungsten–Zinc Nanocages for Efficient Electrocatalytic Hydrogen Oxidation Reaction in Alkali" Nanomaterials 14, no. 9: 808. https://doi.org/10.3390/nano14090808

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