Next Article in Journal
New Insights into the Surfactant-Assisted Liquid-Phase Exfoliation of Bi2S3 for Electrocatalytic Applications
Next Article in Special Issue
Petal-like g-C3N4 Enhances the Photocatalyst Removal of Hexavalent Chromium
Previous Article in Journal
The Effect of Cu(II) Nanoparticle Decoration on the Electron Relaxations and Gaseous Photocatalytic Oxidations of Nanocrystalline TiO2
Previous Article in Special Issue
Unraveling the Mechanism for H2O2 Photogeneration on Polymeric Carbon Nitride with Alkali Metal Modification
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Promotional Effect of Na on Ru for pH-Universal Hydrogen Evolution Reactions

1
College of Materials and Environmental Engineering, Institute of Advanced Magnetic Materials, Hangzhou Dianzi University, Hangzhou 310018, China
2
Key Laboratory of Interface Science and Engineering in Advanced Materials, Ministry of Education, Taiyuan University of Technology, Taiyuan 030024, China
3
Institute of Industrial Catalysis, College of Chemical Engineering, Zhejiang University of Technology, Hangzhou 310032, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(3), 552; https://doi.org/10.3390/catal13030552
Submission received: 15 February 2023 / Revised: 4 March 2023 / Accepted: 7 March 2023 / Published: 9 March 2023

Abstract

:
Alkali metals, as ideal electron donors, can effectively regulate the valence state distribution of the host metals. Nevertheless, no studies have reported the application of alkali metal promoters in the hydrogen evolution reaction (HER). Here, we designed an efficient and wide pH-universal hydrogen evolution catalyst that utilizes alkali metal to control the valence, size, and dispersion of Ru NPs. The experimental results reveal that the alkali metal additives contribute to the dispersion and stabilization of metallic Ru. More importantly, the interaction between Na and Ru regulates the distribution of Ru valence states and helps to form more active components of Ru0. Additionally, NaCl functioned as an in situ template to assist the construction of a porous carbon skeleton promotes mass transfer and exposes more active sites, further promoting the synergistic effect of Ru and Na. As a result, the optimal Ru0.3/C−800 delivers high efficiency for HER with an overpotential as low as 29 mV in 1.0 M KOH and 83 mV in 0.5 M H2SO4 under 10 mA cm−2. Particularly, the catalytic performance of Ru0.3/C−800 even outbalanced that of commercial Pt/C in an alkaline medium. This rational construction strategy opens up new avenues for obtaining superior pH-universal electrocatalysts.

1. Introduction

As a sustainable secondary energy, hydrogen (H2) energy has the advantages of wide raw material sources and zero carbon emission, and it has demonstrated an excellent application foreground in industries such as automotive, chemical, and aviation [1,2]. Compared with traditional steam reforming technology, producing hydrogen by water electrolysis fundamentally eliminates the dependence on fossil fuels and achieves zero carbon emissions [3,4]. Nowadays, the electrolysis of water has become the most competitive hydrogen production technology [5,6]. Alkaline electrolyzer, proton exchange membrane electrolyzer, and other types of electrolyzers are available on the market. The various operating environments put forward higher requirements for the design of catalysts. It is highly desirable to create economical and pH-universal catalysts. Moreover, the practical deployment of water electrolysis technology is heavily restricted by the sluggish kinetics of the cathode catalysts, which results in a high overpotential to drive H2 production [7]. Currently, Pt-based electrocatalysts, as the benchmark for HER, exhibit excellent catalytic performance [8,9,10,11]. Unfortunately, the scarce reserves and exorbitant price largely impede their commercialization. Developing cost-effective electrocatalysts to substitute Pt/C is a central issue in the water industry [12]. As a Pt-group metal, Ru possesses a moderate metal–hydrogen bond strength similar to Pt–H, but the cost is only 1/30 that of Pt [13,14,15,16]. In general, the HER efficiency is determined by the strength of the metal–hydrogen bond [17]. Ru serving as a catalyst’s active center instead of expensive Pt metal is expected to keep the catalytic activity unchanged while reducing the catalyst cost [18,19]. However, the current Ru-based catalysts suffer from complex preparation methods, the high toxicity of reagents, and high metal loading [20]. Therefore, it is urgent and challenging to develop facile, economical, and green synthetic strategies for fabricating efficient Ru-based catalysts.
Alkali metal, serving as a structural or electronic regulator, plays vital role in thermocatalytic reactions [21]. Since the first report on the promotional effect of alkali metal in 1845, the function of alkali metal has been systematically studied in hydrogenation reactions, such as ammonia production [22], Fischer–Tropsch process [23], aromatics [24], hydrogenation of unsaturated hydrocarbons [25,26,27,28], and CO2 [29]. In the meantime, it is generally believed that alkali metals act as electron donors to modulate the electronic structures of the active metals, thereby changing the adsorption of reacting intermediates [30,31,32]. For example, Zhang introduced Na dopant and oxygen vacancy, which made the d-band center of RuO2 far away from the Fermi level, resulting in the weakening of the chemical bond between the oxygen intermediate and the surface of RuO2, thus reducing the activation barrier of the oxygen evolution reaction (OER) to optimize the performance [33]. By regulating the interplay of Na and Co, Wu’s group obtained stable Na-Co2C active sites, formed Na-CO bonds to strengthen the interaction between Na and Co2C, dispersed the Co2C, and reduced the particle size [34]. Despite alkali metal promoters being widely used in thermocatalysis, OER, etc., the application of alkali metal promoters in HER has not been reported, let alone the influencing mechanism of alkali metal on the HER performance.
Herein, “starch strips” derived from potatoes, commercially used in the preparation of puffed food were chosen as the carbon source. The “starch strips” were economical, widely sourced, and rich in NaCl. Utilizing abundant NaCl in “starch strips” as the pore-forming agent, the porous carbon-supported ultra-dispersed Ru-based composite (Ru/C) was synthesized by a simple impregnation–pyrolysis strategy; meanwhile, alkali metal Na promoters were introduced in situ. It was found that the introduction of Na greatly modulated the valence state distribution of Ru species, inducing the formation of more Ru0 states which, in turn, boosted the electrocatalytic performance. Benefitting from the interplay between Na and Ru, as well as the developed texture structure, Ru0.3/C−800 functioned as a super and long-lasting HER catalyst over a broad pH range. In 1.0 M KOH and 0.5 M H2SO4, the overpotentials of Ru0.3/C−800 were only 29 mV and 83 mV to achieve the current density of 10 mA cm−2. Notably, the mass loading of Ru in Ru0.3/C−800 was as low as 0.81 wt%, greatly reducing the synthetic cost compared with that of commercial Pt/C (20 wt%). Furthermore, the HER performance for Ru0.3/C−800 outbalanced Pt/C in alkaline solution. This work provides a new method for alkali metal promoters to tune the electrocatalytic performance, which could lead to the design of novel catalysts for efficient hydrogen evolution.

2. Results and Discussion

2.1. Catalyst Synthesis and Characterization

The Ru/C was synthesized using the method of impregnation adsorption and one-step pyrolysis starting from RuCl3 and “starch strips” derived from potatoes as the feedstock, which was ecofriendly, commercially available, and inexpensive (Figure 1). Benefitting from the confinement effect of the carbon skeleton, Ru species were converted into monodispersed nanoparticles (NPs) [35]. In this study, the as-prepared catalysts were named Rux/C−T, where x refers to the additional amounts of Ru (0, 0.2, 0.3, and 0.4 mmol), and T refers to pyrolysis temperature (700, 800, 900, and 1000 ℃). Transmission electron microscopy (TEM) was performed to observe their morphologies and microstructures. For Ru0.3/C−800, a distinct pore structure was exhibited on the carbon substrate (Figure S1), which was attributed to the pore-forming effect of NaCl. It was not difficult to find that, the Ru NPs of Ru0.3/C−800 were uniformly distributed on the carbon frame with a mean particle size of ~1.90 nm (Figure 2a). When the dose of Ru3+ was lower than 0.4 mmol, the Ru particles of Ru0.2/C−800 had a uniform size. However, upon increasing the addition of Ru3+ to 0.4 mmol, the Ru particles of Ru0.4/C−800 were significantly agglomerated (Figure S2). In addition, the pyrolysis temperature also played a vital part in the distribution of Ru particles. As displayed in Figure S3a, Ru0.3/C−700 displayed the homogeneous distribution of Ru NPs due to the fact of insufficient nucleation under low pyrolysis temperature. However, with the increase in the pyrolysis temperature, the agglomeration of the Ru NPs occurred obviously (Figure S3b,c). Furthermore, to investigate the role of NaCl on the microstructure and metal phase of the catalysts, the “starch strips” were pretreated to wash off NaCl as much as possible, and the obtained catalyst was denoted as Ru0.3/C−800−WF. Compared with Ru0.3/C−800, the pore structure of Ru0.3/C−800−WF was not obvious at the same scale, indicating the pore-forming effect of NaCl. In addition, Ru0.3/C−800−WF only contained a small number of Ru particles, which demonstrates that NaCl also contributed to dispersing and anchoring the metallic Ru (Figure S4). The mass loading of Ru and Na was, furthermore, measured by an inductively coupled plasma-optical emission spectrometer (ICP-OES). As elevating the pyrolysis temperature from 800 ℃ to 1000 ℃, the Na content gradually reduced from 0.36 wt% to 0.004 wt% (Table S1). Accordingly, the content of Ru showed the same variation trend as that of Na content with the decreased content from 0.81 wt% to 0.11 wt%. Similarly, the Ru content of Ru0.3/C−800−WF was reduced as the Na content decreased compared with Ru0.3/C−800. Such results revealed that the stabilizing effect of Na on Ru was due to the interaction between them. Detected by high-resolution transmission electron microscopy (HRTEM), the interplanar spacings of Ru0.3/C−800 were measured to be 0.205 nm and 0.231 nm (Figure 2b), coincident with the (101) and (100) planes of Ru, respectively. In addition, the uniform distribution of metallic Ru in the carbon matrix was further identified under a high-angle annular dark-field scanning transmission electron microscope (HADDF-STEM) (Figure 2c) [36]. The elemental mapping of Ru0.3/C−800 (Figure 2d – i) demonstrated the homogeneous dispersion of C, O, N, Ru, Na, and Cl. Notably, it could be found that the element Ru was distributed around Na, as shown in the Figure S5, indicating the interaction between them, to a certain extent.
More structural details of the catalysts were then confirmed by X-ray diffractometer (XRD). From Figure 3a, a broad peak belonging to amorphous carbon was observed in all samples. Apart from that, the diffraction peaks coincided well with the (100), (002), and (101) planes of Ru (PDF#06-0663) located at ≈38.4°, 42.2°, and 44.0°, which appeared in Ru0.3/C−800 and Ru0.3/C−900. This result was associated with the good crystalline character of Ru at 800 °C and 900 °C (Figure S6). In addition, the typical diffraction peaks of NaCl (PDF#05-0628) at ≈31.7° and 45.5° were presented as well. It was worth noting that the intensity of the NaCl peaks weakened obviously along with the elevated pyrolysis temperature, which was due to the volatilization of NaCl as the temperature increased [37]. As for Ru0.3/C−800−WF, there were only amorphous carbon peaks existing from the XRD pattern, revealing that NaCl was completely removed (Figure S7b). Moreover, as exhibited in Figure S7a, the NaCl peak intensity gradually weakened as the content of in Ru content increased.
X-ray photoelectron spectroscopy (XPS) was further employed to reveal the chemical valence and elemental compositions of the samples [17]. Here, the C 1s peak of 284.5 eV served as the standard for all of the peaks’ calibration [38]. Figure 3b visually showed the Na 1s peaks of Ru0.3/C−800 and C−800. In contrast to C−800, the Na 1s peak of Ru0.3/C−800 was shifted towards higher binding energy owing to the electron transfer from Na to Ru. From the Ru 3p spectra in Figure 3c, the peaks at 462.0 eV, 484.2 eV, 464.6 eV, and 486.5 eV were corresponding to Ru0 3p3/2, Ru0 3p1/2, Run+ 3p3/2, and Run+ 3p1/2, respectively [17,39]. The binding energy of Ru 3p of Ru0.3/C−800 was negatively biased compared with Ru0.3/C−800−WF. Moreover, the peak positions of Na 1s and Ru 3p varied with the pyrolysis temperature. As shown in Figure S8a, with the increase in the temperature, Na 1s shifted to the low binding energy. In Figure S8b, when the content of Na is high, the shift in the Ru 3p peak position was more obvious, because more Na transferred electrons to Ru. It was calculated that the ratio of Ru0/Run+ for Ru0.3/C−800 was 1.68, which was much higher than that of other control groups (Table S2). According to reports, Ru0 was the active center of catalytic hydrogen production, and the high content of Ru0 was helpful for improving catalytic hydrogen production [40]. In this system, Na, as an electronic structure regulator, could effectively adjust the valence distribution of Ru, promoting the formation of more Ru0, and hopefully improved catalytic activity.
Na not only served as an assistant to regulate the valence distribution of Ru and increase the content of Ru0 but also acted as a pore-forming agent that could affect the pore structure of the catalysts [41]. The pore structure was studied using the Brunauer–Emmett–Teller (BET) method, as displayed in Figure 3d. The N2 adsorption/desorption plots exhibited that the specific surface area of Ru0.3/C−800 was 592.0 m2 g−1, significantly larger than other catalysts (Table S3). The decreased specific surface areas of Ru0.3/C−900 and Ru0.3/C−1000 were possibly caused by the collapse of the mesoporous structure when the material calcined at high temperatures (especially over 800 °C). Furthermore, the pore size distribution of Ru0.3/C−800 (inset in Figure 3d) manifested a large number of mesopores. The existence of a large number of mesopores was beneficial for the solution to enter during the reaction, which increased the contact area, exposed more active sites, and accelerated mass transfer [42]. Additionally, the defective nature of the catalysts was investigated by Raman spectroscopy [43]. A typical D band and G band of carbon was manifested in Ru0.3/C−800, and the intensity ratio ID/IG was 0.856. Moreover, the value of ID/IG increased with the increase in the pyrolysis temperatures (Figure S9), possibly because too high pyrolysis temperature would lead to the collapse of the pore structure. It displayed that only an appropriate pyrolysis temperature can optimize the conductivity, as well as the catalytic abilities of the composites.

2.2. Electrochemical Characterization

The Ru powder, commercial Pt/C (20 wt%), Pt electrode, and Rux/C−T were tested for HER in 1.0 M KOH, 0.5 M H2SO4, and 1.0 M PBS, respectively. As is well known, the HER activity was evaluated by the overpotential (η10) at the current density of 10 mA cm−2 [43,44,45,46]. From Figure 4a, the blank carbon (C−800) displayed almost no activity for hydrogen generation within the applied potential window. However, upon introducing trace Ru in the system, the catalytic activity was significantly improved. With the increase in the Ru concentration, the η10 distinctly declined from 73 mV (Ru0.2/C−800) to 29 mV (Ru0.3/C−800). Whereas, the η10 further increased to 35 mV (Ru0.4/C−800) when the Ru dosage was increased to 0.4 mmol. These results indicate that the appropriate Ru content was able to enhance the catalytic activity. It is noteworthy that the HER of Ru0.3/C−800 exhibited an even better performance than Pt/C (20 wt%), Ru powder, and Pt electrode in 1.0 M KOH, as shown in Figure S10 and Figure 4a. Additionally, an effective catalyst for HER was able to initiate proton reduction with minimum overpotential and fast kinetics [46]. Normally, the Tafel slope is a vital indicator to investigate the reaction kinetics for HER [47]. According to the polarization curves of the samples, the corresponding Tafel slopes were obtained (Figure 4b). It was not hard to find that the order of Tafel slopes was Ru0.3/C−800 (34.43 mV dec−1) < Ru0.4/C−800 (36.89 mV dec−1) < Pt/C (59.44 mV dec−1) < Ru0.2/C−800 (78.74 mV dec−1), which was in line with the trend of η10. The smaller Tafel slope value of Ru0.3/C−800 suggested a higher reaction rate, and the HER catalyzed by Ru0.3/C−800 followed a Volmer–Heyrovsky mechanism [12]. Then, the charge transfer kinetics involved in HER was detected through EIS. The subsequent Nyquist diagram (Figure 4c) demonstrated that the curve of the sample was virtually semicircular, with the diameter of the curve standing for the associated charge transfer resistance (Rct). In addition, the lower Rct indicated a rapid electrode reaction rate. The Ru0.3/C−800 had the smallest Rct, indicating that the reaction rate was the fastest, which was consistent with the results of the voltammetry measurements. Furthermore, the electrochemical double-layer capacitance (Cdl), which was measured in the non-Faraday region of the CV curves, was used to evaluate the ECSA. The Cdl value was produced by linearly graphing the capacitance current density response vs. the scanning rate [17]. Figure S11 depicts the ECSA curves with various Ru contents. The Cdl of Ru0.3/C−800 was 135.66 mF cm−2 (Figure 4d), which was significantly higher than the Cdl of Ru0.4/C−800 (113.54 mF cm−2), Ru0.2/C−800 (65.31 mF cm−2), and C−800 (4.60 mF cm−2). The increase in the Cdl value means that Ru0.3/C-800 possesses a large active surface area, which contributes to the exposure of active sites and the charge transfer [48].
On this basis, the role of the pyrolysis temperature on the catalytic activity was subsequently explored [18]. As shown in Figure 5a, Ru0.3/C−700 presented a high overpotential of 214 mV in alkaline solution to achieve 10 mA cm−2. Upon increasing the pyrolysis temperature, the overpotential dropped sharply, and the overpotential of Ru0.3/C−800 only required 29 mV to achieve the same current density. However, further elevating the pyrolysis temperature, the overpotential increased to 52 mV and 90 mV for Ru0.3/C−900 and Ru0.3/C−1000, respectively. Additionally, as exhibited in Figure S12, the Cdl value rose first and then declined with the increase of the pyrolysis temperature. This result was in line with the activity trend. It was further confirmed that the high Cdl was the reason for the improved activity. However, the variation in the Cdl mainly came from the specific surface area, which was ultimately ascribed to the NaCl template for the forming of pores. Furthermore, the catalytic activity was closely correlated with the electrochemical reaction rate and the capability of charge transfer, which was verified by the Tafel (Figure 5b) and EIS (Figure 5c) results. In addition, Figure S13 illustrates the EIS fitting circuit for Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 1.0 M KOH. It can be seen from Table S4 that the smaller Rct value of Ru0.3/C−800 corresponded to the higher reaction rates. More importantly, the overpotential of Ru0.3/C-800-WF increased to 72 mV when the amount of NaCl in the raw material was washed off (Figure 5d), and Figure S14 shows that the corresponding Tafel, Cdl, and EIS results displayed a worsening trend. This was due to the pore-forming effect of NaCl, which regulated the specific surface area of the catalysts and, thus, affected their activity. Meanwhile, it can be seen from Figure 5e that the decrease in the NaCl content in the raw materials was accompanied by the decrease in the Ru0 component and the deterioration of activity. Moreover, the same results can be obtained from the catalysts calcined under varied temperatures. When the pyrolysis temperature rose, the overpotential for Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 at 10 mA cm−2 was inversely proportional to the Ru0 content. To put it simply, the higher the Ru0 content, the smaller the overpotential of the catalyst, and the better the electrochemical performance. These findings again demonstrate the interaction between Na and Ru; that is, Na not only adjusted the valence state of the Ru phase but also stabilized Ru. Simultaneously, an ideal catalyst not only has excellent catalytic activity but also high catalytic stability. Therefore, stability was another vital indicator to assess the performance of the catalyst [49,50,51]. On the one hand, the durability of Ru0.3/C−800 was measured by continuous CV sweeps. Figure 5f reveals that the overpotential only fell by 2 mV after 1000 cycles, suggesting the excellent stability of Ru0.3/C−800. On the other hand, under the corresponding potential, a further measurement was performed for 10 h by chronoamperometry. No significant reduction in the current density could be observed (inset in Figure 5f).
Furthermore, we also studied the electrocatalytic activity of the catalyst in acidic and neutral media, verifying the universality of the designed catalyst in a wide pH [52]. In acidic medium, a current density of 10 mA cm−2 could be reached by the low overpotential of 83 mV for Ru0.3/C−800, which was significantly lower than that of Ru0.3/C−700 (basically inactivated), Ru0.3/C−900 (96 mV), and Ru0.3/C−1000 (131 mV). Meanwhile, the corresponding Tafel slope and EIS plots (Figure 5h,5i) showed the same trend as the activity curves. Additionally, the activity of Ru0.3/C−800 was better than Ru powder in acidic medium (Figure S15a), and Ru0.3/C−800 exhibited good stability (Figure S15b). Moreover, we conducted activity tests in a. neutral environment (Figure S16a). The performance was not ideal, but Ru0.3/C−800 still had advantages compared with the Ru powder. Surprisingly, the performance of Ru0.3/C−800 was excellent in terms of stability (Figure S16b). In summary, the above results strongly support the hypothesis that Ru0.3/C−800 has good activity and stability for HER in a wide pH range.

3. Experimental

3.1. Chemicals

Potassium hydroxides (KOH), ruthenium (III) chloride (RuCl3, Ru content: 45 wt%~55 wt%), sulfuric acid (H2SO4), Ru powder, and ethanol were bought from Aladdin Industrial Inc. The commercial Pt/C with 20 wt% Pt loading was purchased from Johnson Matthey Company (Sigma, Shanghai, China). The main ingredient in the “starch strips” was starch, which was bought from supermarkets and is usually used in the production of puffed foods. Deionized water (DIW) was used to prepare all solutions. The DIW used in the experiment was obtained using a water purifier Hitech-DW, Wenzhou, Zhejiang Province, China).

3.2. Synthesis of Electrocatalysts

3.2.1. Synthesis of Ru0.3/C−800

First, 400 mg of RuCl3 was weighed and dissolved in 20 mL of DIW. Another clean glass beaker was used and 17 mL of DIW was mixed with 3 mL of the prepared RuCl3 solution. After, ~4 g of the “starch strips” were soaked in the above RuCl3 solution at room temperature for 1 h. Furthermore, the treated “starch strips” were transferred to a watchglass and dried at 60 °C to obtain the precursor. Afterward, the precursor was calcined at 800 °C for 2 h in a tube furnace under an Ar environment. Finally, the product was ultrasonically washed several times with DIW. The final product, Ru0.3/C−800, was obtained after drying.

3.2.2. Synthesis of C−800, Ru0.2/C−800, and Ru0.4/C−800

For the comparison of the samples of C−800, Ru0.2/C−800, and Ru0.4/C−800, the total volume of the reaction mixture remained at 20 mL. However, the addition of the RuCl3 solution was changed to 0/2/4 mL, respectively. The other conditions remained unchanged.

3.2.3. Synthesis of Ru0.3/C−700, Ru0.3/C−900, and Ru0.3/C−1000

For the comparison of the samples of Ru0.3/C−700, Ru0.3/C−900, and Ru0.3/C−1000, the preparation procedure followed that of Ru0.3/C−800, except that the calcination temperature varied to 700 °C, 900 °C, and 1000 °C.

3.2.4. Synthesis of Ru0.3/C−800−WF (Wash First)

In addition, to demonstrate the influence of Na on the valence state of Ru, a control sample was set. In the early stage, the NaCl in the “starch strips” were washed as much as possible, and then the washed “starch strips” were used as raw materials for the synthesis of Ru0.3/C−800−WF. The preparation process followed that of Ru0.3/C−800.

3.3. Material Characterization

The HRTEM was carried out on a Talos F200S S-Twin (Thermo Fisher, Czech Republic), and the accelerating voltage was 200 kV. A D/tex Ultima TV wide-angle X-ray diffractometer (NIC, Japan) was employed to measure the metal phase of the samples. The diffractometer was equipped with Cu Kα radiation (1.54 Å), and the scanning rate was 5 °/min. The Raman spectra were collected on a Raman spectrometer (Labram HR800-LS55, HORIBA, France). The specific surface area and PSD curves were calculated by traditional BET and BJH methods, respectively. The X-ray photoelectron spectra (XPS) for the bonding states were obtained using a Thermo ESCALAB 250xi spectrometer (Thermo Fisher, Czech Republic). In addition, the samples were first dissolved with aqua regia. Inductively coupled plasma-optical emission spectrometry (ICP-OES) (PerkinElmer, Optima 5300 DV) was used to measure the contents of Ru and Na. The Ru content of Ru0.3/C-800 was as low as 0.81 wt% as measured by ICP-OES.

3.4. Electrochemical Characterization

To assess the HER performance of the catalysts, all experiments were performed on a conventional three-electrode system with CHI 760E as the electrochemical workstation, and the temperature during the test was kept at room temperature. A saturated calomel electrode (SCE) and a carbon rod was employed as the reference and the counter electrode, respectively. A glassy carbon electrode 5 mm in diameter acted as the working electrode. A 3 mg catalyst was dispersed in the mixture of Nafion (20 μL) and ethanol (300 μL). Whereafter, the prepared ink was ultrasonic treated for 30 min to form a uniform suspension. Then, 30 μL ink was coated on the surface of the glassy carbon electrode and dried at room temperature. The electrochemical tests were operated in 0.5 M H2SO4, 1.0 M KOH, and 1.0 M PBS solution. Linear sweep voltammogram (LSV) curves were acquired at a scan rate of 5 mV s−1 over a potential window of −0.4 to 0.1 V (relative to a reversible hydrogen electrode (RHE)). All polarization curves were IR-corrected. The Tafel equation (η = a + b log(j)) was employed to fit the slope of Tafel. In order to obtain the electrochemical active surface area (ECSA) of the prepared catalysts, we measured a series of cyclic voltammetry (CV) plots at different scanning rates from 10 to 50 mV s−1 in the potential window of 0.1–0.2 V relative to RHE. Additionally, electrochemical impedance spectroscopy (EIS) was collected from the frequency range of 0.01 Hz to 100 kHz, and the test voltage was −0.023 V (vs. RHE).

4. Conclusions

In this work, pH-universal Ru/C electrocatalysts for HER were constructed using a two-step adsorption–pyrolysis strategy. In 1.0 M KOH and 0.5 M H2SO4, the overpotentials of Ru0.3/C−800 were as low as 29 mV and 83 mV to obtain a current density of 10 mA cm−2. This work has several advantages over previously reported results: (1) The “starch strips” containing natural alkali metal additives (NaCl) were used as the raw material, which were green, ecofriendly, and economical. (2) The alkali metal auxiliaries were able to stabilize Ru and regulate the distribution of Ru valence, which contributed to the formation of more active components Ru0. (3) NaCl as a salt template could induce a porous structure, promoting mass transfer and exposing more active sites. The synthetic strategy was not only general and scalable but also emphasized the critical role of alkali metal in promoting HER activity, which provides a new idea for developing efficient non-Pt HER catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13030552/s1, Table S1: Elemental content of Ru and Na measured by ICP-OES, Table S2: Ru0/Run+ ratio of Ru0.3/C−800, Ru0.3/C−900, Ru0.3/C−1000, and Ru0.3/C−800−WF; Table S3: BET surface area (m²/g) of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, Ru0.3/C−1000, and Ru0.3/C−800−WF; Table S4: Rs and Rct of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 1.0 M KOH; Table S5: Summary of the recently reported HER catalysts; Figure S1: TEM image of Ru0.3/C−800, where the yellow circle is a pore structure, mainly mesopores; Figure S2: (a) TEM image of Ru0.2/C−800, (b) TEM image of Ru0.4/C−800; Figure S3: (a) TEM image of Ru0.3/C−700, (b) TEM image of Ru0.3/C−900, (c) TEM image of Ru0.3/C−1000; Figure S4: TEM image of (a) Ru0.3/C−800 and (b) Ru0.3/C−800−WF; Figure S5: HRTEM of Ru0.3/C−800; Figure S6: HRTEM image of (a) Ru0.3/C−800 and (b) Ru0.3/C−900; Figure S7: (a) XRD patterns of C−800, Ru0.2/C−800, Ru0.3/C−800, and Ru0.4/C−800 (inset: XRD enlargement of Ru0.3/C−800), (b) XRD patterns of Ru0.3/C−800 and Ru0.3/C−800−WF; Figure S8: (a) High-resolution XPS spectra of Na 1s of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000, (b) high-resolution XPS spectra of Ru 3p of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; Figure S9: Raman spectra of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; Figure S10: LSV curves of Ru powder, Pt electrode, and Ru0.3/C−800 in 1.0 M KOH; Figure S11: CV curves of (a) C−800, (b) Ru0.2/C−800, (c) Ru0.3/C−800, and (d) Ru0.4/C−800 at different scan rates: 10, 20, 30, 40, and 50 mV s-1 in 1.0 M KOH solution; Figure S12: CV curves of (a) Ru0.3/C−700, (b) Ru0.3/C−900, and (c) Ru0.3/C−1000 at different scan rates: 10, 20, 30, 40, and 50 mV s-1 in 1.0 M KOH solution, (d) linear plot of the capacitive current versus the scan rate for determination of the Cdl; Figure S13: EIS spectra of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 (inset: equivalent circuit models used for fitting the EIS response of HER); Figure S14: (a) Tafel plots of Ru0.3/C−800 and Ru0.3/C−800−WF, (b) CV curves of Ru0.3/C−800−WF at different scan rates: 10, 20, 30, 40, and 50 mV s-1 in 1.0 M KOH solution, (c) linear plot of the capacitive current versus the scan rate for determination of the Cdl, (d) Nyquist plots of Ru0.3/C−800 and Ru0.3/C−800−WF; Figure S15: (a) LSV curves of Ru0.3/C−800, Ru powder, Pt electrode, and 20 wt% Pt/C in 0.5 M H2SO4, (b) The i–t curve of Ru0.3/C−800 in 0.5 M H2SO4; Figure S16: (a) LSV curves of Ru0.3/C−800, Ru powder, Pt electrode, and 20 wt% Pt/C in 1.0 M PBS, (b) The i–t curve of Ru0.3/C−800 in 1.0 M PBS.

Author Contributions

B.G., Writing—original draft, Investigation, and Visualization; C.Z., Investigation; Y.Z.: Investigation; J.G., Project administration; Z.W., Conceptualization, Writing—review, and Funding acquisition; J.W., Writing—review and editing, Investigation, Methodology, and Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (52002102 and 22008213), the Zhejiang Provincial Natural Science Foundation (LQ21B060005 and LQ19B030009), and the Fundamental Research Funds for the Provincial Universities of Zhejiang (RF-A2022011).

Data Availability Statement

Data will be made available upon request.

Conflicts of Interest

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.

References

  1. Jiang, B.; Wang, Z.; Cheng, L.; Chen, Z.; Dong, Y.; Wang, X.; Wang, T.; Mao, X.; Gao, Y.; Xu, Z.; et al. Regulating charge distribution of Ru atoms in ruthenium phosphide/carbon nitride/carbon for promoting hydrogen evolution reaction. J. Alloy. Compd. 2023, 939, 168717. [Google Scholar] [CrossRef]
  2. Abdollahi, A.; Ghaffarinejad, A.; Arabi, M. Electrodeposition of Ni-Fe on graphite rod as an efficient and binder-free electrocatalyst for oxygen and hydrogen evolution reactions. J. Alloy. Compd. 2023, 937, 168400. [Google Scholar] [CrossRef]
  3. Wu, X.; Wang, Z.; Zhang, D.; Qin, Y.; Wang, M.; Han, Y.; Zhan, T.; Yang, B.; Li, S.; Lai, J.; et al. Solvent-free microwave synthesis of ultra-small Ru-Mo2C@CNT with strong metal-support interaction for industrial hydrogen evolution. Nat. Commun. 2021, 12, 4018. [Google Scholar] [CrossRef] [PubMed]
  4. Jiang, K.; Luo, M.; Liu, Z.; Peng, M.; Chen, D.; Lu, Y.-R.; Chan, T.-S.; de Groot, F.M.F.; Tan, Y. Rational strain engineering of single-atom ruthenium on nanoporous MoS2 for highly efficient hydrogen evolution. Nat. Commun. 2021, 12, 1687. [Google Scholar] [CrossRef] [PubMed]
  5. Liu, Y.; Li, W.; Wu, H.; Lu, S. Carbon Dots Enhance Ruthenium Nanoparticles for Efficient Hydrogen Production in Alkaline. Acta Phys. Chim. Sin. 2021, 37, 2009082. [Google Scholar] [CrossRef]
  6. Zhou, S.; Jang, H.; Qin, Q.; Li, Z.; Kim, M.G.; Li, C.; Liu, X.; Cho, J. Three-dimensional hierarchical Co(OH)F nanosheet arrays decorated by single-atom Ru for boosting oxygen evolution reaction. Sci. China Mater. 2021, 64, 1408–1417. [Google Scholar] [CrossRef]
  7. Zhang, J.; Mao, X.; Wang, S.; Liang, L.; Cao, M.; Wang, L.; Li, G.; Xu, Y.; Huang, X. Superlattice in a Ru Superstructure for Enhancing Hydrogen Evolution. Angew. Chem. Int. Ed. 2022, 61, e202116867. [Google Scholar] [CrossRef]
  8. Feng, T.; Yu, G.; Tao, S.; Zhu, S.; Ku, R.; Zhang, R.; Zeng, Q.; Yang, M.; Chen, Y.; Chen, W.; et al. A highly efficient overall water splitting ruthenium-cobalt alloy electrocatalyst across a wide pH range via electronic coupling with carbon dots. J. Mater. Chem. A 2020, 8, 9638–9645. [Google Scholar] [CrossRef]
  9. Anantharaj, S.; Noda, S. Appropriate Use of Electrochemical Impedance Spectroscopy in Water Splitting Electrocatalysis. ChemElectroChem 2020, 7, 2297–2308. [Google Scholar] [CrossRef]
  10. 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]
  11. Tiwari, J.N.; Dang, N.K.; Sultan, S.; Thangavel, P.; Jeong, H.Y.; Kim, K.S. Multi-heteroatom-doped carbon from waste-yeast biomass for sustained water splitting. Nat. Sustain. 2020, 3, 556–563. [Google Scholar] [CrossRef]
  12. Cai, C.; Liu, K.; Zhu, Y.; Li, P.; Wang, Q.; Liu, B.; Chen, S.; Li, H.; Zhu, L.; Li, H.; et al. Optimizing Hydrogen Binding on Ru Sites with RuCo Alloy Nanosheets for Efficient Alkaline Hydrogen Evolution. Angew. Chem. Int. Ed. 2022, 61, e202113664. [Google Scholar] [CrossRef]
  13. Ren, R.; Wang, X.; Chen, H.; Miller, H.A.; Salam, I.; Varcoe, J.R.; Wu, L.; Chen, Y.; Liao, H.; Liu, E.; et al. Reshaping the Cathodic Catalyst Layer for Anion Exchange Membrane Fuel Cells: From Heterogeneous Catalysis to Homogeneous Catalysis. Angew. Chem. Int. Ed. 2021, 60, 4049–4054. [Google Scholar] [CrossRef]
  14. Ha, Y.; Fei, B.; Yan, X.; Xu, H.; Chen, Z.; Shi, L.; Fu, M.; Xu, W.; Wu, R. Atomically Dispersed Co-Pyridinic N-C for Superior Oxygen Reduction Reaction. Adv. Energy Mater. 2020, 10, 2002592. [Google Scholar] [CrossRef]
  15. Rapson, T.D.; Ju, H.; Marshall, P.; Devilla, R.; Jackson, C.J.; Giddey, S.; Sutherland, T.D. Engineering a solid-state metalloprotein hydrogen evolution catalyst. Sci. Rep. 2020, 10, 3774. [Google Scholar] [CrossRef] [Green Version]
  16. Tong, Y.; Sun, Q.; Chen, P.; Chen, L.; Fei, Z.; Dyson, P.J. Nitrogen-Incorporated Cobalt Sulfide/Graphene Hybrid Catalysts for Overall Water Splitting. ChemSusChem 2020, 13, 5112–5118. [Google Scholar] [CrossRef]
  17. Ding, R.; Lin, L.; Pei, C.; Yu, X.; Sun, Q.; Park, H.S. Hierarchical Architectures Based on Ru Nanoparticles/Oxygen-Rich-Carbon Nanotubes for Efficient Hydrogen Evolution. Chem.—A Eur. J. 2021, 27, 11150–11157. [Google Scholar] [CrossRef]
  18. Jiao, J.; Zhang, N.; Zhang, C.; Sun, N.; Pan, Y.; Chen, C.; Li, J.; Tan, M.; Cui, R.; Shi, Z.; et al. Doping Ruthenium into Metal Matrix for Promoted pH-Universal Hydrogen Evolution. Adv. Sci. 2022, 9, 2200010. [Google Scholar] [CrossRef]
  19. Su, P.; Pei, W.; Wang, X.; Ma, Y.; Jiang, Q.; Liang, J.; Zhou, S.; Zhao, J.; Liu, J.; Lu, G.Q. Exceptional Electrochemical HER Performance with Enhanced Electron Transfer between Ru Nanoparticles and Single Atoms Dispersed on a Carbon Substrate. Angew. Chem. Int. Ed. 2021, 60, 16044–16050. [Google Scholar] [CrossRef]
  20. Wang, C.; Qi, L. Heterostructured Inter-Doped Ruthenium–Cobalt Oxide Hollow Nanosheet Arrays for Highly Efficient Overall Water Splitting. Angew. Chem. Int. Ed. 2020, 59, 17219–17224. [Google Scholar] [CrossRef]
  21. Rao, Y.; Xu, L.; Zhou, M.; Yin, B.; Osuka, A.; Song, J. Expanded Azaporphyrins Consisting of Multiple BODIPY Units: Global Aromaticity and High Affinities Towards Alkali Metal Ions. Angew. Chem. Int. Ed. 2022, 61, e202206899. [Google Scholar] [CrossRef] [PubMed]
  22. Nielsen, A. Review of ammonia catalysis. Catal. Rev. 1971, 4, 1–26. [Google Scholar] [CrossRef]
  23. Williams, F.J.; Palermo, A.; Tracey, S.; Tikhov, M.S.; Lambert, R.M. Electrochemical Promotion by Potassium of the Selective Hydrogenation of Acetylene on Platinum: Reaction Studies and XP Spectroscopy. J. Phys. Chem. B 2002, 106, 5668–5672. [Google Scholar] [CrossRef]
  24. García-Bordejé, E.; Dongil, A.B.; Conesa, J.M.; Guerrero-Ruiz, A.; Rodríguez-Ramos, I. Promotion of Ru or Ni on Alumina Catalysts with a Basic Metal for CO2 Hydrogenation: Effect of the Type of Metal (Na, K, Ba). Nanomaterials 2022, 12, 1052. [Google Scholar] [CrossRef] [PubMed]
  25. Su, L.; Liu, L.; Zhuang, J.; Wang, H.; Li, Y.; Shen, W.; Xu, Y.; Bao, X. Creating Mesopores in ZSM-5 Zeolite by Alkali Treatment: A New Way to Enhance the Catalytic Performance of Methane Dehydroaromatization on Mo/HZSM-5 Catalysts. Catal. Lett. 2003, 91, 155–167. [Google Scholar] [CrossRef]
  26. Zhao, B.; Zhai, P.; Wang, P.; Li, J.; Li, T.; Peng, M.; Zhao, M.; Hu, G.; Yang, Y.; Li, Y.-W.; et al. Direct Transformation of Syngas to Aromatics over Na-Zn-Fe5C2 and Hierarchical HZSM-5 Tandem Catalysts. Chem 2017, 3, 323–333. [Google Scholar] [CrossRef]
  27. Chen, I.; Chen, F.L. Effect of alkali and alkaline-earth metals on the resistivity to coke formation and sintering of nickel-alumina catalysts. Ind. Eng. Chem. Res. 1990, 29, 534–539. [Google Scholar] [CrossRef]
  28. Gentner, T.X.; Mulvey, R.E. Alkali-Metal Mediation: Diversity of Applications in Main-Group Organometallic Chemistry. Angew. Chem. Int. Ed. 2021, 60, 9247–9262. [Google Scholar] [CrossRef]
  29. Qin, R.; Zhou, L.; Liu, P.; Gong, Y.; Liu, K.; Xu, C.; Zhao, Y.; Gu, L.; Fu, G.; Zheng, N. Alkali ions secure hydrides for catalytic hydrogenation. Nat. Catal. 2020, 3, 703–709. [Google Scholar] [CrossRef]
  30. Zhang, W.; Sun, M.; Yin, J.; Abou-Hamad, E.; Schwingenschlögl, U.; Costa, P.M.F.J.; Alshareef, H.N. A Cyclized Polyacrylonitrile Anode for Alkali Metal Ion Batteries. Angew. Chem. Int. Ed. 2021, 60, 1355–1363. [Google Scholar] [CrossRef]
  31. Zimmermann, P.; Ar, D.; Rößler, M.; Holze, P.; Cula, B.; Herwig, C.; Limberg, C. Selective Transformation of Nickel-Bound Formate to CO or C−C Coupling Products Triggered by Deprotonation and Steered by Alkali-Metal Ions. Angew. Chem. Int. Ed. 2021, 60, 2312–2321. [Google Scholar] [CrossRef]
  32. Ding, Y.; Guo, X.; Qian, Y.; Gao, H.; Weber, D.H.; Zhang, L.; Goodenough, J.B.; Yu, G. In Situ Formation of Liquid Metals via Galvanic Replacement Reaction to Build Dendrite-Free Alkali-Metal-Ion Batteries. Angew. Chem. Int. Ed. 2020, 59, 12170–12177. [Google Scholar] [CrossRef]
  33. Zhang, L.; Jang, H.; Liu, H.; Kim, M.G.; Yang, D.; Liu, S.; Liu, X.; Cho, J. Sodium-Decorated Amorphous/Crystalline RuO2 with Rich Oxygen Vacancies: A Robust pH-Universal Oxygen Evolution Electrocatalyst. Angew. Chem. Int. Ed. 2021, 60, 18821–18829. [Google Scholar] [CrossRef]
  34. Zhang, S.; Wu, Z.; Liu, X.; Shao, Z.; Xia, L.; Zhong, L.; Wang, H.; Sun, Y. Tuning the interaction between Na and Co2C to promote selective CO2 hydrogenation to ethanol. Appl. Catal. B Environ. 2021, 293, 120207. [Google Scholar] [CrossRef]
  35. Chen, J.; Wang, H.; Gong, Y.; Wang, Y. Directly immobilizing a Ru–tannic acid linkage coordination complex on carbon cloth: An efficient and ultrastable catalyst for the hydrogen evolution reaction. J. Mater. Chem. A 2019, 7, 11038–11043. [Google Scholar] [CrossRef]
  36. Sun, S.-W.; Wang, G.-F.; Zhou, Y.; Wang, F.-B.; Xia, X.-H. High-Performance Ru@C4N Electrocatalyst for Hydrogen Evolution Reaction in Both Acidic and Alkaline Solutions. ACS Appl. Mater. Interfaces 2019, 11, 19176–19182. [Google Scholar] [CrossRef]
  37. Xu, C.; Ming, M.; Wang, Q.; Yang, C.; Fan, G.; Wang, Y.; Gao, D.; Bi, J.; Zhang, Y. Facile synthesis of effective Ru nanoparticles on carbon by adsorption-low temperature pyrolysis strategy for hydrogen evolution. J. Mater. Chem. A 2018, 6, 14380–14386. [Google Scholar] [CrossRef]
  38. Kweon, D.H.; Okyay, M.S.; Kim, S.-J.; Jeon, J.-P.; Noh, H.-J.; Park, N.; Mahmood, J.; Baek, J.-B. Ruthenium anchored on carbon nanotube electrocatalyst for hydrogen production with enhanced Faradaic efficiency. Nat. Commun. 2020, 11, 1278. [Google Scholar] [CrossRef] [Green Version]
  39. Li, J.-S.; Huang, M.-J.; Kong, L.-X.; Chen, X.-N.; Zhou, Y.-W.; Li, J.-L.; Wang, M.-Y. Ruthenium Nanoparticles Anchored on Graphene Hollow Nanospheres Superior to Platinum for the Hydrogen Evolution Reaction in Alkaline Media. Inorg. Chem. 2020, 59, 930–936. [Google Scholar] [CrossRef]
  40. Chithaiah, P.; Binwal, D.C.; Raos, C.N.R. Simple Synthesis of 2D Molybdenum Carbide Nanosheets and Their Application in the Hydrogen Evolution Reaction. Eur. J. Inorg. Chem. 2022, 2022, e202101086. [Google Scholar] [CrossRef]
  41. Song, H.; Wu, M.; Tang, Z.; Tse, J.S.; Yang, B.; Lu, S. Single Atom Ruthenium-Doped CoP/CDs Nanosheets via Splicing of Carbon-Dots for Robust Hydrogen Production. Angew. Chem. Int. Ed. 2021, 60, 7234–7244. [Google Scholar] [CrossRef] [PubMed]
  42. Kong, X.; Xu, K.; Zhang, C.; Dai, J.; Oliaee, S.N.; Li, L.; Zeng, X.; Wu, C.; Peng, Z. Free-Standing Two-Dimensional Ru Nanosheets with High Activity toward Water Splitting. ACS Catal. 2016, 6, 1487–1492. [Google Scholar] [CrossRef]
  43. Zheng, Y.; Jiao, Y.; Zhu, Y.; Li, L.H.; Han, Y.; Chen, Y.; Jaroniec, M.; Qiao, S.-Z. High Electrocatalytic Hydrogen Evolution Activity of an Anomalous Ruthenium Catalyst. J. Am. Chem. Soc. 2016, 138, 16174–16181. [Google Scholar] [CrossRef] [PubMed]
  44. Dong, S.; Li, Y.; Zhao, Z.; Li, R.; He, J.; Yin, J.; Yan, B.; Zhang, X. A Review of the Application of Heterostructure Catalysts in Hydrogen Evolution Reaction. ChemistrySelect 2022, 7, e202104041. [Google Scholar] [CrossRef]
  45. Lu, E.; Zhang, Z.; Tao, J.; Hou, Y.; Zhang, J.; Yu, Z. Enhanced Metal-Semiconductor Interaction for Photocatalytic Hydrogen Evolution Reaction. Chem.—A Eur. J. 2022, 28, e202201590. [Google Scholar] [CrossRef]
  46. Chen, Z.; Liu, X.; Shen, T.; Wu, C.; Zu, L.; Zhang, L. Porous NiFe alloys synthesized via freeze casting as bifunctional electrocatalysts for oxygen and hydrogen evolution reaction. Int. J. Hydrogen Energy 2021, 46, 37736–37745. [Google Scholar] [CrossRef]
  47. Wang, T.; Tao, L.; Zhu, X.; Chen, C.; Chen, W.; Du, S.; Zhou, Y.; Zhou, B.; Wang, D.; Xie, C.; et al. Combined anodic and cathodic hydrogen production from aldehyde oxidation and hydrogen evolution reaction. Nat. Catal. 2022, 5, 66–73. [Google Scholar] [CrossRef]
  48. Liu, L.; Qin, C.; Yu, M.; Wang, Q.; Wang, J.; Hou, B.; Jia, L.; Li, D. Morphology Evolution of Hcp Cobalt Nanoparticles Induced by Ru Promoter. ChemCatChem 2020, 12, 2083–2090. [Google Scholar] [CrossRef]
  49. Niyitanga, T.; Jeong, H.K. Modification of thermally reduced graphite oxide and molybdenum disulfide by solution plasma for hydrogen evolution reaction. Mater. Chem. Phys. 2021, 263, 124345. [Google Scholar] [CrossRef]
  50. Ortiz-Restrepo, J.E.; Loaiza, O.A.; Urresta, J.D.; Velasquez, J.D.; Pastor, E.; Chaur, M.N.; Lizcano-Valbuena, W.H. A comparative study of different carbon materials as metal-free catalysts for oxygen reduction and hydrogen evolution reactions in alkaline media. Diam. Relat. Mater. 2021, 117, 108464. [Google Scholar] [CrossRef]
  51. Pham, N.N.; Kang, S.G.; Kim, H.-J.; Pak, C.; Han, B.; Lee, S.G. Catalytic activity of Ni3Mo surfaces for hydrogen evolution reaction: A density functional theory approach. Appl. Surf. Sci. 2021, 537, 147894. [Google Scholar] [CrossRef]
  52. Zhu, T.; Huang, J.; Huang, B.; Zhang, N.; Liu, S.; Yao, Q.; Haw, S.; Chang, Y.; Pao, C.; Chen, J.; et al. High-Index Faceted RuCo Nanoscrews for Water Electrosplitting. Adv. Energy Mater. 2020, 10, 2002860. [Google Scholar] [CrossRef]
Figure 1. The synthetic procedure for Rux/C−T catalysts.
Figure 1. The synthetic procedure for Rux/C−T catalysts.
Catalysts 13 00552 g001
Figure 2. (a) TEM image of Ru0.3/C−800 (inset: the corresponding Ru NPs size distribution histogram); (b) HRTEM image of Ru0.3/C−800; (c) HAADF-STEM image of Ru0.3/C−800; (di) Elemental mappings for C, N, O, Ru, Na, and Cl in Ru0.3/C−800, respectively.
Figure 2. (a) TEM image of Ru0.3/C−800 (inset: the corresponding Ru NPs size distribution histogram); (b) HRTEM image of Ru0.3/C−800; (c) HAADF-STEM image of Ru0.3/C−800; (di) Elemental mappings for C, N, O, Ru, Na, and Cl in Ru0.3/C−800, respectively.
Catalysts 13 00552 g002
Figure 3. (a) XRD patterns of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900. and Ru0.3/C−1000; (b) XPS spectra of Na 1s of Ru0.3/C−800 and C−800; (c) XPS spectra of Ru 3p of Ru0.3/C−800 and Ru0.3/C−800−WF; (d) N2 adsorption/desorption curves of Ru0.3/C−800 (inset: pore size distribution plot).
Figure 3. (a) XRD patterns of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900. and Ru0.3/C−1000; (b) XPS spectra of Na 1s of Ru0.3/C−800 and C−800; (c) XPS spectra of Ru 3p of Ru0.3/C−800 and Ru0.3/C−800−WF; (d) N2 adsorption/desorption curves of Ru0.3/C−800 (inset: pore size distribution plot).
Catalysts 13 00552 g003
Figure 4. (a) LSV curves of C−800, Ru0.2/C−800, Ru0.3/C−800, Ru0.4/C−800, and 20 wt% Pt/C in 1.0 M KOH; (b) Tafel plots of C−800, Ru0.2/C−800, Ru0.3/C−800, Ru0.4/C−800, and 20 wt% Pt/C in 1.0 M KOH; (c) Nyquist plots of C−800, Ru0.2/C−800, Ru0.3/C−800 and Ru0.4/C−800 in 1.0 M KOH; (d) linear plot of the capacitive current against the scan rate for the estimation of Cdl.
Figure 4. (a) LSV curves of C−800, Ru0.2/C−800, Ru0.3/C−800, Ru0.4/C−800, and 20 wt% Pt/C in 1.0 M KOH; (b) Tafel plots of C−800, Ru0.2/C−800, Ru0.3/C−800, Ru0.4/C−800, and 20 wt% Pt/C in 1.0 M KOH; (c) Nyquist plots of C−800, Ru0.2/C−800, Ru0.3/C−800 and Ru0.4/C−800 in 1.0 M KOH; (d) linear plot of the capacitive current against the scan rate for the estimation of Cdl.
Catalysts 13 00552 g004
Figure 5. (a) LSV curves of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 1.0 M KOH; (b) Tafel plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (c) Nyquist plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (d) LSV curves of Ru0.3/C−800 and Ru0.3/C−800−WF in 1.0 M KOH; (e) relationship diagram of Ru0 content and Run+ content and overpotential of Ru0.3/C−800, Ru0.3/C−900, Ru0.3/C−1000, and Ru0.3/C−800−WF; (f) LSV curves of Ru0.3/C−800 before and after 1000 CV tests (inset: the i–t curve of Ru0.3/C−800); (g) LSV curves of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 0.5 M H2SO4; (h) Tafel plots of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (i) Nyquist plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000.
Figure 5. (a) LSV curves of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 1.0 M KOH; (b) Tafel plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (c) Nyquist plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (d) LSV curves of Ru0.3/C−800 and Ru0.3/C−800−WF in 1.0 M KOH; (e) relationship diagram of Ru0 content and Run+ content and overpotential of Ru0.3/C−800, Ru0.3/C−900, Ru0.3/C−1000, and Ru0.3/C−800−WF; (f) LSV curves of Ru0.3/C−800 before and after 1000 CV tests (inset: the i–t curve of Ru0.3/C−800); (g) LSV curves of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000 in 0.5 M H2SO4; (h) Tafel plots of Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000; (i) Nyquist plots of Ru0.3/C−700, Ru0.3/C−800, Ru0.3/C−900, and Ru0.3/C−1000.
Catalysts 13 00552 g005
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guo, B.; Zhao, C.; Zhou, Y.; Guo, J.; Wei, Z.; Wang, J. The Promotional Effect of Na on Ru for pH-Universal Hydrogen Evolution Reactions. Catalysts 2023, 13, 552. https://doi.org/10.3390/catal13030552

AMA Style

Guo B, Zhao C, Zhou Y, Guo J, Wei Z, Wang J. The Promotional Effect of Na on Ru for pH-Universal Hydrogen Evolution Reactions. Catalysts. 2023; 13(3):552. https://doi.org/10.3390/catal13030552

Chicago/Turabian Style

Guo, Bingxin, Chengfei Zhao, Yingshuang Zhou, Junjie Guo, Zhongzhe Wei, and Jing Wang. 2023. "The Promotional Effect of Na on Ru for pH-Universal Hydrogen Evolution Reactions" Catalysts 13, no. 3: 552. https://doi.org/10.3390/catal13030552

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