Next Article in Journal
Two-Dimensional Hetero- to Homochiral Phase Transition from Dynamic Adsorption of Barbituric Acid Derivatives
Previous Article in Journal
Construction of Type-II Heterojunctions in Crystalline Carbon Nitride for Efficient Photocatalytic H2 Evolution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Construction of an Amethyst-like MoS2@Ni9S8/Co3S4 Rod Electrocatalyst for Overall Water Splitting

1
State Key Laboratory of Silicon and Advanced Semiconductor Materials, School of Materials Science and Engineering, Zhejiang University, Hangzhou 310058, China
2
Hangzhou Global Scientific and Technological Innovation Center, Zhejiang University, Hangzhou 311200, China
*
Author to whom correspondence should be addressed.
Nanomaterials 2023, 13(16), 2302; https://doi.org/10.3390/nano13162302
Submission received: 3 July 2023 / Revised: 7 August 2023 / Accepted: 8 August 2023 / Published: 10 August 2023
(This article belongs to the Special Issue Nanomaterials towards Electrocatalysis)

Abstract

:
Transition metal sulphide electrocatalytic materials possess the bright overall water-splitting performance of practical electrocatalytic technologies. In this study, an amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst was constructed via a one-step hydrothermal method with in-situ-grown ZIF-67 nanoparticles on nickel foam (NF) as a precursor. The rational design and synthesis of MoS2@Ni9S8/Co3S4 endow the catalyst with neat nanorods morphology and high conductivity. The MoS2@Ni9S8/Co3S4/NF with the amethyst-like rod structure exposes abundant active sites and displays fast electron-transfer capability. The resultant MoS2@Ni9S8/Co3S4/NF exhibits outstanding hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) electrocatalytic activities, with low overpotentials of 81.24 mV (HER) at 10 mA cm−2 and 159.67 mV (OER) at 50 mA cm−2 in 1.0 M KOH solution. The full-cell voltage of overall water splitting only achieves 1.45 V at 10 mA cm−2. The successful preparation of the amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst provides a reliable reference for obtaining efficient electrocatalysts for overall water splitting.

1. Introduction

Studies have long focused on developing environmentally friendly and recyclable clean energy materials to replace depleted fossil fuels [1,2,3]. In this respect, hydrogen has become one of the most important material candidates for future energy technologies, owing to its cleanliness, renewability, and high calorific value [4,5,6]. Compared with numerous other hydrogen-production methods (e.g., coal gasification, steam methane reforming, biomass conversion, and photocatalytic hydrogen [7]), water-splitting technology has more potential since it only needs a certain voltage to produce high-purity hydrogen simply and without any pollution [8]. Typically, water splitting produces hydrogen and oxygen via the hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) [9]. In the alkaline solution, the HER usually occurs through Volmer–Heyrovsky or Volmer–Tafel mechanism as follows [10,11]:
H 2 O + e + H * + O H V o l m e r   s t e p
H 2 O + e + H H 2 + O H + H e y r o v s k y   S t e p
H * + H * H 2 + 2 ( T a f e l   s t e p )
wherein   is the active adsorption site of the catalyst and ( H * ) is the adsorbed hydrogen. The OER mechanism can be described as follows [12]:
O H + O H * + e
O H * + O H O * + H 2 O + e
O * + O H O O H * + e
O O H * + O H O 2 + + H 2 O + e
wherein   is the active adsorption site, and  O H * O * , and  O O H *  are the intermediates. Compared with the HER, the OER involves more intermediates and has a more complicated reaction process. Due to the confinement of material active sites and the complex electron-transfer process of the OER, a high potential is typically required to drive the actual water splitting [13]. Electrocatalysts based on noble metals (Pt and Ru/Ir) can effectively reduce the overpotential of the HER and OER in the water-splitting process. However, the rare reserves and high cost of noble-metal-based electrocatalysts prevent them from being applied to long-term and large-scale hydrogen production in the future [14,15,16]. Therefore, the development of low-cost bifunctional non-noble metal electrocatalysts with high activity is key to achieving long-term and efficient hydrogen production through water splitting [17,18,19].
Earth-abundant transition metals have attracted considerable attention because of their good catalytic activity and low cost [11,12]. To date, numerous transition-metal-based electrocatalytic materials have been studied, including Ni-based sulphides [10,20], Co-based phosphides/sulphides [8,21,22,23], and Mo-based sulphides [24,25]. Among these materials, Ni-, Co-, and Mo-based sulphides are considered potential electrocatalysts for water splitting because of their abundant active sites and adjustable electronic properties [26]. However, single-component sulphide materials typically exhibit monofunctional catalytic activity for either the hydrogen evolution reaction (HER) or oxygen evolution reaction (OER). Strategies by combining different kinds of transition metal sulphides can effectively improve the activity of catalysts and achieve bifunctional catalytic performance [27,28,29]. Chen et al. constructed a Ni9S8/MoS2 nanosheet-modified NiMoO4 nanorod electrocatalyst with overpotentials of 190 (HER) and 360 mV (OER) at a current density of 10 mA cm−2 [27]. Yang et al. synthesized a hierarchical nanoassembly MoS2/Co9S8/Ni3S2/Ni electrocatalyst, which exhibited high HER and OER bifunctional catalytic activities in a wide pH range [30].
Moreover, designing and regulating the structural morphology of electrocatalytic materials are also effective strategies to enhance the activity of electrocatalysts [21]. Recently, the construction of unique electrocatalyst morphologies by incorporating zeolitic imidazolate frameworks (ZIFs) has attracted significant interest [24,31,32,33]. The introduction of ZIF-67 during the preparation of electrocatalysts can not only regulate the structure of the electrocatalyst but also introduce the Co element [33,34,35,36,37,38,39,40]. For example, Guo et al. synthesized hollow Co3S4@MoS2 derived from ZIF-67 using a two-step hydrothermal and calcination method, whereby hollow Co3S4@MoS2 exhibited an overpotential of 330 mV (HER) at 10 mA cm−2 [41]. Peng et al. combined carbon nanofibers and ZIF-67 to prepare tubular electrocatalytic materials with rough surfaces that exhibited enhanced catalytic activity for the OER after calcination [42]. However, most electrocatalytic materials derived from ZIF-67 exhibited high overpotentials, owing to the poor conductivity of ZIF-67 [13,24]. Although high-temperature calcination can enhance the conductivity of electrocatalysts, it may destroy the structural characteristics and active catalytic sites of the catalyst [43,44]. According to the previous reports, ZIF-67 can be etched by the sulphide to release Co2+ under hydrothermal conditions, and the conductivity and exposed active sites of the ZIF-derived electrocatalysts can be improved [45]. However, most of the research directly added ZIF-67 powder into the precursor solution to regulate reactions, which may lead to the serious agglomeration of the prepared electrocatalyst [33,46]. The preparation of electrocatalysts by a one-step hydrothermal method using ZIF-67/NF as a precursor is rarely reported. Ge et al. reported a non-agglomerated MoS2/CoP electrocatalyst exhibiting enhanced HER activity prepared by hydrothermal reaction of ZIF-67 precursor grown on titanium foil (TF) [47].
Herein, we demonstrate the preparation of an amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst with high conductivity and rich active sites via a convenient hydrothermal method. Benefiting from the synergistic effects and improved electronic environment, MoS2@Ni9S8/Co3S4/NF shows low overpotentials for the HER at 10 mA cm−2 and the OER at 50 mA cm−2 in alkaline solution and possesses a low full-cell voltage of overall water splitting at 10 mA cm−2 and excellent electrocatalytic stability.

2. Materials and Methods

2.1. Materials

Nickel foam (0.5 mm thickness) was purchased from Saibo Electrochemical Materials (China). Co(NO3)2·6H2O, Na2MoO4·2H2O, 2-methylimidazole (C4H6N2), and thioacetamide (C2H5NS) were purchased from Aladdin Chemical Reagent. Poly(sodium-p-styrene sulfonate) (C8H7NaO3S)n was purchased from Macklin Biochemical Reagent, and methanol (CH3OH) was purchased from Sinopharm Chemical Reagent. All the reagents and chemicals were used without further purification.

2.2. Preparation of the ZIF-67/NF Precursor

The ZIF-67 precursor was directly grown on NF substrate via a simple process according to a previously reported method with some modifications [42]. First, NF (2 × 3 cm2) was cleaned with 3 M HCl solution, deionized water, and ethanol for 10 min under ultrasonication in successive order and dried at 70 °C for 12 h. Next, 40 mg poly(sodium-p-styrene sulfonate) was added to 10 mL deionized water to form solution I. To achieve surface modification, the as-cleaned NF was soaked in solution I for 30 min under ultrasonication and then washed thrice with deionized water. Next, 40 mmol 2-methylimidazole (2-MeIM) was distributed in 50 mL methanol to form solution II, and 5 mmol Co(NO3)2·6H2O was added to 50 mL methanol to form solution III. The surface-modified NF was fully immersed in solution II for 30 min. Subsequently, solution III was poured into the resulting solution II with the surface-modified NF solution mixture and agitated for 30 min. The obtained blend was then allowed to stand at room temperature (25 °C) for 24 h. The desired product was obtained and washed thrice with deionized water. Finally, the product was dried at 70 °C in a vacuum oven for 12 h and named ZIF-67/NF.

2.3. Synthesis of Amethyst-like MoS2@Ni9S8/Co3S4 Rods

First, 1 mmol Na2MoO4·2H2O and 3 mmol thioacetamide (C2H5NS) were dissolved in 40 mL deionized water. The resultant solution was agitated for 30 min, after which one piece of the as-prepared ZIF-67/NF and the above solution was transferred into a 50 mL autoclave, where it was maintained at 180 °C for 6 h. Subsequently, the resultant sample was collected and washed several times with ethanol and deionized water. Finally, the amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst was obtained after drying at 70 °C in a vacuum oven for 12 h. For comparison, MoS2/Ni9S8/NF without a ZIF-67 precursor, Mo-doped ZIF-67/NF without an S source, and S-doped ZIF-67/NF without a Mo source were synthesized using a preparation process similar to that of the amethyst-like MoS2@Ni9S8/Co3S4 rods.

2.4. Material Characterization

The crystalline structures of all materials were characterized using a PANalytical X’Pert X-ray diffractometer (bruker D8 ADVACNCE, Bruker, Mannheim, Germany) with a Cu Kα radiation source at 30 kV. The morphologies and structures of the samples were observed using a Gemini SEM 300 scanning electron microscope (SEM) (Gemini, Friedrichshafen, Germany) and a JEOL JEM-2100F transmission electron microscope (JEOL Co., Ltd., Tokyo, Japan). X-ray photoelectron spectroscopy was conducted using a Thermo Scientific K-Alpha instrument (Waltham, MA, USA).

2.5. Electrochemical Measurements

All electrochemical tests were performed using a three-electrode system in a 1 M KOH electrolyte on a CHI 660E electrochemical workstation (CHI 660E, Shanghai, China). The as-prepared catalyst (1 × 1 cm2) was used as the working electrode, and a graphite rod and Hg/HgO were used as the counter and reference electrodes, respectively. All potentials were converted to the reversible hydrogen electrode (RHE) potential according to the following equation:  E R H E = E H g / H g O + 0.0591 p H + 0.098  [18]. Linear sweep voltammetry (LSV) was performed from −1.85 to −0.70 V (−0.20 to 1.40 V) vs. Hg/HgO for the hydrogen evolution reaction (HER) (oxygen evolution reaction (OER)) at a scan rate of 2 mV/s. Electrochemical impedance spectroscopy (EIS) was performed in the frequency range of 100 KHz to 0.01 Hz at −1.03 V vs. Hg/HgO (0.56 V) for the HER (OER). The double-layer capacitance (Cdl) and electrochemical surface area (ECSA) were obtained using cyclic voltammetry (CV) at scan rates of 20–100 mV s−1 in the range of −0.30 to −0.20 V vs. Hg/HgO. All polarization curves were corrected using iR compensation.

3. Results

An amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst was synthesized via a one-pot hydrothermal reaction (Figure 1). A precursor comprising ZIF-67 deposited on NF was first obtained by the hydrothermal reaction at room temperature. As shown in Figure 2e, the prepared ZIF-67/NF precursor presented a classical uniform polyhedral morphology with a size range of 100–300 nm [48]. Subsequently, in the hydrothermal process, the ZIF-67/NF immersed in the solution first reacted with sulphides to release Co ions and generate Co3S4, resulting in the structural evolution of ZIF-67/NF [45]. Compared with S-ZIF-67/NF (Figure 2g), it can be seen that the structure of ZIF-67/NF became dense with sharp protrusions after reacting with sulphides, which may have a good structure-orienting effect for the further formation of nanorods on it. Then, the Ni from the NF reacted with sulphides to generate Ni9S8 nanorods, and Mo reacted with sulphides to form MoS2 [46,47]. As the reaction proceeded, eventually, an amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst was successfully obtained via this one-step hydrothermal method.
Figure 2 shows SEM images of the different samples. The microstructures of MoS2@Ni9S8/Co3S4/NF at different magnifications are shown in Figure 2a–c. A uniform structure with a neat nanorods array of MoS2@Ni9S8/Co3S4/NF is shown in Figure 2a, whereby the structural characteristics of a single nanorod are similar to those of an amethyst-like rod (Figure 2b). Each nanorod presents a rough surface and polygonal morphology, with a diameter in the range of 200–500 nm (Figure 2c). ZIF-67 particles disappear compared with the precursor before sulphuration, indicating that ZIF-67 completely reacted with sulphides. The non-agglomerated rod-shaped structure of MoS2@Ni9S8/Co3S4/NF can expose rich active sites and facilitate the detachment of bubbles from the catalyst during water splitting [27,49]. The morphologies of the other comparative samples are significantly different from that of MoS2@Ni9S8/Co3S4/NF. Except for the ZIF-67/NF (Figure 2e) and S-ZIF-67/NF (Figure 2g) discussed above, MoS2/Ni9S8/NF without ZIF-67 exhibits an open structure composed of nonuniform-sized blocks (Figure 2d). When the Mo source is introduced to the sample, Mo-ZIF-67/NF exhibits a network structure composed of nanosheets (Figure 2f). The above results fully demonstrate that the Mo, S sources, and ZIF-67/NF precursor are essential for generating an amethyst-like rod structure of MoS2@Ni9S8/Co3S4/NF. The EDS mapping images of MoS2@Ni9S8/Co3S4/NF indicate a uniform distribution of elemental Mo, Ni, and S (Figure 2h). The structure and composition of MoS2@Ni9S8/Co3S4/NF were further observed using TEM (Figure 2i). The results reveal that the MoS2@Ni9S8/Co3S4/NF nanorod comprises a rough surface with a non-hollow structure. An HR-TEM analysis of MoS2@Ni9S8/Co3S4/NF (Figure 2j) reveals the presence of lattice fringes with 0.42 nm plane spacings belonging to the (201) crystal plane of Ni9S8 and ones of approximately 0.62 and 0.26 nm corresponding to the (002) and (101) crystal planes of MoS2.
XRD tests were used to further determine the composition of MoS2@Ni9S8/Co3S4/NF. As shown in Figure 3, all sample patterns display strong diffraction peaks at 44.5°, 51.8°, and 76.4° corresponding to Ni foam (PDF#87-0712) (111), (200), and (220) crystal planes, respectively, which are derived from the NF substrate. Owing to the strong Ni peak, the diffraction peaks of ZIF-67/NF and Mo-ZIF-67/NF are less obvious. S-ZIF-67/NF exhibits clear diffraction peaks at 31.4°, 38.1°, 50.2°, 55.1°, and 77.8° that correspond to the (311), (400), (511), (440), and (731) crystal planes of Co3S4 (PDF#47-1738), and the peaks at 21.5°, 38.6°, 50.5°, 55.5°, 69.6°, and 72.9° could be indexed to (201), (241), (153), (530), (082), and (207) crystal planes of Ni9S8 (PDF#22-1193), respectively. MoS2/Ni9S8/NF and MoS2@Ni9S8/Co3S4/NF present weak peaks of MoS2 (PDF#37-1492) and Ni9S8 (PDF#22-1193). Thus, to reduce the influence of NF-derived Ni, MoS2@Ni9S8/Co3S4/NF powder was scraped from the NF for XRD analysis. As shown in Figure S1, the peaks at 14.4°, 32.7°, 35.9°, and 49.8° correspond to the (002), (100), (102), and (105) crystal planes of MoS2 (PDF#37-1492) [27], whereas the peaks at 15.4°, 21.5°, 24.6°, 27.2°, 31.3°, 32.5°, 37.9°, 38.6°, 50.8°, 55.5°, and 56.6° belong to the (111), (201), (022), (202), (222), (023), (330), (241), (025), (530), and (531) crystal planes of Ni9S8 (PDF#22-1193) [50]. The peaks at 16.3°, 31.5°, 38.2°, and 55.1° are indexed to the (111), (311), (400), and (440) planes of Co3S4 (PDF#47-1738), respectively. The coexistence of the MoS2, Ni9S8, and Co3S4 crystalline phases confirms that MoS2@Ni9S8/Co3S4/NF is successfully prepared.
XPS analysis was performed to characterize the surface element compositions of MoS2@Ni9S8/Co3S4/NF. The full XPS spectrum of MoS2@Ni9S8/Co3S4/NF demonstrates the presence of elemental Mo, Ni, and S (Figure 4a). In the Mo 3d spectra of MoS2@Ni9S8/Co3S4/NF (Figure 4b), the peaks at 229.10 and 232.90 eV correspond to Mo 3d5/2 and 3d3/2, respectively, which may be assigned to Mo4+, whereas those at 232.10 and 235.90 eV can be ascribed to Mo6+ [27,51]. The peak at 226.00 eV is assigned to the Mo–S bond [24]. The Ni 2p peaks of MoS2@Ni9S8/Co3S4/NF at 856.40 and 874.20 eV correspond to Ni 2p3/2 and 2p1/2, respectively, which are derived from Ni3+ [30,52]. Moreover, two satellite peaks (Sat) (Figure 4c) at 862.70 and 880.70 eV are observed. Compared with those observed in the MoS2/Ni9S8/NF without a ZIF-67 spectrum, the Mo 3d and Ni 2p binding energies of MoS2@Ni9S8/Co3S4/NF exhibit positive shifts of 0.32 and 0.20 eV, respectively, which indicates the process of losing electrons in Mo and Ni elements of MoS2@Ni9S8/Co3S4/NF [53]. According to previous reports, transition metal sulphides that lose electrons can generate more positive charges, which is conducive to the adsorption of OH and thus promotes the OER [33,54]. As shown in Figure S2, the peaks at 779.77 and 797.56 eV could correspond to Co 2p3/2 and 2p1/2, respectively, which may be assigned to Co3+ and Co2+ [55], and other peaks are ascribed to satellite peaks. The peaks at 161.80 and 163.10 eV are attributed to S 2p3/2 and S 2p1/2 (Figure 4d), corresponding to the Ni–S bond [54], whereas those at 164.10 and 168.50 eV are ascribed to Mo–S and S–O bonds, respectively [27].
The HER performances of the electrocatalysts were tested using a three-electrode system in 1.0 M KOH. As shown in Figure 5a, the linear sweep voltammetry (LSV) curves reveal that MoS2@Ni9S8/Co3S4/NF has the best HER performance. Figure 5b shows the overpotential values of all the samples at 10 and 200 mA cm−2. At a low current density of 10 mA cm−2, MoS2@Ni9S8/Co3S4/NF exhibits the lowest overpotential of 81.24 mV, which is superior to MoS2/Ni9S8/NF (96.08 mV), Mo-ZIF-67/NF (187.80 mV), ZIF-67/NF (250.41 mV), and S-ZIF-67/NF (256.63 mV). The optimal HER performance of MoS2@Ni9S8/Co3S4/NF originates from the high conductivity and the unique nanorods structure-exposed rich active sites [23,33]. Meanwhile, the improved electronic environment of MoS2@Ni9S8/Co3S4/NF enhances the adsorption of hydrogen-containing species and accelerates the HER [21,53]. At a high current density of 200 mA cm−2, MoS2@Ni9S8/Co3S4/NF also exhibits superior HER catalytic activity, with an overpotential of 161.35 mV. Comparing the previously reported Mo/Co/Ni-S electrocatalysts for the HER shown in Table S1, the amethyst-like MoS2@Ni9S8/Co3S4/NF rod electrocatalyst presents the outstanding HER catalytic performance with a lower potential. The Tafel slope was used to study the electrocatalytic kinetics of electrocatalysts. As shown in Figure 5c, The Tafel slopes of MoS2@Ni9S8/Co3S4/NF, MoS2/Ni9S8/NF, ZIF-67/NF, Mo-ZIF-67/NF, and S-ZIF-67/NF are 50.69, 56.60, 96.48, 116.83, and 118.48 mV dec−1, respectively. The lower Tafel slope of MoS2@Ni9S8/Co3S4/NF indicates more-efficient HER electrocatalytic kinetics [9,23,27]. In addition, the electrocatalytic kinetics were investigated using electrochemical impedance (EIS) analysis. The charge transfer resistance (Rct) is related to the electrocatalytic kinetics at the electrolyte–electrode interface. Generally, a smaller Rct represents a higher electron-transfer velocity [30]. Typically, the semicircle diameter in Nyquist plots is positively correlated with the value of Rct, and the specific Rct value can be obtained through equivalent circuit fitting [56]. Nyquist plots for all the synthesized samples are shown in Figure 5d, wherein MoS2@Ni9S8/Co3S4/NF shows the smallest semicircle diameter. The Rct values of MoS2@Ni9S8/Co3S4/NF, MoS2/Ni9S8/NF, Mo-ZIF-67/NF, and S-ZIF-67/NF are determined as 2.30, 3.27, 8.72, and 23.93 Ω, respectively. ZIF-67/NF presents the highest Rct value (39.49 Ω), which is consistent with the previously reported conclusion that ZIF-67 materials have poor conductivity [8,13,33]. The smallest Rct of MoS2@Ni9S8/Co3S4/NF reflects the higher electron-transfer velocity and improved conductivity of MoS2@Ni9S8/Co3S4/NF [21].
Figure 6a shows the LSV curves of the as-prepared electrocatalysts for the OER. Similar to the HER test results, MoS2@Ni9S8/Co3S4/NF exhibits superior OER activity. As shown in Figure 6b, the overpotential of MoS2@Ni9S8/Co3S4/NF is 159.67 mV at 50 mA cm−2, which is lower than those of MoS2/Ni9S8/NF (194.97 mV), ZIF-67/NF (417.20 mV), Mo-ZIF-67/NF (423.97 mV), and S-ZIF-67/NF (439.77 mV). Moreover, MoS2@Ni9S8/Co3S4/NF only requires an overpotential of 230.05 mV to reach a current density of 100 mA cm−2. The outstanding OER performance of MoS2@Ni9S8/Co3S4/NF is related to two factors: on the one hand, the combination of Ni9S8 and Co3S4 with OER catalytic activity can produce an effective synergistic effect of components [50,57]. on the other hand, according to the XPS test results, the slight shift of Ni binding energy indicates an improved electronic environment around Ni9S8, which makes it easier to adsorb OH and thus promote the OER rate [22,52,53]. Compared with some previously reported electrocatalysts for the OER (Table S2), MoS2@Ni9S8/Co3S4/NF exhibits competitive OER performance. Figure 6c shows the Tafel plots of the electrocatalysts used for the OER. The Tafel slope of MoS2@Ni9S8/Co3S4/NF (48.75 mV dec−1) is lower than those of MoS2/Ni9S8/NF (71.06 mV dec−1), ZIF-67/NF (92.36 mV dec−1), Mo-ZIF-67/NF (101.54 mV dec−1), and S-ZIF-67/NF (104.27 mV dec−1), thus reflecting the faster OER catalytic kinetics of MoS2@Ni9S8/Co3S4/NF. EIS analysis is performed to further study the electrocatalytic kinetics of the OER (Figure 6d). It can be seen that MoS2@Ni9S8/Co3S4/NF shows the smallest semicircle diameter. The Rct values of MoS2@Ni9S8/Co3S4/NF, MoS2/Ni9S8/NF, ZIF-67, Mo-ZIF-67/NF, and S-ZIF-67/NF are determined to be 2.33, 2.80, 248.50, 123.30, and 241.90 Ω, respectively. Notably, the Rct value of MoS2@Ni9S8/Co3S4/NF is significantly lower than those of the other samples, indicating that MoS2@Ni9S8/Co3S4/NF possesses the best electron-transfer ability in the OER process [53].
Electrochemical active surface area (ECSA) is a vital parameter for evaluating the performance of electrocatalysts [58]. Typically, the ECSA can be determined using the formula:   E C S A = C d l / C S , where ( C d l ) is the double-layer capacitance and  C s  is the specific capacitance and is generally calculated by using 40.0 μF cm−2 [59]. The cyclic voltammetry (CV) curves of the as-prepared samples were measured across the potential range from −0.30 to −0.20 V (vs. Hg/HgO) at scanning speeds of 20–100 mV s−1. The CV curves of MoS2@Ni9S8/Co3S4/NF are shown in Figure 7a, whereas those of MoS2/Ni9S8/NF, ZIF-67/NF, Mo-ZIF-67/NF, and S-ZIF-67/NF are shown in Figure S3a–d, respectively. The  C d l  can be obtained by fitting the relationship between half the current density at −0.25 V and the scanning speeds. The  C d l  values of MoS2@Ni9S8/Co3S4/NF, MoS2/Ni9S8/NF, ZIF-67/NF, Mo-ZIF-67/NF, and S-ZIF-67/NF are determined to be 45.32, 17.85, 0.66, 1.54, and 1.68 mF cm−2 (Figure 7b), and the corresponding ECSAs are 1133.0, 446.3, 16.4, 38.5, and 42.0 cm2, respectively. Notably, the ECSA of MoS2@Ni9S8/Co3S4/NF is 2.5 and 69 times larger than those of MoS2/Ni9S8/NF and ZIF-67/NF, respectively. These results demonstrate that the MoS2@Ni9S8/Co3S4/NF comprising an amethyst-like rod structure can provide a large ECSA and expose abundant active sites.
Typically, ideal electrocatalysts must not only exhibit high HER and OER activities but also outstanding electrochemical stability. As shown in Figure 7c, during the chronopotentiometry (CP) test for the HER, the potentials of MoS2@Ni9S8/Co3S4/NF maintain a linear shape after 10, 50, and 100 mA cm−2 for 24 h, displaying excellent HER stability. Similarly, Figure 7d reveals the positive OER durability of MoS2@Ni9S8/Co3S4/NF. The 1000 CV cycles test further evaluated the stability of MoS2@Ni9S8/Co3S4/NF. As shown in Figure S4a, the LSV curves completely coincide before and after 1000 CV cycles for the HER, with almost no change in the overpotential. However, the overpotential of MoS2@Ni9S8/Co3S4/NF increases at the same current density for the OER after 1000 CV cycles (Figure S4b). Compared to that of the HER (Figure S5a), the SEM images reveal a more-pronounced degradation of MoS2@Ni9S8/Co3S4/NF morphology after 1000 CV cycles for the OER (Figure S5b), but the catalyst still attaches to the NF. The relatively weakened OER durability may be derived from the apparent degradation of MoS2@Ni9S8/Co3S4/NF morphology after 1000 CV cycles, which is due to the violent generation of bubbles during the OER process. Moreover, the samples after 1000 CV cycles were used for the XRD test. It can be seen that no other new diffraction peaks of samples appear after 1000 CV cycles (Figure S6). In the sample after 1000 CV cycles for the HER, the peaks at 31.3°, 37.9°, 40.2°, and 55.5° correspond to the (222), (330), (114), and (530) crystal planes of Ni9S8 (PDF#22-1193), respectively. The peaks at 33.5°, 35.8°, and 60.1° correspond to the (101), (102), and (008) crystal planes of MoS2 (PDF#37-1492), respectively. And in the sample after 1000 cycles for the OER, the peaks at 31.3°, 37.9°, and 55.5° correspond to the (222), (330), and (530) crystal planes of Ni9S8 (PDF#22-1193), respectively. The peaks at 39.5°, 60.1°, and 85.1° correspond to the (103), (008), and (206) crystal planes of MoS2 (PDF#37-1492), respectively. Collectively, these results show that the MoS2@Ni9S8/Co3S4/NF electrocatalyst possesses outstanding electrochemical stability.
Owing to the superior electrochemical activities of MoS2@Ni9S8/Co3S4/NF for the HER and OER, MoS2@Ni9S8/Co3S4/NF was applied as a bifunctional catalyst for an overall water-splitting test in a two-electrode system. The LSV curve (Figure 8a) shows that the MoS2@Ni9S8/Co3S4/NF has a low full-cell voltage of 1.45 V at 10 mA cm−2, and the cell voltage increases by only 0.03 V after holding at 10 mA cm−2 for 19 h (Figure 8b). These results demonstrate that MoS2@Ni9S8/Co3S4/NF has excellent bifunctional electrocatalytic activity and stability for overall water splitting.

4. Conclusions

In summary, we successfully constructed an amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst using a ZIF-67/NF precursor and a one-step hydrothermal method. By adopting a simple synthesis strategy, MoS2@Ni9S8/Co3S4/NF possesses high conductivity and numerous active edge sites. Meanwhile, the synergistic effect produced by the composite of MoS2, Ni9S8, and Co3S4 further improves the catalytic activity of the electrocatalyst. Therefore, MoS2@Ni9S8/Co3S4/NF exhibits lower overpotentials and outstanding electrochemical stability in a 1.0 M KOH solution. These results confirm that the as-prepared amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst possesses excellent bifunctional activity for overall water splitting.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano13162302/s1, Figure S1: XRD patterns of MoS2@Ni9S8/Co3S4/NF; Figure S2. Co 2p XPS spectra of MoS2@Ni9S8/Co3S4/NF; Table S1. Comparison of the HER overpotentials of different electrocatalysts; Table S2. Comparison of the OER overpotentials of different electrocatalysts; Figure S3. CV curves of the samples at different scan rates: (a) MoS2/Ni9S8/NF, (b) ZIF-67/NF, (c) Mo-ZIF-67/NF, and (d) S-ZIF-67/NF; Figure S4. Performance of MoS2@Ni9S8/Co3S4/NF during the stability test: (a) LSV curves for the HER after 1000 CV cycles, (b) LSV curves for the OER after 1000 CV cycles; Figure S5. SEM images of MoS2@Ni9S8/Co3S4/NF: (a) after 1000 CV cycles for the HER, (b) after 1000 CV cycles for the OER; Figure S6. XRD patterns of samples after 1000 CV cycles. References [18,20,23,27,30,33,60,61,62] are cited in the supplementary materials.

Author Contributions

Conceptualization, Investigation, Data curation, Writing—Original Draft: Z.P.; Methodology and Formal Analysis: T.Q.; Formal Analysis and Validation: R.T.; Formal Analysis and Visualization: Y.O.; Writing—Review and Editing, Supervision, Resources, and Investigation: X.G. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the National Natural Science Foundation of China (Grants 21875217 and 51372225).

Data Availability Statement

The data presented in this study are available in this article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, J.; An, Y.; Guo, K.; Ren, X.; Jiang, B. Nitrogen doped FeCoNiS nanoparticles on N, S-co-doped vertical graphene as bifunctional electrocatalyst for water splitting. Int. J. Hydrogen Energy 2023, 48, 4143–4157. [Google Scholar] [CrossRef]
  2. Zhu, J.; Hu, L.; Zhao, P.; Lee, L.Y.S.; Wong, K.Y. Recent Advances in Electrocatalytic Hydrogen Evolution Using Nanoparticles. Chem. Rev. 2020, 120, 851–918. [Google Scholar] [CrossRef]
  3. Huang, C.Q.; Yu, L.; Zhang, W.; Xiao, Q.; Zhou, J.Q.; Zhang, Y.L.; An, P.F.; Zhang, J.; Yu, Y. N-doped Ni-Mo based sulfides for high-efficiency and stable hydrogen evolution reaction. Appl. Catal. B Environ. 2020, 276, 119137. [Google Scholar] [CrossRef]
  4. Zhang, S.; Zhang, X.; Rui, Y.; Wang, R.; Li, X. Recent advances in non-precious metal electrocatalysts for pH-universal hydrogen evolution reaction. Green Energy Environ. 2021, 6, 458–478. [Google Scholar] [CrossRef]
  5. El-Refaei, S.M.; Russo, P.A.; Pinna, N. Recent Advances in Multimetal and Doped Transition-Metal Phosphides for the Hydrogen Evolution Reaction at Different pH values. ACS Appl. Mater. Interfaces 2021, 13, 22077–22097. [Google Scholar] [CrossRef]
  6. Chang, Z.; Zhu, L.; Zhao, J.; Chen, P.; Chen, D.; Gao, H. NiMo/Cu-nanosheets/Ni-foam composite as a high performance electrocatalyst for hydrogen evolution over a wide pH range. Int. J. Hydrogen Energy 2021, 46, 3493–3503. [Google Scholar] [CrossRef]
  7. Mamiyev, Z.; Balayeva, N.O. Metal Sulfide Photocatalysts for Hydrogen Generation: A Review of Recent Advances. Catalysts 2022, 12, 1316. [Google Scholar] [CrossRef]
  8. Qian, L.-H.; Dong, W.-W.; Li, H.-B.; Song, X.; Ding, Y. Fe-doped NiCo2S4 catalyst derived from ZIF-67 towards efficient hydrogen evolution reaction. Int. J. Hydrogen Energy 2022, 47, 8820–8828. [Google Scholar] [CrossRef]
  9. Yu, Z.Y.; Duan, Y.; Feng, X.Y.; Yu, X.; Gao, M.R.; Yu, S.H. Clean and Affordable Hydrogen Fuel from Alkaline Water Splitting: Past, Recent Progress, and Future Prospects. Adv. Mater 2021, 33, 2007100. [Google Scholar] [CrossRef]
  10. Guo, Y.; Park, T.; Yi, J.W.; Henzie, J.; Kim, J.; Wang, Z.; Jiang, B.; Bando, Y.; Sugahara, Y.; Tang, J.; et al. Nanoarchitectonics for Transition-Metal-Sulfide-Based Electrocatalysts for Water Splitting. Adv. Mater. 2019, 31, 1807134. [Google Scholar] [CrossRef]
  11. Yu, F.; Yu, L.; Mishra, I.K.; Yu, Y.; Ren, Z.F.; Zhou, H.Q. Recent developments in earth-abundant and non-noble electrocatalysts for water electrolysis. Mater. Today Phys. 2018, 7, 121–138. [Google Scholar] [CrossRef]
  12. Xu, Y.; Wang, C.; Huang, Y.; Fu, J. Recent advances in electrocatalysts for neutral and large-current-density water electrolysis. Nano Energy 2021, 80, 105545. [Google Scholar] [CrossRef]
  13. Peng, W.; Zheng, G.; Wang, Y.; Cao, S.; Ji, Z.; Huan, Y.; Zou, M.; Yan, X. Zn doped ZIF67-derived porous carbon framework as efficient bifunctional electrocatalyst for water splitting. Int. J. Hydrogen Energy 2019, 44, 19782–19791. [Google Scholar] [CrossRef]
  14. Li, Y.; Peng, C.K.; Hu, H.; Chen, S.Y.; Choi, J.H.; Lin, Y.G.; Lee, J.M. Interstitial boron-triggered electron-deficient Os aerogels for enhanced pH-universal hydrogen evolution. Nat. Commun. 2022, 13, 1143. [Google Scholar] [CrossRef]
  15. Pi, C.; Huang, C.; Yang, Y.; Song, H.; Zhang, X.; Zheng, Y.; Gao, B.; Fu, J.; Chu, P.K.; Huo, K. In situ formation of N-doped carbon-coated porous MoP nanowires: A highly efficient electrocatalyst for hydrogen evolution reaction in a wide pH range. Appl. Catal. B Environ. 2020, 263, 118358. [Google Scholar] [CrossRef]
  16. Shao, L.; Qian, X.; Wang, X.; Li, H.; Yan, R.; Hou, L. Low-cost and highly efficient CoMoS4/NiMoS4-based electrocatalysts for hydrogen evolution reactions over a wide pH range. Electrochim. Acta 2016, 213, 236–243. [Google Scholar] [CrossRef] [Green Version]
  17. Chiang, C.H.; Yang, Y.C.; Lin, J.W.; Lin, Y.C.; Chen, P.T.; Dong, C.L.; Lin, H.M.; Chan, K.M.; Kao, Y.T.; Suenaga, K.; et al. Bifunctional Monolayer WSe2/Graphene Self-Stitching Heterojunction Microreactors for Efficient Overall Water Splitting in Neutral Medium. ACS Nano 2022, 16, 18274–18283. [Google Scholar] [CrossRef]
  18. Wei, D.; Tang, W.; Wang, Y. Hairy sphere-like Ni9S8/CuS/Cu2O composites grown on nickel foam as bifunctional electrocatalysts for hydrogen evolution and urea electrooxidation. Int. J. Hydrogen Energy 2021, 46, 20950–20960. [Google Scholar] [CrossRef]
  19. Pi, Y.; Shao, Q.; Wang, P.; Guo, J.; Huang, X. General Formation of Monodisperse IrM (M = Ni, Co, Fe) Bimetallic Nanoclusters as Bifunctional Electrocatalysts for Acidic Overall Water Splitting. Adv. Funct. Mater. 2017, 27, 1700886. [Google Scholar] [CrossRef]
  20. Zhao, D.; Dai, M.; Liu, H.; Chen, K.; Zhu, X.; Xue, D.; Wu, X.; Liu, J. Sulfur-Induced Interface Engineering of Hybrid NiCo2O4@NiMo2S4 Structure for Overall Water Splitting and Flexible Hybrid Energy Storage. Adv. Mater. Interfaces 2019, 6, 1901308. [Google Scholar] [CrossRef]
  21. Tian, R.; Wang, F.; Zou, C.; Pei, Z.; Guo, X.; Yang, H. Modulating organic ligands to construct 2D–3D-hybrid porous P-doped metal-organic frameworks electrocatalyst for overall water splitting. J. Alloys Compd. 2023, 933, 167670. [Google Scholar] [CrossRef]
  22. Wang, F.; Guo, X.; He, F.; Hou, Y.; Liu, F.; Zou, C.; Yang, H. Binder free construction of hollow hierarchical Mn–Co–P nanoarrays on nickel foam as an efficient bifunctional electrocatalyst for overall water splitting. Sustain. Energy Fuels 2022, 6, 851–860. [Google Scholar] [CrossRef]
  23. Su, H.; Du, X.; Zhang, X. NiCoP coated on NiCo2S4 nanoarrays as electrode materials for hydrogen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 30910–30916. [Google Scholar] [CrossRef]
  24. Xie, Y.; Chen, L.; Jin, Q.; Yun, J.; Liang, X. MoS2–Co3S4 hollow polyhedrons derived from ZIF-67 towards hydrogen evolution reaction and hydrodesulfurization. Int. J. Hydrogen Energy 2019, 44, 24246–24255. [Google Scholar] [CrossRef]
  25. Jayabal, S.; Saranya, G.; Wu, J.; Liu, Y.; Geng, D.; Meng, X. Understanding the high-electrocatalytic performance of two-dimensional MoS2 nanosheets and their composite materials. J. Mater. Chem. A 2017, 5, 24540–24563. [Google Scholar] [CrossRef]
  26. Wang, M.; Zhang, L.; He, Y.; Zhu, H. Recent advances in transition-metal-sulfide-based bifunctional electrocatalysts for overall water splitting. J. Mater. Chem. A 2021, 9, 5320–5363. [Google Scholar] [CrossRef]
  27. Chen, L.; Deng, Z.; Chen, Z.; Wang, X. Building Ni9S8/MoS2 Nanosheets Decorated NiMoO4 Nanorods Heterostructure for Enhanced Water Splitting. Adv. Mater. Interfaces 2021, 8, 2101483. [Google Scholar] [CrossRef]
  28. Wang, P.; Zhang, X.; Zhang, J.; Wan, S.; Guo, S.; Lu, G.; Yao, J.; Huang, X. Precise tuning in platinum-nickel/nickel sulfide interface nanowires for synergistic hydrogen evolution catalysis. Nat. Commun. 2017, 8, 14580. [Google Scholar] [CrossRef]
  29. Wang, D.Y.; Gong, M.; Chou, H.L.; Pan, C.J.; Chen, H.A.; Wu, Y.; Lin, M.C.; Guan, M.; Yang, J.; Chen, C.W.; et al. Highly active and stable hybrid catalyst of cobalt-doped FeS2 nanosheets-carbon nanotubes for hydrogen evolution reaction. J. Am. Chem. Soc. 2015, 137, 1587–1592. [Google Scholar] [CrossRef]
  30. Yang, Y.; Yao, H.; Yu, Z.; Islam, S.M.; He, H.; Yuan, M.; Yue, Y.; Xu, K.; Hao, W.; Sun, G.; et al. Hierarchical Nanoassembly of MoS2/Co9S8/Ni3S2/Ni as a Highly Efficient Electrocatalyst for Overall Water Splitting in a Wide pH Range. J. Am. Chem. Soc. 2019, 141, 10417–10430. [Google Scholar] [CrossRef]
  31. Yang, Y.-W.; Song, B.-Y. A Novel Phosphide Derived From Metal-Organic Frameworks as Cost-Effective Electrocatalyst for Oxygen Evolution Reaction. J. Electrochem. Energy Convers. Storage 2022, 19, 010907. [Google Scholar] [CrossRef]
  32. Jadhav, H.S.; Bandal, H.A.; Ramakrishna, S.; Kim, H. Critical Review, Recent Updates on Zeolitic Imidazolate Framework-67 (ZIF-67) and Its Derivatives for Electrochemical Water Splitting. Adv. Mater. 2022, 34, e2107072. [Google Scholar] [CrossRef]
  33. Wu, Q.; Dong, A.; Yang, C.; Ye, L.; Zhao, L.; Jiang, Q. Metal-organic framework derived Co3O4@Mo-Co3S4-Ni3S2 heterostructure supported on Ni foam for overall water splitting. Chem. Eng. J. 2021, 413, 127482. [Google Scholar] [CrossRef]
  34. Zhao, L.; Gong, C.; Chen, X.; He, X.; Chen, H.; Du, X.; Wang, D.; Fang, W.; Zhang, H.; Li, W. ZIF-67 derived Mo-CoS2 nanoparticles embedded in hierarchically porous carbon hollow sphere for efficient overall water splitting. Appl. Surf. Sci. 2023, 623, 157030. [Google Scholar] [CrossRef]
  35. Chen, P.; Duan, X.; Li, G.; Qiu, X.; Wang, S.; Huang, Y.; Stavitskaya, A.; Jiang, H. Construction of ZIF-67/MIL-88(Fe, Ni) catalysts as a novel platform for efficient overall water splitting. Int. J. Hydrogen Energy 2023, 48, 7170–7180. [Google Scholar] [CrossRef]
  36. Wang, J.; Chaemchuen, S.; Chen, C.; Heynderickx, P.M.; Roy, S.; Verpoort, F. N-functionalized hierarchical carbon composite derived from ZIF-67 and carbon foam for efficient overall water splitting. J. Ind. Eng. Chem. 2022, 105, 222–230. [Google Scholar] [CrossRef]
  37. Zhang, Z.; Li, S.; Bu, X.; Dai, Y.; Wang, J.; Bao, X.; Wang, T. Hollow ZIF-67 derived porous cobalt sulfide as an efficient bifunctional electrocatalyst for overall water splitting. N. J. Chem. 2021, 45, 17313–17319. [Google Scholar] [CrossRef]
  38. Li, Y.; Zhu, S.; Kong, X.; Liang, Y.; Li, Z.; Wu, S.; Chang, C.; Luo, S.; Cui, Z. ZIF-67 derived Co@NC/g-C3N4 as a photocatalyst for enhanced water splitting H(2) evolution. Environ. Res. 2021, 197, 111002. [Google Scholar] [CrossRef]
  39. Qin, J.F.; Xie, J.Y.; Wang, N.; Dong, B.; Chen, T.S.; Lin, Z.Y.; Liu, Z.Z.; Zhou, Y.N.; Yang, M.; Chai, Y.M. Surface construction of loose Co(OH)(2) shell derived from ZIF-67 nanocube for efficient oxygen evolution. J. Colloid. Interface Sci. 2020, 562, 279–286. [Google Scholar] [CrossRef]
  40. Liu, H.; Jin, M.; Zhan, D.; Wang, J.; Cai, X.; Qiu, Y.; Lai, L. Stacking faults triggered strain engineering of ZIF-67 derived Ni-Co bimetal phosphide for enhanced overall water splitting. Appl. Catal. B Environ. 2020, 272, 118951. [Google Scholar] [CrossRef]
  41. Guo, Y.; Tang, J.; Qian, H.; Wang, Z.; Yamauchi, Y. One-Pot Synthesis of Zeolitic Imidazolate Framework 67-Derived Hollow Co3S4@MoS2 Heterostructures as Efficient Bifunctional Catalysts. Chem. Mater. 2017, 29, 5566–5573. [Google Scholar] [CrossRef]
  42. Peng, W.; Yang, X.; Mao, L.; Jin, J.; Yang, S.; Zhang, J.; Li, G. ZIF-67-derived Co nanoparticles anchored in N doped hollow carbon nanofibers as bifunctional oxygen electrocatalysts. Chem. Eng. J. 2021, 407, 127157. [Google Scholar] [CrossRef]
  43. Ma, F.; Jin, S.; Li, Y.; Feng, Y.; Tong, Y.; Ye, B.-C. Pyrolysis-derived materials of Mn-doped ZIF-67 for the electrochemical detection of o-nitrophenol. J. Electroanal. Chem. 2022, 904, 115932. [Google Scholar] [CrossRef]
  44. Jiao, L.; Zhou, Y.X.; Jiang, H.L. Metal-organic framework-based CoP/reduced graphene oxide: High-performance bifunctional electrocatalyst for overall water splitting. Chem. Sci. 2016, 7, 1690–1695. [Google Scholar] [CrossRef] [Green Version]
  45. Yu, X.Y.; Feng, Y.; Jeon, Y.; Guan, B.; Lou, X.W.; Paik, U. Formation of Ni-Co-MoS(2) Nanoboxes with Enhanced Electrocatalytic Activity for Hydrogen Evolution. Adv. Mater. 2016, 28, 9006–9011. [Google Scholar] [CrossRef]
  46. Gao, H.; Zang, J.; Wang, Y.; Zhou, S.; Tian, P.; Song, S.; Tian, X.; Li, W. One-step preparation of cobalt-doped NiS@MoS2 core-shell nanorods as bifunctional electrocatalyst for overall water splitting. Electrochim. Acta 2021, 377, 138051. [Google Scholar] [CrossRef]
  47. Ge, Y.C.; Chu, H.; Chen, J.Y.; Zhuang, P.Y.; Feng, Q.Y.; Smith, W.R.; Dong, P.; Ye, M.X.; Shen, J.F. Ultrathin MoS2 Nanosheets Decorated Hollow CoP Heterostructures for Enhanced Hydrogen Evolution Reaction. ACS Sustain. Chem. Eng. 2019, 7, 10105–10111. [Google Scholar] [CrossRef]
  48. Wang, Z.C.; Xu, W.J.; Chen, X.K.; Peng, Y.H.; Song, Y.Y.; Lv, C.X.; Liu, H.L.; Sun, J.W.; Yuan, D.; Li, X.Y.; et al. Defect-Rich Nitrogen Doped Co3O4/C Porous Nanocubes Enable High-Efficiency Bifunctional Oxygen Electrocatalysis. Adv. Funct. Mater. 2019, 29, 1902875. [Google Scholar] [CrossRef]
  49. Kuang, P.; Tong, T.; Fan, K.; Yu, J. In Situ Fabrication of Ni–Mo Bimetal Sulfide Hybrid as an Efficient Electrocatalyst for Hydrogen Evolution over a Wide pH Range. ACS Catal. 2017, 7, 6179–6187. [Google Scholar] [CrossRef]
  50. Tan, X.; Duan, Z.; Liu, H.; Wu, X.; Cho, Y.-R. Core-shell structured MoS2/Ni9S8 electrocatalysts for high performance hydrogen and oxygen evolution reactions. Mater. Res. Bull. 2022, 146, 111626. [Google Scholar] [CrossRef]
  51. Liu, J.; Wang, Z.; Li, J.; Cao, L.; Lu, Z.; Zhu, D. Structure Engineering of MoS2 via Simultaneous Oxygen and Phosphorus Incorporation for Improved Hydrogen Evolution. Small 2020, 16, e1905738. [Google Scholar] [CrossRef]
  52. Yan, Y.; Li, A.; Lu, C.; Zhai, T.; Lu, S.; Li, W.; Zhou, W. Double-layered yolk-shell microspheres with NiCo2S4-Ni9S8-C hetero-interfaces as advanced battery-type electrode for hybrid supercapacitors. Chem. Eng. J. 2020, 396, 125316. [Google Scholar] [CrossRef]
  53. Liu, F.; Guo, X.; Hou, Y.; Wang, F.; Zou, C.; Yang, H. Hydrothermal combined with electrodeposition construction of a stable Co9S8/Ni3S2@NiFe-LDH heterostructure electrocatalyst for overall water splitting. Sustain. Energy Fuels 2021, 5, 1429–1438. [Google Scholar] [CrossRef]
  54. Khani, H.; Wipf, D.O. Iron Oxide Nanosheets and Pulse-Electrodeposited Ni-Co-S Nanoflake Arrays for High-Performance Charge Storage. ACS Appl. Mater. Interfaces 2017, 9, 6967–6978. [Google Scholar] [CrossRef]
  55. Yang, Y.J.; Chen, S.; Jiang, C.; Wang, N.; Liu, M.; Yang, P.; Cheng, Y. Assembly of Co3S4 nanoporous structure on Ni foam for binder-free high-performance supercapacitor electrode. J. Solid State Electrochem. 2023, 27, 1107–1118. [Google Scholar] [CrossRef]
  56. Yuan, F.; Liu, Z.; Qin, G.; Ni, Y. Fe-Doped Co-Mo-S microtube: A highly efficient bifunctional electrocatalyst for overall water splitting in alkaline solution. Dalton Trans. 2020, 49, 15009–15022. [Google Scholar] [CrossRef]
  57. Tang, S.; Wang, X.; Zhang, Y.; Courte, M.; Fan, H.J.; Fichou, D. Combining Co3S4 and Ni:Co3S4 nanowires as efficient catalysts for overall water splitting: An experimental and theoretical study. Nanoscale 2019, 11, 2202–2210. [Google Scholar] [CrossRef]
  58. Dai, Z.; Geng, H.; Wang, J.; Luo, Y.; Li, B.; Zong, Y.; Yang, J.; Guo, Y.; Zheng, Y.; Wang, X.; et al. Hexagonal-Phase Cobalt Monophosphosulfide for Highly Efficient Overall Water Splitting. ACS Nano 2017, 11, 11031–11040. [Google Scholar] [CrossRef]
  59. Yang, Y.; Zhang, W.; Xiao, Y.; Shi, Z.; Cao, X.; Tang, Y.; Gao, Q. CoNiSe2 heteronanorods decorated with layered-double-hydroxides for efficient hydrogen evolution. Appl. Catal. B Environ. 2019, 242, 132–139. [Google Scholar] [CrossRef]
  60. Wu, C.; Du, Y.; Fu, Y.; Feng, D.; Li, H.; Xiao, Z.; Liu, Y.; Yang, Y.; Wang, L. Mo, Co co-doped NiS bulks supported on Ni foam as an efficient electrocatalyst for overall water splitting in alkaline media. Sustain. Energy Fuels 2020, 4, 1654–1664. [Google Scholar] [CrossRef]
  61. Narasimman, R.; Waldiya, M.K.J.; Vemuri, S.K.; Mukhopadhyay, I.; Ray, A. Self-standing, hybrid three-dimensional-porous MoS2/Ni3S2 foam electrocatalyst for hydrogen evolution reaction in alkaline medium. Int. J. Hydrogen Energy 2021, 46, 7759–7771. [Google Scholar] [CrossRef]
  62. Zhai, Z.; Li, C.; Zhang, L.; Wu, H.-C.; Zhang, L.; Tang, N.; Wang, W.; Gong, J. Dimensional construction and morphological tuning of heterogeneous MoS2/NiS electrocatalysts for efficient overall water splitting. J. Mater. Chem. A 2018, 6, 9833–9838. [Google Scholar] [CrossRef]
Figure 1. Schematic of the synthesis process of the amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst.
Figure 1. Schematic of the synthesis process of the amethyst-like MoS2@Ni9S8/Co3S4 rod electrocatalyst.
Nanomaterials 13 02302 g001
Figure 2. SEM images of (ac) MoS2@Ni9S8/Co3S4/NF, (d) MoS2/Ni9S8/NF, (e) ZIF-67, (f) Mo-ZIF-67/NF, and (g) S-ZIF-67/NF. (h) EDS elemental mapping, (i) TEM, and (j) HRTEM images of MoS2@Ni9S8/Co3S4/NF.
Figure 2. SEM images of (ac) MoS2@Ni9S8/Co3S4/NF, (d) MoS2/Ni9S8/NF, (e) ZIF-67, (f) Mo-ZIF-67/NF, and (g) S-ZIF-67/NF. (h) EDS elemental mapping, (i) TEM, and (j) HRTEM images of MoS2@Ni9S8/Co3S4/NF.
Nanomaterials 13 02302 g002
Figure 3. XRD patterns of the as-prepared target and comparison samples.
Figure 3. XRD patterns of the as-prepared target and comparison samples.
Nanomaterials 13 02302 g003
Figure 4. (a) XPS full spectrum of the MoS2@Ni9S8/Co3S4/NF sample and (b) Mo 3d, (c) Ni 2p, and (d) S 2p XPS spectra of MoS2@Ni9S8/Co3S4/NF and MoS2/Ni9S8/NF.
Figure 4. (a) XPS full spectrum of the MoS2@Ni9S8/Co3S4/NF sample and (b) Mo 3d, (c) Ni 2p, and (d) S 2p XPS spectra of MoS2@Ni9S8/Co3S4/NF and MoS2/Ni9S8/NF.
Nanomaterials 13 02302 g004
Figure 5. HER performances of the as-prepared samples in 1 M KOH: (a) LSV curves, (b) overpotential at different current densities, (c) Tafel plots, and (d) Nyquist plots for the different samples.
Figure 5. HER performances of the as-prepared samples in 1 M KOH: (a) LSV curves, (b) overpotential at different current densities, (c) Tafel plots, and (d) Nyquist plots for the different samples.
Nanomaterials 13 02302 g005
Figure 6. OER performances of the as-synthesized samples in 1 M KOH: (a) LSV curves, (b) overpotential at different current densities, (c) Tafel plots, and (d) Nyquist plots for the different samples.
Figure 6. OER performances of the as-synthesized samples in 1 M KOH: (a) LSV curves, (b) overpotential at different current densities, (c) Tafel plots, and (d) Nyquist plots for the different samples.
Nanomaterials 13 02302 g006
Figure 7. (a) CV curves of MoS2@Ni9S8/Co3S4/NF at different scan rates and (b) electrochemical double layer capacitance of the different as-synthesized samples. Performance of MoS2@Ni9S8/Co3S4/NF during the stability test: (c) HER-CP test at 10, 50, and 100 mA cm–2; (d) OER-CP test at 10 and 50 mA cm–2.
Figure 7. (a) CV curves of MoS2@Ni9S8/Co3S4/NF at different scan rates and (b) electrochemical double layer capacitance of the different as-synthesized samples. Performance of MoS2@Ni9S8/Co3S4/NF during the stability test: (c) HER-CP test at 10, 50, and 100 mA cm–2; (d) OER-CP test at 10 and 50 mA cm–2.
Nanomaterials 13 02302 g007
Figure 8. (a) LSV curve of MoS2@Ni9S8/Co3S4/NF in a two-electrode system and (b) CP test of MoS2@Ni9S8/Co3S4/NF at 10 mA cm–2.
Figure 8. (a) LSV curve of MoS2@Ni9S8/Co3S4/NF in a two-electrode system and (b) CP test of MoS2@Ni9S8/Co3S4/NF at 10 mA cm–2.
Nanomaterials 13 02302 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

Pei, Z.; Qin, T.; Tian, R.; Ou, Y.; Guo, X. Construction of an Amethyst-like MoS2@Ni9S8/Co3S4 Rod Electrocatalyst for Overall Water Splitting. Nanomaterials 2023, 13, 2302. https://doi.org/10.3390/nano13162302

AMA Style

Pei Z, Qin T, Tian R, Ou Y, Guo X. Construction of an Amethyst-like MoS2@Ni9S8/Co3S4 Rod Electrocatalyst for Overall Water Splitting. Nanomaterials. 2023; 13(16):2302. https://doi.org/10.3390/nano13162302

Chicago/Turabian Style

Pei, Zhen, Tengteng Qin, Rui Tian, Yangxin Ou, and Xingzhong Guo. 2023. "Construction of an Amethyst-like MoS2@Ni9S8/Co3S4 Rod Electrocatalyst for Overall Water Splitting" Nanomaterials 13, no. 16: 2302. https://doi.org/10.3390/nano13162302

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