RESEARCH ARTICLE

Oxygen-deficient MoOx/Ni3S2 heterostructure grown on nickel foam as efficient and durable self-supported electrocatalysts for hydrogen evolution reaction

  • Zihuan Yu 1 ,
  • Haiqing Yan 1 ,
  • Chaonan Wang 1 ,
  • Zheng Wang 1 ,
  • Huiqin Yao , 2 ,
  • Rong Liu , 3 ,
  • Cheng Li , 4 ,
  • Shulan Ma , 1
Expand
  • 1. Beijing Key Laboratory of Energy Conversion and Storage Materials, College of Chemistry, Beijing Normal University, Beijing 100875, China
  • 2. School of Basic Medical Sciences, Ningxia Medical University, Yinchuan 750004, China
  • 3. Analytical and Testing Center, Beijing Normal University, Beijing 100875, China
  • 4. Shanghai Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China
huiqin_yao@163.com
liur@bnu.edu.cn
licheng@sinap.ac.cn
mashulan@bnu.edu.cn

Received date: 06 May 2022

Accepted date: 27 Jul 2022

Copyright

2022 Higher Education Press

Abstract

High-performance and ultra-durable electrocatalysts are vital for hydrogen evolution reaction (HER) during water splitting. Herein, by one-pot solvothermal method, MoOx/Ni3S2 spheres comprising Ni3S2 nanoparticles inside and oxygen-deficient amorphous MoOx outside in situ grow on Ni foam (NF), to assembly the heterostructure composites of MoOx/Ni3S2/NF. By adjusting volume ratio of the solvents of ethanol to water, the optimized MoOx/Ni3S2/NF-11 exhibits the best HER performance, requiring an extremely low overpotential of 76 mV to achieve the current density of 10 mA∙cm‒2 (η10 = 76 mV) and an ultra-small Tafel slope of 46 mV∙dec‒1 in 0.5 mol∙L‒1 H2SO4. More importantly, the catalyst shows prominent high catalytic stability for HER (> 100 h). The acid-resistant MoOx wraps the inside Ni3S2/NF to ensure the high stability of the catalyst under acidic conditions. Density functional theory calculations confirm that the existing oxygen vacancy and MoOx/Ni3S2 heterostructure are both beneficial to the reduced Gibbs free energy of hydrogen adsorption (|∆GH*|) over Mo sites, which act as main active sites. The heterostructure effectively decreases the formation energy of O vacancy, leading to surface reconstruction of the catalyst, further improving HER performance. The MoOx/Ni3S2/NF is promising to serve as a highly effective and durable electrocatalyst toward HER.

Cite this article

Zihuan Yu , Haiqing Yan , Chaonan Wang , Zheng Wang , Huiqin Yao , Rong Liu , Cheng Li , Shulan Ma . Oxygen-deficient MoOx/Ni3S2 heterostructure grown on nickel foam as efficient and durable self-supported electrocatalysts for hydrogen evolution reaction[J]. Frontiers of Chemical Science and Engineering, 2023 , 17(4) : 437 -448 . DOI: 10.1007/s11705-022-2228-1

1 Introduction

With the fast growth of energy demands and deterioration of environment, it is necessary to develop renewable and clean energy [1]. Hydrogen (H2), thanks to its environmental friendliness and high energy density, has attracted great attention [2]. Electrocatalytic water splitting has be considered as the most eco-friendly and economical route to produce H2, because water is an abundant and renewable source [3]. However, the hydrogen evolution reaction (HER) during water electrolysis suffers from large overpotential and slow kinetics [4]. To speed up the HER kinetics, highly active and durable electrocatalysts are applied to lower the dynamic overpotentials [5]. Pt-based compounds have been considered as the most efficient electrocatalysts toward HER, but the preciousness and scarcity of Pt seriously restrict its practical utilization in water electrolysis [6]. Thus, it is vital to explore low-cost while high-efficiency electrocatalysts to replace Pt-based materials.
Nonprecious metal materials, such as Mo-based sulfides [7], oxides [8], phosphides [9], and carbides [10], have been investigated as potential HER catalysts. Compared with molybdenum sulfides, molybdenum oxides are highly stable and readily available in large scale and can also show acceptable HER property [11]. MoO3 (α-MoO3), as a low-price and nontoxic material, has been reported to be widely applied in HER, exhibiting an overpotential of 112 mV achieving the current density of 10 mA∙cm‒2 (η10 = 112 mV) in acid media [12]. However, because of the less active sites and poor conductivity of intrinsic MoO3, its catalytic performance has been reported to be much lower than that of Pt-based materials [13]. If the MoO3 is combined with other substances to fabricate certain active interfaces, the active sites of MoO3 may be exposed as much as possible [14]. Zhang’s group [15] coupled MoO3 with 1T'-MoS2 to assembly 1T'-MoS2/MoO3 heterostructure nanosheets, in which the MoS2/MoO3 interfaces can facilitate electron transfer and surface hydrogen generation, leading to the improvement of electrocatalytic performance. Besides, loading P element onto MoO3 nanosheets can also improve the HER catalytic performance [16], for which the P component can boost the adsorption/desorption of proton and afterwards increase electrocatalytic performance. A composite electrocatalyst comprising RuO2 supported on MoO3 nanosheets was investigated as an HER electrocatalyst, in which the synergetic effect created through interaction between MoO3 and RuO2 led to the enhancement of catalytic activity [17]. Nevertheless, their HER performance is still inferior to that of precious metals. In addition, the oxygen vacancy in MoO3 may not only produce greater conductivity, but also act as HER active site [13]. For example, MoO3‒x with oxygen vacancies depicted higher HER catalytic activity than com-MoO3 [13]. Density functional theory (DFT) calculation revealed that the oxygen vacancy prominently reduces the adsorption energy of H2O and then improves the catalytic activity. Thus, fabricating oxygen vacancies and active interfaces into molybdenum oxides may be effective ways to enhance the HER performance of molybdenum oxides. The known researches on molybdenum oxides with oxygen vacancy are all about crystalline MoO3, but there are lack of reports focused on amorphous MoOx with oxygen vacancy.
Meanwhile, Ni-based materials have shown a big potential for using as HER catalysts [18]. Ni3S2, because of the existence of Ni–Ni networks throughout its crystal structure, exhibits good metallic conductivity [19]. So far, many studies revealed that Ni3S2 owns HER activity [20], while its catalytic activity and stability are still less competitive relative to noble-metal catalysts. On the other hand, self-supported catalysts, compared to powdery ones, avoid the addition of polymer binder. Meanwhile, the introduction of substrates in self-supporting catalysts can raise electrochemical active areas. The three-dimensional nickel foam (NF) is a low-cost template with large surface area, and is normally used as a conductive substrate to host electrocatalyst materials [21]. Using thiourea, Ni(NO3)2·6H2O and NF as reactants, Ni3S2/NF was fabricated to depict high HER activity and good durability in basic and neutral conditions [22]. Hybridizing Ni3S2 with other electrocatalysts such as MoS2 [23] and/or Co9S8/MoS2 [24] is also a commonly employed method to augment the catalytic activity. Moreover, Ni3S2 is generally combined with other sulfides, while the integration of Ni3S2 with amorphous MoOx is rarely reported. As known, Ni3S2 and NF matrix are unstable in acidic solutions, while MoO3 or MoOx is acid-resistant. Therefore, if NF and the in situ grown Ni3S2 are wrapped by dense layers of MoOx, it is expected to improve the acid resistance of the integrated elelctrocatalysts.
Herein, via a facile solvothermal method, we fabricate MoOx/Ni3S2 spheres which consist of dispersed Ni3S2 nanoparticles and amorphous MoOx wrapped outside. By adjusting ratios of water/ethanol solvents, the as-prepared MoOx/Ni3S2/NF-11 (volume ratio of ethanol:water is 1:1) exhibits optimum HER performance, requiring an extremely low overpotential (η10 = 76 mV) and a much low Tafel slope of 46 mV·dec‒1 in 0.5 mol·L‒1 H2SO4. Benefiting from the existing oxygen vacancy and formed heterojunction, Mo atoms are exposed as much as possible to exhibit reduced Gibbs free energy of hydrogen adsorption (∆GH*) and act as active-sites. Additionally, the stability of the catalyst under acidic conditions is greatly improved because the acid-resistant amorphous MoOx layer is wrapped around Ni3S2/NF. The MoOx/Ni3S2/NF-11 displays extremely high stability (≥ 100 h at 23.5 mA∙cm‒2). This study would give a novel perspective on the fabrication of amorphous molybdenum oxides or oxygen-deficient materials with excellent HER performance.

2 Experimental

2.1 Synthesis of MoOx/Ni3S2/NF composites

A series of MoOx/Ni3S2/NF composites (labeled as MoOx/Ni3S2/NF-mn, where m and n are volume ratios of ethanol:water) were synthesized via a one-pot solvothermal reaction. Typically, two pieces of commercial NF (1 cm × 2 cm) were first immersed in an acid solution (1 mol·L−1 HCl) for 20 min, to remove the oxides on the surface of NF. Then the NF was washed using acetone, water and ethanol in turn, and then was dried in vacuum at 45 °C, followed by weighing. Meanwhile, 2 mmol of Na2MoO4·2H2O was dissolved in 30 mL of ethanol:water mixture solvents. After stirred for 30 min, 10 mmol of thiourea was dissolved in the solutions. The abovementioned solutions and pre-treated NF were simultaneously transferred to a Teflon-lined autoclave (50 mL) and reacted at 140 °C for 18 h. After cooled, the obtained MoOx/Ni3S2/NF-mn products were washed with excess deionized water and ethanol several times, and then dried in vacuum at 45 °C overnight.

2.2 Electrochemical measurements

The electrochemical tests were conducted in a three-electrode cell in a 0.5 mol·L−1 H2SO4 solution. Ag/AgCl (saturated KCl) electrode and a graphite rod electrode worked as reference and counter electrodes, respectively. The as-prepared samples were used directly as working electrodes. For a reliable comparison, we loaded Pt−C and commercial MoO3 (com-MoO3) on NF with the same mass loading as MoOx/Ni3S2/NF-11. Table S1 (cf. Electronic Supplementary Material, ESM) listed the mass loading of all catalysts. The details of electrochemical measurements are shown in ESM.

3 Results and discussion

3.1 Material characterization

The fabrication procedure of the MoOx/Ni3S2/NF composites is described in . In the solvothermal reaction, the thiourea may break down to release HS ions (Eq. (1)); meantime, the NF can release Ni2+ into the solution, and the Ni2+ would react with the active HS to generate the Ni3S2 particles (Eq. (2)) [24]:
Scheme1 Schematic illustration of formation of oxygen-deficient MoOx/Ni3S2/NF composites.

Full size|PPT slide

NH2CSNH2+3H2O2NH4++HS+HCO3
3Ni+2HS+2H2ONi3S2+2OH+2H2
The structures of the MoOx/Ni3S2/NF-mn were studied firstly by X-ray diffraction (XRD) measurements. As shown in Fig.1, for all samples, there appeared diffraction peaks at 21.7°, 31.1°, 37.8°, 49.7° and 55.2° respectively assigned to the (101), (110), (003), (113) and (122) planes of hexagonal Ni3S2 (JCPDS no. 44-1418) [25]. This indicates that the partial NF surface was converted to Ni3S2 under the solvothermal conditions. Three strong diffraction peaks of Ni were derived from the NF substrate. However, no distinguishable diffractions related to MoOx were found, suggesting that if there is certain MoOx phase, it would exist in amorphous form. The transmission electron microscope (TEM) characterization discussed below with the absence of lattice fringes would provide important evidence of the amorphous characteristic.
Fig.1 XRD patterns of (a) MoOx/Ni3S2/NF-01, (b) MoOx/Ni3S2/NF-11 and (c) MoOx/Ni3S2/NF-21.

Full size|PPT slide

The morphologies of NF (Fig. S1, cf. ESM) and MoOx/Ni3S2/NF-mn (Fig.2) were characterized by scanning electron microscopy (SEM). Figure S1(a) shows the 3D porous characteristic of NF, and at higher magnification, the surface of NF can be observed to be very smooth (Fig. S1(a′)). The special skeleton of NF provides large surface area and ideal mechanical stability. As seen in Fig.2, the entire NF surface is tightly covered by the MoOx/Ni3S2. When ethanol solvent was absent in the reaction (no ethanol), the resulting product of MoOx/Ni3S2/NF-01 (Fig.2(a) and Fig.2(a′)) has some cracks on its surface, which may reduce the conductivity of the catalyst [26]. At the ethanol:water ratio of 1:1, the as-formed MoOx/Ni3S2/NF-11 presents irregular small spheres grown on the substrate (Fig.2(b) and Fig.2(b′)). In addition, it can be seen from Fig.2(b) that there is no crack on the surface of the sample, which means that the acid-resistant NF matrix is completely coated by MoOx/Ni3S2, so that NF cannot be corroded by acid, paving the way for the excellent stability of the catalyst under acidic conditions. As the ethanol:water ratio is increased to 2:1 (with more ethanol), as found in MoOx/Ni3S2/NF-21, the formed MoOx/Ni3S2 spheres on the NF surface become larger (Fig.2(c) and Fig.2(c′)).
Fig.2 SEM images of (a, a′) MoOx/Ni3S2/NF-01, (b, b′) MoOx/Ni3S2/NF-11 and (c, c′) MoOx/Ni3S2/NF-21.

Full size|PPT slide

For the MoOx/Ni3S2/NF-11 with the most regular and perfect appearance, detailed morphology and structure were further explored by TEM (Fig.3 and S3, cf. ESM) via scratching the samples from the NF substrate. TEM images (Fig.3(a) and Fig.3(b)) show that the MoOx/Ni3S2 spheres are comprised of nanoparticles. Fig.3(c) reveals the interface region of the Ni3S2 nanoparticles, indicating that there are two clearly distinguishable phases between the inside and outside of the particle, and the Ni3S2 nanoparticle is closely connected to the amorphous MoOx. For clarity, the original image (without scaling) of Fig.3(c) is shown in Fig. S2 (cf. ESM). The TEM investigations of MoOx/Ni3S2-11 (see Fig.3(d) and Fig.3(e)) further exhibit the tight connection between the Ni3S2 particles and MoOx. The lattices of 0.24 (Fig.3(d)) and 0.29 nm (Fig.3(e)) are ascribed to (003) and (110) crystal planes of Ni3S2 [27], further proving its presence. The above results can be supported by the SAED pattern (Fig.3(f)), which displays characteristic crystal planes of Ni3S2 and diffuse rings of the amorphous MoOx. The amorphous characteristic of MoOx leads to the absence of XRD diffractions (see Fig.1). The MoOx surrounding for Ni3S2 further ensures the acid resistance of the electrocatalyst. The compositional distributions of the MoOx/Ni3S2-11 sphere are confirmed by elemental mapping analyses (Fig.3(g)), high angle annular dark field scanning transmission electron microscope (HAADF-STEM) image (Fig. S3(a)) and line scans (Fig. S3(b)). In Fig.3(g), uneven distribution of Mo/O and Ni/S is clearly observed throughout the whole sphere. The elements of Mo and O appear in the same region, while Ni and S elements locate in other regions, indicating that the two phases of MoOx and Ni2S3 have distinct interfaces. From Fig.3(g), the atom ratio of Mo:O is 1:2.94, which is close to MoO3, with the oxygen content slightly reduced. All these results prove the existence of Ni3S2 crystalline phase along with amorphous MoOx.
Fig.3 (a, b) TEM images of MoOx/Ni3S2-11, (c, d, e) HR-TEM images of MoOx/Ni3S2-11, and the visible lattice fringes images obtained from the blue and yellow square regions; (f) SAED pattern of MoOx/Ni3S2-11; (g) HAADF-STEM and energy dispersive spectroscopy elemental mappings of Mo, O and Ni of MoOx/Ni3S2-11; (h) electron paramagnetic resonance (EPR) spectrum of MoOx/Ni3S2/NF-11 before and after 60 h It test, and the control sample of com-MoO3.

Full size|PPT slide

To clarify the detailed structure of MoOx/Ni3S2/NF-11, EPR spectroscopy and X-ray photoelectron spectroscopy (XPS) were carried out and the results were shown in Fig.4, with com-MoO3 as a control sample. In Fig.3(h), the EPR spectra of com-MoO3 (black line) shows no signal, while MoOx/Ni3S2/NF-11 (red line) displays a signal at about 3512 Guass (g = 2.0036), indicating the existence of oxygen vacancy. For XPS spectra of MoOx/Ni3S2/NF-11 (Fig.4(a)), the binding energies of 235.6 and 232.4 eV are related to Mo 3d3/2 and 3d5/2 of Mo6+, the same as found in com-MoO3 [28]. In contrast to com-MoO3, MoOx/Ni3S2/NF-11 shows a pair of peaks of Mo5+ (234.0 eV for Mo 3d3/2 and 230.8 eV for Mo 3d5/2) [29]. Comparison of O 1s spectra of MoOx/Ni3S2/NF-11 and com-MoO3 is presented in Fig.4(b). As shown, for com-MoO3, the binding energy of 531.9 eV is assigned to lattice oxygen (O2‒) [30], while in MoOx/Ni3S2/NF-11, the energy of O2‒ is reduced to a lower level (531.5 eV). Also, the half-peak width of O 1s core-level spectra of the MoOx/Ni3S2/NF-11 is widened in comparison with that of the com-MoO3. The difference of half-peak width of O 1s core-level spectra in the materials with and without oxygen vacancies has also been observed in other literature [31]. For MoOx/Ni3S2/NF-11, the appearance of Mo5+ and the shift of O 1s simultaneously prove the presence of oxygen vacancy, which is consistent with EPR results. All these indicate a changed coordination configuration between Mo and O, as discussed in literatures [13,16]. It was reported that the shift of O 1s to a lower energy level means the electron transfer to the neighboring oxygen vacancies [32]. Meanwhile, one weak peak detected at 533.4 eV can be assigned to O 1s of surface adsorbed species (here is OH) [33]. In Fig.4(c), for Ni 2p spectra of MoOx/Ni3S2/NF-11, the peaks at 873.8 and 856.0 eV are related to Ni 2p1/2 and Ni 2p3/2 of Ni2+, accompanied by two satellite peaks at 879.8 and 861.7 eV [34]. The peak at 853.4 eV is assigned to Ni0, which belongs to Ni3S2 or NF [35]. In the spectra of S 2p (Fig.4(d)), the two signals at 162.9 and 161.7 eV belong to 2p1/2 and 2p3/2 of S2‒, and the other two signals at 164.4 and 163.2 eV are attributed to 2p1/2 and 2p3/2 of S22‒ [36], suggesting the presence of terminal unsaturated S of Ni–S bonds [37]. The peak at 168.2 eV is assigned to oxidized sulfur species (here is SO42‒) owing to surface oxidation [38]. In Raman spectra of MoOx/Ni3S2/NF-11 (Fig. S4, cf. ESM), the band at 324 cm‒1 is associated with A1 vibration mode of the Ni3S2 phase [39], and the peaks ranging from 800 to 1000 cm–1 are related to Mo=O modes [16]. We also measured the infrared spectrum of MoOx/Ni3S2/NF-11 (Fig. S5, cf. ESM). The bands at 975 and 914 cm−1 correspond to the stretching vibration of Mo=O, and the bands at 615 and 514 cm−1 are assigned to stretching and bending vibrations of Mo–O, respectively [40]. No Mo–S related spectral peaks can be observed, which proves that the sample does not contain Mo–S bond.
Fig.4 X-ray photoelectron spectra with deconvolution of (a) Mo 3d, (b) O 1s for MoOx/Ni3S2/NF-11 and com-MoO3, (c) Ni 2p, and (d) S 2p for MoOx/Ni3S2/NF-11.

Full size|PPT slide

3.2 Electrocatalytic performance for HER

HER performance of MoOx/Ni3S2/NF-mn synthesized at different ethanol:H2O ratios and control samples of com-MoO3/NF, Ni3S2/NF and Pt–C/NF was measured (Fig.5 and Fig. S6, cf. ESM). As observed (Fig. S6), among the three samples, the MoOx/Ni3S2/NF-11 at an ethanol:H2O ratio of 1:1 exhibits an extremely low η10 value of 76 mV, demonstrating HER performance much better than MoOx/Ni3S2/NF-01 (η10 = 127 mV) without ethanol added during synthesis and MoOx/Ni3S2/NF-21 (η10 = 114 mV) prepared at an ethanol:H2O ratio of 2:1. From the SEM images mentioned above, the MoOx/Ni3S2/NF-11 has a relatively small particle size and no crack on the surface, which is conducive to increasing the contact area between the catalyst and the electrolyte and accelerating the electron transmission. Therefore, the MoOx/Ni3S2/NF-11 exhibits the optimal HER performance. Furthermore, the η10 of 76 mV of MoOx/Ni3S2/NF-11 is extremely lower than that (235 mV) of com-MoO3/NF, and is close to that (37 mV) of Pt–C/NF, showing performance exceeding com-MoO3 and comparable to Pt–C. As listed in Tab.1 [16,17,4146], the η10 of MoOx/Ni3S2/NF-11 is obviously lower compared with the reported Mo-related electrocatalysts. The excellent catalytic performance of MoOx/Ni3S2/NF-11 may be mainly attributed to the oxygen vacancy. For the P-MoO3 [16], though it also contains oxygen vacancies coming from P doping, it has a larger η10 of 166 mV, meaning an inferior HER activity than MoOx/Ni3S2/NF-11. The outstanding HER activity of MoOx/Ni3S2/NF-11 suggests that there may be heterojunctions formed in the two integrated phases of MoO3 and Ni3S2. The heterojunction and oxygen vacancy would offer rapid electron transfer favorable to the HER activity.
Tab.1 HER performance of MoOx/Ni3S2/NF-11 and some reported electrocatalysts in 0.5 mol·L−1 H2SO4
Catalystsη10/mVTafel slope/(mV∙dec‒1)Ref.
MoOx/Ni3S2/NF-117646This work
P-MoO3‒xa)16642[11]
MoO3@RuO2b)11062[17]
Pd NDs/DR MoS2c)10341[41]
UDSL-MoS2-rGOd)ca. 21035[42]
Mo/Mo2Ce)8962[43]
(1T/2H) MoS2/α-MoO3f)23281[44]
MoP/NGg)9450[45]
MoCxh)14253[46]

a) P doped MoO3−x, prepared via two-step intercalation method by using dodecylamine (DDA) and 4-bromine benzyl phosphoric acid; b) RuO2 nanoparticles supported on MoO3 nanosheets, prepared by a sonochemical method followed by calcination in air; c) Pd nanodisks (NDs) assembled on the basal plane of defect-rich MoS2 nanosheets (DR-MoS2); d) ultradispersed and single-layered MoS2 nanoflakes coupled with reduced graphene oxide sheets (UDSL-MoS2-rGO), synthesized by a hydrothermal method using (NH4)6Mo7O24·4H2O, L-cysteine and GO under pH = 1; e) Mo/Mo2C heteronanosheets, obtained by a NaCl template method followed by the reduction and carbonization under H2 and CH4, respectively; f) (1T/2H) MoS2/α-MoO3 prepared by hydrothermal method using thiourea, MoO3 and N2H4·H2O; g) MoP nanoflakes intercalated nitrogen-doped graphene nanobelts (MoP/NG), synthesized by inserting DDA into MoO3 nanobelts followed by phosphorization; h) MoCx obtained by MOFs-assisted strategy followed by annealing under N2 flow.

Fig.5 (a) Polarization curves of MoOx/Ni3S2/NF-11, com-MoO3/NF, Ni3S2/NF and Pt–C/NF; (b) Tafel slopes of MoOx/Ni3S2/NF-11, com-MoO3/NF and Pt–C/NF derived from polarization curves; (c) Nyquist plots of MoOx/Ni3S2/NF-11, com-MoO3/NF and Pt–C/NF at ‒200 mV versus RHE measured from electrochemical impedance spectroscopy (EIS) in the frequency range from 105 to 0.01 Hz; (d) plots of current density as a function of scan rates for MoOx/Ni3S2/NF-11, com-MoO3/NF and Pt–C/NF; (e) chronoamperometric curve of MoOx/Ni3S2/NF-11 at a constant applied potential of ‒180 mV versus RHE; (f) polarization curves of before and after 100 h I–t test of MoOx/Ni3S2/NF-11. All the measurements were performed in a 0.5 mol·L−1 H2SO4 electrolyte.

Full size|PPT slide

To explore the catalytic mechanism of the electrocatalysts, Tafel plots were calculated (Fig.5(b)). The Tafel slope of MoOx/Ni3S2/NF-11 is 46 mV∙dec‒1, which is close to the value (30 mV∙dec‒1) of Pt–C/NF and much lower than that (115 mV∙dec‒1) of com-MoO3/NF. Tab.1 listed the Tafel slopes of reported Mo-based HER catalysts. The low Tafel slope of MoOx/Ni3S2/NF-11 demonstrates its effectively enhanced kinetics during the H2O dissociation [17]. On the basis of kinetic models, the Tafel slopes of 120, 40 and 30 mV∙dec‒1 are assigned to Volmer (Eq. (3)), Heyrovsky (Eq. (4)) and Tafel (Eq. (5)) reactions, respectively:
H3O++*+eHads*+H2O
Hads*+H3O++e*+H2+H2O
Hads*+Hads*2*+H2
The first step is Volmer reaction (Eq. (3)), for which one proton (H3O+) first adsorbs on the surface of the catalyst and accepts one electron to generate one hydrogen atom (Hads*). The second step involves two kinds of reactions: one is Heyrovsky reaction and the other is Tafel reaction. For the Heyrovsky reaction (Eq. (4)), which is an electrochemical desorption, another proton (H3O+) accepts one electron and desorbs from catalyst surface and react with the former H atom to generate one H2. For the Tafel step (Eq. (5)), two Hads desorb from catalyst surface and combine directly and then release one H2. The slope of 46 mV∙dec‒1 for MoOx/Ni3S2/NF-11 indicates it follows the Volmer–Heyrovsky mechanism, while Pt–C/NF with the much lower slope of 30 mV∙dec‒1 follows the Volmer–Tafel mechanism. Nyquist plots were obtained by measuring EIS and the charge-transfer resistance (Rct) values were fitted by the equivalent circuit model (Fig.5(c)). It can be seen that the MoOx/Ni3S2/NF-11 shows an extremely small Rct of 6.51 Ω, which is much lower than the com-MoO3/NF (21.51 Ω), indicating that the MoOx/Ni3S2/NF-11 has faster charge transfer dynamics in HER process (see Table S2, cf. ESM). The fast charge transfer is probably attributed to synergetic effect of the amorphous MoOx and the Ni3S2 nanoparticles. The abundant interfaces of the two phases make the hydrogen adsorption and desorption more effective [17].
In general, good electrocatalytic activity is accompanied by a large electrochemically active surface area, which is normally proportional to the electrochemical double-layer capacitance (Cdl). From the cyclic voltammetry curves carried out at varied scan rates (Fig. S7, cf. ESM), the Cdl of MoOx/Ni3S2/NF-11 is calculated as 42 mF∙cm‒2 (Fig.5(d)). This Cdl value is much larger than those many reported catalysts, such as MoO3 (13.58 mF∙cm‒2) [14], UDSL-MoS2-rGO (24 mF∙cm‒2) [42], MoP/NG (9.1 mF∙cm‒2) [45] and MoO2/MoSe2 (18.68 µF∙cm‒2) [47]. This large Cdl unveils the large amount of electrochemical active sites in the MoOx/Ni3S2/NF-11. These results reveal that the MoOx/Ni3S2/NF-11 with oxygen vacancies generates more accessible active sites, thereby promoting the H2 evolution process.

3.3 DFT calculations

DFT calculations (Fig.6) were employed to well understand the influence of oxygen vacancy and Ni3S2 component on the catalytic activity. According to TEM characterization, MoOx was found to be wrapped outside the Ni3S2. Therefore, when building the optimized models, the possibility of H* adsorption on MoO3 is given priority, and Ni3S2 is regarded as the “substrate” of MoO3. As shown in Fig.6(A), we calculate the ΔGH* of MoO3 (on O site), oxygen-deficient MoO3 (noted as Ov-MoO3, on Mo site), MoO3/Ni3S2 (on O site), and Ov-MoO3/Ni3S2 (on O and Mo sites). The MoO3 (Fig.6(A-a)) shows a ΔGH* of ‒2.37 eV, for which the negative value indicates easy H adsorption (Volmer reaction, Eq. (3) discussed above) but difficult H desorption (Heyrovsky reaction, Eq. (4) discussed above). Because the surface of MoO3 is all O atoms and the O–H* bond is strong, the H* is difficult to be desorbed. At this time, desorption step becomes the rate-determining step. When oxygen vacancy is present, as in Ov-MoO3 (Fig.6(A-b)), Mo sites are exposed to exhibit a smaller |∆GH*| of 2.30 eV. When MoO3 and Ni3S2 form the heterojunction (Fig.6(A-c), the MoO3/Ni3S2 has a much smaller |∆GH*| of 1.82 eV, which means weakened binding between O and H, thus facilitating the H desorption. While for the heterojunction of Ov-MoO3/Ni3S2 containing oxygen vacancy, |∆GH*| on the O site adjacent to oxygen vacancy is slightly increased to 1.98 eV (Fig.6(A-d)), indicating strengthened O–H bonding (difficult H desorption). More importantly and interestingly, the Mo site of Ov-MoO3/Ni3S2 has a much lower |∆GH*| of 1.54 eV (Fig.6(A-e)), which is the most close to the value of 0. As known an electrocatalyst with |∆GH*| approaching to zero would have high HER activity. So the exposed Mo due to oxygen defect serves as active sites. By comparing (a) and (c), or (b) and (d) in Fig.6(A), it can be seen that the heterojunction greatly improves the catalytic activity of both O and Mo sites, while comparison of (c), (d) and (e) in Fig.6(A) shows that although oxygen deficiency slightly reduces the activity of O sites, the exposed Mo sites result in much better catalytic activity.
Fig.6 (A) Optimized structures of (a) MoO3, (b) Ov-MoO3, (c) MoO3/Ni3S2, (d) Ov-MoO3/Ni3S2 with H* on O site and (e) Ov-MoO3/Ni3S2 with H* on Mo site and the corresponding H adsorption free energy (ΔGH*) at a potential U = 0 V relative to standard hydrogen electrode at pH = 0; (B) differential charge density distribution of Ov-MoO3/Ni3S2; (C) vacancy formation energy of Ov-MoO3/Ni3S2 and Ov-MoO3 and corresponding models. Symbols for atoms: Mo is cyan in A and C while purple in B, S is yellow, Ni is blue in A and C while grey in B, O is red and H* is white. The dotted box shows the position of oxygen vacancy.

Full size|PPT slide

Differential charge density distribution of the Ov-MoO3/Ni3S2 has been calculated, with results shown in Fig.6(B). It is found that there is a strong charge transfer between Ni3S2 and MoO3. Compared with MoO3/Ni3S2 (without oxygen vacancy, Fig. S8, cf. ESM), the O atoms in Ov-MoO3/Ni3S2 (Fig.6(B)) gain more electrons. The increase of surface charge of oxygen will strengthen the binding of O–H and make the desorption more difficult. This is the reason why the |∆GH*| on O site of Ov-MoO3/Ni3S2 is larger than that of MoO3/Ni3S2. In addition, Fig.6(B) depicts electron deficiency of exposed Mo site in Ov-MoO3/Ni3S2. During the Heyrovsky reaction, Mo–H* bond needs to be dissociated prior to H2 release. In this process, electrophilic Mo will pull the electrons between Mo and H, which will weaken the Mo–H* bond thus promoting hydrogen evolution via the cleavage of Mo–H* [48].
Long-term durability is another crucial requirement for the catalysts in real applications. To measure the durability of MoOx/Ni3S2/NF-11, the time-dependent current density was measured at a constant potential of ‒0.18 V for 100 h. The excellent stability is due to that the acid-resistant amorphous MoOx encases the Ni3S2 and NF substrate which are unstable under acidic conditions. As seen in Fig.5(e), the MoOx/Ni3S2/NF-11 was continuously operated for 100 h with no significant degradation observed, confirming the ultra-durable stability. Fig.5(f) shows the linear sweep voltammetry before and after 100 h I–t HER test. As observed, the catalytic activity of the MoOx/Ni3S2/NF-11 was even increased after the stability test. This may be thankful to the surface reconfiguration of the catalyst [49]. In order to deeply learn about the reaction mechanism in the HER process, we have carried out SEM and XPS for MoOx/Ni3S2/NF-11 after stability test. From the SEM images shown in Fig. S9 (cf. ESM), we see after the HER test, the microstructure of MoOx/Ni3S2/NF-11 has not changed significantly, and many irregular small spheres are still attached to the surface of NF. From XPS (Fig.7), after long-term stability test (100 h), the valence states of Ni and S do not change [50]. For Mo, the mixed valence state of Mo5+ and Mo6+ is still reflected after the stability test. The peaks at 235.6 and 232.4 eV correspond to Mo6+ [28], and the binding energies of 234.0 and 230.8 eV are related to Mo5+ [29]. At the same time, it can be seen from the integral area of the curve that the content of Mo5+ increases compared with that before. In addition, the O 1s peaks shift towards lower binding energy by 0.2 eV after stability test. The increase of Mo5+ content and the shift of O 1s with lower binding energy also prove that the oxygen vacancy content increased after long-term stability. The in situ surface reconfiguration may result in more oxygen vacancies on catalyst surface, which exposes more Mo active sites. EPR spectroscopy (Fig.3(h)) shows that the MoOx/Ni3S2/NF-11 after 60 h It test displays a much stronger signal, proving increased content of oxygen vacancy in the catalyst after HER testing, which is consistent with the XPS characterization. To further confirm the in situ surface reconfiguration of the catalyst, we calculated the formation energy of oxygen vacancy in MoO3 and MoO3/Ni3S2. As shown in Fig.6(C), when MoO3 and Ni3S2 form heterojunction, the formation energy of oxygen vacancy is significantly reduced, and thus decreases the energy barrier for structural reconfiguration. This will enable MoO3 to have more oxygen vacancies, thus exposing more active sites of Mo. The low |∆GH*| of exposed Mo sites is beneficial to the improvement of the HER activity. With the continuation of the electrolysis reactions, the performance of the catalyst will gradually improve. From the long-term durability test (Fig.5(e)), the initial current density is 18.2 mA∙cm‒2, while until 100 h, the current density increases to 23.5 mA∙cm‒2. This phenomenon confirms the rationality of the above conjecture. The higher durability of the MoOx/Ni3S2/NF-11 than that of the existing catalysts endows it a broad application.
Fig.7 XPS spectra of (a) Ni 2p, (b) O 1s, and (c) S 2p for MoOx/Ni3S2/NF-11 before and after HER testing, and (d) Mo 3d for MoOx/Ni3S2/NF-11 after HER testing.

Full size|PPT slide

Overall, from the above results we see clearly that the electrocatalytic performance of the MoOx/Ni3S2/NF-11 material exceeds most known transition metal sulfides and oxides catalysts. Benefiting from the oxygen vacancy and heterostructure, the exposed Mo sites exhibit much reduced |∆GH*|, which endows the function as active sites, leading to the excellent HER performance. The outstanding catalytic activity and wonderful durability of MoOx/Ni3S2/NF-11 demonstrate it is a promising electrocatalyst for HER.

4 Conclusions

In summary, by a facile and cost-effective solvothermal method, amorphous and oxygen-deficient MoOx integrated with Ni3S2 nanoparticles are embedded in the Ni foam to fabricate the new electrocatalysts of MoOx/Ni3S2/NF-mn. The optimum oxygen-deficient MoOx/Ni3S2/NF-11 exhibits remarkable HER activity, requiring a much low η10 value of 76 mV and an extremely small Tafel slope of 46 mV∙dec‒1 in acidic media. The oxygen vacancy leaves the Mo sites exposed, and the reduced |∆GH*| of Mo sites ensures the enhancement of HER activity. Moreover, the formed MoOx/Ni3S2 heterostructure further reduces the |∆GH*| of Mo and O sites, which is another key factor in improving the catalytic activity. The high activity of MoOx/Ni3S2/NF-11 can be maintained for more than 100 h at constant potential of 180 mV. The outside acid-resistant MoOx layer encases the inside Ni3S2 and NF which are unstable under acidic conditions, thus increasing the acid resistance of the overall catalyst. In addition, the heterostructure can effectively reduce the formation energy of oxygen vacancy, which leads to the in situ reconstruction of the catalysts during the electrocatalytic process, thus promoting HER performance. As a high-performance and ultra-durable electrocatalyst, MoOx/Ni3S2/NF-11 might be regarded as a cost-effective candidate to take the place of noble metals catalysts used in HER. This work could provide new ideas for the design of other oxygen-deficient materials, therefore starting new opportunities to develop high-performance materials for HER or other applications.

Acknowledgements

Experimental work is supported by the National Natural Science Foundation of China (Grant No. 22176017), Scientific Research Project of the Ningxia Higher Education Department of China (Grant No. NGY2020034) and CAS “Light of West China Program (Grant No. XAB2020YW16)”.

Electronic Supplementary Material

Supplementary material is available in the online version of this article at https://dx.doi.org/10.1007/s11705-022-2228-1 and is accessible for authorized users.
1
Zou X, Zhang Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chemical Society Reviews, 2015, 44(15): 5148–5180

DOI

2
Turner J A. Sustainable hydrogen production. Science, 2004, 305(5686): 972–974

DOI

3
Holladay J D, Hu J, King D L, Wang Y. An overview of hydrogen production technologies. Catalysis Today, 2009, 139(4): 244–260

DOI

4
Chen W F, Muckerman J T, Fujita E. Recent developments in transition metal carbides and nitrides as hydrogen evolution electrocatalysts. Chemical Communications, 2013, 49(79): 8896–8909

DOI

5
Zhang J, Wang T, Liu P, Liao Z, Liu S, Zhuang X, Chen M, Zschech E, Feng X. Efficient hydrogen production on MoNi4 electrocatalysts with fast water dissociation kinetics. Nature Communications, 2017, 8(1): 15437

DOI

6
Stamenkovic V R, Mun B S, Arenz M, Mayrhofer K J J, Lucas C A, Wang G F, Ross P N, Markovic N M. Trends in electrocatalysis on extended and nanoscale Pt-bimetallic alloy surfaces. Nature Materials, 2007, 6(3): 241–247

DOI

7
Wang K W, She X L, Chen S, Liu H L, Li D H, Wang Y, Zhang H W, Yang D J, Yao X D. Boosting hydrogen evolution via optimized hydrogen adsorption at the interface of CoP3 and Ni2P. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2018, 6(14): 5560–5565

DOI

8
Lu W, Song Y, Dou M, Ji J, Wang F. Ni3S2@MoO3 core/shell arrays on Ni foam modified with ultrathin CdS layer as a superior electrocatalyst for hydrogen evolution reaction. Chemical Communications, 2018, 54(6): 646–649

DOI

9
Wang X, Ma W, Ding C, Xu Z, Wang H, Zong X, Li C. Amorphous multi-elements electrocatalysts with tunable bifunctionality toward overall water splitting. ACS Catalysis, 2018, 8(11): 9926–9935

DOI

10
Liang Q, Jin H, Wang Z, Xiong Y, Yuan S, Zeng X, He D, Mu S. Metal–organic frameworks derived reverse-encapsulation Co-NC@Mo2C complex for efficient overall water splitting. Nano Energy, 2019, 57: 746–752

DOI

11
Li L, Zhang T, Yan J, Cai X, Liu S. P doped MoO3−x nanosheets as efficient and stable electrocatalysts for hydrogen evolution. Small, 2017, 13(25): 1700441

DOI

12
Sinaim H, Ham D J, Lee J S, Phuruangrat A, Thongtem S, Thongtem T. Free-polymer controlling morphology of α-MoO3 nanobelts by a facile hydrothermal synthesis, their electrochemistry for hydrogen evolution reactions and optical properties. Journal of Alloys and Compounds, 2012, 516: 172–178

DOI

13
Zhu Y H, Yao Y, Luo Z, Pan C Q, Yang J, Fang Y R, Deng H T, Liu C X, Tan Q, Liu F D, Guo Y. Nanostructured MoO3 for efficient energy and environmental catalysis. Molecules, 2020, 25(1): 26

14
Li J, Cheng Y, Zhang J, Fu J, Yan W, Xu Q. Confining Pd nanoparticles and atomically dispersed Pd into defective MoO3 nanosheet for enhancing electro- and photocatalytic hydrogen evolution performances. ACS Applied Materials & Interfaces, 2019, 11(31): 27798–27804

DOI

15
Xue X, Zhang J, Saana I A, Sun J, Xu Q, Mu S. Rational inert-basal-plane activating design of ultrathin 1T′ phase MoS2 with a MoO3 heterostructure for enhancing hydrogen evolution performances. Nanoscale, 2018, 10(35): 16531–16538

DOI

16
Liu P T, Zhu J Y, Zhang J Y, Xi P X, Tao K, Gao D Q, Xue D S. P dopants triggered new basal plane active sites and enlarged interlayer spacing in MoS2 nanosheets toward electrocatalytic hydrogen evolution. ACS Energy Letters, 2017, 2(4): 745–752

DOI

17
Sadhanala H K, Harika V K, Penki T R, Aurbach D, Gedanken A. Ultrafine ruthenium oxide nanoparticles supported on molybdenum oxide nanosheets as highly efficient electrocatalyst for hydrogen evolution in acidic medium. ChemCatChem, 2019, 11(5): 1495–1502

DOI

18
Lim K J H, Yilmaz G, Lim Y F, Ho G W. Multi-compositional hierarchical nanostructured Ni3S2@MoSx/NiO electrodes for enhanced electrocatalytic hydrogen generation and energy storage. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2018, 6(41): 20491–20499

DOI

19
Kou T, Smart T, Yao B, Chen I, Thota D, Ping Y, Li Y. Theoretical and experimental insight into the effect of nitrogen doping on hydrogen evolution activity of Ni3S2 in alkaline medium. Advanced Energy Materials, 2018, 8(19): 1703538

DOI

20
Feng L L, Yu G, Wu Y, Li G D, Li H, Sun Y, Asefa T, Chen W, Zou X. High-index faceted Ni3S2 nanosheet arrays as highly active and ultrastable electrocatalysts for water splitting. Journal of the American Chemical Society, 2015, 137(44): 14023–14026

DOI

21
Chang Y H, Lin C T, Chen T Y, Hsu C L, Lee Y H, Zhang W, Wei K H, Li L J. Highly efficient electrocatalytic hydrogen production by MoSx grown on graphene-protected 3D Ni foams. Advanced Materials, 2013, 25(5): 756–760

DOI

22
Tang C, Pu Z, Liu Q, Asiri A M, Luo Y, Sun X. Ni3S2 nanosheets array supported on Ni foam: a novel efficient three-dimensional hydrogen-evolving electrocatalyst in both neutral and basic solutions. International Journal of Hydrogen Energy, 2015, 40(14): 4727–4732

DOI

23
Cao J, Zhou J, Zhang Y, Wang Y, Liu X. Dominating role of aligned MoS2/Ni3S2 nanoarrays supported on three-dimensional Ni foam with hydrophilic interface for highly enhanced hydrogen evolution reaction. ACS Applied Materials & Interfaces, 2018, 10(2): 1752–1760

DOI

24
Yang Y, Yao H, Yu Z, Islam S M, He H, Yuan M, Yue Y, Xu K, Hao W, Sun G, Li H, Ma S, Zapol P, Kanatzidis M G. Hierarchical nanoassembly of MoS2/Co9S8/Ni3S2/Ni as a highly efficient electrocatalyst for overall water splitting in a wide pH range. Journal of the American Chemical Society, 2019, 141(26): 10417–10430

DOI

25
Li T T, Zuo Y P, Lei X M, Li N, Liu J W, Han H Y. Regulating the oxidation degree of nickel foam: a smart strategy to controllably synthesize active Ni3S2 nanorod/nanowire arrays for high-performance supercapacitors. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2016, 4(21): 8029–8040

DOI

26
Tang T, Jiang W J, Niu S, Liu N, Luo H, Chen Y Y, Jin S F, Gao F, Wan L J, Hu J S. Electronic and morphological dual modulation of cobalt carbonate hydroxides by Mn doping toward highly efficient and stable bifunctional electrocatalysts for overall water splitting. Journal of the American Chemical Society, 2017, 139(24): 8320–8328

DOI

27
He W, Wang C, Li H, Deng X, Xu X, Zhai T. Ultrathin and porous Ni3S2/CoNi2S4 3D-network structure for superhigh energy density asymmetric supercapacitors. Advanced Energy Materials, 2017, 7(21): 1700983

DOI

28
Yang J, Zhang F J, Wang X, He D S, Wu G, Yang Q H, Hong X, Wu Y, Li Y D. Porous molybdenum phosphide nano-octahedrons derived from confined phosphorization in UIO-66 for efficient hydrogen evolution. Angewandte Chemie International Edition, 2016, 55(41): 12854–12858

DOI

29
Weber T, Muijsers J C, Niemantsverdriet J W. Structure of amorphous MoS3. Journal of Physical Chemistry, 1995, 99(22): 9194–9200

DOI

30
Li L D, Yan J Q, Wang T, Zhao Z J, Zhang J, Gong J L, Guan N J. Sub-10 nm rutile titanium dioxide nanoparticles for efficient visible-light-driven photocatalytic hydrogen production. Nature Communications, 2015, 6(1): 10

DOI

31
Li J, Cheng Y, Zhang J, Fu J, Yan W, Xu Q. Confining Pd Nanoparticles and atomically dispersed Pd into defective MoO3 nanosheet for enhancing electro- and photocatalytic hydrogen evolution performances. ACS Applied Materials & Interfaces, 2019, 11(31): 27798–27804

DOI

32
Kang Q, Cao J Y, Zhang Y J, Liu L Q, Xu H, Ye J H. Reduced TiO2 nanotube arrays for photoelectrochemical water splitting. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2013, 1(18): 5766–5774

DOI

33
Yan J Q, Zhang Y X, Liu S Z, Wu G J, Li L D, Guan N J. Facile synthesis of an iron doped rutile TiO2 photocatalyst for enhanced visible-light-driven water oxidation. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2015, 3(43): 21434–21438

DOI

34
Xie F, Wu H, Mou J, Lin D, Xu C, Wu C, Sun X. Ni3N@Ni-Ci nanoarray as a highly active and durable non-noble-metal electrocatalyst for water oxidation at near-neutral pH. Journal of Catalysis, 2017, 356: 165–172

DOI

35
Zhou W, Wu X J, Cao X, Huang X, Tan C, Tian J, Liu H, Wang J, Zhang H. Ni3S2 nanorods/Ni foam composite electrode with low overpotential for electrocatalytic oxygen evolution. Energy & Environmental Science, 2013, 6(10): 2921

DOI

36
Chia X, Sutrisnoh N A A, Pumera M. Tunable Pt-MoSx hybrid catalysts for hydrogen evolution. ACS Applied Materials & Interfaces, 2018, 10(10): 8702–8711

DOI

37
Kuang P Y, Tong T, Fan K, Yu J G. In situ fabrication of Ni−Mo bimetal sulfide hybrid as an efficient electrocatalyst for hydrogen evolution over a wide pH range. ACS Catalysis, 2017, 7(9): 6179–6187

DOI

38
Wang B, Huang H, Sun T, Yan P, Isimjan T T, Tian J, Yang X. Dissolution reconstruction of electron-transfer enhanced hierarchical NiSx−MoO2 nanosponges as a promising industrialized hydrogen evolution catalyst beyond Pt/C. Journal of Colloid and Interface Science, 2020, 567: 339–346

DOI

39
Cheng Z, Abernathy H, Liu M L. Raman spectroscopy of nickel sulfide Ni3S2. Journal of Physical Chemistry C, 2007, 111(49): 17997–18000

DOI

40
Li Z, Ma J, Zhang B, Song C, Wang D. Crystal phase- and morphology-controlled synthesis of MoO3 materials. CrystEngComm, 2017, 19(11): 1479–1485

DOI

41
Qi K, Yu S S, Wang Q Y, Zhang W, Fan J C, Zheng W T, Cui X Q. Decoration of the inert basal plane of defect-rich MoS2 with Pd atoms for achieving Pt-similar HER activity. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2016, 4(11): 4025–4031

DOI

42
Huang H L, Huang J Y, Liu W P, Fang Y P, Liu Y. Ultradispersed and single-layered MoS2 nanoflakes strongly coupled with graphene: an optimized structure with high kinetics for the hydrogen evolution reaction. ACS Applied Materials & Interfaces, 2017, 9(45): 39380–39390

DOI

43
Xiong J, Li J, Shi J W, Zhang X L, Suen N T, Liu Z, Huang Y J, Xu G X, Cai W W, Lei X R, Feng L, Yang Z, Huang L, Cheng H. In situ engineering of double-phase interface in Mo/Mo2C heteronanosheets for boosted hydrogen evolution reaction. ACS Energy Letters, 2018, 3(2): 341–348

DOI

44
Manikandan A, Ilango P R, Chen C W, Wang Y C, Shih Y C, Lee L, Wang Z M M, Ko H, Chueh Y L. A superior dye adsorbent towards the hydrogen evolution reaction combining active sites and phase-engineering of (1T/2H) MoS2/MoO3 hybrid heterostructured nanoflowers. Journal of Materials Chemistry A: Materials for Energy and Sustainability, 2018, 6(31): 15320–15329

DOI

45
Huang C, Pi C R, Zhang X M, Ding K, Qin P, Fu J J, Peng X, Gao B, Chu P K, Huo K F. In situ synthesis of MoP nanoflakes intercalated N-doped graphene nanobelts from MoO3-amine hybrid for high-efficient hydrogen evolution reaction. Small, 2018, 14(25): 7

DOI

46
Wu H B, Xia B Y, Yu L, Yu X Y, Lou X W. Porous molybdenum carbide nano-octahedrons synthesized via confined carburization in metal−organic frameworks for efficient hydrogen production. Nature Communications, 2015, 6(1): 8

DOI

47
Chen X, Liu G, Zheng W, Feng W, Cao W, Hu W, Hu P. Vertical 2D MoO2/MoSe2 core-shell nanosheet arrays as high-performance electrocatalysts for hydrogen evolution reaction. Advanced Functional Materials, 2016, 26(46): 8537–8544

DOI

48
Zhu L F, Liu L J, Huang G M, Zhao Q. Hydrogen evolution over N-doped CoS2 nanosheets enhanced by superaerophobicity and electronic modulation. Applied Surface Science, 2020, 504: 144490

DOI

49
He L, Zhang W, Mo Q, Huang W, Yang L, Gao Q. Molybdenum carbide-oxide heterostructures: in situ surface reconfiguration toward efficient electrocatalytic hydrogen evolution. Angewandte Chemie International Edition, 2020, 59(9): 3544–3548

DOI

50
Yilmaz G, Yang T, Du Y, Yu X, Feng Y, Shen L, Ho G. Stimulated electrocatalytic hydrogen evolution activity of MOF-derived MoS2 basal domains via charge injection through surface functionalization and heteroatom doping. Advancement of Science, 2019, 6(15): 1900140

Outlines

/