Shaista Zubaida,
Misbah Jabeena,
Hirra Ahmada,
Abid Ur Rehman Chaudharya,
Syed Ali Raza Naqvib,
Aliya Tufaila,
Atta Ullah Shahc,
Yaqoob Khan*d and
Tauqir A. Sherazi‡
*a
aDepartment of Chemistry, COMSATS University Islamabad, Abbottabad Campus, Abbottabad 22060, Pakistan. E-mail: sherazi@cuiatd.edu.pk
bDepartment of Chemistry, Government College University Faisalabad, Faisalabad 38040, Pakistan
cNational Institute of Lasers and Optronics College, Pakistan Institute of Engineering and Applied Sciences, Islamabad 45650, Pakistan
dNanosciences and Technology Department, National Centre for Physics, Islamabad 45320, Pakistan. E-mail: yaqoob@ncp.edu.pk
First published on 17th July 2025
Transition-metal chalcogenides have emerged as strong candidates for the hydrogen evolution reaction (HER) due to their exceptional catalytic activity. A facile synthetic method of Ti3C2Tx MXene, and transition-metal chalcogenides (specifically nickel-molybdenum sulfide), and their hybrid formation via in situ growth of nickel-molybdenum sulfide within MXene 2D nanosheets (NiMo3S4-MXene) is reported. The synthesis of these electrocatalysts was confirmed through X-ray diffraction spectroscopy (XRD), scanning electron microscopy (SEM), and energy-dispersive X-ray spectroscopy (EDS). Electrochemical studies of NiMo3S4 and NiMo3S4-MXene were conducted after depositing these materials onto nickel foam, which served as the “current collector”. The HER performance of NiMo3S4 was significantly enhanced in an alkaline medium after forming a hybrid with MXene (NiMo3S4-MXene). The overpotential was found to be 104 mV at 10 mA cm−2, and the Tafel slope for the HER was 52 mV dec−1. NiMo3S4-MXene exhibited remarkable durability and a low charge-transfer resistance (Rct = 2.18 Ω). The high electrical conductivity of MXene ensured efficient electron transport to active sites, whereas NiMo3S4 offered excellent catalytic activity, facilitating proton adsorption and hydrogen generation. Thus, a synergistic effect arises from the complementary properties of the components in the hybrid electrocatalyst, resulting in favourable characteristics for the HER. These findings suggest that the transition-metal chalcogenides, in combination with MXene, could serve as efficient and durable electrocatalysts for electrochemical water-splitting.
Hydrogen fuel is considered one of the most promising non-carbonaceous, environmentally friendly energy carriers.2,3 Water-splitting is an emerging technique for hydrogen production, requiring an ideal potential of 1.23 V to split water into hydrogen and oxygen gases. In practice, this potential reaches up to 1.8 V, which is quite high and makes the thermodynamic process expensive.4,5 The efficiency of water-splitting is limited due to inevitable kinetic overpotentials. These overpotentials can be minimized by applying efficient electrocatalysts. Noble metal-based electrocatalysts, such as Pt/C, are commonly used for the hydrogen evolution reaction (HER) via electrochemical water-splitting.6,7 Although Pt/C exhibits excellent performance in both acidic and alkaline media, its high cost and scarcity make it unsuitable for large-scale HER applications. Therefore, there is a strong need to explore efficient yet cost-effective non-noble metal-based electrocatalysts for the HER.8,9 However, non-noble metals are unstable in acidic environments; therefore, using an alkaline medium is a preferable alternative for non-noble metal-based electrocatalysts. These electrocatalysts can be rationally designed to possess high electrical conductivity, enhanced surface hydrophilicity, superior electrocatalytic activity, good structural stability, and abundant active sites.10–14 In this regard, the MXene family (e.g., Ti3C2, Ti2C, and V2C) has attracted considerable interest as a promising class of HER electrocatalysts due to their two dimensional (2D) morphology, excellent metallic conductivity, strong mechanical strength, and highly hydrophilic surfaces.15 MXene formation relies on the etching of the MAX phase (e.g., Ti3AlC2), where various terminating groups (–F, –O and –OH) lead to the formation of the stable nanosheets of the corresponding MXene, such as Ti3C2Tx if Ti3AlC2 is used as the MAX phase.16 These ultrathin MXene sheets are prone to restacking. Therefore, it is necessary to introduce a spacer or incorporate them into a 2D or 3D framework to minimize this tendency.17–19 Conversely, cost-effective transition-metal chalcogenides (TMCs)-based electrocatalysts, such as Ni3S2, NiS, CoS, Fe2S, and MoS2, have been developed with promising properties, including high HER activity and abundant active sites.20 A heterogeneous bimetallic sulfide/phosphide-based electrocatalyst supported on nickel foam (NiFeSP/NF) has been reported, exhibiting remarkable activity in alkaline media with a low overpotential of 91 mV at a current density of 10 mA cm−2 for the HER. A few reports have also highlighted advances in bimetallic sulfide-based hybrid electrocatalysts coupled with 2D nanomaterials.21,22 Additionally, a hollow Chevrel-phase NiMo3S4-based electrocatalyst has also been studied, offering an overpotential of 257 mV at a current density of 10 mA cm−2 and a Tafel slope of 98 mV dec−1.23 MXene-based co-catalysts, in combination with metal sulfides, such as ZnS or Zn–CdS, have exhibited enhanced water-splitting characteristics.24 These encouraging results have motivated researchers to further explore MXene in combination with bimetallic sulfides. Hence, the integration of MXene sheets with TMCs or bimetallic sulfides could lead to remarkable HER performance through a synergistic effect.
Herein, we present a facile synthesis of a bimetallic sulfide (NiMo3S4)-based electrocatalyst and its composite with MXene (NiMo3S4-MXene) via in situ hydrothermal synthesis. NiMo3S4-MXene exhibited a significantly enhanced HER performance in alkaline media, as evidenced by its lower overpotential and Tafel slope values, which are very close to those of the noble metal Pt/C. The synergistic effect between titanium-carbide MXene (Ti3C2Tx) and nickel-molybdenum sulfide (NiMo3S4) in the HER arises from their complementary properties, significantly enhancing the overall catalytic performance. The high electrical conductivity of MXenes ensured efficient electron transport to the active sites, while NiMo3S4 provided excellent catalytic activity by facilitating proton adsorption and hydrogen generation. The 2D layered structure of MXene increased the surface area, exposing more NiMo3S4 active sites and promoting faster reaction kinetics. This combination also led to improved charge transfer, thereby lowering the overpotential required for the HER. Additionally, the structural stability of MXenes supported the durability of the catalyst by preventing the degradation of NiMo3S4 in alkaline media. This synergy resulted in a highly efficient, durable, and scalable catalyst for the HER, outperforming the individual components. The large specific surface area provided by MXene in the hybrid electrocatalyst increased the exposure of NiMo3S4 active sites, thereby boosting catalytic activity. It is anticipated that bimetallic sulfide-based hybrid electrocatalysts will open up new avenues developing other rationally designed electrochemically active nanostructures aimed at promising applications in the field of energy storage and conversion.
The mixture was sonicated for 1 h and stirred for 30 min. The solution was then transferred to a 500 mL Teflon™-lined autoclave, and the hydrothermal reaction was conducted at 180 °C for 5 h. The resulting product was centrifuged, washed, and dried, following the same procedure described for the synthesis of NiMo3S4. Scheme 1 presents an illustration of the electrocatalyst preparation.
![]() | (1) |
The pH of the electrolyte (KOH) was 14, and the standard electrode potential (E°) of the Ag/AgCl electrode was 0.197 V vs. RHE at 25 °C. Overpotential for the HER was calculated using the following equation:
η = −ERHE | (2) |
Linear sweep voltammetry (LSV) was performed for the HER at a scan rate of 5 mV s−1, within a potential window of 0 to −0.677 V. To assess stability, LSV was recorded before and after 1000 cyclic voltammetry (CV) cycles at scan rate of 50 mV s−1 in the same potential window. Electrochemical impedance spectroscopy (EIS) was carried out over a frequency range from 100 kHz to 0.01 Hz. The obtained data were well fitted to a constant phase element (CPE) with a diffusion circuit; using this fitting model, the charge transfer resistance and solution resistance were calculated. The electrochemically active surface area (ECSA) was estimated by calculating double-layer capacitance (Cdl) using CV at different scan rates ranging from 5 mV s−1 to 100 mV s−1, within a non-Faradic region. A linear slope corresponding to Cdl was obtained by plotting various scan speeds versus the difference in current density (Δj) between anodic and cathodic current (janodic − jcathodic). All electrochemical studies were repeated thrice, and data represent the average of these replicates.
![]() | ||
Fig. 1 SEM images of (a) titanium aluminium carbide MAX phase, (b) Ti3C2Tx MXene synthesized by etching of the MAX phase, (c) NiMo3S4, and (d) NiMo3S4-MXene and (e–j) EDS mapping of NiMo3S4-MXene. |
Energy dispersive X-ray spectroscopy (EDS) confirmed the presence of all relevant elements in NiMo3S4 and NiMo3S4-MXene, verifying the synthesis of electrocatalysts with the proposed composition. In the EDS spectrum, the abscissa represents the ionization energy, while the ordinate corresponds to counts. The EDS spectrum of the MAX phase (Fig. 2a) showed that it is composed of three main elements: titanium, aluminium, and carbon. After etching of the MAX phase with HF, the aluminium signal was significantly suppressed, indicating its removal from the structure and the subsequent formation of MXene. Etched MXene contained titanium and carbon as major elements, as shown in Fig. 2b. Fluorine was also observed as the terminal group in the EDS spectrum of MXene and originated from HF during the synthesis of MXene.
![]() | ||
Fig. 2 EDS spectrum of (a) titanium aluminium carbide MAX phase, (b) Ti3C2Tx MXene synthesized by etching of the MAX phase, (c) NiMo3S4, and (d) NiMo3S4-MXene. |
There was a possibility of the newly synthesized MXene reacting with HF, as illustrated in eqn (3)–(5). Although the reaction of MXene with water is dominant, it also reacts with HF, resulting in the formation of –OH− and –F− terminal groups on the MXene surface.27
![]() | (3) |
Ti3C2 + 2H2O → Ti3C2(OH)2 + H2 | (4) |
Ti3C2 + 2HF → Ti3C2F2 + H2 | (5) |
The EDX spectrum of NiMo3S4 confirmed the presence of nickel and molybdenum metals along with sulfur (Fig. 2c). In the NiMo3S4-MXene hybrid structure, the elements detected were nickel, molybdenum, sulfur, titanium, and carbon, as shown in Fig. 2d. Moreover, the terminal fluorine on MXene was eliminated after its composite formation with NiMo3S4.
XRD analysis was performed to confirm the synthesis and determine structural information. The overlay of the XRD pattern of all samples is shown in Fig. 3. The black curve is the XRD of the MAX phase (titanium aluminium carbide), a precursor of MXene. It shows diffraction peaks at, 9.6°, 19.2°, 34.1°, 39°, and 41.8° corresponding to the (002), (004), (101), (104), and (105) crystalline planes of hexagonal Ti3AlC2, (JCPDS card # = 00-052-0875). After etching with HF (49%), the MAX phase was converted to MXene, as shown by the XRD pattern presented in green in Fig. 3.
The diffraction peaks appeared at 7°, 14.2°, and 16.8°, corresponding to the (002), (004), and (006) planes of cubic Ti3C2Tx (MXene). XRD analysis provided clear confirmation that Ti3C2 MXene had been synthesized via etching. In particular, the transformation from the Ti3AlC2 MAX phase to Ti3C2 MXene was evidenced by the shifting of (002) from 9.6° to 7°. This downward shift reflects expansion of the MXene lattice and an increase in d-spacing caused by the removal of aluminum and its partial replacement with –F and –OH groups during etching. Additionally, the disappearance of the characteristic Al-related peak at 39° further confirmed the complete removal of the Al layers from the Ti3AlC2 precursor, which aligns with the literature.28 Moreover, the absence of the TiO2 peak in the XRD pattern of MXene indicated that synthesis did not involve oxidation. The peaks appeared in the XRD curve of NiMo3S4 at 2θ values of 20.4°, 30°, 37.9°, 45°, 45.7°, and 53.8° and were attributed to the (110), (−120), (−220), (−311), (221), and (222) planes of rhombohedral NiMo3S4, respectively (JCPDS card # = 01-071-0011). NiMo3S4-MXene showed peaks of both MXene (indicated by green leaf symbol) and NiMO3S4 (indicated by a red leaf symbol), which confirmed the synthesis of NiMo3S4-MXene. Some planes of NiMO3S4 disappeared in the composite. New peaks appeared at 2θ of 14.2°, 28.8°, 40.6°, and 46.8° and corresponded to the (010), (−120), (−221), and (−131) planes of rhombohedral NiMo3S4, respectively (JCPDS card # = 01-071-0011). Hence, some planes were suppressed while others became prominent after composite formation. It is obvious from Fig. 3 that the sharp peaks of NiMo3S4 broadened when it combined with MXene to form the hybrid nanostructure NiMo3S4-MXene. The reason behind this is etching or delamination of MXene that affected the overall crystallinity of MXene nanosheets.29
X-ray photoelectron spectroscopy (XPS) was executed to investigate the surface electronic states and chemical composition of the NiMo3S4-MXene hybrid. XPS spectra (Fig. 4) confirmed the presence of Ti, C, O, Mo, Ni, S elements in the prepared electrode material. The XPS spectrum for Ti 2p (Fig. 4a) shows typical doublet peaks corresponding to Ti 2p3/2 and Ti 2p1/2, located at binding energies (BE) of 457.6 eV and 464.5 eV, respectively.30 The Ni 2p spectrum exhibited a high-energy band at of 870.6 eV for Ni 2p1/2 and a low-energy band at 854.4 eV for Ni 2p3/2 (Fig. 4e). Additionally, two satellite peaks at 860.3 and 873.8 eV were observed. The peaks featured at 161.2 eV and 162.8 eV (Fig. 4f) represent S 2p3/2 and S 2p1/2 in NiS, respectively. The partial S oxidation resulted in the appearance of one satellite peak at 168.2.31
The peaks observed in Fig. 4d at binding energies of 224.7 eV, 227.5 eV, and 231.6 eV corresponded to S 2s, Mo 3d5/2, and Mo 3d3/2, respectively, thereby confirming the presence of Mo4+ in the NiMo3S4-MXene hybrid. Additionally, three peaks at 228 eV, 229 eV, and 232 eV could be attributed to MoSxOγ species, indicating partial surface oxidation of MoS2 due to air exposure.32 MXene often contains carbon originating from the MAX-phase precursor or surface functionalization, which give rise to C 1s signal (Fig. 4b). The C 1s spectrum revealed distinct chemically shifted components, which could be deconvoluted as CC/C–C (284 eV), C–OH (285 eV), and C
O (288 eV) (Fig. 4(b)). In Fig. 4c, the O 1s signal appears at a binding energy in the range of 530–534 eV.33
The coexistence of mixed valence states provides numerous active sites and facilitates rich redox reactions, thereby significantly enhancing the overall electrochemical performance of the hybrid electrode. These combined features make the NiMo3S4-MXene a promising candidate for high-performance applications in overall water-splitting.
To evaluate the electrocatalytic HER performance of the prepared electrocatalysts, NiMo3S4 and NiMo3S4-MXene were tested in N2-saturated 1.0 M KOH solution using the three-electrode system. In this setup, nickel foam was used as support material for depositing the active material to prepare the working electrode. For comparison, pristine Ni foam, and a working electrode prepared using pristine MXene on Ni-foam, were used with approximately the same loading as used for other electrocatalytic materials. Platinum wire was also employed as a benchmark reference for comparison. Fig. 5a shows the polarization curves of electrocatalysts obtained through linear sweep voltammetry (LSV), performed at a scan rate of 5 mV s−1. The NiMo3S4-MXene-based electrode exhibited remarkable HER activity with an onset potential of 21 mV, as compared with NiMo3S4 (33 mV), MXene (46 mV), and bare NF (106 mV). Overpotential is another important parameter and direct indicator to determine the HER performance of an electrocatalyst. NiMo3S4-MXene demonstrated the lowest overpotential of 104 mV at a current density of 10 mA cm−2, outperforming the individual components used in our study (i.e., NiMo3S4 (241 mV) and MXene (282 mV)). Its performance approached that of Pt wire, a well-established benchmark for the HER. The overpotential of bare NF at 10 mA cm−2 was also measured and found to be 397 mV.
The enhanced catalytic activity of NiMo3S4-MXene could be attributed to a synergistic effect, whereby NiMo3S4 provides excellent electrocatalytic activity with the high electrical conductivity and 2D surface area of MXene. This was evidenced by a significant reduction in both overpotential and onset potential, indicating improved catalytic performance. As a result, more active sites are exposed, which effectively promotes the HER.34,35 The significantly enhanced HER activity of NiMo3S4-MXene indicated that the designed electrocatalyst performed comparably with, or even better than, several other non-noble metals-based electrocatalysts (Table S2†). Pt/C was used as the noble metal-based reference electrocatalyst and exhibited excellent HER performance, with an onset potential of 14 mV, a value not yet achieved by non-noble metal electrocatalysts. However, replacing the noble metal with non-noble metals exhibiting a comparable performance is a key objective in advancing technology for widespread applications.
Fig. 5b presents the Tafel plots corresponding to polarization curves. The enhanced HER kinetics of NiMo3S4-MXene was supported by examining the Tafel slope. NiMo3S4-MXene exhibited a low Tafel slope of 52 mV dec−1, which was superior to those of individual components (NiMo3S4 (68 mV dec−1) and MXene (98 mV dec−1)). The small Tafel slope value of NiMo3S4-MXene suggested the presence of more active sites and indicated a rapid increase in the hydrogen generation rate along with applied overpotential. This finding is consistent with the high activity observed in its polarization curve.
MXene notably enhances catalytic kinetics due to its metallic conductivity and water-interactive terminations (±O, –F, –OH), which facilitate rapid electron transport at the catalyst–electrolyte interface and promote efficient proton adsorption. For instance, recent work on alkaline Ti3C2 MXene-based catalysts demonstrated significantly improved HER current densities and reduced overpotentials, underscoring their role in accelerating reaction rates. Its 2D structure and tunable terminal chemistry also increase active-site exposure and electrolyte penetration, collectively boosting HER kinetics.36–38
According to classical theory, the HER in alkaline media proceeds through two main steps (eqn (6)–(8)). The first step involves electron transfer leading to adsorbed hydrogen via water dissociation and is known as the Volmer step. The second step can proceed in two ways to produce one hydrogen molecule. If the surface coverage of adsorbed hydrogen content on the electrode is low, another H2O molecule acquires an electron and reacts with the adsorbed hydrogen to produce molecular hydrogen (Heyrovsky step). In contrast, if the surface coverage of adsorbed hydrogen content is sufficiently high, direct coupling takes place between two adsorbed hydrogen to give molecular hydrogen (Tafel step).39 Therefore, the HER follows a Volmer–Heyrovsky or Volmer–Tafel mechanism as per the following equations.
Volmer step:
H2O + M + e− → M–Hads + OH− | (6) |
Heyrovsky step:
H2O + M–Hads + e− → H2 + M + OH− | (7) |
Tafel step:
2M–Hads → H2 + 2M | (8) |
Typically, Tafel slopes of 120, 40, 30 mV dec−1 correspond to the Volmer, Heyrovsky, and Tafel rate-determining steps, respectively. The Tafel slope determines the rate-limiting step in HER kinetics, with a lower slope indicating a faster HER reaction rate. The Tafel slope of the NiMo3S4-MXene was measured to be 52 mV dec−1, which was higher than that of the noble metal-based electrocatalyst Pt/C (30 mV dec−1), but lower than that of pristine MXene (98 mV dec−1) and NiMo3S4 (68 mV dec−1). This result indicated that the superior HER reaction kinetics of NiMo3S4-MXene were through the water-dissociation reaction and followed the Volmer–Heyrovsky mechanism for hydrogen production, whereas Heyrovsky was the rate-determining step. The overpotential values of all prepared electrocatalysts at specific current densities (10, 50, 100 and 200 mA cm−2) are presented in Fig. S4.† NiMo3S4-MXene exhibited the lowest overpotential (104 mV) at a current density of 10 mA cm−2.
To further explore the catalytic activity of designed electrocatalysts, the ECSA was measured, which is proportional to the double-layer capacitance (Cdl). Using Cdl, the ECSA of NiMo3S4, MXene, and NiMo3S4-MXene catalysts can be calculated by the following equation:
![]() | (9) |
To calculate Cdl, a series of CV curves under alkaline conditions were obtained at multiple scan rates in the non-faradaic region (Fig. S5a and b†). Fig. 5c shows plots of Δj as a function of a scan rate that yields a straight line whose slope is taken as Cdl. The specific capacitance (Cs) was ∼35 μF cm−2 based on variously reported capacitances of metal-based electrodes in alkaline solutions.40 NiMo3S4-MXene exhibited a relatively higher Cdl (60.83 μF cm−2), greater than those of MXene (47.63 μF cm−2), and NiMo3S4 (22.38 μF cm−2). As a result, the ECSA of NiMo3S4-MXene, MXene, and NiMo3S4 was calculated to be 1.73 cm−2, 1.36 cm−2, and 0.31 cm−2, respectively, indicating that due to its high conductivity, NiMo3S4-MXene had the highest exposure of effective active sites, which enhanced its contact with the electrolyte. Surface topology can influence catalyst performance, so LSV data were normalized by the ECSA to determine the intrinsic HER activity of each catalyst, as presented in Fig. 5d. A similar trend of catalytic activity for these three electrocatalysts compared with the geometric area-normalized LSV curves (presented in Fig. 5a) was observed. The ECSA of NiMo3S4 was significantly enhanced upon forming its hybrid with MXene. Hence, its electrocatalytic performance was attributed to a synergistic effect arising from the high ECSA of MXene and superior catalytic performance of NiMo3S4. These findings are in line with the literature, which highlights the prominent effect of ECSA on HER activity and kinetics.41
EIS was carried out for the measurement of the electrochemical impedance and electron transfer properties of the prepared electrode material in frequency range of 100 kHz to 0.01 Hz. EIS results and the equivalent circuit-fitting model are presented in Fig. 6. Pristine NF was best fitted with the CPE circuit model. The obtained Nyquist plots of all electrocatalysts were best fitted to a contact phase element (CPE) with a diffusion-equivalent circuit model. An equivalent circuit was presented that is consist of a reference electrode (R.E.), along with the solution resistance (Rs), which represents the resistance of the electrolyte. Y0 denotes CPE admittance, and Wd is the Warburg element, which represents the diffusion or mass transport resistance. Additionally, the circuit included the charge transfer resistance (Rct) and working electrode (W.E.), reflecting the interfacial electrochemical dynamics at the electrode surface. The relevant parameters are summarized in ESI as Table S1.† Information about the Rs of electrode materials as well as the Rct corresponded to the electrocatalytic kinetics at the electrode/electrolyte interface (Fig. 5). All electrocatalysts experienced an Rs value close to each other, but a significant difference in their Rct values was observed. The estimated Rct for NiMo3S4, MXene, and NiMo3S4-MXene was 11.06 Ω, 2.84 Ω, and 2.18 Ω, respectively. The smallest value for NiMo3S4-MXene indicated a fast electrocatalytic process and, therefore, superior HER kinetics. The low charge transfer resistance of NiMo3S4-MXene revealed that the hybrid structure had a key role in electrocatalytic performance. The synergistic effect in NiMo3S4-MXene arises from the excellent electrocatalytic activity of NiMo3S4 and high conductivity of MXene, which collectively offered higher water oxidation and efficient charge transfer, thereby improving the overall performance.42 A summary of the HER kinetics of designed electrocatalysts is given in Table 1.
![]() | ||
Fig. 6 Nyquist plot of bare Ni-foam, MXene, NiMo3S4, and NiMo3S4-MXene measured at a potential of −0.2 V vs. RHE (the unit of Rct and Rs values is Ω). |
Electrodes | Onset potential (mV) | η10 (mV) | Tafel slope (mV dec−1) | Cdl (μF cm−2) | Rct (ohm) | Rs (ohm) |
---|---|---|---|---|---|---|
NF | 106 | 397 | 115 | 0.94 | 18 | 0.5 |
MXene | 46 | 287 | 98 | 47.63 | 2.84 | 1.08 |
NiMo3S4 | 33 | 241 | 68 | 22.38 | 11.06 | 1.12 |
NiMo3S4-MXene | 21 | 104 | 52 | 60.83 | 2.18 | 0.8 |
The stability of all prepared electrocatalysts was evaluated by performing 1000 CV cycles at a scan rate of 50 mV s−1 for the HER in 1.0 M KOH electrolyte (Fig. S6†). The electrodes were pre-activated by immersion in an electrolyte solution for 1 h, followed by 50 cycles of CV prior to recording data for the stability test. After activation, the initial LSV scan was recorded. Subsequently, another LSV scan was conducted after 1000 cycles, and the resulting polarization curves were compared, as shown in Fig. 7a and b. The LSV curve of NiMo3S4 before and after 1000 cycles of CV showed slight loss of performance as the overpotential at a certain current density increased after running 1000 cycles. However, this performance loss significantly diminished after formation of the composite of NiMo3S4 with MXene. This finding indicated that the in situ synthesis of NiMo3S4 covered the surface of MXene 2D nanosheets, preventing the restacking of MXene layers and preventing MXene from oxidation because no oxides of titanium were observed in EDX analysis. Additionally, the interaction between NiMo3S4 with MXene after hybrid formation improved the stability and preserved the active sites for long-term use to perform electrocatalytic reactions.
![]() | ||
Fig. 7 LSV curves of electrocatalysts before and after 1000 cycles. (a) NiMo3S4 and (b) NiMo3S4-MXene. (c) LSV of NiMo3S4-MXene and (d) chronoamperometry data of NiMo3S4-MXene. |
The stability of the NiMo3S4-MXene electrode was further evaluated using chronoamperometry at a constant potential of 104 mV (corresponding to a current density of approximately −10 mA cm−2) in 1.0 M KOH for ∼14 h, as shown in Fig. 7(c), with the results displayed in Fig. 7(d). Following a slight initial drop (likely due to temporary blockade of active sites by accumulated hydrogen bubbles), the current density remained stable or slightly increased over time. Continuous stirring helped release the bubbles, keeping the active sites accessible for the reaction. Overall, NiMo3S4-MXene demonstrated excellent stability over the 14 h testing period.
The overpotential of NiMo3S4-MXene compared with other MXene-based hybrid electrocatalysts reported in the literature is presented in Fig. 8, and the comparison is summarized in Table S2.† NiMo3S4-MXene outperformed several competing structures, making it a promising candidate for the HER in alkaline media, with notably low overpotential (104 mV) at current density of 10 mA cm−2.
![]() | ||
Fig. 8 Comparison of the overpotential of NiMo3S4-MXene (at a current density of 10 mA cm−2) with some relevant materials reported in the literature.26,43–48 |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra02903a |
‡ Visiting affiliation: INM, Leibniz Institute for New Materials, Campus D2.2, 66123 Saarbrücken, Germany. |
This journal is © The Royal Society of Chemistry 2025 |