Next Article in Journal
Study on the Performance and Service Life Prediction of Corrosion-Resistant Concrete Cut-Corner Square Piles
Previous Article in Journal
Effect of Ultraviolet Light on the Shear Bond Strength of Commercial Dental Adhesives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis

1
Division of System Semiconductor, Dongguk University, Seoul 04620, Republic of Korea
2
Quantum-Functional Semiconductor Research Center, Dongguk University, Seoul 04620, Republic of Korea
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(16), 3775; https://doi.org/10.3390/ma18163775
Submission received: 11 July 2025 / Revised: 5 August 2025 / Accepted: 7 August 2025 / Published: 12 August 2025
(This article belongs to the Special Issue Advanced Nanomaterials for Energy Storage and Conversion)

Abstract

For future clean and renewable energy technology, designing highly efficient and robust electrocatalysts is of great importance. Particularly, creating efficient bifunctional electrocatalysts capable of effectively catalyzing both hydrogen- and oxygen-evolution reactions (HERs and OERs) is vital for overall water electrolysis. In this study, we employ 2D molybdenum disulfide (MoS2) nanosheets and pyrolytically fabricated 2D graphitic carbon nitride (gC3N4) nanosheets to create 2D gC3N4-decorated 2D MoS2 (2D–2D gC3N4–MoS2) nanocomposites using a facile sonochemical method. The 2D–2D gC3N4–MoS2 nanocomposites show an interconnected and agglomerated structure of 2D gC3N4 nanosheets decorated on 2D MoS2 nanosheets. For water electrolysis, the gC3N4–MoS2 nanocomposites exhibit low overpotentials (OER: 225 mV, HER: 156 mV), small Tafel slope values (OER: 49 mV/dec, HER: 101 mV/dec), and excellent durability (up to 100 h for both OER and HER) at 10 mA/cm2 in 1 M KOH. Furthermore, the gC3N4–MoS2 nanocomposites show excellent overall water electrolysis performance with a low full-cell voltage (1.52 V at 10 mA/cm2) and outstanding long-term cell stability. The superb bifunctional activities of the gC3N4–MoS2 nanocomposites are attributed to the synergistic effects of 2D gC3N4 (i.e., low charge-transfer resistance) and 2D MoS2 (i.e., a large electrochemically active surface area). These findings suggest that the 2D–2D gC3N4–MoS2 nanocomposites could serve as excellent bifunctional catalysts for overall water electrolysis.

Graphical Abstract

1. Introduction

Fossil fuel scarcity, increasing energy demands, and environmental pollution concerns have stimulated significant interest in pursuing renewable, eco-friendly, and clean energy sources [1,2,3,4]. Recently, hydrogen has emerged as a leading alternative renewable energy resource because of its excellent mass–energy density, pollution-free characteristics, sustainability, renewability, and superior cleanliness [5,6,7,8]. Among all the available hydrogen production methods, electrocatalytic water electrolysis is considered a fascinating method for harvesting oxygen and hydrogen energies from water (via oxygen- and hydrogen-evolution reactions (OERs and HERs), respectively) because of its environmental friendliness, reusability, scalability, and feasibility [9,10,11,12]. Currently, Pt (HER catalyst) and IrO2/RuO2-based (OER catalyst) materials are widely believed to be highly efficient, stable, and reliable catalysts for water splitting, but their material scarcity and high cost hinder their practical applications [13,14]. Most of the reported electrocatalysts have not yet exhibited outstanding OER and HER activities simultaneously in a single electrolyte because of incompatible activity and stability [15,16,17,18]. Therefore, developing alternative catalysts with superior durability, excellent catalytic active sites, low cost, and earth-abundant resources is vital.
Recently, transition metal dichalcogenides (TMDs)-based materials have garnered considerable attention as prospective electrocatalysts for OERs and HERs because of their distinctive two-dimensional (2D) layered structure, low cost, and unique physicochemical characteristics [19,20,21,22]. Among the various TMDs, molybdenum disulfide (MoS2) is considered a substantial catalyst for water splitting owing to its high intrinsic edge activity, layered structure, excellent stability, nontoxicity, natural abundance, low cost, unique electronic structures, and good electrochemical catalytic activity [23,24,25]. Generally, MoS2 comprises two phases, i.e., 2H (semiconducting) and 1T (metallic) [26]. However, pure 2H phase MoS2 demonstrates insufficient electrocatalytic performance due to its inferior electrical conductivity, insufficient active edge sites, sluggish charge-transfer kinetics, and high hydrogen adsorption–desorption energy [27,28]. Therefore, various methods have been used to hybridize 2H MoS2 nanostructures with transition metal/metal oxide-based materials [29,30,31,32,33] and carbonaceous materials [34,35,36,37,38] in order to improve their electrochemical catalytic behavior. Recently, graphitic carbon nitride (gC3N4) has gained significant interest as a good co-catalyst for enhancing the catalytic activity of host catalysts owing to its impressive chemical stability, facile synthesis, affordability, high earth abundance, unique 2D layered structure, high nitrogen concentration, and easily adaptable structure [39,40]. The coupling of 2D MoS2 and 2D gC3N4 can notably enhance the specific surface area, electrical conductivity, and density of active catalytic sites, all of which are beneficial for improving electrocatalytic performance [34,41]. For example, Fageria et al. [34] hydrothermally synthesized the highly efficient electrocatalyst of MoS2-decorated gC3N4, demonstrating an HER overpotential of 240 mV at −10 mA/cm2 in 0.5 M H2SO4. Additionally, Zhang et al. [35] synthesized MoS2 nanosheets anchored on a gC3N4 substrate, exhibiting an HER overpotential of 158 mV at −10 mA/cm2 in 0.5 M H2SO4. Recently, Liu et al. [42] used the hydrothermal method, followed by ultrasonic techniques, to fabricate a hybrid structure of the MoS2/gC3N4 nanojunction, which exhibited effective HER activity with an overpotential of −200 mV to achieve a current density of 10 mA/cm2 in 0.5 M H2SO4. He et al. [43] fabricated a CoOx/mC@MoS2@gC3N4 composite that showed the HER overpotential of 31 mV at −10 mA/cm2 in 0.5 M H2SO4. Furthermore, Mehtab et al. [44] synthesized the MoS2/gC3N4 heterostructure through a ball-milling technique and exhibited bifunctional OER and HER activities with overpotential values of 410 and 262 mV at 20 mA/cm2 in 0.1 M KOH, respectively. Despite all the benefits of 2D MoS2 and 2D gC3N4, hybrid 2D gC3N4-decorated 2D MoS2 (2D–2D gC3N4–MoS2) nanocomposites have rarely been investigated for bifunctional water electrolysis, particularly in regards to substantial OER and HER performance in alkaline medium.
Motivated by the abovementioned backgrounds, we synthesized 2D–2D gC3N4–MoS2 nanocomposites using a facile sonochemical method and examined their bifunctional water electrolysis performances. The gC3N4–MoS2 nanocomposites demonstrated excellent electrocatalytic water-splitting activities in 1 M KOH with small overpotential values (i.e., 225 mV for OER and 156 mV for HER at 10 mA/cm2). Furthermore, the assembled gC3N4–MoS2||gC3N4–MoS2 electrolyzer demonstrated a low full cell voltage of 1.52 V at a current density of 10 mA/cm2, maintaining superior stability during prolonged operation for up to 100 h. Herein, the synthesis-to-electrocatalytic characteristics of the 2D–2D gC3N4–MoS2 nanocomposites are assessed and deliberated in detail.

2. Experiment

2.1. MoS2 Nanosheet Synthesis

Figure 1a depicts the facile sonochemical process for producing the hybrid 2D–2D gC3N4–MoS2 nanocomposites. The 2D MoS2 nanosheets were derived from commercial bulk MoS2 (Sigma-Aldrich, St. Louis, MO, USA). Initially, 2 g of bulk MoS2 was dispersed into 100 mL of deionized water and stirred continuously for 30 min. Next, the blended solution was sonicated (fultra = 35 kHz and Pultra = 240 W) for 60 min. During the ultrasonication process, the ultrasonic waves induced alternating low and high pressures in the bulk MoS2. This process facilitated cavitation-driven exfoliation of the MoS2 sub-lattices, resulting in the successful formation of layered MoS2 nanosheets. Afterward, the colloidal suspension was collected, cleaned, filtered, and parched in a 150 °C electric oven for 480 min. Finally, the MoS2 nanosheets were collected in powder form.

2.2. gC3N4 Nanosheet Synthesis

The 2D gC3N4 nanosheets were synthesized from melamine using a facile pyrolysis method [39,45,46]. Initially, 5 g of melamine (Sigma-Aldrich, St. Louis, MO, USA) was transferred to an alumina crucible and encapsulated by the crucible cover. Thereafter, the crucible setup was loaded in a muffle furnace and thermally annealed at 550 °C for 240 min in an air atmosphere. Finally, yellow-colored powdered gC3N4 nanosheets were obtained.

2.3. gC3N4–MoS2 Nanocomposite Fabrication

The 2D–2D gC3N4–MoS2 nanocomposites were fabricated via an ultrasonication process using the pyrolytically fabricated 2D gC3N4 nanosheets and the sonochemically derived 2D MoS2 nanosheets. Prior to discussing the experimental procedures, it is important to note that the gC3N4 to MoS2 ratio of 1:0.3 was selected based on insights from previous literature [47,48,49,50,51,52], as this composition was found to provide optimal material properties and enhanced electrochemical performance. For this, first, the MoS2 nanosheets (1 g) were interspersed in deionized water (100 mL) through vigorous stirring for 20 min. Then, the gC3N4 nanosheets (0.3 g) were subsequently added and mixed with the abovementioned MoS2 solution through additional stirring for 20 min. Thereafter, the gC3N4–MoS2 blended aqueous solution was ultrasonicated for 60 min (fultra = 35 kHz and Pultra = 240 W). Finally, the ultrasonicated colloidal suspension was washed, accumulated, and parched at 150 °C for 6 h to obtain the nanopowder form of 2D–2D gC3N4–MoS2.

2.4. Material Characterization

The morphological and elemental properties of MoS2 and gC3N4–MoS2 were monitored through field-emission scanning electron microscopy (FE-SEM, Clara LMH, Tescan Brno, Czech Republic) and in situ energy-dispersive X-ray spectroscopy (EDX) measurements, respectively. Moreover, the topographic features of the prepared catalysts were further assessed via transmission electron microscopy (TEM, JEM 2100F, JEOL, Tokyo, Japan). Crystallographic information was investigated through the X-ray diffraction (XRD; D8-Advance, Bruker, Billerica, MA, USA) analysis. The textural characteristics were assessed via the Brunauer–Emmett–Teller (BET, BELSORP-mini II system, MicrotracBEL, Osaka, Japan) and Barrett–Joyner–Halenda (BJH) techniques. The surface chemical states of the synthesized catalysts were examined through X-ray photoelectron spectroscopy (XPS, ESCALab250Xi system, Thermos Fisher Scientific, Waltham, MA, USA).

2.5. Electrocatalytic Measurements

The water splitting performance of the prepared MoS2 nanosheets and gC3N4–MoS2 nanocomposites was evaluated using the typical three-electrode method in an alkaline electrolyte (1 M KOH) solution using the VersaSTAT3 workstation (Ametek Scientific Company, Mahwah, NJ, USA). First, we assembled two different working electrodes using the prepared 2D MoS2 nanosheets and 2D–2D gC3N4–MoS2 nanocomposites. For this, 3 mg of each synthesized catalyst (i.e., either MoS2 or gC3N4–MoS2) was amalgamated with a 3 mL of N-methyl-2-pyrrolidinone solution and coated on nickel foam substrates (1 cm × 1 cm, 110 ppi) that were purchased from the MTI Korea Group, Seoul, Republic of Korea. Next, each substrate (i.e., catalyst-coated nickel foam) was dried at 180 °C for 480 min. Moreover, a saturated calomel electrode (SCE) (i.e., reference electrode) and a coiled Pt wire (i.e., counter electrode) were also prepared to set up the three-electrode system. After preparing all the electrode setups, the electrocatalytic characteristics were examined using the electrochemical workstation. First, the electrochemical cyclic voltammetry (CV) characteristics of the synthesized catalysts were examined in a 0–0.5 V potential range, where the scan rate was also varied from 10 to 100 mV/s. Next, the linear sweep voltammetry (LSV) characteristics were assessed at a fixed scan rate of 1 mV/s within certain potential windows (i.e., OER: −0.1–1.2 V, HER: −1–−1.8 V). Additionally, the electrochemical impedance spectroscopy (EIS) tests were performed using an AC signal amplitude of 10 mV in the 1 Hz to 10 kHz frequency range in 1 M KOH. Furthermore, the chronopotentiometry measurements were conducted at different current densities (i.e., OER: 10–100 mA/cm2, HER: −10–−100 mA/cm2), for which each step was maintained for 10 min. Here, it should be noted that the reference was standardized to the reversible hydrogen electrode (RHE) scale. Furthermore, all the polarization values and their corresponding curves were iR-corrected. The electrochemical double-layer capacitance (CDL) and its corresponding electrochemically active surface area (ECSA) of the catalysts were determined from the non-Faradaic CV curves by using the following relationships [53,54,55]:
J D L = C D L × v / A
E C S A = C D L / C e ,
where JDL, Ce, v, and A are the double-layer charging current, capacitance of the used electrolyte (0.04 mF/cm2 for KOH), the potential scan rate, and the area of electrodes, respectively. The overpotential (η) and the Tafel slope (ST) for the OER and HER were calculated using the following equations [14,56,57,58,59,60]:
E R H E = E S C E + E S C E 0 + 0.059 p H
η = E R H E 1.23   V   (for OER)
η = E R H E   (for OER)
η = S T log ( J ) + c
where E0SCE is the SCE’s standard potential, J is the current density, c is the fitting parameter, and ERHE is the RHE’s standard potential.

3. Results and Discussion

The morphology and chemical composition of the bare 2D MoS2 nanosheets and 2D–2D gC3N4–MoS2 nanocomposites were characterized via FE-SEM and in situ EDX measurements. Figure 1b–e display the FE-SEM images of the pristine MoS2 nanosheets and the hybrid gC3N4–MoS2 nanocomposites. The MoS2 sample exhibited stacked-layer nanosheet-like structures with an average length of 200–400 nm (Figure 1b,c). After ultrasonicating the bare MoS2 nanosheets together with the gC3N4 nanosheets, the sample exhibited the interconnected and agglomerated structure of the 2D gC3N4-decorated 2D MoS2 nanocomposites (Figure 1d,e). Next, the elemental composition of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites was investigated. As shown in the EDX profiles (Figure 1f,g), both samples revealed their own intrinsic Mo, S, N, and C constituents, demonstrating that the prepared materials were highly pure and free of other impurities. Here, it should be noted that Pt detected in both samples arose from the conductive coating for FE-SEM measurements, which was applied to avoid the electron charging effect.
The topography of both the bare 2D MoS2 nanosheets and 2D–2D gC3N4–MoS2 nanocomposites was further analyzed by TEM measurements. Figure 2a,b show the bright-field TEM images of the MoS2 nanosheets. The bare MoS2 exhibited a stacked-layer nanosheet morphology. In the high-resolution image (Figure 2c), the bare MoS2 shows an interlayer spacing of 0.62 nm, which corresponds to the (002) plane of hexagonal MoS2 [61]. The well-defined diffraction lattice indicates that the bare MoS2 nanosheets were single crystals corresponding to the hexagonal phase of MoS2 (Figure 2d). Unlike the bare MoS2 nanosheets, the hybrid gC3N4–MoS2 nanocomposites exhibited an interconnected and agglomerated structure of gC3N4 nanosheets decorated on stacked-layer MoS2 nanosheets (Figure 2e,f). From the high-resolution image of gC3N4–MoS2 (Figure 2g), the lattice spacings of MoS2 and gC3N4 were confirmed to be 0.62 and 0.32 nm, which are consistent with those of (002) MoS2 and (002) gC3N4, respectively. Moreover, the SAED pattern of gC3N4–MoS2 exhibited polycrystalline phases (Figure 2h) because of its microstructural hybridization.
To further understand the formation kinetics of the gC3N4–MoS2 nanocomposites, we discuss the chemical mechanisms implicated during the ultrasonication process. The sonochemical technique possesses several advantages, including rapid synthesis, low energy consumption, and the ability to generate highly dispersed nanostructures with uniform morphology. Ultrasonic irradiation induces acoustic cavitation, producing localized high temperatures and pressures that promote fast nucleation and prevent agglomeration. These features result in materials with high surface area, better crystallinity, and enhanced active site exposure—critical for improving electrocatalytic performance. In an aqueous solution, ultrasonication generates two significant radicals from water (i.e., hydrogen (H*) and hydroxyl (OH*) radicals). During the sonication of bulk materials, these H* and OH* radicals serve as reductants [62,63,64]. Consequently, bulk MoS2 (nMoS2) could be diminished into stacked-layer MoS2 nanosheets (MoS2(n)) under the high ultrasonic power in water. This sonochemical exfoliation process (i.e., micro-cavitation and shock waves) can be described by the following reactions [63,65,66,67,68]:
H 2 O S o n i c a t i o n O H * + H *
n M o S 2 + O H * + H * S o n i c a t i o n   M o S 2 ( n )
n M o S 2 + g C 3 N 4 + O H * + H * S o n i c a t i o n M o S 2 ( n ) g C 3 N 4 M o S 2 ( n ) .
Figure 2. (a,b) Bright-field TEM images, (c) high-resolution TEM image, and (d) SAED pattern of MoS2. (e,f) Bright-field TEM image, (g) high-resolution TEM image, and (h) SAED pattern of gC3N4–MoS2.
Figure 2. (a,b) Bright-field TEM images, (c) high-resolution TEM image, and (d) SAED pattern of MoS2. (e,f) Bright-field TEM image, (g) high-resolution TEM image, and (h) SAED pattern of gC3N4–MoS2.
Materials 18 03775 g002
Next, the crystallographic phases of pristine MoS2 nanosheets and gC3N4–MoS2 nanocomposites were investigated using powder XRD measurements. Figure 3a shows the XRD patterns of MoS2 and gC3N4–MoS2. Both samples revealed diffraction angles at 14.38°, 29.14°, 32.72°, 33.58°, 35.94°, 39.67°, 44.25°, 49.94°, 56.07°, 58.36°, and 60.23°, which are attributed to the (002), (004), (100), (101), (102), (103), (006), (105), (106), (110), and (008) planes of 2H-phase hexagonal MoS2, respectively (JCPDS No. 75-1539) [69,70,71]. The gC3N4–MoS2 nanocomposites displayed two additional diffraction peaks at 12.88° and 25.77°, attributed to the (100) and (002) planes of hexagonal gC3N4, respectively (JCPDS No. 87-1526) [72,73]. The XRD pattern of gC3N4–MoS2 displayed diffraction peaks from both MoS2 and gC3N4, confirming the successful fabrication of the gC3N4–MoS2 hybrid composite system. Moreover, no other secondary phases were observed in the fabricated materials, implying that the materials reflected high purity.
After confirming the successful hybridization of 2D gC3N4 and 2D MoS2, the textural characteristics were assessed through BET and BJH measurements using N2 adsorption–desorption isotherms (N2-ADIs). Figure 3b shows the N2-ADI curves of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites. Both materials exhibited the Type-II isotherm characteristics, with a typical Type-H3 hysteresis loop (classified from IUPAC), demonstrating the distinctive features of mesoporous materials [74,75,76]. Based on the BET measurements, the specific surface areas of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites were calculated to be 14 and 46 m2/g, respectively. Furthermore, BJH analysis revealed that the pore surface areas of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites were 6 and 14 m2/g, respectively (Figure 3c). Additionally, the average pore diameter of MoS2 and gC3N4–MoS2 were 23.91 and 18.48 nm, respectively. Owing to the incorporation of layered gC3N4 nanosheets into the gC3N4–MoS2 composite system, the total pore volume was greater for gC3N4–MoS2 (0.09136 cm3/g) compared to that for MoS2 (0.06124 cm3/g). The unique porosity and high surface area of the gC3N4–MoS2 nanocomposites could be beneficial for enhancing their bifunctional water-splitting performance in alkaline media, which will be discussed later in detail.
The surface composition and ionic state interaction for both MoS2 and gC3N4–MoS2 were examined using XPS measurements. The XPS full survey spectrum of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites clearly revealed their intrinsic elements, including Mo, S, C, N, and O (see Figure S1a,b). For the Mo 3d core-level spectra (Figure 4a), the bare MoS2 displayed two predominant peaks at 229.46 and 232.58 eV, associating with the Mo 3d5/2 and Mo 3d3/2 spin-orbit splitting states of Mo4+, respectively [77]. In addition, the small peak observed at 235.91 eV is correlated with the orbital electron of Mo6+, indicating the existence of small amounts of MoO3 due to the surface oxidation of MoS2 [78]. The S 2s peak observed at 226.64 eV corresponds to the hexagonal phase of MoS2 [79]. For S 2p (Figure 4b), the two peaks at 162.41 and 163.62 eV arise from the S 2p3/2 and S 2p1/2 orbit splitting states of the S2− ion, respectively [80,81]. The hybrid gC3N4–MoS2 nanocomposites clearly revealed similar features of the Mo 3d (Figure 4c) and S 2p (Figure 4d) core-level spectra. In the case of C 1s (Figure 4e), gC3N4–MoS2 clearly exhibited two carbon peaks at 284.61 and 288.36 eV, attributed to the C–C and N–C=N bonds, respectively [41,82]. Moreover, as shown in Figure 4f, the N 1s core-level spectra contained two peaks at 398.24 and 399.93 eV, ascribed to the pyridinic C–N=C and triazine N–(C3) bonds, respectively [34,83]. Additionally, the oxygen-related peaks were observed at 532.15 and 533.92 eV, attributed to the Mo–O bond and chemisorbed oxygen [34,41], respectively (see Figure S1c,d). These results clearly indicate that the gC3N4 nanosheets are well-decorated on the MoS2 nanosheets in the composite system.
To investigate the impact of 2D–2D gC3N4–MoS2 hybridization on its electrocatalytic performance, we assessed the electrochemical CV characteristics of the bare MoS2 nanosheets and gC3N4–MoS2 nanocomposites. As shown in Figure 5a,b, both the MoS2 and gC3N4–MoS2 catalysts display noticeable redox peaks in their CV curves. These peaks are mainly due to the electrochemical activity of the NF substrate, which undergoes a reversible Ni2+ to Ni3+ oxidation and Ni3+ to Ni2+ reduction in alkaline solution [84], while the catalysts may slightly influence the peak intensity or shape of the CV curves. These redox features are an indicative of the pseudocapacitive behavior that may affect both OER and HER activities. Here, it is noteworthy that the current density increased with an increasing scan rate. This means that the active catalyst material possesses a low diffusion resistance. Compared to the bare MoS2 catalyst, the gC3N4–MoS2 hybrid catalyst exhibited a wider CV window with a greater current response. This implies that the 2D–2D gC3N4–MoS2 catalyst possessed a higher number of active sites than did the MoS2 catalyst. We believe that the enhanced electrochemical activity of the 2D–2D gC3N4–MoS2 hybrid catalyst is due to two possible reasons: the increased number of active sites [67,85,86] and the increased electrical conductivity [14,35,36,60,87]. The former will be explained below, while the latter is discussed later in the EIS section.
Figure 4. (a) Mo 3d and (b) S 2p core-level spectra of MoS2 nanosheets. (c) Mo 3d, (d) S 2p, (e) C 1s, and (f) N 1s core-level spectra of gC3N4–MoS2 nanocomposites.
Figure 4. (a) Mo 3d and (b) S 2p core-level spectra of MoS2 nanosheets. (c) Mo 3d, (d) S 2p, (e) C 1s, and (f) N 1s core-level spectra of gC3N4–MoS2 nanocomposites.
Materials 18 03775 g004
Figure 5. CV curves of (a) MoS2 and (b) gC3N4–MoS2. Non-faradaic JDL at 1.10 V as a function of the potential scan rate for (c) MoS2 and (d) gC3N4–MoS2. Nyquist plots of (e) MoS2 and (f) gC3N4–MoS2. The insets of (e,f) illustrate the equivalent circuits of the fabricated working electrodes.
Figure 5. CV curves of (a) MoS2 and (b) gC3N4–MoS2. Non-faradaic JDL at 1.10 V as a function of the potential scan rate for (c) MoS2 and (d) gC3N4–MoS2. Nyquist plots of (e) MoS2 and (f) gC3N4–MoS2. The insets of (e,f) illustrate the equivalent circuits of the fabricated working electrodes.
Materials 18 03775 g005
To elucidate the improved electrocatalytic performance of the gC3N4–MoS2 hybrid catalyst, we first calculated the ECSA of the catalyst materials using Equations (1) and (2). The ECSA values were determined from the non-faradic CV regions (1.05 to 1.15 V) for both MoS2 and gC3N4–MoS2 (see Figure S2a,b). From the JDL vs. v curves obtained from the non-faradaic CV region at approximately 1.10 V (Figure 5c,d), the CDL values were calculated to be 1.33 and 6.29 mF/cm2 for MoS2 and gC3N4–MoS2, respectively. Using Equations (1) and (2), the corresponding ECSA values of 33 and 158 cm2 were estimated for MoS2 and gC3N4–MoS2, respectively. This indicates that the gC3N4–MoS2 hybrid catalyst displayed a larger ECSA than did the bare MoS2 catalyst (see Figure S3). Hence, it can be conjectured that the hybridization of gC3N4 and MoS2 increased the number of electrochemically active sites in the entire composite medium of gC3N4–MoS2. In short, hybridizing MoS2 with gC3N4 enhanced the specific surface area compared to that of bare MoS2. gC3N4 helps prevent the natural tendency of MoS2 to restack by acting as a spacer, leading to a more open and porous structure [34,88]. This exposes more active edge sites, which are crucial for water splitting. The nitrogen-rich surface of gC3N4 also provides anchoring points for MoS2, ensuring strong interaction and uniform dispersion [89,90]. In addition, electronic coupling between the two facilitates efficient charge transfer across the interface, while the rough, crumpled morphology created during assembly further increases surface roughness and accessibility [91,92,93].
Next, the electrochemical resistive behavior of the fabricated catalysts was examined through the EIS measurements. As shown in Figure 5e,f, the Nyquist plots of both MoS2 and gC3N4–MoS2 revealed straight lines at the low-frequency region, while they exhibited no parabolic curves at the high-frequency region. The former is relevant to the charge-transfer characteristics, which are directly associated with the series resistance (Rs) of the working electrode [34,94,95,96], and the latter is attributed to the electrolyte dispersion characteristics [36,67,97]. Through fitting the EIS data to the equivalent circuit model (Figure 5e,f, insets), the Rs values of MoS2 and gC3N4–MoS2 were estimated to be 0.71 and 0.48 Ω, respectively. Thus, it can be inferred that gC3N4–MoS2 possess a better charge-transfer characteristic than does MoS2. This can be interpreted as resulting from the large porosity and high electrical conductivity of the incorporated gC3N4 nanosheets.
The increased ECSA and the decreased Rs help to enhance rapid ion diffusion and swift electron transport, resulting in improved OER/HER performance. To verify this hypothesis, we performed LSV measurements for MoS2, gC3N4, and gC3N4–MoS2. Figure 6a shows the iR-corrected OER LSV curves of the bare MoS2 and gC3N4–MoS2 hybrid catalysts at 1 mV/s (see also Figure S4a for bare gC3N4). Using Equations (3) and (4), the η10 values of MoS2, gC3N4, and gC3N4–MoS2 were calculated to be 297, 325, and 225 mV, respectively, from the measured LSV curves at J = 10 mA/cm2. The gC3N4–MoS2 hybrid catalyst exhibited lower overpotential values (η50 = 254 mV and η100 = 271 mV), even at high current density (J = 50 and 100 mA/cm2), compared to those for MoS2 (η50 = 339 mV and η100 = 378 mV) and gC3N4 (η50 = 358 mV and η100 = 381 mV) (Figure 6b). Notably, the η10 values determined for MoS2 and gC3N4–MoS2 are in line with, and occasionally better than, established data from earlier studies (see Table S1). This means that the 2D–2D gC3N4–MoS2 catalyst exhibits outstanding intrinsic reaction kinetics, resulting in significant OER performance in alkaline electrolytes [98,99,100]. Improved OER performance can also be confirmed by determining the ST value. Using the Tafel equation in Equation (6), the small ST values of MoS2 (55 mV/dec), gC3N4 (58 mV/dec), and gC3N4–MoS2 (49 mV/dec) were determined from their Tafel curves (Figure 6c,d and Figure S4b). The achieved ST values suggest that the MoS2, gC3N4, and gC3N4–MoS2 catalysts follow the combined Volmer–Heyrovsky mechanism. In particular, gC3N4–MoS2 exhibited a smaller ST value than the values in the literature (Table S1). Namely, the 2D–2D gC3N4–MoS2 catalyst exhibited outstanding intrinsic reaction kinetics owing to its decreased charge-transfer resistance, large porosity, and increased active surface area. Compared to MoS2, the gC3N4–MoS2 catalyst exhibited lower η and smaller ST values, signifying the faster reaction rate of OH over the catalyst surface. The catalytic OER mechanism in an alkaline medium typically involves a four-electron process:
M + O H M O H a d s + e
M O H a d s + O H M O a d s + e + H 2 O
M O a d s + O H M O O H a d s + e
M O O H a d s + O H M + e + O 2 + H 2 O
where M specifies an active site of the prepared catalysts, and M-OOH, M-O, and M-OH are the reaction intermediates on the catalysts’ surface.
The improved intrinsic reaction kinetics of the catalyst can also impact its chronopotentiometric characteristics. As shown in Figure 6e, the gC3N4–MoS2 catalyst displayed a smaller overpotential at each step (i.e., at different current densities) compared to that of the bare MoS2 catalyst. This proves that the incorporation of gC3N4 could aid in enhancing ion storage performance and catalytic activity. The long-term OER stability of the MoS2 and gC3N4–MoS2 catalysts was systematically evaluated through CP measurements conducted at 10 and 100 mA/cm2 for each 100 h, respectively. As shown in Figure 6f, both catalysts maintained stable performance and increased activity at diverse current densities. Furthermore, both samples exhibited almost indistinguishable LSV characteristic curves before and after the 100 h stability evaluation. (Figure S6). Compared to MoS2, however, gC3N4–MoS2 demonstrated superior long-term durability due to its trivial resistance and increased catalytic active areas. This indicates the excellent endurance of gC3N4–MoS2 in an alkaline medium.
Figure 6. OER performance of fabricated electrocatalysts. (a) iR-corrected LSV curves, (b) overpotential comparison, (c) Tafel plots, (d) Tafel slope comparison, (e) chronopotentiometric profiles at 10–100 mA/cm2, and (f) long-term stability characteristics of MoS2 and gC3N4–MoS2.
Figure 6. OER performance of fabricated electrocatalysts. (a) iR-corrected LSV curves, (b) overpotential comparison, (c) Tafel plots, (d) Tafel slope comparison, (e) chronopotentiometric profiles at 10–100 mA/cm2, and (f) long-term stability characteristics of MoS2 and gC3N4–MoS2.
Materials 18 03775 g006
Next, we assessed the HER performance of the fabricated catalysts to examine their bifunctional water-splitting activity. Figure 7a shows the iR-corrected LSV curves of gC3N4 and gC3N4–MoS2 obtained during the HER process, which was conducted at a 1 mV/s scan rate in the KOH electrolyte (see also Figure S5a for bare gC3N4). Using Equations (3) and (5), the obtained η10 values of MoS2, gC3N4, and gC3N4–MoS2 were 228, 236, and 156 mV, respectively, at a current density of −10 mA/cm2. In addition, the gC3N4–MoS2 catalyst revealed smaller η50 (=244 mV) and η100 (=275 mV) values compared to those for the MoS2 catalyst (η50 = 327 mV and η100 = 374 mV) and gC3N4 catalyst (η50 = 321 mV and η100 = 378 mV) (Figure 7b). Furthermore, from the Tafel curves (Figure 7c,d and Figure S5b), the obtained ST values for MoS2, gC3N4, and gC3N4–MoS2 were 145, 158, and 101 mV/dec, respectively. Notably, both η and ST values obtained from gC3N4–MoS2 compare favorably with and, in certain cases, surpass those reported in prior literature. (Table S2). In both the chronopotentiometric HER test (Figure 7e) and the long-term HER stability test (Figure 7f and Figure S7), gC3N4–MoS2 also revealed better electrocatalytic activity than that of bare MoS2.
The outstanding bifunctional OER and HER performances of the 2D–2D gC3N4–MoS2 catalyst may allow for superior overall water-splitting performance. To verify this, we tested the overall water-splitting activities of gC3N4–MoS2 using a two-electrode configuration in 1M KOH (Figure 8a). Figure 8b displays the LSV curve of the gC3N4–MoS2|| gC3N4–MoS2 electrolyzer. Remarkably, the bifunctional gC3N4–MoS2||gC3N4–MoS2 electrolyzer could drive specific current densities with low full-cell voltage values (e.g., 10 mA/cm2 with 1.52 V and 100 mA/cm2 with 1.85 V). These results depict an adequate catalytic activity of gC3N4–MoS2 for efficient overall water electrolysis. The full-cell voltages achieved from the current gC3N4–MoS2||gC3N4–MoS2 electrolyzer are on par with or even lower than those previously reported for other metal oxide-based electrocatalysts (Table S3). Moreover, gC3N4–MoS2||gC3N4–MoS2 exhibited constant stability performance for 100 h at both 10 and 100 mA/cm2 (Figure 8c). After overall water electrolysis stability tests, the morphology of both MoS2 and gC3N4–MoS2 maintained its initial structure (see Figure S8a,b). In EDX analysis, it was confirmed that the intrinsic elements Mo, S, C, and N clearly appear, consistent with the composition of the synthesized catalysts. However, the detection of additional Ni, K, and O signals suggests interfacial interactions between the catalyst surface, the electrolyte, and the underlying Ni foam substrate during electrochemical testing (see Figure S8c,d). In XRD testing, the gC3N4–MoS2 still exhibited higher intensity compared to that of the bare MoS2, although both samples showed decreased XRD intensities after the completion of all the water electrolysis tests (see Figure S9a,b). The above results suggest that the sonochemically hybridized 2D–2D gC3N4–MoS2 nanocomposites exhibit excellent potential for use as a bifunctional electrocatalyst for superior overall water electrolysis.
Figure 7. HER performance of fabricated electrocatalysts. (a) iR-corrected LSV curves, (b) overpotential comparison, (c) Tafel plots, (d) Tafel slope comparison, (e) chronopotentiometric profiles at −10–−100 mA/cm2, and (f) long-term stability characteristics of MoS2 and gC3N4–MoS2.
Figure 7. HER performance of fabricated electrocatalysts. (a) iR-corrected LSV curves, (b) overpotential comparison, (c) Tafel plots, (d) Tafel slope comparison, (e) chronopotentiometric profiles at −10–−100 mA/cm2, and (f) long-term stability characteristics of MoS2 and gC3N4–MoS2.
Materials 18 03775 g007
Figure 8. Overall water-splitting performance of gC3N4–MoS2||gC3N4–MoS2: (a) illustration of the two-electrode cell setup for OWS measurements, (b) LSV curves before and after stability testing, and (c) long-term stability characteristics over 100 h under current densities of 10 and 100 mA/cm2.
Figure 8. Overall water-splitting performance of gC3N4–MoS2||gC3N4–MoS2: (a) illustration of the two-electrode cell setup for OWS measurements, (b) LSV curves before and after stability testing, and (c) long-term stability characteristics over 100 h under current densities of 10 and 100 mA/cm2.
Materials 18 03775 g008

4. Conclusions

High-performance 2D–2D gC3N4–MoS2 nanocomposites were successfully fabricated via a facile ultrasonication process using sonochemically synthesized MoS2 and pyrolytically derived gC3N4. The gC3N4–MoS2 nanocomposites exhibited excellent bifunctional OER and HER performance in 1 M KOH. Namely, the gC3N4–MoS2 catalyst demonstrated excellent OER performance (i.e., a low η of 225 mV and a small ST of 49 mV/dec), as well as outstanding HER performance (i.e., a low η of 156 mV and a small ST of 101 mV/dec). Moreover, the gC3N4–MoS2 electrocatalyst revealed the superb overall water splitting with a low full-cell voltage of 1.52 V at 10 mA/cm2. These results were attributed to both low Rs (0.48 Ω) and large ECSA (158 cm2), resulting from the hybridization of 2D MoS2 and highly conductive gC3N4. Additionally, gC3N4–MoS2 also demonstrated good long-term stability up to 100 h for overall water splitting. This work marks a significant improvement over previously reported MoS2-based catalysts by introducing a scalable, energy-efficient synthesis strategy that simultaneously improves catalytic activity and long-term durability. The adoption of a sonochemical approach offers a green, cost-effective, and industrially viable pathway for the fabrication of high-performance bifunctional electrocatalysts. Future research could explore compositional optimization, integration into membrane-based electrolyzer systems, and rigorous evaluation under industrially relevant conditions to facilitate the practical deployment of this catalyst in sustainable hydrogen production technologies.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma18163775/s1, Figure S1: Full-survey XPS spectra of (a) MoS2 nanosheets and (b) gC3N4–MoS2 nanocomposites. O 1s core-level spectra of (c) MoS2 nanosheets and (d) gC3N4–MoS2 nanocomposites; Figure S2: Non-faradic CV curves of (a) MoS2 and (b) gC3N4–MoS2 catalysts; Figure S3: (a) iR-corrected OER LSV curve and (b) Tafel plot of the gC3N4 catalyst. Figure S4: (a) iR-corrected HER LSV curve and (b) Tafel plot of the gC3N4 catalyst. Figure S5: ECSA of the MoS2 and gC3N4–MoS2 catalysts; Figure S6: LSV curves of MoS2 and gC3N4–MoS2 before and after the OER stability test; Figure S7: LSV curves of MoS2 and gC3N4–MoS2 before and after the HER stability test; Figure S8: FE-SEM images of (a) MoS2 and (b) gC3N4–MoS2 after the stability test. EDX spectra of (c) MoS2 and (d) gC3N4–MoS2 after the stability test; Figure S9: XRD pattern of (a) MoS2 and (b) gC3N4–MoS2 after the stability test.; Table S1: Comparison of OER performance for MoS2, gC3N4, and gC3N4–MoS2 with that of previously reported electrocatalysts; Table S2: Comparison of HER performance for MoS2, gC3N4, and gC3N4–MoS2 with that of previously reported electrocatalysts; Table S3: Comparison of overall water splitting performance for MoS2, gC3N4, and gC3N4–MoS2 with that of previously reported electrocatalysts. References [101,102,103,104,105,106,107,108,109,110,111,112,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131] have been cited in Supplementary file.

Author Contributions

S.S.: methodology, formal analysis, investigation, and writing—original draft. A.S.: methodology and formal analysis. Y.L.: data curation, validation, and supervision. S.L.: conceptualization, supervision, and writing—review and editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the National Research Foundation (NRF) of Korea through the basic science research program (RS-2023-NR076644) funded by the Korean Government.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Xiong, Q.; Wang, Y.; Liu, P.-F.; Zheng, L.-R.; Wang, G.; Yang, H.-G.; Wong, P.-K.; Zhang, H.; Zhao, H. Cobalt Covalent Doping in MoS2 to Induce Bifunctionality of Overall Water Splitting. Adv. Mater. 2018, 30, 1801450. [Google Scholar] [CrossRef] [PubMed]
  2. Pan, A.; Xu, S.; Zaidi, S.A.H. Environmental impact of energy imports: Natural resources income and natural gas production profitability in the Asia-Pacific Economic Cooperation Countries. Geosci. Front. 2024, 15, 101756. [Google Scholar] [CrossRef]
  3. Manikandan, R.; Sadhasivam, S.; Lee, S.; Seung-Cheol, C.; Ashok Kumar, K.; Bathula, C.; Gopalan Sree, V.; Young Kim, D.; Sekar, S. Deep eutectic solvents assisted synthesis of AC-decorated NiO nanocomposites for hydrogen evolution reaction. J. Mol. Liq. 2023, 375, 121338. [Google Scholar] [CrossRef]
  4. Li, Z.; Liu, X.; Yu, Q.; Qu, X.; Wan, J.; Xiao, Z.; Chi, J.; Wang, L. Recent advances in design of hydrogen evolution reaction electrocatalysts at high current density: A review. Chin. J. Catal. 2024, 63, 33–60. [Google Scholar] [CrossRef]
  5. Han, N.; Yang, K.R.; Lu, Z.; Li, Y.; Xu, W.; Gao, T.; Cai, Z.; Zhang, Y.; Batista, V.S.; Liu, W.; et al. Nitrogen-doped tungsten carbide nanoarray as an efficient bifunctional electrocatalyst for water splitting in acid. Nat. Commun. 2018, 9, 924. [Google Scholar] [CrossRef]
  6. 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]
  7. Ma, N.; Zhao, W.; Wang, W.; Li, X.; Zhou, H. Large scale of green hydrogen storage: Opportunities and challenges. Int. J. Hydrogen Energy 2024, 50, 379–396. [Google Scholar] [CrossRef]
  8. Ceballos-Alvarez, C.; Maziar, J.; Moradizadeh, L.; Siaj, M.; Shahgaldi, S.; Izquierdo, R. Enhanced Graphene Oxide-Nafion® membranes for proton exchange membrane water electrolysis. J. Membr. Sci 2025, 734, 124267. [Google Scholar] [CrossRef]
  9. Sun, H.; Yan, Z.; Liu, F.; Xu, W.; Cheng, F.; Chen, J. Self-Supported Transition-Metal-Based Electrocatalysts for Hydrogen and Oxygen Evolution. Adv. Mater. 2020, 32, 1806326. [Google Scholar] [CrossRef] [PubMed]
  10. Xu, Y.; Cheng, J.; Ding, L.; Lv, H.; Zhang, K.; Hu, A.; Yang, X.; Sun, W.; Mao, Y. Interfacial engineering for promoting charge transfer in MoS2/CoFeLDH heterostructure electrodes for overall water splitting. Int. J. Hydrogen Energy 2024, 49, 897–906. [Google Scholar] [CrossRef]
  11. Sekar, S.; Lee, E.; Yun, J.; Arumugasamy, S.K.; Choi, M.-J.; Lee, Y.; Lee, S. High-performance electrocatalyst of activated carbon-decorated molybdenum trioxide nanocomposites for effective production of H2 and H2O2. Sep. Purif. Technol. 2025, 361, 131614. [Google Scholar] [CrossRef]
  12. Sekar, S.; Sadhasivam, S.; Lee, Y.; Lee, S. Enhanced bifunctional water electrolysis activities of activated carbon-decorated trimetallic Ni-Al-La nanocomposites. Appl. Surf. Sci. 2025, 698, 163098. [Google Scholar] [CrossRef]
  13. Slobodkin, I.; Davydova, E.; Sananis, M.; Breytus, A.; Rothschild, A. Electrochemical and chemical cycle for high-efficiency decoupled water splitting in a near-neutral electrolyte. Nat. Mater. 2024, 23, 398–405. [Google Scholar] [CrossRef]
  14. Sekar, S.; Park, S.; Jung, J.; Lee, S. Superb Bifunctional Water Electrolysis Activities of Carbon Nanotube-Decorated Lanthanum Hydroxide Nanocomposites. Int. J. Energy Res. 2023, 2023, 6685726. [Google Scholar] [CrossRef]
  15. Rong, C.; Dastafkan, K.; Wang, Y.; Zhao, C. Breaking the Activity and Stability Bottlenecks of Electrocatalysts for Oxygen Evolution Reactions in Acids. Adv. Mater. 2023, 35, 2211884. [Google Scholar] [CrossRef]
  16. Xi, C.; Xu, W.; Zhou, S.; Wang, Y.; Han, S.; Jiang, J. Heterogeneous interface construction of P-doped MoS2 based on the N-doped graphene oxide aerogels for efficient hydrogen evolution. Int. J. Hydrogen Energy 2024, 58, 1596–1607. [Google Scholar] [CrossRef]
  17. Sankar, B.D.; Sekar, S.; Vignesh, V.; Ruan, J.; Nirmala, R.; Lee, Y.; Lee, S.; Tsai, P.-C.; Chen, S.-C.; Lin, Y.-C.; et al. A dual-purpose binder-free FeNiS2-Decorated Ti3C2Tx nanocomposite for supercapacitor and catalytic hydrogen evolution reaction. J. Power Sources 2025, 649, 237412. [Google Scholar] [CrossRef]
  18. Sekar, S.; Shanmugam, A.; Senthilkumar, G.; Thangasami, K.; Jung, H.; Lee, Y.; Lee, S. Enhanced Hydrogen Evolution Reaction Using Biomass-Activated Carbon Nanosheets Derived from Eucalyptus Leaves. Materials 2025, 18, 670. [Google Scholar] [CrossRef] [PubMed]
  19. Ray, C.; Lee, S.C.; Sankar, K.V.; Jin, B.; Lee, J.; Park, J.H.; Jun, S.C. Amorphous Phosphorus-Incorporated Cobalt Molybdenum Sulfide on Carbon Cloth: An Efficient and Stable Electrocatalyst for Enhanced Overall Water Splitting over Entire pH Values. ACS Appl. Mater. Interfaces 2017, 9, 37739–37749. [Google Scholar] [CrossRef] [PubMed]
  20. Prabhu, P.; Jose, V.; Lee, J.-M. Design Strategies for Development of TMD-Based Heterostructures in Electrochemical Energy Systems. Matter 2020, 2, 526–553. [Google Scholar] [CrossRef]
  21. Chhowalla, M.; Shin, H.S.; Eda, G.; Li, L.-J.; Loh, K.P.; Zhang, H. The chemistry of two-dimensional layered transition metal dichalcogenide nanosheets. Nat. Chem. 2013, 5, 263–275. [Google Scholar] [CrossRef] [PubMed]
  22. Sekar, S.; Manikandan, R.; Arumugasamy, S.K.; Sekar, S.; Lee, Y.; Chang, S.-C.; Lee, S. Rapid Microwave-Assisted Synthesis of CuSe Nanoparticles for High-Sensitivity Serotonin Biosensing in Serum. Chemosensors 2025, 13, 264. [Google Scholar] [CrossRef]
  23. Huang, X.; Xu, H.; Cao, D.; Cheng, D. Interface construction of P-Substituted MoS2 as efficient and robust electrocatalyst for alkaline hydrogen evolution reaction. Nano Energy 2020, 78, 105253. [Google Scholar] [CrossRef]
  24. Chen, W.; Gu, J.; Du, Y.; Song, F.; Bu, F.; Li, J.; Yuan, Y.; Luo, R.; Liu, Q.; Zhang, D. Achieving Rich and Active Alkaline Hydrogen Evolution Heterostructures via Interface Engineering on 2D 1T-MoS2 Quantum Sheets. Adv. Funct. Mater. 2020, 30, 2000551. [Google Scholar] [CrossRef]
  25. Ren, X.; Yang, F.; Chen, R.; Ren, P.; Wang, Y. Improvement of HER activity for MoS2: Insight into the effect and mechanism of phosphorus post-doping. New J. Chem. 2020, 44, 1493–1499. [Google Scholar] [CrossRef]
  26. Nayana, K.; Sunitha, A.P. MoS2−x/GCD-MoS2−x nanostructures for tuning the overpotential of Volmer-Heyrovsky reaction of electrocatalytic hydrogen evolution. Int. J. Hydrogen Energy 2024, 55, 422–431. [Google Scholar] [CrossRef]
  27. Sheelam, A.; Bell, J.G. Uniform electrodeposition of Pd nanoparticles on MoS2-x nanosheets using magnetic fields for improved mass utilization in hydrogen evolution. Int. J. Hydrogen Energy 2024, 56, 348–357. [Google Scholar] [CrossRef]
  28. Thomas, S.A.; Roy, N.; Sharaf Saeed, W.; Sreedhar, A.; Cherusseri, J. Hydrothermally synthesized MoS2/ZnS heterostructure with efficient catalytic performance for hydrogen evolution. Int. J. Hydrogen Energy 2023, in press. [Google Scholar] [CrossRef]
  29. Sathiyan, K.; Mondal, T.; Mukherjee, P.; Patra, S.G.; Pitussi, I.; Kornweitz, H.; Bar-Ziv, R.; Zidki, T. Enhancing the catalytic OER performance of MoS2 via Fe and Co doping. Nanoscale 2022, 14, 16148–16155. [Google Scholar] [CrossRef] [PubMed]
  30. Long, A.; Li, W.; Zhou, M.; Gao, W.; Liu, B.; Wei, J.; Zhang, X.; Liu, H.; Liu, Y.; Zeng, X. MoS2 nanosheets grown on nickel chalcogenides: Controllable synthesis and electrocatalytic origins for the hydrogen evolution reaction in alkaline solution. J. Mater. Chem. A 2019, 7, 21514–21522. [Google Scholar] [CrossRef]
  31. Hu, J.; Zhang, C.; Zhang, Y.; Yang, B.; Qi, Q.; Sun, M.; Zi, F.; Leung, M.K.H.; Huang, B. Interface Modulation of MoS2/Metal Oxide Heterostructures for Efficient Hydrogen Evolution Electrocatalysis. Small 2020, 16, 2002212. [Google Scholar] [CrossRef]
  32. Zhong, Y.; Wang, S.; Zhang, L. In situ electrochemical metal (Co, Ni) oxide deposition on MoS2 nanosheets for highly efficient electrocatalytic water splitting. New J. Chem. 2023, 47, 4430–4438. [Google Scholar] [CrossRef]
  33. Guo, Y.; Tang, J.; Henzie, J.; Jiang, B.; Xia, W.; Chen, T.; Bando, Y.; Kang, Y.-M.; Hossain, M.S.A.; Sugahara, Y.; et al. Mesoporous Iron-doped MoS2/CoMo2S4 Heterostructures through Organic–Metal Cooperative Interactions on Spherical Micelles for Electrochemical Water Splitting. ACS Nano 2020, 14, 4141–4152. [Google Scholar] [CrossRef]
  34. Fageria, P.; Sudharshan, K.Y.; Nazir, R.; Basu, M.; Pande, S. Decoration of MoS2 on g-C3N4 surface for efficient hydrogen evolution reaction. Electrochim. Acta 2017, 258, 1273–1283. [Google Scholar] [CrossRef]
  35. Zhang, X.; Ding, P.; Sun, Y.; Wang, Y.; Li, J.; Guo, J. MoS2 nanosheets on C3N4 realizing improved electrochemical hydrogen evolution. Mater. Lett. 2017, 197, 41–44. [Google Scholar] [CrossRef]
  36. Khan, M.S.; Noor, T.; Pervaiz, E.; Iqbal, N.; Zaman, N. Fabrication of MoS2/rGO hybrids as electrocatalyst for water splitting applications. RSC Adv. 2024, 14, 12742–12753. [Google Scholar] [CrossRef] [PubMed]
  37. Hou, Y.; Zhou, C.; Bai, S.; Yang, S.; Yang, F.; Yang, B.; Zeng, X. Multi-interface and rich-defects 1T phase MoS2 combine with FeS anchored on reduced graphene oxide for efficient alkaline hydrogen evolution. Int. J. Hydrogen Energy 2024, 90, 191–198. [Google Scholar] [CrossRef]
  38. Sankar, S.; Saravanan, S.; Ahmed, A.T.A.; Inamdar, A.I.; Im, H.; Lee, S.; Kim, D.Y. Spherical activated-carbon nanoparticles derived from biomass green tea wastes for anode material of lithium-ion battery. Mater. Lett. 2019, 240, 189–192. [Google Scholar] [CrossRef]
  39. Ong, W.-J.; Tan, L.-L.; Ng, Y.H.; Yong, S.-T.; Chai, S.-P. Graphitic Carbon Nitride (g-C3N4)-Based Photocatalysts for Artificial Photosynthesis and Environmental Remediation: Are We a Step Closer to Achieving Sustainability? Chem. Rev. 2016, 116, 7159–7329. [Google Scholar] [CrossRef]
  40. Pan, F.; Khan, M.; Lei, T.; Kamli, M.R.; M Sabir, J.S.; Khan, I.; Ansari, M.Z. Visible light driven hydrogen generation and pollutant degradation with Au loaded 2D/2D heterojunctional nanocomposite of MoS2 and g-C3N4. Int. J. Hydrogen Energy 2024, 51, 1141–1153. [Google Scholar] [CrossRef]
  41. Azhar, A.; Basit, M.A.; Mehmood, W.; Ali, M.A.; Zahid, S.; Ahmad, M.; Zaidi, S.J.A.; Park, T.J. Synchronized wet-chemical development of 2-dimensional MoS2 and g-C3N4/MoS2 QDs nanocomposite as efficient photocatalysts for detoxification of aqueous dye solutions. Colloids Surf. A Physicochem. Eng. Asp. 2023, 657, 130581. [Google Scholar] [CrossRef]
  42. Liu, Y.; Xu, X.; Zhang, J.; Zhang, H.; Tian, W.; Li, X.; Tade, M.O.; Sun, H.; Wang, S. Flower-like MoS2 on graphitic carbon nitride for enhanced photocatalytic and electrochemical hydrogen evolutions. Appl. Catal. B Environ. 2018, 239, 334–344. [Google Scholar] [CrossRef]
  43. He, L.; Cui, B.; Liu, J.; Wang, M.; Zhang, Z.; Zhang, H. Fabrication of Porous CoOx/mC@MoS2 Composite Loaded on g-C3N4 Nanosheets as a Highly Efficient Dual Electrocatalyst for Oxygen Reduction and Hydrogen Evolution Reactions. ACS Sustain. Chem. Eng. 2018, 6, 9257–9268. [Google Scholar] [CrossRef]
  44. Mehtab, A.; Ali, S.A.; Ingole, P.P.; Mao, Y.; Alshehri, S.M.; Ahmad, T. MoS2 Nanoflower-Deposited g-C3N4 Nanosheet 2D/2D Heterojunction for Efficient Photo/Electrocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2023, 6, 12003–12012. [Google Scholar] [CrossRef]
  45. Devi, S.B.; Sekar, S.; Kowsuki, K.; Maiyalagan, T.; Preethi, V.; Nirmala, R.; Lee, S.; Navamathavan, R. Graphitic carbon nitride encapsulated sonochemically synthesized β-nickel hydroxide nanocomposites for electrocatalytic hydrogen generation. Int. J. Hydrogen Energy 2022, 47, 40349–40358. [Google Scholar] [CrossRef]
  46. Wang, L.; Hou, Y.; Xiao, S.; Bi, F.; Zhao, L.; Li, Y.; Zhang, X.; Gai, G.; Dong, X. One-step, high-yield synthesis of g-C3N4 nanosheets for enhanced visible light photocatalytic activity. RSC Adv. 2019, 9, 39304–39314. [Google Scholar] [CrossRef]
  47. Ul Hassan, H.M.; Tawab, S.A.; Khan, M.I.; Hussain, A.; Elsaeedy, H.I.; Almutairi, B.S.; Hassan, A.; Raza, A. Reduce the recombination rate by facile synthesis of MoS2/g-C3N4 heterostructures as a solar light responsive catalyst for organic dye degradation. Diam. Relat. Mater 2023, 140, 110420. [Google Scholar] [CrossRef]
  48. Cui, Z.; Wu, H.; Bai, K.; Chen, X.; Li, E.; Shen, Y.; Wang, M. Fabrication of a g-C3N4/MoS2 photocatalyst for enhanced RhB degradation. Phys. E Low-Dimens. Syst. Nanostructures 2022, 144, 115361. [Google Scholar] [CrossRef]
  49. Li, X.; Zhang, Y.; Wei, T.; Wang, C.; Wan, J.; Tang, Y.; Guo, M.; Ma, Y.; Yang, Y. Boosting the Photocatalytic Performance of g-C3N4 through MoS2 Nanotubes with the Cavity Enhancement Effect. Langmuir 2024, 40, 11160–11172. [Google Scholar] [CrossRef]
  50. Ryaboshapka, D.; Bargiela, P.; Piccolo, L.; Afanasiev, P. On the electronic properties and catalytic activity of MoS2–C3N4 materials prepared by one-pot reaction. Int. J. Hydrogen Energy 2022, 47, 34012–34024. [Google Scholar] [CrossRef]
  51. Singh, J.; Akhtar, S.; Tran, T.T.; Kim, J. MoS2 nanoflowers functionalized with C3N4 nanosheets for enhanced photodecomposition. J. Alloys Compd. 2023, 954, 170206. [Google Scholar] [CrossRef]
  52. Lan, Z.; Yu, Y.; Yao, J.; Cao, Y. The band structure and photocatalytic mechanism of MoS2-modified C3N4 photocatalysts with improved visible photocatalytic activity. Mater. Res. Bull. 2018, 102, 433–439. [Google Scholar] [CrossRef]
  53. Sekar, S.; Sadhasivam, S.; Shanmugam, A.; Saravanan, S.; Pugazhendi, I.; Lee, Y.; Kim, D.Y.; Manikandan, R.; Chang, S.-C.; Lee, S. Enhanced bifunctional water electrolysis performance of spherical ZnMn2O4 nanoparticles. Int. J. Hydrogen Energy 2025, 141, 721–728. [Google Scholar] [CrossRef]
  54. Ahmed, A.T.A.; Ansari, A.S.; Sree, V.G.; Jana, A.; Meena, A.; Sekar, S.; Cho, S.; Kim, H.; Im, H. Nitrogen-Doped CuO@CuS Core–Shell Structure for Highly Efficient Catalytic OER Application. Nanomaterials 2023, 13, 3160. [Google Scholar] [CrossRef]
  55. Manikandan, R.; Sekar, S.; Mani, S.P.; Lee, S.; Kim, D.Y.; Saravanan, S. Bismuth tungstate-anchored PEDOT: PSS materials for high performance HER electrocatalyst. Int. J. Hydrogen Energy 2023, 48, 11746–11753. [Google Scholar] [CrossRef]
  56. Karmakar, A.; Durairaj, M.; Madhu, R.; De, A.; Dhandapani, H.N.; Spencer, M.J.S.; Kundu, S. From Proximity to Energetics: Unveiling the Hidden Compass of Hydrogen Evolution Reaction. ACS Mater. Lett. 2024, 6, 3050–3062. [Google Scholar] [CrossRef]
  57. Shinagawa, T.; Garcia-Esparza, A.T.; Takanabe, K. Insight on Tafel slopes from a microkinetic analysis of aqueous electrocatalysis for energy conversion. Sci. Rep. 2015, 5, 13801. [Google Scholar] [CrossRef]
  58. Pazhamalai, P.; Krishnamoorthy, K.; Natraj, V.; Mohan, V.; Chennakrishnan, J.; Kim, S.J. Unveiling the bi-functional electrocatalytic properties of rhenium di-sulfide nanostructures towards the development of high-rate alkaline water electrolyzer. Chem. Eng. J. 2024, 500, 156356. [Google Scholar] [CrossRef]
  59. Krishnamoorthy, K.; Pazhamalai, P.; Swaminathan, R.; Mohan, V.; Kim, S.-J. Unravelling the Bi-Functional Electrocatalytic Properties of {Mo72Fe30} Polyoxometalate Nanostructures for Overall Water Splitting Using Scanning Electrochemical Microscope and Electrochemical Gating Methods. Adv. Sci. 2024, 11, 2401073. [Google Scholar] [CrossRef] [PubMed]
  60. Sekar, S.; Ilanchezhiyan, P.; Ahmed, A.T.A.; Lee, Y.; Lee, S. Substantial Electrocatalytic Oxygen Evolution Performances of Activated Carbon-Decorated Vanadium Pentoxide Nanocomposites. Int. J. Energy Res. 2024, 2024, 9953038. [Google Scholar] [CrossRef]
  61. Dong, G.; Zhou, T.; Zhang, C.; Wang, Y.; Wang, Y.; Shi, W.; Su, X.; Zeng, T.; Chen, Y. Role of Charge Accumulation in MoS2/Graphitic Carbon Nitride Nanostructures for Photocatalytic N2 Fixation. ACS Appl. Nano Mater. 2024, 7, 15406–15415. [Google Scholar] [CrossRef]
  62. Mizukoshi, Y.; Oshima, R.; Maeda, Y.; Nagata, Y. Preparation of Platinum Nanoparticles by Sonochemical Reduction of the Pt (II) Ion. Langmuir 1999, 15, 2733–2737. [Google Scholar] [CrossRef]
  63. Bang, J.H.; Suslick, K.S. Applications of Ultrasound to the Synthesis of Nanostructured Materials. Adv. Mater. 2010, 22, 1039–1059. [Google Scholar] [CrossRef]
  64. Zhang, J.; Du, J.; Han, B.; Liu, Z.; Jiang, T.; Zhang, Z. Sonochemical Formation of Single-Crystalline Gold Nanobelts. Angew. Chem. Int. Ed. 2006, 45, 1116–1119. [Google Scholar] [CrossRef] [PubMed]
  65. Nemamcha, A.; Rehspringer, J.-L.; Khatmi, D. Synthesis of Palladium Nanoparticles by Sonochemical Reduction of Palladium (II) Nitrate in Aqueous Solution. J. Phys. Chem. B 2006, 110, 383–387. [Google Scholar] [CrossRef]
  66. Su, C.-H.; Wu, P.-L.; Yeh, C.-S. Sonochemical Synthesis of Well-Dispersed Gold Nanoparticles at the Ice Temperature. J. Phys. Chem. B 2003, 107, 14240–14243. [Google Scholar] [CrossRef]
  67. Sekar, S.; Kim, D.Y.; Lee, S. Excellent Oxygen Evolution Reaction of Activated Carbon-Anchored NiO Nanotablets Prepared by Green Routes. Nanomaterials 2020, 10, 1382. [Google Scholar] [CrossRef]
  68. Zuo, L.-X.; Jiang, L.-P.; Zhu, J.-J. A facile sonochemical route for the synthesis of MoS2/Pd composites for highly efficient oxygen reduction reaction. Ultrason. Sonochem. 2017, 35, 681–688. [Google Scholar] [CrossRef]
  69. Wang, W.; Kapitanova, O.O.; Ilanchezhiyan, P.; Xi, S.; Panin, G.N.; Fu, D.; Kang, T.W. Self-assembled MoS2/rGO nanocomposites with tunable UV-IR absorption. RSC Adv. 2018, 8, 2410–2417. [Google Scholar] [CrossRef] [PubMed]
  70. Li, Y.; Nakamura, R. Structural change of molybdenum sulfide facilitates the electrocatalytic hydrogen evolution reaction at neutral pH as revealed by in situ Raman spectroscopy. Chin. J. Catal. 2018, 39, 401–406. [Google Scholar] [CrossRef]
  71. Govindasamy, M.; Wang, S.-F.; Alothman, A.A.; Alshgari, R.A.; Ganesh, P.S. Synergetic effect of the ultrasonic-assisted hydrothermal process on the photocatalytic performance of MoS2 and WS2 nanoparticles. J. Mater. Sci. Mater. Electron. 2022, 33, 8858–8867. [Google Scholar] [CrossRef]
  72. Chen, K.; Feng, H.; Li, L.; Luo, M.; Xue, S.; Sang, J. Synthesis of dual Z-scheme flower-like spherical Ag3PO4/MoS2/g-C3N4 photocatalyst with high photocatalytic performance. J. Mater. Sci. Mater. Electron. 2022, 33, 16077–16098. [Google Scholar] [CrossRef]
  73. Elavarasan, N.; Palanisamy, G.; Kumar, P.S.; Venkatesh, G.; Vignesh, S.; Bhuvaneswari, K.; Rangasamy, G. Construction of a ternary g-C3N4/MoS2/MWCNTs nanocomposite for the enhanced photocatalytic performance against organic dye. Appl. Nanosci. 2023, 13, 5851–5863. [Google Scholar] [CrossRef]
  74. Baig, M.M.; Pervaiz, E.; Yang, M.; Gul, I.H. High-Performance Supercapacitor Electrode Obtained by Directly Bonding 2D Materials: Hierarchal MoS2 on Reduced Graphene Oxide. Front. Mater. 2020, 7, 580424. [Google Scholar] [CrossRef]
  75. Lu, X.; Lin, Y.; Dong, H.; Dai, W.; Chen, X.; Qu, X.; Zhang, X. One-Step Hydrothermal Fabrication of Three-dimensional MoS2 Nanoflower using Polypyrrole as Template for Efficient Hydrogen Evolution Reaction. Sci. Rep. 2017, 7, 42309. [Google Scholar] [CrossRef]
  76. Liu, Z.; Xu, J.; Li, Y.; Yu, H. Hydrothermal synthesis of Co2P-modified MoS2: A highly efficient non-precious metal catalyst of light. Ionics 2019, 25, 5003–5011. [Google Scholar] [CrossRef]
  77. Wei, Z.; Shen, X.; Ji, Y.; Yang, Z.; Wang, T.; Li, S.; Zhu, M.; Tian, Y. Synthesis of novel MoS2/g-C3N4 nanocomposites for enhanced photocatalytic activity. J. Mater. Sci. Mater. Electron. 2020, 31, 15885–15895. [Google Scholar] [CrossRef]
  78. Shi, Q.; Qin, Q.; Zhou, Y.; Wan, J.; Hu, Z. Preparation and high-efficient hydrogen evolution reaction of hydrangea-like MoS2 hollow microspheres modified by needle-like g-C3N4. J. Mater. Sci. Mater. Electron. 2019, 30, 16446–16451. [Google Scholar] [CrossRef]
  79. Yan, Z.; Zhao, J.; Gao, Q.; Lei, H. A 2H-MoS2/carbon cloth composite for high-performance all-solid-state supercapacitors derived from a molybdenum dithiocarbamate complex. Dalton Trans. 2021, 50, 11954–11964. [Google Scholar] [CrossRef]
  80. Li, N.; Zhou, J.; Sheng, Z.; Xiao, W. Molten salt-mediated formation of g-C3N4-MoS2 for visible-light-driven photocatalytic hydrogen evolution. Appl. Surf. Sci. 2018, 430, 218–224. [Google Scholar] [CrossRef]
  81. Archana, C.; Abinaya, R.; Mani, N.; Santhana Krishnan, H. Strain Effect in the Layered MoS2-rGO Heterostructure with Enhanced Performance for Flexible Thermoelectric Applications. ACS Appl. Energy Mater. 2025, 8, 1589–1597. [Google Scholar]
  82. Cao, Y.; Gao, Q.; Li, Q.; Jing, X.; Wang, S.; Wang, W. Synthesis of 3D porous MoS2/g-C3N4 heterojunction as a high efficiency photocatalyst for boosting H2 evolution activity. RSC Adv. 2017, 7, 40727–40733. [Google Scholar] [CrossRef]
  83. Ye, L.; Wang, D.; Chen, S. Fabrication and Enhanced Photoelectrochemical Performance of MoS2/S-Doped g-C3N4 Heterojunction Film. ACS Appl. Mater. Interfaces 2016, 8, 5280–5289. [Google Scholar] [CrossRef] [PubMed]
  84. Watson, N.I.; Keegan, M.; van den Bosch, B.; Yan, N.; Rothenberg, G. The Influence of Metal Impurities on NiOOH Electrocatalytic Activity in the Oxygen Evolution Reaction. ChemElectroChem 2024, 11, e202400223. [Google Scholar] [CrossRef]
  85. Dhanda, M.; Arora, R.; Yadav, M.; Pahuja, P.; Ahlawat, S.; Nehra, S.P.; Lata, S. Calibration of varying MoS2 amount with dual fabrication GCN/CQDs for supercapacitor’s electrode and strengthening of the energy parameters. Mater. Sci. Eng. B 2024, 300, 117055. [Google Scholar] [CrossRef]
  86. Wang, Y.; Williams, T.; Gengenbach, T.; Kong, B.; Zhao, D.; Wang, H.; Selomulya, C. Unique hybrid Ni2P/MoO2@MoS2 nanomaterials as bifunctional non-noble-metal electro-catalysts for water splitting. Nanoscale 2017, 9, 17349–17356. [Google Scholar] [CrossRef]
  87. Zhou, R.; Wei, S.; Liu, Y.; Gao, N.; Wang, G.; Lian, J.; Jiang, Q. Charge Storage by Electrochemical Reaction of Water Bilayers Absorbed on MoS2 Monolayers. Sci. Rep. 2019, 9, 3980. [Google Scholar] [CrossRef]
  88. Bai, J.; Lv, W.; Ni, Z.; Wang, Z.; Chen, G.; Xu, H.; Qin, H.; Zheng, Z.; Li, X. Integrating MoS2 on sulfur-doped porous g-C3N4 iostype heterojunction hybrids enhances visible-light photocatalytic performance. J. Alloys Compd. 2018, 768, 766–774. [Google Scholar] [CrossRef]
  89. Chen, L.; Maigbay, M.A.; Li, M.; Qiu, X. Synthesis and modification strategies of g-C3N4 nanosheets for photocatalytic applications. Adv. Powder Mater. 2024, 3, 100150. [Google Scholar] [CrossRef]
  90. Song, X.-L.; Chen, L.; Gao, L.-J.; Ren, J.-T.; Yuan, Z.-Y. Engineering g-C3N4 based materials for advanced photocatalysis: Recent advances. Green Energy Environ. 2024, 9, 166–197. [Google Scholar] [CrossRef]
  91. Jo, W.-K.; Lee, J.Y.; Selvam, N.C.S. Synthesis of MoS2 nanosheets loaded ZnO–g-C3N4 nanocomposites for enhanced photocatalytic applications. Chem. Eng. J. 2016, 289, 306–318. [Google Scholar] [CrossRef]
  92. Hou, Y.; Li, J.; Wen, Z.; Cui, S.; Yuan, C.; Chen, J. N-doped graphene/porous g-C3N4 nanosheets supported layered-MoS2 hybrid as robust anode materials for lithium-ion batteries. Nano Energy 2014, 8, 157–164. [Google Scholar] [CrossRef]
  93. Ling, G.Z.S.; Oh, V.B.-Y.; Haw, C.Y.; Tan, L.-L.; Ong, W.-J. g-C3N4 Photocatalysts: Utilizing Electron–Hole Pairs for Boosted Redox Capability in Water Splitting. Energy Mater. Adv. 2023, 4, 0038. [Google Scholar] [CrossRef]
  94. Ghanashyam, G.; Jeong, H.K. Size Effects of MoS2 on Hydrogen and Oxygen Evolution Reaction. J. Electrochem. Sci. Technol. 2022, 13, 120–127. [Google Scholar] [CrossRef]
  95. Sekar, S.; Sim, D.H.; Lee, S. Excellent Electrocatalytic Hydrogen Evolution Reaction Performances of Partially Graphitized Activated-Carbon Nanobundles Derived from Biomass Human Hair Wastes. Nanomaterials 2022, 12, 531. [Google Scholar] [CrossRef]
  96. Chen, B.; Hu, P.; Yang, F.; Hua, X.; Yang, F.F.; Zhu, F.; Sun, R.; Hao, K.; Wang, K.; Yin, Z. In Situ Porousized MoS2 Nano Islands Enhance HER/OER Bifunctional Electrocatalysis. Small 2023, 19, 2207177. [Google Scholar] [CrossRef]
  97. Sadighi, Z.; Liu, J.; Zhao, L.; Ciucci, F.; Kim, J.-K. Metallic MoS2 nanosheets: Multifunctional electrocatalyst for the ORR, OER and Li–O2 batteries. Nanoscale 2018, 10, 22549–22559. [Google Scholar] [CrossRef]
  98. Sekar, S.; Yun, J.-S.; Park, S.; Kim, D.Y.; Lee, Y.; Lee, S. Excellent Bifunctional Water Electrolysis Activities of α-MoO3/AC Nanocomposites. Int. J. Energy Res. 2024, 2024, 3167699. [Google Scholar] [CrossRef]
  99. Ahmed, I.; Biswas, R.; Patil, R.A.; Halder, K.K.; Singh, H.; Banerjee, B.; Kumar, B.; Ma, Y.-R.; Haldar, K.K. Graphitic Carbon Nitride Composites with MoO3-Decorated Co3O4 Nanorods as Catalysts for Oxygen and Hydrogen Evolution. ACS Appl. Nano Mater. 2021, 4, 12672–12681. [Google Scholar] [CrossRef]
  100. Xue, Z.; Zhang, X.; Qin, J.; Liu, R. Constructing MoS2/g-C3N4 heterojunction with enhanced oxygen evolution reaction activity: A theoretical insight. Appl. Surf. Sci. 2020, 510, 145489. [Google Scholar] [CrossRef]
  101. Cheng, P.; Yuan, C.; Zhou, Q.; Hu, X.; Li, J.; Lin, X.; Wang, X.; Jin, M.; Shui, L.; Gao, X.; et al. Core–Shell MoS2@CoO Electrocatalyst for Water Splitting in Neural and Alkaline Solutions. J. Phys. Chem. C 2019, 123, 5833–5839. [Google Scholar] [CrossRef]
  102. Li, Y.; Xu, H.; Huang, H.; Wang, C.; Gao, L.; Ma, T. One-dimensional MoO2–Co2Mo3O8@C nanorods: A novel and highly efficient oxygen evolution reaction catalyst derived from metal–organic framework composites. Chem. Commun. 2018, 54, 2739–2742. [Google Scholar] [CrossRef] [PubMed]
  103. Tang, B.; Yu, Z.G.; Seng, H.L.; Zhang, N.; Liu, X.; Zhang, Y.-W.; Yang, W.; Gong, H. Simultaneous edge and electronic control of MoS2 nanosheets through Fe doping for an efficient oxygen evolution reaction. Nanoscale 2018, 10, 20113–20119. [Google Scholar] [CrossRef]
  104. Kwon, I.S.; Debela, T.T.; Kwak, I.H.; Park, Y.C.; Seo, J.; Shim, J.Y.; Yoo, S.J.; Kim, J.-G.; Park, J.; Kang, H.S. Ruthenium Nanoparticles on Cobalt-Doped 1T′ Phase MoS2 Nanosheets for Overall Water Splitting. Small 2020, 16, 2000081. [Google Scholar] [CrossRef] [PubMed]
  105. Li, X.; Wang, Y.; Wang, J.; Da, Y.; Zhang, J.; Li, L.; Zhong, C.; Deng, Y.; Han, X.; Hu, W. Sequential Electrodeposition of Bifunctional Catalytically Active Structures in MoO3/Ni–NiO Composite Electrocatalysts for Selective Hydrogen and Oxygen Evolution. Adv. Mater. 2020, 32, 2003414. [Google Scholar] [CrossRef]
  106. Ji, X.; Lin, Y.; Zeng, J.; Ren, Z.; Lin, Z.; Mu, Y.; Qiu, Y.; Yu, J. Graphene/MoS2/FeCoNi(OH)x and Graphene/MoS2/FeCoNiPx multilayer-stacked vertical nanosheets on carbon fibers for highly efficient overall water splitting. Nat. Commun. 2021, 12, 1380. [Google Scholar] [CrossRef]
  107. Nguyen, D.C.; Tran, D.T.; Doan, T.L.L.; Kim, D.H.; Kim, N.H.; Lee, J.H. Rational Design of Core@shell Structured CoS@Cu2MoS4 Hybridized MoS2/N,S-Codoped Graphene as Advanced Electrocatalyst for Water Splitting and Zn-Air Battery. Adv. Energy Mater. 2020, 10, 1903289. [Google Scholar] [CrossRef]
  108. Sekar, S.; Aqueel Ahmed, A.T.; Pawar, S.M.; Lee, Y.; Im, H.; Kim, D.Y.; Lee, S. Enhanced water splitting performance of biomass activated carbon-anchored WO3 nanoflakes. Appl. Surf. Sci. 2020, 508, 145127. [Google Scholar] [CrossRef]
  109. Doan, T.L.L.; Nguyen, D.C.; Prabhakaran, S.; Kim, D.H.; Tran, D.T.; Kim, N.H.; Lee, J.H. Single-Atom Co-Decorated MoS2 Nanosheets Assembled on Metal Nitride Nanorod Arrays as an Efficient Bifunctional Electrocatalyst for pH-Universal Water Splitting. Adv. Funct. Mater. 2021, 31, 2100233. [Google Scholar] [CrossRef]
  110. Li, C.; Liu, M.; Ding, H.; He, L.; Wang, E.; Wang, B.; Fan, S.; Liu, K. A lightly Fe-doped (NiS2/MoS2)/carbon nanotube hybrid electrocatalyst film with laser-drilled micropores for stabilized overall water splitting and pH-universal hydrogen evolution reaction. J. Mater. Chem. A 2020, 8, 17527–17536. [Google Scholar] [CrossRef]
  111. Xiong, D.; Zhang, Q.; Li, W.; Li, J.; Fu, X.; Cerqueira, M.F.; Alpuim, P.; Liu, L. Atomic-layer-deposited ultrafine MoS2 nanocrystals on cobalt foam for efficient and stable electrochemical oxygen evolution. Nanoscale 2017, 9, 2711–2717. [Google Scholar] [CrossRef]
  112. Denisdon, S.; Senthil Kumar, P.; Rangasamy, G. High-Performance NiO/GCN Nanohybrids Synthesized under Different Thermal Conditions for Efficient Electrocatalytic Oxygen Evolution Reaction. Ind. Eng. Chem. Res. 2024, 63, 6951–6959. [Google Scholar] [CrossRef]
  113. Zabielaite, A.; Balciunaite, A.; Upskuviene, D.; Simkunaite, D.; Levinas, R.; Niaura, G.; Vaiciuniene, J.; Jasulaitiene, V.; Tamasauskaite-Tamasiunaite, L.; Norkus, E. Investigation of Hydrogen and Oxygen Evolution on Cobalt-Nanoparticles-Supported Graphitic Carbon Nitride. Materials 2023, 16, 5923. [Google Scholar] [CrossRef] [PubMed]
  114. Bai, X.; Cao, T.; Xia, T.; Wu, C.; Feng, M.; Li, X.; Mei, Z.; Gao, H.; Huo, D.; Ren, X.; et al. MoS2/NiSe2/rGO Multiple-Interfaced Sandwich-like Nanostructures as Efficient Electrocatalysts for Overall Water Splitting. Nanomaterials 2023, 13, 752. [Google Scholar] [CrossRef]
  115. Liu, P.F.; Yang, S.; Zhang, B.; Yang, H.G. Defect-Rich Ultrathin Cobalt–Iron Layered Double Hydroxide for Electrochemical Overall Water Splitting. ACS Appl. Mater. Interfaces 2016, 8, 34474–34481. [Google Scholar] [CrossRef]
  116. Sekar, S.; Sadhasivam, S.; Nangai, E.K.; Saravanan, S.; Kim, D.Y.; Lee, S. Enhanced Hydrogen Evolution Reaction Performances of Ultrathin CuBi2O4 Nanoflakes. Int. J. Energy Res. 2023, 2023, 5038466. [Google Scholar] [CrossRef]
  117. Kashif, M.; Thangarasu, S.; Oh, T.-H. Enriching the active sites of nanosheets assembled as flower-like MoS2 microstructure by controllable morphology for electrochemical hydrogen evolution reaction. Int. J. Hydrog. Energy 2024, 82, 47–52. [Google Scholar] [CrossRef]
  118. Srividhya, G.; Viswanathan, C.; Ponpandian, N. Interfacing NiV layered double hydroxide with sulphur-doped g-C3N4 as a novel electrocatalyst for enhanced hydrogen evolution reaction through Volmer–Heyrovský mechanism. Energy Adv. 2023, 2, 1464–1475. [Google Scholar] [CrossRef]
  119. Nguyen, T.V.; Tekalgne, M.; Tran, C.V.; Nguyen, T.P.; Dao, V.; Le, Q.V.; Ahn, S.H.; Kim, S.Y. Synthesis of MoS2/WS2 Nanoflower Heterostructures for Hydrogen Evolution Reaction. Int. J. Energy Res. 2024, 2024, 3192642. [Google Scholar] [CrossRef]
  120. Jiang, J.; Gao, M.; Sheng, W.; Yan, Y. Hollow Chevrel-Phase NiMo3S4 for Hydrogen Evolution in Alkaline Electrolytes. Angew. Chem. Int. Ed. 2016, 55, 15240–15245. [Google Scholar] [CrossRef]
  121. Ji, S.; Guo, J.; Yang, N.; Song, J.; Wang, Y.; Xing, G. Improvement of hydrogen evolution catalytic performance of MoS2 nanoflowers by constructing MoS2/MXene Ti3C2 heterostructure petals. Funct. Mater. Lett. 2024, 17, 2451011. [Google Scholar] [CrossRef]
  122. Feng, Y.-Y.; Deng, G.; Wang, X.-Y.; Zhu, M.; Bian, Q.-N.; Guo, B.-S. MoS2/NiFeS2 heterostructure as a highly efficient electrocatalyst for overall water splitting at high current densities. Int. J. Hydrogen Energy 2023, 48, 12354–12363. [Google Scholar] [CrossRef]
  123. Ilyas, T.; Raziq, F.; Ali, S.; Zada, A.; Ilyas, N.; Shaha, R.; Wang, Y.; Qiao, L. Facile synthesis of MoS2/Cu as trifunctional catalyst for electrochemical overall water splitting and photocatalytic CO2 conversion. Mater. Des. 2021, 204, 109674. [Google Scholar]
  124. Zhou, F.; Zhang, X.; Sa, R.; Zhang, S.; Wen, Z.; Wang, R. The electrochemical overall water splitting promoted by MoS2 in coupled nickel–iron (oxy)hydride/molybdenum sulfide/graphene composite. J. Chem. Eng. 2020, 397, 125454. [Google Scholar] [CrossRef]
  125. Zhang, J.; Wang, T.; Pohl, D.; Rellinghaus, B.; Dong, R.; Liu, S.; Zhuang, X.; Feng, X. Interface Engineering of MoS2/Ni3S2 Heterostructures for Highly Enhanced Electrochemical Overall-Water-Splitting Activity. Ed. Angew. Chem. Int. Ed. 2016, 55, 6702–6707. [Google Scholar] [CrossRef]
  126. Liu, Y.; Jiang, S.; Li, S.; Zhou, L.; Li, Z.; Li, J.; Shao, M. Interface engineering of (Ni, Fe)S2@MoS2 heterostructures for synergetic electrochemical water splitting. Appl. Catal. B 2019, 247, 107–114. [Google Scholar] [CrossRef]
  127. Zhang, B.; Xu, K.; Fu, X.; Guan, S.; Li, X.; Peng, Z. Novel three-dimensional Ni2P-MoS2 heteronanosheet arrays for highly efficient electrochemical overall water splitting. J. Alloys Compd. 2021, 856, 158094. [Google Scholar] [CrossRef]
  128. Mekete Meshesha, M.; Gautam, J.; Chanda, D.; Gwon Jang, S.; Lyong Yang, B. Enhancing the electrochemical activity of zinc cobalt sulfide via heterojunction with MoS2 metal phase for overall water splitting. J. Alloys Compd. 2023, 652, 272–284. [Google Scholar] [CrossRef]
  129. Liu, J.; Wang, J.; Zhang, B.; Ruan, Y.; Wan, H.; Ji, X.; Xu, K.; Zha, D.; Miao, L.; Jiang, J. Mutually beneficial Co3O4@MoS2 heterostructures as a highly efficient bifunctional catalyst for electrochemical overall water splitting. J. Mater. Chem. A 2018, 6, 2067–2072. [Google Scholar] [CrossRef]
  130. An, L.; Feng, J.; Zhang, Y.; Wang, R.; Liu, H.; Wang, G.-C.; Cheng, F.; Xi, P. Epitaxial Heterogeneous Interfaces on N-NiMoO4/NiS2 Nanowires/Nanosheets to Boost Hydrogen and Oxygen Production for Overall Water Splitting. Adv. Funct. Mater. 2019, 29, 1805298. [Google Scholar] [CrossRef]
  131. Banti, B.F.; Goddati, M.; Nwaji, N.; Gwak, J.; Gicha, B.B.; Kang, H.; Asgaran, S.; Chun, H.-J.; Lee, J. Defect Engineered Ru-CoMOF@MoS2 Heterointerface Facilitate Water Oxidation Process. ChemSusChem 2024, e202402533. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (a) Sonochemical fabrication process of the MoS2 nanosheets and gC3N4–MoS2 nanocomposites, (b,c) FE-SEM images of MoS2, (d,e) FE-SEM images of gC3N4–MoS2, and EDX spectra of (f) MoS2 and (g) gC3N4–MoS2.
Figure 1. (a) Sonochemical fabrication process of the MoS2 nanosheets and gC3N4–MoS2 nanocomposites, (b,c) FE-SEM images of MoS2, (d,e) FE-SEM images of gC3N4–MoS2, and EDX spectra of (f) MoS2 and (g) gC3N4–MoS2.
Materials 18 03775 g001
Figure 3. (a) XRD patterns, (b) N2-ADI characteristic curves, and (c) pore size distributions of MoS2 nanosheets and gC3N4–MoS2 nanocomposites.
Figure 3. (a) XRD patterns, (b) N2-ADI characteristic curves, and (c) pore size distributions of MoS2 nanosheets and gC3N4–MoS2 nanocomposites.
Materials 18 03775 g003
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

Sekar, S.; Shanmugam, A.; Lee, Y.; Lee, S. Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis. Materials 2025, 18, 3775. https://doi.org/10.3390/ma18163775

AMA Style

Sekar S, Shanmugam A, Lee Y, Lee S. Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis. Materials. 2025; 18(16):3775. https://doi.org/10.3390/ma18163775

Chicago/Turabian Style

Sekar, Sankar, Atsaya Shanmugam, Youngmin Lee, and Sejoon Lee. 2025. "Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis" Materials 18, no. 16: 3775. https://doi.org/10.3390/ma18163775

APA Style

Sekar, S., Shanmugam, A., Lee, Y., & Lee, S. (2025). Highly Efficient Electrocatalyst of 2D–2D gC3N4–MoS2 Composites for Enhanced Overall Water Electrolysis. Materials, 18(16), 3775. https://doi.org/10.3390/ma18163775

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