3.1. XRD analysis
The diffraction pattern of the as-prepared MoS2-based sample was determined using X-ray crystallography. Only hexagonal MoS2 crystalline structures were apparent in the XRD patterns of MoS2, MoS2/rGO hybrid, and MoS2/rGO hybrid in Fig. 3a. The crystal planes of MoS2 (100), (103), and (110) were found using sharp patterning at 2 = 32.8o, 39.8o, and 58.5o, respectively (JCPDS No.37-1492). There were also no further peaks in the sequence, indicating that the commodity is extremely pure. The presence of (002) peak along with MoS2 related peaks in the MoS2@RGO composite sample demonstrate that successful formation of heterostructure composites [26]. Moreover, the intensity of RGO to efficiently inhibit stacking of the MoS2 layers was demonstrated by a significant drop in the diffraction peaks following its addition [27].
3.2. Morphological studies
The hierarchical MoS2@RGO composites were examined using SEM and are displayed in Fig. 2 (a&b) at various magnifications. The MoS2 flowers actually indicate a nanosheet construction throughout. Furthermore, the MoS2 has a multitude of accessible active sites, which is helpful for sonocatalytic activity. Because graphene oxide possesses such a high surface area, MoS2 nanoparticles may be distributed equally across the RGO. According to these findings, it is highly likely that the MoS2 nanosheets are hierarchical structures organised and securely attached to the RGO. The TEM photograph of MoS2 (Fig. 2c) revealed that the material had clumped together into massive clusters and taken on the shape of nanoplates. As illustrated in Fig. 2d, the evolution of pure MoS2 was more scattered; however, the MoS2 contained in the MoS2@rGO hybrid formed in a fashion that was both homogeneous and firmly compacted on the face of the rGO. The lattice gap of MoS2's (002) and (101) planes was measured to be 0.63 nm, while the (002) plane measured 0.28 nm (see Fig. 2e). The EDS mapping of the MoS2@RGO compound is shown in Fig. 2 (f-i). Pure MoS2 is depicted in the photo as a nanoparticle with the shape of a flower. MoS2 nanosheets, on the other hand, are evenly distributed among RGO sheets.
3.3.FTIR and Raman spectra analysis
FTIR spectroscopy was utilised to keep monitor of the likelihood that the GO in the hybrids had been transformed to RGO via the hydrothermal pathway (Fig. 3a). The GO sample's spectrum analysis displays a broad and widespread peak at 3000–3750 cm− 1, suggesting the existence of the O-H stretching vibration [28]. This happens because the GO sample was not completely dried before being analysed. The faint signal observed between 1700 and 1900 cm-1 suggests C = O or COOH. Peaks at 1550–1650 and 1350 cm− 1 [29] gave solid evidence for the presence of sp2-hybridized C = C and C = O-C. However, the absence of a clear peak at 1700–1900 cm1 in the spectra of MoS2@RGO hybrids indicates that the C-O and COOH groups have been removed. The utilisation of FTIR spectra revealed that graphene oxide can be successfully converted into reduced graphene oxide via a hydrothermal method. The Raman spectra (Fig. 4b) show two strong peaks at 375 cm− 1 and 405 cm− 1, which correspond to the hexagonal MoS2 well as well as the in-plane E12g and out-of-plane A1g translational modes [30]. Furthermore, the D peak at 1352 cm− 1 and the G peak at 1597 cm− 1 are apparent in all spectrums. The ID/IG substantially improve, which measures the intensity of the D peak in relation to the intensity of the G peak, could be used to assess flaws in graphene-based materials. This ratio reflects the D and G peaks. As a result, we know that the ID/IG ratio of the test specimens is 1.02 (for RGO) and 1.13 (for MoS2@RGO), indicating that the instability of MoS2@RGO is more visible than that of pure RGO. It is widely acknowledged that MoS2@RGO shown exceptional sonocatalytic activity. This is due to the fact that a larger ID/IG ratio allows for higher levels of electrical conductivity and liquid immersion.
3.4. BET and XPS analysis
The N2 sorption profiles (Fig. 4(c)) demonstrate the differences in the porous nature of 1T MoS2 and its RGO hybrid. Often, rGO hybrids display type IV N2 sorption properties, characterised by mesoporous behaviour and rapid adsorption and desorption enrichment [31–33]. It was determined that the surface area of MoS2@RGO was 106.5 m2/g, whereas that of pure MoS2 was 55.2 m2/g. Whereas MoS2 and MoS2 have pore sizes of around 12 and 7 nm (Fig. 4d). The addition of rGO to the MoS2 nanostructure boosts its surface area, which in turn boosts the likelihood of accessing more active surface area in the combination nanostructure [34]. This, in turn, raises the level of onset and active corners, and it also improves the reflexes at the electrode/electrolyte interaction. XPS (Fig. 5) are conducted to determine the elemental composition and the chemical states of the elements by scanning spectra (Fig. 5a). Likewise, a large C 1 s signal from CC at 285.0 eV and a tiny CO peak at 286.9 eV were seen in the MoS2@RGO heterostructure (Fig. 5(b)). The significantly reduced graphene oxide to rGO is supported by the somewhat lower signals corresponding to O = C-O and CO. Figure 5(c) shows the overlaid MoS2@RGO XPS spectrum, which reveals that Mo mostly occurs as Mo 3d5/2 due to its binding energy of 229.5 eV and as Mo 3d3/2 due to its binding energy of 232.7 eV. (IV). S 2 s peak at 226.6 eV is typical of MoS2 [35], hence this further demonstrates the element's existence. Good agreement was found between the sulphide phase and the S2p peaks in Fig. 3(d) at 162.4 (2p3/2) and 163.5 (2p1/2) eV with an energy gap of 1.1 eV [36].
3.5. Electrochemical HER analysis
In a 3-electrode configuration at room temperature with electrolytes consisting of 0.5 M H2SO4 and 1 M KOH, pure NF, Pt/C, MoS2, and MoS2@RGO were studied. The LSV polarization profiles of Pt/C, bare NF, MoS2, and MoS2@RGO are shown in Fig. 6(a) and (b), respectively. Figure 6(c) depicts the voltage rise that was recorded throughout all three electrocatalysts in the experiment. Pt/C electrocatalysts have a low redox value (41 and 45 mV against RHE, respectively), while bare NF requires an overpotential of 453 and 464 mV versus RHE to generate a current density of 10 mA.cm2 in H2SO4 and KOH. Electrocatalysts may be manufactured, and one instance is MoS2, which needs 118 mV against RHE in acidic medium and 101 mV against RHE in alkaline media to reach 10 at the appropriate potential, accordingly. The contains the following is superior to earlier MoS2 HER results such as 229 mV@10 mA/cm2 [49], 119 mV@15 mA/cm2 [37], 232 mV@20 mA/cm2 [38], 178 mV@100 mA/cm2 [39], and 346 mV@20 mA/cm2 [40]. Inactive or substandard active edges for H2 evolution explain why pure phase MoS2 performs worse in HER than mixed phase MoS2. Tafel plots (shown in Fig. 6(d) and e) were also employed to investigate electrocatalyst properties. The following Tafel slopes were derived via a linear fit: In acid media, 32 mV.dec− 1; in bare NF, 159 mV.dec− 1; in 1T MoS2, 128 mV.dec− 1; and in 1T MoS2@RGO, 44 mV.dec− 1. This catalyst clearly outperforms the findings reported with previous published MoS2 catalysts (Table 1) [41–45]. The excellent electrocatalytic behavior of the 1T MoS2/rGO combination is further enhanced by its low Tafel slopes. EIS was used in this investigation to decipher the HER dynamics and interfacial effects. Pt/C, bare NF, 1T-MoS2, and 1T MoS2@rGO EIS measurements were carried out in acidic and alkaline electrolytes across a frequency range of 0.01 Hz to 100 kHz. Figures 6(f) and (g) illustrate the comparable Nyquist plots in acidic and alkaline environments, respectively. Lower than expected initial resistances confirmed improved NF interaction with electrocatalysts (Rs). Under acidic circumstances, Rs values of 2.63, 2.81, 2.15, and 1.95 were recorded; for Pt/C, bare NF, 1T-MoS2, and 1T MoS2@rGO, Rs values of 1.96, 2.22, 2.38, and 1.83 were evaluated, correspondingly. In both acidic and basic solutions, all electrocatalytic displayed tiny semicircles, which can be interpreted as charge-transfer resistance (Rct). The Rct values are 3.56 for acidic, 3.96 for basic, 2.85 for 1T-MoS2, and 2.30 for 1T-MoS2@rGO. A low CTR, which implies a high rate of interaction among the electrolyte and the electrode's surface, may be caused by a large number of functional edges, which may additionally contribute to the low CTR. Figure 8(h) and Fig. 8(i), respectively, illustrate the density variations over time for MoS2/rGO hybrids in acidic and alkaline conditions, etc. The creation of hydrogen by an acidic electrolyte remains unchanged over the whole 24 hours, but the production of hydrogen by an alkaline electrolyte experiences a significant and dramatic reduction after the first 12 hours. Hydrogen ion absorption or H2 bubble accumulation on the electrocatalyst surface may be responsible for inducing inconsistencies in alkaline medium, which ultimately results in a reduction in HER [46]. UPS examined the energy levels of the developed catalysts, MoS2 and MoS2@RGO. This was accomplished using the vacuum amount and the Fermi level. Energy levels of 5.96 and 5.47 eV have been calculated for the aforementioned catalysts (Fig. 7a & c). The matching band structure of the electrocatalyst sample is also shown in Fig. 7c. Overall, 24 hours of chronoamperometric studies on 1T MoS2@rGO electrocatalysts in acid and alkaline solutions revealed the Robust HER characteristic. As a result, the suggested simple and trustworthy approach to synthesise the TMDs MoS2 layers decorated RGO establishes a fresh route to developing next-generation HER electrocatalysts.