Open Access Article
Asad Ullah Khana,
Syed Haider Ali Shaha,
Fariah Salamb,
Afzal Shah
*a,
Faiza Jan Iftikharc,
Muhammad Umar Farooqd and
Muhammad Abdullah Khan
b
aDepartment of Chemistry, Quaid-i-Azam University, Islamabad, 45320, Pakistan. E-mail: afzals_qau@yahoo.com
bRenewable Energy Advancement Laboratory, Department of Environmental Sciences, Quaid-i-Azam University, Islamabad, 45320, Pakistan
cNUTECH School of Applied Science & Humanities, National University of Technology, Islamabad, 44000, Pakistan
dNational Center for Physics, Islamabad, 45320, Pakistan
First published on 18th September 2025
Developing efficient electrocatalysts for both the hydrogen evolution reaction (HER) and the oxygen evolution reaction (OER) is essential for advancing a sustainable energy future. In this study, bimetallic NiMo–LDH nanoflakes were successfully integrated onto reduced graphene oxide using a straightforward hydrothermal method. The composite material was characterized by FTIR, XRD, SEM, EDS, and XPS. Electrochemical evaluations highlighted the impact of rGO content, revealing that the NiMo–LDH with 7% rGO exhibited superior performance in OER, requiring only 230 mV overpotential and a Tafel slope of 60 mV dec−1 at a current density of 10 mA cm−2 in 1.0 M KOH. This enhanced performance is attributed to improved charge separation and transfer at the electrocatalyst/electrolyte interfaces, driven by the synergistic effects of rGO and NiMo–LDH. Furthermore, the composite electrocatalyst showcased bifunctional capabilities, with promising HER performance characterized by favorable overpotential and Tafel slope in 0.5 M H2SO4. Long-term stability was confirmed through chronoamperometry, while electrochemical impedance spectroscopy revealed efficient charge transport across the modified glassy carbon electrode represented as NiMo–LDH@rGO/GCE. Additionally, the synthesized catalyst demonstrated good recoverability, exhibiting appealing onset and overpotentials.
In principle, electrocatalytic water splitting comprises two-half reactions, the HER and OER. OER exhibits more sluggish kinetics compared to HER due to its four-electron-proton transfer process. Therefore, it is considered as the bottleneck in water splitting processes. To facilitate the kinetics of these reactions, a strategic design of electrocatalysts is crucial for enhancing reaction kinetics, particularly for OER.4 Noble metal-based catalysts such as Pt exhibit exemplary HER performance characterized by low overpotentials and small Tafel slopes, while Ru and Ir-based catalysts are used for OER due to their exceptional durability.5,6 The main issues associated with these electrocatalysts stem from their high cost and scarcity in the Earth's crust, which significantly hamper their extensive applications in strongly alkaline and acidic environments found in electrolyzers and proton exchange membranes (PEMs), respectively. Consequently, persistent endeavors have been directed toward development of economical bifunctional electrocatalysts that are efficient for both HER and OER.7 These electrocatalysts include transition metal (TM)-based oxides, hydroxides, sulphides, phosphides, oxyhydroxides, and selenides. Among the transition-based catalysts, 2D layered-double hydroxides (LDHs) have emerged as significant alternatives for OER due to the presence of OH− in the host layers.8 The structure of LDH provides active centers for OER to take place, where electrons occupy the d-orbitals of the TMs, thus enhancing the redox properties during the water oxidation process. The exposure of these sites in LDHs promotes coordination between the d-orbitals of the TM cations and oxygen species that play a vital role in lowering the energy barriers and speeding up the electrochemical reaction and ion-trapping ability. The anions between the layers maintain the equilibrium charge on surface hydroxides while at the same time preserving structural integrity.9 As a result, 2D LDH nanomaterials have emerged as prominent electrocatalysts in recent years owing to their distinctive layered architecture and adaptable properties. Further, the TM ions in LDH materials can provide vacant orbitals or lone electron pairs that influences the electron migration rate and the reconfiguration of the electrocatalyst surface.10 For example, Ni species offer a substantial number of occupied orbitals containing lone pair electrons, whereas high-valent molybdenum (Mo3+/Mo4+) seems to be an exceptional candidate for supplying empty orbitals with minimal or no occupation of d orbitals, asserting it as crucial for catalytic and ligand bonding reactions. Consequently, Ni and Mo-based nanomaterials can enhance active sites and inherent catalytic activity. Moreover, modifying the LDH intercalation anions can significantly mitigate the poisoning of the electrocatalyst surface.11 Ma and his coworkers successfully adopted a co-doping strategy by using Ru and Mn bimetallics to facilitate the systematic construction of electronic structure of both NiCo–LDH and NiCoP.12 Typically, LDH nanomaterials tend to agglomerate, with the active sites situated at the edges, which influences the interaction between the active sites and electrolyte leading to the constrained activity of LDHs for overall water splitting. However, the electrical conductance, functionality, surface area, and stability of TM hydroxides are significantly enhanced by introducing carbon-based materials due to their synergistic effects.
Carbon materials, such as graphene and other nanocarbon forms, have gained interest due to their surface functionality and electrocatalytic applications. This is largely attributed to their strong π–π interactions, which facilitate rapid electron mobility and enhance stability.13,14 Specifically, graphene oxide (GO) is an economical carbon-based conductor that improves catalytic characteristics. The additional reduction of GO to reduced graphene oxide (rGO) leads to improved electrical conductivity and significantly increases the surface area of metal hydroxides, facilitating enhancement in current and enabling rapid charge transfer. Moreover, the π-network in rGO contributes to the characteristic properties of graphene. Further, there has been a growing trend in the development of bifunctional electrocatalysts.15 A notable example is the rational design of a 2D heterostructure composed of graphene and hexagonal boron nitride (h-BN), which facilitates the formation of abundant interfaces and electroactive sites through π–π interactions. This catalyst demonstrated impressive performance, achieving a low overpotential of 28 mV at 10 mA cm−2 for the HER and 360 mV at 50 mA cm−2 for the OER. Additionally, it effectively served as both the cathode and anode in a quasi-triangulated bipolar electrolysis system, efficiently promoting overall water splitting.16 In addition, Rashid et al., employed an environmentally safe, stable, and cost-effective NiCo nanocomposite functionalized over rGO electrocatalyst to achieve efficient HER and OER activities at low overpotentials, resulting in a benchmark density of 10 mA cm−2.17 Sim and his colleagues synthesized a heterostructure composed of covalently bound 3D MoSe2 and rGO, which resulted in increased number of edge sites for improved catalytic properties for both OER and HER.18 Moreover, Sannegowda and his research team synthesized iron phthalocyanine by incorporating 4-amino-3-napthalene-2-sulfonic acid tri-functional monomers, and characterized it by employing a range of spectroscopic and analytical techniques. The prepared supramolecule was integrated with rGO, which was then immobilized onto a GCE for assessment of its bifunctional activity for both HER and OER during water electrolysis.19 Similarly, Gaddaerappa and his research group reported a highly promising bifunctional electrocatalyst synthesized by using quinone substituted cobalt(II) phthalocyanine (HQCoPc) in conjunction with carbon nanoparticles, which demonstrated high electrocatalytic performance for HER as well as OER.20
This study aims to synthesize a bimetallic NiMo–LDH nanomaterials, which is a hollow configuration superstructure synthesized by employing a hydrothermal method to accelerate water splitting reactions. With careful tuning of Ni and Mo content, the resulting bimetallic NiMo–LDH exhibited notable catalytic performance for water splitting. To make bimetallic LDH most efficient, stable and bifunctional, it was supported on rGO to form a hybrid nanomaterial. Experimental investigations indicated that the nanostructured heterostructure rGO-anchored, NiMoO4/NiFe–LDH nanosheet catalyst exhibited favorable kinetics and rapid charge and mass transfer while providing active sites with optimized adsorption energy and a tailored electronic structure while heteroatom doping with Mo in NiFe–LDH and heterostructuring with NiSx boosts its bifunctional activity for OER and HER.21,22 Consequently, it is expected that bimetallic Mo–NiFe LDH nanomaterial can demonstrate significantly enhanced HER and OER performance. Furthermore, extensive characterization techniques were utilized to examine physical and chemical properties. Further, the morphology, electronic structure, catalytic performance, and mechanism of the rGO-anchored bimetallic electrocatalysts were investigated by varying the quantity of rGO (3% and 7%). This novel material exhibits high efficiency for hydrogen production-a greener fuel that can contribute toward a carbon-free hydrogen economy. Furthermore, to the best of our knowledge, 3% and 7% NiMo–LDH@rGO hybrid nanomaterial have not been utilized as an electrocatalyst for water splitting before, nor have they been recorded for any other applications in scientific literature.
Similarly, in a typical method, 2 mmol of NiCl2·6H2O, and 1 mmol of Na2MoO4·2H2O were mixed in 50 mL of distilled water as shown in Fig. 1. The mixture was then treated with x wt% rGO (where x = 3 and 7) along with 12 mmol of urea, and continuously stirred. Subsequently, the solution was placed in an autoclave and subjected to a temperature of 120 °C for 12 hours, leading to the formation of a black precipitate. Upon cooling to room temperature, the synthesized samples were washed with a water/ethanol mixture and dried overnight at 60 °C in a vacuum drying oven. Finally, NiMo–LDH@rGO nanoflakes were obtained. Additionally, pristine NiMo–LDH was synthesized using the hydrothermal method, without adding the rGO, as a control.
The eqn (1) and (2) were employed to determine the overpotential (η) for OER and HER.
| ηOER = ERHE − 1.23 | (1) |
| ηHER = 0 − ERHE | (2) |
η = a + b log j
| (3) |
In the above equation, a and b demonstrate the charge transfer coefficient and Tafel slope, and j and η are the current density and overpotential, respectively.
![]() | (4) |
![]() | (5) |
The normalization of the curve was performed for the ECSA to find the true efficiency of the prepared electrode as porosity and roughness of the surface affect the catalytic efficiency. Thus, the electrochemical surface area, was calculated using the prescribed formula given in eqn (6) by using normalized current density (js) for determining the catalytic activity of the electrodes.
![]() | (6) |
![]() | (7) |
![]() | (8) |
![]() | (9) |
![]() | ||
| Fig. 2 (a) XRD patterns of NiMo–LDH, NiMo–LDH@3%rGO, and NiMo–LDH@7%rGO. (b) FTIR spectra of NiMo–LDH, NiMo–LDH@3%rGO, and NiMo–LDH@7%rGO. | ||
FE-SEM is most commonly used for characterization of solid materials and analysis of surface morphology. It offers magnification ranging from 10 to 300
000 times and enables non-destructive analysis. With a spatial resolution of up to 500 nm, FE-SEM was employed to analyze the surface morphology of the synthesized materials at various resolutions. The FE-SEM images presented in Fig. 3 illustrate the morphological changes of NiMo–LDH as the rGO content varies from 3% to 7%. The pure NiMo–LDH in Fig. 3(a) exhibits a characteristic nanoflake-like aggregates. Upon addition of 3% rGO (Fig. 3(d)–(f)), the texture transitions to a more porous form, suggesting the integration of rGO and an increase in the spacing between LDH layers. At 7% rGO (Fig. 3(g)–(i)), a significantly more open and interconnected network is observed, indicating that the higher rGO content successfully prevents the restacking of LDH. Similarly, EDS analysis was performed, which confirmed the presence of Ni, Mo, C and O in the synthesized nanomaterials as show in Fig. 3(j)–(p).
Similarly, the elemental composition and electronic state of the materials were examined by using XPS. The XPS spectrum clearly depicts the presence of various elements including Ni, Mo, C and O, which exactly aligns with the results obtained from EDS mapping as shows in Fig. 4(a). As depicted in Fig. 4(b) the Ni 2p spectrum of NiMo–LDH@rGO nanomaterial can be deconvoluted into four peaks, which consist of two main peaks due to spin–orbit doublets and two satellite peaks. The main peaks correspond to the Ni 2p3/2 and Ni 2p1/2 orbitals, respectively, which confirm the coexistence of Ni3+ and Ni2+. Moreover, the peak separation obtained between Ni 2p3/2 and Ni 2p1/2 is around 17.5 eV signifying their presence in the respective states.27,28 Likewise, in the Mo 3d Spectrum shown in Fig. 4(c) the two distinctive peaks were ascribed to the Mo 3d5/2 and Mo 3d3/2, respectively, which confirms the oxidation state of Mo6+. These states are found at binding energy levels of approximately 231.1 eV and 234.2 eV, respectively. The energy separation between these two peaks, amounts to 3.1 eV, highlighting the typical splitting between the spin states.29 Fig. 4(d) depicts a high-resolution C 1s spectrum, in which various oxygen functional groups are clearly associated with carbon. The non-oxygenated ring link with carbon (C
C bond), has a peak at 284.1 eV; the epoxy and hydroxyl groups have a peak at 285.2 eV; and the O
C–OH linkage associated with the carboxyl group has a peak at 288.1 eV.30 Fig. 4(e) shows the O 1s spectra of the NiMo–LDH@rGO nanocomposite, which is deconvoluted into three separate peaks at 531.2 eV (O–Ni/O–Mo), 532.1 eV (O–C), and 533.8 eV (O
C). The presence of these oxygen species implies the presence of metal oxygen and oxygen-containing groups in both the LDH and rGO components.31
![]() | ||
| Fig. 4 (a) XPS full survey spectrum of NiMo–LDH@rGO. High resolution spectra of (b) Ni 2p. (c) Mo 3d. (d) C 1s. (e) O 1s. | ||
| Catalysts | Electrolyte | η (mV) | Tafel slope (mV dec−1) | Ref. |
|---|---|---|---|---|
| NiFe–LDH/NiMo | 1 M KOH | 550 | 61 | 44 |
| NiCoFe–LDH | 1 M KOH | 233 | 29.39 | 45 |
| NiCo2O4 | 1 M KOH | 340 | 75 | 46 |
| NiFe–LDH/rGO | 1 M KOH | 245 | — | 47 |
| Exfoliated CoCo–LDHs | 1 M KOH | 353 | 45 | 48 |
| FeMoSe@NiCo–LDH | 1 M KOH | 240 | 78.4 | 49 |
| CuCo2O4/NrGO | 1 M KOH | 360 | 64 | 50 |
| Ni1Co2–LDH | 1 M KOH | 317 | 96.47 | 51 |
| FeCo2S4@NiCo–LDH | 1 M KOH | 262 | 142.1 | 52 |
| Co9S8/N,S-rGO | 1 M KOH | 266 | 75.5 | 53 |
| Co3O4/Ppy/C | 1 M KOH | 340 | 87 | 54 |
| NiMo–LDH@7%rGO | 1 M KOH | 230 | 60 | This work |
The kinetics of water oxidation can be evaluated using the Tafel slope, calculated by eqn (3): η = a + b
log
j. The Tafel slope determination is a commonly employed technique for finding the rate-determining step for OER. The Tafel slope values indicate the particular reaction route employed by the electrocatalytic process. The Tafel values for various catalysts are shown in Fig. 5(b). The modified GCE with NiMo–LDH@7%rGO demonstrates a minimum Tafel slope value of 60 mV dec−1, outperforming NiMo–LDH@3%rGO/GCE, NiMo–LDH@12%rGO/GCE and bare GCE, indicating that a smaller value of overpotential is required to sustain the reaction. Furthermore, a lower Tafel slope value indicates efficient OER at the surface as a consequence of fewer barriers to the progression of the reaction than the presence of numerous charged intermediates competing for the release of oxygen gas on the electrocatalyst surface that can lead to a higher Tafel slope.34 Similarly, Fig. 5(c) compares overpotential with reported electrocatalysts for OER. The Tafel slope analysis of NiMo–LDH@7%rGO/GCE indicated a slope close to the standard 40 mV dec−1, implying that this reaction is crucial in determining the overall rate of the reaction. This observation corresponds with the η, wherein a reduced Tafel slope value indicates a significant rise in j with a slight change in overpotential, as shown in Fig. 5(d).
Fig. 6 illustrates the overall proposed mechanism of the OER occurring on the surface of NiMo–LDH@7% rGO. In an alkaline environment, nickel sites undergo electrochemical oxidation to form Ni3+/Ni4+ species, which are recognized as active sites for the adsorption and oxidation of oxygen intermediates, thereby reducing the overpotential. For instance, research by Edvinsson et al. has shown that the redox activity of NiMoO4 nanorods on nickel foam is linked to γ-NiO(OH), which serves as the active surface species for efficient OER performance.35 Additionally, Huang et al. provided evidence through operando Raman spectroscopy and ex situ XRD that the unstable Ni(OH)2, when subjected to an applied potential, is reduced to form amorphous NiO, which subsequently converts into γ-NiO(OH).36 In a related study, Rajput et al. analyzed NiMoO4/NF in a 1 M KOH solution and identified a redox peak associated with the Ni(II)/Ni(III) transition. As cycling progressed, the current density increased, and the peak shifted anodically, indicating a reformation of the catalyst. After ten cycles, Raman spectra revealed bands at 483 and 562 cm−1 corresponding to NiO(OH), while immersion in KOH without an applied potential produced a 500 cm−1 Ni–O band. This suggests that both the alkaline environment and the applied potential facilitate the conversion of NiMoO4 to NiO/NiO(OH).37 Notably, some studies have identified NiMoO4 nanostructures on nickel foam (NF) as effective and stable catalysts for the OER.38–40 Zhang et al. demonstrated that within NiMoO4, molybdenum is oxidized to MoO42−, which subsequently polymerizes and re-adsorbs onto the catalyst surface, thereby enhancing its stability. Furthermore, recent findings indicate that the Mo6+ ions located in the octahedral interstitial sites of the monoclinic NiMoO4 structure do not participate directly in the redox reactions. Instead, they play a crucial role in improving the material's electronic properties by enhancing conductivity, reducing electron density, promoting OH− adsorption, and facilitating electron transfer, which supports effective Ni2+/Ni3+ redox reactions without faradaic contributions.41,42 In addition the incorporation of rGO not only improves conductivity and reduces charge-transfer resistance but also prevents the stacking of LDH nanosheets, resulting in an increased electrochemically active surface area. This effect was demonstrated by Chen et al., who investigated the performance of NiFe–LDH@NF and NiFe–LDH/rGO@NF as catalysts for the OER. Their findings revealed that the rGO-enhanced material exhibited more pronounced absorption peaks for M–OH, M–O, and M–OOH intermediates compared to NiFe–LDH@NF. This indicates a higher concentration of intermediates and improved electron transfer, which collectively contribute to enhanced OER activity.43 Consequently, the overall mechanism for OER may need to be revised as follows,
| NiMo–LDH@7%rGO + OH− → NiMo@7%rGO–OH* + e− | (10) |
| NiMo–LDH@7%rGO–OH* + OH− → NiMo@7%rGO–O* + H2O + e− | (11) |
| NiMo–LDH@7%rGO–O* + OH− → NiMo@7%rGO–OOH* + e− | (12) |
| NiMo–LDH@7%rGO–OOH* + OH− → O2 + H2O + e− | (13) |
Similarly, Table 1 compares the current work with the previous reported catalysts, clearly showing that NiMo–LDH@7%rGO/GCE has the lowest Tafel slope and overpotential among the reported catalysts, exhibiting exceptional catalytic capabilities.
Besides overpotential and Tafel slope another crucial factor in assessing the efficiency of LDHs as electrocatalysts is their ECSA, defined as the accessible surface area for electrochemical processes. The porous nature of the LDHs nanomaterial determines the surface area for the adsorption of reactants and products, affecting the catalytic property of the LDHs. For example, an LDH with larger pores may exhibit a higher ECSA due to its capacity to absorb more reactants and products. The integration of rGO into the structure can influence the ECSA of the electrocatalysts since its high surface area and conductivity are shown to enhance the electron transport and augment the catalyst's overall efficiency. The ECSA of NiMo–LDH@7%rGO on a GCE in 1 M KOH solution at varying scan rates was investigated within a limited voltage range (1.00 V–1.17 V), depicted in Fig. 7(a)–(c). The relationship between scan rate and anodic–cathodic current difference (Δj) in CV polarization curves, shown in Fig. 7(d)–(f), indicates that the Cdl value, which is half the slope, approximates the ECSA of NiMo–LDH and rGO supported NiMo–LDH. The Cdl of the NiMo–LDH@7%rGO is significantly greater than the NiMo–LDH and the NiMo–LDH@3%rGO. The ECSA of the NiMo–LDH@7%rGO exhibits a 1200.5 cm2 surface area, which is exceedingly greater than NiMo–LDH@3%rGO (1112.5 cm2) and NiMo–LDH (865 cm2). The results indicate that rGO promotes the transport of electrolytic ions to reactive sites, resulting in a reasonable ECSA value. An increase in its value indicates a higher number of reactive sites that facilitate the catalysis of water.
To evaluate the intrinsic properties of synthesized electrocatalysts, mass activity (mA mg−1) was calculated at a fixed overpotential of 1.9 V by dividing the current density by mass loading. The values of MA of NiMo–LDH@7%rGO/GCE (10.4 mA mg−1) were higher than NiMo–LDH@3%rGO/GCE (8.5 mA mg−1) and NiMo–LDH/GCE (4.2 mA mg−1) as illustrated in Fig. 8(a). Additionally, the TOF was employed to determine the intrinsic activity of the catalyst as shown in Fig. 8(b), which was plotted against the applied voltage vs. RHE. The TOF value for electrocatalyst NiMo–LDH@7%rGO/GCE was calculated to be 0.0012 s−1, while the values for NiMo–LDH@3%rGO/GCE and NiMo–LDH/GCE were 0.0010 s−1, and 0.00049 s−1, respectively. This signifies that NiMo–LDH@7%rGO/GCE demonstrates a higher intrinsic activity dependent on surface reactions. These findings further highlight the fact that adding rGO enhanced the mass activity, specific activity, and turnover frequency. In addition, the exceptional efficiency of NiMo–LDH@7%rGO/GCE facilitates the OER because of the larger surface area of the nanomaterial, unobstructed channels, and porous layered structure. These variables enhance the transfer of ions to the active sites, leading to a significant enhancement in OER performance.
![]() | ||
| Fig. 8 (a) Mass activity of synthesized catalysts: NiMO–LDH, NiMo–LDH@3%rGO, NiMo–LDH@7%rGO (b) relation of TOF with overpotentials. | ||
Furthermore, EIS was conducted to assess the electron transfer rate of each catalyst during the electrochemical reaction and to elucidate the processes behind OER activity for rGO supported bimetallic LDHs hybrid nanomaterial. The AC impedance of all materials was performed at 1.5 V in a voltage range of 0.1 Hz to 106 Hz. Fig. 9(a) shows the fitted EIS spectra of the three as-prepared catalysts and bare GCE. The charge-transfer resistance Rct of NiMo–LDH@7%rGO/GCE, NiMo–LDH@3%rGO/GCE, NiMo–LDH/GCE, and bare GCE was found to be 20.99 Ω, 29.47 Ω, 41.77 Ω, and 46.95 Ω, respectively. The results demonstrate that the NiMo–LDH@7%rGO/GCE exhibits much lower Rs of 8.95 Ω and Rct of 20.99 Ω compared to other catalysts, signifying effective charge transfer between the electrode and catalyst. Fig. 9(a) depicts the intersection of Nyquist impedance curve with Z′ at high frequencies, pointing out Rs associated with the reaction rate, and at low frequencies, the Rct.55
Electrochemical stability is a pertinent parameter for evaluating catalyst performance, as operational conditions can affect the efficiency of the reaction that directly influences the stability of the catalyst. Chronoamperometry and CV experiments were conducted to assess the stability of the electrocatalyst. The stability of NiMo–LDH@7%rGO/GCE was assessed through a cycling endurance test, by subjecting the catalyst to repeated LSV cycles within the potential range of 0–1.5 V vs. RHE at a scan rate of 50 mV s−1. These curves were subsequently acquired on the modified electrode in a freshly prepared electrolyte. The LSV curves showed that current density decreased by only 3% of the original current density after 1500 cycles, as illustrated in Fig. 9(b). This indicated that the structure remained intact, and no variation in current density was observed. Further, chronoamperometry of NiMo–LDH@7%rGO/GCE was conducted in 1 M KOH solution at a corresponding potential of 1.46 V relative to Ag/AgCl. The 29 hour time frame indicates that the current density of 15 mA cm−2 remains constant up to a 29 hour period, which is illustrated in Fig. 9(c). The lack of change in current density shows the stability of the electrocatalyst during the operational conditions, offering sustained performance.
The stability evaluation was performed on NiMo–LDH@7%rGO/GCE by recording LSV before and after 29 hours of chronoamperometric operation. Fig. 9(d) shows notable stability, with a minor divergence in the LSV curve compared to the original curve after 29 hours of operation. This remarkable long-term stability can be attributed to the strong synergistic effect between Ni and Mo ions and the incorporation of rGO, which serves as an electronic scaffold to the electrocatalysts. These results indicate that NiMo–LDH@7%rGO/GCE displays exceptional catalytic performance, surpassing most TM-based nanomaterials employed for the OER in terms of overpotential, Tafel slope and current density.56
The electrocatalytic HER performance was evaluated for NiMo–LDH@rGO/GCE in 0.5 M H2SO4 solution for assessing its water-splitting efficiency as a catalyst. The LSV polarization curves are shown in Fig. 10(a). The NiMo–LDH@7%rGO/GCE has superior HER performance as compared to NiMo–LDH@3%rGO/GCE and NiMo–LDH/GCE. This can be ascribed to its high surface area and increased number of reactive sites, resulting in lower overpotential (225 mV) obtained for NiMo–LDH@7%rGO/GCE at 10 mA cm−2. Another parameter used to evaluate the efficiency of the HER was the current density. Here, maximum current density was achieved by NiMo–LDH@7%rGO/GCE due to enhanced exposure of active sites for electron transfer, accelerating reaction kinetics.57
The Tafel slope is illustrated in Fig. 10(b) which elucidates the mechanism of HER in acidic medium. The Tafel slope of NiMo–LDH@7%rGO/GCE is markedly lower at 136 mV dec−1, in comparison to NiMo–LDH/GCE (138 mV dec−1), NiMo–LDH@3%rGO/GCE (146 mV dec−1), and bare GCE (211 mV dec−1). This indicates that the HER for NiMo–LDH@7%rGO/GCE occurs through a Volmer–Heyrovsky mechanism with enhanced kinetics.58 In the Volmer phase, a proton from the electrolyte discharges towards the electrode by capturing electrons from the surface to form the H* adsorbed species (eqn (14)). During the Heyrovsky step, the adsorbed hydrogen atom is liberated from the surface and reacts with the proton (in the acidic medium) to generate a hydrogen gas molecule (eqn (15)). While in a separate step (Tafel step), two adsorbed H* species combine to form the H2 molecule (eqn (16)). Hence, the electrochemical conversion of the H2O molecule into H2 gas takes place through various distinct steps.59 The following mechanism is followed by HER in an acidic medium,
| H+ + e− → H* | (14) |
| H+ + H* + e− → H2 | (15) |
| 2H* → H2 | (16) |
To gain deeper insights, the HER mechanism was investigated, which entails the reduction of protons through a series of steps that ultimately yield hydrogen gas.60 The HER mechanism can vary depending on whether the conditions are acidic or alkaline. In their study, Li et al. utilized voltage-dependent operando Raman spectroscopy to analyze the active site of N–NiMoO4/Ni/CNTs. Their findings revealed that the NiO peaks between 400 and 600 cm−1 remained unchanged with varying potential, indicating that NiO does not contribute to the reaction.61 In contrast, the spectral features related to Ni (1000–1200 cm−1) and C (1400–1600 cm−1) showed significant variations, with transient peaks for Ni2+–OH* and C–H* appearing at approximately 1150 and 1520 cm−1 at −0.1 V. This behavior suggests that water dissociation and its subsequent adsorption on Ni2+ are taking place.62 The disappearance of these peaks after the HER confirms their transient nature, while the MoO42− peaks at 814, 889, and 938 cm−1 remain stable, suggesting that MoO42− enhances conductivity without directly participating in the reaction. Conducting the HER study in an acidic environment provides sufficient H+ ions to adsorb onto the Ni2+ ions, facilitating their direct involvement in hydrogen evolution. Additionally, the rGO component functions as a co-catalyst, promoting hydrogen adsorption and stabilizing NiMo–LDH through effective electron transfer to active sites, thereby increasing the accessibility of the surface area.63,64 Fig. 10(c) and (d) illustrate the comparison of reported electrocatalysts with the as-synthesized material and the relationship graph (between Tafel slopes and overpotentials of the as-prepared LDHs) respectively. Table 2 presents the performance of various electrocatalysts for the HER, highlighting the effectiveness of our synthesized rGO-based bimetallic NiMo–LDH.
| Catalysts | Electrolyte | Overpotential η (mV) | Tafel slope (mV dec−1) | Ref. |
|---|---|---|---|---|
| Co3O4/WO3/C nanorods | 0.5 M H2SO4 | 400 | 105 | 65 |
| NiCo–LDH/CFP | 1 M KOH | 295 | 195.51 | 66 |
| NiFe–LDH/CMT | 1 M KOH | 333 | 140 | 67 |
| RuO2/MgFe–LDH | 0.5 M H2SO4 | 360 | 76.06 | 68 |
| Fe–NiS2 | 0.5 M H2SO4 | 121 | 37 | 69 |
| Ni–MoS2 | 0.5 M H2SO4 | 417 | 139 | 70 |
| NiFePd–LDH | 1 M KOH | 412 | 171.4 | 71 |
| C doped CoFe2O4/Fe2O3 | 1 M KOH | 236 | 146 | 72 |
| CeO2–CuO–Mn3O4 | 1 M KOH | 380 | 175 | 73 |
| Co3S4 doped MoS2–Ni3S2 | 1 M KOH | 317 | 109 | 74 |
| NiMo–LDH@7%rGO | 0.5 M H2SO4 | 225 | 136 | This work |
Moreover, EIS measurements were conducted to examine the kinetics of the electrocatalysts at the electrode/electrolyte interface in 0.5 M H2SO4, as illustrated in Fig. 11(a). All the fitted curves of EIS data were analyzed using a Randles equivalent circuit depicted in the inset of Fig. 11(a), which comprises of a Rs resistor and a series arrangement with resistor Rct and constant phase element (CPE) that accounts for inhomogeneity at the electrode. According to the EIS results, the NiMo–LDH@7%rGO/GCE, NiMo–LDH@3%rGO/GCE, and NiMo–LDH/GCE exhibit an Rct value of 2.54 kΩ, 2.84 kΩ, and 4.93 kΩ, respectively, while bare GCE has a value of 7.35 kΩ. From this EIS data, NiMo–LDH@7%rGO/GCE possesses the smallest Rct = 2.54 kΩ, suggesting a fast charge transfer kinetics. This value is consistent with the low overpotential and Tafel slope for NiMo–LDH@7%rGO/GCE.75 Additionally, the intrinsic activity was evaluated by TOF for HER at an overpotential of 280 mV. A bar graph is presented in Fig. 11(c) to illustrate the practical application of NiMo–LDH/GCE, NiMo–LDH@3%rGO/GCE, and NiMo–LDH@7%rGO/GCE. The layered structure of NiMo–LDH@7%rGO/GCE demonstrated a significant TOF value of 0.0018 s−1, while the values for NiMo–LDH@3%rGO/GCE and NiMo–LDH/GCE were 0.0015 s−1 and 0.0007 s−1, respectively. The data indicates that NiMo–LDH@7%rGO/GCE exhibits significant intrinsic activity in a surface-dependent process. Similarly, long-term electrocatalytic stability is essential for the catalyst's suitability for diverse applications, including energy storage devices. Further chronoamperometry was performed to assess the long-term stability of the catalyst. Fig. 11(b) depicts chronoamperometric results of NiMo–LDH@7%rGO/GCE, conducted at a constant potential of −0.225 V relative to Ag/AgCl for 27 hours to test the stability of the electrocatalyst. The investigation indicates that the NiMo–LDH@7%rGO/GCE exhibits long-term stability at 10 mA cm−2 for up to 27 hours. Thus, the LSV curves for HER of NiMo–LDH@7%rGO/GCE illustrated in Fig. 11(d) demonstrate its stability with a negligible decrease in current density.
Supplementary information contains the LSV polarization curve of NiMo-LDH@12% rGO. See DOI: https://doi.org/10.1039/d5ra04536c.
| This journal is © The Royal Society of Chemistry 2025 |