0
Views
0
CrossRef citations to date
0
Altmetric
Research Article

Facile synthesis of MoSe2 embedded in La2O3 crystal for enhancing hydrogen evolution reaction

Article: 2367552 | Received 12 Apr 2024, Accepted 06 Jun 2024, Published online: 18 Jun 2024

ABSTRACT

Molybdenum diselenide (MoSe2) is a potential catalyst for the electrocatalytic hydrogen evolution reaction (HER). However, MoSe2 exhibits a high overpotential in HER due to its low density of active sites, which may limit its practical applicability. Tuning the assisted metal interaction is an effective method for altering the atomic structure and catalytic efficiency of transition metal catalysts. Herein, we report a simple hydrothermal process for synthesising hybrid MoSe2-La2O3 heterostructures for HER. The electrocatalytic HER of nanoflower MoSe2 was improved in the presence of a supported lanthanide-oxide interaction (La2O3), proving that La2O3 materials produce additional activity sites in the MoSe2 materials. The nanocomposite was drop-cast onto screen-printed carbon electrodes (SPCE) and showed reduced overpotential values compared to pristine MoSe2 and La2O3 materials. Across all ratios tested, the MoSe2-La2O3 (M-L2) nanocomposites with a molar ratio of 1:2 exhibit rapid charge transfer during HER performance, with a Tafel slope value of 66 mV/dec and an onsetpotential of 270 mV.

1. Introduction

Hydrogen is considered one of the most likely replacements for fossil fuels in light of the depletion and pollution of traditional fossil fuels (Dresselhaus and Thomas Citation2001; Nicoletti et al. Citation2015; Sharma, Agarwal, and Jain Citation2021). An excellent and highly effective method to produce hydrogen is by electrocatalytic water-splitting process (Shinagawa and Takanabe Citation2017). When choosing a catalyst material for hydrogen evolution, it is essential to consider the efficiency of the electricity-to-hydrogen conversion (Laursen et al. Citation2012). Platinum (Pt) noble metals have been identified as the most efficient hydrogen evolution catalysts due to their almost negligible Gibbs free energy for hydrogen absorption (X. Liu et al. Citation2023; Yan et al. Citation2021; Zou and Zhang Citation2015). However, their limited availability and expensive price hinder their widespread application. Consequently, developing effective and economical catalysts that may substitute noble metals for HER is imperative to potentially lower the cost and increase the accessibility of hydrogen generation.

Recently, there has been considerable interest in developing a variety of Pt-free catalysts to lessen the Pt dependence of HER catalysts. Transition metal dichalcogenides (TMDC) have sparked a lot of interest as electrocatalysts due to their unusual structure, excellent electrochemical activity, and wide availability (Alahmadi et al. Citation2021; Ren et al. Citation2018). It is worth noting that all transition metal dichalcogenide crystals are composed of three layers held together by Van der Waals forces; the metal forms the core atom, while chalcogenide atoms (X) constitute the top and bottom layers. The chalcogenide atoms improve the electrical properties and structural stability of the crystals (Manzeli et al. Citation2017; Vikraman et al. Citation2018). Unlike other TMDC materials, molybdenum diselenide (MoSe2) materials, which have two phases: 2H, a semiconductor, and IT, a metal, have attracted interest in research as HER electrocatalysts (Deng et al. Citation2019; Li et al. Citation2018). It is well known that electrocatalytic HER activity can be influenced by active sites, intrinsic activity, and conductivity. Typically, the basal plane of the majority of TMDC materials is rather inert, but the outer edges provide catalytically active sites (Wu et al. Citation2021). Additionally, due to the MoSe2 nanoflakes having several inert basal planes, the electrochemical activity of these flacks is caused by unsaturated Se atoms at their edge positions (Masurkar, Thangavel, and Arava Citation2018; Tang et al. Citation2014). As a result, the electrocatalytic efficiency of MoSe2 can be limited by its small number of active sites.

To address this critical issue, numerous attempts have been undertaken to increase electron delivery in the catalyst. By combining MoSe2 with other materials such as highly conductive carbon materials graphene, carbon nanotubes, and gold, electron transport in the catalyst might be enhanced (Mabuea, Erasmus, and Swart Citation2023). Additionally, researchers have explored the use of doping techniques to improve the electron transport properties of MoSe2 catalysts (Alahmadi and Aoun Citation2024). These techniques include introducing metallic or nonmetallic element dopants or creating defects in the MoSe2 structure. The HER performance can be also enhanced by changing the growth conditions of MoSe2, which results in the production of MoSe2 nanosheets with specific structures and high active edge sites.

In the current study, a nanocomposite composed of MoSe2 and La2O3 materials was produced using a simple one-step hydrothermal process to improve the HER activity of MoSe2. These hybrid mixtures were evaluated as HER catalysts using a screen-printed carbon electrode. We believe that various interactions in the material, such as those between Mo and Se atoms and La2O3 component atoms, might lead to the development of additional catalytic sites, which are critical for improving the electrocatalytic performance of MoSe2. With the inclusion of La2O3 components, MoSe2 nanoflower structures may self-assemble into nanosheets, resulting in more active sites and, hence, increased electrochemical activity. Additionally, the investigation results demonstrated that M-L2 is much superior and more efficient than M-L1 and M-L3 for electrocatalytic HER performance. Similarly, the other electrocatalytic data studied at acidic conditions approved low charge transfer resistance and the high stability of M-L2. Several techniques were conducted to analyze the phase structure, composition, and morphology of the hybrid nanocomposite.

2. Experimental section

2.1. Material

Ammonium molybdate tetrahydrate ((NH4)6Mo7O24·4H2O), lanthanum oxide (La2O3) and selenium (Se) were purchased from Sigma-Aldrich. The screen-printed carbon electrodes (SPCE) are obtained from Metrohm. Without additional purification, all chemicals were utilised simply as purchased.

2.2. Synthesis of the hybrid MoSe2/La2O3 nanocomposite

The MoSe2-La2O3 nanocomposite was synthesised in a single step using hydrothermal synthesis. This affordable approach aids in producing well-dispersed, scalable, and easy-to-handle nanomaterials. illustrates the procedure used for producing the MoSe2/La2O3 heterogeneous structure. Briefly, 1 mmol of (NH4)6Mo7O24·4H2O powdered was dissolved in 30 mL of distilled water. The same solution was subsequently supplemented with 2 mmol of selenium precursor (Se). Next, the mixture was agitated at room temperature for an hour to ensure homogeneity. To increase the limited number of active edge sites in MoSe2 material, La2O3 has been selected to augment these sites at different concentrations. Thus, the same solution was mixed with varying amounts of commercial La2O3 (1, 2, and 3 mmol) and stirred for two hours to create a homogenous mixture. After the homogeneous mixture was placed into a 100 mL autoclave, it was heated for 24 hours to 200°C. After hydrothermal heating, the resulting mixture was cooled to the ambient temperature. The synthesised black precipitates were finally rinsed with distilled water and ethanol. Finally, the resultant nanocomposite was then dried for 6 hours. For comparison, the additional three MoSe2/La2O3 heterojunctions were synthesised using the same method with mmole ratios of (NH4)6Mo7O24.4H2O and La2O3 of 1:1 (denote as M-L1), 1:2 (denote as M-L2), and 1:3 (denote as M-L3). Additionally, the same processes were employed for producing pure MoSe2 yet without the inclusion of La2O3.

Figure 1. Schematic illustration of MoSe2/La2O3 heterostructure fabrication.

Figure 1. Schematic illustration of MoSe2/La2O3 heterostructure fabrication.

2.3. Characterisation

To investigate the phase composition of the obtained materials, X-ray powder diffraction (XRD) was carried out using a SHIMADZU-MAXima-XRD-7000 diffractometer that was outfitted with Cu-Ka radiation. The topographical structure of the materials was studied using a field emission scanning electron microscope (FESEM-JSM-IT700HR). To investigate the chemical component of the produced nanocomposites, a FESEM-JSM-IT700HR was also used to perform energy-dispersive X-ray (EDX) spectra and elemental mapping studies. Raman spectra of the prepared materials were recorded using Senterra at a wavelength of 532 nm laser excitation. Transmission electron microscopy (TEM) (QUANTA FEG 250) was involved to study the composite material. Details regarding elemental composition and their valence were obtained utilising an XPS Thermo Fisher Scientific, USA (with monochromatic X-ray Al K-Alpha radiation from 10 to 1350 eV and a spot size of 400 mm at a pressure of 10–9 mbar).

2.4. Electrochemical measurements

Autolab (PGSTAT204) was used to conduct the electrochemistry measurements at room temperature. The disposable screen-printed carbon electrode (SPCE, Metrohm) is manufactured and constructed with a carbon serving as the counter electrode, an Ag/AgCl serving as the reference electrode, and a carbon serving as the working electrode. In all experiments, sulphuric acid (H2SO4, 0.5M) was used as the electrolyte. The equation ERHE = EAg/AgCl + 0.1976 + 0.056 pH was applied in this work to determine the potentials with regard to the reversible hydrogen electrode. Next, 400 mg of catalytic material was dispersed in 3 mL of a 1:3 water-to-propanol solution. After that, the resultant suspension was ultrasonicated for at least an hour to produce a uniform solution. Subsequently, 2 µL of the slurry ink was delicately dropped casted onto the surface working electrode (4 mm diameter), followed by the addition of 2 µL of Nafion solution (5%). All experiment measurements were performed in an N2-purge 0.5 M H2SO4 electrolyte at room temperature. Before assessing the electrocatalytic activity of the MoSe2-La2O3 catalyst, many cyclic voltammetry (CV) cycles were performed. Linear sweep voltammetry (LSV) was conducted at a scan rate of 1 mV/s throughout a potential range of 0.2 to −0.8 V vs. RHE. The electrochemical impedance spectroscopy (EIS) experiment was performed using a constant 400 mV voltage and frequencies ranging from 100 kHz to 1 Hz.

3. Results and discussion

3.1. General characterisation of nanomaterials

X-ray powder diffraction (XRD) patterns in (a) demonstrate the phase structure of the pristine La2O3 and MoSe2 nanostructures, as well as the hybrid M-L2 nanostructure. The XRD peaks obtained from pristine La2O3 match the JCPDS data for La2O3 Card No. 04-0856 (Mustofa et al. Citation2020). The typical peaks at 2θ of 15.7, 27.3, 27.9, 39.5, and 48.7 relate to the La2O3 planes (100), (002), (101), (102), and (211), respectively (Wang et al. Citation2014). In the case of MoSe2 material, all of the diffraction peaks observed in the XRD pattern of the produced MoSe2 are perfectly aligned with the 2H-phase MoSe2 (JCPDS Card No. 77-1715) (Yuan Liu, Zhu, and Chen Citation2015). The MoSe2 phase planes (002), (100), (103), and (110) are indicated by prominent peaks at 13.3, 31.6, 37.5, and 56.4, respectively. For the mixed M-L2 sample, peak diffractions were observed for pristine La2O3 and MoSe2. Raman spectroscopy was also utilised to examine the crystal structures of pure MoSe2, pristine La2O3, and the M-L2 hybrid structures. The Raman spectra were recorded via a 532 nm laser excitation wavelength. The La2O3 material exhibits the typical Raman peak at 337 and 445 cm−1, which matches to Egv1 vibration mode of the La-O, respectively (Bilel et al. Citation2021). In the resulting MoSe2, a prominent Raman peak is observed at 240 cm−1, corresponding to the out-of-plane mode (A1g), which is consistent with earlier results (Markeev et al. Citation2021; Nayak et al. Citation2017). The hybrid M-L2 nanostructures (top spectrum in (b)) contain the typical Raman peaks of MoSe2 and La2O3. In fact, the Raman spectrum of the M-L2 composite structure has no extra peaks when compared to its constituents, La2O3 nanoparticles and MoSe2 nanoflowers.

Figure 2. (a) XRD patterns; (b) Raman spectra of pure MoSe2, La2O3, and hybrid M-L2.

Figure 2. (a) XRD patterns; (b) Raman spectra of pure MoSe2, La2O3, and hybrid M-L2.

FESEM was used to investigate the structure and morphology of pure MoSe2, pristine La2O3, and M-L2 nanocomposite materials. (a,b) show SEM images of pure La2O3 in the form of dense nanoparticles or nanocrystals with an inconsistent shape. These nanoparticles have a wide variety of sizes, which can reach several hundred nanometers. In contrast, microscopy images of pristine MoSe2 at various magnifications are shown in (c,d). Pristine MoSe2 has the morphology of a microflowers-like structure with a large number of petals. (e,f) demonstrate the unique nanosheet shape of the M-L2 nanocomposite. The rose-like structures of MoSe2 are unaffected by the formation of the M-L2 nanocomposite, which has a more porous and opened structure with outstanding uniformity, indicating successful synthesis and formation of the nanocomposite. The nanosheet structure exhibits a high degree of uniformity in size and shape, indicating the high quality and crystallinity of the material. The edges of the nanosheets appear smooth and well defined. Furthermore, a significant number of edge nanosheets may provide a large number of active sites as well as a large specific surface area, which can boost electrocatalytic HER activity. The distribution and elemental composition of La, O, Mo, and Se were examined using energy dispersive spectroscopy (EDS) and element mapping. According to the EDS study shown in (a), the atomic ratio of Mo:Se is estimated to be 1:2, which is about the same as the atomic ratio of MoSe2. From (c–f), elemental mapping images reveal that the W, Se, La, and O components were distributed consistently, indicating a homogeneous distribution of La2O3 in the MoSe2 microflower structures. This leads to a close contact between MoSe2 and La2O3, which promotes a synergistic impact of M-L2 during the electrocatalytic process and could boost catalytic HER performance. TEM and HRTEM images were obtained to investigate the crystal structure of the M-L2 catalyst ((g–i)). The M-L2 HRTEM image reveals uniformly aligned lattice planes, which correspond to La2O3. (i) shows an enlarged view of the lattice distances. The interplanar lattice spacing matches the [100] plane of La2O3.

Figure 3. FESEM images of produced products at different magnifications: (a, b) pristine La2O3, (c, d) Pure MoSe2, and (e, f) M-L2 nanohybrid.

Figure 3. FESEM images of produced products at different magnifications: (a, b) pristine La2O3, (c, d) Pure MoSe2, and (e, f) M-L2 nanohybrid.

Figure 4. (a) EDX spectrum of the as prepared M-L2 sample. (b) SEM mage of elemental mapping. (c–f) EDX elemental mapping of La, O, Mo, and Se, respectively. (g) TEM image and (h and i) HRTEM along with magnified view of lattice plane of as prepared M-L2 sample.

Figure 4. (a) EDX spectrum of the as prepared M-L2 sample. (b) SEM mage of elemental mapping. (c–f) EDX elemental mapping of La, O, Mo, and Se, respectively. (g) TEM image and (h and i) HRTEM along with magnified view of lattice plane of as prepared M-L2 sample.

The elemental composition and valence state of the MoSe2, La2O3, and M-L2 nanohybrid surfaces were investigated using high-resolution X-ray photoelectron spectroscopy (XPS). (a) demonstrates the entire scan spectra of the MoSe2, La2O3, and M-L2 nanostructures. The XPS analysis of M-L2 revealed that La, O, Mo, and Se are the four primary elements detected in the synthesised materials, confirming the incorporation of these elements into the hybrid structure. The primary source of the C 1S signal at 284.6 eV is the carbon-conductive tape, which was employed for holding the sample during XPS measurements. Furthermore, the La 3d core level photoelectron spectra collected from pristine La2O3 nanoparticles and the M-L2 nanohybrid are depicted in (b). The most prominent XPS peaks of elemental lanthanum are the La 3d3/2 and La 3d5/2 peaks, which are attributed to 851.6 and 834.78 eV, respectively (Agrawal et al. Citation2020; Sarkar et al. Citation2016). These peaks shifted by 0.3 eV to high-binding energy after being combined with MoSe2, as shown in the M-L2 nanohybrid spectrum. Additionally, (c) demonstrates the XPS spectrum of O 1s core level for La2O3 nanoparticles and M-L2 nanohybrid in which the binding energy at 530 eV is attributed to the La – O bond (Karthikeyan, Selvapandiyan, and Sankar Citation2022). Moreover, the XPS spectrum of Mo 3d derived from pure MoSe2 and M-L2 nanohybrid displays four primary peaks, as shown in (d). The oxidation state of Mo4+ is illustrated by two peaks at 228.6 and 231.7 eV, which agree to Mo 3d5/2 and Mo 3d3/2, respectively (Alahmadi and Sami Citation2023; J. Yang et al. Citation2016). In contrast to pristine MoSe2, the electronic structure of MoSe2 in the nanohybrid (M-L2 spectra) was changed, as demonstrated by the 0.8 eV shift of the Mo (IV) 3d5/2 peak. Additionally, two peaks at 231.2 and 235.4 eV match Mo 3d5/2 and Mo 3d3/2, indicating the +6 valence state of Mo (Zhang et al. Citation2015). In the case of M-L2 spectrum, Mo (IV) 3d5/2 binding energy shift by 0.8 eV. On the other hand, the hexavalent state peak at 235 is significantly pronounced in the M-L2 spectrum but is barely apparent in the pristine MoSe2 spectrum. This peak could point to a stronger interaction of components in the M-L2 mixture. Similarly, (e) shows the high-resolution XPS of Se 3d. The two identical peaks correspond to 3d5/2 (55 eV) and Se 3d3/2 (56.3 eV). This implies that the element selenium is in a −2 valence state (Tang et al. Citation2014; Vikraman et al. Citation2021). As indicated by (e), the peaks (Se 3d5/2 and Se 3d3/2) in the M-L2 spectrum shift by 0.5 eV to low-binding energy, proving that the crystal structure of MoSe2 in the nanohybrid has been changed. The findings support the successful production of the hybrid composite.

Figure 5. XPS results of pristine MoSe2, La2O3, and an M-L2 nanocomposite sample. (a) Scan spectra for the three materials. High-resolution XPS spectra of (b) La 3d, (c) O 1s, (d) Mo 3d, and (e) Se 3d.

Figure 5. XPS results of pristine MoSe2, La2O3, and an M-L2 nanocomposite sample. (a) Scan spectra for the three materials. High-resolution XPS spectra of (b) La 3d, (c) O 1s, (d) Mo 3d, and (e) Se 3d.

3.2. Electrocatalytic performance

In the following section, the electrochemical performance of bare and modified SPCE electrodes made from pristine MoSe2 and pristine La2O3 are compared to that of modified SPCE electrodes formed from different nanocomposites (M-L1, M-L2, and M-L3). Besides, the electrochemical characteristics of Pt were employed as a control experiment. First, to verify the fabricated electrode activity, a few cyclic voltammetry (CV) cycles were performed before measuring HER performance in nitrogen-saturated 0.5 M H2SO4. First, the HER catalytic activity of as-fabricated nanocatalysts modified SPCE was examined using LSV at room temperature and a scan rate of 1 m/Vs as shown in (a). The prepared catalysts are labelled as MoSe2/SPCE, La2O3/SPCE, M-L1/SPCE, M-L2/SPCE and M-L3/SPCE. In acid electrolytes, MoSe2/SPCE and La2O3/SPCE materials do not display significant catalytic activity for HER. The effect of varying the La2O3 catalyst's precursor concentration on MoSe2 activity was investigated. When the molar ratio of La2O3 is raised from 1 to 2 mmol, the M-L2/SPCE electrode exhibits enhanced current density and a decreased overpotential ((a)), indicating improved HER activity. The M-L2/SPCE hybrid exhibited the highest HER catalytic activity with a current density of 10 mA cm−2 at an overpotential of around 400 mV. The presence of active sites generated by metal oxides at Mo and Se atoms may improve the nanocomposite's catalytic activity. However, the polarisation curve and overpotential show a reduction in activity after adding 3 mmol of La2O3 (M-L3) catalysts. Noteworthy is the fact that the amount of precursor used can significantly alter the phases of nanostructures, resulting in reduced overpotential and Rct. This indicates that selecting appropriate precursor concentrations can improve the catalytic performance in HER applications. shows that MoSe2/La2O3 exhibits better HER activity than previously described TMDC MoSe2-based catalysts. It has long been recognised that the quantity of catalyst on the supporting electrode influences the performance of HER. Additionally, according to a recent study, Tafel slopes are strongly reliant on catalyst mass and fluctuate with the mass loading (Anantharaj and Kundu Citation2019).

Figure 6. (a) LSV polarisation curves. (b) The corresponding Tafel plots obtained from polarisation curves for the bare carbon, MoSe2, La2O3, M-L1, M-L2 and M-L3.

Figure 6. (a) LSV polarisation curves. (b) The corresponding Tafel plots obtained from polarisation curves for the bare carbon, MoSe2, La2O3, M-L1, M-L2 and M-L3.

Table 1. A comparison of HER catalytic performance of MoSe2 materials.

The Tafel slope of each electrocatalyst is illustrated in (b). The Tafel equation (k = b log (j) + a) can be applied to derive Tafel slopes, where g is the overpotential, b is the Tafel slope, and j is the current density. MoSe2, having Tafel slopes of 160 mV/dec, is a more potent HER catalyst than pristine La2O3 and bare carbon, which possessed Tafel slopes of 171 and 198 mV/dec, respectively. The hybrid M-L2 nanocomposite has a superior catalytic response as compared to its predecessors, with a value of 66.8 mV/dec. The synergistic impact of integrating different material components (MoSe2 and La2O3) may be responsible for the improved HER efficiency and water-splitting ability of nanocomposite materials. As a result, the nanocomposites have higher HER activity compared to their pure components. Based on the lowest Tafel slope value across electrocatalysts, the electrochemical desorption process is expected to be the limiting step for M-L2/SPCE. We may deduce from the Tafel slope values that La2O3 is a promising catalyst for improving HER performance. Additionally, our M-L2@SPCE combination exhibits better electrocatalytic activity and more efficient charge transfer kinetics than other MoSe2 materials previously examined, including tube-like MoSe2/CoSe2 microcages (73 mV/dec) (Mu et al. Citation2016), hetero-dimensional hybrid MoSe2 (72 mV/dec) (B. Mao et al. Citation2017), and Ni-doped MoSe2 (83 mV/dec) (Y. Yang et al. Citation2020).

To determine the electrical resistance of the catalysts, further investigation was conducted on the as-synthesis materials MoSe2, La2O3, M-L1, M-L2, and M-L3 using electrochemical impedance spectroscopy (EIS). All electrocatalysts were measured using EIS in 0.5 M H2SO4. (a) depicts the Nyquist plots produced from EIS measurements for all nanocatalysts modified SPCE. The diameter of a semicircle was utilised to calculate the catalyst charge transfer resistance R(ct). Among all samples, M-L2/SPCE sample exhibits the lowest Rct (∼250 Ω) and the highest electrical conductivity, indicating that electrons are moving rapidly across the electrocatalyst-electrolyte interface. The EIS data supports the earlier results that M-L2 was the most effective electrocatalyst for the HER since it showed a rapid rate of reaction. Furthermore, durability is a significant factor to consider when evaluating electrocatalytic HER performance, whether the catalyst will be used for industrial or commercial uses. The long-term stability of the optimised electrode towards HER was assessed by testing ideal catalysts (M-L2/SPCE) over 1000 CV cycles in 0.5 M H2SO4, as shown in (b). This result suggests that the M-L2 catalyst demonstrated stability during HER performance. With a slight drop in the M-L2 electrocatalyst's catalytic activity, this data implies that the M-L2 catalyst is durable during HER performance. It is likely that during the HER event, the hybrid M-L2 mixture delaminated from the SPCE's surface, leading to the observed drop in HER catalytic activity (Benck et al. Citation2012).

Figure 7. (a) Electrochemical impedance spectra of the bare carbon, MoSe2, La2O3, M-L1, M-L2 and M-L3. (b) Stability results for the M-L2 sample before and after 1000 CV cycles in 0.5 M H2SO4 under N2 atmosphere.

Figure 7. (a) Electrochemical impedance spectra of the bare carbon, MoSe2, La2O3, M-L1, M-L2 and M-L3. (b) Stability results for the M-L2 sample before and after 1000 CV cycles in 0.5 M H2SO4 under N2 atmosphere.

4. Conclusion

In conclusion, a hybrid nanocomposite of MoSe2-La2O3 with high catalytic activity toward HER can be prepared using one-step hydrothermal techniques. Metal oxide (La2O3) has significantly improved MoSe2 HER performance by increasing the possibility of H+ adsorption at active edge sites. It has been proven that a small quantity of La2O3 is critical for increasing the activity of MoSe2 even though La2O3 and MoSe2 alone have low activity. In acidic conditions, among several developed nanocomposites, the M-L2 nanocomposite sample demonstrates considerably better electrocatalytic capabilities for HER, with a low overpotential of around 400 mV at 10 mA/cm2, which is significantly lower than that of the pure MoSe2. Finally, these catalysts are stable, making them excellent for electrocatalytic hydrogen production.

Disclosure statement

No potential conflict of interest was reported by the author(s).

References

  • Agrawal, K.S., V.N. Barhate, V.S. Patil, L.S. Patil, and A.M. Mahajan. 2020. “Plasma-enhanced Atomic Layer-Deposited La2O3 Ultra-Thin Films on Si and 6H–SiC: A Comparative Study.” Applied Physics A 126 (8): 1–10. https://doi.org/10.1007/s00339-020-03684-2.
  • Alahmadi, M., and S. Ben Aoun. 2024. “A Single-Step Synthesis of Aluminum-Doped WS2 for Oxygen Evolution Reaction with Loading Influence on Catalytic Performance.” Journal of Saudi Chemical Society 28 (3): 101845. https://doi.org/10.1016/j.jscs.2024.101845.
  • Alahmadi, M., F. Mahvash, T. Szkopek, and M. Siaj. 2021. “A Two-Step Chemical Vapor Deposition Process for the Growth of Continuous Vertical Heterostructure WSe 2/h-BN and its Optical Properties.” RSC Advances 11 (28): 16962–16969. https://doi.org/10.1039/D1RA02523F.
  • Alahmadi, M., and B.E.N. Sami. 2023. “One-Pot In-Situ Hydrothermal Synthesis of VSe2/MoSe2 Nanocomposite for Enhanced Hydrogen Evolution Reaction.” Arabian Journal of Chemistry 16 (7): 104846.
  • Anantharaj, S., and S. Kundu. 2019. “Do the Evaluation Parameters Reflect Intrinsic Activity of Electrocatalysts in Electrochemical Water Splitting?” ACS Energy Letters 4 (6): 1260–1264. https://doi.org/10.1021/acsenergylett.9b00686.
  • Benck, J.D., Z. Chen, L.Y. Kuritzky, A.J. Forman, and T.F. Jaramillo. 2012. “Amorphous Molybdenum Sulfide Catalysts for Electrochemical Hydrogen Production: Insights Into the Origin of Their Catalytic Activity.” Acs Catalysis 2 (9): 1916–1923. https://doi.org/10.1021/cs300451q.
  • Bilel, C., R. Jbeli, I. Ben Jemaa, Y. Dabaki, M. Alzaid, F. Saadallah, M. Bouaicha, and M. Amlouk. 2021. “Synthesis and Physical Characterization of Ni-Doped La2O3 for Photocatytic Application Under Sunlight.” Journal of Materials Science: Materials in Electronics 32 (5): 5415–5426. https://doi.org/10.1007/s10854-021-05264-3.
  • Chen, X., Y. Qiu, G. Liu, W. Zheng, W. Feng, F. Gao, W. Cao, Y. Fu, W. Hu, and P. Hu. 2017. “Tuning Electrochemical Catalytic Activity of Defective 2D Terrace MoSe2 Heterogeneous Catalyst via Cobalt Doping.” Journal of Materials Chemistry A 5 (22): 11357–11363. https://doi.org/10.1039/C7TA02327H.
  • Dai, T., Z. Zhou, H. Xiao, Y. Luo, Y. Xu, and X. Wang. 2022. “In Situ Growth of NiSe2-MoSe2 Heterostructures on Graphene Nanosheets as High-Performance Electrocatalyst for Hydrogen Evolution Reaction.” Catalysts 12 (7): 701. https://doi.org/10.3390/catal12070701.
  • Deng, S., C. Ai, M. Luo, B. Liu, Y. Zhang, Y. Li, S. Lin, G. Pan, Q. Xiong, Q. Liu, X. Wang, X. Xia, and J. Tu. 2019. “Coupled Biphase (1T-2H)-MoSe2 on Mold Spore Carbon for Advanced Hydrogen Evolution Reaction.” Small 15 (30): 1901796. https://doi.org/10.1002/smll.201901796.
  • Dresselhaus, M.S., and I.L. Thomas. 2001. “Alternative Energy Technologies.” Nature 414 (6861): 332–337. https://doi.org/10.1038/35104599.
  • Hwang, Y., and N. Shin. 2021. “Colloidal Synthesis of MoSe2/WSe2 Heterostructure Nanoflowers via Two-Step Growth.” Materials 14 (23): 7294. https://doi.org/10.3390/ma14237294.
  • Karthikeyan, S., M. Selvapandiyan, and A. Sankar. 2022. “Electrochemical Performance of Reduced Graphene Oxide (rGO) Decorated Lanthanum Oxide (La2O3) Composite Nanostructure as Asymmetric Supercapacitors.” Inorganic Chemistry Communications 139: 109331. https://doi.org/10.1016/j.inoche.2022.109331.
  • Laursen, A.B., S. Kegnæs, S. Dahl, and I. Chorkendorff. 2012. “Molybdenum Sulfides—Efficient and Viable Materials for Electro-and Photoelectrocatalytic Hydrogen Evolution.” Energy & Environmental Science 5 (2): 5577–5591. https://doi.org/10.1039/c2ee02618j.
  • Lei, Z., S. Xu, and P. Wu. 2016. “Ultra-thin and Porous MoSe 2 Nanosheets: Facile Preparation and Enhanced Electrocatalytic Activity Towards the Hydrogen Evolution Reaction.” Physical Chemistry Chemical Physics 18 (1): 70–74. https://doi.org/10.1039/C5CP06483J.
  • Li, N., J. Wu, Y. Lu, Z. Zhao, H. Zhang, X. Li, Y.-Z. Zheng, and X. Tao. 2018. “Stable Multiphasic 1 T/2H MoSe2 Nanosheets Integrated with 1D Sulfide Semiconductor for Drastically Enhanced Visible-Light Photocatalytic Hydrogen Evolution.” Applied Catalysis B: Environmental 238: 27–37. https://doi.org/10.1016/j.apcatb.2018.07.002.
  • Liu, X., Y. Jiang, J. Huang, W. Zhong, B. He, P. Jin, and Y. Chen. 2023. “Bifunctional PdPt Bimetallenes for Formate Oxidation-Boosted Water Electrolysis.” Carbon Energy 5 (12): e367. https://doi.org/10.1002/cey2.367.
  • Liu, Y., L. Ren, Z. Zhang, X. Qi, H. Li, and J. Zhong. 2016. “3D Binder-Free MoSe2 Nanosheets/Carbon Cloth Electrodes for Efficient and Stable Hydrogen Evolution Prepared by Simple Electrophoresis Deposition Strategy.” Scientific Reports 6 (1): 22516. https://doi.org/10.1038/srep22516.
  • Liu, Y., M. Zhu, and D. Chen. 2015. “Sheet-like MoSe 2/C Composites with Enhanced Li-ion Storage Properties.” Journal of Materials Chemistry A 3 (22): 11857–11862. https://doi.org/10.1039/C5TA02100F.
  • Mabuea, B.P., E. Erasmus, and H.C. Swart. 2023. “Molybdenum-Tungsten Carbides Based Electrocatalysts for Hydrogen Evolution Reaction.” International Journal of Sustainable Energy 42 (1): 91–102. https://doi.org/10.1080/14786451.2023.2176700.
  • Manzeli, S., D. Ovchinnikov, D. Pasquier, O.V. Yazyev, and A. Kis. 2017. “2D Transition Metal Dichalcogenides.” Nature Reviews Materials 2 (8): 1–15. https://doi.org/10.1038/natrevmats.2017.33.
  • Mao, B., T. Bao, J. Yu, L. Zheng, J. Qin, W. Yin, and M. Cao. 2017. “One-pot Synthesis of MoSe2 Hetero-Dimensional Hybrid Self-Assembled by Nanodots and Nanosheets for Electrocatalytic Hydrogen Evolution and Photothermal Therapy.” Nano Research 10 (8): 2667–2682. https://doi.org/10.1007/s12274-017-1469-7.
  • Mao, Z., C. Wang, H. Lu, K. Tang, Q. Li, C. Yan, and X. Wang. 2020. “Boron-Modified Electron Transfer in Metallic 1 T MoSe2 for Enhanced Inherent Activity on Per-Catalytic Site Toward Hydrogen Evolution.” Advanced Materials Interfaces 7 (1): 1901560. https://doi.org/10.1002/admi.201901560.
  • Markeev, P.A., E. Najafidehaghani, Z. Gan, K. Sotthewes, A. George, A. Turchanin, and M.P. de Jong. 2021. “Energy-level Alignment at Interfaces Between Transition-Metal Dichalcogenide Monolayers and Metal Electrodes Studied with Kelvin Probe Force Microscopy.” The Journal of Physical Chemistry C 125 (24): 13551–13559. https://doi.org/10.1021/acs.jpcc.1c01612.
  • Masurkar, N., N.K. Thangavel, and L.M.R. Arava. 2018. “CVD-grown MoSe2 Nanoflowers with Dual Active Sites for Efficient Electrochemical Hydrogen Evolution Reaction.” ACS Applied Materials & Interfaces 10 (33): 27771–27779. https://doi.org/10.1021/acsami.8b07489.
  • Mu, C.H., H.X. Qi, Y.Q. Song, Z.P. Liu, L.X. Ji, J.G. Deng, Y.B. Liao, and F. Scarpa. 2016. “One-pot Synthesis of Nanosheet-Assembled Hierarchical MoSe2/CoSe2 Microcages for the Enhanced Performance of Electrocatalytic Hydrogen Evolution.” RSC Advances 6 (1): 23–30. https://doi.org/10.1039/C5RA21638A.
  • Mustofa, K., Y. Yulizar, A. Saefumillah, and D.O.B. Apriandanu. 2020. “La2O3 Nanoparticles Formation Using Nothopanax Scutellarium Leaf Extract in Two-Phase System and Photocatalytic Activity Under UV Light Irradiation.” Dans IOP Conference Series: Materials Science and Engineering, 12018. 902 vols. IOP Publishing.
  • Nayak, P.K., Y. Horbatenko, S. Ahn, G. Kim, J.-U. Lee, K.Y. Ma, A-Rang Jang, et al. 2017. “Probing Evolution of Twist-Angle-Dependent Interlayer Excitons in MoSe2/WSe2 van der Waals Heterostructures.” ACS Nano 11 (4): 4041–4050. https://doi.org/10.1021/acsnano.7b00640.
  • Nicoletti, G., N. Arcuri, G. Nicoletti, and R. Bruno. 2015. “A Technical and Environmental Comparison Between Hydrogen and Some Fossil Fuels.” Energy Conversion and Management 89: 205–213. https://doi.org/10.1016/j.enconman.2014.09.057.
  • Ren, X., Q. Ma, P. Ren, and Y. Wang. 2018. “Synthesis of Nitrogen-Doped MoSe2 Nanosheets with Enhanced Electrocatalytic Activity for Hydrogen Evolution Reaction.” International Journal of Hydrogen Energy 43 (32): 15275–15280. https://doi.org/10.1016/j.ijhydene.2018.06.122.
  • Sarkar, P.K., M. Prajapat, A. Barman, S. Bhattacharjee, and A. Roy. 2016. “Multilevel Resistance State of Cu/La2O3/Pt Forming-Free Switching Devices.” Journal of Materials Science 51 (9): 4411–4418. https://doi.org/10.1007/s10853-016-9753-6.
  • Sharma, S., S. Agarwal, and A. Jain. 2021. “Significance of Hydrogen as Economic and Environmentally Friendly Fuel.” Energies 14 (21): 7389. https://doi.org/10.3390/en14217389.
  • Shinagawa, T., and K. Takanabe. 2017. “Towards Versatile and Sustainable Hydrogen Production Through Electrocatalytic Water Splitting: Electrolyte Engineering.” ChemSusChem 10 (7): 1318–1336. https://doi.org/10.1002/cssc.201601583.
  • Tang, H., K. Dou, C.-C. Kaun, Q. Kuang, and S. Yang. 2014. “MoSe 2 Nanosheets and Their Graphene Hybrids: Synthesis, Characterization and Hydrogen Evolution Reaction Studies.” Journal of Materials Chemistry A 2 (2): 360–364. https://doi.org/10.1039/C3TA13584E.
  • Tran, X.T., S. Poorahong, and M. Siaj. 2017. “One-pot Hydrothermal Synthesis and Selective Etching Method of a Porous MoSe 2 Sand Rose-Like Structure for Electrocatalytic Hydrogen Evolution Reaction.” RSC Advances 7 (82): 52345–52351. https://doi.org/10.1039/C7RA10001A.
  • Vikraman, D., S. Hussain, K. Akbar, K. Adaikalam, S.H. Lee, S.-H. Chun, Jongwan Jung, Hyun-Seok Kim, and H.J. Park. 2018. “Facile Synthesis of Molybdenum Diselenide Layers for High-Performance Hydrogen Evolution Electrocatalysts.” ACS Omega 3 (5): 5799–5807. https://doi.org/10.1021/acsomega.8b00459.
  • Vikraman, D., S. Hussain, S.A. Patil, L. Truong, A.A. Arbab, S.H. Jeong, S.-H. Chun, J. Jung, and H.-S. Kim. 2021. “Engineering MoSe2/WS2 Hybrids to Replace the Scarce Platinum Electrode for Hydrogen Evolution Reactions and dye-Sensitized Solar Cells.” ACS Applied Materials & Interfaces 13 (4): 5061–5072. https://doi.org/10.1021/acsami.0c19890.
  • Wang, K., Y. Wu, H. Li, M. Li, F. Guan, and H. Fan. 2014. “A Hybrid Antioxidizing and Antibacterial Material Based on Ag–La2O3 Nanocomposites.” Journal of Inorganic Biochemistry 141: 36–42. https://doi.org/10.1016/j.jinorgbio.2014.08.009.
  • Wu, X., H. Zhang, J. Zhang, and X.W. Lou. 2021. “Recent Advances on Transition Metal Dichalcogenides for Electrochemical Energy Conversion.” Advanced Materials 33 (38): 2008376. https://doi.org/10.1002/adma.202008376.
  • Yan, Y., C. Wang, Z. Huang, J. Fu, Z. Lin, X. Zhang, J. Ma, and J. Shen. 2021. “Highly Efficient and Robust Catalysts for the Hydrogen Evolution Reaction by Surface Nano Engineering of Metallic Glass.” Journal of Materials Chemistry A 9 (9): 5415–5424. https://doi.org/10.1039/D0TA10235K.
  • Yang, J., F. Zhang, X. Wang, D. He, G. Wu, Q. Yang, X. Hong, Y. Wu, and Y. Li. 2016. “Porous Molybdenum Phosphide Nano-Octahedrons Derived from Confined Phosphorization in UIO-66 for Efficient Hydrogen Evolution.” Angewandte Chemie 128 (41): 13046–13050. https://doi.org/10.1002/ange.201604315.
  • Yang, Y., X. Zhao, H. Mao, R. Ning, X. Zheng, J. Sui, and W. Cai. 2020. “Nickel-doped MoSe2 Nanosheets with Ni–Se Bond for Alkaline Electrocatalytic Hydrogen Evolution.” International Journal of Hydrogen Energy 45 (18): 10724–10728. https://doi.org/10.1016/j.ijhydene.2019.12.212.
  • Zhang, K., C. Li, Y. Zhao, X. Yu, and Y. Chen. 2015. “Porous One-Dimensional Mo 2 C–Amorphous Carbon Composites: High-Efficient and Durable Electrocatalysts for Hydrogen Generation.” Physical Chemistry Chemical Physics 17 (25): 16609–16614. https://doi.org/10.1039/C5CP02028J.
  • Zou, X., and Y. Zhang. 2015. “Noble Metal-Free Hydrogen Evolution Catalysts for Water Splitting.” Chemical Society Reviews 44 (15): 5148–5180. https://doi.org/10.1039/C4CS00448E.