Next Article in Journal
GPpred: A Novel Sequence-Based Tool for Predicting Glutamic Proteases Using Optimized Hybrid Encodings
Next Article in Special Issue
Effect of Calcination-Induced Oxidation on the Photocatalytic H2 Production Performance of Cubic Cu2O/CuO Composite Photocatalysts
Previous Article in Journal
One-Pot Synthesis of Highly Dispersed VO2 on g-C3N4 Nanomeshes for Advanced Oxidation
Previous Article in Special Issue
Aqueous Phase Reforming by Platinum Catalysts: Effect of Particle Size and Carbon Support
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Synthesis of La-MoS2 and Its Catalytic Activity for Improved Hydrogen Evolution Reaction

1
Department of Chemistry, Medi-Caps University, Indore 453331, India
2
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
3
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(12), 893; https://doi.org/10.3390/catal14120893
Submission received: 31 October 2024 / Revised: 27 November 2024 / Accepted: 3 December 2024 / Published: 5 December 2024

Abstract

:
Herein, we report the synthesis and characterization of lanthanum-doped MoS2 (La-MoS2) via a hydrothermal route. The synthesized La-MoS2 was characterized using powder X-ray diffraction (PXRD), scanning electron microscopy (SEM), and energy-dispersive X-ray (EDX) techniques. The band gap of La-MoS2 was observed to be 1.68 eV, compared to 1.80 eV for synthesized MoS2. In the photoluminescence (PL) spectra, a decrease in the intensity was observed for La-MoS2 compared to MoS2, which suggests that due to doping with charged La3+, separation increases. The as-synthesized MoS2 and La-MoS2 were used for photocatalytic hydrogen evolution reactions (HERs), exhibiting 928 µmol·g−1 evolution of H2 in five hours for a 10 mg dose of La-MoS2, compared to 612 µmol·g−1 for MoS2. A 50 mg mass of the catalyst (La-MoS2) exhibited enhanced H2 production of 1670 µmol·g−1 after five hours. The higher rate of the HER for La-MoS2 is because of doping with La3+. The photocatalytic hydrogen evolution performance of La-MoS2 was also evaluated for different doses of La-MoS2 exhibiting reusability up to the fourth cycle, showing potential applications of La-MoS2 in hydrogen evolution reactions. Mechanistic aspects of the HER on the surface of La-MoS2 have also been discussed.

1. Introduction

In the past few decades, rapid expansion of industries and a growing global population have posed significant challenges for humanity in meeting its energy needs [1]. Over the past few years, energy demand has been risen exponentially [2]. At present, the primary source of energy is fossil fuels, which are rapidly depleting and have harmful environmental effects when utilized [2]. To overcome this challenge, researchers and scientists are working on the development of sustainable technologies designed to reduce environmental impact. Solar energy has been proved to be one of the most efficient and green of all energy sources. The sun provides an enormous amount of solar energy, around 1022 joules, which can contribute to the development of next generation energy technologies [3]. Thus, it is essential to utilize solar energy for the development of green and renewable energy technology for the future. In connection with this, hydrogen energy is considered one of the most efficient energy technologies to fulfill energy requirements [4,5]. Photocatalytic hydrogen generation by the use of solar energy on the surface of a catalyst (semiconductor) is a very propitious alternative for resolving the present energy and environmental crisis [6,7,8]. Hydrogen fuel provides a high yield of energy upon combustion, i.e., 122 kJ/mol, which is far better than any other fossil fuel (gasoline, coal, etc.) [9,10]. The process of hydrogen generation by photocatalysis is environmentally benign, producing no harmful byproducts [11]. The use of photocatalytic materials for the generation of hydrogen was demonstrated for the first time in 1972 [12]. Thereafter, a great number of semiconductor materials have been designed and demonstrated for the efficient generation of hydrogen [13,14,15]. A series of MoS2-TiO2 photocatalysts have been designed by Zhu et al. [16], using a ball-milling process. Due to the introduction of MoS2, a decrease in the recombination of photogenerated charge carriers was observed, enhancing the photocatalytic performance compared to pure TiO2. The rate of the HER was enhanced to 150.7 μmol/h for the 4.0 wt% MoS2-TiO2 catalyst. Kumar et al. designed defect-rich MoS2 nanosheets and decorated them with nitrogen-doped ZnO nanorod composites [17]. For this photocatalyst, optimal hydrogen evolution (17.3 mmol h−1 g−1) was recorded for 15 wt% defect-rich MoS2 nanosheets coated on N-ZnO. The results also demonstrated that a defect-induced interfacial contact region developed on encapsulation of defect-rich MoS2 across N-ZnO, resulting in a composite that further increased the photocatalytic activity of the designed semiconductor system.
Doping of MoS2 with a metal or nonmetal may generates defects along with alteration of the optical bandgap. For a Co-doped MoS2/g-C3N4 composite, a very good photocatalytic reduction rate of water has been reported by Hu et al. [18] with a 0.31 mmol g−1h−1 rate of H2 evolution due to the generation of an edge-enriched 1T phase as a result of Co doping. Additionally, the rate of photocatalysis can be enhanced by depositing a metal on MoS2. With metal deposition on MoS2, the interfacial charge transfer rate is enhanced along with the generation of a local electric field via the Schottky junction. It may also broaden the absorption spectrum to the near-infrared region. Scientists have studied the deposition of several metals, viz., copper [19], silver [20] gold [21], platinum [22], and palladium [23], on MoS2, showing the creation of plasmonic photocatalysts absorbing strong visible light along with size- and shape-dependent surface plasmon resonance (SPR) effects.
Metal dichalcogenide compounds have received a great deal of attention because of their excellent optoelectronic properties. Metal dichalcogenide compounds including metal sulfides (such as MoS₂), selenides, and tellurides exhibit tunable band gaps and high surface activity, making them suitable for semiconductors, photocatalysis, and energy storage. Metal chalcogenides, especially in two-dimensional forms, have gained attention for hydrogen evolution reactions (HERs) and other electrochemical applications, as they offer excellent electron mobility and catalytic efficiency. MoS₂ is a 2D layered material characterized by a large surface area, tunable electronic properties, and structural flexibility, which make it suitable for catalytic processes. Its catalytic activity in HERs is primarily driven by the active edge sites of the MoS₂ layers, where sulfur vacancies and exposed Mo atoms enhance electron transfer and improve reaction kinetics. One key property of MoS₂ is its relatively low overpotential, especially when engineered with additional defects or doped with metals such as Ni or Co, which introduce more active sites and increase conductivity. The basal plane, which is generally inert, can also be activated through exfoliation, creating more active edge sites. Additionally, MoS₂ has excellent stability at acidic and neutral pH values, making it a durable option for water-splitting applications. Recent research focuses on synthesizing MoS₂ heterostructures with materials such as graphene or carbon nanotubes, further enhancing conductivity and stability. These modifications make MoS₂-based catalysts a promising alternative to noble metals in efficient and cost-effective H₂ production. MoS2, a layered transition metal dichalcogenide, is generated by the stacking of S–Mo–S atomic layers that are held together by van der Waals forces present among them [24]. When MoS2 is changed from bulk to the nanoscale, it exhibits unique physiochemical and optical properties and becomes a potential candidate for photocatalysis [25]. Bulk MoS2 is optically inactive with an indirect bandgap of 1.29 eV; it also demonstrates a low photoluminescence response [26]. However, when the thickness of bulk MoS2 is reduced to a single layer or a few layers, it exhibits a peak at 1.8–1.9 eV for the direct band gap [27]. In MoS2, speedy/fast transfer of charge carriers occurs when visible light strikes its surface because of the narrow band gap; therefore, it is a promising and valuable photocatalyst candidate. However, the fast rate of recombination of charge carriers limits it applications in photocatalysis reactions [27,28].
Herein, we report the fabrication of a lanthanum-doped MoS2 photocatalyst for a hydrogen evolution reaction. La-MoS2 was prepared using a hydrothermal method, and its photocatalytic activity was examined for a hydrogen evolution reaction under visible light irradiation. The proposed photocatalyst La-MoS2 also demonstrated decent stability for hydrogen evolution under visible light irradiation.

2. Results and Discussion

2.1. Material Characterization

To characterize the generated phase, the powder X-ray diffraction patterns of pure MoS2 and La-MoS2 were recorded and are shown in Figure 1. Four distinctive peaks at 2θ values of 13.1°, 32.4°, 35.8°, and 57.9° were observed in the samples, corresponding to the (002), (100), (103), and (110) planes of MoS2. The PXRD patterns confirm the presence of a hexagonal crystal structure (2H-MoS2), which is in accordance with the existing literature (JCPDS No.: 0037-1492). On comparing the PXRD patterns of both the samples of pure MoS2 and La-MoS2, it was confirmed that almost the same peak position could be observed in both cases, with a slight shift in La-MoS2. The slight shift in the peak position after doping may be due to the larger ionic radius of lanthanum (La3+) as compared to molybdenum (Mo4+), which results in the expansion of the crystalline lattice and is consistent with Bragg’s equation [29]. In the PXRD pattern of La-MoS2, a higher peak intensity could be observed as compared to MoS2; this increase in the peak intensity may be ascribed to the different electron densities of La3+ (dopant) and MoS2 (host), which primarily relate to the scattering factor, structure factor, etc. In the case of La-MoS2, the increase in the structure factor may be attributed to increased crystallite size, which results in an increase in the intensity of PXRD peaks on doping with La3+ [30]. The absence of peaks for elemental La3+ confirmed that the crystal structure of MoS2 had not been changed due to doping. To calculate the particle sizes of both MoS2 and La-MoS2, the Debye–Scherrer formula (D = K λ/β cosθ, where K = 0.89, λ = 0.154, D is the average crystalline size, is β is full width half maxima (FWHM) and θ is the Bragg angle) has been used in a previous study. The average crystallite sizes of MoS2 and La-Mos2 samples have been calculated as 17.5 and 21.2 nm, respectively, showing an increase in the crystallite size of MoS2 on doping with La3+ [31].
Figure 2 presents surface morphology of synthesized MoS2 (Figure 2a) and La-MoS2 (Figure 2c). For both of the samples, spherical morphologies were observed, with an average diameter of approximately one nanometer (Figure 2b,d). The average sizes of the MoS2 and La-MoS2 particles were found to be 648.26 nm and 676.01 nm, respectively, using particle size distribution curves (Figure 2b,d).
From the FESEM images, it appeared that the surfaces of the MoS2 and La-MoS2 microspheres were made up of nanoflakes. Hollow microspheres of MoS2 have also been reported by Afanasiev and Bezverkhy [32], who produced them via a heat treatment method, and by Chen et al. [33], who produced them via a direct sulfidization route; these results are similar to ours. Moreover, in these synthesized microspheres, numerous thin-stretched, folding flakes have also been observed, which are helpful in hydrogen evolution reactions.
When these active sites are exposed to light, they assist in the transportation of charge carriers along with participating in oxidation and reduction reactions in the course of the photocatalytic procedure [34]. EDX spectra of MoS2 and La-MoS2 sub-microspheres are shown in Figure S1. As shown in the EDX spectrum of MoS2, two signals indicating the elements Mo and S in the elemental composition were observed, while in the EDX spectrum of La-MoS2, signals for the element La were also detected. The obtained results again indicated the phase purity of the synthesized samples.
In order to check the distribution of the elements in the samples, elemental mapping was also performed. Images of element mapping of MoS2 revealed that the sample contained Mo and S as the main elements, and both elements were uniformly distributed in the sample (Figure 3a–c). Images of elemental mapping along with the corresponding overlay and FESEM images for La-MoS2 sample are presented in Figure 3d–g, showing the presence of the element La in the MoS2 sample. It is evident from the images that the element La was distributed in lower amounts than Mo and S.
To calculate the optical bandgaps of the prepared samples, UV–visible spectra were been recorded for both the as-prepared MoS2 and La-MoS2 samples. Figure 4a depicts the UV–visible spectra of both the samples. The results of UV–visible spectroscopy indicated that there was a significant shift in the absorption band towards visible region in the case of La-MoS2. To calculate the band gap energy of both the samples, a Tauc plot was created using the formula (αhυ)2 = A (hυ − Eg). A plot of (αhν)2 against hν is shown in Figure 4b. The band gap of as-prepared MoS2 was calculated as 1.80 eV, similar to the band gap of MoS2 monolayer material [35], whereas a significant reduction in the band gap (1.68 eV) was noticed after doping with La3+, which further affirms that upon doping with La3+, the visible light absorption property of the synthesized photocatalyst was enhanced.
PL spectra of MoS2 and La-MoS2 are shown in Figure S2. The main role of photoluminescence (PL) spectroscopy is to investigate migration, electron transfer efficiency, and electron trapping in semiconductor materials [36]. When the rate of recombination of electron–hole pairs is high, the peaks in the PL spectrum are intense. Thus, it is clear from the recorded PL spectra of MoS2 and La-MoS2 that on doping with La3+, the rate of recombination of electron–hole pairs decreases, as the intensity La-MoS2 is less than that of pure MoS2. The decrease in the recombination rate is due to the introduction of a discrete energy level due to doping with La3+.

2.2. Photocatalytic Performance of La-MoS2

To analyze the photocatalytic efficiency, the synthesized MoS2 and La-MoS2 samples were suspended in water using lactic acid as a sacrificial agent, and the whole suspensions were stimulated with solar irradiation. Figure 5a shows the results of hydrogen production using 10 mg of photocatalyst. With increasing time, the photocatalytic efficiency increases for both MoS2 and La-MoS2, but for La-MoS2, the rate of the photocatalytic hydrogen evolution rection (HER) was much higher than for MoS2. After 5 h, the amount of H2 was found to be 612 µmol·g−1 for MoS2 and 918 µmol·g−1 for La-MoS2. Thus, it may be concluded that La-MoS2 works better for the HER as compared to MoS2 due to charge separation. Hence, for further experiments, La-MoS2 was taken into consideration. The effects of different doses of La-MoS2 are shown in Figure 5b. The doses of La-MoS2 were varied between 10–50 mg. The results were collected at intervals of one hour for a total of five hours. As shown in Figure 5b, with an increasing dose of the photocatalyst, the amount of H2 production also increased.
The highest amount of H2, 1670 µmol·g−1, was obtained for 50 mg catalyst. The rates of H2 evolution for different catalyst doses are summarized in Figure 5c. Figure 5c shows that when the dose of La-MoS2 was increased, the rate of H2 production increased more than 1.5-folds. It reached from 185.6 µmol·h−1·g−1 to 334 µmol·h−1·g−1. In order to use a catalyst in real-time applications, it is essential to test the stability and reusability of the photocatalyst; accordingly, for La-MoS2, a reusability test was performed, and it showed almost consistent results up to four consecutive cycles (Figure 5d). The XRD results after the stability test also reflected the acceptable stability of the La-MoS2 (Figure 1).
A potential mechanism for the hydrogen evolution reaction on the surface of La-MoS2 is shown in Scheme 1. The H2 evolution is initiated when visible light strikes the surface of La-MoS2 and it absorbs photons with energy (hυ) either equal to or greater than its energy band gap [1]. On absorbing the energy from solar radiation, electrons jump from the valence band to the conduction band, creating electron–hole pairs and leaving the holes in the valence band. The photogenerated electrons that are present in the conduction band reduce H+ into H2, whereas the holes present in the valence band combine with H2O and break it down into O2 and H+ [37]. These holes also react with the scavenger lactic acid, changing it to pyruvic acid. The enhanced rate of photocatalytic production over the surface of La-MoS2 may be attributed to the narrow band gap and synergistic interactions. This may improve the electron transport process, and enhanced H2 evolution activity can be observed. Therefore, it can be concluded that controlling the transfer and migration of charge carriers may enhance the efficiency of photocatalysts, such as the spatial separation of carriers, elongating their lifetime and thus increasing their photocatalytic performance.
In previous years, various photocatalysts have been reported for photocatalytic H2 evolution. In particular, MoS2 exhibited acceptable performance for the generation of H2 under visible light irradiation. Xin et al. [38] reported the use of MoS2 as a photocatalyst; it achieved a demonstrated H2 evolution rate of 99.4 µmol·h−1·g−1 with lactic acid as a sacrificial reagent. The authors further adopted P-doped MoS2 as a photocatalyst and explored it for H2 evolution. An enhanced H2 evolution rate of 278.8 µmol·h−1·g−1 was observed with a similar environment and conditions. The authors stated that doping with the element P significantly improved the catalytic activity of P-MoS2. In another published article, MoS2 was used as a photocatalyst, and its photocatalytic activity was evaluated in presence of SO32− sacrificial reagent [39]. The authors found that pristine MoS2 had lower photocatalytic activity for H2 evolution and a low H2 evolution rate of 39 µmol·h−1·g−1. The authors also used pristine ZnO as a photocatalyst, which achieved a demonstrated H2 evolution rate of 22 µmol·h−1·g−1. However, an MoS2/ZnO composite demonstrated a significant improvement in its H2 evolution rate, and a decent H2 evolution rate of 235 µmol·h−1·g−1 was obtained. Pristine MoS2 also exhibited H2 evolution at a rate of 185 µmol·h−1·g−1 in the presence of methanol as a sacrificial reagent [40]. In another work, MoS2/g-C3N4 composite-based investigations revealed that H2 evolution at a rate of 441.3 µmol·h−1·g−1 could be observed in presence of triethanolamine as a sacrificial reagent [41]. Wei et al. [42] also proposed the synthesis of MoN1.2xS2−1.2x@g-C3N4 as a photocatalyst for H2 production applications. An interesting H2 evolution rate of 360.4 µmol·h−1·g−1 was observed for the photocatalyst MoN1.2xS2−1.2x@g-C3N4 in the presence of triethanolamine as a sacrificial reagent. In 2023, Wang et al. [43] proposed the hydrothermal synthesis of O, P–MoS2 for H2 production application under visible light. The authors used triethanolamine/acetonitrile as sacrificial reagents and reported that the H2 evolution rate was 339.3 µmol·h−1·g−1. The aforementioned studies showed that MoS2-based photocatalysts have a significant role in photocatalytic H2 evolution studies. The results obtained in the present study are compared with previously reported articles in Table 1. The comparison of results showed that La-MoS2 showed a decent response for hydrogen production compared to the many previously reported photocatalysts (Table 1).

3. Materials and Methods

3.1. Materials

Ammonium molybdate tetrahydrate (NH4)6Mo7O24·4H2O; 99.98% trace metal basis), thiourea (ACS reagent, ≥99.0%), and lanthanum (III) nitrate hexahydrate (99.999% trace metal basis) were purchased from Merck (Mumbai, India). All the other used chemicals and materials were of analytical grade and used as received from Sigma (Mumbai, India), Alfa Aesar (Indore, India), and Merck (Mumbai, India). No further treatment or purification was carried out.

3.2. Synthesis of La-MoS2

In this study, we adopted hydrothermal synthesis method for the preparation of pristine MoS2 and La-MoS2 materials. In brief, hydrothermal treatment of the reaction solution of NH4)6Mo7O24·4H2O and thiourea yielded MoS2. In brief, 0.6 g of thiourea was slowly added to an aqueous solution of 0.5 g ammonium molybdate (30 mL), and this reaction solution was stirred for 25–30 min at room temperature. A clear solution was obtained, which was then poured into a Teflon reactor and sealed tightly in a stainless steel autoclave (the autoclave needed to be carefully tightened to prevent the melting of the Teflon cup). This autoclave was heated at 180 °C for 18 h in a muffle furnace. After cooling down the furnace, the autoclave was opened; the MoS2 was then collected, washed with DI water and ethanol, and dried at 70 °C for 8 h in a vacuum oven. In further investigations, La-MoS2 was also prepared using a hydrothermal method as illustrated in Scheme 2.
Typically, 0.6 g of thiourea was added to an aqueous solution of 0.5 g ammonium molybdate. Furthermore, 5 wt% (0.055 g) lanthanum nitrate was added to the above reaction solution and stirred for 25–30 min at room temperature. After stirring, this solution was transferred to a Teflon reactor, sealed tightly in a stainless steel autoclave, and heated at 180 °C for 18 h in a muffle furnace (Scheme 2). After cooling down the furnace, the autoclave was opened, and La-MoS2 was collected using centrifugation, washed with DI water and ethanol, and dried at 70 °C for 8 h in a vacuum oven.

3.3. Instruments

A Rigaku powder X-ray diffractometer (model Rint 2500, manufactured in Rigaku, Tokyo, Japan) with wavelength of 1.5406 Å and Cu/Kα radiation was used to record the PXRD results of the synthesized samples. A Supra Zeiss-55 field-emission scanning electron microscope was used to record the surface morphological images of the synthesized samples (Zeiss, Jena, Germany). The energy-dispersive X-ray spectroscopy (EDX) results were obtained on a Horiba EDX spectroscope (Horiba, Kyoto, Japan). An ultraviolet–visible spectrophotometer (UV-vis spectrophotometer model Cary 100 (Varian, Palo Alto, CA, USA) was used to capture the UV-vis spectra of the synthesized samples. The photoluminescence (PL) spectra of the samples were obtained on a Dongwoo Optron PL spectrometer. The hydrogen evolution results were obtained on a gas chromatograph with a thermal conductivity detector (TCD).

3.4. Photocatalytic H2 Evolution

We used a quartz tube reactor as the photocatalytic H2 production set-up. A mass of 10 mg of the catalyst (La-MoS2 or MoS2) was added to a mixture of 80 mL DI water and 20 mL lactic acid. The reaction solution was purged with nitrogen gas for 40 min to remove unnecessary gases or oxygen. The quartz tube containing the catalyst and reaction solution was closed with an airtight seal and used for photocatalytic H2 evolution processes under visible light irradiation (Figure S3). A 300 W xenon lamp with wavelength = 420 nm was used as the light source. Different doses of the catalyst (La-MoS2; 10, 20, 30, 40, and 50 mg) were used to optimize the performance of the photocatalyst for improved H2 evolution. The generated H2 was taken out using a syringe and measured by employing a gas chromatograph (in combination with a TCD).

4. Conclusions

In conclusion, lanthanum-doped molybdenum disulfide (La-MoS2) was synthesized via a hydrothermal route. To characterize La-MoS2, powder X-ray diffraction (PXRD), scanning electron microscopy (SEM), and energy-dispersive X-ray (EDX) techniques were used. The band gap of La-MoS2 was calculated as 1.68 eV, which is less than the band gap of synthesized MoS2 (i.e., 1.80 eV). In the photoluminescence (PL) spectra, a decrease in intensity was observed for La-MoS2 compared to MoS2, indicating increased charge separation for La-MoS2. The as-synthesized MoS2 and La-MoS2 were used for photocatalytic hydrogen evolution reactions (HERs), and La-MoS2 exhibited an improved H2 production rate compared to the MoS2. The increased photocatalytic performance may be because of the generation of another discrete energy level due to doping with La3+. Mechanistic aspects of the HER on the surface of La-MoS2 have also been discussed.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal14120893/s1, Figure S1: EDX spectra of MoS2 and La-MoS2 sub-microspheres; Figure S2. PL spectra of MoS2 and La-MoS2 sub-microspheres; Figure S3: Schematic illustration of H2 evolution reaction.

Author Contributions

Conceptualization: A.C. and R.A.K.; methodology: K.A. and A.C.; validation: R.A.K., S.S.A. and A.A.; formal analysis: A.C.; investigation: K.A.; resources, R.A.K.; writing—original draft preparation: K.A. and A.C.; writing—review and editing: T.H.O. All authors have read and agreed to the published version of the manuscript.

Funding

Researchers Supporting Project number (RSP2024R400), King Saud University, Riyadh, Saudi Arabia.

Data Availability Statement

The authors elect not to share the data.

Acknowledgments

Authors gratefully acknowledged Researchers Supporting Project number (RSP2024R400), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Rahman, A.; Jennings, J.R.; Tan, A.L.; Khan, M.M. Molybdenum Disulfide-Based Nanomaterials for Visible-Light-Induced Photocatalysis. ACS Omega 2022, 7, 22089–22110. [Google Scholar] [CrossRef] [PubMed]
  2. Christoforidis, K.C.; Fornasiero, P. Photocatalytic Hydrogen Production: A Rift into the Future Energy Supply. Chem. Cat. Chem. 2017, 9, 1523–1544. [Google Scholar] [CrossRef]
  3. Lewis, N.S.; Nocera, D.G. Powering the planet: Chemical challenges in solar energy utilization. Proc. Natl. Acad. Sci. USA 2006, 103, 15729–15735. [Google Scholar] [CrossRef] [PubMed]
  4. Furukawa, H.; Yaghi, O.M. Storage of hydrogen, methane, and carbon dioxide in highly porous covalent organic frameworks for clean energy applications. J. Am. Chem. Soc. 2009, 131, 8875–8883. [Google Scholar] [CrossRef]
  5. Ganguly, P.; Harb, M.; Cao, Z.; Cavallo, L.; Breen, A.; Dervin, S.; Dionysiou, D.D. 2D Nanomaterials for Photocatalytic Hydrogen Production. ACS Energy Lett. 2019, 4, 1687–1709. [Google Scholar] [CrossRef]
  6. Gupta, U.; Rao, C. Hydrogen generation by water splitting using MoS2 and other transition metal dichalcogenides. Nano Energy 2017, 41, 49–65. [Google Scholar] [CrossRef]
  7. Teets, T.S.; Nocera, D.G. Photocatalytic hydrogen production. Chem. Commun. 2011, 47, 9268–9274. [Google Scholar] [CrossRef]
  8. Kumaravel, V.; Mathew, S.; Bartlett, J.; Pillai, S.C. Photocatalytic hydrogen production using metal doped TiO2: A review of recent advances. Appl. Catal. B 2019, 244, 1021–1064. [Google Scholar] [CrossRef]
  9. Noyan, O.F.; Hasan, M.M.; Pala, N. A Global Review of the Hydrogen Energy Eco-System. Energies 2023, 16, 1484. [Google Scholar] [CrossRef]
  10. Li, Y.; Somorjai, G.A. Nanoscale advances in catalysis and energy applications. Nano Lett. 2010, 10, 2289–2295. [Google Scholar] [CrossRef]
  11. Omer, A.M. Energy, environment and sustainable development. Renew. Sustain. Energy Rev. 2008, 12, 2265–2300. [Google Scholar] [CrossRef]
  12. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  13. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338. [Google Scholar] [CrossRef] [PubMed]
  14. Wang, X.; Maeda, K.; Thomas, A.; Takanabe, K.; Xin, G.; Carlsson, J.M.; Domen, K.; Antonietti, M. A metal-free polymeric photocatalyst for hydrogen production from water under visible light. Nat. Mater. 2009, 8, 76–80. [Google Scholar] [CrossRef]
  15. Zhou, X.; Liu, N.; Schmidt, J.; Kahnt, A.; Osvet, A.; Romeis, S.; Zolnhofer, E.M.; Marthala, V.R.R.; Guldi, D.M.; Peukert, W. Noble-Metal-Free Photocatalytic Hydrogen Evolution Activity: The Impact of Ball Milling Anatase Nanopowders with TiH2. Adv. Mater. 2017, 29, 1604747. [Google Scholar] [CrossRef]
  16. Zhu, Y.; Ling, Q.; Liu, Y.; Wang, H.; Zhu, Y. Photocatalytic H2 evolution on MoS2–TiO2 catalysts synthesized via mechanochemistry. Phys. Chem. Chem. Phys. 2015, 17, 933–940. [Google Scholar] [CrossRef]
  17. Kumar, S.; Kumar, A.; Rao, V.N.; Kumar, A.; Shankar, M.V.; Krishnan, V. Defect-Rich MoS2 Ultrathin Nanosheets-Coated Nitrogen-Doped ZnO Nanorod Heterostructures: An Insight into in-Situ-Generated ZnS for Enhanced Photocatalytic Hydrogen Evolution. ACS Appl. Energy Mater. 2019, 2, 5622–5634. [Google Scholar] [CrossRef]
  18. Hu, Y.; Yu, X.; Liu, Q.; Wang, L.; Tang, H. Highly Metallic Co-Doped MoS2 Nanosheets as an Efficient Cocatalyst to Boost Photoredox Dual Reaction for H2 Production and Benzyl Alcohol Oxidation. Carbon 2022, 188, 70–80. [Google Scholar] [CrossRef]
  19. Yusuf, M.; Song, S.; Park, S.; Park, K.H. Multifunctional Core-Shell Pd@Cu on MoS2 as a Visible Light-Harvesting Photocatalyst for Synthesis of Disulfide by S S Coupling. Appl. Catal. A 2021, 613, 118025. [Google Scholar] [CrossRef]
  20. Hu, J.; Zhang, C.; Li, X.; Du, X. An Electrochemical Sensor Based on Chalcogenide Molybdenum Disulfide-Gold-Silver Nanocomposite for Detection of Hydrogen Peroxide Released by Cancer Cells. Sensors 2020, 20, 6817. [Google Scholar] [CrossRef]
  21. Lai, H.; Ma, G.; Shang, W.; Chen, D.; Yun, Y.; Peng, X.; Xu, F. Multifunctional Magnetic Sphere-MoS2@Au Hybrid for Surface-Enhanced Raman Scattering Detection and Visible Light Photo-Fenton Degradation of Aromatic Dyes. Chemosphere 2019, 223, 465–473. [Google Scholar] [CrossRef] [PubMed]
  22. Gottam, S.R.; Tsai, C.-T.; Wang, L.-W.; Lin, C.-C.; Chu, S.-Y. Highly Sensitive Hydrogen Gas Sensor Based on a MoS2-Pt Nanoparticle Composite. Appl. Surf. Sci. 2020, 506, 144981. [Google Scholar] [CrossRef]
  23. Chi, M.; Zhu, Y.; Jing, L.; Wang, C.; Lu, X. Fabrication of Ternary MoS2-Polypyrrole-Pd Nanotubes as Peroxidase Mimics with a Synergistic Effect and Their Sensitive Colorimetric Detection of L-Cysteine. Anal. Chim. Acta 2018, 1035, 146–153. [Google Scholar] [CrossRef] [PubMed]
  24. Lin, T.; Zhong, L.; Guo, L.; Fu, F.; Chen, G. Seeing Diabetes: Visual Detection of Glucose Based on the Intrinsic Peroxidase-like Activity of MoS2 Nanosheets. Nanoscale 2014, 6, 11856–11862. [Google Scholar] [CrossRef] [PubMed]
  25. Liu, Y.; Xie, Y.; Liu, L.; Jiao, J. Sulfur vacancy induced high performance for photocatalytic H2 production over 1T@2H phase MoS2 nanolayers. Catal. Sci. Technol. 2017, 7, 5635–5643. [Google Scholar] [CrossRef]
  26. Okuno, Y.; Lancry, O.; Tempez, A.; Cairone, C.; Bosi, M.; Fabbri, F.; Chaigneau, M. Probing the nanoscale light emission properties of a CVD-grown MoS2 monolayer by tip-enhanced photoluminescence. Nanoscale 2018, 10, 14055–14059. [Google Scholar] [CrossRef]
  27. Deokar, G.; Rajput, N.S.; Vancsó, P.; Ravaux, F.; Jouiad, M.; Vignaud, D.; Cecchet, F.; Colomer, J.F. Large area growth of vertically aligned luminescent MoS2 nanosheets. Nanoscale 2017, 9, 277–287. [Google Scholar] [CrossRef]
  28. Pradhan, G.; Sharma, A.K. Anomalous Raman and photoluminescence blue shift in mono- and a few layered pulsed laser deposited MoS2 thin films. Mater. Res. Bull. 2018, 102, 406–411. [Google Scholar] [CrossRef]
  29. Marin, R.; Back, M.; Mazzucco, N.; Enrichi, F.; Frattini, R.; Benedetti, A.; Riello, P. Unexpected optical activity of cerium in Y2O3:Ce3+, Yb3+, Er3+ up and down-conversion system. Dalton Trans. 2013, 42, 16837. [Google Scholar] [CrossRef]
  30. Revathy, J.S.; Priya, N.S.C.; Sandhya, K.; Rajendran, D.N. Structural and optical studies of cerium doped gadolinium oxide phosphor. Bull. Mater. Sci. 2021, 44, 13. [Google Scholar] [CrossRef]
  31. Kumar, M.J.K.; Kalathi, J.T. Low-temperature sonochemical synthesis of high dielectric Lanthanum doped Cerium oxide nanopowder. J. Alloy. Compd. 2018, 748, 348–354. [Google Scholar] [CrossRef]
  32. Afanasiev, P.; Bezverkhy, I. Synthesis of MoSx (5 > x > 6) Amorphous Sulfides and Their Use for Preparation of MoS2 Monodispersed. Chem. Mater. 2002, 14, 2826–2830. [Google Scholar] [CrossRef]
  33. Chen, X.; Wang, X.; Wang, Z.; Yu, W.; Qian, Y. Direct sulfidization synthesis of high-quality binary sulfides (WS2, MoS2, and V5S8) from the respective oxides. Mater. Chem. Phys. 2004, 87, 327–331. [Google Scholar] [CrossRef]
  34. Riaz, K.N.; Yousaf, N.; Tahir, M.B.; Israr, Z.; Iqbal, T. Facile hydrothermal synthesis of 3D flower-like La-MoS2 nanostructure for photocatalytic hydrogen energy production. Int. J. Energy Res. 2019, 43, 491–499. [Google Scholar] [CrossRef]
  35. Radisavljevic, B.; Radenovic, A.; Brivio, J. Single-layer MoS2 transistors. Nat. Nanotechnol. 2011, 6, 147–150. [Google Scholar] [CrossRef]
  36. Schneider, J.; Matsuoka, M. Takeuchi Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef]
  37. Galińska, A.; Walendziewski, J. Photocatalytic water splitting over Pt–TiO2 in the presence of sacrificial reagents. Energy Fuels 2005, 19, 1143–1147. [Google Scholar] [CrossRef]
  38. Xin, X.; Song, Y.; Guo, S.; Zhang, Y.; Wang, B.; Wang, Y.; Li, X. One-step synthesis of P-doped MoS2 for efficient photocatalytic hydrogen production. J. Alloys Compd. 2020, 829, 154635. [Google Scholar] [CrossRef]
  39. Hunge, Y.M.; Yadav, A.A.; Kang, S.-W.; Lim, S.J.; Kim, H. Visible light activated MoS2/ZnO composites for photocatalytic degradation of ciprofloxacin antibiotic and hydrogen production. J. Photochem. Photobiol. A Chem. 2023, 434, 114250. [Google Scholar] [CrossRef]
  40. Li, D.; Zeng, G.; Wu, Y.; Zhou, Z.; Guo, W.; Li, C. Au induced in-situ formation of ultra-stable 1T-MoS2 on polymeric carbon nitride toward promoted photocatalytic hydrogen production. Int. J. Hydrogen Energy 2023, 48, 34363–34369. [Google Scholar] [CrossRef]
  41. Xu, Y.; Ouyang, J.; Zhang, L.; Long, H.; Song, Y.; Cui, Y. Embedded 1T-rich MoS2 into C3N4 hollow microspheres for effective photocatalytic hydrogen production. Chem. Phys. Lett. 2023, 814, 140331. [Google Scholar] [CrossRef]
  42. Wei, X.; Wang, M.; Ali, S.; Wang, J.; Zhou, Y.; Zuo, R.; Zhong, Q.; Zhan, C. Enhanced photocatalytic H2 evolution on g-C3N4 nanosheets loaded with nitrogen-doped MoS2 as cocatalysts. Int. J. Hydrogen Energy 2024, 89, 691–702. [Google Scholar] [CrossRef]
  43. Wang, L.; Hu, Y.; Xu, J.; Huang, Z.; Lao, H.; Xu, X.; Xu, J.; Tang, H.; Yuan, R.; Wang, Z.; et al. Dual non-metal atom doping enabled 2D 1T-MoS2 cocatalyst with abundant edge-S active sites for efficient photocatalytic H2 evolution. Int. J. Hydrogen Energy 2023, 48, 16987–16999. [Google Scholar] [CrossRef]
Figure 1. PXRD patterns of as-obtained MoS2, La-MoS2, and La-MoS2 after stability test.
Figure 1. PXRD patterns of as-obtained MoS2, La-MoS2, and La-MoS2 after stability test.
Catalysts 14 00893 g001
Figure 2. SEM image (a) and particle size distribution (b) of MoS2. SEM image (c) and particle size distribution (d) of La-MoS2.
Figure 2. SEM image (a) and particle size distribution (b) of MoS2. SEM image (c) and particle size distribution (d) of La-MoS2.
Catalysts 14 00893 g002
Figure 3. (a) Electron micrograph and mapping images of the elements (b) Mo and (c) S in the prepared MoS2. (d) Electron micrograph and mapping images of the elements (e) La, (f) Mo, and (g) S in the prepared La-MoS2.
Figure 3. (a) Electron micrograph and mapping images of the elements (b) Mo and (c) S in the prepared MoS2. (d) Electron micrograph and mapping images of the elements (e) La, (f) Mo, and (g) S in the prepared La-MoS2.
Catalysts 14 00893 g003
Figure 4. (a) UV-vis spectra and (b) Tauc plots of MoS2 and La-MoS2.
Figure 4. (a) UV-vis spectra and (b) Tauc plots of MoS2 and La-MoS2.
Catalysts 14 00893 g004
Figure 5. (a) H2 evolution activity of MoS2 and La-MoS2 (10 mg catalyst) in lactic acid. (b) Effect of different doses of catalyst (La-MoS2; 10–50 mg) in lactic acid. (c) H2 evolution rates for different catalyst doses. (d) Reusability test.
Figure 5. (a) H2 evolution activity of MoS2 and La-MoS2 (10 mg catalyst) in lactic acid. (b) Effect of different doses of catalyst (La-MoS2; 10–50 mg) in lactic acid. (c) H2 evolution rates for different catalyst doses. (d) Reusability test.
Catalysts 14 00893 g005
Scheme 1. Potential mechanism of H2 generation.
Scheme 1. Potential mechanism of H2 generation.
Catalysts 14 00893 sch001
Scheme 2. Representation of the synthetic process for the preparation of La-MoS2.
Scheme 2. Representation of the synthetic process for the preparation of La-MoS2.
Catalysts 14 00893 sch002
Table 1. Comparison of H2 evolution rate of La-MoS2 with published articles.
Table 1. Comparison of H2 evolution rate of La-MoS2 with published articles.
PhotocatalystsH2 Evolution Rate (µmol·h−1·g−1)Light SourceSacrificial AgentReference
La-MoS2334300 W; Xe lamp (λ = 420 nm)Lactic acidThis study
P-MoS2278.8300 W; Xe lamp (λ = 420 nm)Lactic acid38
MoS299.4300 W; Xe lamp (λ = 420 nm)Lactic acid38
MoS239300 W; Xe lamp (λ = 420 nm)SO32−39
MoS2/ZnO235300 W; Xe lamp (λ = 420 nm)SO32−39
ZnO22300 W; Xe lamp (λ = 420 nm)SO32−39
g-C3N454300 W; Xe lamp (λ = 420 nm)Methanol40
MoS2185300 W; Xe lamp (λ = 420 nm)Methanol40
MoS2/g-C3N4441.3300 W; Xe lamp (λ = 420 nm)Triethanolamine41
MoN1.2xS2−1.2x@g-C3N4360.4300 W; simulated solar lightTriethanolamine42
O, P–MoS2339.3LED light; λ = 420 nmTriethanolamine/acetonitrile43
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

Chaudhary, A.; Khan, R.A.; Almadhhi, S.S.; Alsulmi, A.; Ahmad, K.; Oh, T.H. Hydrothermal Synthesis of La-MoS2 and Its Catalytic Activity for Improved Hydrogen Evolution Reaction. Catalysts 2024, 14, 893. https://doi.org/10.3390/catal14120893

AMA Style

Chaudhary A, Khan RA, Almadhhi SS, Alsulmi A, Ahmad K, Oh TH. Hydrothermal Synthesis of La-MoS2 and Its Catalytic Activity for Improved Hydrogen Evolution Reaction. Catalysts. 2024; 14(12):893. https://doi.org/10.3390/catal14120893

Chicago/Turabian Style

Chaudhary, Archana, Rais Ahmad Khan, Sultan Saad Almadhhi, Ali Alsulmi, Khursheed Ahmad, and Tae Hwan Oh. 2024. "Hydrothermal Synthesis of La-MoS2 and Its Catalytic Activity for Improved Hydrogen Evolution Reaction" Catalysts 14, no. 12: 893. https://doi.org/10.3390/catal14120893

APA Style

Chaudhary, A., Khan, R. A., Almadhhi, S. S., Alsulmi, A., Ahmad, K., & Oh, T. H. (2024). Hydrothermal Synthesis of La-MoS2 and Its Catalytic Activity for Improved Hydrogen Evolution Reaction. Catalysts, 14(12), 893. https://doi.org/10.3390/catal14120893

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