Next Article in Journal
Optical Absorption on Electron Quantum-Confined States of Perovskite Quantum Dots
Previous Article in Journal
Lanthanide-Ion-Doping Effect on the Morphology and the Structure of NaYF4:Ln3+ Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ex Situ Synthesis and Characterizations of MoS2/WO3 Heterostructures for Efficient Photocatalytic Degradation of RhB

1
Department of Physics, University of Lahore, Lahore 54000, Pakistan
2
Department of Physics, University of the Punjab, Lahore 54590, Pakistan
3
School of Materials Science and Engineering, Zhejiang University, Hangzhou 310027, China
4
Institute of Molecular Biology and Biotechnology, University of Lahore, Lahore 54000, Pakistan
5
Department of Nanoengineering, Kyonggi University, Suwon 16227, Korea
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(17), 2974; https://doi.org/10.3390/nano12172974
Submission received: 26 July 2022 / Revised: 21 August 2022 / Accepted: 22 August 2022 / Published: 28 August 2022

Abstract

:
In this study, novel hydrothermal ex situ synthesis was adopted to synthesize MoS2/WO3 heterostructures using two different molar ratios of 1:1 and 1:4. The “bottom-up” assembly was successfully developed to synthesize spherical and flaky-shaped heterostructures. Their structural, morphological, compositional, and bandgap characterizations were investigated through XRD, EDX, SEM, UV-Visible spectroscopy, and FTIR analysis. These analyses help to understand the agglomerated heterostructures of MoS2/WO3 for their possible photocatalytic application. Therefore, prepared heterostructures were tested for RhB photodegradation using solar light irradiation. The % efficiency of MoS2/WO3 composites for 30 min irradiation of 1:1 was 91.41% and for 1:4 was 98.16%. Similarly, the % efficiency of 1:1 MoS2/WO3 heterostructures for 60 min exposure was 92.68%; for 1:4, it was observed as 98.56%; and for 90 min exposure, the % efficiency of 1:1 was 92.41%, and 98.48% was calculated for 1:4 composites. The photocatalytic efficiency was further verified by reusability experiments (three cycles), and the characterization results afterward indicated the ensemble of crystalline planes that were responsible for the high efficiency. Moreover, these heterostructures showed stability over three cycles, indicating their future applications for other photocatalytic applications.

1. Introduction

The unprecedented growth of industrialization has led to organic contaminants in the environment. The degradation of pollutants by using semiconductor photocatalysis has been regarded as a promising technology [1,2]. Photocatalysis has gained increasing attention recently due to its potential to provide an ecologically friendly method for converting sunlight into chemical energy in mild reaction environments through photocatalytic processes and photochemical reactions. Several heterojunction photocatalysts have been developed and are being employed for water-splitting and the photocatalytic degradation of organic compounds. Among them, solar-responsive metallic oxides/sulfides have gained attention due to their effective redox processes [3,4].
For visible light photocatalysis, tungsten trioxide (WO3) is an excellent choice because of its nontoxicity and distinctive optical properties. In addition, it is inexpensive, easy to synthesize, has strong stability in both acidic and basic environments, and has high electron carrier efficiency [5]. Due to its limited light absorption range and the quick recombination of photo-generated electron-hole pairs, pure WO3 photocatalytic activity is severely limited. To improve solar light absorption, impurities and higher conduction band semiconducting materials such as TiO2 (−0.29 eV), Ag3PO4 (+0.3 eV), and BiVO4 (0 eV) were used [6]. The addition of impurities by doping, on the other hand, reduces the thermal stability of the material and provides defect sites that act as additional recombination centers. A unique approach for separating photo-generated electron–hole pairs driven by a self-built electric field while also broadening the spectrum absorption range is currently being investigated [5].
Molybdenum disulfide (MoS2) is a widely used photocatalyst due to its large specific surface area, abundant unsaturated active sites, and visible spectrum absorption. The photochemical performance of MoS2 can be improved by coupling it with other semiconductors [6]. MoS2 possesses a more powerful negative conduction band (−0.06 eV) than WO3; the addition of MoS2 might make it possible to overcome WO3’s low conduction band position. Few studies have reported on the use of MoS2 on WO3 to increase photocatalytic performance because of the very small contact area between these two semiconductors [7,8].
Contrary to the single-component system, heterojunction photocatalysts have become a more feasible option for the breakdown of toxic contaminants. Here, the ex situ approach was used to synthesize binary composites of MoS2/WO3 in various weight ratios [9]. Combining two catalysts with similar band gaps increases photocatalytic activity by creating a heterojunction, which increases the promotion of photo-generated charges and decreases the recombination rate. Therefore, a heterojunction can be formed using several semiconducting materials, such as WO3/MoS2, WO3/CuBi2O4, and WO3/CdS, which would improve the catalyst effectiveness and ability to absorb light. Given that MoS2 has a sizable surface area and strong electrical mobility, the MoS2/WO3 (MSW) pair is an excellent choice for creating the heterojunction [10].
We have successfully synthesized WO3, MoS2, and MoS2/WO3 composites and evaluated photocatalytic degradation of Rhodamine (RhB) dye under solar light. A significant increase in the rate of photodegradation of RhB in the presence of heterojunction photocatalysts has been observed. To investigate the photocatalytic effectiveness of ex situ hydrothermally synthesized MoS2/WO3 nanocomposites of varied weight ratios, Rh B photodegradation studies were carried out in this study. Several investigations, including pH, concentration, and reusability assays, were conducted to evaluate the effectiveness and performance of the catalyst prepared. To demonstrate the photocatalyst’s superiority, its degrading effectiveness was also compared to that of the literature.
Additionally, the mechanism of the composite-enhanced photocatalytic activity was investigated, which could be useful for understanding the hydrogen evolution process for future applications.

2. Materials and Methods

All the chemicals used were of analytical grade, purchased from Sigma-Aldrich, (St. Louis, MO, USA), and were used without further purification.

2.1. Synthesis of MoS2

MoS2 was produced utilizing a two-step hydrothermal technique under acidic circumstances, employing MoO3 and Ammonium Thiocyanate (NH4SCN) as starting materials. A total of 1.5 mmol MoO3 (0.22 g) was dissolved in 40 mL of deionized water and sonicated for 30 min, after which the pH value of the solution was adjusted to 1 by stirring for 30 min with a 1 mol/L HCl solution. These samples were then transferred to a 50 mL Teflon-lined stainless steel autoclave for hydrothermal treatment at 180 °C for 12 h, which was cooled down at room temperature, yielding a black powder of MoS2 by centrifugation for 10–15 min, washed multiple times with ethanol and deionized water, and dried for 12 h in an oven at 80 °C.

2.2. Synthesis of WO3

The sodium tungstate dehydrates (Na2WO4) (0.1 M) were dissolved in distilled water (10 mL) to maintain a pH value of ~8. Then, hydrochloric acid (HCl) (0.5 M) was added dropwise at 50 °C to form a homogeneous solution. After continuous stirring for 15 min, the pH of the resulting solution was set to ~1. The resulting solution was then shifted to a 50 mL Teflon-lined Stainless-Steel Autoclave and placed in the oven at 180 °C for 6 h for hydrothermal treatment. After naturally cooling down, blue precipitates were separated, centrifuged, and washed several times simultaneously with deionized water and absolute ethanol, then dried in an oven at 80 °C for 12 h. The resultant powder was obtained after annealing at 400 °C for 2 h in a muffle furnace.

2.3. Synthesis of MoS2/WO3

Already-prepared MoS2 and WO3 samples were used for this process. MoS2/WO3 (1:1) and MoS2/WO3 (1:4) nanocomposites were prepared by using an ex situ synthesis in which the first MoS2 solution was prepared with 10 mL of ethanol, and the second WO3 solution was prepared with 10 mL of ethanol. Following that, WO3 solution was added dropwise to the MoS2 solution while continuous stirring was maintained for the next 30 min. The resulting solution was then centrifuged and washed simultaneously with deionized water and ethanol. The resultant dark greyish powder was obtained after drying in an oven at 80 °C for 12 h. After drying, the agglomerated material was ground in mortar and pestle to a fine powder of MoS2/WO3 (1:1) and MoS2/WO3 (1:4).

2.4. Photocatalytic Activity

The RhB dye was used to investigate photocatalytic activity. The research was carried out under the presence of a solar simulator, namely, the “Abet Technologies Sunlight TM Solar Simulator.” The catalysts (MoS2/WO3) were washed by a series of centrifugation and washing steps before being reused for the subsequent degradation. Three cycles were performed to test the catalysts’ reusability and stability. The dye solution catalytic decolorization is a pseudo-first-order reaction, and the degradation rate was estimated using Equation (1) [11]:
Degradation   ( % ) = ( 1 C t / C 0 ) × 100 %
where C t and C 0 are the dye’s concentration at time t and initial concentration, respectively. The degradation of 100 mL of aqueous RhB dye (10 ppm) was examined at dark, 30 min, 60 min, and 90 min. All the tests were carried out with 0.1 mg of the catalyst at its natural pH. During the photocatalytic degradation process, the distinctive absorption peak of RhB at 554 nm was set. In the absence of a catalyst, there was essentially no degradation, or at the very least, the rate of degradation was minimal. The catalytic activity of MoS2/WO3 was also tested in the dark for one hour.

3. Results and Discussion

3.1. Structural Analysis

Structural properties of MoS2/WO3 heterostructures were examined out using XRD analysis, as shown in Figure 1. The XRD was performed through Analytical X’Pert emitting CuK (Alpha) X-rays at the scanning range of 10–70°. All the diffracted peaks for MoS2/WO3-(1:1) and MoS2/WO3-(1:4) prepared heterostructures were matched with JCPDS 73-1508 for MoS2 and JCPDS 89-4480 for WO3, confirming the successful synthesis of heterostructures. The corresponding hkl parameters are marked in Figure 1. The observed pattern shows the increase in the WO3 ratio, and the intensity of WO3 peaks is also increased compared to MoS2 peaks. Moreover, the composition of the MoS2/WO3 heterostructures was also confirmed by using the EDX spectra, as shown in Figure S1.

3.2. Morphological Analysis

SEM analysis of MoS2/WO3 heterostructures (SEM) is shown in Figure 2a–d. Figure 2a,b are SEM images of MoS2/WO3 (1:4) heterostructures, where the prepared nanostructures have a relatively small size in comparison to Figure 2c,d for MoS2/WO3 (1:1). Figure 2b spherical agglomerates are smaller than Figure 2d agglomerates due to changed concentration, which may result in a decrease in crystalline size and an increase in the surface area [12]. SEM images of MoS2 and WO3 has been provided in Figure S2.

3.3. FTIR Analysis

For WO3 (in Figure 3), a very weak shoulder peak observed at 930 cm−1 was attributed to the W=O stretching vibration. As strong absorption peaks of W-O vibrational frequency near 1100 cm−1 in MoS2/WO3 1:1 and 1:4 heterostructures, this peak was slightly shifted to 1100 cm−1 [12], and it overlaps with W-O at 1100 cm−1 in MoS2/WO3 (1:1) and MoS2/WO3 (1:4) heterostructure peaks. Due to the overlap of S-O and W-O peaks, the intense sharp peaks of both heterostructures appear as stronger peaks than the parent WO3 and MoS2 molecules. Another peak at 1365 cm−1 was absent in WO3 but appeared as a prominent peak in MoS2/WO3 1:1 and 1:4 heterostructures corresponding to W-OH vibrational frequency in the form of a bending peak, but slight and broad and intensely sharp peaks were observed at 3450 cm−1 and 1612 cm−1 corresponding to OH stretching and bending peaks due to water molecules present in the WO3 crystal, respectively. Both became broader, more intense, and sharper in MoS2/WO3 1:1 and 1:4, which shows the presence of string H-bonding. A sharp peak of Mo-OH vibration was observed for both heterostructures spectra, but it was absent in the WO3 spectrum and appeared as a slight impression in the MoS2 spectrum. It is inferred that both heterostructures were prepared by strong forces, which help composite molecules bind firmly together, and H-bonding due to water crystallization plays a vital role in this regard [12,13].

3.4. UV-Visible Spectroscopy Analysis

The UV-Vis spectroscopy model Shimadzu UV-1800 (Kyoto, Japan) was used to record absorption. Furthermore, by using Tauc’s plot, the optical bandgaps of prepared heterostructures of MoS2/WO3 were calculated by using Equation (2) [14].
α h ϑ = A   ( h ϑ E g ) n ,
where n depends on electron transitions, and n = 1/2 corresponds to indirect electron transitions with an indirect optical bandgap, and where is the photon energy, α is a constant, and Eg is the optical bandgap energy. MoS2/WO3 (1:1) exhibited a bandgap of 2.06 eV in Figure 4a, while MoS2/WO3 (1:4) exhibited a narrow bandgap of 1.59 eV in Figure 4b. Their absorption spectra have been provided in Figure S3.

3.5. RhB Photodegradation

Photocatalytic activity of the prepared MoS2/WO3 heterostructures at 1:1 was performed by studying the degradation of RhB under solar light irradiation. The absorption intensity peak of RhB was centered at 554 nm and gradually decreased with an increase in the irradiation time, as shown in Figure 5a. Figure 5b represents the absorption and photodegradation curves of an aqueous solution of RhB photodegraded by a photocatalyst at a concentration of 0.1 gm over time. Photodegradation of MoS2/WO3 heterostructures with 1:4 under dark conditions was also observed, and the characteristic peak was taken as 554 nm of RhB dye used as a standard parameter during the photocatalytic degradation process. It was also observed that the concentration of RhB decreased with time, which was attributed to its degradation by a catalyst [14].
The photocatalytic activity was observed under solar light irradiation for 30 min, 60 min, and 90 min in Figure 6a, showing the relationship between % efficiency and irradiation time for MoS2/WO3 (1:1) and (1:4) heterostructures. Efficiencies of 91.41%, 92.41%, and 92.68% were observed for MoS2/WO3 (1:1), and efficiencies of 98.16%, 98.48%, and 98.56% were observed for MoS2/WO3 (1:4).
Figure 6b shows the degradation rate (C/C0) versus time (minutes), i.e., under dark, it was 0.0448, while for 30 min of solar light irradiation, it was 0.0858; for 60 min, it was 0.7314, and for 90 min it was 0.0758 for MoS2/WO3 (1:1). Likewise, the degradation rate (C/C0) under dark for MoS2/WO3 (1:4) was observed as 1, for 30 min under solar light irradiation it was 0.018, for 60 min it was 0.014, and for 90 min it was 0.015.
Figure 6c shows the graph between Ln(C/C0) vs. time in minutes and vs. % efficiency. The values of Ln(C/C0) for MoS2/WO3 (1:1), for 30, 60, and 90 min are 2.454, 2.615, and 2.579, respectively. Meanwhile, MoS2/WO3 (1:4) Ln(C/C0) values for 30, 60, and 90 min are 3.996, 4.244, and 4.189, respectively. The rate constant k/h for MoS2/WO3 (1:1) is calculated as 0.0278 and 0.0425 MoS2/WO3 (1:4), indicating first-order kinetics with excellent photocatalytic activity. The stability of MoS2/WO3 prepared heterostructures was tested for up to three cycles, as shown in Figure 6d. After 90 min of solar irradiation, prepared heterostructures showed excellent degradation and stability.

3.6. Proposed Photocatalytic Mechanism

The MoS2/WO3 heterostructure demonstrated enhanced photocatalytic potential in two ways: (i) generation of p-n heterojunction by band alignment through the close contact interface and (ii) efficiency as a co-catalyst. Both mechanisms reduce incoming light by forming electron–hole pairs. Under thermodynamic conditions, electrons in the MoS2 conduction band (CB) migrate to WO3 due to its bigger negative Fermi level, leaving holes in MoS2. Electrons in MoS2 CB can partially react with O2 to generate •O2, and holes in WO3’s VB can partially react with H2O to form •OH. Due to MoS2 having slightly higher CB potential than O2/•O2, the electron’s reduction ability was so feeble that the production rate of •O2 was substantially slower. The VB potential of WO3 was larger than that of H2O/•OH; therefore, enough •OH could be formed. The active species, •O2 and OH, interacted with organic molecules, causing them to oxidize and produce CO2 and H2O as by-products. In addition, the holes acted as active specie directs, oxidizing the RhB to the final product. Figure 7 depicts the photocatalytic mechanism of MoS2/WO3 [15,16].

4. Conclusions

A heterostructure binary nanocomposite, MoS2/WO3, at 1:1 and 1:4 ratios, was successfully prepared via ex situ synthesis. These prepared heterostructures were characterized, and the formation of heterojunctions was confirmed through XRD analysis. The optical bandgap energies for the MoS2/WO3 heterostructure were calculated as 2.06 eV and 1.59 eV for 1:1 and 1:4, respectively. An efficient RhB photodegradation was observed under solar light irradiation for prepared heterostructures. Among the already reported photocatalysts, MoS2/WO3 prepared heterostructures showed enhanced photocatalytic activity. The maximum photodegradation efficiency for MoS2/WO3 (1:4) was observed as 98.5%, indicating it is a suitable material for efficient photocatalytic degradation. The value of the K/h rate constant was observed as 0.0425, indicating the best photocatalytic activity was found when using RhB dye at 1:4. The relation between irradiation time to the rate constant Ln(C/C0) for MoS2/WO3 heterostructures was observed as 0.0278 for MoS2/WO3 at 1:1. MoS2/WO3 showed a quick and overall degradation ability for RhB in just 30 min with good reusability behavior and photostability. With suitable bandgap engineering, the current heterostructure could be used for molecular hydrogen production. All of the mentioned findings indicate that the MoS2/WO3 composites made would have a wide range of potential applications for eliminating organic dyes from wastewater.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano12172974/s1, Figure S1: EDX spectra of MoS2/WO3 heterostructures; Figure S2: SEM image of (a,b) MoS2 and (c) WO3; Figure S3: Absorption Spectra of MoS2/WO3 (1:1) and MoS2/WO3 (1:4).

Author Contributions

W.S. designed and performed the experiments and wrote the manuscript. F.I. proposed, designed, and supervised all of the research work; M.U.T. helped with writing and graphing; S.S. and M.A.I. helped in reviewing the manuscript; and J.R.C. helped with the final revision. All authors have read and agreed to the published version of the manuscript.

Funding

This study is supported by the Pakistan Science Foundation (PSF) grant funded by the government of Pakistan (no. PSF-NSFC-IV/Phy/P-PU(31)) and by the National Research Foundation of Korea (NRF) grant no. NRF-2021R1F1A1062849.

Data Availability Statement

Not applicable.

Acknowledgments

Pakistan Science Foundation (PSF), Department of Physics, University of the Punjab, Lahore 54590, Pakistan and The University of Lahore, Pakistan, National Research Foundation of Korea and China Scholarship Council (CSC) from the Ministry of Education of the People’s Republic of China and Zhejiang University, China.

Conflicts of Interest

The authors declare no conflict of interest as well as competing interest, either financial or non-financial.

References

  1. Wang, R.; Zhang, W.; Zhu, W.; Yan, L.; Li, S.; Chen, K.; Hu, N.; Suo, Y.; Wang, J. Enhanced visible-light-driven photocatalytic sterilization of tungsten trioxide by surface-engineering oxygen vacancy and carbon matrix. Chem. Eng. J. 2018, 348, 292–300. [Google Scholar] [CrossRef]
  2. Oller, I.; Malato, S.; Sánchez Pérez, J.A. Combination of Advanced Oxidation Processes and biological treatments for wastewater decontamination—A review. Sci. Total Environ. 2011, 409, 4141–4166. [Google Scholar] [CrossRef] [PubMed]
  3. Bai, X.; Yang, L.; Hagfeldt, A.; Johansson, E.M.; Jin, P. D35-TiO2 nano-crystalline film as a high performance visible-light photocatalyst towards the degradation of bis-phenol A. Chem. Eng. J. 2018, 355, 999–1010. [Google Scholar] [CrossRef]
  4. Ullattil, S.G.; Narendranath, S.B.; Pillai, S.C.; Periyat, P. Black TiO2 Nanomaterials: A Review of Recent Advances. Chem. Eng. J. 2018, 343, 708–736. [Google Scholar] [CrossRef]
  5. Mohite, S.; Ganbavle, V.; Rajpure, K. Solar photoelectrocatalytic activities of rhodamine-B using sprayed WO3 photoelectrode. J. Alloys Compd. 2015, 655, 106–113. [Google Scholar] [CrossRef]
  6. Fu, Y.-D.; Feng, X.-X.; Yan, M.-F.; Wang, K.; Wang, S.-Y. First principle study on electronic structure and optical phonon properties of 2H-MoS2. Phys. B Condens. Matter 2013, 426, 103–107. [Google Scholar] [CrossRef]
  7. Lee, C.; Yan, H.; Brus, L.E.; Heinz, T.F.; Hone, J.; Ryu, S. Anomalous Lattice Vibrations of Single- and Few-Layer MoS2. ACS Nano 2010, 4, 2695–2700. [Google Scholar] [CrossRef] [PubMed]
  8. Rakibuddin, M.; Kim, H. Fabrication of MoS2/WO3 nanocomposite films for enhanced electrochromic performance. New J. Chem. 2017, 41, 15327–15333. [Google Scholar] [CrossRef]
  9. Singla, S.; Sharma, S.; Basu, S. MoS2/WO3 heterojunction with the intensified photocatalytic performance for decomposition of organic pollutants under the broad array of solar light. J. Clean. Prod. 2021, 324, 129290. [Google Scholar] [CrossRef]
  10. Guo, M.; Xing, Z.; Zhao, T.; Li, Z.; Yang, S.; Zhou, W. WS2 quantum dots/MoS2@WO3-x core-shell hierarchical dual Z-scheme tandem heterojunctions with wide-spectrum response and enhanced photocatalytic performance. Appl. Catal. B Environ. 2019, 257, 117913. [Google Scholar] [CrossRef]
  11. Tanveer, M.; Cao, C.; Aslam, I.; Ali, Z.; Idrees, F.; Tahir, M.; Khan, W.S.; Butt, F.K.; Mahmood, A. Effect of the morphology of CuS upon the photocatalytic degradation of organic dyes. RSC Adv. 2014, 4, 63447–63456. [Google Scholar] [CrossRef]
  12. Simelane, S.; Ngila, J.C.; Dlamini, L.N. The effect of humic acid on the stability and aggregation kinetics of WO3 nanoparticles. Part. Sci. Technol. 2017, 35, 632–642. [Google Scholar] [CrossRef]
  13. Zheng, Y.; Wang, J.; Wang, Y.; Zhou, H.; Pu, Z.; Yang, Q.; Huang, W. The combination of MoS2/WO3 and its adsorption properties of methylene blue at low temperatures. Molecules 2019, 25, 2. [Google Scholar] [CrossRef] [PubMed]
  14. Idrees, F.; Cao, C.; Butt, F.K.; Tahir, M.; Tanveer, M.; Aslam, I.; Ali, Z.; Mahmood, T.; Hou, J. Facile synthesis of novel Nb3O7 F nanoflowers, their optical and photocatalytic properties. CrystEngComm 2013, 15, 8146–8152. [Google Scholar] [CrossRef]
  15. Idrees, F.; Cao, C.; Ahmed, R.; Butt, F.; Butt, S.; Tahir, M.; Tanveer, M.; Aslam, I.; Ali, Z. Novel Nano-Flowers of Nb2O5 by Template Free Synthesis and Enhanced Photocatalytic Response Under Visible Light. Sci. Adv. Mater. 2015, 7, 1298–1303. [Google Scholar] [CrossRef]
  16. Ganesh, I.; Kumar, P.P.; Annapoorna, I.; Sumliner, J.M.; Ramakrishna, M.; Hebalkar, N.Y.; Padmanabham, G.; Sundararajan, G. Preparation and characterization of Cu-doped TiO2 materials for electrochemical, photoelectrochemical, and photocatalytic applications. Appl. Surf. Sci. 2014, 293, 229–247. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of MoS2/WO3 heterostructures; blue represents the molar ratio (1:1), while red shows (1:4).
Figure 1. XRD patterns of MoS2/WO3 heterostructures; blue represents the molar ratio (1:1), while red shows (1:4).
Nanomaterials 12 02974 g001
Figure 2. SEM Analysis of MoS2/WO3 with (a) molar ratio 1:4 at 5 µm, (b) molar ratio 1:4 at 1 µm, (c) molar ratio 1:1 at 5 µm, and (d) molar ratio 1:1 at 1 µm.
Figure 2. SEM Analysis of MoS2/WO3 with (a) molar ratio 1:4 at 5 µm, (b) molar ratio 1:4 at 1 µm, (c) molar ratio 1:1 at 5 µm, and (d) molar ratio 1:1 at 1 µm.
Nanomaterials 12 02974 g002
Figure 3. FTIR transmission spectra for (black) MoS2/WO3 (1:1), (red) MoS2/WO3 (1:4), (blue) MoS2, and (green) WO3.
Figure 3. FTIR transmission spectra for (black) MoS2/WO3 (1:1), (red) MoS2/WO3 (1:4), (blue) MoS2, and (green) WO3.
Nanomaterials 12 02974 g003
Figure 4. Optical bandgap of (a) MoS2/WO3 (1:1) and (b) MoS2/WO3 (1:4).
Figure 4. Optical bandgap of (a) MoS2/WO3 (1:1) and (b) MoS2/WO3 (1:4).
Nanomaterials 12 02974 g004
Figure 5. Graphs between wavelength vs. absorbance for MoS2/WO3 under dark conditions and at varying times of 30, 60, and 90 min for molar ratios: (a) (1:1) and (b) (1:4).
Figure 5. Graphs between wavelength vs. absorbance for MoS2/WO3 under dark conditions and at varying times of 30, 60, and 90 min for molar ratios: (a) (1:1) and (b) (1:4).
Nanomaterials 12 02974 g005
Figure 6. Graphs of: (a) % efficiency and time; (b) irradiation time vs. degradation (C/C0); (c) irradiation times versus Ln(C/C0) and % efficiency; (d) cycles vs. % efficiency for MoS2/WO3 prepared heterostructures.
Figure 6. Graphs of: (a) % efficiency and time; (b) irradiation time vs. degradation (C/C0); (c) irradiation times versus Ln(C/C0) and % efficiency; (d) cycles vs. % efficiency for MoS2/WO3 prepared heterostructures.
Nanomaterials 12 02974 g006
Figure 7. Schematic of photodegradation for MoS2/WO3.
Figure 7. Schematic of photodegradation for MoS2/WO3.
Nanomaterials 12 02974 g007
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shahid, W.; Idrees, F.; Iqbal, M.A.; Tariq, M.U.; Shahid, S.; Choi, J.R. Ex Situ Synthesis and Characterizations of MoS2/WO3 Heterostructures for Efficient Photocatalytic Degradation of RhB. Nanomaterials 2022, 12, 2974. https://doi.org/10.3390/nano12172974

AMA Style

Shahid W, Idrees F, Iqbal MA, Tariq MU, Shahid S, Choi JR. Ex Situ Synthesis and Characterizations of MoS2/WO3 Heterostructures for Efficient Photocatalytic Degradation of RhB. Nanomaterials. 2022; 12(17):2974. https://doi.org/10.3390/nano12172974

Chicago/Turabian Style

Shahid, Wajeehah, Faryal Idrees, Muhammad Aamir Iqbal, Muhammad Umair Tariq, Samiah Shahid, and Jeong Ryeol Choi. 2022. "Ex Situ Synthesis and Characterizations of MoS2/WO3 Heterostructures for Efficient Photocatalytic Degradation of RhB" Nanomaterials 12, no. 17: 2974. https://doi.org/10.3390/nano12172974

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