Next Article in Journal
Ovicidal Effect on Haemonchus contortus of Extract Partitions Shrubby Plants of the Tropical Dry Forest and Potentially Active Compounds Identification by UHPLC-Q/Orbitrap/MS/MS
Previous Article in Journal
Computational-Intelligence-Based Spectrum-Sharing Scheme for NOMA-Based Cognitive Radio Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hydrothermal Fabrication of GO Decorated Dy2WO6-ZnO Ternary Nanocomposites: An Efficient Photocatalyst for the Degradation of Organic Dye

by
Karuppaiah Selvakumar
1,*,
Tae Hwan Oh
1,*,
Muthuraj Arunpandian
1,
Kanakaraj Aruchamy
1 and
Veerababu Polisetti
2,*
1
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
2
Wallenberg Wood Science Center, Department of Fibre and Polymer Technology, School of Engineering Sciences in Chemistry, Biotechnology and Health, KTH Royal Institute of Technology, SE-100 44 Stockholm, Sweden
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2023, 13(12), 7145; https://doi.org/10.3390/app13127145
Submission received: 28 May 2023 / Revised: 10 June 2023 / Accepted: 12 June 2023 / Published: 14 June 2023

Abstract

:
Environmental and human health are seriously threatened by organic dye pollution. Many efforts have been made to find effective and safe methods of eliminating these contaminants. To mitigate these effects, the hydrothermal method was used to effectively generate a ternary kind of Dy2WO6-ZnO embedded in graphene oxide (DWZG) nanocomposites, which were used to degrade the pollutant. Powder X-ray diffraction (XRD) investigation confirms the crystalline character of the as-prepared DWZG nanocomposite. The Dy2WO6-ZnO composition on the graphene oxide (GO) layer is shaped like a combination of algae (Dy2WO6) and clusters (ZnO), as shown by scanning electron microscopy (SEM). X-ray photoelectron spectroscopy (XPS) investigation revealed the composition of elements and oxidation state of C, Dy, O, W and Zn elements. Methylene blue (MB) was chosen as the organic dye target for photocatalytic degradation using the produced nanocomposites. MB is degraded with a photocatalytic efficiency of 98.2% in about 30 min using a DWZG catalyst. Based on the result of the research entitled “Reactive Oxidative Species,” the primary reactive species involved in the MB degradation are photo-generated OH and O2•− radicals. The recycle test was also successful in evaluating the catalysts’ long-term viability as well as their reusability.

1. Introduction

The primary hazard to human existence is environmental pollution, specifically the contamination of the water supply. Industrialization is increasing pollutants every year, harming the ecosystem. The ecology and aquatic wildlife are seriously endangered when industrial wastewater is improperly dumped into the environment [1,2,3]. The process of cleaning up industrial effluent for discharge is complicated. Since certain methods of treating wastewater are too expensive, scientists have been hard at work developing more efficient methods in recent years [4]. One such method is the advanced oxidation process (AOP). Photocatalysis is an AOP technique used to clean water and other environmental contaminants. Recent research has shown that semiconductor photocatalysts can be utilized to degrade organic pollutants in the presence of light irradiation [4,5,6]. These photocatalytic methods have excellent efficiency and little impact on the environment. Photocatalysis has been recognized as a green chemistry method since it is not only efficient at converting organic pollutants into non-hazardous molecules but also because of its low cost [7,8].
Recently, researchers have made great strides in developing extremely efficient and innovative catalytic materials syntheses [9,10]. Semiconductor photocatalysts have been shown to be effective at degrading organic pollutants, making them a potentially useful tool in the fight against environmental pollution. TiO2, ZnO, WO3, ZnS, SnO2, etc., are just some of the many inorganic semiconductors (metal oxides) employed in the photodegradation process [11,12,13,14]. The photocatalytic destruction of organic contaminants and the solar-powered synthesis of hydrogen from water are two promising new uses for semiconductor photocatalysts that have gained a lot of attention recently [7]. TiO2 and ZnO, because of their broad bandgap and higher photosensitivity, are the most effective materials for the degradation process [15,16]. Although both of these semiconductor photocatalysts have been put to use in environmental detoxification, ZnO has been the more popular choice due to its superior degrading efficiency [17,18,19], and it has a lot going for it, such as the fact that it’s cheap, stable, very active, and its effects vary with size in comparison to other catalysts [20,21]. When compared to TiO2, ZnO nanoparticles are more effective in blocking both UVA and UVB radiation [22]. The photocatalytic degradation of various organic molecules by ZnO has been shown to be as effective as that by TiO2; in other scenarios, ZnO has even been shown to be more active than TiO2 in this regard [23,24].
ZnO photocatalysts have only recently been produced through the combination of other semiconductor oxide materials. The photocatalytic efficiency is growing as a result of the support of metal oxides (two or more) that are photocatalytic materials. Because photons form electron-hole pairs that are then separated, and the interfacial charge transfer efficiency is increased [25,26]. Researchers are drawn to the rare earth metal tungstate compounds because of their unique structure, interesting physio-chemical behavior, and great potential for industry applications [27,28]. Nano-structured doped metal oxides are growing in popularity due to their enhanced physicochemical properties, according to recent research, in order to tailor the doped metal oxide to a specific application, its physicochemical properties can be doped or co-doped with transition metals at varying concentrations [29,30]. The exceptional electrical conductivity of rare earth elements can be further enhanced by doping with semiconductor transition metals [31].
In addition, carbon-based materials such as GO have very appealing attributes such as effective surface area, distinctive architectures, and exceptional charge-transfer capabilities [32,33]. These are only a few of their many advantages. They have earned a reputation as a superb substrate for the development of nanocomposites that enhance photocatalytic activity.
The Dy2WO6-ZnO/GO (DWZG) nanocomposite is totally synthesized in this work by the hydrothermal process, and spectroscopic techniques are used to verify that it has been successfully synthesized. Under the influence of ultraviolet light, the nanocomposites that were created have a photocatalytic capability that is extremely effective for the destruction of organic pollutants, such as methylene blue (MB) dye [34,35]. The photocatalytic degradation process is an effective method for the removal of MB from the environment. The use of textile dyes such as MB, which is a plentiful source of colored organic compound and an increasingly widespread environmental danger, is recommended. In the textile industry, MB is used for dyeing and printing, but it is also a major contributor to water pollution in aquatic environments [36,37,38,39,40]. In order to lessen the impact that the MB had on the surrounding ecosystem, the photocatalytic procedure used the DWZG catalyst to remove it from the water. The dangerous organic pollutants are totally mineralized into non-hazardous compounds using this semiconductor photocatalysis process. These substances include carbon dioxide, water and other mineral acids, in addition to the process of deterioration.
This rare earth coupled ternary photocatalytic Dy2WO6-ZnO/GO (DWZG) nanocomposite was used for the degradation of MB with a superior degradation efficiency than the other previous reports. Because of its impact on the optical characteristics, morphology is a crucial factor in nanomaterial synthesis. Both changes in the band gap and improvements in photocatalytic performance contribute to the significance of morphological control [41]. Noting that these DWZG are mesoporous is crucial because pores improve photocatalytic activity and light dispersion, leading to more efficient pollutant decomposition. According to these conclusions, Dy2WO6-coupled ZnO systems have the potential to significantly improve their photocatalytic activity.

2. Materials and Methods

Raw materials for degradation include methylene blue from Merck Chemicals, as well as sodium hydroxide, zinc acetate, sodium tungstate, urea, oxalic acid, graphite powder, sulphuric acid, hydrochloric acid, hydrogen peroxide, and potassium permanganate from Sigma Aldrich and Merck Chemicals, respectively. All chemicals used in this work were extra-pure grade reagents. All solutions were made with de-ionized water, which is extremely pure and clean.

2.1. Preparation of Dy2WO6 (DW)

For 10 min, while stirring constantly, 3 mmol of Na2WO4·2H2O was dissolved in 100 mL of purified water. The aforementioned solution was then slowly added to 20 mL of dysprosium nitrate hexahydrate (6 mmol) while being stirred. After bringing the pH of the solution down to 10 with NaOH, 0.5 g/L of urea was added, and the mixture was agitated for 1 h to completely produce a precipitate of Dy2WO6. Hydrothermal treatment was performed on the aforementioned suspension mixture in a Teflon-lined autoclave at 180 °C for 24 h. The product was hydrothermally treated, then filtered and washed three times with water and ethanol after the autoclave had cooled to room temperature. The precipitate was dried in a vacuum oven at 60 °C for 12 h after being treated hydrothermally. After that, it spent 8 h in a muffle furnace at 550 °C.

2.2. Preparation of Dy2WO6-ZnO/GO (DWZG)

The modified Hummers method was used to produce graphene oxide. After 30 min of stirring, 100 mL of a 0.4 M Zn (CH3CO2)2·2H2O solution was added to the aforementioned Dy2WO6 (10%) suspension combination, followed by 100 mL of a 0.6 M oxalic acid solution added drop by drop over the course of 4 h. After finishing, the suspension mixer went through a 12 h hydrothermal treatment in a Teflon-lined stainless-steel autoclave heated to 115 °C. After 12 h the sample was heated in a 70 °C oven, the samples that had been hydrothermally treated were annealed for 3 h at 650 °C.
For 20 min, GO (10%) and de-ionized water were sonicated together. The aforementioned GO solution was then mixed with the catalyst Dy2WO6-ZnO (90%) for a further 2 h. The samples were cleaned via ultra-centrifugation and washed in distilled water, ethanol, and acetone once they had been collected during the completion reaction. The sample was dried overnight in a 60 °C oven after being collected.

2.3. Evaluation of Catalytic Degradation of Pollutants

Heber multi lamp photo reactor (model HML-MP 88), operating at 365 nm (UV light), was used to irradiate organic pollutants (MB). About 30 mg of the catalyst was combined with 100 mL of MB solution (1 × 10−5 M) for the degradation process. The adsorption–desorption equilibrium of the reaction solution was maintained by stirring the aforesaid mixture in the dark for 30 min. The dispersed solution was put into a borosilicate glass tube photo reactor vessel (40 cm height, 20 mm in diameter) after being thoroughly mixed. Using a UV visible spectrometer, the absorption peak of MB at 665 nm was tracked in 5 mL samples taken every 5 min during the degrading phase. Ultracentrifugation was used to isolate the catalyst, and then it was washed three times in de-ionized water before being dried at 60 °C. Finally, its reusability was put to the test.

3. Results and Discussion

3.1. Analysis of Crystal Pattern of the Photocatalysts

The phase structure can be analyzed using XRD, which is a non-destructive technique. The XRD pattern of the processed ZnO is displayed in Figure 1. Wurtzite ZnO (red line) has diffraction peaks at 31.78°, 34.32°, 36.23°, 47.48°, 56.46°, 62.69°, 66.32°, 67.88°, 69.09°, 72.55° and 76.87° corresponding to the (100), (002), (101), (101), (110), (103), (220), (112), (201), (004) and (202) planes (JCPDS card: 36-1451) [26]. The monoclinic body centered structure of Dy2WO6 is confirmed by JCPDS card: 26-0595 [42,43], which accounts for the 2θ values of 22.82°, 23.51°, 24.20°, 28.18°, 30.26°, 38.04°, 41.02°, 53.60°, 64.67° and 74.87° and the corresponding planes (011), (211), (130), (321), (231), (320), (132), (741), (441) and (622), respectively. It can be seen in DWZ (pink line) and DWZG patterns (dark yellow line) that loading of Dy2WO6, the 2θ angles of the (100), (002) and (101) planes for pure ZnO were displaced to the lower diffraction values (31.51°, 34.15°, 35.98°). Since there was such a small amount of GO mass loading, the GO peak had a relatively weak intensity [44]. After Dy2WO6 and GO loading, the crystalline structure of the ZnO-containing DWZG remained unmodified, demonstrating that no changes were observed in the final composites.

3.2. Morphology and Elemental Composition of the Photocatalysts

The surface shape of materials is predicted using a scanning electron microscope (EVO-80, CARL ZEISS), and the presenting elements of nanocomposites are located using energy dispersive X-ray spectroscopy (EDX) (AMETEK-EDAX (Z2e Analyzer)). Scanning electron microscopy (SEM) was used to examine the structure and morphology of the prepared nanocomposites. From SEM analysis, Dy2WO6, ZnO, Dy2WO6-ZnO and Dy2WO6-ZnO/GO nanocomposite, with their respective morphologies, are shown in Figure 2a–f. The unmodified Dy2WO6 catalyst has a structure resembling that of algae (Figure 2a,b). Figure 2c shows an uneven cluster form within the ZnO materials. Figure 2d shows the Dy2WO6-ZnO (DWZ) composite, which observed two different structures like coupled algae with a cluster shape. The SEM imaging of a Dy2WO6-ZnO/GO nanocomposite (shown in Figure 2e,f) confirms the presence of Dy2WO6 (algae), ZnO (cluster) and GO (layer) nanosheets. The photocatalytic performance of the catalyst may be improved by these various Dy2WO6-ZnO shapes [42]. Furthermore, the elemental composition of DWZG is shown in Figure 3 using EDS analysis. When properly deposited on the surface of the GO layer, the composite Dy2WO6-ZnO can boost the photocatalytic activity of the material. The elemental distribution of carbon (C), oxygen (O), dysprosium (Dy), tungsten (W) and zinc (Zn) in Dy2WO6-ZnO/GO nanocomposite photocatalysts is depicted in Figure 3 using EDS mapping. The EDS investigation verified that C, O, Dy, W and Zn elements were present in the proper ratios as DWZG’s nanocomposites.

3.3. Optical Properties by DRS-UV and Raman Spectroscopy

We used a Shimadzu UV-2450 spectrophotometer to acquire UV-Vis diffuse reflection spectra (DRS). Band energy values for samples and detectable light for pollutant degradation via band gap determination are both determined using UV-diffused reflectance spectroscopy (DRS-UV). The high UV absorption of the as-synthesized sample suggests that UV light is an essential activator for photocatalysis in this system (Figure 4a). The Tauc plot was used to calculate the band gap of the generated samples. Tauc’s plot of photon energy (hγ) and (αhγ)2 reveals values of 3.67 eV for DW, 3.22 eV for ZnO and 3.02 eV for the band energy of the DWZG composite (Figure 4b). The deterioration process is greatly influenced by the determination of the band energy value. Light degrades the pollutant by causing the production of e-h+ pairs, which occur when electrons are stimulated from the valence band to the conduction band. The photocatalytic degradation reaction in the UV range has been facilitated by band gap value 3.02 eV, which has also increased the efficiency of the reaction.
Raman spectroscopy was used to identify the GO layer material. The Raman spectra of DWZG materials are shown in Figure S2, where two distinct bands can be seen at 1345 and 1592 cm−1, which correspond to the D and G bands. It is feasible to calculate the values of the ID/IG ratio, and they are 0.952 for GO and 1.17 for rGO [45], respectively. The G band is located at 1592 cm−1 due to the sp2 hybridization. In comparison to rGO, GO has a lower ID/IG ratio of 0.952. GO has formed entirely, as evidenced here [46,47].

3.4. XPS Analysis

A comprehensive investigation into the presence of metals and their valence state in the components of as-prepared samples relies heavily on the results of X-ray photo electron spectroscopy. X-ray photoelectron spectroscopy (XPS) revealed the presence of carbon (C), dysprosium (Dy), oxygen (O), tungsten (W) and zinc (Zn) in the forming DWZG nanocomposite, as shown in Figure 5a–f. The XPS spectroscopic element formation is consistent with the EDX analysis. The survey spectrum of DWZG shows the Zn 3d, W 4f, Zn 3p, Dy 4d, C 1s, O 1s, W 4s and Zn 2p elements in Figure 5a. High magnification XPS spectra for C 1s (284.8, 286.4 and 288.4 eV), Dy 4d (158.7 eV), O 1s (530.3 and 533.1 eV), W 4f (34.6 and 36.9 eV) and Zn 2p (1022.5 and 1045.6 eV) are shown in Figure 5b–f. The XPS peak in the C 1s spectrum (Figure 5b) is proportional to the binding energy value of 284.8 eV (C-C) (yellow), 286.4 eV (C-O) (pink) and 288.4 eV (C=O) (cyan), respectively [48]. Having a valence of Dy4+, the core spectrum peak for Dy 4d5/2 is seen at 158.7 eV [42] in Figure 5c. The adsorbed oxygen species causes the high intensity peak in O 1s to occur at 530.3 eV (pink) and 533.1 eV (yellow) (Figure 5d) [48]. Figure 5e displays the 35.59 eV (cyan) and 38.03 eV (pink) [44] ranges for the W 4f5/2 and W 4f7/2 peaks corresponds to the W6+ chemical state, respectively. However, a new peak at 39.09 eV (yellow line) was observed for the W(4f5/2) peak, which was attributed to the creation of a W-C bond in the nanomaterials [49]. The binding energy of Zn 2p3/2 and Zn 2p1/2 is shown to be 1022.5 eV and 1045.6 eV [50], respectively, in the XPS spectra shown in Figure 5f. The effective synthesis of DWZG nanocomposite proves the binding energy values for the presented ingredients.

3.5. Estimation of Photodegradation Performance

The optimal conditions for MB pollutant degradation need an investigation into factors such as the number of catalysts used, the mass of the supporting material, the concentration of the MB and the presence of reactive oxidative species. Figure 6a shows, using different catalysts (such as DW, ZnO, DWZ, and DWZG), the affected MB degradation. The photodegradation efficiency of the different catalysts is 78.5% for DW, 86.3% for ZnO, 88.6% for DWZ and 98.2% for DWZG nanocomposites. In DWZG, the MB degrading process is finished in 30 min (98.2%), which is significantly faster than the other photocatalysts. Dye degradation rate kinetics can be calculated by graphing ln (C0/C) versus time. The results show that the rate constants for MB dye degradation for the host Dy2WO6 photocatalyst, ZnO photocatalyst, DWZ photocatalyst and DWZG photocatalyst are 0.0431 min−1, 0.0624 min−1, 0.0718 min−1 and 0.1215 min−1, respectively. The rate constant values of the obtained sample were illustrated in Figure S1a. Degradation of MB dye as a function of DWZG nanocomposite in the dark is shown in Figure S1b. When light is absent, adsorption can take place. The adsorption ratio for the breakdown of MB dye increases linearly with the amount of impurities present in the Dy2WO6 and ZnO nanomaterial. The ternary type DWZG nanocomposite is preferable because its adsorption percentage value is higher than those of the host DW, ZnO and DWZ nanomaterials. From the results, it demonstrates that the graphene oxide doped DWZG photocatalyst has a higher degradation performance and shows a high capacity for the adsorption of dye effluents [51]. In Figure 6b, absorption spectra of MB dye degradation using DWZG shows that the absorbance of MB decreased to nearly zero. The inefficiency of others also prevented the decline from being fully realized. Therefore, DWZG photocatalyst efficiency is significantly higher than that of alternative catalysts.
In heterogeneous photocatalytic degradation, the amount of materials utilized were critical. The catalyst dosage was varied from 20 mg to 40 mg in MB to determine its efficacy (Figure 6c). Optimization of the DWZG catalyst dose revealed an optimum dose between 10 and 30 mg, beyond which the rate of MB degradation is increased. In Figure 6c, the 30 mg of the catalyst system increases the surface area of the material and also increase the number of reactive sites [52]. From Figure S1c, the kinetics value for MB dye degradation under various catalyst dosage levels is illustrated. From the results, the 30 mg dosage level has a higher rate constant value than the other dosage level. Degradation efficiency drops if photocatalysts aggregate due to the loading of a large volume (40 mg) of reactive catalyst, reducing their active surface area. Additionally, higher solution turbidity may also reduce light penetration. One of the most important aspects of pollutant degradation is optimizing the concentration of the pollutants. The optimum range for MB concentration was found to be 10 µM to 30 µM (Figure 6d). The degradation of pollutants at 10 µM is optimal and steady relative to other factors. The kinetics value for different dye concentrations are shown in Figure S1d; the rate constant value for low dye concentration level is higher than the other concentration level. If the pollutant load were to grow, photodegradation would occur at a slower rate. Adsorption of molecules onto the catalyst’s surface could be the reason for the drop in performance. As can be seen in Figure 6d, a concentration of 10 µM is optimal for MB photodecomposition.
The reactive oxygen species (ROS) play a crucial role in photocatalysis; we have studied the effects of various scavengers on the photocatalytic degradation rate. Scavengers were added to the methylene blue degradation solution to detect the reactive oxygen species (ROS) produced during the photocatalytic activity. For species such as h+, O2•−, and OH radicals, we used triethanolamine (TEOA), benzoquinone (BQ), and isopropyl alcohol (IPA) [53,54]. As shown in Figure 7a, the MB dye degradation clearly decreased the percentage of degradation efficiency with the addition of TEOA (68.5%), IPA (38.2%) and BQ (50.3%) scavengers. The photocatalytic reactions are crucial to the degradation rates of MB, and the primary active species is OH radicals. This is in accordance with the suggested mechanism for the photocatalytic degradation of DWZG shown in Figure 8. In the diagram, the positive holes leave the valence band (VB), and the electrons go into the conduction band (CB) after being activated by the light of ZnO/GO. Electron-hole recombination occurs when electrons from Dy2WO6 in CB react with oxygen in the water, creating superoxide radical anion (O2•−) [42,55]. The hydroxyl radical (OH) is produced in VB when holes combine with water molecules. Nevertheless, the coupled h+ were making it harder for the water molecule to produce OH radicals in an indirect manner [55]. In conclusion, the MB photodegradation was caused by the reactive species OH and O2•− radicals.
The residual colloidal catalyst was filtered, dried and reused for the degradation of MB dye after the first run was completed. Under six iterations of photodegradation, the collective catalyst was continuously engaged in the reaction. Figure 7b shows that the MB dye can be recycled. In the reusability experiment, the effectiveness of degradation somewhat decreases as the number of counts per run increases (98.2% to 87.91%). Nonetheless, the effectiveness of MB degradation by DWZG plateaued after the sixth cycle. The data indicates that the catalyst is mechanically stable [56]. During the photocatalytic process, the GO content of Dy2WO6-ZnO/GO photocatalysts can undergo a chemical transformation, changing from an oxygenated to a reduced state [57,58,59,60].

4. Conclusions

In conclusion, the unique DWZG nanocomposite was synthesized using the hydrothermal approach, and its spectral information was described in detail. The DWZG photocatalyst shows extremely high degradation efficiency for MB dye when exposed to UV light. When compared to undoped ZnO, Dy2WO6 and DWZ, the catalytic efficiency of the DWZG nanocomposite is significantly higher. The XRD, XPS and EDS analysis confirmed the formation of DWZG nanocomposites. After being exposed to UV light for 30 min, the MB was completely degraded. According to research on reactive active species, hydroxide radicals have a key role in the degradation of MB dye.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/app13127145/s1, Figure S1. (a) Kinetic plot of different catalysts, (b) Adsorption study of different catalysts, (c) Kinetic plot of different dosage and (d) Kinetic plot of different concentration of MB, Figure S2. Raman spectra of DWZG nanocomposites.

Author Contributions

Conceptualization, K.S.; methodology, K.S.; software, M.A. and K.A.; validation, K.S., T.H.O. and V.P.; formal analysis, K.S. and K.A.; investigation, K.S. and M.A.; resources, T.H.O.; data curation, T.H.O. and V.P.; writing—original draft preparation, K.S.; writing—review and editing, T.H.O.; visualization, K.S.; supervision, T.H.O.; project administration, T.H.O.; funding acquisition, T.H.O. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIT), grant number 2022R1A2C1004283 and Korea Basic science Institute (National research Facilities and Equipment Center) grant funded by the Ministry of Education, grant number 2019R1A6C1010046. V.P. is grateful for the postdoctoral opportunities provided by the Knut and Alice Wallenberg Foundations. V.P. wished to thank KTH Royal Institute of Technology for the financial support to publish this article with open access.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Pastrana-Martínez, L.M.; Faria, J.L.; Doña-Rodríguez, J.M.; Fernández-Rodríguez, C.; Silva, A.M. Degradation of diphenhydramine pharmaceutical in aqueous solutions by using two highly active TiO2 photocatalysts: Operating parameters and photocatalytic mechanism. Appl. Catal. B Environ. 2012, 113, 221–227. [Google Scholar] [CrossRef]
  2. Ambigadevi, J.; Kumar, P.S.; Vo, D.V.N.; Haran, S.H.; Raghavan, T.S. Recent developments in photocatalytic remediation of textile effluent using semiconductor based nanostructured catalyst: A review. J. Environ. Chem. Eng. 2021, 9, 104881. [Google Scholar] [CrossRef]
  3. Fagan, R.; McCormack, D.E.; Dionysiou, D.D.; Pillai, S.C. A review of solar and visible light active TiO2 photocatalysis for treating bacteria, cyanotoxins and contaminants of emerging concern. Mater. Sci. Semicond. Process. 2016, 42, 2–14. [Google Scholar] [CrossRef] [Green Version]
  4. Hunge, Y.; Yadav, A.; Mathe, V. Oxidative degradation of phthalic acid using TiO2 photocatalyst. J. Mater. Sci. Mater. Electron. 2018, 29, 6183–6187. [Google Scholar] [CrossRef]
  5. Ali, T.; Hunge, Y.M.; Venkatraman, A. UV assisted photoelectrocatalytic degradation of reactive red 152 dye using spray deposited TiO2 thin films. J. Mater. Sci. Mater. Electron. 2018, 29, 1209–1215. [Google Scholar] [CrossRef]
  6. Hunge, Y.; Yadav, A.; Mathe, V. Ultrasound assisted synthesis of WO3-ZnO nanocomposites for brilliant blue dye degradation. Ultrason. Sonochem. 2018, 45, 116–122. [Google Scholar] [CrossRef] [PubMed]
  7. Hoffmann, M.R.; Martin, S.T.; Choi, W.Y.; Bahnemann, D.W. Environmental applications of semiconductor photocatalysis. Chem. Rev. 1995, 95, 69–96. [Google Scholar] [CrossRef]
  8. Chen, C.C.; Ma, W.H.; Zhao, J.C. Semiconductor-mediated photodegradation of pollutants under visible-light irradiation. Chem. Soc. Rev. 2010, 39, 4206–4219. [Google Scholar] [CrossRef]
  9. Wang, Y.; Jiang, T.; Meng, D.; Yang, J.; Li, Y.; Ma, Q.; Han, J. Fabrication of nanostructured CuO films by electrodeposition and their photocatalytic properties. Appl. Surf. Sci. 2014, 317, 414–421. [Google Scholar] [CrossRef]
  10. Chen, Q.; Wang, Y.; Zheng, M.; Fang, H.; Meng, X. Nanostructures confined self-assembled in biomimetic nanochannels for enhancing the sensitivity of biological molecules response. J. Mater. Sci. Mater. Electron. 2018, 29, 19757–19767. [Google Scholar] [CrossRef]
  11. Lee, D.H.; Kim, J.E.; Han, T.H.; Hwang, J.W.; Jeon, S.; Choi, S.Y.; Hong, S.H.; Lee, W.J.; Ruoff, R.S.; Kim, S.O. Versatile carbon hybrid films composed of vertical carbon nanotubes grown on mechanically compliant graphene films. Adv. Mater. 2010, 22, 1247–1252. [Google Scholar] [CrossRef]
  12. Özgür, Ü.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.; Doğan, S.; Avrutin, V.; Cho, S.J.; Morkoç, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 041301–0413103. [Google Scholar] [CrossRef] [Green Version]
  13. Chao, L.C.; Huang, J.W.; Chang, C.W. Annealing effects on the properties of Nd containing ZnO nanoparticles prepared by sol–gel process. Phys. B Condens. Matter. 2009, 404, 1301–1304. [Google Scholar] [CrossRef]
  14. Wang, R.; Xin, J.H.; Yang, Y.; Liu, V.; Xu, L.; Hu, J. The characteristics and photocatalytic activities of silver doped ZnO nanocrystallites. Appl. Surf. Sci. 2004, 227, 312–317. [Google Scholar] [CrossRef]
  15. Hunge, Y.; Yadav, A.; Kulkarni, S.; Mathe, V. A multifunctional ZnO thin film-based devices for photoelectrocatalytic degradation of terephthalic acid and CO2 gas sensing applications. Sens. Actuators B Chem. 2018, 274, 1–9. [Google Scholar] [CrossRef]
  16. Hunge, Y.M. Sunlight assisted photoelectrocatalytic degradation of benzoic acid using stratified WO3/TiO2 thin films. Ceram. Int. 2017, 43, 10089–10096. [Google Scholar] [CrossRef]
  17. El Bekkali, C.; Bouyarmane, H.; El Karbane, M.; Masse, S.; Saoiabi, A.; Coradin, T.; Laghzizil, A. Zinc oxide-hydroxyapatite nanocomposite photocatalysts for the degradation of ciprofloxacin and ofloxacin antibiotics. Colloids Surf. A Physicochem. Eng. Asp. 2018, 539, 364–370. [Google Scholar] [CrossRef] [Green Version]
  18. Priyadharsan, A.; Shanavas, S.; Vasanthakumar, V.; Balamuralikrishnan, B.; Anbarasan, P. Synthesis and investigation on synergetic effect of rGO-ZnO decorated MoS2 microflowers with enhanced photocatalytic and antibacterial activity. Colloids Surf. A Physicochem. Eng. Asp. 2018, 559, 43–53. [Google Scholar]
  19. Menon, N.G.; Tatiparti, S.S.V.; Mukherji, S. Synthesis, characterization and photocatalytic activity evaluation of TiO2–ZnO nanocomposites: Elucidating effect of varying Ti: Zn molar ratio. Colloids Surf. A Physicochem. Eng. Asp. 2019, 565, 47–58. [Google Scholar] [CrossRef]
  20. Lu, M.Y.; Lu, M.P.; Chung, Y.A.; Chen, M.J.; Wang, Z.L.; Chen, L.J. Intercrossed sheet-like Ga-doped ZnS nanostructures with superb photocatalytic activity and photo response. J. Phys. Chem. C 2009, 113, 12878–12882. [Google Scholar] [CrossRef]
  21. Wan, Q.; Li, Q.; Chen, Y.; Wang, T.H.; He, X.; Li, J.; Lin, C. Fabrication and ethanol sensing characteristics of ZnO nanowire gas sensors. Appl. Phys. Lett. 2004, 84, 3654–3656. [Google Scholar] [CrossRef] [Green Version]
  22. Schneider, S.L.; Lim, H.W. A review of inorganic UV filters zinc oxide and titanium dioxide. Photodermatol. Photoimmunol. Photomed. 2019, 35, 442–446. [Google Scholar] [CrossRef] [Green Version]
  23. Zheng, G.; Shang, W.; Xu, L. Enhanced photocatalytic activity of ZnO thin films deriving from a porous structure. Mater. Lett. 2015, 150, 1–4. [Google Scholar] [CrossRef]
  24. Nourmohammadi, A.; Rahighi, R.; Akhavan, O.; Moshfegh, A. Graphene oxide sheets involved in vertically aligned zinc oxide nanowires for visible light photoinactivation of bacteria. J. Alloys Compd. 2014, 612, 380–385. [Google Scholar] [CrossRef]
  25. Balachandran, S.; Swaminathan, M. The simple, template free synthesis of a Bi2S3–ZnO heterostructure and its superior photocatalytic activity under UV-A light. Dalton Trans. 2013, 42, 5338–5347. [Google Scholar] [CrossRef] [PubMed]
  26. Balachandran, S.; Prakash, N.; Thirumalai, K.; Muruganandham, M.; Sillanpää, M.; Swaminathan, M. Facile construction of heterostructured BiVO4–ZnO and its dual application of greater solar photocatalytic activity and self-cleaning property. Ind. Eng. Chem. Res. 2014, 53, 8346–8356. [Google Scholar] [CrossRef]
  27. Song, S.; Zhang, Y.; Xing, Y.; Wang, C.; Feng, J.; Shi, W.; Zheng, G.; Zhang, H. Rectangular AgIn(WO4)2 nanotubes: A promising photoelectric material. Adv. Funct. Mater. 2008, 18, 2328–2334. [Google Scholar] [CrossRef]
  28. Zhang, L.W.; Wang, Y.J.; Cheng, H.Y.; Yao, W.Q.; Zhu, Y.F. Synthesis of porous Bi2WO6 thin films as efficient visible-light-active photocatalysts. Adv. Mater. 2009, 21, 1286–1290. [Google Scholar] [CrossRef]
  29. Rahman, M.M.; Balkhoyor, H.B.; Asiri, A.M. Removal of a melamine contaminant with Ag-doped ZnO nanocomposite materials. New J. Chem. 2019, 43, 18848–18859. [Google Scholar] [CrossRef]
  30. Kerli, S. Boron-doped cobalt oxide thin films and its electrochemical properties. Mod. Phys. Lett. B 2016, 30, 1650343. [Google Scholar] [CrossRef]
  31. Raja, A.; Selvakumar, K.; Swaminathan, M.; Kang, M. Redox additive based rGO-Dy2WO6-ZnO nanocomposite for enhanced electrochemical supercapacitor applications. Synth. Met. 2021, 276, 116753. [Google Scholar] [CrossRef]
  32. Stankovich, S.; Dikin, D.A.; Dommett, G.B.; Kohlhaas, K.M.; Zimney, E.J.; Stach, E.A.; Piner, R.D.; Nguyen, S.T.; Ruoff, R.S. Graphene-based composite materials. Nature 2006, 442, 282–286. [Google Scholar] [CrossRef]
  33. Geim, A.K. Graphene: Status and prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [Green Version]
  34. Arunpandian, M.; Selvakumar, K.; Raja, A.; Rajasekaran, P.; Thiruppathi, M.; Nagarajan, E.R.; Arunachalam, S. Fabrication of novel Nd2O3/ZnO-GO nanocomposite: An efficient photocatalyst for the degradation of organic pollutants. Colloids Surf. A Physicochem. Eng. Asp. 2019, 567, 213–227. [Google Scholar] [CrossRef]
  35. Arunpandian, M.; Selvakumar, K.; Raja, A.; Thiruppathi, M.; Rajasekaran, P.; Rameshkumar, P.; Nagarajan, E.R.; Arunachalam, S. Development of novel Nd2WO6/ZnO incorporated on GO nanocomposite for the photocatalytic degradation of organic pollutants and biological studies. J. Mater. Sci. Mater. Electron. 2019, 30, 18557–18574. [Google Scholar] [CrossRef]
  36. Basumatary, B.; Basumatary, R.; Ramchiary, A.; Konwar, D. Evaluation of Ag@TiO2/WO3 heterojunction photocatalyst for enhanced photocatalytic activity towards methylene blue degradation. Chemosphere 2022, 286, 131848. [Google Scholar] [CrossRef]
  37. Yulizar, Y.; Apriandanu, D.O.B.; Surya, R.M. Fabrication of novel SnWO4/ZnO using Muntingia calabura L. leaf extract with enhanced photocatalytic methylene blue degradation under visible light irradiation. Ceram. Int. 2022, 48, 3564–3577. [Google Scholar]
  38. George, A.; Raj, D.M.A.; Venci, X.; Raj, A.D.; Irudayaraj, A.A.; Josephine, R.; Sundaram, S.J.; Al-Mohaimeed, A.M.; Al Farraj, D.A.; Chen, T.W. Photocatalytic effect of CuO nanoparticles flower-like 3D nanostructures under visible light irradiation with the degradation of methylene blue (MB) dye for environmental application. Environ. Res. 2022, 203, 111880. [Google Scholar] [CrossRef]
  39. Perumal, V.; Inmozhi, C.; Uthrakumar, R.; Robert, R.; Chandrasekar, M.; Mohamed, S.B.; Honey, S.; Raja, A.; Al-Mekhlafi, F.A.; Kaviyarasu, K. Enhancing the photocatalytic performance of surface-Treated SnO2 hierarchical nanorods against methylene blue dye under solar irradiation and biological degradation. Environ. Res. 2022, 209, 112821. [Google Scholar] [CrossRef] [PubMed]
  40. Mahanta, U.; Khandelwal, M.; Deshpande, A.S. TiO2@SiO2 nanoparticles for methylene blue removal and photocatalytic degradation under natural sunlight and low-power UV light. Appl. Surf. Sci. 2022, 576, 151745. [Google Scholar] [CrossRef]
  41. Estrada-Flores, S.; Martínez-Luévanos, A.; María Perez-Berumen, C.; Alfonso García-Cerda, L.; Emmanuel Flores-Guia, T. Relationship between morphology, porosity, and the photocatalytic activity of TiO2 obtained by sol–gel method assisted with ionic and nonionic surfactants. Bol. Soc. Esp. Ceram. Vidr. 2020, 59, 209–218. [Google Scholar] [CrossRef]
  42. Thirumalai, K.; Shanthi, M.; Swaminathan, M. Hydrothermal fabrication of natural sun light active Dy2WO6 doped ZnO and its enhanced photo-electrocatalytic activity and self-cleaning properties. RSC Adv. 2017, 7, 7509–7518. [Google Scholar] [CrossRef] [Green Version]
  43. Josephine, G.S.; Sivasamy, A. Nanocrystalline ZnO doped on lanthanide oxide Dy2O3: A novel and UV light active photocatalyst for environmental remediation. Environ. Sci. Technol. Lett. 2014, 1, 172–178. [Google Scholar] [CrossRef]
  44. Arunpandian, M.; Selvakumar, K.; Raja, A.; Rajasekaran, P.; Ramalingan, C.; Nagarajan, E.R.; Pandikumar, A.; Arunachalam, S. Rational design of novel ternary Sm2WO6/ZnO/GO nanocomposites: An affordable photocatalyst for the mitigation of carcinogenic organic pollutants. Colloids Surf. A Physicochem. Eng. Asp. 2020, 596, 124721. [Google Scholar]
  45. Wang, Y.; Selvakumar, K.; Oh, T.H.; Arunpandian, M.; Swaminathan, M. Boosting photocatalytic activity of single metal atom oxide anchored on TiO2 nanoparticles: An efficient catalyst for photodegradation of pharmaceutical pollutants. J. Alloys Compd. 2023, 950, 169821. [Google Scholar] [CrossRef]
  46. Akhavan, O. Bacteriorhodopsin as a superior substitute for hydrazine in chemical reduction of single-layer graphene oxide sheets. Carbon 2015, 81, 158–166. [Google Scholar] [CrossRef]
  47. Calizo, I.; Balandin, A.A.; Bao, W.; Miao, F.; Lau, C.N. Temperature Dependence of the Raman Spectra of Graphene and Graphene Multilayers. Nano Lett. 2007, 7, 2645–2649. [Google Scholar] [CrossRef] [PubMed]
  48. Selvakumar, K.; Raja, A.; Arunpandian, M.; Stalindurai, K.; Rajasekaran, P.; Sami, P.; Nagarajan, E.R.; Swaminathan, M. Efficient photocatalytic degradation of ciprofloxacin and bisphenol A under visible light using Gd2WO6 loaded ZnO/bentonite nanocomposite. Appl. Surf. Sci. 2019, 481, 1109–1119. [Google Scholar] [CrossRef]
  49. Akhavan, O.; Choobtashani, M.; Ghaderi, E. Protein Degradation and RNA Efflux of Viruses Photocatalyzed by Graphene−Tungsten Oxide Composite under Visible Light Irradiation. J. Phys. Chem. C 2012, 116, 9653–9659. [Google Scholar] [CrossRef]
  50. Zhang, P.; Shao, C.; Li, X.; Zhang, M.; Zhang, X.; Sun, Y.; Liu, Y. In situ assembly of well-dispersed Au nanoparticles on TiO2/ZnO nanofibers: A three-way synergistic heterostructure with enhanced photocatalytic activity. J. Hazard. Mater. 2012, 237, 331–338. [Google Scholar] [CrossRef]
  51. Meidanchi, A.; Akhavan, O. Superparamagnetic zinc ferrite spinel–graphene nanostructures for fast wastewater purification. Carbon 2014, 69, 230–238. [Google Scholar] [CrossRef]
  52. Zhang, L.; Yang, H.; Ma, J.; Li, L.; Wang, X.; Zhang, L.; Tian, S.; Wang, X. Controllable synthesis and shape-dependent photocatalytic activity of ZnO nanorods with a cone and different aspect ratios and of short-and-fat ZnO microrods by varying the reaction temperature and time. Appl. Phys. A 2010, 100, 1061–1067. [Google Scholar] [CrossRef]
  53. Raja, A.; Selvakumar, K.; Rajasekaran, P.; Arunpandian, M.; Ashokkumar, S.; Kaviyarasu, K.; Bahadur, S.A.; Swaminathan, M. Visible active reduced graphene oxide loaded titania for photodecomposition of ciprofloxacin and its antibacterial activity. Colloids Surf. A Physicochem. Eng. Asp. 2019, 564, 23–30. [Google Scholar] [CrossRef]
  54. Arunpandian, M.; Khan, M.R.; Busquets, R.; Selvakumar, K.; Raja, A.; Arunpandiyan, S.; Arivarasan, A.; Arunachalam, S.; Nagarajan, E.R. Fabrication of novel Ce2(WO4)3/ZnO@GO nanocomposite for superior photocatalytic performance under visible light and supercapacitor applications. Diam. Relat. Mater. 2022, 125, 109026. [Google Scholar] [CrossRef]
  55. Velmurugan, K.; Prakash, K.; Ponmari, G.; Selvam, V.; Anitha, C.; Joy Sinthiya, A.S.I. Hierarchical fabrication of GO@ Dy2MoO6 heterojunction for catalytic performance and effective wastewater treatment. Opt. Mater. 2023, 136, 113422. [Google Scholar] [CrossRef]
  56. Priya, A.; Arunachalam, P.; Selvi, A.; Madhavan, J.; Al-Mayouf, A.M. Synthesis of BiFeWO6/WO3 nanocomposite and its enhanced photocatalytic activity towards degradation of dye under irradiation of light. Colloids Surf. A Physicochem. Eng. Asp. 2018, 559, 83–91. [Google Scholar]
  57. Akhavan, O.; Ghaderi, E. Photocatalytic Reduction of Graphene Oxide Nanosheets on TiO2 Thin Film for Photoinactivation of Bacteria in Solar Light Irradiation. J. Phys. Chem. C 2009, 113, 20214–20220. [Google Scholar] [CrossRef]
  58. Akhavan, O. Photocatalytic reduction of graphene oxides hybridized by ZnO nanoparticles in ethanol. Carbon 2011, 49, 11–18. [Google Scholar] [CrossRef]
  59. Akhavan, O.; Abdolahad, M.; Esfandiar, A.; Mohatashamifar, M. Photodegradation of Graphene Oxide Sheets by TiO2 Nanoparticles after a Photocatalytic Reduction. J. Phys. Chem. C 2010, 114, 12955–12959. [Google Scholar] [CrossRef]
  60. Akhavan, O. Graphene Nanomesh by ZnO Nanorod Photocatalysts. ACS Nano 2010, 4, 4174–4180. [Google Scholar] [CrossRef]
Figure 1. X-ray powder diffraction patterns of graphene oxide (GO), ZnO, Dy2WO6-ZnO (DWZ) and Dy2WO6-ZnO/GO (DWZG) nanocomposite.
Figure 1. X-ray powder diffraction patterns of graphene oxide (GO), ZnO, Dy2WO6-ZnO (DWZ) and Dy2WO6-ZnO/GO (DWZG) nanocomposite.
Applsci 13 07145 g001
Figure 2. SEM images of (a,b) Dy2WO6, (c) ZnO, (d) Dy2WO6-ZnO, (e,f) Dy2WO6-ZnO/GO (DWZG) nanocomposite at different magnification.
Figure 2. SEM images of (a,b) Dy2WO6, (c) ZnO, (d) Dy2WO6-ZnO, (e,f) Dy2WO6-ZnO/GO (DWZG) nanocomposite at different magnification.
Applsci 13 07145 g002
Figure 3. EDS spectrum and elemental mapping of DWZG nanocomposites and weight percentage of presenting elements.
Figure 3. EDS spectrum and elemental mapping of DWZG nanocomposites and weight percentage of presenting elements.
Applsci 13 07145 g003
Figure 4. (a) DRS-UV absorption spectra and (b) Tauc plot of DW, ZnO and DWZG nanocomposites.
Figure 4. (a) DRS-UV absorption spectra and (b) Tauc plot of DW, ZnO and DWZG nanocomposites.
Applsci 13 07145 g004
Figure 5. XPS analysis of DWZG nanocomposite, (a) Survey spectrum, (b) C 1s, (c) Dy 4d, (d) O 1s, (e) W 4f and (f) Zn 2p elements.
Figure 5. XPS analysis of DWZG nanocomposite, (a) Survey spectrum, (b) C 1s, (c) Dy 4d, (d) O 1s, (e) W 4f and (f) Zn 2p elements.
Applsci 13 07145 g005
Figure 6. Photodegradation of MB solution under different conditions (a) different catalysts (catalysts: 30 mg, MB conc.: 1 × 10−5 M), (b) absorption spectrum of degradation of MB dye, (c) different dosage of DWZG (MB conc.: 1 × 10−5) and (d) different concentration of MB (DWZG: 30 mg).
Figure 6. Photodegradation of MB solution under different conditions (a) different catalysts (catalysts: 30 mg, MB conc.: 1 × 10−5 M), (b) absorption spectrum of degradation of MB dye, (c) different dosage of DWZG (MB conc.: 1 × 10−5) and (d) different concentration of MB (DWZG: 30 mg).
Applsci 13 07145 g006
Figure 7. (a) Photodegradation of MB dye solution under different scavengers and (b) recycle test for DWZG photocatalyst under UV light illumination.
Figure 7. (a) Photodegradation of MB dye solution under different scavengers and (b) recycle test for DWZG photocatalyst under UV light illumination.
Applsci 13 07145 g007
Figure 8. Schematic representation for photocatalytic mechanism of DWZG nanocomposite.
Figure 8. Schematic representation for photocatalytic mechanism of DWZG nanocomposite.
Applsci 13 07145 g008
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

Selvakumar, K.; Oh, T.H.; Arunpandian, M.; Aruchamy, K.; Polisetti, V. Hydrothermal Fabrication of GO Decorated Dy2WO6-ZnO Ternary Nanocomposites: An Efficient Photocatalyst for the Degradation of Organic Dye. Appl. Sci. 2023, 13, 7145. https://doi.org/10.3390/app13127145

AMA Style

Selvakumar K, Oh TH, Arunpandian M, Aruchamy K, Polisetti V. Hydrothermal Fabrication of GO Decorated Dy2WO6-ZnO Ternary Nanocomposites: An Efficient Photocatalyst for the Degradation of Organic Dye. Applied Sciences. 2023; 13(12):7145. https://doi.org/10.3390/app13127145

Chicago/Turabian Style

Selvakumar, Karuppaiah, Tae Hwan Oh, Muthuraj Arunpandian, Kanakaraj Aruchamy, and Veerababu Polisetti. 2023. "Hydrothermal Fabrication of GO Decorated Dy2WO6-ZnO Ternary Nanocomposites: An Efficient Photocatalyst for the Degradation of Organic Dye" Applied Sciences 13, no. 12: 7145. https://doi.org/10.3390/app13127145

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