Next Article in Journal
Fire Extinguishing Performance of Chemically Bonded Struvite Ceramic Powder with High Heat-Absorbing and Flame Retardant Properties
Next Article in Special Issue
An Experimental and Numerical Study of the Influence of Temperature on Mode II Fracture of a T800/Epoxy Unidirectional Laminate
Previous Article in Journal
Special Issue: “Structural and Thermal Properties of Polymeric Microspheres”
Previous Article in Special Issue
Bimetallic Thin-Walled Box Beam Thermal Buckling Response
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Effect of Different Morphologies of WO3/GO Nanocomposite on Photocatalytic Performance

by
Banu Esencan Türkaslan
1,*,
Aziz Kerim Çelik
1,
Ayça Dalbeyler
1 and
Nicholas Fantuzzi
2
1
Department of Chemical Engineering, Faculty of Engineering, University of Süleyman Demirel, Isparta 32260, Turkey
2
Department of Civil, Chemical, Environmental, and Materials Engineering, University of Bologna, 40126 Bologna, Italy
*
Author to whom correspondence should be addressed.
Materials 2022, 15(22), 8019; https://doi.org/10.3390/ma15228019
Submission received: 7 October 2022 / Revised: 7 November 2022 / Accepted: 9 November 2022 / Published: 14 November 2022
(This article belongs to the Special Issue Feature Papers in Materials Simulation and Design)

Abstract

:
Tungsten trioxide/graphene oxide (WO3/GO) nanocomposites have been successfully synthesized using in situ and ex situ chemical approaches. Graphite and tungsten carbide (WC) were employed to perform in situ synthesis, and WO3 and GO were employed to perform the ex situ synthesis of WO3/GO nanocomposites. GO, which was required for ex situ synthesis, is synthesized via the modified and improved Hummers method. XRD, SEM/EDS, and FTIR are used for the characterization of the nanocomposite. From the XRD of the WO3/GO nanocomposites, it was observed that WO3 distributed uniformly on graphene oxide sheets or was incorporated between the sheets. The photocatalytic activities of WO3/GO nanocomposites were evaluated by methylene blue (MB) adsorption and visible light photocatalytic degradation activities by UV-vis spectroscopy. The results showed that the efficiency of the photocatalytic activity of the nanocomposite depends on different synthesis methods and the morphology resulting from the changed method. WO3/GO nanocomposites synthesized by both methods exhibited much higher photocatalytic efficiencies than pure WO3, and the best degradation efficiencies for MB was 96.30% for the WO3/GO in situ synthesis nanocomposite.

1. Introduction

In recent years, due to a notable increase in industrial activities that produce wastewater, the development of easily accessible adsorbent materials for cleaning water resources gained great momentum. Organic dyes have a dangerous and toxic effect on humans, which are found in the wastewater of various industries [1].
Among the different methods, degradation by photocatalysis has become a promising technique as it is simple and sustainable, and it enables the conversion of dyes to non-hazardous waste.
Nanocomposite materials, including metal oxide nanoparticles, with their advanced physicochemical properties, are common photoactive semiconductor materials that make them suitable for use in photocatalysis [2,3].
Tungsten oxide (WO3), a nanostructured metal oxide, is widely used in the fields of gas detection [4,5], lithium-ion batteries [6,7], smart windows [8], and photocatalysis [9,10] practices. However, minimal photonic efficiency, the fast recombination rate of the charge carriers (photogenerated hole–electron), and low absorption ranges restrict the performance of WO3 as an efficient photocatalyst [11,12,13,14]. The most important features that distinguish an ideal adsorbent from others are high adsorption capacity, rapid adsorption rate, and high selectivity. In this sense, doping WO3 with other elements or compounds to improve its photocatalytic activity is deemed as an effective approach [15,16,17]. This may help improve the morphology of WO3 material as well as the affected band levels and characteristics of the charge carriers [18].
In particular, graphene and its derivatives are frequently preferred in improving the performance of metal oxide nanostructures due to their extraordinary properties [19,20,21].
Contrary to graphene, the use of hydrophilic GO, which contains various proportions of carbon, oxygen, and hydrogen in its structure, is expanding day by day due to its easy dispersion in solutions, its dielectric properties, transparency and adjustable electronic properties [22,23]. However, it is difficult to separate GO from water after adsorption. To overcome this drawback, the hybridization of GO with other inorganic or organic materials is an alternative [24].
Hummers and Offeman synthesized GO via the method of oxidizing graphite in H2SO4, NaNO3, and KMnO4 atmospheres in 1958 [25]. Hummers’ methodology was widely accepted, yet many disadvantages of this method have been reported, such as the production of toxic gas (NO2 and N2O4), the residual nitrate, and low efficiency levels. In the last 20 years, alternative methods have been attempted relative to Hummer’s method, including the addition of a peroxidation phase before KMnO4 oxidation (without NaNO3), increasing the amount of KMnO4 rather than NaNO3, and replacing KMnO4 with K2FeO4 when NaNO3 is extracted [26,27].
Studies conducted on WO3/GO show that while the photocatalytic degradation activities of composites formed by combining WO3 and GO structures that are synthesized separately via the ex situ method [28,29,30,31,32], there are no studies conducted regarding the photocatalytic degradation activities of nanocomposite structures, which are synthesized via in situ methods. It is thought that the composite structures synthesized using different methods are in the form of nanoparticles and nanowires [33,34,35,36], and this will affect the photocatalytic efficiency level.
In this study, highly efficient photocatalysts were developed, which is a method of producing renewable energy. To this end, both ex situ and in situ syntheses of WO3/GO composites were performed, and the photocatalytic activity of the formed structure was examined. GO was synthesized via the modified and improved Hummers method without using NaNO3. Given that there is no other WO3/GO study conducted to examine the photocatalytic degradation activity of nanocomposite structures synthesized via in situ methods, our study bears the characteristics of the first study conducted in this field, and it is also the first study that compares photocatalytic activities of ex-situ and in-situ synthesized WO3/GO composite structures.

2. Materials and Methods

Graphite flake (≥75% min), sulfuric acid (H2SO4, 98%), potassium permanganate (KMnO4, 99%), hydrogen peroxide (H2O2, 30%), hydrochloric acid (HCI, 37%), and tungsten (VI) oxide (WO3, <100 nm) were obtained from Sigma-Aldrich, and tungsten carbide (WC 45 nm, 99%) was obtained for the synthesis process from Nanokar company (İstanbul, Turkey).

2.1. Graphene Oxide (GO) Synthesis

GO was synthesized from layered graphite via the modified and improved Hummers method. Firstly, graphite (2 g) and then KMnO4 (6 g) were gradually added into H2SO4 within the ice bath and mixed. Later, 300 mL of deionized water was added to the mixture. In order to stop the oxidation process and remove the impurities in the structure, the mixture was filtered by adding H2O2 and HCl, respectively. Synthesized graphite oxide measuring 1 g was taken into 350 mL of pure water and dispersed for 3 h. After two hours of sonication in order to facilitate the exfoliation of clumped graphite oxide layers on GO layers, the mixture was centrifuged, and GO was produced.

2.2. In Situ and Ex Situ Synthesis of WO3/GO Composites

A total of 10 mg WO3 and 30 mg GO were separately mixed for 2 h in 10 mL and 30 mL water, respectively, for the ex situ synthesis of WO3/GO nanocomposites. Later, the two solutions were combined and mixed for another 2 h in a magnetic stirrer. The achieved solution was centrifuged and finally dried at 60 °C for 24 h to obtain the WO3/GO composite.
WC powder measuring 1 g and 10 mL H2SO4 were mixed in ice bath for the in situ synthesis of WO3/GO nanocomposite. KMnO4 measuring 3 g was gradually added into this mixture by continuously stirring. After adding KMnO4, the mixture was stirred for an additional 2 h; then, 15 mL H2O2 (%30 w/w) was added, and it was observed that the color of the mixture turned bright yellow. The achieved solution was centrifuged and finally dried at 60 °C for 24 h.

2.3. Photocatalytic Activities of WO3/GO Composites

Methylene blue (MB) was used as typical pollutants to study the photocatalytic activity of the synthesized WO3/GO composites, which are synthesized with in situ and ex situ methods. In order to simulate the coloring agent, 75 mL of 20 ppm MB solutions was prepared. WO3 was added to one of the solutions, and 15 mg of in situ WO3/GO and 15 mg of ex situ WO3/GO were added to other solutions as catalysts. Then, the solution was deposited into tubes in equal amounts in order to be able to make measurements at different time intervals.
After mixing for 30 min in the dark to ensure an adsorption–desorption balance, the solution tubes that were placed in the UV cabinet were exposed to a total of 2 xenon lamps, each of which was 150 Watt (Figure 1).
The distance between the lamp and the center of the tubes was measured as 8 cm. Samples were then taken at regular intervals to observe the degradation of methylene blue at 660 nm. The first measurement was made in the 15th minute; the next measurements were set to be made every 30 min, and regular measurements were made.

3. Results

The GO characterization was performed with X-ray diffraction (XRD) and the scanning electron microscopy (SEM/EDS) technique. WO3/GO composites were also evaluated by using a scanning electron microscope (SEM, Quanta Feg 250; FEI, Eindhoven, the Netherlands). WO3/GO composites were examined with a low-vacuum at 20.00 kV and 12.7–13.2 mm working distance at 5000× and 10000× magnifications, respectively. The elemental analysis of nanocomposites was carried out using an SEM microscope equipped with an energy-dispersive X-ray spectroscopy (EDX, Quanta Feg 250; FEI, Eindhoven, The Netherlands). The distribution and atomic composition of WO3/GO was examined using elemental mappings at an accelerating voltage of 20 kV. The crystalline phase of WO3/GO were examined by X-ray diffraction (XRD, Bruker D8 Advance Twin-Twin; Bruker, Karlsruhe, Germany) at 40 kV, 40 mA, and 1600 watts. In order to determine the photocatalytic capacity of WO3/GO nanocomposites, a UV-Vis spectrophotometer (UV-Vis Carry 60) device was used.

3.1. Characterization of GO

When the XRD spectrum of the GO structure was examined, it was observed that the peak formed at 2θ = 11.52° for GO, which is consistent with the results in the literature and it shows that the GO structure is obtained properly (Figure 2) [37].
SEM analysis results showed that the GO structure was formed by layered wavy structures piled on top of each other (Figure 3).

3.2. Characterization of WO3/GO Composites

Figure 4 provides the comparative FTIR spectra of GO and WO3/GO composites. The peaks of <1000 cm−1 within the composite structures show the existence of pure WO3 [28].
The FTIR spectrum of GO shows that the hydroxyl bond (-OH) is at 3425 cm−1, the carbonyl bond (C=O) is at 1719 cm−1, the aromatic bond (C=C) is at 1627cm−1, the epoxy bond is at 1627 cm−1, (C-O) is at 1400 cm−1, and the alkoxy bond (C-O) is at 1064 cm−1. The bands at 1715 cm−1 and 1617 cm−1 in the in situ synthesized composite structure and the bands at 1713 cm−1 and 1614 cm−1 in the ex situ synthesized composite structure belong to C=O and C=O vibrations, indicating the presence of GO in the composite’s structure.
Peaks under <1000 cm−1 observed in composites that are not observed in GO structure are caused by O-W-O stretch vibrations and show that nanoparticles bind to GO nanolayers strongly [38]. O-W-O vibrations observed approximately at 820 and 758 cm−1 in the in situ synthesized composite structure were observed less in ex situ structures.
SEM micrographs were used to identify the morphology of the synthesized composites and the location of the metal oxide in the carbon matrix. It was observed that while the pure WO3 structure comprised spongy structures of various sizes placed in such a way as to form spaces between them, WC consisted of randomly distributed and irregularly shaped coarse grains (Figure 5).
It is observed that the wavy interlayer spaces in GO structures are randomly dispersed by some spherical WO3 particles to form smoother surfaces in the WO3/GO ex situ synthesized nanocomposite (Figure 5e,f) structures. In WO3/GO insitu synthesis morphologies, however, it is seen that a single-phase homogeneous composite morphology formed with good interfacial interaction between GO and WC (Figure 5g,h). The homogeneous coating of the GO surface with WO3 as a result of good interfacial interaction in the in situ synthesis structure shows parallelisms with the O-W-O vibration bands observed in the FTIR results, while the lesser observation of these bands in the ex situ synthesis also supports the WO3 particle structure observed between the GO layers.
The EDS analysis of WO3 powders shows a tungsten atomic percentage of 79.35% and an oxygen atomic percentage of 20.65%. WC powders show tungsten at 90.34% and a carbon atomic percentage of 9.66%. After the synthesis of nanocomposites, a notable decrease in the atomic percentages of W elements was observed. Furthermore, the presence of GO was confirmed via EDS analyses, which showed additional carbon elements and oxygen elements in WO3/GO ex situ and in situ, respectively (Table 1).
Figure 6 shows the XRD patterns of the ex situ and in situ WO3/GO, WO3/GO nanocomposites, which confirms the presence of both WO3 and GO.
A low intensity peak at 10.8° indicates the formation of GO sheets in the in situ synthesis of WO3/GO, which is due to the poor crystalline nature of carbon. The other peaks at 2θ values of 23.3, 24.5, 34.2, 42.1, 47.49, and 50.22 confirm the presence of WO3 particles [39].
The WC diffraction spectrum shows three major intense peaks located at 2θ = 30.42°, 38.98°, and 47.03°, which correspond well to the crystallographic planes (001, 100, and 101) of WC, respectively [40].

3.3. Photocatalytic Degradation

Figure 7 shows the adsorption capacities (qe) of WO3 and WO3/GO ex situ and WO3/GO in situ nanocomposites, which were calculated using the following formula:
q e = C 0   C e   m V
where C0 (mg/L) and Ce (mg/L) refer, respectively, to the initial concentration of the coloring agent and the concentration of the coloring agent remaining in the solution after adsorption, m (g) refers to the amount of adsorbents, and V (mL) represents the volume of the solution. Accordingly, WO3 showed adsorption capacities in the range of 12.65–15.59, while WO3/GO ex situ and WO3/GO in situ nanocomposites showed adsorption capacities in the range of 15.18–19.84 and 19.46–23.91, respectively. During the experiment, the adsorption maximum capacity was determined with WO3/GO in situ nanocomposites at t = 270, and the lowest was determined with WO3 at t = 0. Within the increasing time intervals, the adsorption capacity of WO3/GO in situ nanocomposites showed a significant increase after 3 h. This shows that this situation can be associated with the surface area of the nanocomposite, thus leading to the understanding that the photocatalytic effect increases with time.
The degradation efficiency of was calculated using the following formula:
η % = ( 1 C C 0 ) × 100
where C0 is the absorption maximum at t = 0, and C is the absorption maximum after complete degradation.
WO3 and WO3/GO nanocomposites synthesized ex situ/in situ showed the maximum degradation of 75.79%, 90.52%, and 96.30% respectively (Figure 8). It was observed that the degradation amount of the WO3/GO in situ catalyst was higher than the one of ex situ synthesis, and it was concluded that this difference depends on the synthesis method. When GO is added to the matrix, it increased the photocatalytic effect, and in parallel with this, the synthesized nanocomposite structures reach higher percentage degradation efficiency values in a shorter time compared to WO3.
The chemical structure of MB has cationic atoms and aromatic rings. The degradation mechanism starts with the MB dye adsorption on the nanocomposite’s surface followed by its photodegradation, which can be summarized in the following steps. Figure 9 illustrates the mechanism of the photocatalytic degradation of MB via the WO3/GO nanocomposite’s UV irradiation. First, visible light radiation allows the transfer of electrons in in the valence band WO3 to the conduction band of GO. Therefore, holes (h) and electrons (e-) are formed on the surface of the WO3 photocatalyst. GO behaves as an electron acceptor via electrostatic and π−π stacking interactions. Then, while the holes react with the hydroxide ion, the electrons react with dissolved oxygen to produce OH-, which degrades MB dyes into non-toxic gases such as carbon dioxide and water. In addition, hydrogen peroxide reacts with electrons to produce more OH- to increase the degradation of the dye.
As shown in Table 2, prepared photocatalysts were compared with other WO3-based nanocomposites. Their photocatalytic efficiency is usually at 80–97%. In this work, WO3/GO nanocomposites prepared by in situ synthesis were higher than that of the tungsten-oxide-based counterparts. In addition, the method is simple and does not require synthesizing GO separately; the in situ oxidation of graphite is provided while the nanocomposite is formed.

4. Discussion

The photocatalytic activities of the WO3 and WO3/GO nanocomposites were evaluated by the degradation of MB in aqueous solutions. Compared with WO3, the photocatalytic activity of the WO3/GO nanocomposites was enhanced, and the best degradation efficiencies for MB was 96.30% for the WO3/GO in situ synthesis nanocomposite. It was attributed to the large surface area of GO, which served as an acceptor of the electrons generated in the WO3 and effectively decreased the recombination. To overcome the rapid recombination and slow migration of charge carriers, different morphologies have been developed, such as nanoplates, nanotubes, and nano-sheets. The homogeneous nanoplate structure formed in in situ syntheses showed higher photocatalytic effects due to its large surface area than the nanorod-like structure formed as a result of ex situ syntheses.
Although there are various different pollutant sources in the environment [42], the application of composite-based photocatalysts is limited to water treatments. Expanding the application areas of GO-based photocatalysts with different studies is necessary. Laboratory equipment is mostly used for the degradation processes in the laboratory. To prepare photocatalysts on a commercial scale, cost studies should be carried out for the large-scale degradation of pollutants, and systems should be modified with appropriate strategies. It is seen that suitable morphologies can be obtained by adjusting different methods and reaction conditions, and morphology control is an important parameter for photocatalytic activities.

5. Conclusions

In summary, we reported a simple chemical in situ and ex situ synthesis process and the physical properties and photocatalytic activities of WO3/GO composite structures. We observed that different synthesis methods affect WO3/GO’s morphology, while different WO3/GO morphologies affect photocatalytic performances. The ex situ preparation of the composite leads to the formation of well-dispersed WO3 with smoother surface in the WO3/GO. However, the in situ-prepared WO3 nanostructures have showed that single-phase homogeneous composite morphologies formed with good interfacial interactions between GO and WC. The prepared WO3 and its nanocomposite with GO was evidenced for the dye degradation of MB. The best degradation efficiencies for MB were 96.30% for the WO3/GO in situ synthesis nanocomposite, which are much better than that of WO3. The results showed that WO3/GO composites exhibited an enhanced WO3 photocatalysis efficiency in visible light. This study gave a new perspective for applications of WO3/GO nanocomposite photocatalysts for various areas.

Author Contributions

Conceptualization, B.E.T. and A.K.Ç.; methodology B.E.T., A.K.Ç. and N.F.; validation, B.E.T., A.D. and A.K.Ç.; formal analysis, N.F.; investigation, B.E.T. and A.D.; resources, B.E.T. and A.D.; data curation, B.E.T. and A.D.; writing—original draft preparation, B.E.T., A.D. and N.F.; writing—review and editing, B.E.T. and N.F.; visualization, B.E.T. and A.D.; supervision, B.E.T. and N.F.; project administration, B.E.T., A.K.Ç., A.D. and N.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by grants from the Scientific and Technological Research Council of Turkey (TUBITAK, project code: 1919B012001762).

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. Alam, M.W.; Aamir, M.; Farhan, M.; Albuhulayqah, M.; Ahmad, M.M.; Ravikumar, C.; Dileep Kumar, V.; Ananda Murthy, H. Green Synthesis of Ni-Cu-Zn Based Nanosized Metal Oxides for Photocatalytic and Sensor Applications. Crystals 2021, 11, 1467. [Google Scholar] [CrossRef]
  2. Hossain, S.; Chu, W.S.; Lee, C.S.; Ahn, S.H.; Chun, D.M. Photocatalytic performance of few-layer Graphene/WO3 thin films prepared by a nano-particle deposition system. Mater. Chem. Phys. 2019, 226, 141–150. [Google Scholar] [CrossRef]
  3. Alam, M.W.; Al Qahtani, H.S.; Souayeh, B.; Ahmed, W.; Albalawi, H.; Farhan, M.; Abuzir, A.; Naeem, S. Novel Copper-Zinc-Manganese Ternary Metal Oxide Nanocomposite as Heterogeneous Catalyst for Glucose Sensor and Antibacterial Activity. Antioxidants 2022, 11, 1064. [Google Scholar] [CrossRef] [PubMed]
  4. Chatterjee, S.G.; Chatterjee, S.; Ray, A.K.; Chakraborty, A.K. Graphene–metal oxide nanohybrids for toxic gas sensor. Sens. Actuators B Chem. 2015, 221, 1170–1181. [Google Scholar] [CrossRef]
  5. Chang, X.; Zhou, Q.; Sun, S.; Shao, C.; Lei, Y.; Liu, T.; Yin, Y. Graphene-tungsten oxide nanocomposites with highly enhanced gas-sensing performance. J. Alloys Compd. 2017, 705, 659–667. [Google Scholar] [CrossRef]
  6. Tian, W.; Liu, X.; Yu, W. Research Progress of Gas Sensor Based on Graphene and Its Derivatives. Appl. Sci. 2018, 8, 1118. [Google Scholar] [CrossRef] [Green Version]
  7. Zhang, J.; Liu, X.; Neri, G.; Pinna, N. Nanostructured materials for room-temperature gas sensors. Adv. Mater. 2016, 28, 795–831. [Google Scholar] [CrossRef]
  8. Le, X.V.; Luu, T.L.A.; Nguyen, H.L.; Nguyen, C.T. Synergistic enhancement of ammonia gassensing properties at low temperature by compositing carbon nanotubes with tungsten oxide nanobricks. Vacuum 2019, 168, 108861. [Google Scholar] [CrossRef]
  9. Korotcenkov, G.; Brinzari, V.; Ivanov, M.; Cerneavschi, A.; Rodríguez, J.R.; Cirera, A.; Cornet, A.; Morante, J.R. Structural stability of indium oxide films deposited by spray pyrolysis during thermal annealing. Thin Solid Films 2005, 479, 38–51. [Google Scholar] [CrossRef]
  10. Righettoni, M.; Amann, A.; Pratsinis, S.E. Breath analysis by nanostructured metal oxides as chemoresistive gas sensors. Mater. Today 2015, 18, 163–171. [Google Scholar] [CrossRef]
  11. Sayama, K.; Hayashi, H.; Arai, T.; Yanagida, M.; Gunji, T.; Sugihara, H. Highly active WO3 semiconductor photocatalyst prepared from amorphous peroxo-tungstic acid for the degradation of various organic compounds. Appl. Catal. B Environ. 2010, 94, 150–157. [Google Scholar] [CrossRef]
  12. Guo, Y.; Quan, X.; Lu, N.; Zhao, H.; Chen, S. High photocatalytic capability of self-assembled nanoporous WO3 with preferential orientation of (002) planes. Environ. Sci. Technol. 2007, 41, 4422–4427. [Google Scholar] [CrossRef] [PubMed]
  13. Sonia, S.; Kumar, P.S.; Mangalaraj, D.; Ponpandian, N.; Viswanathan, C. Influence of growth and photocatalytic properties of copper selenide (CuSe) nanoparticles using reflux condensation method. Appl. Surf. Sci. 2013, 283, 802–807. [Google Scholar] [CrossRef]
  14. Deng, F.; Pei, X.; Luo, Y.; Luo, X.; Dionysiou, D.D.; Wu, S.; Luo, S. Fabrication of hierarchically porous reduced graphene oxide/SnIn4S8 composites by a low-temperature coprecipitation strategy and their excellent visible-light photocatalytic mineralization performance. Catalysts 2016, 6, 113. [Google Scholar] [CrossRef] [Green Version]
  15. Sun, S.; Wang, W.; Zeng, S.; Shang, M.; Zhang, L. Preparation of ordered mesoporous Ag/WO3 and its highly efficient degradation of acetaldehyde under visible-light irradiation. J. Hazard. Mater. 2010, 178, 427–433. [Google Scholar] [CrossRef]
  16. Zheng, H.; Mathe, M. Hydrogen evolution reaction on single crystal WO3/C nanoparticles supported on carbon in acid and alkaline solution. Int. J. Hydrogen Energy 2011, 36, 1960–1964. [Google Scholar] [CrossRef]
  17. Cao, L.; Yuan, J.; Chen, M.; Shangguan, W. Photocatalytic energy storage ability of TiO2-WO3 composite prepared by wet-chemical technique. Res. J. Environ. Sci. 2010, 22, 454–459. [Google Scholar] [CrossRef]
  18. Kalanur, S.S.; Seo, H. Aligned nanotriangles of tantalum doped tungsten oxide for improved photoelectrochemical water splitting. J. Alloys Compd. 2019, 785, 1097–1105. [Google Scholar] [CrossRef]
  19. Chen, H.; Chen, Z.; Yang, H.; Wen, L.; Yi, Z.; Zhou, Z.; Dai, B.; Zhang, J.; Wu, X.; Wu, P. Multi-mode surface plasmon resonance absorber based on dart-type single-layer graphene. RSC Adv. 2022, 12, 7821–7829. [Google Scholar] [CrossRef]
  20. Zhang, Z.; Cai, R.; Long, F.; Wang, J. Development and application of tetrabromobisphenol A imprinted electrochemical sensor based on graphene/carbon nanotubes three-dimensional nanocomposites modified carbon electrode. Talanta 2015, 134, 435–442. [Google Scholar] [CrossRef]
  21. Shangguan, Q.; Chen, Z.; Yang, H.; Cheng, S.; Yang, W.; Yi, Z.; Wu, X.; Wang, S.; Yi, Y.; Wu, P. Design of Ultra-Narrow Band Graphene Refractive Index Sensor. Sensors 2022, 22, 6483. [Google Scholar] [CrossRef] [PubMed]
  22. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of graphene oxide (GO) by modified hummers method and its thermal reduction to obtain reduced graphene oxide (rGO). Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef] [Green Version]
  23. Khan, Z.U.; Kausar, A.; Ullah, H.; Badshah, A.; Khan, W.U. A review of graphene oxide, graphene buckypaper, and polymer/graphene composites: Properties and fabrication techniques. J. Plast. Film Sheeting 2016, 32, 336–379. [Google Scholar] [CrossRef]
  24. Cheng, Z.; Liao, J.; He, B.; Zhang, F.; Zhang, F.; Huang, X.; Zhou, L. One-Step Fabrication of Graphene Oxide Enhanced Magnetic Composite Gel for Highly Efficient Dye Adsorption and Catalysis. ACS Sustain. Chem. Eng. 2015, 3, 1677–1685. [Google Scholar] [CrossRef]
  25. Hummers, W.S., Jr.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  26. Yu, H.; Zhang, B.; Bulin, C.; Li, R.; Xing, R. High-efficient synthesis of graphene oxide based on improved hummers method. Sci. Rep. 2016, 6, 36143. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Yazıcı, M.; Tiyek, İ.; Ersoy, M.S.; Alma, M.H.; Dönmez, U.; Yıldırım, B.; Salan, T.; Karataş, Ş.; Uruş, S.; Karteri, İ.; et al. Modifiye Hummers Yöntemiyle Grafen Oksit (GO) Sentezi ve Karakterizasyonu. GU J. Sci. Part C 2016, 4, 43–50. [Google Scholar]
  28. Jeevitha, G.; Abhinayaa, R.; Mangalaraj, D.; Ponpandian, N. Tungsten oxide-graphene oxide (WO3-GO) nanocomposite as an efficient photocatalyst, antibacterial and anticancer agent. J. Phys. Chem. Solids 2018, 116, 137–147. [Google Scholar] [CrossRef]
  29. Fu, L.; Xia, T.; Zheng, Y.; Yang, J.; Wang, A.; Wang, Z. Preparation of WO3-reduced graphene oxide nanocomposites with enhanced photocatalytic property. Ceram. Int. 2015, 41, 5903–5908. [Google Scholar] [CrossRef]
  30. Kofuji, Y.; Isobe, Y.; Shiraishi, Y.; Sakamoto, H.; Ichikawa, S.; Tanaka, S.; Hirai, T. Hydrogen peroxide production on a carbon nitride–boron nitride-reduced graphene oxide hybrid photocatalyst under visible light. ChemCatChem 2018, 10, 2070–2077. [Google Scholar] [CrossRef]
  31. Ismail, A.A.; Faisal, M.; Al-Haddad, A. Mesoporous WO3-graphene photocatalyst for photocatalytic degradation of Methylene Blue dye under visible light illumination. Res. J. Environ. Sci. 2018, 66, 328–337. [Google Scholar] [CrossRef]
  32. Zhou, M.; Yan, J.; Cui, P. Synthesis and enhanced photocatalytic performance of WO3 nanorods@ graphene nanocomposites. Mater. Lett. 2012, 89, 258–261. [Google Scholar] [CrossRef]
  33. Bragaglia, M.; Paleari, L.; Lamastra, F.R.; Puglia, D.; Fabbrocino, F.; Nanni, F. Graphene nanoplatelet, multiwall carbon nanotube, and hybrid multiwall carbon nanotube–graphene nanoplatelet epoxy nanocomposites as strain sensing coatings. J. Reinf. Plast. Compos. 2021, 40, 632–643. [Google Scholar] [CrossRef]
  34. Pittella, E.; D’Alvia, L.; Palermo, E.; Piuzzi, E. Microwave Characterization of 3D Printed PLA and PLA/CNT Composites. In Proceedings of the 2021 IEEE 6th International Forum on Research and Technology for Society and Industry (RTSI), Rome, Italy, 7–9 June 2021. [Google Scholar] [CrossRef]
  35. Paleari, L.; Bragaglia, M.; Fabbrocino, F.; Nanni, F. Structural Monitoring of Glass Fiber/Epoxy Laminates by Means of Carbon Nanotubes and Carbon Black Self-Monitoring Plies. Nanomaterials 2021, 11, 1543. [Google Scholar] [CrossRef]
  36. Izadi, R.; Tuna, M.; Trovalusci, P.; Ghavanloo, E. Torsional Characteristics of Carbon Nanotubes: Micropolar Elasticity Models and Molecular Dynamics Simulation. Nanomaterials 2021, 11, 453. [Google Scholar] [CrossRef] [PubMed]
  37. Hanifah, M.F.R.; Jaafar, J.; Aziz, M.; Ismail, A.F.; Rahman, M.A.; Othman, M.H.D. Synthesis of kaphene oxide nanosheets via modified hummers method and ıts physicochemical properties. J. Technol. 2015, 74, 189–192. [Google Scholar] [CrossRef] [Green Version]
  38. Murugan Vadivel, A.; Muraliganth, T.; Manthiram, A. Rapid, Facile Microwave-Solvothermal Synthesis of Graphene Nanosheets and Their Polyaniline Nanocomposites for Energy Strorage. Chem. Mater. 2009, 21, 5004–5006. [Google Scholar] [CrossRef]
  39. Hu, X.; Xu, P.; Gong, H.; Yin, G. Synthesis and Characterization of WO3/Graphene Nanocomposites for Enhanced Photo-catalytic Activities by One-Step In-Situ Hydrothermal Reaction. Materials 2018, 11, 147. [Google Scholar] [CrossRef] [Green Version]
  40. Vakhrushev, A.V. Synthesis of WO3 Nanostructures and Their Nanocomposites with Graphene Derivatives via Novel Chemical Approach. In Nanomechanics—Theory and Application; IntechOpen: London, UK, 2021; 146p. [Google Scholar] [CrossRef]
  41. Tie, L.; Yu, C.; Zhao, Y.; Chen, H.; Yang, S.; Sun, J.; Dong, S.; Sun, J. Fabrication of WO3 na-norods on reduced graphene oxide sheets with augmented visible light photocatalytic ac-tivity for efficient mineralization of dye. J. Alloys Compd. 2018, 769, 83–91. [Google Scholar] [CrossRef]
  42. Derradji, M.; Mehelli, O.; Liu, W.; Fantuzzi, N. Sustainable and Ecofriendly Chemical Design of High Performance Bio-Based Thermosets for Advanced Applications. Front. Chem. 2021, 9, 691117. [Google Scholar] [CrossRef]
Figure 1. Designed UV cabinet (a) and tubes prepared for measurements to be made with different time intervals (b).
Figure 1. Designed UV cabinet (a) and tubes prepared for measurements to be made with different time intervals (b).
Materials 15 08019 g001
Figure 2. XRD patterns of GO.
Figure 2. XRD patterns of GO.
Materials 15 08019 g002
Figure 3. SEM/EDS image of (a) graphite and (b) GO.
Figure 3. SEM/EDS image of (a) graphite and (b) GO.
Materials 15 08019 g003
Figure 4. FTIR Spectra of GO and WO3/GO Composites.
Figure 4. FTIR Spectra of GO and WO3/GO Composites.
Materials 15 08019 g004
Figure 5. SEM images of WO3 (a,b), WC (c,d), WO3/GO ex situ synthesis (e,f), and WO3/GO in situ synthesis (g,h).
Figure 5. SEM images of WO3 (a,b), WC (c,d), WO3/GO ex situ synthesis (e,f), and WO3/GO in situ synthesis (g,h).
Materials 15 08019 g005
Figure 6. XRD patterns of WO3/GO ex situ synthesis (a) and WO3/GO in situ synthesis (b).
Figure 6. XRD patterns of WO3/GO ex situ synthesis (a) and WO3/GO in situ synthesis (b).
Materials 15 08019 g006
Figure 7. Relationship between adsorption capacity and time.
Figure 7. Relationship between adsorption capacity and time.
Materials 15 08019 g007
Figure 8. Degradation efficiency of MB.
Figure 8. Degradation efficiency of MB.
Materials 15 08019 g008
Figure 9. Schematic representation of photocatalytic degradation of methylene blue by WO3/GO nanocomposites under UV irradiation illumination.
Figure 9. Schematic representation of photocatalytic degradation of methylene blue by WO3/GO nanocomposites under UV irradiation illumination.
Materials 15 08019 g009
Table 1. Elemental composition of WO3, WC and WO3/GO in ex situ synthesis and in WO3/GO in situ synthesis.
Table 1. Elemental composition of WO3, WC and WO3/GO in ex situ synthesis and in WO3/GO in situ synthesis.
Samples% at.
OxygenTungstenCarbon
WO320.6579.35-
WC-90.349.66
WO3/GO ex situ41.108.6250.28
WO3/GO in situ26.9716.7156.32
Table 2. Comparison of photocatalytic performance between this work and reported references.
Table 2. Comparison of photocatalytic performance between this work and reported references.
PhotocatalystMethods of SynthesisPhotodecompositionPhotocatalytic EffciencyReferences
WO3/GOIn situ
ex situ chemical oxidation
MB96.30%
90.52%
Current work
WO3/GOUltrasonication MethodMB97.03%[28]
WO3/GOSol-gel methodMB82%[31]
WO3/GOPhoto-reduction methodMO92.7%[32]
WO3/GRHydrothermal methodMB83%[39]
WO3/rGOIn situ slvothermal methodMB94%[41]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Esencan Türkaslan, B.; Çelik, A.K.; Dalbeyler, A.; Fantuzzi, N. The Effect of Different Morphologies of WO3/GO Nanocomposite on Photocatalytic Performance. Materials 2022, 15, 8019. https://doi.org/10.3390/ma15228019

AMA Style

Esencan Türkaslan B, Çelik AK, Dalbeyler A, Fantuzzi N. The Effect of Different Morphologies of WO3/GO Nanocomposite on Photocatalytic Performance. Materials. 2022; 15(22):8019. https://doi.org/10.3390/ma15228019

Chicago/Turabian Style

Esencan Türkaslan, Banu, Aziz Kerim Çelik, Ayça Dalbeyler, and Nicholas Fantuzzi. 2022. "The Effect of Different Morphologies of WO3/GO Nanocomposite on Photocatalytic Performance" Materials 15, no. 22: 8019. https://doi.org/10.3390/ma15228019

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