Next Article in Journal
Effects of Surface Softening on the Mechanical Properties of an AISI 316L Stainless Steel under Cyclic Loading
Next Article in Special Issue
Plasmonic Gold Nanorod Size-Controlled: Optical, Morphological, and Electrical Properties of Efficiency Improved Tin Disulfide Vacuum-Free Hybrid Solar Cells
Previous Article in Journal
Deformation Characteristics in a Stretch-Based Dimensional Correction Method for Open, Thin-Walled Extrusions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Environmentally Benign Organic Dye Conversion under UV Light through TiO2-ZnO Nanocomposite

by
Sandip M. Deshmukh
1,
Sudhir S. Arbuj
2,
Santosh B. Babar
1,
Shoyebmohamad F. Shaikh
3,
Asiya M. Tamboli
4,*,
Nguyen Tam Nguyen Truong
4,
Chang-Duk Kim
5,
Sanjay M. Khetre
6,*,
Mohaseen S. Tamboli
4,* and
Sambhaji R. Bamane
7
1
Department of Chemistry, VNBN Mahavidyalaya, Shirala 415408, Maharashtra, India
2
Centre for Materials for Electronics Technology, Materials of Renewable Energy and Sensor Division, Panchawati Off Pahsan Road, Pune 411008, Maharashtra, India
3
Department of Chemistry, College of Science, King Saud University, Riyadh 11451, Saudi Arabia
4
School of Chemical Engineering, Yeungnam University, 280 Daehak-ro, Gyeongsan 38541, Korea
5
Department of Physics, Kyungpook National University, 80 Daehakro, Bukgu, Daegu 41566, Korea
6
Nanomaterials Research Laboratory, Department of Chemistry, Dahiwadi College, Dahiwadi 415508, Maharashtra, India
7
Department of Chemistry, Sushila Shankarrao Gadhave Mahavidyalaya, Khandala 412802, Maharashtra, India
*
Authors to whom correspondence should be addressed.
Metals 2021, 11(11), 1787; https://doi.org/10.3390/met11111787
Submission received: 5 October 2021 / Revised: 23 October 2021 / Accepted: 2 November 2021 / Published: 6 November 2021

Abstract

:
In this work, we developed a very simple and novel approach for synthesizing TiO2-ZnO nanocomposites via the urea-assisted thermal decomposition of titanium oxysulphate and zinc acetate at different weight ratios. The synthesized nanocomposite samples were studied by means of HR-TEM, XRD, STEM, UV–Vis DRS, PL and EDS. The observed results demonstrate that the TiO2-ZnO nanocomposite consists of an anatase crystal phase of TiO2 with a crystallite size of 10–15 nm. Combined characterization, including UV–Vis DRS, STEM, EDS and HR-TEM, revealed the successful formation of a heterojunction between TiO2 and ZnO and an improvement in UV spectrum absorption. The photocatalytic activity was explored using MO degradation under ultraviolet light illumination. The results of the optimized TiO2-ZnO nanocomposite show excellent photocatalytic activity and photostability over a number of degradation reaction cycles. In addition, the current approach has immense potential to be used as a proficient method for synthesizing mixed metal oxide nanocomposites.

1. Introduction

The major pollutants released into the ecosystem due to increased population and industrialization have turned out to be an environmental concern. These pollutants have been converted into nontoxic as well as less toxic substances using various physical, chemical and biological techniques. Recently, advanced oxidative methods such as semiconductor metal oxide-based photocatalysis have received enormous attention for their effective degradation of pollutants present in air, water and soil. This technique is low cost, is environmentally friendly, can function at ambient reaction conditions and is useful for a wide range of pollutants [1,2,3].
Numerous metal oxides such as ZnO, TiO2, Fe2O3, Ta2O5 and Bi2O3 have been used as photocatalysts for the degradation of various aqueous dye pollutants [4,5,6,7,8]. Among the various metal oxides, TiO2 has been used as a photocatalytic material with great potential due to its excellent photocatalytic activity, wide energy band gap (3.2 eV), nontoxic nature, chemical stability and cost effectiveness [9]. However, the main problem with TiO2 is its wide band gap of 3.2 eV, which requires UV light for performing photocatalytic reactions [10].
To date, different approaches have been tested in order to improve the photocatalytic activity of TiO2 such as cation and anion doping to alter the band gap, loading of noble metal nanoparticles and coupling with other semiconductor-based metal oxides/sulfides for the effective separation of photogenerated charge carriers [11,12,13]. The coupling of TiO2 with different metal oxides such as CuO, NiO, Ta2O5, Nb2O5 and ZnO shows a great enhancement in photocatalytic activity due to the increase in the lifetime of photogenerated charge carriers, the effective electron–hole pair separation and the modification of optical properties [14,15,16,17,18].
Among the various studied metal oxides, the coupling of ZnO with TiO2 creates one of the most efficient photocatalysts because of the development of heterojunction structures. Notably, ZnO possesses a wide band gap of 3.37 eV, high electron mobility, large exciton binding energy (60 meV) and similar band gap energies to TiO2 [19]. It is suggested that the formation of the type II heterojunction in the TiO2-ZnO composite might possibly facilitate the transfer of electrons from the conduction band (CB) of TiO2 to the CB of ZnO under UV light irradiation. Consequently, the electron–hole recombination rate is suppressed and photocatalytic activity is enhanced [20]. Further, the TiO2-ZnO nanocomposite’s activity can be enhanced by changing the synthesis method.
More recently, some methods have been developed for the fabrication of TiO2-ZnO nanocomposites including sol–gel, hydrothermal, atomic layer deposition and microwave methods [21,22,23,24]. Despite these advanced methods, these strategies possess some drawbacks such as the fact that they are time consuming and costly and involve hazardous substrates and complicated reaction procedures. In this regard, we have developed a thermal decomposition approach for the large-scale synthesis of the nanostructured TiO2-ZnO heterostructure as a rapid, easy, low-cost and eco-friendly technique. TiO2-ZnO nano-heterostructures were prepared by varying the ZnO concentration, and we studied their photocatalytic performance towards the degradation of aqueous methylene blue dye. To the best of our knowledge, this thermal decomposition approach for the one-step synthesis of TiO2-ZnO nano-heterostructures is the first attempt of its kind and could be the easiest possible method.

2. Experimental Section

2.1. Materials

Titanium oxysulphate (TiOSO4.H2O) zinc acetate (Zn(CH3CO2)2·2H2O) and urea (CON2H4) were purchased from Sigma-Aldrich (Bangalore, India). Methyl orange purchased from Molychem (Mumbai, India) was of analytical grade and was used as received, without further purification.

2.2. Preparation of TiO2, ZnO and TiO2-ZnO Composite

TiO2 and ZnO nanostructures were separately prepared using a urea-assisted thermal decomposition method in a similar manner to our previous report [25]. In the typical synthesis of the TiO2-ZnO nanocomposite, grinding of titanium oxysulphate, zinc acetate and urea was carried out for 15 min using a mortar and pestle. The obtained powder was heated at 600 °C in a muffle furnace with a heating rate of 15 °C per minute for 3 h. Then, the final product was washed with distilled water (DW) and dried in the oven at 80 °C. The obtained products were finally ground into a fine powder in a mortar and pestle. TiO2 and ZnO were also synthesized with the addition of urea and denoted as TU and ZU, respectively. The precursors (titanium oxysulphate and zinc acetate) had weight ratios of 0.95:0.05, 0.90:0.10, 0.85:0.15, 0.80:0.20, 0.75:0.25 and 0.70:0.30 and were denoted as TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70, respectively. The as-prepared products were used for further characterization.

2.3. Characterization

The crystalline structures of the as-prepared TiO2 and ZnO nanomaterials and TiO2-ZnO nanocomposites were identified by the X-ray diffraction technique (Pan analytical diffractometer, Almelo, The Netherlands) using CuKα radiation (λ = 1.5406 Å), which provides detection limits in the range of 0.1 to 1 wt.% per phase. With the help of field-emission scanning electron microscopy (FE-SEM, Hitachi S-4800 with an accelerating voltage range of 10.0 kV), field-emission transmission electron microscopy (Titan G2 ChemiSTEM Cs Probe, FEI Company, Hillsboro, OR, USA) the morphological features of the samples were investigated. The chemical composition and element allocation in a selected area were analyzed by scanning transmission electron microscopy elemental mapping using an X-ray energy-dispersive spectrometer (Titan G2 ChemiSTEM Cs Probe, FEI Company, Hillsboro, OR, USA). The absorption profile and band gap of the composites were studied using a UV−Vis spectrophotometer (Shimadzu, Model-UV-3600, Kyoto, Japan), and photoluminescence spectra (PL) were analyzed using a spectrofluorophotometer (JASCO, Model FP.750, Tokyo, Japan).

2.4. Evaluation of Photocatalytic Performance

The photocatalytic activities of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70 were evaluated by observing the degradation of methyl orange (MO), as an organic dye, at room temperature. For this purpose, 0.1 g of the photocatalyst was dispersed in 100 mL of aqueous solution with 20 ppm of MO dye. A Philips HPL-N (Amsterdam, Netherlands, 250 W, wavelength range of 200–600 nm) lamp was utilized as a UV–visible light source, at room temperature. Before the irradiation, the reaction mixture was stirred (on a magnetic stirrer) in the dark for half an hour, in order to achieve adsorption–desorption equilibrium between the methyl orange molecules and catalyst samples.
At definite time intervals, 3 mL of solution was removed, and the catalyst was separated by centrifugation. The concentration of the dye solution was quantitatively calculated by monitoring the absorbance of MO at 464 nm through the UV–Vis spectrophotometer (Shimadzu, Model-UV-3600, Kyoto, Japan). The recyclability study was also carried out using the used catalyst by maintaining the same reaction conditions.

3. Results and Discussion

3.1. XRD Analysis

The crystalline structures of as-prepared materials were studied using XRD analysis. Figure 1 exhibits the XRD pattern of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70. As shown in Figure 1, the characteristic peaks of TU, observed at 2θ = 25.3°, 37.8°, 48.0°, 53.9°, 55.1°, 62.7°, 68.9°, 70.3° and 75.1°, correspond to the reflection planes of (101), (004), (200), (105), (211), (204), (116), (220) and (215), respectively, which indicates the formation of TiO2 in the anatase phase and clearly matches with the reported pattern (JCPDS 21-1272) [25]. For the ZU sample shown in Figure 1, the characteristic peaks of ZnO appear at 31.63°, 34.29°, 36.32°, 47.54°, 56.72°, 62.83° and 67.93° and can be indexed to the planes of (100), (002), (101), (102), (110), (103) and (112), which indicates the existence of the hexagonal wurtzite structure of ZnO (JCPDS 36-1451) [26]. For the TZU95 and TZU90 nanocomposite samples, no characteristic diffraction peaks of ZnO are found; this may be because the lower weight ratio of the ZnO precursor is below the detection limit of XRD. On the other hand, the TZU85, TZU80 and TZU75 nanocomposite samples show the slight appearance of the characteristic peaks of wurtzite ZnO, ranging from 30° to 38°, which correspond to the reflection planes of (100), (002) and (101). XRD confirmed that the thermal decomposition of titanium oxysulphate, zinc acetate and urea is a very proficient method for the successful formation of the TiO2-ZnO nanocomposite. In addition, no impurity peaks are observed in the TZU95, TZU90, TZU85, TZU80 and TZU75 samples, indicating that the coupling of the TiO2 anatase phase and the ZnO hexagonal wurtzite forms the TiO2-ZnO nanocomposite. Furthermore, coupling of ZnO with TiO2 in the TZU95, TZU90, TZU85, TZU80 and TZU75 samples still maintains the anatase phase of TiO2. However, it should be noted that the TZU70 sample shows weak peaks at 28.8°, suggesting the formation of a new Zn2TiO4 phase [27].
The crystallite sizes (D) of the TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70 samples were calculated using Scherrer’s equation [28].
D = 0.9 λ β cos θ
where ‘λ’ indicates the wavelength of the X-ray employed, ‘θ’ stands for the angle of diffraction for a more intense diffraction peak and ‘β’ stands for the full width at half maximum of the most intense diffraction peak (FWHM). The diffraction peak (101) was used to calculate the crystallite sizes of the samples. The crystallite sizes of the TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70 samples are about 14, 13.9, 12.8, 15.2, 14.8, 15.3, 13.3 and 15 nm, respectively.

3.2. Structure and Morphological Analysis

The morphology and structure of the TiO2-ZnO nanocomposite were studied using SEM, EDX, HRTEM and STEM. Figure 2a–d portray the SEM images of the TZU75 nanocomposite with various magnifications. It can be clearly seen that the TZU75 nanocomposite is composed of porous and aggregated nanoparticles (NPs) of TiO2 and ZnO. In addition, spherical-shaped NPs are interconnected which may lead to interfacial interaction and to the formation of a heterojunction between TiO2 and ZnO NPs.
Figure 2a–d and Figure 3a–d show the SEM and TEM images of TZU75, which reveal no distinct boundary between the TiO2 and ZnO phases; this might be because of the similar structure and morphology of the TiO2 and ZnO NPs. For further proof of the existence of the nanocomposite, TEM was used to examine the morphology of the samples. Figure 3a–d show that TiO2 NPs with ZnO are found in the form of interconnected, aggregated, spherical, slightly hierarchical and porous NPs with a higher surface area. The size of the NPs is well within the range of 10 to 30 nm and, notably, matches with the XRD results. HR-TEM, STEM and elemental mapping of the TZU75 nanocomposite were carried out to check the presence of Zn and Ti. Lattice fringes of 0.356 and 0.262 nm, measured in the HRTEM image of the TZU75 sample, correspond to the interplanar distances of the (101) and (002) crystallographic planes of TiO2 and ZnO, respectively [29]. Furthermore, the elemental mapping (Figure 3e–h) demonstrated the uniform distribution of Ti, O and Zn atoms throughout the nanocomposite. Interestingly, the elemental mapping showed that the ZnO NPs consistently spread over the entire surface of the TiO2 nanostructure, signifying the formation of the nano-heterostructure.

3.3. Optical Properties

The activity of the photocatalyst is highly dependent on the absorption of light together with its quantity and wavelength. To investigate the optical properties, we used UV–Vis diffuse reflectance spectra for the TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70 samples. As shown in Figure 4, TU and ZU exhibit pronounced absorption in the UV region. Remarkably, the addition of ZnO to TiO2, the TZU75 nanocomposite exhibited increased absorption in both the UV and visible regions, which is ascribed to the transfer of electrons from the valence band (VB) to the conduction band (CB) [30]. The successful formation of TiO2-ZnO heterojunctions reduces the recombination rate of the photoinduced electron–hole pair and enhances the photocatalytic activity of the TiO2-ZnO nanocomposite.

3.4. PL Spectra

Photoluminescence (PL) spectra were implemented to study the recombination rate of the photoinduced electron–hole pair and the scope of charge separation. The photoluminescence spectra of the photocatalysts had an excitation wavelength of 295 nm for TiO2 and the TiO2-ZnO nanocomposite, at room temperature, as depicted in Figure 5. The PL spectra depict two broad emission peaks centered at 410 and 470 nm, which correspond to the high rate of electron–hole pair recombination or band edge emission, and to oxygen vacancies, respectively. However, in the TiO2-ZnO nanocomposite, the PL intensity of the TZU75 nanocomposite was considerably reduced with the introduction of ZnO and the formation of the TiO2-ZnO heterojunction which has interfacial charge transport from TiO2 to ZnO, or vice versa.
Due to this, the electron–hole pair recombination rate extensively decreased, and the photocatalytic activity of the TiO2-ZnO nanocomposite increased.

3.5. Photocatalytic Degradation of Methyl Orange Using UV–Visible Light

The photocatalytic performances of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70 were independently evaluated for the degradation of aqueous methyl orange as a pollutant in the presence of a UV–visible light source. The quantity of the aqueous MO solution was calculated from the absorbance (λmax = 464 nm), using UV–Vis spectra. The percent photodegradation (D%) of MO under UV light irradiation over the photocatalyst was estimated using the following equation:
D % = ( C 0 C ) C 0 × 100 %
where C and C0 are the final and initial concentrations of MO, respectively. The photocatalysis by the as-prepared photocatalysts for MO degradation under UV light illumination is exhibited in Figure 6a–b. The photocatalytic activity of MO is 79.0, 70, 8, 87, 90, 94, 99.1 and 88% for TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70, after being irradiated with UV light for 60 min. It was found that TZU75 showed a much better performance than its counter parts TU (79.03%) and ZU (70%) within 60 min, which is ascribed to the constructive heterojunction and interfacial contact between TiO2 and ZnO. It should be noted that the TU and ZU photocatalysts exhibit lower photocatalytic efficiency due to the higher recombination rate of the photogenerated electron–hole pair. Moreover, the photodegradation performance increased with the increasing content of ZnO, due to the formation of the effective heterojunction and the inhibition of electron–hole pair recombination [31]. However, when the ZnO content was exceeded in TZU70, there was a saturation value and the overall heterostructure was unstable, leading to a decrease in the effective charge transfer process in the heterostructure and a decrease in photocatalytic efficiency [29].
The most noteworthy principles of a perfect photocatalyst in a real-world application are its photostability and recyclability. To verify the photostability of the TZU75 nanocomposite, a number of reaction cycles were performed for the degradation of MO under the same reaction conditions. As shown in Figure 7, after five reaction cycles, the TZU75 nanocomposite did not display any noteworthy loss of activity. It was observed that the MO photocatalytic activity dropped to 6% after five consecutive reaction cycles. This result indicates that the TZU75 nanocomposite shows good photostability and recyclability.
On the basis of the above result, the comprehensive plausible mechanism for the photodegradation of MO dye over the TiO2-ZnO nanocomposite is delineated in Scheme 1. As per the process of the photodegradation of the dye, the photocatalytic activity of the TiO2-ZnO nanocomposite correlated to the formation of a suitable type II heterojunction between the interface of ZnO and TiO2, with band gaps of 3.37 eV and 3.2 eV, respectively. Under UV–visible light illumination, both ZnO and TiO2 were excited and produced electrons (e) in the CB and holes (h+) in the VB. More significantly, due to the establishment of the heterojunction between ZnO and TiO2, the CB level of ZnO is lower than the CB level of TiO2. Photogenerated e− in the CB of TiO2 have affinity to transfer into the CB of ZnO through the interface; conversely, photogenerated h+ in the VB of TiO2 could travel into the VB of ZnO [32]. This route could efficiently enhance the lifespan and separation ability of the photoinduced electron–hole pair. Consequently, the photoinduced and separated e in the VB of TiO2 capture the surface-adsorbed molecular oxygen to form superoxide radicals (•O2), whereas the h+ migrate into the VB of ZnO to oxidize H2O and to generate hydroxyl radicals (•OH). Finally, both •O2 and •OH radicals act as strong oxidizing agents to oxidize MO into carbon dioxide and water molecules.

4. Conclusions

In summary, we rationally developed a simple, inexpensive and novel technique for the synthesis of TiO2-ZnO nanocomposites, using a thermal decomposition method. The as-synthesized TiO2-ZnO nanocomposite was employed for the photodegradation of methyl orange (MO) in the presence of UV–visible light illumination. The degradation study revealed that the TiO2-ZnO nanocomposite showed superior photocatalytic activity to pure TiO2 and ZnO. More importantly, the photocatalytic activity of the TiO2-ZnO nanocomposite did not exhibit any noticeable decrease and remained stable even after five recycles. The remarkable photocatalytic activity can be ascribed to enhanced charge separation at the interface and to the formation of a heterojunction between TiO2 and ZnO. It is hoped that our unique approach for the synthesis of mixed metal oxide heterostructure nanocomposites may open up a new opportunity for wastewater treatment.

Author Contributions

Writing—original draft and methodology, S.M.D.; writing—review and editing, and validation, S.S.A.; data curation and formal analysis, S.B.B.; funding acquisition and visualization, S.F.S.; project administration and data interpretation, A.M.T.; validation and resources, N.T.N.T.; formal analysis and data collection, C.-D.K.; supervision and investigation, S.M.K.; conceptualization and writing—review and editing, M.S.T.; supervision and project administration, S.R.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was funded by the Researchers Supporting Project (number RSP-2021/370), King Saud University, Riyadh, Saudi Arabia.

Institutional Review Board Statement

Not Applicable.

Informed Consent Statement

Not Applicable.

Data Availability Statement

Not Applicable.

Acknowledgments

The authors are thankful to the Researchers Supporting Project (number RSP-2021/370), King Saud University, Riyadh, Saudi Arabia.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could appear to influence the work reported in this paper.

References

  1. Li, Q.; Guo, B.; Yu, J.; Ran, J.; Zhang, B.; Yan, H.; Gong, J.R. Highly efficient visible-light-driven photocatalytic hydrogen production of CdS-cluster-decorated graphene nanosheets. J. Am. Chem. Soc. 2011, 133, 10878–10884. [Google Scholar] [CrossRef]
  2. Zhou, M.; Lou, X.W.; Xie, Y. Two-dimensional nanosheets for photoelectrochemical water splitting: Possibilities and opportunities. Sci. Direct 2013, 8, 598–618. [Google Scholar] [CrossRef]
  3. Zhang, A.Y.; Long, L.L.; Liu, C.; Li, W.W.; Yu, H.Q. Chemical recycling of the waste anodic electrolyte from the TiO2 nanotube preparation process to synthesize facet-controlled TiO2 single crystals as an efficient photocatalyst. Green Chem. 2014, 16, 2745–2753. [Google Scholar] [CrossRef]
  4. Shaikh, A.F.; Arbuj, S.S.; Tamboli, M.S.; Naik, S.D.; Rane, S.B.; Kale, B.B. ZnSe/ZnO Nano-Heterostructures for Enhanced Solar Light Hydrogen Generation. Chem. Sel. 2017, 2, 9174–9180. [Google Scholar] [CrossRef]
  5. Zhang, L.; Wang, W.; Yang, J.; Chen, Z.; Zhan, W.; Zhou, L.; Liu, S. Sonochemical synthesis of nanocrystallite Bi2O3 as a visible-light-driven photocatalyst. Appl. Catal. A 2006, 308, 105–110. [Google Scholar] [CrossRef]
  6. Li, L.; Chu, Y.; Liu, Y.; Dong, L. Template-Free Synthesis and Photocatalytic Properties of Novel Fe2O3 Hollow Spheres. J. Phys. Chem. C 2007, 111, 2123–2127. [Google Scholar] [CrossRef]
  7. Deshpande, S.P.; Tamboli, M.S.; Arbuj, S.S.; Mulik, U.P.; Amalnerkar, D.P. Synthesis of Nanostructured Ta2O5 and Its Photocatalytic Performance Study. J. Nanoeng. Nanomanuf. 2014, 4, 215–220. [Google Scholar] [CrossRef]
  8. Smith, Y.R.; Kar, A.; Subramanian, V.R. Investigation of physicochemical parameters that influence photocatalytic degradation of methyl orange over TiO2 nanotubes. Ind. Eng. Chem. Res. 2009, 48, 10268–10276. [Google Scholar] [CrossRef]
  9. Jin, Z.; Duan, W.; Liu, B.; Chen, X.; Yang, F. Fabrication of efficient visible light activated Cu–P25–graphene ternary composite for photocatalytic degradation of methyl blue. J. Appl. Surf. Sci. 2015, 356, 707–718. [Google Scholar] [CrossRef]
  10. Margan, P.; Haghighi, M. Hydrothermal-assisted sol–gel synthesis of Cd-doped TiO2 nanophotocatalyst for removal of acid orange from wastewater. J. Sol-Gel Sci. Technol. 2017, 81, 556–718. [Google Scholar] [CrossRef]
  11. Wang, P.; Yi, X.; Lu, Y.; Yu, H.; Yu, J. In-situ synthesis of amorphous H2TiO3-modified TiO2 and its improved photocatalytic H2-evolution performance. J. Colloid Interface Sci. 2018, 532, 272–279. [Google Scholar] [CrossRef]
  12. Wang, Y.; Yu, J.G.; Xiao, W.; Li, Q. Microwave-assisted hydrothermal synthesis of graphene based Au–TiO2 photocatalysts for efficient visible-light hydrogen production. J. Mater. Chem. A 2014, 2, 3847–3855. [Google Scholar] [CrossRef]
  13. Dong, Z.; Wu, M.H.; Wu, J.Y.; Ma, Y.Y.; Ma, Z.Z. In situ synthesis of TiO2/SnOx–Au ternary heterostructures effectively promoting visible-light photocatalysis. Dalton Trans. 2015, 44, 11901–11910. [Google Scholar] [CrossRef]
  14. Muzakki, A.; Shabrany, H.; Saleh, R. Synthesis of ZnO/CuO and TiO2/CuO nanocomposites for light and ultrasound assisted degradation of a textile dye in aqueous solution. AIP Conf. Proc. 2016, 1725, 020051. [Google Scholar] [CrossRef]
  15. Vinoth, R.; Karthik, P.; Devan, K.; Neppolian, B.; Ashokkumar, M. TiO2–NiO p–n nanocomposite with enhanced sonophotocatalytic activity under diffused sunlight. Ultrason Sonochem. 2017, 35, 655–663. [Google Scholar] [CrossRef] [PubMed]
  16. Lachom, V.; Poolcharuansin, P.; Laokul, P. Preparation, characterizations and photocatalytic activity of a ZnO/TiO2 nanocomposite. Mater. Res. Express. 2017, 4, 035006. [Google Scholar] [CrossRef]
  17. Arbuj, S.S.; Mulik, U.P.; Amalnerkar, D.P. Synthesis of Ta2O5/TiO2 Coupled Semiconductor Oxide Nanocomposites with High Photocatalytic Activity. Nanosci. Nanotechnol. Lett. 2013, 5, 968–973. [Google Scholar] [CrossRef]
  18. Arbuj, S.S.; Hawaldar, R.R.; Mulik, U.P.; Amalnerkar, D.P. Preparation, Characterisation and Photocatalytic Activity of Nb2O5/TiO2 Coupled Semiconductor Oxides. J. Nanoeng. Nanomanuf. 2013, 3, 79–83. [Google Scholar] [CrossRef]
  19. Sun, C.; Xu, Q.; Xie, Y.; Ling, Y.; Hou, Y. Designed synthesis of anatase–TiO2 (B) biphase nanowire/ZnO nanoparticle heterojunction for enhanced photocatalysis. J. Mater. Chem. A 2018, 6, 8289–8298. [Google Scholar] [CrossRef]
  20. Wang, J.; Wang, G.; Wei, X.; Liu, G.; Li, J. ZnO nanoparticles implanted in TiO2 macrochannels as an effective direct Z-scheme heterojunction photocatalyst for degradation of RhB. Appl. Surf. Sci. 2018, 456, 666–675. [Google Scholar] [CrossRef]
  21. Liao, M.-H.; Hsu, C.-H.; Chen, D.-H. Preparation and properties of amorphous titania-coated zinc oxide nanoparticles. J. Solid State Chem. 2006, 179, 2020–2026. [Google Scholar] [CrossRef]
  22. Chakma, S.; Moholkar, V.S. Synthesis of bi-metallic oxides nanotubes for fast removal of dye using adsorption and sonocatalysis process. J. Ind. Eng. Chem. 2016, 37, 84–89. [Google Scholar] [CrossRef]
  23. King, D.M.; Liang, X.; Zhou, Y.; Carney, C.S.; Hakim, L.F.; Li, P.; Weimer, A.W. Atomic layer deposition of TiO2 films on particles in a fluidized bed reactor. Powder Technol. 2008, 183, 356–363. [Google Scholar] [CrossRef]
  24. Bahadur, N.M.; Furusawa, T.; Sato, M.; Kurayama, F.; Suzuki, N. Rapid synthesis, characterization and optical properties of TiO2 coated ZnO nanocomposite particles by a novel microwave irradiation method. Mater. Res. Bull. 2010, 45, 1383–1388. [Google Scholar] [CrossRef]
  25. Deshmukh, S.M.; Tamboli, M.S.; Shaikh, H.; Babar, S.B.; Hiwarale, D.P.; Thate, A.G.; Shaikh, A.F.; Alam, M.A.; Khetre, S.M.; Bamane, S.R. A Facile Urea-Assisted Thermal Decomposition Process of TiO2 Nanoparticles and Their Photocatalytic Activity. Coatings 2021, 11, 165. [Google Scholar] [CrossRef]
  26. He, Y.; Wang, Y.; Zhang, L.; Teng, B.; Fan, M. High-efficiency conversion of CO2 to fuel over ZnO/g-C3N4 photocatalyst. Appl. Catal. B 2015, 168, 1–8. [Google Scholar] [CrossRef]
  27. Houskova, V.; Stengl, V.; Bakardjieva, S.; Murafa, N. Photoactive materials prepared by homogeneous hydrolysis with thioacetamide: Part 2—TiO2/ZnO nanocomposites. J. Phys. Chem. Solids 2008, 69, 1623–1631. [Google Scholar] [CrossRef] [Green Version]
  28. Cullity, B.D.; Stock, S.R. Elements of X-ray Diffraction Reading; Addison-Wesley: London, UK, 1978; Available online: https://www.worldcat.org/title/elements-of-x-ray-diffraction/oclc/3672627 (accessed on 10 September 2021).
  29. Xiao, F.-X. Construction of Highly Ordered ZnO–TiO2 Nanotube Arrays (ZnO/TNTs) Heterostructure for Photocatalytic Application. ACS Appl. Mater. Interfaces 2012, 4, 7055–7063. [Google Scholar] [CrossRef] [PubMed]
  30. Mane, R.S.; Lee, W.J.; Pathan, H.M.; Han, S.H. Nanocrystalline TiO2/ZnO Thin Films:  Fabrication and Application to Dye-Sensitized Solar Cells. J. Phys. Chem. B 2005, 109, 24254–24259. [Google Scholar] [CrossRef]
  31. Hashemi, M.M.; Nikfarjam, A.; Hajghassem, H.; Salehifar, N. Hierarchical Dense Array of ZnO Nanowires Spatially Grown on ZnO/TiO2 Nanofibers and Their Ultraviolet Activated Gas Sensing Properties. J. Phys. Chem. C 2020, 124, 322–335. [Google Scholar] [CrossRef]
  32. Zha, R.; Nadimicherla, R.; Guo, X. Ultraviolet photocatalytic degradation of methyl orange by nanostructured TiO2/ZnO heterojunctions. J. Mater. Chem. A 2015, 3, 6565. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Figure 1. XRD patterns of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Metals 11 01787 g001
Figure 2. SEM images of (ad) TZU75 nanocomposite.
Figure 2. SEM images of (ad) TZU75 nanocomposite.
Metals 11 01787 g002
Figure 3. (a,b) TEM, (c,d) HRTEM and (e) STEM images and the corresponding (fh) elemental mapping of the TZU75 nanocomposite.
Figure 3. (a,b) TEM, (c,d) HRTEM and (e) STEM images and the corresponding (fh) elemental mapping of the TZU75 nanocomposite.
Metals 11 01787 g003
Figure 4. UV–Vis diffuse reflectance spectra of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Figure 4. UV–Vis diffuse reflectance spectra of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Metals 11 01787 g004
Figure 5. Photoluminescence (PL) spectra of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Figure 5. Photoluminescence (PL) spectra of TU, ZU, TZU95, TZU90, TZU85, TZU80, TZU75 and TZU70.
Metals 11 01787 g005
Figure 6. Photocatalytic activity of MO solution over TZU75 under UV light irradiation: (a) absorbance against wavelength; (b) degradation against irradiation time.
Figure 6. Photocatalytic activity of MO solution over TZU75 under UV light irradiation: (a) absorbance against wavelength; (b) degradation against irradiation time.
Metals 11 01787 g006
Figure 7. Reusability of TZU75 for MO degradation under UV light irradiation.
Figure 7. Reusability of TZU75 for MO degradation under UV light irradiation.
Metals 11 01787 g007
Scheme 1. Plausible mechanism for the photocatalytic degradation of MO over TZU75 under UV light irradiation.
Scheme 1. Plausible mechanism for the photocatalytic degradation of MO over TZU75 under UV light irradiation.
Metals 11 01787 sch001
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Deshmukh, S.M.; Arbuj, S.S.; Babar, S.B.; Shaikh, S.F.; Tamboli, A.M.; Nguyen Truong, N.T.; Kim, C.-D.; Khetre, S.M.; Tamboli, M.S.; Bamane, S.R. Environmentally Benign Organic Dye Conversion under UV Light through TiO2-ZnO Nanocomposite. Metals 2021, 11, 1787. https://doi.org/10.3390/met11111787

AMA Style

Deshmukh SM, Arbuj SS, Babar SB, Shaikh SF, Tamboli AM, Nguyen Truong NT, Kim C-D, Khetre SM, Tamboli MS, Bamane SR. Environmentally Benign Organic Dye Conversion under UV Light through TiO2-ZnO Nanocomposite. Metals. 2021; 11(11):1787. https://doi.org/10.3390/met11111787

Chicago/Turabian Style

Deshmukh, Sandip M., Sudhir S. Arbuj, Santosh B. Babar, Shoyebmohamad F. Shaikh, Asiya M. Tamboli, Nguyen Tam Nguyen Truong, Chang-Duk Kim, Sanjay M. Khetre, Mohaseen S. Tamboli, and Sambhaji R. Bamane. 2021. "Environmentally Benign Organic Dye Conversion under UV Light through TiO2-ZnO Nanocomposite" Metals 11, no. 11: 1787. https://doi.org/10.3390/met11111787

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