Next Article in Journal
Mayenite-Based Electride C12A7e: A Reactivity and Stability Study
Next Article in Special Issue
Photocatalytic Oxidation of Chlorantraniliprole, Imidacloprid, Pirimicarb, Thiamethoxam and Their Main Photoreaction InterMediates as Impacted by Water Matrix Composition under UVA-LED Exposure
Previous Article in Journal
One-Pot Synthesis of Ultra-Small Pt Dispersed on Hierarchical Zeolite Nanosheet Surfaces for Mild Hydrodeoxygenation of 4-Propylphenol
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

rGO Functionalized ZnO–TiO2 Core-Shell Flower-Like Architectures for Visible Light Photocatalysis

by
Evangelia Vasilaki
1,2,3,*,
Nikos Katsarakis
1,3,
Spyros Dokianakis
1 and
Maria Vamvakaki
2,3
1
Center of Materials Technology and Photonics, Hellenic Mediterranean University, 714 10 Heraklion, Crete, Greece
2
Department of Materials Science and Technology, University of Crete, 700 13 Heraklion, Crete, Greece
3
Institute of Electronic Structure and Laser, Foundation for Research & Technology-Hellas, 700 13 Heraklion, Crete, Greece
*
Author to whom correspondence should be addressed.
Submission received: 28 January 2021 / Revised: 26 February 2021 / Accepted: 1 March 2021 / Published: 5 March 2021
(This article belongs to the Special Issue Photocatalytic Activity of TiO2 and Its Applications)

Abstract

:
Core-shell heterostructures with a complex, flower-like morphology, comprising a ZnO core and a TiO2 shell decorated with reduced graphene oxide (rGO) sheets by hydrothermal wrapping, are reported to extend the absorption properties of the semiconductors toward the visible light range. The ternary photocatalysts were characterized by X-ray diffraction, field emission scanning electron microscopy, Raman spectroscopy, diffuse reflectance UV–Vis, and attenuated total reflectance-Fourier transform infrared spectroscopy. Its photocatalytic performance was evaluated under visible light irradiation using methylene blue dye as a model pollutant. The rGO-modified ZnO–TiO2 photocatalyst exhibited superior photoactivity compared to that of the parent ZnO–TiO2 core-shell structures, which was dependent on its graphene content. The enhanced photocatalytic response was attributed to the higher absorption in the visible light range, as well as the pronounced electron and hole separation in the ternary system.

Graphical Abstract

1. Introduction

The intense industrialization of the last decades has resulted in the continuous introduction of industrial and agro-industrial wastewaters in aqueous matrices; heterogeneous photocatalysis has been proposed as an efficient alternative to conventional methods for their effective treatment. TiO2 has been the most widely used photocatalyst, but the maximization of its photocatalytic response is hindered by its inadequate absorption in the visible light range, as well as the high degree of recombination of the photogenerated charge carrier pairs. The synthesis of core-shell heterostructures comprising TiO2 and ZnO has been evaluated as an effective approach to diminish charge recombination due to the favorable energy differences between the two semiconductors [1,2,3,4,5]. The modification of TiO2 or ZnO with reduced graphene oxide (rGO) sheets has been shown to present synergistic advantages such as a suppressed recombination of the photogenerated charge carriers and the extension of the photoresponse toward the visible light range [6,7,8,9,10,11].
The preparation of ternary photocatalysts, comprising TiO2, ZnO and rGO, has been scarcely studied. Previous reports on ternary TiO2, ZnO, and rGO photocatalysts focused on the nanocomposites of mixed inorganic particles prepared by conventional sol-gel, hydrothermal methods, or other modern synthetic approaches such as sonochemistry or microwave treatment [12,13,14,15,16,17,18,19,20]. In an important study, Johra et al. synthesized a mixed photocatalyst comprising TiO2 and ZnO particles that were surface-modified with graphene oxide (GO) via a hydrothermal reaction [12]. The mixed photocatalyst exhibited increased photocatalytic performance in the UV degradation of Cr(VI) compared with the parent TiO2–ZnO analogues that were assigned to enhance light absorption and better the separation of electron-hole pairs. Potle et al. also compared the preparation of a mixed photocatalyst, based on TiO2 and ZnO modified by rGO, via conventional stirring and a sonochemical approach. The application of ultrasound suppressed the aggregation of the inorganic nanoparticles and resulted in a better dispersion on the rGO sheets, which was reflected by more effective photocatalytic activity [13]. Simsek et al. immobilized rGO sheets on fluorine-doped tin oxide (FTO) substrates, followed by the hydrothermal growth of ZnO and TiO2 nanoparticles on graphene sheets. They observed a band-gap narrowing for the ternary photocatalyst and superior photocatalytic performance against estrogen bisphenol-A and pharmaceuticals under UV and visible-light irradiation.
Architectures of a higher complexity have not been examined, yet it is known that the morphology of the ZnO–TiO2 heterostructures heavily affects significant properties of the composite material such as the surface area and the contact between the two interfaces; thus, its control is important [21]. The contact between the rGO sheets and the inorganic component has been reported to significantly influence the photocatalytic performance of the composite photocatalyst; however, the degree of contact in the ternary photocatalysts presented so far is rather poor [22,23,24,25]. Therefore, the tailored synthesis of rGO functionalized ZnO–TiO2 core-shell heterostructures with a complex and well-defined morphology, as well as a visible light photocatalytic activity, is expected to attract great scientific interest and will find wide applicability in wastewater treatment.
In the present work, we report the development of complex ZnO–TiO2 core-shell, flower-like heterostructures and the extension of its photocatalytic response toward the visible light range via wrapping the inorganic surface with rGO sheets. First, well-defined core-shell, flower-like architectures, with a wurtzite ZnO core and an anatase TiO2 shell, were synthesized by controlled sol-gel reactions. Next, the ZnO–TiO2 heterostructures were surface-functionalized with 3-aminopropyltrimethoxysilane (APTMS) and hydrothermally wrapped with rGO sheets at tunable mass ratios, resulting in a strong rGO–inorganic surface contact. The photocatalytic performance of the obtained rGO–ZnO–TiO2 ternary photocatalysts was evaluated in the decoloration of aqueous solutions of methylene blue (MB) under visible light irradiation.

2. Results and Discussion

2.1. Graphene Oxide

The GO was synthesized by a modified Hummers’ method using graphite powder [26] and was characterized by XRD and Raman spectroscopy, and FTIR (Figure 1). A single peak at ~10° can be observed in the XRD pattern of the GO (Figure 1a), which can be assigned to the (001) crystalline plane of the GO, verifying the successful oxidation of graphite [27,28,29].
The Raman spectrum of the GO (Figure 1b) exhibited peaks at ~1355 and 1600 cm−1, assigned to the typical D and G bands of graphitic materials, respectively [30,31]. Specifically, the D band can be attributed to the modes of the sp2 atoms in the rings, while the G band is due to the stretching vibration of the sp2-hybridized C–C bonds [32].
The ATR-FTIR spectrum of the GO presented in Figure 1c shows typical peaks associated with the oxygen functionalities introduced onto the rGO sheets upon the oxidation of graphite. In particular, the broad band at ~3400 cm−1 can be assigned to O–H stretching vibration, while the peaks at ~1720 and 1040 cm−1 are ascribed to C=O and C–O stretching vibrations, respectively. Moreover, the peak at 1615 cm−1 is derived from the vibrations of the remaining unoxidized graphitic domains [33,34].

2.2. ZnO Flower-Like Structures and ZnO–TiO2 Core-Shell Heterostructures

The SEM and TEM images of the ZnO and the ZnO–TiO2 core-shell architectures are shown in Figure 2. Well-defined flower-like ZnO structures with a smooth surface (Figure 2a,c and Figure S1a) were observed. After coating with a TiO2 shell, non-agglomerated core-shell structures were obtained with a uniform porous titania shell of 59 ± 7 nm thickness (Figure 2b,d and Figure S1b). Based on the findings of our previous work, where the photoactivity of ZnO–TiO2 core-shell, flower-like architectures with variable titania shell thickness were evaluated, a shell thickness of ~59 nm was found to exhibit the highest photocatalytic performance [2].

2.3. rGO-Modified ZnO–TiO2 Core-Shell Heterostructures

To investigate the effect of APTMS pre-treatment on the morphology of the ternary photocatalysts, the synthesis of rGO-modified ZnO–TiO2 structures (denoted as ZT-rGO) was first carried out using ZnO–TiO2 heterostructures with and without amine functionalization. As shown in Figure 3, when the surface of the ZnO–TiO2 flower-like structures were not amine-functionalized before the introduction of the GO at a 2 wt % content, the rGO sheets did not wrap the inorganic material well (indicated by the white arrow in Figure 3a). On the contrary, the amine pre-functionalized flower-like structures were completely wrapped with the rGO sheets, which attached very well onto the inorganic surface (Figure 3b).
The observed increase in the adhesion of the rGO sheets onto the surface of the ZnO–TiO2 heterostructures can be assigned to the favorable electrostatic interactions between the amine-modified inorganic surface and the rGO sheets. The APTMS functionalization of the ZnO–TiO2 core-shell, flower-like heterostructures resulted in the formation of a positively charged inorganic surface onto which the negatively charged rGO sheets could readily attach [34,35,36,37,38]. The high contact degree between the rGO layers and the semiconductor resulted in an effective charge separation during the photocatalytic process, which is highly desirable [22,23,24,25]. Therefore, the APTMS functionalized ZnO–TiO2 core-shell, flower-like structures were employed in all the experiments discussed below.
The rGO sheets were attached onto the APTMS-modified ZnO–TiO2 core-shell, flower-like structures at different GO loadings. The obtained ternary catalyst structures are shown in Figure 4. The attachment of the rGO sheets on the flower-like structures was heavily affected by the rGO content of the composite. The rGO sheets completely wrapped the inorganic structures at 0.5 and 2 wt % graphitic loading (Figure 4a,b and Figure S1c); above this plateau value (Figure 4c), unattached rGO sheets (indicated by white arrows) were also observed.
The rGO-modified core-shell semiconductors were characterized by ATR-FTIR (Figure 4d). The spectra of the ZT-rGO structures exhibited new peaks at ~1560 and 1230 cm−1, which were attributed to the C=C skeletal vibration mode and C–O–C stretching vibration of rGO, respectively [39,40]. The peak at 1400 cm−1 can be attributed to the C–OH vibration mode of the graphitic material. As expected, the intensity of the peaks decreased with the reduction in the rGO loading.
Raman spectroscopy was also employed to verify the wrapping of the rGO sheets on the surface of the ZnO–TiO2 core-shell, flower-like structures (Figure 5). The peaks at 142, 391, 517, and 630 cm−1 observed for TiO2 can be attributed to the Eg, B1g, A1g + B1g, and Eg modes of the anatase phase, respectively [41]. Two peaks of low intensity can be observed in the Raman spectrum of the ZnO flowers at 435 and 1145 cm−1, corresponding to the E2high and 2LO modes of wurtzite ZnO [42,43,44]. The peaks were also detected in the spectrum of the ZnO–TiO2 core-shell, flower-like structures (Figure 5a), whereas the ZT-rGO ternary material exhibited two additional Raman lines at 1355 and 1590 cm−1, attributed to the D and G peaks of rGO, respectively.
The band gap energy values (Eg) of the synthesized ZT-rGO materials were quantified by UV–Vis diffuse reflectance measurements, employing the Tauc equation:
α h ν = A ( h ν Eg ) 1 / n
where hν is the photon energy, A is an energy-independent constant, and n = 2, as the indirect band gap, TiO2 is the outer material in the core-shell heterostructures [1].
The normalized and transformed Kubelka–Munk plots as a function of light energy are shown in Figure 6, while the calculated Eg values are presented in Table 1. A narrowing in the band-gap values and a redshift in the absorbance edge were observed in the ZT-rGO structures compared with their parent ZnO–TiO2 analogue, in agreement with the literature [45,46,47]. An Eg of 3.08 eV was calculated for the ZnO–TiO2 heterostructures, lower than the values reported for the bare ZnO and TiO2 (~3.2–3.3 eV) nanoparticles [2]. This phenomenon can be assigned to the generation of new energy states at the ZnO and TiO2 heterojunction [48]. The Eg values of the ternary photocatalysts decreased monotonically with the increase in the rGO content due to the generation of Ti–O–C bonds during the thermal treatment, as we reported in our earlier work [6].
Next, the photoactivity of the parent core-shell and the ZT-rGO structures was investigated in the decoloration of aqueous MB solutions under visible-light irradiation (Figure 7).
The kinetic rate constants, k, for the decoloration process of MB were calculated from the first-order equation:
ln C Co = kt
The regression coefficient of the linear fits, R2, was greater than 0.99 for all studied samples.
The photocatalytic activity of the rGO-modified ZnO–TiO2 core-shell, flower-like photocatalysts was, for all studied samples, superior compared to that of the parent ZnO–TiO2 core-shell semiconductor. A graphitic-content-dependent performance was observed, which was optimum for the ZT-rGO sample with 2 wt % rGO loading, and was reduced upon further increase in the rGO loading to 5 wt %.
The parent ZnO–TiO2 photocatalyst exhibited a ~51% dye decoloration (rate constant k = 0.0056 min−1) after a 120 min irradiation time. The decrease in the dye concentration for the ZT-rGO sample with 0.5 wt % rGO was 76% (k = 0.011 min−1), while the ZT-rGO photocatalyst with 2 wt % rGO exhibited the highest photocatalytic performance with an 86% dye removal (k = 0.015 min−1). Lastly, upon increasing the rGO loading to 5 wt % a decrease in the performance of the ZT-rGO photocatalyst to a 60% dye decoloration, with a rate constant k = 0.007 min−1, was found.
The reusability of the ZT-rGO catalyst with 2 wt % rGO, which exhibited the highest photocatalytic activity, was evaluated for three photocatalytic cycles. As shown in Figure 8, the photocatalyst exhibited excellent reusability, with its performance decreasing only by ~8% after three photocatalytic cycles. This is indirect evidence that the intimate contact between the inorganic component and the rGO sheets is maintained during photocatalysis. Arguably, the rGO content would decrease since each photocatalytic cycle is followed by centrifugation and washing with water; thus, a rapid decrease in the photocatalytic performance after each cycle would be observed.
As presented above, the unmodified ZnO–TiO2 core-shell photocatalyst exhibited a low, nonnegligible photoactivity in the visible light range. As mentioned in the experimental section, the lamp used in all photocatalytic experiments exhibited emission lines between 390 and 600 nm. The observed photoactivity of the ZnO–TiO2 core-shell photocatalyst can be attributed to its band-gap value, which was measured as 3.08 eV (Table 1), corresponding to ~400 nm. Therefore, the ZnO–TiO2 core-shell photocatalyst could be effectively activated in the 390–400 nm emission range of the lamp. The photocatalytic decoloration of MB by the ZT-rGO photocatalysts was significantly promoted compared to that of the parent core-shell photocatalyst. The observed superior performance of the ZT-rGO photocatalysts can be assigned to the rGO sheets acting as sensitizers in the visible light, inducing band-gap narrowing and a shift in the absorption edge of the ZT-rGO materials toward the visible light range [49,50].
The superior photocatalytic performance of the ZT-rGO core-shell architectures can also be attributed to the appropriate energy level differences between ZnO, TiO2, and rGO. These differences result in a more efficient separation and migration of the photogenerated electrons and holes, significantly suppressing its recombination and promoting the photocatalytic rates. According to the literature, the conduction and valence bands of ZnO are more negatively charged (~0.48 eV) and are located above those of TiO2 [1,17,51,52], while rGO has a work function of −4.42 eV [38,53,54,55,56]. Therefore, rGO can act as a trap for the electrons transferred to the conduction band of TiO2 (−4.2 eV) from ZnO and hinder the charge recombination, which enhances the overall performance of the ternary photocatalyst [50,57,58]. However, a very high rGO loading (above the optimum value of 2 wt %) decreases the active sites on the surface of the photocatalyst, which become adversely blocked by the excessive rGO, resulting in the observed reduction of the photoactivity of the ternary photocatalyst [59,60].
To elucidate the radical species that are responsible for the photocatalytic removal of MB, radical trapping experiments were conducted by the addition of tert-butanol, (2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (TEMPO), and ethylenediaminetetraacetic acid (EDTA) as •OH, •O2, and h+ traps, respectively, during the photocatalytic experiments (Figure 9).
The addition of EDTA, and thus the quenching of photogenerated holes, had an insignificant effect on the photocatalytic rates induced by the ZT-rGO photocatalyst with 2 wt % rGO. However, when quenching the OH radical with the alcohol, the MB removal efficiency decreased from 86% to 58% after 120 min of irradiation. This suggested that the highly reactive hydroxyl radicals produced by the photogenerated holes participate in the redox reactions. Moreover, an even more pronounced suppression of the photocatalytic performance (from 86% down to 40%) was observed upon scavenging the O2 species with TEMPO, which indicated that the superoxide anion radicals were the main active species in the removal of MB, in agreement with the literature [16,19,61]. As discussed above, the modification of ZnO–TiO2 with the rGO sheets increased the separation of electrons and holes, resulting in an increased migration of electrons on the graphene sheets, which reacted further with dissolved O2 present in the reaction solution to generate the highly responsive O2 species.

3. Materials and Methods

Zinc acetate dihydrate (Zn(CH3COO)2∙2H2O, 99%), hydrochloric acid (HCl, 37 wt %), APTMS (H2N(CH2)3Si(OCH3)3), sodium nitrate (NaNO3), and graphite were supplied by Sigma Aldrich (St. Louis, Missouri United States); while MB (82%) and titanium isopropoxide (Ti[OCH(CH3)2]4, 97%) were received from Aldrich. Potassium permanganate (KMnO4) was obtained from Scharlau (Barcelona, Spain). Sulfuric acid (H2SO4) was purchased from PENTA (Prague, Czech Republic), while hydrogen peroxide (H2O2, 30% solution in H2O) was provided by Chem-Lab. Ethanol (98%) was received from Honeywell (Charlotte, NC, USA) and was dried overnight over CaH2 prior to use. Finally, Milli-Q water, obtained from a Millipore apparatus, with a resistivity of 18.2 MΩ × cm at 298 K, was used for the preparation of all samples.

3.1. Synthesis of GO

For the synthesis of the GO, graphite powder (1.0 g) was first stirred into a mixture of H2SO4 (23 mL, 98%) and NaNO3 (0.5 g) in an ice bath. Then, KMnO4 (3 g) was slowly added over a period of 2 h. After 4 h, the reaction mixture was heated to 35 °C for 30 min and was added into H2O (50 mL), followed by heating at 70 °C for 15 min. The mixture was then poured into 250 mL H2O and the unreacted KMnO4 was removed by the addition of 3% H2O2. The GO was purified by several centrifugation/redispersion cycles in H2O and was dried under a vacuum overnight. An aqueous GO suspension (0.5 mg GO/mL H2O) was prepared by ultrasonication for 1 h.

3.2. Synthesis of ZnO Flower-Like Structures and ZnO–TiO2 Core-shell Heterostructures

The synthesis of the ZnO flower-like structures, as well as the ZnO-TiO2 core-shell, flower-like heterostructures, has been described in detail in our previous work [2]. Briefly, ZnO, with a flower-like morphology, was prepared via the reaction of 40 mL of a 0.035 M Zn(CH3COO)2·2H2O and 0.055 M NaOH aqueous solution at 70 °C for 24 h. For the synthesis of the ZnO-TiO2 core-shell, flower-like heterostructures, 0.1 g ZnO flowers were first dispersed in 35 mL ethanol by ultrasonication. A solution of 0.25 mL Ti[OCH(CH3)2]4 in 5 mL ethanol was prepared separately and was added dropwise in the ZnO dispersion. The dispersion was transferred in a preheated oil bath at 80 °C and a 10 mL mixture of water in ethanol (50 mL/L) was added via dropwise. After 2 h, the ZnO-TiO2 core-shell catalyst was isolated by centrifugation, washed thoroughly with ethanol, and dried under a vacuum.

3.3. Synthesis of ZnO–TiO2 Core-Shell Structures Wrapped with rGO

First, the ZnO–TiO2 core-shell structures (400 mg) were ultrasonically dispersed in 200 mL dry ethanol. Then 3 mL APTMS was injected in the dispersion, followed by an overnight reflux at 90 °C. Finally, the APTMS-modified ZnO–TiO2 flowers were rinsed thrice with ethanol and dried under a vacuum.
Next, 300 mg of the APTMS-modified ZnO–TiO2 heterostructures were dispersed in 100 mL H2O by ultrasonication, and an appropriate amount of the GO suspension was added under vigorous stirring to obtain ternary photocatalysts with 0.5, 2, and 5 wt % theoretical GO content. After stirring for 1 h, the GO-modified inorganic heterostructures were washed repeatedly with H2O and dried under a vacuum. This process was repeated for the ternary photocatalyst with 2 wt % GO, using core-shell heterostructures without the APTMS functionalization to elucidate the effect of the amine groups on the interactions between the GO sheets and the inorganic core-shell structures.
To induce the reduction of GO to rGO, the GO-modified ZnO–TiO2 flower-like particles were dispersed in a mixture of H2O/EtOH (2:1) and were hydrothermally treated at 120 °C for 24 h in a Teflon-lined, stainless-steel autoclave. Finally, the ternary photocatalysts were calcined at 500 °C for 2 h under N2, to further reduce rGO and crystallize TiO2.

3.4. Characterization

The ternary core-shell structures were characterized by Raman spectroscopy using a Nicolet Almega XR Raman spectrometer (Thermo Scientific, Waltham, MA, USA) with a 473 nm blue laser as the excitation source in high resolution mode. ATR-FTIR spectra were recorded on a Thermo Scientific Nicolet 6700 spectrometer in the 400–4000 cm−1 range. We collected 64 scans at a resolution of 4 cm−1. XRD patterns were measured on a PANalytical (Almelo, Netherlands) Xpert Pro X-ray diffractometer, using Cu-Kα radiation (45 kV and 40 mA). UV–Vis diffuse reflectance spectra in the 300–800 nm wavelength range were obtained using a Shimadzu UV-2401 PC spectrophotometer with an ISR-240A integrating sphere, with BaSO4 powder used as a 100% reflectance standard. Finally, the morphology of the samples was investigated by field emission scanning electron microscopy (FE-SEM, JEOL JSM-7000F).

3.5. Photocatalytic Study

The photocatalytic performance of the parent and the rGO-modified ZnO–TiO2 core-shell semiconductors was quantified via the decoloration of MB dye solutions (20 mg/L) under visible-light irradiation at pH 5 (corresponding to the natural pH of the solutions). The photocatalytic reactions were conducted at a photocatalyst loading of 160 ppm, while no catalyst was added for the photolytic experiments. The photocatalyst dispersion in the dye solutions was first stirred in the dark for ca. 40 min to establish the MB adsorption/desorption equilibrium, followed by irradiation using a medium-pressure mercury lamp emitting in the 200–600 nm range. The lamp was placed in a glass jacket that filtered out the UV lines at λexc < 390 nm (Table S1) [2,6]. All photocatalytic experiments were carried out in duplicate. To evaluate the reusability of the most photoactive photocatalyst, three photocatalytic cycles were conducted by recovering the catalyst after each cycle by centrifugation and washing thrice with H2O before the next cycle.

4. Conclusions

In summary, complex ZnO–TiO2 core-shell, flower-like structures wrapped with rGO sheets were prepared by a simple and well-controlled synthetic approach. First, core-shell flowers with a wurtzite ZnO core and an anatase TiO2 shell of 59 ± 7 nm thickness were prepared and their surfaces were functionalized with primary amino groups using APTMS. The surface-modified ZnO–TiO2 core-shell structures were then hydrothermally wrapped with rGO sheets at 0.5, 2, and 5 wt % loading. The amine functionalities on the inorganic surface induced strong adhesion of the rGO sheets on the TiO2 shell due to favorable electrostatic interactions. The rGO-modified ZnO–TiO2 photocatalysts exhibited superior photoactivity in the decoloration of aqueous solutions of MB under visible-light irradiation compared with that of the parent ZnO–TiO2 core-shell structures, which was dependent on the rGO loading. Specifically, the ternary photocatalyst with a 2 wt % rGO content resulted in ~86% dye removal after 120 min of irradiation, which was significantly higher than the ~51% decoloration found for the parent ZnO–TiO2. Overall, the enhanced photocatalytic response of the ternary photocatalysts was attributed to the increase in the absorption in the visible-light range, as well as the promoted electron and hole separation due to the favorable energy level differences between ZnO, TiO2, and rGO. Notably, excessive rGO loading above the optimum value of 2 wt % diminished the photoactivity of the ternary heterostructures to ~60% MB removal because of the decrease in free active sites on the surface of the photocatalyst.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2073-4344/11/3/332/s1, Figure S1. Higher magnification SEM images of the (a) bare ZnO flowers, (b) ZnO-TiO2 core-shell heterostructures, and (c) ZT-rGO structures at a 2 wt % rGO loading, Table S1. Radiation flux Φ for the immersion lamp used in this study, Figure S2. Decoloration of MB in the absence of a photocatalyst (●) and in the presence of the parent ZnO–TiO2 core-shell flowers (◼), and ZT-rGO at 0.5 wt % (◆), 2 wt % (◀), and 5 wt % (▲) rGO. All values were corrected for the adsorption component.

Author Contributions

Conceptualization, E.V., N.K., and M.V.; methodology and experiments, E.V.; data curation, E.V. and M.V.; writing-original draft preparation, E.V.; writing-review and editing, E.V., N.K., S.D., and M.V.; funding acquisition, N.K.; All authors have read and agreed to the published version of the manuscript.

Funding

This research was co-financed by the European Regional Development Fund of the European Union and Greek national funds through the Operational Program Competitiveness, Entrepreneurship and Innovation, under the call RESEARCH–CREATE–INNOVATE (project code: T1EDK–01633).

Data Availability Statement

The authors agree to the terms and conditions mentioned at https://0-www-mdpi-com.brum.beds.ac.uk/ethics (accessed on 5 March 2021).

Acknowledgments

The authors would like to gratefully acknowledge Aleka Manousaki and Dimitrios Theodoridis for assistance with the SEM and TEM measurements, respectively.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Kwiatkowski, M.; Bezverkhyy, I.; Skompska, M. ZnO nanorods covered with a TiO2 layer: Simple sol–gel preparation, and optical, photocatalytic and photoelectrochemical properties. J. Mater. Chem. A 2015, 3, 12748–12760. [Google Scholar] [CrossRef]
  2. Vasilaki, E.; Vamvakaki, M.; Katsarakis, N. Complex ZnO-TiO2 Core–Shell Flower-Like Architectures with Enhanced Photocatalytic Performance and Superhydrophilicity without UV Irradiation. Langmuir 2018, 34, 9122–9132. [Google Scholar] [CrossRef] [PubMed]
  3. Boyadjiev, S.I.; Kéri, O.; Bárdos, P.; Firkala, T.; Gáber, F.; Nagy, Z.K.; Baji, Z.; Takács, M.; Szilágyi, I.M. TiO2/ZnO and ZnO/TiO2 core/shell nanofibers prepared by electrospinning and atomic layer deposition for photocatalysis and gas sensing. Appl. Surf. Sci. 2017, 424, 190–197. [Google Scholar] [CrossRef] [Green Version]
  4. Pan, L.; Shen, G.-Q.; Zhang, J.-W.; Wei, X.-C.; Wang, L.; Zou, J.-J.; Zhang, X. TiO2–ZnO Composite Sphere Decorated with ZnO Clusters for Effective Charge Isolation in Photocatalysis. Ind. Eng. Chem. Res. 2015, 54, 7226–7232. [Google Scholar] [CrossRef]
  5. Chu, Y.-M.; Javed, H.M.; Awais, M.; Khan, M.I.; Shafqat, S.; Khan, F.S.; Mustafa, M.S.; Ahmed, D.; Khan, S.U.; Khalil, R.M. Photocatalytic Pretreatment of Commercial Lignin Using TiO2-ZnO Nanocomposite-Derived Advanced Oxidation Processes for Methane Production Synergy in Lab Scale Continuous Reactors. Catalysts 2021, 11, 54. [Google Scholar] [CrossRef]
  6. Vasilaki, E.; Georgaki, I.; Vernardou, D.; Vamvakaki, M.; Katsarakis, N. Ag-loaded TiO2/reduced graphene oxide nanocomposites for enhanced visible-light photocatalytic activity. Appl. Surf. Sci. 2015, 353, 865–872. [Google Scholar] [CrossRef]
  7. Lv, W.; Wei, B.; Xu, L.; Zhao, Y.; Gao, H.; Liu, J. Photocatalytic properties of hierarchical ZnO flowers synthesized by a sucrose-assisted hydrothermal method. Appl. Surf. Sci. 2012, 259, 557–561. [Google Scholar] [CrossRef]
  8. Qin, J.; Zhang, X.; Yang, C.; Cao, M.; Ma, M.; Liu, R. ZnO microspheres-reduced graphene oxide nanocomposite for photocatalytic degradation of methylene blue dye. Appl. Surf. Sci. 2017, 392, 196–203. [Google Scholar] [CrossRef]
  9. Zhang, B.; Yang, F.; Liu, H.; Yan, L.; Yang, W.; Xu, C.; Huang, S.; Li, Q.; Bao, W.; Liu, B.; et al. Assembling Graphene-Encapsulated Pd/TiO2 Nanosphere with Hierarchical Architecture for High-Performance Visible-Light-Assisted Methanol Electro-Oxidation Material. Ind. Eng. Chem. Res. 2019, 58, 19486–19494. [Google Scholar] [CrossRef]
  10. Sookhakian, M.; Amin, Y.M.; Zakaria, R.; Baradaran, S.; Mahmoudian, M.R.; Rezayi, M.; Tajabadi, M.T.; Basirun, W.J. Enhanced Photovoltaic Performance of Polymer Hybrid Nanostructure Heterojunction Solar Cells Based on Poly(3-hexylthiophene)/ZnS/ZnO/Reduced Graphene Oxide Shell–Core Nanorod Arrays. Ind. Eng. Chem. Res. 2014, 53, 14301–14309. [Google Scholar] [CrossRef]
  11. Ong, W.-J.; Voon, S.-Y.; Tan, L.-L.; Goh, B.T.; Yong, S.-T.; Chai, S.-P. Enhanced Daylight-Induced Photocatalytic Activity of Solvent Exfoliated Graphene (SEG)/ZnO Hybrid Nanocomposites toward Degradation of Reactive Black 5. Ind. Eng. Chem. Res. 2014, 53, 17333–17344. [Google Scholar] [CrossRef]
  12. Johra, F.T.; Jung, W.-G. RGO–TiO2–ZnO composites: Synthesis, characterization, and application to photocatalysis. Appl. Catal. A Gen. 2015, 491, 52–57. [Google Scholar] [CrossRef]
  13. Potle, V.D.; Shirsath, S.R.; Bhanvase, B.A.; Saharan, V.K. Sonochemical preparation of ternary rGO-ZnO-TiO2 nanocomposite photocatalyst for efficient degradation of crystal violet dye. Optik 2020, 208, 164555. [Google Scholar] [CrossRef]
  14. Liu, X.; Pan, L.; Lv, T.; Sun, Z. Investigation of photocatalytic activities over ZnO–TiO2–reduced graphene oxide composites synthesized via microwave-assisted reaction. J. Colloid Interface Sci. 2013, 394, 441–444. [Google Scholar] [CrossRef]
  15. Raghavan, N.; Thangavel, S.; Venugopal, G. Enhanced photocatalytic degradation of methylene blue by reduced graphene-oxide/titanium dioxide/zinc oxide ternary nanocomposites. Mater. Sci. Semicond. Process. 2015, 30, 321–329. [Google Scholar] [CrossRef]
  16. Divya, K.S.; Xavier, M.M.; Vandana, P.V.; Reethu, V.N.; Mathew, S. A quaternary TiO2/ZnO/RGO/Ag nanocomposite with enhanced visible light photocatalytic performance. New J. Chem. 2017, 41, 6445–6454. [Google Scholar] [CrossRef]
  17. Malekshoar, G.; Pal, K.; He, Q.; Yu, A.; Ray, A.K. Enhanced Solar Photocatalytic Degradation of Phenol with Coupled Graphene-Based Titanium Dioxide and Zinc Oxide. Ind. Eng. Chem. Res. 2014, 53, 18824–18832. [Google Scholar] [CrossRef]
  18. Van Bao, H.; Dat, N.M.; Giang, N.T.H.; Thinh, D.B.; Tai, L.T.; Trinh, D.N.; Hai, N.D.; Khoa, N.A.D.; Huong, L.M.; Nam, H.M.; et al. Behavior of ZnO-doped TiO2/rGO nanocomposite for water treatment enhancement. Surf. Interfaces 2021, 23, 100950. [Google Scholar] [CrossRef]
  19. Nguyen, C.H.; Tran, M.L.; Van Tran, T.T.; Juang, R.-S. Enhanced removal of various dyes from aqueous solutions by UV and simulated solar photocatalysis over TiO2/ZnO/rGO composites. Sep. Purif. Technol. 2020, 232, 115962. [Google Scholar] [CrossRef]
  20. Pan, X.; Yang, P.; Nan, H.; Yang, L.; Chen, H.; Zhao, X. Preparation and enhanced visible-light photoelectrocatalytic activity of ternary TiO2ZnO/RGO nanocomposites. Electrochim. Acta 2018, 261, 284–288. [Google Scholar] [CrossRef]
  21. Li, Q.; Li, X.; Swelm, W.; Al-Ghamdi, A.A.; Yu, J. CdS/Graphene Nanocomposite Photocatalysts. Adv. Energy Mater. 2015, 5, 1500010. [Google Scholar] [CrossRef]
  22. Liang, Y.; Wang, H.; Casalongue, H.S.; Chen, Z.; Dai, H. TiO2 nanocrystals grown on graphene as advanced photocatalytic hybrid materials. Nano Res. 2010, 3, 701–705. [Google Scholar] [CrossRef] [Green Version]
  23. Pei, F.; Xu, S.; Zuo, W.; Zhang, Z.; Liu, Y.; Cao, S. Effective improvement of photocatalytic hydrogen evolution via a facile in-situ solvothermal N-doping strategy in N-TiO2/N-graphene nanocomposite. Int. J. Hydrogen Energy 2014, 39, 6845–6852. [Google Scholar] [CrossRef]
  24. Cao, X.; Tian, G.; Chen, Y.; Zhou, J.; Zhou, W.; Tian, C.; Fu, H. Hierarchical composites of TiO2 nanowire arrays on reduced graphene oxide nanosheets with enhanced photocatalytic hydrogen evolution performance. J. Mater. Chem. A 2014, 2, 4366–4374. [Google Scholar] [CrossRef]
  25. Shen, J.; Shi, M.; Yan, B.; Ma, H.; Li, N.; Ye, M. Ionic liquid-assisted one-step hydrothermal synthesis of TiO2-reduced graphene oxide composites. Nano Res. 2011, 4, 795–806. [Google Scholar] [CrossRef]
  26. Poh, H.L.; Šaněk, F.; Ambrosi, A.; Zhao, G.; Sofer, Z.; Pumera, M. Graphenes prepared by Staudenmaier, Hofmann and Hummers methods with consequent thermal exfoliation exhibit very different electrochemical properties. Nanoscale 2012, 4, 3515–3522. [Google Scholar] [CrossRef]
  27. Johra, F.T.; Lee, J.-W.; Jung, W.-G. Facile and safe graphene preparation on solution based platform. J. Ind. Eng. Chem. 2014, 20, 2883–2887. [Google Scholar] [CrossRef]
  28. Stobinski, L.; Lesiak, B.; Malolepszy, A.; Mazurkiewicz, M.; Mierzwa, B.; Zemek, J.; Jiricek, P.; Bieloshapka, I. Graphene oxide and reduced graphene oxide studied by the XRD, TEM and electron spectroscopy methods. J. Electron. Spectrosc. Relat. Phenom. 2014, 195, 145–154. [Google Scholar] [CrossRef]
  29. Zhang, K.; Zhang, Y.; Wang, S. Enhancing thermoelectric properties of organic composites through hierarchical nanostructures. Sci. Rep. 2013, 3, 3448. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  30. Kudin, K.N.; Ozbas, B.; Schniepp, H.C.; Prud’Homme, R.K.; Aksay, I.A.; Car, R. Raman Spectra of Graphite Oxide and Functionalized Graphene Sheets. Nano Lett. 2008, 8, 36–41. [Google Scholar] [CrossRef]
  31. Kim, H.J.; Lee, S.-M.; Oh, Y.-S.; Yang, Y.-H.; Lim, Y.S.; Yoon, D.H.; Lee, C.; Kim, J.-Y.; Ruoff, R.S. Unoxidized Graphene/Alumina Nanocomposite: Fracture- and Wear-Resistance Effects of Graphene on Alumina Matrix. Sci. Rep. 2015, 4, 5176. [Google Scholar] [CrossRef] [PubMed]
  32. Wang, Y.; Li, Y.; Tang, L.; Lu, J.; Li, J. Application of graphene-modified electrode for selective detection of dopamine. Electrochem. Commun. 2009, 11, 889–892. [Google Scholar] [CrossRef]
  33. Shen, J.; Yan, B.; Shi, M.; Ma, H.; Li, N.; Ye, M. One step hydrothermal synthesis of TiO2-reduced graphene oxide sheets. J. Mater. Chem. 2011, 21, 3415–3421. [Google Scholar] [CrossRef]
  34. Tan, L.-L.; Ong, W.-J.; Chai, S.-P.; Mohamed, A.R. Reduced graphene oxide-TiO2 nanocomposite as a promising visible-light-active photocatalyst for the conversion of carbon dioxide. Nanoscale Res. Lett. 2013, 8, 465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Zhang, J.; Xiao, F.-X.; Xiao, G.; Liu, B. Self-assembly of a Ag nanoparticle-modified and graphene-wrapped TiO2 nanobelt ternary heterostructure: Surface charge tuning toward efficient photocatalysis. Nanoscale 2014, 6, 11293–11302. [Google Scholar] [CrossRef] [PubMed]
  36. Yang, X.; Chen, W.; Huang, J.; Zhou, Y.; Zhu, Y.; Li, C. Rapid degradation of methylene blue in a novel heterogeneous Fe3O4 @rGO@TiO2-catalyzed photo-Fenton system. Sci. Rep. 2015, 5, 10632. [Google Scholar] [CrossRef]
  37. Zheng, P.; Liu, T.; Su, Y.; Zhang, L.; Guo, S. TiO2 nanotubes wrapped with reduced graphene oxide as a high-performance anode material for lithium-ion batteries. Sci. Rep. 2016, 6, 36580. [Google Scholar] [CrossRef] [Green Version]
  38. Lee, J.S.; You, K.H.; Park, C.B. Highly Photoactive, Low Bandgap TiO2 Nanoparticles Wrapped by Graphene. Adv. Mater. 2012, 24, 1084–1088. [Google Scholar] [CrossRef] [PubMed]
  39. Lui, G.; Liao, J.-Y.; Duan, A.; Zhang, Z.; Fowler, M.; Yu, A. Graphene-wrapped hierarchical TiO2 nanoflower composites with enhanced photocatalytic performance. J. Mater. Chem. A 2013, 1, 12255–12262. [Google Scholar] [CrossRef] [Green Version]
  40. Liu, C.; Hao, F.; Zhao, X.; Zhao, Q.; Luo, S.; Lin, H. Low temperature reduction of free-standing graphene oxide papers with metal iodides for ultrahigh bulk conductivity. Sci. Rep. 2015, 4, 1–6. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  41. Shu, W.; Liu, Y.; Peng, Z.; Chen, K.; Zhang, C.; Chen, W. Synthesis and photovoltaic performance of reduced graphene oxide–TiO2 nanoparticles composites by solvothermal method. J. Alloys Compd. 2013, 563, 229–233. [Google Scholar] [CrossRef]
  42. Ngo-Duc, T.; Singh, K.; Meyyappan, M.; Oye, M.M. Vertical ZnO nanowire growth on metal substrates. Nanotechnology 2012, 23, 194015. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, R.; Yin, P.-G.; Wang, N.; Guo, L. Photoluminescence and Raman scattering of ZnO nanorods. Solid State Sci. 2009, 11, 865–869. [Google Scholar] [CrossRef]
  44. Singh, A.K.; Thool, G.S.; Bangal, P.R.; Madhavendra, S.S.; Singh, S.P. Low Temperature Mn Doped ZnO Nanorod Array: Synthesis and Its Photoluminescence Behavior. Ind. Eng. Chem. Res. 2014, 53, 9383–9390. [Google Scholar] [CrossRef]
  45. Min, Y.; Zhang, K.; Zhao, W.; Zheng, F.; Chen, Y.; Zhang, Y. Enhanced chemical interaction between TiO2 and graphene oxide for photocatalytic decolorization of methylene blue. Chem. Eng. J. 2012, 193–194, 203–210. [Google Scholar] [CrossRef]
  46. Zhang, H.; Lv, X.; Li, Y.; Wang, Y.; Li, J. P25-Graphene Composite as a High Performance Photocatalyst. ACS Nano 2009, 4, 380–386. [Google Scholar] [CrossRef]
  47. Sakthivel, S.; Kisch, H. Daylight Photocatalysis by Carbon-Modified Titanium Dioxide. Angew. Chem. Int. Ed. 2003, 42, 4908–4911. [Google Scholar] [CrossRef] [PubMed]
  48. Kwiatkowski, M.; Chassagnon, R.; Heintz, O.; Geoffroy, N.; Skompska, M.; Bezverkhyy, I. Improvement of photocatalytic and photoelectrochemical activity of ZnO/TiO2 core/shell system through additional calcination: Insight into the mechanism. Appl. Catal. B Environ. 2017, 204, 200–208. [Google Scholar] [CrossRef]
  49. Nguyen-Phan, T.-D.; Pham, V.H.; Shin, E.W.; Pham, H.-D.; Kim, S.; Chung, J.S.; Kim, E.J.; Hur, S.H. The role of graphene oxide content on the adsorption-enhanced photocatalysis of titanium dioxide/graphene oxide composites. Chem. Eng. J. 2011, 170, 226–232. [Google Scholar] [CrossRef]
  50. Chen, C.; Cai, W.; Long, M.; Zhou, B.; Wu, Y.; Wu, D.; Feng, Y. Synthesis of Visible-Light Responsive Graphene Oxide/TiO2 Composites with p/n Heterojunction. ACS Nano 2010, 4, 6425–6432. [Google Scholar] [CrossRef] [PubMed]
  51. Scanlon, D.O.; Dunnill, C.W.; Buckeridge, J.; Shevlin, S.A.; Logsdail, A.J.; Woodley, S.M.; Catlow, C.R.A.; Powell, M.J.; Palgrave, R.G.; Parkin, I.P.; et al. Band alignment of rutile and anatase TiO2. Nat. Mater. 2013, 12, 798–801. [Google Scholar] [CrossRef]
  52. Lei, Y.; Zhao, G.; Liu, M.; Zhang, Z.; Tong, X.; Cao, T. Fabrication, Characterization, and Photoelectrocatalytic Application of ZnO Nanorods Grafted on Vertically Aligned TiO2 Nanotubes. J. Phys. Chem. C 2009, 113, 19067–19076. [Google Scholar] [CrossRef]
  53. Zhang, J.; Xiong, Z.; Zhao, X.S. Graphene–metal–oxide composites for the degradation of dyes under visible light irradiation. J. Mater. Chem. 2011, 21, 3634–3640. [Google Scholar] [CrossRef]
  54. Liu, L.; Bai, H.; Liu, J.; Sun, D.D. Multifunctional graphene oxide-TiO2-Ag nanocomposites for high performance water disinfection and decontamination under solar irradiation. J. Hazard. Mater. 2013, 261, 214–223. [Google Scholar] [CrossRef] [PubMed]
  55. Wang, P.; Wang, J.; Ming, T.; Wang, X.; Yu, H.; Yu, J.; Wang, Y.; Lei, M. Dye-Sensitization-Induced Visible-Light Reduction of Graphene Oxide for the Enhanced TiO2 Photocatalytic Performance. ACS Appl. Mater. Interfaces 2013, 5, 2924–2929. [Google Scholar] [CrossRef]
  56. Wei, L.; Chen, S.; Yang, Y.; Dong, Y.; Song, W.; Fan, R. Reduced graphene oxide modified TiO2 semiconductor materials for dye-sensitized solar cells. RSC Adv. 2016, 6, 100866–100875. [Google Scholar] [CrossRef]
  57. Shang, Q.; Huang, X.; Tan, X.; Yu, T. High Activity Ti3+-Modified Brookite TiO2/Graphene Nanocomposites with Specific Facets Exposed for Water Splitting. Ind. Eng. Chem. Res. 2017, 56, 9098–9106. [Google Scholar] [CrossRef]
  58. Huang, Y.; Li, R.; Chen, D.; Hu, X.; Chen, P.; Chen, Z.; Li, D. Synthesis and Characterization of CNT/TiO2/ZnO Composites with High Photocatalytic Performance. Catalysts 2018, 8, 151. [Google Scholar] [CrossRef] [Green Version]
  59. Zhang, Y.; Tang, Z.-R.; Fu, X.; Xu, Y.-J. TiO2−Graphene Nanocomposites for Gas-Phase Photocatalytic Degradation of Volatile Aromatic Pollutant: Is TiO2−Graphene Truly Different from Other TiO2−Carbon Composite Materials? ACS Nano 2010, 4, 7303–7314. [Google Scholar] [CrossRef] [PubMed]
  60. Zhou, K.; Zhu, Y.; Yang, X.; Jiang, X.; Li, C. Preparation of graphene–TiO2 composites with enhanced photocatalytic activity. New J. Chem. 2011, 35, 353–359. [Google Scholar] [CrossRef]
  61. Simsek, E.B.; Kilic, B.; Asgin, M.; Akan, A. Graphene oxide based heterojunction TiO2 –ZnO catalysts with outstanding photocatalytic performance for bisphenol-A, ibuprofen and flurbiprofen. J. Ind. Eng. Chem. 2018, 59, 115–126. [Google Scholar] [CrossRef]
Figure 1. (a) XRD pattern, (b) Raman, and (c) FTIR spectra of graphene oxide (GO).
Figure 1. (a) XRD pattern, (b) Raman, and (c) FTIR spectra of graphene oxide (GO).
Catalysts 11 00332 g001
Figure 2. (a,b) SEM and (c,d) TEM images of the ZnO flower-like structures (a,c) and the ZnO–TiO2 core-shell heterostructures (b,d).
Figure 2. (a,b) SEM and (c,d) TEM images of the ZnO flower-like structures (a,c) and the ZnO–TiO2 core-shell heterostructures (b,d).
Catalysts 11 00332 g002
Figure 3. SEM images of rGO-modified ZnO–TiO2 structures (ZT-rGO) structures at a 2 wt % reduced GO (rGO) loading without (a) and with (b) 3-aminopropyltrimethoxysilane (APTMS) functionalization.
Figure 3. SEM images of rGO-modified ZnO–TiO2 structures (ZT-rGO) structures at a 2 wt % reduced GO (rGO) loading without (a) and with (b) 3-aminopropyltrimethoxysilane (APTMS) functionalization.
Catalysts 11 00332 g003
Figure 4. SEM images of ZT-rGO with (a) 0.5, (b) 2, and (c) 5 wt % rGO content; (d) ATR-FTIR spectra of the parent ZnO-TiO2 core-shell flowers (black line, solid) and ZT-rGO at 0.5 (red line, dash), 2 green line, dot), and 5 wt % (blue line, dash dot) rGO loading. The white arrows indicate unattached rGO sheets.
Figure 4. SEM images of ZT-rGO with (a) 0.5, (b) 2, and (c) 5 wt % rGO content; (d) ATR-FTIR spectra of the parent ZnO-TiO2 core-shell flowers (black line, solid) and ZT-rGO at 0.5 (red line, dash), 2 green line, dot), and 5 wt % (blue line, dash dot) rGO loading. The white arrows indicate unattached rGO sheets.
Catalysts 11 00332 g004
Figure 5. Raman spectra of (a) anatase TiO2 (black dotted line), the wurtzite ZnO flowers (red dashed line), and the ZnO-TiO2 core-shell flowers (blue solid line) and (b) ZT-rGO with 0.5 (red dotted line), 2 (green dashed line), and 5 wt % (blue solid line) rGO (the TiO2 peaks are indicated with a rhombus, the ZnO peaks with an asterisk and the rGO peaks with a circle).
Figure 5. Raman spectra of (a) anatase TiO2 (black dotted line), the wurtzite ZnO flowers (red dashed line), and the ZnO-TiO2 core-shell flowers (blue solid line) and (b) ZT-rGO with 0.5 (red dotted line), 2 (green dashed line), and 5 wt % (blue solid line) rGO (the TiO2 peaks are indicated with a rhombus, the ZnO peaks with an asterisk and the rGO peaks with a circle).
Catalysts 11 00332 g005
Figure 6. Normalized and transformed Kubelka–Munk function versus the light energy for the parent ZnO-TiO2 core-shell flowers (black line, solid) and ZT-rGO at 0.5 (red line, dash), 2 (green line, dot), and 5 wt % (blue line, dash dot) rGO loading.
Figure 6. Normalized and transformed Kubelka–Munk function versus the light energy for the parent ZnO-TiO2 core-shell flowers (black line, solid) and ZT-rGO at 0.5 (red line, dash), 2 (green line, dot), and 5 wt % (blue line, dash dot) rGO loading.
Catalysts 11 00332 g006
Figure 7. Decoloration of methylene blue (MB) in the absence of a photocatalyst (●) and in the presence of parent ZnO–TiO2 core-shell flowers (◼) and ZT-rGO at 0.5 (◆), 2 (◀), and 5 wt % (▲) rGO.
Figure 7. Decoloration of methylene blue (MB) in the absence of a photocatalyst (●) and in the presence of parent ZnO–TiO2 core-shell flowers (◼) and ZT-rGO at 0.5 (◆), 2 (◀), and 5 wt % (▲) rGO.
Catalysts 11 00332 g007
Figure 8. Decoloration of MB in the presence of the ZT-rGO catalyst with 2 wt % rGO for three photocatalytic cycles (cycle 1 ◀, cycle 2 ◼ and cycle 3 ●).
Figure 8. Decoloration of MB in the presence of the ZT-rGO catalyst with 2 wt % rGO for three photocatalytic cycles (cycle 1 ◀, cycle 2 ◼ and cycle 3 ●).
Catalysts 11 00332 g008
Figure 9. Decoloration of MB in the presence of ZT-rGO at 2 wt % rGO without scavenger (◀) and in the presence of 2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (TEMPO) (●), ethylenediaminetetraacetic acid (EDTA) (◼), and tert-butanol (▲) at 1.0 mM.
Figure 9. Decoloration of MB in the presence of ZT-rGO at 2 wt % rGO without scavenger (◀) and in the presence of 2,2,6,6-Tetramethylpiperidin-1-yl)oxyl (TEMPO) (●), ethylenediaminetetraacetic acid (EDTA) (◼), and tert-butanol (▲) at 1.0 mM.
Catalysts 11 00332 g009
Table 1. Calculated optical band-gap values (Eg) for the ZnO–TiO2 and the ZT-rGO core-shell structures.
Table 1. Calculated optical band-gap values (Eg) for the ZnO–TiO2 and the ZT-rGO core-shell structures.
SampleEg (eV)
ZnO–TiO23.08
ZT-rGO 0.5 wt %2.97
ZT-rGO 2 wt %2.84
ZT-rGO 5 wt %2.72
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Vasilaki, E.; Katsarakis, N.; Dokianakis, S.; Vamvakaki, M. rGO Functionalized ZnO–TiO2 Core-Shell Flower-Like Architectures for Visible Light Photocatalysis. Catalysts 2021, 11, 332. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11030332

AMA Style

Vasilaki E, Katsarakis N, Dokianakis S, Vamvakaki M. rGO Functionalized ZnO–TiO2 Core-Shell Flower-Like Architectures for Visible Light Photocatalysis. Catalysts. 2021; 11(3):332. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11030332

Chicago/Turabian Style

Vasilaki, Evangelia, Nikos Katsarakis, Spyros Dokianakis, and Maria Vamvakaki. 2021. "rGO Functionalized ZnO–TiO2 Core-Shell Flower-Like Architectures for Visible Light Photocatalysis" Catalysts 11, no. 3: 332. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11030332

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