Next Article in Journal
Performance of Catalytic Fast Pyrolysis Using a γ-Al2O3 Catalyst with Compound Modification of ZrO2 and CeO2
Next Article in Special Issue
Extending the Hierarchy of Heterogeneous Catalysis to Substituted Derivatives of Benzimidazole Synthesis: Transition Metals Decorated CNTs
Previous Article in Journal
Oxidized Palladium Supported on Ceria Nanorods for Catalytic Aerobic Oxidation of Benzyl Alcohol to Benzaldehyde in Protic Solvents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Indium Tin-Oxide Wrapped 3D rGO and TiO2 Composites: Development, Characterization, and Enhancing Photocatalytic Activity for Methylene Blue

1
School of Resources, Environmental, and Chemical Engineering, Nanchang University, No.999 Xuefu Avenue, Nanchang 330031, China
2
Key Lab of Poyang Lake Ecology and Bio-resource Utilization (Ministry of Education), Nanchang University, No.999 Xuefu Avenue, Nanchang 330031, China
*
Authors to whom correspondence should be addressed.
Catalysts 2019, 9(10), 848; https://doi.org/10.3390/catal9100848
Submission received: 2 September 2019 / Revised: 30 September 2019 / Accepted: 8 October 2019 / Published: 12 October 2019
(This article belongs to the Special Issue Hybrid Metal/Metal Oxide-Carbon Nanomaterials Catalysts)

Abstract

:
A hybrid material of indium tin-oxide (ITO) wrapped titanium dioxide and reduced graphene oxide (ITO-rGO and TiO2) was prepared using a facile hydrothermal technique. TiO2 nanorods were in situ grown on the surface of rGO (rGO and TiO2), and which was then assembled onto ITO substrate layer by layer with formation of a 3D structure. ITO-rGO and TiO2 exhibit low charge transfer resistance at the electrode-electrolyte interface and have good photoresponsive ability. Methylene blue (MB) can be effectively adsorbed and enriched onto ITO-rGO and TiO2 surface. The adsorption kinetics and thermodynamics of ITO-rGO and TiO2 were evaluated, showing that the exothermic and entropy-driven reaction were the main thermodynamic processes, and the Langmuir isotherm was the ideal model for adsorption fitting. Meanwhile, ITO greatly improved degradation of rGO and TiO2 because electrons can be collected by ITO before recombination and MB can easily enter into the 3D structure of rGO and TiO2. The highest photodegradation rate of MB reached 93.40% for ITO-rGO and TiO2 at pH 9. Additionally, ITO-rGO and TiO2 successfully solved the problems of being difficult to recycle and causing secondary pollution of traditional TiO2 catalysts. Therefore, ITO-rGO and TiO2 may be a potential photocatalyst for degrading organic pollutants in water.

Graphical Abstract

1. Introduction

Recently, photocatalysis have become one of the advanced processes for the degrading organic pollutants in wastewater. Among various semiconductor photocatalytic investigations, due to many advantages of titanium dioxide (TiO2), such as low cost, non-toxicity, and high stability, it has received extensive attention [1,2,3,4]. Unfortunately, the application of conventional TiO2 photocatalysts still face lots of drawbacks because of low electrical conductivity [5], wide band gap and high recombination rate of electron-hole pairs [6], difficulty separation, and secondary pollution. Therefore, to overcome these shortages of conventional TiO2 photocatalysts, different hybrid TiO2 materials have been investigated over the past years.
Among these hybrid materials, graphene-based TiO2 have been considered as highly photoactive materials [7,8,9,10,11,12]. Graphene has been well-known as a good catalyst for optimizing the photocatalytic performance of semiconductors. [12,13,14,15,16]. Reduced graphene oxide (rGO) modified with carboxylic acids, hydroxyl groups, and epoxy groups has obvious advantages such as sensitivity to chemical doping, species adsorbed or bound, and structural deformation [17]. Nowadays, many researchers have been devoted to preparation of rGO doping TiO2 photocatalysts for improving TiO2 catalyst performance [18,19]. However, there is a clear disadvantage that the intermolecular force between rGO and TiO2 is relatively weak and the way they combine via physisorption or van der Waals forces. [20,21]. Its weak connection can cause difficulties in electronic transmission between the rGO and TiO2. Therefore, finding a strong way to combine is necessary for rGO and TiO2 and the C–Ti bond could be a reliable way through a facile hydrothermal method [22].
As an adsorbent and catalyst, nanomaterials have some obvious shortcomings in suspension adsorption and photocatalytic systems due to their difficult solid-liquid separation, which may cause secondary pollution [19,20]. In order to solve the problem of difficult recycling and tackle dispersion, magnetic nanoparticles are used as a substrate to combine with TiO2 [23]. However, the formation of Fe2+ is due to the electrons obtained from Fe3+ which will cause the iron to be dissolved in the solution. [24]. Such a photochemical dissolution of Fe3+ might cause huge damage to the catalyst. In this study, indium tin-oxide (ITO) has many advantages as a promising material such as a low resistivity, and transparency in the visible region [25,26].
Methylene blue (MB) is the most common contaminant in nature and has been studied by many researchers [27,28]. Soma et al. reported removal of MB by rGO-metal oxide (TiO2/Fe3O4) based nanocomposites [29]. Wu et al. prepared of a tough TiO2-rGO-PDMAA nanocomposite hydrogel by one-pot UV photopolymerization for photodegradation of MB [30]. Unfortunately, these catalysts have drawbacks of being difficult to separate and causing secondary pollution. In this study, TiO2 and rGO can be easily recycled and taken out from water after bonding to ITO substrate. However, as far as we know, the adsorption and photocatalytic degradation of MB by ITO-TiO2 and rGO in pollution water have never been reported.
In this study, the indium tin-oxide wrapped 3D rGO and TiO2 composites (ITO-TiO2 and rGO) were prepared using a facile hydrothermal method. TiO2 and rGO have a strong electronic correlation as unpaired π electrons were bonded with unsaturated Ti atoms through the hydrothermal method. ITO-TiO2 and rGO exhibited synergy advantage of adsorption and photocatalysis. The most important is that ITO-TiO2 and rGO can be easily recycled and avoide secondary pollution. Meanwhile, ITO-TiO2 and rGO characterizations of morphology, composition, stability, and photoresponsive property were investigated. The static and dynamic adsorption processes were evaluated for ITO-TiO2 and rGO. Finally, photocatalytic performance and mechanism of ITO-TiO2 and rGO were proposed. This work shows that the combination of rGO and TiO2 can greatly reduce the wide band gap of the composite, which makes the electrons more prone to transition in the case of illumination. The combination of composite materials and ITO substrate provides a novel way of consideration for the preparation of environmentally friendly materials. Compared to a single catalyst, our composite catalysts have higher photocatalytic activity and better photoelectric response.

2. Results and Discussion

2.1. Construction of ITO-rGO and TiO2

The general preparation procedure of ITO-rGO and TiO2 is shown in Figure 1. GO can easily disperse into water due to many hydrophilic functional groups and can be reduced to rGO by DEG. TiO2 was prepared using PTO as a titanium source under hydrothermal conditions. The presence of PVP hinders the contact between titanium dioxide molecules, thus TiO2 was well dispersed in water solution to form nanorods with smaller particle sizes. TiO2 can be coated by rGO via electrostatic interaction and Ti–C bond. Finally, rGO and TiO2 was assembled onto the ITO substrate layer by layer via the annealing method. rGO and TiO2 showed 3D structure on the ITO substrate with excellent adsorption ability and photocatalytic activity for MB in pollution water. Importantly, ITO-rGO and TiO2 can be easily separated and recycled without secondary pollution.

2.2. Characterization of ITO-rGO and TiO2

2.2.1. XRD

The XRD patterns of different photocatalysts are shown in Figure 2. A strong diffraction peak appeared at 2θ of 10.5°, which correspond to the (001) plane of GO (Figure 2a) [31]. And from XRD pattern of TiO2 and rGO and ITO-TiO2 and rGO, many diffraction peaks appeared at 2θ of 25.307°, 37.792°, 48.043°, 53.886°, 55.068°, 62.689° and 75.051°, which correspond to the (101), (004), (200), (105), (211), (204) and (215) planes of anatase TiO2, respectively. Meanwhile, many diffraction peaks appeared at 2θ of 27.436°, 36.080°, 41.240° according to the XRD standard card (PDF NO. 78–2485), these diffraction peaks correspond to the (110), (101), and (111) planes of rutile TiO2. The XRD results of the pure TiO2 showed that the diffraction peaks all belonged to the anatase phase. However, TiO2 contained two phases of anatase and rutile after coating with rGO.
Figure 2b shows the XRD of ITO-TiO2 and rGO prepared at different reaction times. The diffraction peaks were located at 2θ of 25.307°, 37.792°, 48.043°, and 53.886°, which correspond to the (101), (004), (200), and (105) planes of anatase TiO2, respectively. These diffraction peaks significantly enhanced with increased reaction time, and new diffraction peaks appeared at 2θ of 27.436°, 36.080°, and 41.240°, which corresponded to the (110), (101), and (111) planes of rutile TiO2. Thus, anatase and rutile phase of TiO2 were also present in ITO-TiO2 and rGO. The diffraction peak intensity of TiO2 became stronger with increasing reaction time. Both types of diffraction peaks constantly became stronger under solvothermal condition in the presence of GO. Figure 2c shows the XRD of samples prepared by different DEG concentrations after reaction for 7 h. We can see that the XRD diffraction peaks are the same XRD diffraction peaks when DEG concentration was 3–7 mol/L. However, the diffraction peaks at 2θ of 27.436°, 36.080°, and 41.240° disappeared when DEG was 9 mol/L, which shows that TiO2 has only one crystal form of anatase phase.

2.2.2. SEM

To understand the formation mechanism of ITO-TiO2 and rGO during hydrothermal treatment, the morphology of ITO-TiO2 and rGO was investigated using SEM. Figure 3a shows that GO exists in the form of a sheet. The morphology of pure TiO2 represented nanorods (Figure 3b) and the length and diameter were approximately 0.8 μm and 20 nm, respectively. A previous report indicated that microsphere TiO2 were easily formed due to the aggregation of TiO2 nanoparticles to reduce the surface energy in the hydrothermal treatment under presence of PTO [32]. Meanwhile, the morphologies of ITO-TiO2 and rGO were investigated under different hydrothermal times. The nanorods were not observed when the reaction time was shorter than 3 h (Figure 3c). In addition, some TiO2 nanorods can be seen on the rGO surface after 5 h reaction. When the reaction time reached 7 h, the rGO was completely wrapped by TiO2 nanorods owing to the strong force between TiO2 and rGO. However, TiO2 nanorods on the rGO surface were connected together due to the increase in diameter of the nanorods when the reaction time reached 9 h, which leads to the specific surface area decrease.
Meanwhile, the morphology of ITO-TiO2 and rGO was also investigated when the variable factor was the concentration of DEG shown in Figure 4. Obviously, the TiO2 nanorods structure is the most obvious, has the largest specific surface area, and was highest when DEG concentration was 7 mol/L. Thus, the optimized condition for ITO-TiO2 and rGO preparation was reaction time of 7 h and DEG concentration of 7 mol/L.

2.2.3. TEM

Under such optimized conditions, TEM images of ITO-TiO2 and rGO were detected and results are shown in Figure 5. Figure 5a,b indicates that TiO2 nanorods are grown on the surface of rGO. Additionally, the SAED diffraction patterns of rGO showed that the reduction of GO was relatively complete because of six symmetrically diffraction points that can be clearly observed (Figure 5a inset). Meanwhile, a close-up image of TiO2 nanorods show that the diameter was 20 nm (Figure 5c) and the lattice spacing (d101 = 0.35 nm) was observed in TEM images (Figure 5d).

2.2.4. XPS

To investigate the surface elements of samples, XPS measurements of GO, TiO2, and TiO2 and rGO were carried out (Figure 6). It is obvious that GO contained only C and O elements. However, it also contained K element for TiO2 and TiO2 and rGO samples due to PTO (K2TiO(C2O4)2) (Figure 6a). Figure 6b shows the C1s spectra, which indicated that the binding energies at 284.4 eV (C–C/C=C) and 290.1 eV (O–C=O) were observed [33]. The two main C1s peaks for TiO2 and rGO sharply decreased, confirming the reduction of GO to rGO (Figure 6b). Figure 6c exhibits two main peaks located at 457.9 and 464 eV, which could correspond to Ti 2p3/2 and Ti 2p1/2, respectively [34]. However, these two peaks of TiO2 and rGO shifted towards higher binding energies compared with TiO2. Due to the d–π electron orbital overlap, the binding energy of Ti4+ shifted by 0.4 eV [35]. This makes the transmission of electrons more convenient [36] and makes the recombination rate of electron-hole pairs greatly reduced. In Figure 6d the two peaks at 529.1 and 532.6 eV correspond to lattice oxygen (Ti–O) and hydroxyl groups (–OH), respectively.

2.2.5. Raman Spectra

The results of Raman spectra are shown in Figure 7a, four peaks located at frequencies of 149, 400, 515, and 634 cm–1 were consistent with the anatase phase of TiO2 with the D4h19 (I41/amd) space group [37]. The intensity of these peaks sharply decrease under the presence of GO for the TiO2 and rGO. It can be determined that TiO2 nanorods were actually grown on the rGO. The weak peak that corresponds to rutile TiO2 was also observed at 230 cm–1, and it is also compounded with the results of XRD. In Figure 7b, the D and G bands of GO appear at 1351 and 1602 cm–1 [38]. The disorder and defects are generated when the O-rich groups above the rGO are removed which can reduce the recombination of electron-hole pairs [39]. The ID/IG means the ratio of the intensity of two peaks which could reflect the number of defects. The ID/IG ratio of GO and TiO2 and rGO were 1.12 and 1.14, respectively. This increase shows that the sample has few defects in the complete sp2 hybrid structure. Moreover, the peak intensity of ID for the TiO2 and rGO sample increased compared with GO, showing that the amounts of rGO were reduced.

2.2.6. FT-IR and UV-Vis Spectra

The reduction of GO to rGO could be further confirmed in FT-IR spectra (Figure S1). The peak of 1730cm–1 corresponding to the chemical bond C=O has disappeared for TiO2 and rGO compared with GO, showing that DEG have the ability to reduce GO to rGO. A new peak at 987 cm–1 appeared because of Ti–O–C bonds between TiO2 and rGO [40,41].
The light-harvesting ability of semiconductors played a vital role in its photocatalytic activity. The absorption spectrum of TIO-TiO2 and rGO was investigated using a diffuse reflectance UV-Vis recording spectrophotometer. We can see that the photocatalysts exhibit similar diffuse reflectance spectra (Figure 8a). The absorption spectrum of TiO-TiO2 and rGO and TiO2 and rGO composite was decreased compared with pure TiO2, which was due to the band gap transition [42]. The band gap could obtain according to Kubelka–Munk equation [43].
F(R) = (1 − R)2/2R
F(R)hv = C1(hv − Eg)2
As shown in Figure 8b, the band gap of TiO2 was 3.14 eV, and that of ITO-rGO and TiO2 and rGO and TiO2 was about 2.78 and 2.79eV, respectively. The optical band gaps were decreased from 3.14 to 2.78 eV. The addition of rGO resulted in the reduced band gaps and improved absorption and photocatalytic activity of ITO-rGO and TiO2.

2.3. Photoelectrical Performance

Figure 9 shows the photocurrent response of ITO-rGO and TiO2 at four on-off cycles, which showed that the photocurrents of electron-hole pairs immediately generated when ITO-rGO and TiO2 electrodes were illuminated. The photocurrent value was increased after coated with rGO. The separated e/h+ pairs immediately recombined when light was turned off, which caused the photocurrent to rapidly drop density to almost zero. The reduction in the band gap of ITO-rGO and TiO2 made the transition of the electrons from the valence band to the conducting band easy. Subsequently, the separated electrons migrated to the ITO substrate [44]. Compared to its excellent electrical conductivity, foam nickel has poor light transmission. The ITO glass has a good light transmission property, so that the sample can be illuminated in all directions. The value of the photocurrent is the largest compared with the sample loaded on the foamed nickel. When the light is turned on again, the photocurrent density was not significantly reduced compared to the previous steady photocurrent density value. Therefore, ITO-rGO and TiO2 exhibited stronger photoresponsive property with good reproducibility.

2.4. Adsorption and Photocatalytic Properties

2.4.1. Adsorption Kinetics

Figure 10 shows that the adsorption capacity of MB increased with increasing MB concentration. At the beginning, there were more unadsorbed MB molecules that provide a larger adsorption driving force, so the initial adsorption rate was faster. Meanwhile, the Figure 10a results also indicate that the adsorption rate was fast and reached the highest adsorption ability within 30 min. Kinetic studies can use the following equation:
Pseudo-first-order-model :   log ( q e q t )   =   log q e k 1 t 2.303
Pseudo-second-order-model :   t q t   = 1   k 2 q e 2   +   t q e
As shown in Table S1 and Figure S3, the pseudo-second-order model show a good linearity with higher R2 of 0.9960, and the values of qe,cal and qe,exp were close, which suggests that the dynamic adsorption processes fitted the Pseudo-second-order model for ITO-rGO and TiO2.
Stability and regenerability are an important indicator of evaluation for the adsorbent. As we all know, TiO2 based photocatalysts are usually used under the condition of powders, which leads to difficulty separating the power from the water environment. Thus, the reusable ability of ITO-rGO and TiO2 was investigated using adsorption-desorption experiments (Figure 10b). The adsorption capacity of the sample did not decrease significantly after five cycles showing that ITO-TiO2 and rGO has excellent regeneration and stability ability. The most important is that ITO-TiO2 and rGO was easily recovered from pollution water and thus prevent secondary pollution.
The adsorption capacity of MB increases first and then decreases with the increase of pH value as shown in Figure 10c. As a typical cationic dye, MB is ionized and positively charged, so coulomb rejection occurs between the positive charge on the surface of TiO2 and MB at a relatively low pH range, which results in a relatively low adsorption at pH = 3 for ITO-rGO and TiO2. When pH increased, the electrostatic attraction between the TiO2 and MB increased, leading to an increase in the adsorption capacity [45]. The adsorbability of ITO-rGO and TiO2 reached its maximum when the pH was 9.

2.4.2. Adsorption Thermodynamics

As we all know, thermodynamic studies provide a deeper reference for the adsorption process. The effect of temperature on the adsorption of MB was also investigated. Gibbs free energy can be calculated by the following equation [45]:
Δ G 0 = RT   lnK d
A high negative value of ΔG0 represents that the adsorption process is spontaneous. This result is shown in Figure S4 and Table S2, indicating that ΔG0 values were negative and thus the adsorption process was spontaneous. The equations of enthalpy change and entropy change can be expressed by the following equation [45].
lnK d = Δ S 0 R     Δ H 0 RT

2.4.3. Adsorption Isotherm

Langmuir isotherms are often applied to the adsorption surface homogenization, and all adsorption sites have the same adsorption affinity. In addition, the Freundlich isotherm model is more focused on homogeneous.
Langmuir   isotherm :   C e q e   = 1   bq m   +   C e q m
Freundlich   isotherm :   logq e   =   logK f   +   1 n logC e
Figure S5 shows the application of nonlinear Langmuir and Freundlich isotherms to the MB adsorption on ITO-rGO and TiO2, and the calculated coefficients are listed in Table S3. It can be seen that the Langmuir isotherm was better for fitting the experimental data with higher regression coefficient (R2) than that of the Freundlich model, which indicates that the active sites on the surface of ITO-rGO and TiO2 were uniform with homogenous adsorption. Furthermore, the essential feature of the Langmuir isotherm can be expressed in terms of a dimensionless constant separation factor (RL) as the following equation:
R L   =   1 1   +   bC 0
As we all know, the RL values show the shape of the isotherm to be either unfavorable (RL > 1), linear (RL = 1), favorable (0 < RL < 1), or irreversible (RL = 0), respectively. The RL in this study was in the range of 0.166–0.375 for MB adsorption on ITO-rGO and TiO2, which indicated that the shape of isotherm was favorable adsorption.

2.4.4. Photocatalytic Activity

Photocatalytic degradation of MB by ITO-rGO and TiO2 were also investigated (Figure 11). It can be seen that in the absence of a photocatalyst (blank), MB is hardly degraded, indicating that MB has high stability under ultraviolet irradiation. However, a great MB degradation was observed in the presence of the photocatalyst of rGO and TiO2 and ITO-rGO and TiO2. Interestedly, the photocatalytic activity of rGO and TiO2 was greatly enhanced after bonding onto the ITO substrate. About 93.40% of MB was decomposed by ITO-rGO and TiO2 within 120 min. The higher photocatalytic degradation of ITO-rGO and TiO2 may be due to the following reasons. Firstly, the electrons from rGO can be collected by the ITO substrate, which lead to improved electron transfer from MB to rGO. Secondly, the electron/hole recombination could be inhibited by ITO-rGO and TiO2 and then extend the holes (h+) existence time.
An apparent first-order equitation was utilized for comparison of the degradation rate constant of MB on the ITO-rGO and TiO2:
log ( C 0 C )   =   K app t
where Kapp is the apparent first-order rate constant. As shown in Figure 11b, after modification with rGO and assembly onto ITO, the photocatalytic activity of titanium dioxide is greatly improved. The rate constant of ITO-rGO and TiO2 was 0.0088 min–1, which are 1.7 and 1.16 times larger than the rate constants of pure TiO2 and rGO and TiO2, respectively, which further indicates that the photocatalytic activity of TiO2 combined with ITO increases. The most important is that ITO-rGO and TiO2 can be easily recycled and avoid secondary pollution. As we all know, the pH plays an important role in the photocatalytic degradation of contaminants. Thus, the effect of pH on the photodegradation of MB by ITO-rGO and TiO2 was investigated in the pH range from 3 to 11 and the results are shown in Figure 11c,d. According to Equation (10), the pseudo first order rate constants for ITO-rGO and TiO2 were 0.00371, 0.00549, 0.00669, 0.1275, and 0.1241 cm−1 in pH solutions of 3, 5, 7, 9, and 11, respectively, showing that ITO-rGO and TiO2 has the highest photocatalytic ability when the pH value was 9, which was due to the strong adsorption of MB on the ITO-rGO and TiO2 at such condition.
Reusability tests and a degradation stability test were also investigated and are shown in Figures S6 and S7. The catalyst was recycled five times to degrade the new undegraded MB solution. The degradation rate did not significantly decrease. In addition, in order to confirm that the MB solution is truly degraded, the degraded MB solution detects the absorbance at fixed time intervals. Figure S7 show that the MB solution is completely degraded.

2.5. Mechanism of Degradation

Based on the above systematic studies, the photocatalytic degradation mechanism of MB by ITO-rGO and TiO2 is proposed in Figure 12. Firstly, cationic MB molecules were efficiently adsorbed and enriched on the surface of ITO-rGO and TiO2 due to electrostatic attraction as well as π–π conjugation between MB and rGO. Then, the hydrophilic functional groups on rGO can improve the wettability of ITO-rGO and TiO2 in aqueous solution, and the 3D structure of rGO and TiO2 can shorten the mass transfer distance for the degradation process, which results in enhancement of MB concentration on the surface of ITO-rGO and TiO2. In addition, the typical work functions of TiO2, graphene, and MB are −4.40, −4.42, and −3.60 eV, respectively [42], which showed that the electrons of MB and TiO2 can be accepted by rGO. The adsorbed MB can provide electrons to the graphene (or directly to the photocatalyst) in a photoexcited state (MB*). The electrons transfer through the rGO can be collected by the ITO substrate. ITO can inhibit electron/cloudy recombination, thereby extending the length (h+) of the time of existence. Finally, MB can be converted to CO2, H2O, and some small molecules. Therefore, the synergy of adsorption-enrichment and photocatalytic reaction was the main mechanism for the high decomposition rate of ITO-rGO and TiO2.

3. Experimental Section

3.1. Materials

Graphene oxide (GO) was purchased from Shanghai Yuanye Biotechnology Co. Ltd (Shanghai, China). Diethylene glycol (DEG) was obtained from Shanghai Aladdin Biochemical Technology Co. Ltd (Shanghai, China). Tartaric acid was purchased from Shanghai Zhanyun Chemical Co., Ltd (Shanghai, China). Dipotassium Titanium Oxide Dioxalate (PTO) was purchased from Shanghai D&B Biological Science and Technology Co. Ltd (Shanghai, China). Methylene blue (MB) powder was purchased from Sigma Aldrich. The system with the water outlet operated at 18.2 MΩ (Millipore, Bedford, MA, USA). ITO was purchased from Kaiwei (Zhuhai, China) for the Optoelectronics Technology Co. Ltd.

3.2. Preparation of rGO-TiO2

GO powder (0.008 g) was dispersed into water and ultrasonic treated for 3 h at room temperature. The suspensions were then added into a reactor containing a certain amount of DEG (3–9 mol/L), and the total volume of the solution was 80 mL. After adding PTO powder (0.5 g), the suspensions were stirred for 12 h at 60 °C. Polyvinyl pyrrolidone (PVP) (0.5 g) was also added into the solutions and dried at 180 °C for (2–10 h). The powder (rGO and TiO2) was washed with distilled water after the reaction was over. Next, it was dried in a vacuum drying oven at 50 °C and then further calcined in nitrogen at 650 °C for 2 h.

3.3. Fabrication of ITO-rGO and TiO2

The powder (rGO and TiO2) were firstly evenly dispersed in 5 mg/mL ethanol. A total of 5 μL of the suspension was dripped on ITO conductive substrate surface. A film formed on the ITO substrate when ethanol evaporated. Six cycles of dropping were carried out. The ITO-rGO and TiO2 was then placed at 450 °C under a nitrogen flow for 1 h.

3.4. Characterizations

The morphology of TiO2 powders was identified using an environmental scanning electron microscope (SEM) (JSM-6701F, JEOL, Tokyo, Japan) and transmission electron microscope (TEM) and high-resolution TEM (HRTEM) (JEOL, Tokyo, Japan). X-ray diffractometry (XRD, D8 ADVANCE, Bruker, Karlsruhe, Germany) was employed to characterize the phase structure of the samples. The photocatalytic reaction was carried out in a photochemistry reaction instrument (BL-GHX-V, BILON, Shanghai, China). Raman spectra were recorded on a micro-Raman spectrometer (XploRA PLUS, HORIBA, Beijing, China). X-ray photoelectron spectroscopy (XPS) measurements were conducted by using a Multilab 2000 XPS system (ESCALAB250Xi, ThermoFisher Scientific, Waltham, USA). The Fourier transform infrared (FT-IR) spectra were recorded on an infrared spectrometer (Nicolet5700, Thermo Nicolet, Waltham, USA). Photocurrent measurements were conducted on an electrochemical station (CHI660, Chen Hua, China). A 3 W LED lamp (UVEC-4 II, lamplic, Shenzhen, China) was used as the light source. The UV–vis absorption spectra of the samples were obtained by using a UV–vis spectrophotometer (Cary 100, Agilent Technologies Inc., Beijing, China).

3.5. Photoelectrical Performance

In order to test the photoelectrochemical performance of the sample, a three-electrode system that consisted of ITO-rGO and TiO2 (the work electrode), platinum wire (the counter electrode), and Ag/AgCl electrode (the reference electrode) was used. The 3 W UV lamp was used as light source and the responding photocurrent was recorded by electrochemical workstation. The electrolyte was 0.4 mol/L Na2SO4 solution.

3.6. Photocatalytic Oxidation Reactions of ITO-rGO and TiO2

First, 0.01 g of ITO-rGO and TiO2 was added into a quartz reaction tube with 70 mL MB for the photocatalytic reaction. In order to achieve the adsorption-desorption equilibrium, the suspensions were stirred in the dark for 2 h. The UV absorbance of the MB solution was tested every 20 min. The effect of pH (1–11) and temperature (298, 308, and 318 K) on the removal of MB were also investigated. The amount of MB adsorbed per unit mass can be expressed by this equation
q = (C0 − Ce)V/m.

4. Conclusions

In this study, rGO and TiO2 was successfully fabricated by the hydrothermal treatment, which was then assembled onto ITO substrate to form a novel photocatalyst of ITO-rGO and TiO2. MB can be efficiently enriched on the surface of ITO-rGO and TiO2 and then easily enter into the 3D structure of rGO and TiO2. Due to the efficient transfer and separation of photogenerated electrons and holes, rGO can be used as an effective promoter for improving the photocatalytic performance of TiO2. Meanwhile, ITO substrate can collect electrons from rGO by inhibiting the recombination of electrons. The ITO-rGO and TiO2 has the advantage of higher adsorption and photocatalytic ability. The most important is that ITO-rGO and TiO2 can be easily recycled with high stability, which prevents secondary pollution from traditional TiO2 photocatalysts. Therefore, ITO-rGO and TiO2 may be a promising photocatalyst in the application of water pollution.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/9/10/848/s1, Figure S1: FTIR of GO, TiO2 and ITO-rGO and TiO2, Figure S2: Graph of an ideal photocurrent response, Figure S3: Pseudo-first-order kinetics and pseudo-second-order kinetics of MB adsorption on the ITO-rGO and TiO2 (prepared by 9 mol/L DEG concentrations at 25 °C), Table S1: Adsorption isotherm parameters for MB adsorption on the TiO2 and rGO composite, Figure S4: Van’t Hoff plot of lnKd vs. 1/T, Table S2: Adsorption thermodynamic parameters for MB adsorption on the TiO2 and rGO composite, Figure S5: Isotherms of MB adsorption on the ITO-rGO and TiO2 composite, Table S3: Adsorption isotherm parameters for MB adsorption on the TiO2 and rGO composite, Figure S6: Degradation ratio of MB on ITO-rGO and TiO2 in five successive cycles of complete photocatalytic process, Figure. S7: Degradation rate of each interval test sample after the end of the photocatalytic degradation.

Author Contributions

Conceptualization, C.G.; Methodology, J.D. and H.P.; Software, S.X., P.X., X.W. and F.L.; Formal analysis, C.G.; Investigation, C.G., Y.X. and J.Y.; Data curation, R.L.; Writing—original draft preparation, C.G.; writing—review and editing, J.D. and H.P; Supervision, J.D. and H.P.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 51162022 and 31660482), Outstanding Youth Foundation of Jiangxi (20171BCB23010), and the Test Foundation of Nanchang University (Grant No.2012019).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sambur, J.B.; Chen, T.; Choudhary, E.; Chen, G.; Nissen, E.J.; Thomas, E.M.; Zou, N.; Chen, P. Sub-particle reaction and photocurrent mapping to optimize catalyst-modified photoanodes. Nature 2016, 530, 77–80. [Google Scholar] [CrossRef] [PubMed]
  2. Butburee, T.; Bai, Y.; Wang, H.; Chen, H.; Wang, Z.; Liu, G.; Zou, J.; Khemthong, P.; Lu, G.Q.M.; Wang, L. 2D Porous TiO2 Single-Crystalline Nanostructure Demonstrating High Photo-Electrochemical Water Splitting Performance. Adv. Mater. 2018, 30, 1705666. [Google Scholar] [CrossRef] [PubMed]
  3. Menny, S.; Snir, D.; Sven, R.; Larissa, G.; Arie, Z. Core/CdS Quantum Dot/Shell Mesoporous Solar Cells with Improved Stability and Efficiency Using an Amorphous TiO2 Coating. J. Phys. Chem. C 2009, 113, 3895–3898. [Google Scholar]
  4. Wang, Y.; Gu, L.; Guo, Y.; Li, H.; He, X.; Tsukimoto, S.; Ikuhara, Y.; Wan, L. Rutile-TiO2 Nanocoating for a High-Rate Li4Ti5O12 Anode of a Lithium-Ion Battery. J. Am. Chem. Soc. 2012, 134, 7874–7879. [Google Scholar] [CrossRef] [PubMed]
  5. 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]
  6. Tan, L.; Chai, S.; Mohamed, A.R. Synthesis and Applications of Graphene-Based TiO2 Photocatalysts. ChemSusChem 2012, 5, 1868–1882. [Google Scholar] [CrossRef] [PubMed]
  7. Faraldos, M.; Bahamonde, A. Environmental applications of titania-graphene photocatalysts. Catal. Today 2017, 285, 13–28. [Google Scholar] [CrossRef]
  8. Zhang, N.; Yang, M.; Liu, S.; Sun, Y.; Xu, Y. Waltzing with the Versatile Platform of Graphene to Synthesize Composite Photocatalysts. Chem. Rev. 2015, 115, 10307–10377. [Google Scholar] [CrossRef]
  9. Han, C.; Zhang, N.; Xu, Y. Structural diversity of graphene materials and their multifarious roles in heterogeneous photocatalysis. Nano Today 2016, 11, 351–372. [Google Scholar] [CrossRef]
  10. Lu, K.; Chen, Y.; Xin, X.; Xu, Y. Rational utilization of highly conductive, commercial Elicarb graphene to advance the graphene-semiconductor composite photocatalysis. Appl. Catal. B Environ. 2018, 224, 424–432. [Google Scholar] [CrossRef]
  11. Geim, A.K. Graphene: Status and Prospects. Science 2009, 324, 1530–1534. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Huang, X.; Yin, Z.; Wu, S.; Qi, X.; He, Q.; Zhang, Q.; Yan, Q.; Boey, F.; Zhang, H. Graphene-based materials: Synthesis, characterization, properties, and applications. Small 2011, 7, 1876–1902. [Google Scholar] [CrossRef] [PubMed]
  13. Tu, W.; Zhou, Y.; Zou, Z. Versatile Graphene-Promoting Photocatalytic Performance of Semiconductors: Basic Principles, Synthesis, Solar Energy Conversion, and Environmental Applications. Adv. Funct. Mater. 2013, 23, 4996–5008. [Google Scholar] [CrossRef]
  14. Xiang, Q.; Yu, J.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2012, 41, 782–796. [Google Scholar] [CrossRef] [PubMed]
  15. Li, Y.; Chen, J.; Huang, L.; Li, C.; Hong, J.; Shi, G. Highly Compressible Macroporous Graphene Monoliths via an Improved Hydrothermal Process. Adv. Mater. 2014, 26, 4789–4793. [Google Scholar] [CrossRef] [PubMed]
  16. Low, J.; Yu, J.; Ho, W. Graphene-Based Photocatalysts for CO2 Reduction to Solar Fuel. J. Phys. Chem. Lett. 2015, 6, 4244–4251. [Google Scholar] [CrossRef] [PubMed]
  17. 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]
  18. Tolosana-Moranchel, Á.; Manassero, A.; Satuf, M.L.; Alfano, O.M.; Casas, J.A.; Bahamonde, A. Influence of TiO2-rGO optical properties on the photocatalytic activity and efficiency to photodegrade an emerging pollutant. Appl. Catal. B Environ. 2019, 246, 1–11. [Google Scholar] [CrossRef]
  19. Pu, S.; Zhu, R.; Ma, H.; Deng, D.; Pei, X.; Qi, F.; Chu, W. Facile in-situ design strategy to disperse TiO2 nanoparticles on graphene for the enhanced photocatalytic degradation of rhodamine 6G. Appl. Catal. B Environ. 2017, 218, 208–219. [Google Scholar] [CrossRef]
  20. Yang, N.; Zhai, J.; Wang, D.; Chen, Y.; Jiang, L. Two-Dimensional Graphene Bridges Enhanced Photoinduced Charge Transport in Dye-Sensitized Solar Cells. ACS Nano 2010, 4, 887–894. [Google Scholar] [CrossRef]
  21. Dai, Y.; Jing, Y.; Zeng, J.; Qi, Q.; Wang, C.; Goldfeld, D.; Xu, C.; Zheng, Y.; Sun, Y. Nanocables composed of anatase nanofibers wrapped in UV-light reduced graphene oxide and their enhancement of photoinduced electron transfer in photoanodes. J. Mater. Chem. 2011, 21, 18174. [Google Scholar] [CrossRef]
  22. Huang, Q.; Tian, S.; Zeng, D.; Wang, X.; Song, W.; Li, Y.; Xiao, W.; Xie, C. Enhanced Photocatalytic Activity of Chemically Bonded TiO2/Graphene Composites Based on the Effective Interfacial Charge Transfer through the C-Ti Bond. ACS Catal. 2013, 3, 1477–1485. [Google Scholar] [CrossRef]
  23. Li, Z.; Huang, Z.; Guo, W.; Wang, L.; Zheng, L.; Chai, Z.; Shi, W. Enhanced Photocatalytic Removal of Uranium (VI) from Aqueous Solution by Magnetic TiO2/Fe3O4 and Its Graphene Composite. Environ. Sci. Technol. 2017, 51, 5666–5674. [Google Scholar] [CrossRef] [PubMed]
  24. Zhang, P.; Xiang, M.; Liu, H.; Yang, C.; Deng, S. Novel Two-Dimensional Magnetic Titanium Carbide for Methylene Blue Removal over a Wide pH Range: Insight into Removal Performance and Mechanism. ACS Appl. Mater. Interfaces 2019, 11, 24027–24036. [Google Scholar] [CrossRef] [PubMed]
  25. Le, Q.T.; Nüesch, F.; Rothberg, L.J.; Forsythe, E.W.; Gao, Y. Photoemission study of the interface between phenyl diamine and treated indium–tin–oxide. Appl. Phys. Lett. 1999, 75, 1357–1359. [Google Scholar] [CrossRef]
  26. Monamary, A.; Vijayalakshmi, K.; Jereil, S.D. Fe overlayered hybrid TiO2/ITO nanocomposite sensor for enhanced hydrogen sensing at room temperature by novel two step process. Sens. Actuators B Chem. 2019, 287, 278–289. [Google Scholar] [CrossRef]
  27. Payra, S.; Challagulla, S.; Bobde, Y.; Chakraborty, C.; Ghosh, B.; Roy, S. Probing the photo- and electro-catalytic degradation mechanism of methylene blue dye over ZIF-derived ZnO. J. Hazard. Mater. 2019, 373, 377–388. [Google Scholar] [CrossRef]
  28. Wolski, L.; Ziolek, M. Insight into pathways of methylene blue degradation with H2O2 over mono and bimetallic Nb, Zn oxides. Appl. Catal. B Environ. 2018, 224, 634–647. [Google Scholar] [CrossRef]
  29. Banerjee, S.; Benjwal, P.; Singh, M.; Kar, K.K. Graphene oxide (rGO)-metal oxide (TiO2/Fe3O4) based nanocomposites for the removal of methylene blue. Appl. Surf. Sci. 2018, 439, 560–568. [Google Scholar] [CrossRef]
  30. Wu, D.; Yi, M.; Duan, H.; Xu, J.; Wang, Q. Tough TiO2 -rGO-PDMAA nanocomposite hydrogel via one-pot UV polymerization and reduction for photodegradation of methylene blue. Carbon 2016, 108, 394–403. [Google Scholar] [CrossRef]
  31. Ranjan, P.; Agrawal, S.; Sinha, A.; Rao, T.R.; Balakrishnan, J.; Thakur, A.D. A Low-Cost Non-explosive Synthesis of Graphene Oxide for Scalable Applications. Sci. Rep. 2018, 8. [Google Scholar] [CrossRef] [PubMed]
  32. Lv, K.; Fang, S.; Si, L.; Xia, Y.; Ho, W.; Li, M. Fabrication of TiO2 nanorod assembly grafted rGO (rGO@TiO2 -NR) hybridized flake-like photocatalyst. Appl. Surf. Sci. 2017, 391, 218–227. [Google Scholar] [CrossRef]
  33. Dai, Y.; Sun, Y.; Yao, J.; Ling, D.; Wang, Y.; Long, H.; Wang, X.; Lin, B.; Zeng, T.H.; Sun, Y. Graphene-wrapped TiO2 nanofibers with effective interfacial coupling as ultrafast electron transfer bridges in novel photoanodes. J. Mater. Chem. A 2014, 2, 1060–1067. [Google Scholar] [CrossRef]
  34. Fortes, A.D.; Suard, E.; Knight, K.S. Negative Linear Compressibility and Massive Anisotropic Thermal Expansion in Methanol Monohydrate. Science 2011, 331, 742–746. [Google Scholar] [CrossRef] [PubMed]
  35. Manga, K.K.; Zhou, Y.; Yan, Y.; Loh, K.P. Multilayer Hybrid Films Consisting of Alternating Graphene and Titania Nanosheets with Ultrafast Electron Transfer and Photoconversion Properties. Adv. Funct. Mater. 2009, 19, 3638–3643. [Google Scholar] [CrossRef]
  36. 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]
  37. Lin, Y.P.; Lin, S.Y.; Lee, Y.C.; Chen-Yang, Y.W. High surface area electrospun prickle-like hierarchical anatase TiO2 nanofibers for dye-sensitized solar cell photoanodes. J. Mater. Chem. A 2013, 1, 9875. [Google Scholar] [CrossRef]
  38. Li, J.; Ishigaki, T.; Sun, X. Anatase, Brookite, and Rutile Nanocrystals via Redox Reactions under Mild Hydrothermal Conditions: Phase-Selective Synthesis and Physicochemical Properties. J. Phys. Chem. C 2007, 111, 4969–4976. [Google Scholar] [CrossRef]
  39. Yang, J.; Heo, M.; Lee, H.J.; Park, S.; Kim, J.Y.; Shin, H.S. Reduced Graphene Oxide (rGO)-Wrapped Fullerene (C60) Wires. ACS Nano 2011, 5, 8365–8371. [Google Scholar] [CrossRef]
  40. Lui, G.; Liao, J.; 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. [Google Scholar] [CrossRef]
  41. 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]
  42. Lv, K.; Zuo, H.; Sun, J.; Deng, K.; Liu, S.; Li, X.; Wang, D. (Bi, C and N) codoped TiO2 nanoparticles. J. Hazard. Mater. 2009, 161, 396–401. [Google Scholar] [CrossRef] [PubMed]
  43. Zhang, Y.; Tang, Z.; Fu, X.; Xu, Y. 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]
  44. Bell, N.J.; Ng, Y.H.; Du, A.; Coster, H.; Smith, S.C.; Amal, R. Understanding the Enhancement in Photoelectrochemical Properties of Photocatalytically Prepared TiO2-Reduced Graphene Oxide Composite. J. Phys. Chem. C 2011, 115, 6004–6009. [Google Scholar] [CrossRef]
  45. Aluigi, A.; Rombaldoni, F.; Tonetti, C.; Jannoke, L. Study of Methylene Blue adsorption on keratin nanofibrous membranes. J. Hazard. Mater. 2014, 268, 156–165. [Google Scholar] [CrossRef]
Figure 1. The preparation of ITO-reduced graphene oxide (rGO) and titanium dioxide (TiO2) and photocatalytic processes.
Figure 1. The preparation of ITO-reduced graphene oxide (rGO) and titanium dioxide (TiO2) and photocatalytic processes.
Catalysts 09 00848 g001
Figure 2. X-ray diffractometry (XRD) patterns of (a) TiO2, rGO and TiO2, and ITO-rGO and TiO2, (b) ITO-rGO and TiO2 prepared at different times, and (c) ITO-rGO and TiO2 prepared in different diethylene glycol (DEG) concentrations.
Figure 2. X-ray diffractometry (XRD) patterns of (a) TiO2, rGO and TiO2, and ITO-rGO and TiO2, (b) ITO-rGO and TiO2 prepared at different times, and (c) ITO-rGO and TiO2 prepared in different diethylene glycol (DEG) concentrations.
Catalysts 09 00848 g002
Figure 3. Scanning electron microscope (SEM) images of (a) rGO, (b) TiO2, (c) ITO-rGO and TiO2 prepared in 1 h, (d) ITO-rGO and TiO2 prepared in 3 h, (e) ITO-rGO and TiO2 prepared in 5 h, (f) ITO-rGO and TiO2 prepared in 7 h, (g) ITO-rGO and TiO2 prepared in 9 h.
Figure 3. Scanning electron microscope (SEM) images of (a) rGO, (b) TiO2, (c) ITO-rGO and TiO2 prepared in 1 h, (d) ITO-rGO and TiO2 prepared in 3 h, (e) ITO-rGO and TiO2 prepared in 5 h, (f) ITO-rGO and TiO2 prepared in 7 h, (g) ITO-rGO and TiO2 prepared in 9 h.
Catalysts 09 00848 g003
Figure 4. SEM images of ITO-rGO and TiO2 prepared in different DEG concentrations: (a) 3 mol/L, (b) 5 mol/L, (c) 7 mol/L, and (d) 9 mol/L.
Figure 4. SEM images of ITO-rGO and TiO2 prepared in different DEG concentrations: (a) 3 mol/L, (b) 5 mol/L, (c) 7 mol/L, and (d) 9 mol/L.
Catalysts 09 00848 g004
Figure 5. Transmission electron microscope (TEM) images and magnification of (a,b) TiO2 and rGO and (c) TiO2, and high-resolution TEM (HRTEM) image of (d) TiO2 nanorods (insets are the corresponding SAED diffraction patterns of rGO and TiO2).
Figure 5. Transmission electron microscope (TEM) images and magnification of (a,b) TiO2 and rGO and (c) TiO2, and high-resolution TEM (HRTEM) image of (d) TiO2 nanorods (insets are the corresponding SAED diffraction patterns of rGO and TiO2).
Catalysts 09 00848 g005
Figure 6. X-ray photoelectron spectroscopy (XPS) of (a) XPS survey, (b) C1s, (c) Ti2p, and (d) O1s.
Figure 6. X-ray photoelectron spectroscopy (XPS) of (a) XPS survey, (b) C1s, (c) Ti2p, and (d) O1s.
Catalysts 09 00848 g006
Figure 7. Raman spectra of (a) TiO2 and (b) rGO.
Figure 7. Raman spectra of (a) TiO2 and (b) rGO.
Catalysts 09 00848 g007
Figure 8. (a) UV–vis diffuse reflectance spectra (DRS) of the pure TiO2, rGO and TiO2, and ITO-rGO and TiO2, and (b) curve of [F (R) hv]1/2 versus photon energy of pure TiO2, rGO and TiO2, and ITO-rGO and TiO2.
Figure 8. (a) UV–vis diffuse reflectance spectra (DRS) of the pure TiO2, rGO and TiO2, and ITO-rGO and TiO2, and (b) curve of [F (R) hv]1/2 versus photon energy of pure TiO2, rGO and TiO2, and ITO-rGO and TiO2.
Catalysts 09 00848 g008
Figure 9. Transient photocurrent density of TiO2, Nickel foam-rGO and TiO2, and ITO-rGO and TiO2.
Figure 9. Transient photocurrent density of TiO2, Nickel foam-rGO and TiO2, and ITO-rGO and TiO2.
Catalysts 09 00848 g009
Figure 10. (a) The adsorption on methylene blue (MB) solution with different initial concentrations (the time range of the inset is 0–30min), (b) adsorption capacity of MB on ITO-rGO and TiO2 in five successive cycles, and (c) effect of pH value on sample adsorption.
Figure 10. (a) The adsorption on methylene blue (MB) solution with different initial concentrations (the time range of the inset is 0–30min), (b) adsorption capacity of MB on ITO-rGO and TiO2 in five successive cycles, and (c) effect of pH value on sample adsorption.
Catalysts 09 00848 g010
Figure 11. Photocatalytic degradation of MB by pure TiO2, rGO and TiO2, and ITO-rGO and TiO2 (a,c), and the corresponding first-order rate constant (k) of catalysts in MB decoloring (b,d).
Figure 11. Photocatalytic degradation of MB by pure TiO2, rGO and TiO2, and ITO-rGO and TiO2 (a,c), and the corresponding first-order rate constant (k) of catalysts in MB decoloring (b,d).
Catalysts 09 00848 g011
Figure 12. Schematic mechanisms of MB decomposition through photodegradation by ITO-rGO and TiO2.
Figure 12. Schematic mechanisms of MB decomposition through photodegradation by ITO-rGO and TiO2.
Catalysts 09 00848 g012

Share and Cite

MDPI and ACS Style

Gong, C.; Xu, S.; Xiao, P.; Liu, F.; Xu, Y.; Yang, J.; Li, R.; Wang, X.; Du, J.; Peng, H. Indium Tin-Oxide Wrapped 3D rGO and TiO2 Composites: Development, Characterization, and Enhancing Photocatalytic Activity for Methylene Blue. Catalysts 2019, 9, 848. https://doi.org/10.3390/catal9100848

AMA Style

Gong C, Xu S, Xiao P, Liu F, Xu Y, Yang J, Li R, Wang X, Du J, Peng H. Indium Tin-Oxide Wrapped 3D rGO and TiO2 Composites: Development, Characterization, and Enhancing Photocatalytic Activity for Methylene Blue. Catalysts. 2019; 9(10):848. https://doi.org/10.3390/catal9100848

Chicago/Turabian Style

Gong, Cheng, Shiyin Xu, Peng Xiao, Feifan Liu, Yunhui Xu, Jin Yang, Rui Li, Xuan Wang, Jun Du, and Hailong Peng. 2019. "Indium Tin-Oxide Wrapped 3D rGO and TiO2 Composites: Development, Characterization, and Enhancing Photocatalytic Activity for Methylene Blue" Catalysts 9, no. 10: 848. https://doi.org/10.3390/catal9100848

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