Next Article in Journal
Role of Humic Acid Chemical Structure Derived from Different Biomass Feedstocks on Fe(III) Bioreduction Activity: Implication for Sustainable Use of Bioresources
Previous Article in Journal
Complete Hydrodesulfurization of Dibenzothiophene via Direct Desulfurization Pathway over Mesoporous TiO2-Supported NiMo Catalyst Incorporated with Potassium
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Photocatalytic Degradation of 2-Butanone Using Hybrid Nanostructures of Gallium Oxide and Reduced Graphene Oxide Under Ultraviolet-C Irradiation

1
Department of Materials Engineering, Korea Aerospace University, Goyang 10540, Korea
2
School of Electrical Engineering, KAIST, Daejeon 34141, Korea
*
Author to whom correspondence should be addressed.
Catalysts 2019, 9(5), 449; https://doi.org/10.3390/catal9050449
Submission received: 20 April 2019 / Revised: 9 May 2019 / Accepted: 10 May 2019 / Published: 15 May 2019
(This article belongs to the Section Environmental Catalysis)

Abstract

:
Hybrid nanostructures made of gallium oxide (Ga2O3) and reduced graphene oxide (rGO) are synthesized using a facile hydrothermal process method, where the Ga2O3 nanostructures are well dispersed on the rGO surface. The formed Ga2O3-rGO hybrids are characterized via Field emission scanning electron microscopy (FESEM), X-ray diffraction (XRD), Fourier transform infrared spectroscopy (FTIR), a diffuse reflectance Ultraviolet-visible-near infrared (UV-Vis-NIR) spectrophotometer, Brunauer–Emmett–Teller (BET), and photoluminescence (PL). Gas chromatography mass spectrometry (GC-MS) was used for analyzing volatile organic compounds (VOCs). The photocatalytic activity of the hybrid nanostructures is evaluated via the degradation of the 2-butanone, representing the VOCs under 254-nm radiation in the atmosphere. That activity is then compared to that of the Ga2O3 and commercial TiO2-P25. The Ga2O3-rGO hybrid shows enhanced photocatalytic degradation of 2-butanone compared to Ga2O3 and TiO2-P25, which is attributed to the enhanced specific surface area. The results indicate that the Ga2O3-rGO hybrid could be a promising method of enhancing photocatalytic activity and thereby effectively degrading VOCs, including the 2-butanone.

1. Introduction

Volatile organic compounds (VOCs) with high vapor pressures at room temperature are everywhere because they are emitted from a wide range of consumer products, building materials and adhesives. Many VOCs are toxic and adversely affect the human body in both indoor and outdoor environments [1,2,3]. As human beings spend more time at home and in the work place, the control of VOCs in indoor environments is a critical concern [4]. While absorption and incineration methods have traditionally been used to remove VOCs, photocatalytic oxidation (PCO) using metal oxide semiconductors has emerged as an attractive alternative due to their low cost and non-toxicity [5]. The photo-induced generation of electron-hole pairs promotes reduction and oxidation upon the surface of the semiconducting transition metal oxides [6,7]. Among various semiconductors, TiO2 is preferred in terms of its photocatalytic activity under UV light absorption [8]. It has been widely applied to photocatalysts due to its nontoxicity, low cost and high reactivity. Nevertheless, its long-term instability greatly limits the practical applications, due to the deactivation of the photocatalytic sites resulting from the accumulation of a stable intermediate on the TiO2 surface from the oxidation of aromatics in a dry atmosphere [9,10]. Recently, interest has increased in Ga2O3 (Eg = 4.8 eV) as a wide band gap semiconductor with its high redox ability and long-term stability, compared to commercial TiO2 (Eg = 3.2 eV), for the degradation of aromatic pollutants, such as the benzene series and toluene [11,12,13,14,15]. In addition to its strong redox ability, due to a higher conduction band (Ecb = −1.55 eV) and lower valence band (Evb = 3.25 eV) than TiO2 (Ecb = −0.5 eV, Evb = 2.7 eV) [16], Ga2O3 is composed of a valence band of O 2p orbitals and a dispersive conduction band of s and p orbitals from the p-block metal oxide atom, which promotes the mobility of photo-induced electrons in the Ga2O3 [13,17,18]. This improves the separation efficiency of the photo-induced electrons and holes, resulting in an increase in photocatalytic activity. Advanced strategies to enhance photocatalytic activity have been reported via the formation of a heterogeneous structure with noble metals, semiconductors and/or graphene allotropes [19,20,21,22]. Among these materials, the addition of graphene oxide has been demonstrated to be an effective method due to its large specific surface area and excellent charge carrier mobility [5,21,22,23,24,25,26]. However, to the best of our knowledge, few studies thus far have looked at the photocatalytic degradation of gas-phase contaminants for the purification of enclosed atmospheres with the use of Ga2O3-reduced graphene oxide (rGO) hybrids under ultraviolet C (UVC) irradiation. The present study focuses on the photocatalytic degradation of 2-butanone, which is used in adhesives and cleaning agents, as an indoor VOC source [27,28]. Herein, we report on the synthesis of the Ga2O3-rGO hybrids formed using the hydrothermal method, and the photocatalytic degradation of 2-butanone is discussed as compared to TiO2 and Ga2O3.

2. Results and Discussion

Figure 1a shows the Field emission scanning electron microscope (FESEM) images of the rGO, Ga2O3-rGO hybrids and Ga2O3. The results showed that the rGO flakes were wrinkled, partially folded and stacked. This was attributed to the weakening of the hydrogen bonding network between the flakes and the generation of various structural defects during the hydrothermal process [29,30]. The images of the Ga2O3-rGO hybrids showed that the Ga2O3 nanorods were well dispersed and uniformly anchored on the rGO. The chemical bonds between the Ga2O3 and rGO were discussed via the Fourier transform infrared spectra (FTIR) analysis, thereby promoting the charge transfer of photo-excited carriers from the Ga2O3 to rGO and eventually enhancing the photocatalytic activity. The magnified image (inset of Figure 1a) of the Ga2O3 revealed that the Ga2O3 possessed a mesoporous surface that could contribute to the improvement of photocatalytic performance [31,32]. Figure 1b shows the X-ray diffraction (XRD) patterns of the pristine GO, rGO, Ga2O3-rGO hybrids and Ga2O3 nanorods. The pristine GO showed a very strong intensity at 2θ = 10.55°, representing a typical inter-planar d-spacing (0.84 nm) of the GO stacks [33]. After the GO was reduced to rGO via the hydrothermal process, the peaks originating from the GO d-spacing diminished, but the broad peak at 2θ = 24° appeared, corresponding to the inter-planar d-spacing (0.37 nm) of the rGO [34]. The decreased d-spacing in the rGO compared to the GO was attributed to the removal of oxygen-containing functional groups, which also indicated the reestablishment of the sp2 carbon network in the rGO [35]. The results also revealed that the hydrothermal process was an effective method for the reduction of a GO without strong acidic conditions [25,36]. For Hybrid-2, the most noticeable peaks were indexed to the β-Ga2O3 with a broad rGO peak. As the ratio of Ga2O3 over rGO increased, the peaks representing Ga2O3 increased while the broad peak representing rGO diminished, as shown with Hybrid-1. These results showed that the Ga2O3 retained its crystal structure even after the hydrothermal process and dispersion on the rGO.
Figure 2 shows the FTIR spectra in the wavenumber range from 4000 to 650 cm−1, along with a schematic drawing representing the responsible chemical bonds of the pristine GO, rGO, Ga2O3-rGO hybrids and Ga2O3. The spectrum of the pristine GO showed typical transmittance peaks. The bands in the range of 1700 to 1000 cm−1 were associated with oxygen functional groups such as the carboxyl groups, hydroxyl groups and epoxy groups on both sides of the GO surface [24,36,37]. These bands of oxygen functional groups in the GO significantly decreased after the formation of rGO via the hydrothermal process [34,38]. This revealed that most oxygen functional groups had been removed from the GO surface while the GO was reduced [29]. These eliminated peaks associated with the oxygen functional group were recovered in the spectra of the Ga2O3-rGO hybrids. The presence of these oxygen functional groups indicated the formation of the chemical bonds between the Ga2O3 nanorods and the rGO, which allowed the carrier transfer from the Ga2O3 to rGO, and thereby enhanced the photocatalytic activity.
The Ultraviolet-visible-near infrared (UV-Vis-NIR) absorption of the Ga2O3 and hybrids were characterized in the wavelength range of 240 to 1000 nm, and the results are shown in Figure 3a. These results showed that the Ga2O3 was transparent in the visible light and ultraviolet C. As the ratio of the rGO over Ga2O3 increased, the overall absorption from the visible and near infrared (400–1000 nm) into the UV wavelength regions (260 nm ~) increased. In detail, a red shift in the absorption edge was observed in Hybrid-1 compared with Ga2O3, indicating a narrowing of the Ga2O3 band gap due to the coupling in the Ga2O3–rGO hybrids. The inset in Figure 3a shows the Tauc plot for the Ga2O3 and Ga2O3–rGO hybrids, where (ahv) 2 was plotted against the photo energy (hv), and the band gap energies were calculated from the extrapolated lines as shown. The optical energy band gap of the Ga2O3 and Hybrid-1 was calculated as 4.63 and 4.25 eV, respectively. However, it was difficult to calculate the exact optical energy band gap for Hybrid-2 because the large amount of rGO induced a very significant background absorption in the range of 400 to 800 nm. Also, while the absorption of Hybrid-1 around 254 nm was comparable to that of Ga2O3, the absorption of Hybrid-2 (a greater rGO ratio) was greatly reduced. This indicated that the photogenerated electrons from the Ga2O3 in Hybrid-1 were able to move along the rGO and then interact with the absorbed O2 to produce OH·, which facilitated the strong oxidative ability to degrade the organic pollutants. In this process, the rGO acted as an electron collector and transporter to effectively interfere with the electron-hole recombination and extend the lifetime of the photogenerated charge carriers formed by Ga2O3, leading to the higher photocatalytic efficiency in Hybrid-1. However, with the excessive addition of rGO (Hybrid-2), the rGO nanosheets blocked the UVC light from reaching the surface of the Ga2O3, resulting in the lower generation rate of the charge carriers. Figure 3b shows the PL emission spectra of the Ga2O3, Hybrid-1 and -2 when excited at 260 nm. A broad emission band ranging from 300 to 500 nm was observed in the Ga2O3 spectrum, while considerable PL quenching was observed in the hybrids with higher ratios of rGO over Ga2O3. The decreased intensity of the PL emission might have been due to the reduction in the recombination process of electrons and holes because of the charge carrier transfer from the excited Ga2O3 to the rGO [23,39,40]. To further discuss the charge carrier transfer from the excited Ga2O3 to the rGO, the PL decay characteristics of the Ga2O3 and hybrids were investigated, as shown in the inset of Figure 3b. The results were fitted to multi-exponential curves to derive the PL decay parameters that were obtained using Equations (1) and (2) [41]. The lifetime, pre-exponential factors and average lifetime of the Ga2O3 and hybrids are summarized in Table 1.
f ( t ) = A 1 e t τ 1 + A 2 e t τ 2 + A 3 e t τ 3 + A 4 e t τ 4
τ = i = 1 n A i τ i 2 A i τ i
where A i denotes the pre-exponential factor, τ i is the lifetime decay constant, and τ is the average lifetime.
For the Ga2O3, the fast and slow component of the PL decay had a lifetime of 2.5 ns and 2876 ns, respectively. As the ratio of the rGO over Ga2O3 increased, the fast lifetime and average lifetime decreased, but the slow lifetime increased. The analysis revealed that the charge carrier transfer between the excited Ga2O3 and rGO was dominated by the fast component. Furthermore, from the fast decay component, the electron-transfer rate could be estimated using Equation (3).
k e t = 1 τ ( G a 2 O 3 r G O ) 1 τ ( G a 2 O 3 )
The electron-transfer rate of Hybrid-1 was 0.017 × 10 9   s 1 , while that of Hybride-2 was 0.225 × 10 9   s 1 . This showed that the electron-transfer rate of Hybrid-2 was 13.5 times higher than that of Hybrid-1, while the number of electron-hole pairs was lower in Hybrid-2. Based on this analysis, it was concluded that the chemical bonds between the Ga2O3 and rGO were very effective in transferring the photo-excited electrons before the recombination process.
The identification of the adhesive component was measured with a GC-MS analysis, which was identified by comparing retention times and mass spectra with library spectra. Figure 4a shows that air, water and 2-butanone were detected before the photodegradation. After UVC irradiation of the 2-butanone for 30 min using a Ga2O3-rGO hybrid (Figure 4b), it was found that air and water were the main by-products. Figure 4c shows the degradation of 2-butanone as a function of exposure time under the 254 nm UVC irradiation time, at intervals of 30 min when different catalysts were used. The TiO2-P25, which is known to be a good photocatalyst, was also compared with its counterparts under the identical reaction conditions. With conventional TiO2-25, the 2-butanone degraded to 62% within 1 h. On the other hand, the 2-butanone degraded to 93% within 1 h when Ga2O3 was used as a catalyst. The hybrids showed a much-enhanced photodegradation efficiency compared to the Ga2O3 and TiO2. The 2-butanone degraded to 99.9% and 98.2% under the identical conditions when Hybrid-1 and Hybrid-2 were used, respectively. The additional rGO in the Ga2O3 played a role as a transfer pathway of the photo-generated carriers. In addition, the rGO provided a much higher specific surface area, improving the photocatalytic performance of the Ga2O3. It was found that Hybrid-1 had a higher specific surface area (SBET = 149 m2·g−1) than the pure Ga2O3 (SBET = 3.4 m2·g−1) and Hybrid-2 (SBET = 36 m2·g−1), using the Brunauer–Emmett–Teller (BET) surface area measurement [42,43]. While the additional rGO in the Ga2O3 improved the photocatalytic performance via the reasons above mentioned, it adversely degraded the photocatalytic performance due to the lower UVC absorption blocked by excessive rGO [20,34], which could be observed in Figure 3a. This explained the experimental results; the degradation performance of different catalysts was in the order of Hybrid-1 (1 wt. % of rGO) > Hybrid-2 (5 wt. % of rGO) > Ga2O3 > TiO2-P25. In addition to the photocatalytic activity, the stability and repeatability of the 2-butanone degradation was investigated using Hybrid-1, with the results shown in Figure 4d. The results showed that Hybrid-1 retained its photocatalytic activity even after prolonged operation.
Figure 5 shows the optical transmittance of Hybrid-1 before and after 2-butanone degradation. The transmittance decreased after 2-butanone degradation particularly around 1710 cm−1 and 1560 cm−1, representing C=O bonding [12] and C=C bonding [44], respectively. This indicated that the decomposed residue of the 2-butanone remained on the Hybrid-1 surface. These residual byproducts on the surface need to be investigated further.

3. Materials and Methods

3.1. Synthesis of Hybrids

The GO was prepared using a modified Hummers’ method [45], and the Ga2O3 powder (particle diameter 2.5–6 μm, 99.999%) was purchased from Taewon Scientific Co., Ltd. (Republic of Korea). Ga2O3-rGO hybrids were synthesized via a conventional hydrothermal method. First, a GO solution in which the GO (lateral size > 3 μm, 99%) was dispersed in distilled water (DI water, 50 mL) was ultrasonicated for 3 h to obtain uniform dispersions of exfoliated GO. Next, Ga2O3 powders were added into the prepared GO solutions. After being vigorously stirred for 2 h, the mixture of Ga2O3 and GO solutions was transferred to a Teflon-lined autoclave for the hydrothermal process. The mixed solution was heated in an oven at 150 °C for 6 h (6 °C min−1) and cooled at room temperature. The unreacted solvent and the resulting precipitate were separated by centrifugal force at 4000 rpm for 30 min. The collected precipitate was washed and rinsed with DI water and ethanol several times and then dried at 60 °C for 12 h (6 °C min−1). Finally, two Ga2O3-rGO hybrids with 1 wt. % and 5 wt. % of rGO were formed, which were called Hybrid-1 and Hybrid-2, respectively. It was noted that the GO was successfully converted into rGO without strong acidic conditions via the hydrothermal process.

3.2. Materials Characterization

The microstructures of the synthesized Ga2O3-rGO hybrids were observed using a field emission scanning electron microscope (FESEM, JEOL JSM-7100F, Japan) at 15 kV acceleration voltage; the samples were distributed on the sticky-carbon tape followed by a Pt coating. The crystallographic structures of the synthesized hybrids were determined using a powder X-ray diffraction system (XRD, Rigaku Ultima IV XRD, Japan) over the 2θ range of 5–80° with Cu Kα radiation (λ = 0.15405 nm). Middle-infrared (4000–650 cm−1) transmission spectra were collected with a Fourier-transform infrared spectra (FTIR, Agilent Cary 670, U.S.A.) spectrometer with an ATR accessory (ZnSe and Ze crystal). UV-Vis-NIR spectroscopy (DRS, Shimadzu SolidSpec-3700, Japan) was measured in the range from 240 to 1000 nm using BaSO4 as a reference. Photoluminescence (PL) spectra were recorded under 260 nm UV laser excitation (PL, HORIBA LabRAM HR-800, Japan). Furthermore, Time-resolved PL measurements were performed on a HORIBA Fluorolog-3 via a time-correlated single photon counting (TCSPC) spectrometer equipped with a Pulsed Nano LED 260-nm excitation source. The specific surface area of the product was measured using a nitrogen adsorption-desorption isotherm on a BELSORP-max instrument and calculated via the Brunauer–Emmett–Teller (BET) method. Prior to measurement, the samples were degassed at 473 K for 4 h.

3.3. Photocatalytic Activity

The photocatalytic performance of the Ga2O3-rGO hybrids with different rGO wt. % were evaluated via the photodegradation of 2-butanone using a commercial VOC air detector (KHALDER: KD-001) with TVOC values ranging between 0.00 to 9.99 mg/m3. For every experiment, 200 mg of the photocatalyst powder and 18.2 mg of a commercial adhesive (as a 2-butanone source) were used with a UVC source (λ = 254 nm, 4W, irradiance of 340 μW·cm−2 with a distance of 15 cm, Model: LF-104.S). No noticeable degradation of the 2-butanone was observed in the absence of a photocatalyst under the UVC irradiation. In the presence of a photocatalyst under UVC exposure, the photodegradation efficiency was calculated using Equation (4):
Degradation efficiency (%) = (C0 − C)/C0 × 100
where C0 is the initial concentration and C is the final concentration of the 2-butanone. Changes in the 2-butanone concentration were measured at intervals of 30 min. For comparison, photodegradation efficiency of Ga2O3 and TiO2-P25 was also carried out. To further explore the long-term stability of the Ga2O3-rGO hybrids, repeated recycling experiments were carried out using Hybrid-1 under identical conditions. The gaseous products generated before and after the photocatalytic degradation were collected and analyzed with a gas chromatograph (GC, Agilent 5975C, U.S.A.) equipped with a HP-5 MS column (with dimensions of 60 m × 0.25 μm × 0.25 mm). A product sample was injected into the GC column with a split ratio 20:1. Helium was used as the carrier gas at a flow rate of 1.0 mL·min−1. The column temperature program was as follows: 50 °C for 10 min, increasing at a rate of 10 °C min−1 to 300 °C. The product identification was confirmed with a gas chromatograph-mass spectrometer. Major reaction products were identified by comparing retention times and mass spectra with reference compounds.

4. Conclusions

Ga2O3-rGO hybrids were synthesized using a facile hydrothermal process method without using any additives or surfactants. The Ga2O3 nanorods were well dispersed on the surface of the rGO nanosheet through strong coupling between the Ga2O3 and rGO. Compared with the pure Ga2O3 and conventional TiO2-P25, the photocatalytic activity of the hybrids was significantly enhanced. The enhanced photocatalytic activity of the hybrids was attributed to the effective carrier transfer from the Ga2O3 to rGO and enhanced specified surface area. The best photocatalytic activity was observed with Hybrid-1 (1 wt. % rGO) instead of Hybrid-2 (5 wt. % rGO), indicating that the presence of rGO with Ga2O3 can also degrade the photocatalytic performance via the lower absorption of UVC light blocked by the excessive rGO. The Ga2O3-rGO hybrid offers a promising method of enhancing photocatalytic activity and thereby effectively degrading 2-butanone.

Author Contributions

Conceptualization, B.J.C., W.S.H.; formal analysis, H.J.B., T.H.Y., S.K., W.C.; methodology and resources, Y.S.S., D.-K.K., B.J.C., W.S.H.; writing, review and editing, H.J.B., T.H.Y., S.K., W.C., Y.S.S., D.-K.K., B.J.C., W.S.H.

Funding

This work was funded by the National Research Foundation of Korea (NRF) through the Basic Science Research Program (2017R1A2B4012278), and by the Center for Advanced Soft-Electronics that is funded by the Ministry of Science, ICT and Future Planning, through the Global Frontier Project (CASE-2011-0031638).

Acknowledgments

Gas chromatography mass spectrometry (GC-MS) analysis was conducted at Korea Basic Science Institute (Busan Center).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dolai, S.; Bhunia, S.K.; Beglaryan, S.S.; Kolusheva, S.; Zeiri, L.; Jelinek, R. Colorimetric Polydiacetylene−Aerogel Detector for Volatile Organic Compounds (VOCs). ACS Appl. Mater. Interfaces 2017, 9, 2891–2898. [Google Scholar] [CrossRef] [PubMed]
  2. Konvalina, G.; Haick, H. Sensors for Breath Testing: From Nanomaterials to Comprehensive Disease Detection. Acc. Chem. Res. 2014, 47, 66–76. [Google Scholar] [CrossRef] [PubMed]
  3. Suh, J.M.; Sohn, W.B.; Shim, Y.S.; Choi, J.S.; Song, Y.G.; Kim, T.M.; Jeon, J.M.; Kwon, K.C.; Choi, K.S.; Kang, C.Y.; et al. p−p Heterojunction of Nickel Oxide-Decorated Cobalt Oxide Nanorods for Enhanced Sensitivity and Selectivity toward Volatile Organic Compounds. ACS Appl. Mater. Interfaces 2018, 10, 1050–1058. [Google Scholar] [CrossRef] [PubMed]
  4. McCurdy, T.; Glen, G.; Smith, L.; Lakkadi, Y. The National Exposure Research Laboratory’s Consolidated Human Activity Database. J. Expo. Anal. Environ. Epidemiol. 2000, 10, 566–578. [Google Scholar] [CrossRef]
  5. Kim, S.D.; Han, K.I.; Lee, I.G.; Park, W.K.; Yoon, Y.J.; Yoo, C.S.; Yang, W.S.; Hwang, W.S. A Gallium Oxide-Graphene Oxide Hybrid Composite for Enhanced Photocatalytic Reaction. Nanomaterials 2016, 6, 127. [Google Scholar] [CrossRef]
  6. Devahasdin, S.; Fan, C., Jr.; Li, K.; Chen, D.H. TiO2 photocatalytic oxidation of nitric oxide: Transient behavior and reaction kinetics. J. Photochem. Photobiol. A 2003, 156, 161170. [Google Scholar] [CrossRef]
  7. Kamal, M.S.; Razzak, S.A.; Hossain, M.M. Catalytic oxidation of volatile organic compounds (VOCs)—A review. Atmos. Environ. 2016, 140, 117–134. [Google Scholar] [CrossRef]
  8. Tatsuma, T.; Saitoh, S.; Ngaotrakanwiwat, P.; Ohko, Y.; Fujishima, A. Energy Storage of TiO2-WO3 Photocatalysis Systems in the Gas Phase. Langmuir 2002, 18, 7777–7779. [Google Scholar] [CrossRef]
  9. Martra, G.; Coluccia, S.; Marchese, L.; Augugliaro, V.; Loddo, V.; Palmisano, L.; Schiavello, M. The Role of H2O in the Photocatalytic Photocatalytic Oxidation of Toluene in Vapour Phase on Anatase TiO2 Catalyst: A FTIR Study. Catal. Today 1999, 53, 695–702. [Google Scholar] [CrossRef]
  10. Mendez-Roman, R.; Cardona-Martinez, N. Relationship between the Formation of Surface Species and Catalyst Deactivation during the Gas-Phase Photocatalytic Oxidation of Toluene. Catal. Today 1998, 40, 353–365. [Google Scholar] [CrossRef]
  11. Girija, K.; Thirumalairajan, S.; Patra, A.K.; Mangalaraj, D.; Ponpandian, N.; Viswanathan, C. Enhanced photocatalytic performance of novel self-assembled floral β-Ga2O3 nanorods. Curr. Appl. Phys. 2013, 13, 652–658. [Google Scholar] [CrossRef]
  12. Hou, Y.; Wang, X.; Wu, L.; Ding, Z.; Fu, X. Efficient Decomposition of Benzene over a β-Ga2O3 Photocatalyst under Ambient Conditions. Environ. Sci. Technol. 2006, 40, 5799–5803. [Google Scholar] [CrossRef] [PubMed]
  13. Hou, Y.; Wu, L.; Wang, X.; Ding, Z.; Li, Z.; Fu, X. Photocatalytic performance of α-, β-, and γ -Ga2O3 for the destruction of volatile aromatic pollutants in air. J. Catal. 2007, 250, 12–18. [Google Scholar] [CrossRef]
  14. Hou, Y.; Zhang, J.S.; Ding, Z.X.; Wu, L. Synthesis, characterization and photocatalytic activity of β-Ga2O3 nanostructures. Powder Technol. 2010, 203, 440–446. [Google Scholar] [CrossRef]
  15. Sun, M.; Li, D.; Zhang, W.; Fu, X.; Shao, Y.; Li, W.; Xiao, G.; He, Y. Rapid microwave hydrothermal synthesis of GaOOH nanorods with photocatalytic activity toward aromatic compounds. Nanotechnology 2010, 21, 355601. [Google Scholar] [CrossRef] [PubMed]
  16. Liu, Q.; Yu, Z.; Li, M.; Hou, Y.; Sun, L.; Wang, L.; Peng, Z.; Chen, D.; Liu, Y. Fabrication of Ag/AgBr/Ga2O3 heterojunction composite with efficient photocatalytic activity. Mol. Catal. 2017, 432, 57–63. [Google Scholar] [CrossRef]
  17. Ikarashi, K.; Sato, J.; Kobayashi, H.; Saito, N.; Nishiyama, H.; Inoue, Y. Photocatalysis for Water Decomposition by RuO2-Dispersed ZnGa2O4 with d10 Configuration. J. Phys. Chem. B 2002, 106, 9048–9053. [Google Scholar] [CrossRef]
  18. Sato, J.; Kobayashi, H.; Ikarashi, K.; Saito, N.; Nishiyama, H.; Inoue, Y. Photocatalytic Activity for Water Decomposition of RuO2-Dispersed Zn2GeO4 with d10 Configuration. J. Phys. Chem. B 2004, 108, 4369–4375. [Google Scholar] [CrossRef]
  19. Ge, M.Z.; Cao, C.Y.; Li, S.H.; Tang, Y.X.; Wang, L.N.; Qi, N.; Huang, J.Y.; Zhang, K.Q.; Al-Deyab, S.S.; Lai, Y.K. In situ plasmonic Ag nanoparticle anchored TiO2 nanotube arrays as visible-light-driven photocatalysts for enhanced water splitting. Nanoscale 2016, 8, 5226–5234. [Google Scholar] [CrossRef] [PubMed]
  20. Shengyan, P.; Rongxin, Z.; Hui, M.; Daili, D.; Xiangjun, P.; Fei, Q.; Wei, C. Facile in-situ design strategy to disperse TiO2 nanoparticles on graphene for the enhanced photocatalytic degradation of rhodamine 6G. Appl. Catal. B 2017, 218, 208–219. [Google Scholar] [CrossRef]
  21. Singh, A.; Sinha, A.S.K. Active CdS/rGO photocatalyst by a high temperature gas-solid reaction for hydrogen production by splitting of water. Appl. Surf. Sci. 2018, 430, 184–197. [Google Scholar] [CrossRef]
  22. 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] [Green Version]
  23. Pan, X.; Yi, Z. Graphene Oxide Regulated Tin Oxide Nanostructures: Engineering Composition, Morphology, Band Structure, and Photocatalytic Properties. ACS Appl. Mater. Interfaces 2015, 7, 27167–27175. [Google Scholar] [CrossRef]
  24. 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]
  25. Wang, P.; Wang, J.; Wang, X.; Yu, H.; Yu, J.; Lei, M.; Wang, Y. One-step synthesis of easy-recycling TiO2-rGO nanocomposite photocatalysts with enhanced photocatalytic activity. Appl. Catal. B 2013, 132–133, 452–459. [Google Scholar] [CrossRef]
  26. Zeng, X.; Wang, Z.; Wang, G.; Gengenbach, T.R.; McCarthy, D.T.; Deletic, A.; Yu, J.; Zhang, X. Highly dispersed TiO2 nanocrystals and WO3 nanorods on reduced graphene oxide: Z-scheme photocatalysis system for accelerated photocatalytic water disinfection. Appl. Catal. B 2017, 218, 163–173. [Google Scholar] [CrossRef]
  27. Koss, A.R.; Gouw, J.D.; Warneke, C.; Gilman, J.B.; Lerner, B.M.; Graus, M.; Yuan, B.; Edwards, P.; Brown, S.S.; Wild, R.; et al. Photochemical aging of volatile organic compounds associated with oil and natural gas extraction in the Uintah Basin, UT, during a wintertime ozone formation event. Atmos. Chem. Phys. 2015, 15, 5727–5741. [Google Scholar] [CrossRef] [Green Version]
  28. Torpy, F.; Clements, N.; Pollinger, M.; Dengel, A.; Mulvihill, I.; He, C.; Irga, P. Testing the single-pass VOC removal efficiency of an active green wall using methyl ethyl ketone (MEK). Air Qual. Atmos. Health 2018, 11, 163–170. [Google Scholar] [CrossRef] [PubMed]
  29. Mungse, H.P.; Sharma, O.P.; Sugimura, H.; Khatri, O.P. Hydrothermal deoxygenation of graphene oxide in sub- and supercritical water. RSC Adv. 2014, 4, 22589–22595. [Google Scholar] [CrossRef]
  30. Zhou, Y.; Bao, Q.; Tang LA, L.; Zhong, Y.; Loh, K.P. Hydrothermal Dehydration for the “Green” Reduction of Exfoliated Graphene Oxide to Graphene and Demonstration of Tunable Optical Limiting Properties. Chem. Mater. 2009, 21, 2950–2956. [Google Scholar] [CrossRef]
  31. Li, D.; Duan, X.; Qin, Q.; Fan, H.; Zheng, W. Ionic liquid-assisted synthesis of mesoporous α-Ga2O3 hierarchical structures with enhanced photocatalytic activity. J. Mater. Chem. A 2013, 1, 12417–12421. [Google Scholar] [CrossRef]
  32. Zhao, W.; Yang, Y.; Hao, R.; Liu, F.; Wang, Y.; Tan, M.; Tang, J.; Ren, D.; Zhao, D. Synthesis of mesoporous β-Ga2O3 nanorods using PEG as template: Preparation, characterization and photocatalytic properties. J. Hazard. Mater. 2011, 192, 1548–1554. [Google Scholar] [CrossRef] [PubMed]
  33. Shen, J.F.; 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. Hu, K.; Xie, X.; Szkopek, T.; Cerruti, M. Understanding Hydrothermally Reduced Graphene Oxide Hydrogels: From Reaction Products to Hydrogel Properties. Chem. Mater. 2016, 28, 1756–1768. [Google Scholar] [CrossRef]
  35. Xu, Y.; Sheng, K.; Li, C.; Shi, G. Self-Assembled Graphene Hydrogel via a One-Step Hydrothermal Process. ACS Nano 2010, 4, 4324–4330. [Google Scholar] [CrossRef]
  36. Bai, X.; Wang, L.; Zong, R.; Lv, Y.; Sun, Y.; Zhu, Y. Performance Enhancement of ZnO Photocatalyst via Synergic Effect of Surface Oxygen Defect and Graphene Hybridization. Langmuir 2013, 29, 3097–3105. [Google Scholar] [CrossRef] [PubMed]
  37. Li, Q.; Guo, B.; Yu, J.; Ran, J.; Zhang, B.; Yan, H.; Gong, J.R. Highly Efficient Visible-Light-Driven Photocatalytic Hydrogen Production of CdS-Cluster-Decorated Graphene Nanosheets. J. Am. Chem. Soc. 2011, 133, 10878–10884. [Google Scholar] [CrossRef]
  38. Soltani, T.; Lee, B.K. A benign ultrasonic route to reduced graphene oxide from pristine graphite. J. Colloid Interface Sci. 2017, 486, 337–343. [Google Scholar] [CrossRef]
  39. Gollu, S.R.; Sharma, R.; Srinivas, G.; Kundu, S.; Gupta, D. Incorporation of silver and gold nanostructures for performance improvement in P3HT: PCBM inverted solar cell with rGO/ZnO nanocomposite as an electron transport layer. Org. Electron. 2016, 29, 79–87. [Google Scholar] [CrossRef]
  40. Xu, T.; Zhang, L.; Cheng, H.; Zhu, Y. Significantly enhanced photocatalytic performance of ZnO via graphene hybridization and the mechanism study. Appl. Catal. B 2011, 101, 382–387. [Google Scholar] [CrossRef]
  41. Williams, G.; Kamat, P.V. Graphene-Semiconductor Nanocomposites: Excited-State Interactions between ZnO Nanoparticles and Graphene Oxide. Langmuir 2009, 25, 13869–13873. [Google Scholar] [CrossRef]
  42. Wu, F.; Wang, X.; Li, M.; Xu, H. A high capacity NiFe2O4/RGO nanocomposites as superior anode materials for sodium-ion batteries. Ceram. Int. 2016, 42, 16666–16670. [Google Scholar] [CrossRef]
  43. Khatamian, M.; Khodakarampoor, N.; Oskoui, M.S.; Kazemian, N. Synthesis and characterization of RGO/zeolite composites for the removal of arsenic from contaminated water. RSC Adv. 2015, 5, 35352–35360. [Google Scholar] [CrossRef]
  44. Li, J.; Xing, X.; Li, J.; Shi, M.; Lin, A.; Xu, C.; Zheng, J.; Li, R. Preparation of thiol-functionalized activated carbon from sewage sludge with coal blending for heavy metal removal from contaminated water. Environ. Pollut. 2018, 234, 677–683. [Google Scholar] [CrossRef] [PubMed]
  45. Hummers, W.S., Jr.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
Figure 1. (a) FESEM images of rGO, Hybrid-2, Hybrid-1 and Ga2O3 (the scale bar is 3 μm); the scale bar of the inset is 1 μm), and (b) XRD patterns of the pristine GO, rGO, Hybrid-2, Hybrid-1 and Ga2O3 with the reference from Joint Committee on Powder Diffraction Standards (JCPDS) card no. 760573 (monoclinic Ga2O3 phase).
Figure 1. (a) FESEM images of rGO, Hybrid-2, Hybrid-1 and Ga2O3 (the scale bar is 3 μm); the scale bar of the inset is 1 μm), and (b) XRD patterns of the pristine GO, rGO, Hybrid-2, Hybrid-1 and Ga2O3 with the reference from Joint Committee on Powder Diffraction Standards (JCPDS) card no. 760573 (monoclinic Ga2O3 phase).
Catalysts 09 00449 g001
Figure 2. FTIR spectra of pristine GO, rGO, Hybrid-2, Hybrid-1 and Ga2O3, along with schematic chemical structure of GO, rGO and Ga2O3-rGO hybrids.
Figure 2. FTIR spectra of pristine GO, rGO, Hybrid-2, Hybrid-1 and Ga2O3, along with schematic chemical structure of GO, rGO and Ga2O3-rGO hybrids.
Catalysts 09 00449 g002
Figure 3. (a) UV-Vis-NIR spectroscopy and (b) PL spectroscopy of Ga2O3 and Ga2O3-rGO hybrids at room temperature. The inset in Figure 3a shows a Tauc plot, which is one method of extracting an optical energy bandgap in semiconductors. The inset in Figure 3b shows the normalized PL decay of the Ga2O3 and Ga2O3-rGO hybrids, which was measured at the emission wavelengths of 455 nm and 459 nm, respectively.
Figure 3. (a) UV-Vis-NIR spectroscopy and (b) PL spectroscopy of Ga2O3 and Ga2O3-rGO hybrids at room temperature. The inset in Figure 3a shows a Tauc plot, which is one method of extracting an optical energy bandgap in semiconductors. The inset in Figure 3b shows the normalized PL decay of the Ga2O3 and Ga2O3-rGO hybrids, which was measured at the emission wavelengths of 455 nm and 459 nm, respectively.
Catalysts 09 00449 g003
Figure 4. Gas chromatography graph (a) before and (b) after photodegradation under 254 nm UVC irradiation for 30 min. (The inset in (a) is the mass spectrum of 2-butanone, and the inset in (b) is the mass spectra of H2O and air), (c) photodegradation of 2-butanone under 254 nm UVC irradiation with different catalysts such as TiO2-P25, Ga2O3 and Ga2O3-rGO hybrids. TVOC refers to the total volatile organic compounds and a group of a wide range of organic chemical compounds. The guideline below 0.3 mg/m3 provides practical guidance to manage health and safety risks, which are regulated under Health and Safety, and subject to change (inset: Schematic drawing of the photocatalytic reactor), and (d) repeatability test of photocatalytic degradation of Hybrid-1 under 254 nm UVC irradiation at intervals of 15 min.
Figure 4. Gas chromatography graph (a) before and (b) after photodegradation under 254 nm UVC irradiation for 30 min. (The inset in (a) is the mass spectrum of 2-butanone, and the inset in (b) is the mass spectra of H2O and air), (c) photodegradation of 2-butanone under 254 nm UVC irradiation with different catalysts such as TiO2-P25, Ga2O3 and Ga2O3-rGO hybrids. TVOC refers to the total volatile organic compounds and a group of a wide range of organic chemical compounds. The guideline below 0.3 mg/m3 provides practical guidance to manage health and safety risks, which are regulated under Health and Safety, and subject to change (inset: Schematic drawing of the photocatalytic reactor), and (d) repeatability test of photocatalytic degradation of Hybrid-1 under 254 nm UVC irradiation at intervals of 15 min.
Catalysts 09 00449 g004
Figure 5. Optical transmittance of fresh Hybrid-1 and used Hybrid-1 as a function of wavelength.
Figure 5. Optical transmittance of fresh Hybrid-1 and used Hybrid-1 as a function of wavelength.
Catalysts 09 00449 g005
Table 1. Summary of the fitting parameters for the PL decay components of the Ga2O3 and hybrids.
Table 1. Summary of the fitting parameters for the PL decay components of the Ga2O3 and hybrids.
A1τ1 (ns)A2τ2 (ns)A3τ3 (ns)A4τ4 (ns) τ   ( ns )
Ga2O389732.50 603110 201287668.3
Hybrid-195382.40 236126 81308130.7
Hybrid-295971.618643.659419 49417125.7

Share and Cite

MDPI and ACS Style

Bae, H.J.; Yoo, T.H.; Kim, S.; Choi, W.; Song, Y.S.; Kwon, D.-K.; Cho, B.J.; Hwang, W.S. Enhanced Photocatalytic Degradation of 2-Butanone Using Hybrid Nanostructures of Gallium Oxide and Reduced Graphene Oxide Under Ultraviolet-C Irradiation. Catalysts 2019, 9, 449. https://doi.org/10.3390/catal9050449

AMA Style

Bae HJ, Yoo TH, Kim S, Choi W, Song YS, Kwon D-K, Cho BJ, Hwang WS. Enhanced Photocatalytic Degradation of 2-Butanone Using Hybrid Nanostructures of Gallium Oxide and Reduced Graphene Oxide Under Ultraviolet-C Irradiation. Catalysts. 2019; 9(5):449. https://doi.org/10.3390/catal9050449

Chicago/Turabian Style

Bae, Hyun Jeong, Tae Hee Yoo, Seungdu Kim, Wonhyeok Choi, Yo Seung Song, Do-Kyun Kwon, Byung Jin Cho, and Wan Sik Hwang. 2019. "Enhanced Photocatalytic Degradation of 2-Butanone Using Hybrid Nanostructures of Gallium Oxide and Reduced Graphene Oxide Under Ultraviolet-C Irradiation" Catalysts 9, no. 5: 449. https://doi.org/10.3390/catal9050449

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