DOI:
10.1039/C5RA20314G
(Paper)
RSC Adv., 2015,
5, 106644-106650
Controlled synthesis of Zn(1−1.5x)FexS nanoparticles via a microwave route and their photocatalytic properties†
Received
1st October 2015
, Accepted 10th December 2015
First published on 11th December 2015
Abstract
Fe-doped ZnS photocatalysts synthesized by a conventional hydrothermal method usually have poor crystallinity and low photocatalytic activity. In this study, the Zn(1−1.5x)FexS particles were first directly synthesized by the microwave irradiation method without additional heat treatments. The prepared Zn(1−1.5x)FexS catalysts were characterized using X-ray diffraction (XRD), UV-Vis absorption spectra, scanning electron microscopy (SEM), and Brunauer–Emmett–Teller (BET) analyzer, etc. The characterization results show that the morphology and physico-chemical properties of samples are changed depending on the ratio of Fe and Zn. The absorption edges of Zn(1−1.5x)FexS were red-shifted as the value of x decreased. The band gaps were estimated to be from 2.74 to 3.64 eV from the onset of the UV-Vis absorption edges. The results indicated that the photocatalyst of Zn0.97Fe0.02S has the highest photocatalytic activity of dimethyl phthalate (DMP) with a removal of 97.5%. In addition, the crystallite size, band gap and structure for the Zn(1−1.5x)FexS samples have a strong influence on the degradation of DMP from wastewater.
1. Introduction
DMP is one of the most important endocrine disruptors which are a major class of environmental contaminants and widely used as plasticizers to improve the flexibility and durability of consumer products, food packaging materials, and polyvinyl chloride plastics.1,2 Nowadays, DMP are suspected to be mutagens, endocrine disruptors and hepatotoxic agents, and can accumulate in the human body, and might pose a risk to human health.3,4 Due to their widespread presence, persistence, and difficulty in degradation, the United States Environmental Protection Agency has listed DMP as a priority pollutant.5,6 As a result, different treatment methods for DMP-contaminated water have attracted research attention.
Over the past decades, various chemical, physical and biological techniques, for the removal of DMP have been developed, including advanced oxidation processes, constructed wetlands, anaerobic degradation, microwave and photocatalysis.1,6–9 Recently, a great deal of attention has been paid to the photodegradation of DMP by inorganic materials, especially ZnS-based semiconductors.10–12 There are many preparation methods for Fe-doped ZnS, such as hydrothermal,13,14 microemulsion,15 and chemical co-precipitation method.16–19 Contrasted with the above-mentioned methods, microwave irradiation method is a environmentally friendly and economical method to synthesize materials,20,21 probably due to its outstanding advantages of fast reaction speed and clean reaction way.22,23 Consequently, some photocatalytic materials such as β-Ga2O3 (ref. 24) and ZnS21 have been synthesized using microwave method instead of the conventional methods.
In the study, Fe-doped ZnS materials with different Fe concentration were synthesized through microwave irradiation. The physical properties, crystal structure and photocatalytic performance of the prepared nanoparticles were manipulated. Particularly, we analyze the photocatalyst band structures under different x values according to characterization results and experimental data.
2. Experimental
2.1. Chemicals
The starting materials for the synthesis of Zn(1−1.5x)FexS nanoparticles were Zn(NO3)2·6H2O and Fe(NO3)3·9H2O as zinc and iron sources, respectively. Thioacetamide (TAA) was applied as a sulfur source, and 2-mercaptoethanol (2-hydroxyethanthiol, HOCH2CH2SH) was applied as a stabilizing agent. All chemicals were of the highest purity available and used without further purification. All solutions were prepared using deionized (DI) water at room temperature.
2.2. Synthesis
Microwave irradiation hydrothermal synthesis procedure was as follows. Briefly, appropriate amounts of Zn(NO3)2·6H2O and Fe(NO3)3·9H2O (totaling to 10 mmol) were dissolved in 100 mL of deionized (DI) water using a mechanical stirrer, in which specific amounts of the two compounds were calculated in Table S1 in ESI.† And then 10 mmol of TAA and 1.0 mL of 2-mercaptoethanol were into the above-mentioned solution, stirring and obtaining a mixture. The flask was moved into the microwave reactor (COOLPEX-E with output power 1200 W, purchased from PreeKem Scientific Instruments Co., Ltd., China). Fig. 1 shows the picture of the microwave reactor. As the volume of the mixed solution was less than 100 mL, all experiments were conducted at PMW = 500 W according to the instruction of MW reactor. The reaction was carried out at 388 K for 30 min in the MW reactor with a reflux condenser and a mechanical stirrer. After it was finished, the mixture was cooled till room temperature. The precipitated nanoparticles were separated in a centrifuge (at 7000 rpm) within 15 min, and washed by distillated water and absolute ethyl alcohol rigorously. Finally, the mixture was vacuum dried at 358 K for 60 min. The photocatalysts were further characterized using methods as described in following section.
 |
| Fig. 1 The picture of the microwave reactor. | |
2.3. Characterization
XRD analysis was conducted on a Philips X'pert diffractometer equipped with an X'celerator module using Cu Kα radiation. Diffractograms were obtained for 2θ = 10–70° with a step size of 0.0167°.
UV-Vis absorption spectrum of the photocatalyst was measured using an absorption spectrophotometer (Bechman, DU7000). The spectral region was from 200 to 800 nm operating at a resolution of 2 nm. The sample cell was a quartz cuvette (1 cm by 1 cm).
The metallic elemental composition of the photocatalyst was measured with a Perkin Elmer Optima NexIon™ 300 ICP-MS (Inductively Coupled Plasma Mass Spectrometry) system. Briefly, 50 mg of the powder catalyst were digested by adding 2 mL HNO3 (67–70%, v/v) and 1 mL H2O2 (30 wt% in H2O). Then the suspension was heated on a hotplate at 150 °C for 20 min with a watch glass covering the beaker. The S composition of the photocatalyst was analyzed by Ion Chromatography (IC) on a Dionex AS50 equipped with an ED50 electrochemical detector and an AS9-HC column. The mobile phase was a solution of 9 mM Na2CO3, and the flow rate was set at 1.0 mL min−1.
BET surface area was determined by the N2 physisorption measurement performed on a Micromeritics ASAP 2020 physisorption analyzer. Prior to analysis, samples were degassed at 200 °C for 4 h under vacuum. The BET surface area was calculated from the adsorption isotherm in the region 0.05 < P/P0 < 0.3.
SEM images were obtained on a Zeiss Ultra60 microscope with an accelerating voltage of 20 kV. The sample powder was spread on a carbon-coated sample mount and coated with gold to prevent surface charging effects. Optimum images were taken using the “inlens” detector mode and 5.5 mm of working distance. Elemental analysis of the powder catalyst was performed using EDS with the Zeiss Ultra60 FE-SEM.
The morphology of the sample was determined by TEM on an FEI-Tecnai F20 FEG-TEM with an accelerating voltage of up to 200 kV. TEM samples were prepared by placing 5 μL of the aqueous Zn(1−1.5x)FexS water suspension on copper grids with a continuous carbon film coating, followed by solvent evaporation at room temperature.
2.4. Photocatalytic activities of Zn(1−1.5x)FexS nanoparticles
Photocatalytic removal experiments were conducted in a cylindrical reactor (Fig. 2). In the experiment, the solution was made by adding 0.1 g of Zn(1−1.5x)FexS powder to 300 mL aqueous solution of DMP with a concentration of 10 mg L−1 at pH 6.5. The pH value of the initial reaction solution was adjusted by adding 0.1 M HCl or 0.1 M NaOH solutions. Before irradiation, the suspension was magnetically stirred in the dark for 30 min to ensure adsorption equilibrium of DMP on the catalysts. Simulated solar light (SSL) irradiation was provided by a 500 W xenon lamp (dominant wavelength is 250–1000 nm) that was positioned in the cylindrical quartz cold trap. The irradiation intensity was about 1.2 W cm−2 measured by a FZ-A spectroradiometer (the photoelectric instrument factory of Beijing Normal University, China). The system was cooled by circulating water and maintained at room temperature (20 °C). Approximately 3.0 mL of reaction solution was taken at given time intervals and centrifuged to separate the catalyst powder for DMP analysis.
 |
| Fig. 2 Schematic diagram of photochemical reaction device. | |
DMP concentration was analyzed by HPLC instrument (LC-20AD; SHIMADZU) equipped with an electrolytic conductivity detector (CDD-10AVP; SHIMADZU). Mobile phase A was methanol and phase B was the DI water containing 20 mmol L−1 of KH2PO4, and flow rate was 0.8 mL min−1. The limit of detection for DMA is 0.01 mg L−1 in this experiment. All measurements of the DMP degradation at different irradiation times were performed three times to confirm their reproducibility. The presented data points are mean values with SDs as error bars. The removal efficiency of DMP can be calculated by:
|
 | (1) |
where
R,
C0 and
C are the removal efficiency of DMP, initial concentration of solution and concentration of DMP after irradiation at various time interval (
t), respectively.
3. Results and discussion
3.1. X-ray diffraction and specific surface areas of Zn(1−1.5x)FexS
The XRD patterns were used for characterization and evaluated of crystallite sizes of the synthesized Zn(1−1.5x)FexS. Fig. 3 shows the XRD patterns of Zn(1−1.5x)FexS (x = 0, 0.006, 0.01, 0.02 and 0.03). Despite the different x values, the XRD patterns of these photocatalysts are similar, which indicates that the x value had a negligible effect on the crystallinity. Three major peaks are clearly observed at 2θ values of 28.5°, 47°, and 55.5° with indexed as (111), (200), and (222), which they well correspond to the standard card JCPDS no. 5-566.25,26 The Fe metal peaks were not observed in the XRD patterns of all obtained Zn(1−1.5x)FexS, indicating the added Fe ions should entered into ZnS host lattice as substituent.13
 |
| Fig. 3 XRD patterns of undoped and Fe doped (x = 0.006, 0.01, 0.02 and 0.03) ZnS nanoparticles. | |
With the decreasing Fe content in the Zn(1−1.5x)FexS, the diffraction peaks gradually shift toward the smaller angle, which agrees with previous findings.16 The successive shift of XRD patterns also implies that crystals of the Zn(1−1.5x)FexS photocatalyst were not a simple physical mixture of FeS and ZnS.
The crystallite size was determined using the Scherrer equation:27,28
|
 | (2) |
where
D is the crystallite size (nm),
λ is the wavelength of the incident X-ray (0.15406 nm),
θ is the diffraction angle of the (111) peak plane (degree), and
B corresponds to the full width at half-maximum of the crystalline plane (radian). The size of Zn
(1−1.5x)Fe
xS particles was calculated accordance to the Scherrer equation and shown in
Table 1. The crystallite sizes of all obtained samples were from 48.2 to 65.5 nm.
Table 1 Characterization results of Zn(1−1.5x)FexS (0 ≤ x ≤ 0.03) photocatalysts
Value of x |
Crystallite size (nm) |
Surface area (m2 g−1) |
Band gap (eV) |
Absorption edge (nm) |
0 |
65.5 |
85.7 |
3.64 |
341 |
0.006 |
56.4 |
61.8 |
3.28 |
378 |
0.01 |
49.7 |
68.9 |
2.96 |
419 |
0.02 |
48.2 |
75.6 |
2.85 |
435 |
0.03 |
48.5 |
73.8 |
2.74 |
453 |
The calculated BET surface areas of Zn(1−1.5x)FexS were also listed in Table 1. The highest BET surface area of 85.7 m2 g−1 was found in ZnS (x = 0). Introducing Fe3+ into the lattice ZnS resulted in a significant decrease of BET surface areas, possibly owing to particle agglomeration upon the microwave synthesis process, the blocking of the greater accumulation of Fe on ZnS surface.16
3.2. Elemental analysis of Zn(1−1.5x)FexS
The chemical compositions of the catalysts determined with ICP-MS and IC are shown in Table 2. The measured element compositions closely match the theoretical formula of Zn(1−1.5x)FexS, particularly for the catalyst that was synthesized at x = 0.02. Fig. 4 shows the EDS analysis result that also indicates the presence of chemical elemental Zn, Fe, and S in the Zn0.97Fe0.02S catalyst. It is clear that Fe ions have been incorporated in Zn2+ lattice sites.
Table 2 Element compositions of the catalysts synthesized at different x values
Value of x |
Chemical formula |
0 |
Zn(0.995±0.002)S |
0.006 |
Zn(0.990±0.005)Fe(0.006±0.001)S |
0.01 |
Zn(0.985±0.004)Fe(0.01±0.002)S |
0.02 |
Zn(0.970±0.003)Fe(0.02±0.001)S |
0.03 |
Zn(0.956±0.004)Fe(0.03±0.002)S |
 |
| Fig. 4 Energy-dispersive X-ray spectroscopy spectrum of the Zn0.97Fe0.02S photocatalysts. | |
3.3. UV-vis absorption spectra of Zn(1−1.5x)FexS
Fig. 5(a) shows the UV-Vis absorption spectra of as-prepared Zn(1−1.5x)FexS photocatalysts. ZnS had no light absorption in visible-light region, whereas the absorption edge for the Zn(1−1.5x)FexS (0 < x ≤ 0.03) samples is red-shifted relative to ZnS, which is attributed to the incorporation of Fe into the lattice of ZnS. These Zn(1−1.5x)FexS (x ≥ 0.01) photocatalysts had intense absorption bands with steep edges in the visible light region, indicating that the visible-light absorption was due to a band gap transition rather than to the transition of impurity energies to the CB of Zn(1−1.5x)FexS; this phenomenon was also observed for metal-ion-doped ZnS photocatalysts.29,30
 |
| Fig. 5 (a) Optical absorption spectra and (b) Tauc's plot for the band gap values for Zn(1−1.5x)FexS photocatalysts. | |
According to the Kubelka–Munk function, the band gaps of the Zn(1−1.5x)FexS sample scan be determined from the plot of (αhν)2 versus the energy of the excitation light (hν), where α is the absorption coefficient (the absorbance in Fig. 5(a)) and hν is the incident photon energy.31,32 The plots of (αhν)2 versus hν for the Zn(1−1.5x)FexS samples are shown in Fig. 5(b). Extrapolation of linear regions of the plots to zero value can give the direct band gap values, which are listed in Table 1. These samples have optical absorption gaps of 3.64 eV, 3.28 eV, 2.96 eV, 2.85 eV and 2.74 eV for ZnS, Zn0.991Fe0.006S, Zn0.985Fe0.01S, Zn0.97Fe0.02S and Zn0.955Fe0.03S, respectively. The decreasing band gap suggests that incorporating Fe ions in ZnS can effectively reduce its band gap into the visible light absorption region.
3.4. Morphology of Zn(1−1.5x)FexS
TEM and SEM images were performed to assess the size, morphology and microstructure of the nanoparticles. The SEM and TEM micrographs of the Zn0.97Fe0.02S (x = 2) photocatalyst were shown in Fig. 6. Fig. 6(a) shows that the agglomerates of particles and not the crystallite size. It was not possible by SEM image to calculate the crystallite size due to the resolution limit. Although the particles were aggregated, it is clear that the particle morphology is cubic. The more precise size distribution of nanocrystallites was performed by TEM. As shown in Fig. 6(b), the average particle size of the Zn0.97Fe0.02S photocatalysts based on the statistical measurement (no less than 30 particles were counted) was 50 nm, which agrees well with the crystallite size derived from the XRD results.
 |
| Fig. 6 (a) SEM and (b) TEM images of the Zn0.97Fe0.02S photocatalysts. | |
3.5. Calculation of band edge levels of Zn(1−1.5x)FexS
The positions of conduction band and valence band of the Zn(1−1.5x)FexS samples in relation to the normal hydrogen electrode (NHE) potential can be calculated using the following equations.33–35 |
 | (3) |
|
 | (4) |
|
 | (5) |
|
 | (6) |
|
 | (7) |
where X is the absolute electronegativity of a pristine semiconductor and is expressed as the geometric mean of x of the constituent atoms; x is the electronegativity of a neutral atom; E0 is the energy of a free electron on the hydrogen scale (∼4.5 eV); Eg is the semiconductor band gap energy (eV); A is the atom's electron affinity; I is the first ionization energy. The data of A and I obtained from ref. 36.
Using eqn (3)–(7), the band positions for ECB, and EVB can be calculated in shown Tables 3 and 4. It should be noted that the band edges calculated are approximate. The ECB becomes more negative with decreasing x value in the Zn(1−1.5x)FexS samples, indicating that the Zn-doped sulfide has stronger oxidation activity for DMP removal than those sulfides doped with Fe. However, ZnS itself has a large band gap (3.64 eV) and thus cannot utilize visible light for DMP removal. Thus, the best photocatalytic activity should be achieved by keeping the balance between the oxidation power and light absorption.37
Table 3 Values of the electron affinity, ionization energy and element electronegativity
Constituent elements |
Electron affinity (eV) |
Ionization energy (eV) |
Element electronegativity (eV) |
Fe |
0.151 |
7.902 |
4.026 |
Zn |
−0.87 |
9.394 |
4.262 |
S |
2.077 |
10.36 |
6.2182 |
Table 4 X, Eg, ECB, and EVB at the point of zero charge for Zn(1−1.5x)FexS
Value of x |
X (eV) |
Eg (eV) |
ECB (eV) |
EVB (eV) |
0 |
5.148 |
3.64 |
−1.17 |
2.47 |
0.006 |
5.149 |
3.28 |
−0.99 |
2.29 |
0.01 |
5.149 |
2.96 |
−0.83 |
2.13 |
0.02 |
5.150 |
2.85 |
−0.78 |
2.08 |
0.03 |
5.151 |
2.74 |
−0.72 |
2.02 |
3.6. Photocatalytic response of Zn(1−1.5x)FexS nanoparticles
Fig. 7 shows the removal efficiency of DMP in the presence of undoped and Fe-doped ZnS under Xe lamp irradiation. No DMP was removed without the Zn(1−1.5x)FexS photocatalysts under Xe lamp irradiation. As shown in Fig. 7(a), the removal efficiency of DMP increased gradually with x value from 0 and reached a maximum level (97.5%) at x = 0.02. According to the calculation results, the Zn0.97Fe0.02S had the highest crystallinity and the highest specific surface area, which explained its high photocatalytic activity. Further increase in x value from 0.02 to 0.03 decreased the removal efficiency of DMP, probably because of the decreased specific surface area and low crystallinity structure. Here, many crystal defects on Zn0.955Fe0.03S appeared, which may serve as recombination centers to decrease the degradation efficiency. Moreover, the removal efficiency of DMP using ZnS (x = 0) was only 18% that of the Zn0.97Fe0.02S (x = 0.02), indicating that the ratio of Fe and Zn is critical for improving the photocatalyst activity.
 |
| Fig. 7 (a) Photocatalytic degradation of DMP in the presence of undoped and Fe-doped ZnS under Xe lamp irradiation and (b) the corresponding kinetics of DMP oxidation during photocalysis. | |
Fig. 7(b) shows that the photocatalytic degradation of DMP followed the first-order decay kinetics. The experimental data indicate that the photocatalytic degradation of DMP can be described by the first-order kinetic model, ln(C0/C) = kt, where C0 and C are initial concentration of solution, concentration of DMP after irradiation at various time interval (t), respectively. The ln(C0/C) vs. t plot shows a linear relation ship with the irradiation time. The calculated rate constant (k) for ZnS was 1.4 × 10−3 min−1 and the k values of Zn0.991Fe0.006S, Zn0.985Fe0.01S, Zn0.97Fe0.02S and Zn0.955Fe0.03S were 7.0 × 10−3, 1.9 × 10−2, 2.9 × 10−3, and 2.1 × 10−3 min−1, respectively. It is clear that the doping of Fe in ZnS can increase the photocatalytic degradation of DMP under Xe lamp irradiation. In the meanwhile, the photocatalytic activity of Fe doped ZnS is strongly dependent on the dopant concentration.
4. Conclusions
In conclusion, Zn(1−1.5x)FexS synthesized by the microwave irradiation method is found to be high active for the DMP removal. In the synthesized method, the reaction time under microwave irradiation was only 30 min at 388 K and the vacuum drying time was only 60 min at 358 K, and the subsequent procedure did not need the calcination. The crystallite size, BET specific surface area, band gaps and band edge positions of the photocatalysts could be controlled by adjusting the atomic ratios of the Zn and Fe components. The photocatalyst of Zn0.97Fe0.02S has the best photocatalytic activity to remove DMP from wastewater. This work not only provides a facile and friendly synthesized method to produce highly active Zn(1−1.5x)FexS materials for removing DMP from water or wastewater, but also demonstrates new insights into understanding the photocatalytic activity through controlling synthesis process.
Acknowledgements
The work was supported by the Program of Marine Biological Resources Exploitation and Utilization of Science and Technology Innovation Team of Taizhou (Document of CPC Taizhou Municipal Committee Office of Zhejiang Province No. [2012]58), School of Biological and Chemical Engineering, Taizhou Vocational & Technical College. This work was also supported by the National Natural Science Foundation of China (No. 21301155) and Zhejiang Provincial Natural Science Foundation of China (LY13E020012).
Notes and references
- X. Y. Tang, S. Y. Wang, Y. Yang, R. Tao, Y. Dai, A. Dan and L. Li, Chem. Eng. J., 2015, 275, 198–205 CrossRef CAS.
- M. A. Surhio, F. N. Talpur, S. M. Nizamani, F. Amin, C. W. Bong, C. W. Lee, M. A. Ashraf and M. R. Shah, RSC Adv., 2014, 4, 55960–55966 RSC.
- Z. W. Xu, W. M. Zhang, L. Lv, B. C. Pan, P. Lan and Q. X. Zhang, Environ. Sci. Technol., 2010, 44, 3130–3135 CrossRef CAS PubMed.
- L. J. Xu, W. Chu and L. Gan, Chem. Eng. J., 2015, 263, 435–443 CrossRef CAS.
- B. Xu, N. Y. Gao, H. F. Cheng, S. J. Xia, M. Rui and D. D. Zhao, J. Hazard. Mater., 2009, 162, 954–959 CrossRef CAS PubMed.
- B. L. Yuan, X. Z. Li and N. Graham, Water Res., 2008, 42, 1413–1420 CrossRef CAS PubMed.
- Y. Wang, Y. Liu, T. Liu, S. Song, X. Gui, H. Liu and P. Tsiakaras, Appl. Catal., B, 2014, 156, 1–7 CrossRef.
- D. Han, J. Li, H. Cao, M. He, J. Hu and S. Yao, Chemosphere, 2014, 95, 50–57 CrossRef CAS PubMed.
- D. W. Liang, T. Zhang and H. H. P. Fang, Water Res., 2007, 41, 2879–2884 CrossRef CAS PubMed.
- Y. H. Chen, N. C. Shang, L. L. Chen, C. Y. Chang, P. C. Chiang, C. Y. Hu and C. H. Chang, Desalin. Water Treat., 2014, 52, 3377–3383 CrossRef CAS.
- Z. J. Huang, P. X. Wu, Y. H. Lu, X. R. Wang, N. W. Zhu and Z. Dang, J. Hazard. Mater., 2013, 246, 70–78 CrossRef PubMed.
- W. C. Liao, T. Zheng, P. Wang, S. S. Tu and W. Q. Pan, Environ. Eng. Sci., 2010, 27, 1001–1007 CrossRef CAS.
- N. Wetchakun, B. Incessungvorn, K. Wetchakun and S. Phanichphant, Mater. Res. Bull., 2013, 48, 1668–1674 CrossRef CAS.
- S. Kumar and N. K. Verma, J. Mater. Sci.: Mater. Electron., 2015, 26, 2754–2759 CrossRef CAS.
- Y. Li, C. Cao and Z. Chen, Chem. Phys. Lett., 2011, 517, 55–58 CrossRef CAS.
- P. C. Patel and P. C. Srivastava, J. Mater. Sci., 2014, 49, 6012–6019 CrossRef CAS.
- S. Sambasivam, D. P. Joseph, D. R. Reddy, B. K. Reddy and C. K. Jayasankar, Mater. Sci. Eng., B, 2008, 150, 125–129 CrossRef CAS.
- R. Chauhan, A. Kumar and R. P. Chaudhary, Spectrochim. Acta, Part A, 2013, 113, 250–256 CrossRef CAS PubMed.
- R. Chauhan, A. Kumar and R. P. Chaudhary, Appl. Surf. Sci., 2013, 270, 655–660 CrossRef CAS.
- W. C. Liao and P. Wang, J. Braz. Chem. Soc., 2009, 20, 866–872 CrossRef CAS.
- L. X. Yin, D. Wang, H. J. Feng, L. Y. Cao, H. B. Ouyang, J. P. Wu and X. Yong, Ceram. Int., 2015, 41, 3288–3292 CrossRef CAS.
- T. Prakash, G. Neri, A. Bonavita, E. R. Kumar and K. Gnanamoorthi, J. Mater. Sci.: Mater. Electron., 2015, 26, 4913–4921 CrossRef CAS.
- K. Vijayalakshmi and S. D. Jereil, J. Mater. Sci.: Mater. Electron., 2014, 25, 5089–5094 CrossRef CAS.
- B. X. Zhao, X. Li, L. Yang, F. Wang, J. C. Li, W. X. Xia, W. J. Li, L. Zhou and C. l. Zhao, Photochem. Photobiol., 2015, 91, 42–47 CrossRef CAS PubMed.
- H. R. Rajabi, O. Khani, M. Shamsipur and V. Vatanpour, J. Hazard. Mater., 2013, 250–251, 370–378 CrossRef CAS PubMed.
- O. Khani, H. R. Rajabi, M. H. Yousefi, A. A. Khosravi, M. Jannesari and M. Shamsipur, Spectrochim. Acta, Part A, 2011, 79, 361–369 CrossRef CAS.
- D. H. Wang, L. Wang and A. W. Xu, Nanoscale, 2012, 4, 2046–2053 RSC.
- G. S. Zhang, W. Zhang, D. Minakata, Y. S. Chen, J. Crittenden and P. Wang, Int. J. Hydrogen Energy, 2013, 38, 11727–11736 CrossRef CAS.
- G. S. Zhang, W. Zhang, D. Minakata, P. Wang, Y. S. Chen and J. Crittenden, Int. J. Energy Res., 2014, 38, 1513–1521 CrossRef CAS.
- Q. Li, H. Meng, P. Zhou, Y. Q. Zheng, J. Wang, J. G. Yu and J. R. Gong, ACS Catal., 2013, 3, 882–889 CrossRef CAS.
- K. Li, B. Chai, T. Peng, J. Mao and L. Zan, RSC Adv., 2013, 3, 253–258 RSC.
- E. S. Agorku, M. A. Mamo, B. B. Mamba, A. C. Pandey and A. K. Mishra, J. Porous Mater., 2015, 22, 47–56 CrossRef CAS.
- M. A. Butler and D. S. Ginley, J. Electrochem. Soc., 1978, 125, 228–232 CrossRef CAS.
- Y. Xu and M. A. A. Schoonen, Am. Mineral., 2000, 85, 543–556 CrossRef CAS.
- G. S. Zhang, W. Zhang, J. C. Crittenden, Y. S. Chen, D. Minakata and P. Wang, Chin. J. Catal., 2013, 34, 1926–1935 CrossRef CAS.
- D. R. Lida, Handbook of Chemistry and Physics, CRC Press, Florida, 87th edn, 2006-2007 Search PubMed.
- S. H. Mohamed, J. Phys. D: Appl. Phys., 2010, 43, 035406 CrossRef.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra20314g |
|
This journal is © The Royal Society of Chemistry 2015 |
Click here to see how this site uses Cookies. View our privacy policy here.