DOI:
10.1039/C6RA08316A
(Paper)
RSC Adv., 2016,
6, 49966-49972
Magnetic BiFeO3 grafted with MWCNT hybrids as advanced photocatalysts for removing organic contamination with a high concentration
Received
31st March 2016
, Accepted 11th May 2016
First published on 12th May 2016
Abstract
This work focuses on a novel visible light responsive photocatalysis system for removing organic contamination in a high concentration, employing MWCNTs as a scaffold and an efficient electron relay mediator and visible light and magnetically responsive BiFeO3. The MWCNT grafted BiFeO3 composites are fabricated by annealing an acid-functional MWCNT and Bi–Fe complex mixture at 800 °C, which induces the in situ formation of BiFeO3 particles that are grafted with the MWCNT matrix. The optimal BiFeO3/MWCNT photocatalyst exhibits enhanced photocatalytic activity under visible light, ascribed to efficient charge separation and transport which greatly extends the electron life time using MWCNTs as an electron acceptor. Simultaneously, the high surface area of the BiFeO3/MWCNT photocatalyst can act as an adsorbent to reduce organic contamination of a high concentration under an external magnetic field. The two processes composed of physical adsorption and photocatalytic degradation result in the fast removal of organic contamination during a short reaction time compared to N-doped TiO2. This work not only provides a new strategy for the fabrication of high performance multimetallic oxide grafted MWCNT photocatalysts, but also facilitates their practical application in environmental issues.
1. Introduction
The increasing emissions of organic contamination from the papermaking and textile industries greatly threaten human health. A sustainable, green and recyclable system is required to purify and separate these organic contaminations. In recent years, semiconductor photocatalysis, as a “green” technology, has been investigated for treating polluted water via degradation or splitting processes.1–12 In view of solar energy utilization, researchers have been focusing on the development of visible-light-responsive photocatalysts, because ultraviolet (UV) light only accounts for about 4% whereas visible light contributes to about 43%.13–15 Stable multimetallic oxides have attracted much attention as novel visible light responsive semiconductors for the removal of organic contamination under visible light irradiation.16–19 Electronic structures calculated using density functional theory (DFT) show that the d orbital of the transition metals contributes the bottom of the conduction band (CB), while the top of the valence band (VB) consists of an O 2p orbital.20 In particular, a perovskite-type BiFeO3 with 2.18 eV bandgap has received considerable attention in recent decades because of its good photochemical stability and low cost.21–23 Moreover, BiFeO3 is a magnetic semiconductor material, meaning that it could be magnetically separable in aqueous organic contamination.16,24 However, due to its low surface area (nanoparticle: 8.3 m2 g−1, bulk: 1.2 m2 g−1) and high ratio of recombination of electron/hole pairs, it needs to be combined with foreign materials that could offer a high surface area and electron transport ability.25–28 For instance, Li et al. synthesized a BiFeO3/graphene nanohybrid via a hydrothermal reaction that presented a six times enhancement in photocatalytic Congo red degradation compared to pure BiFeO3 particles.25 Dai also combined BiFeO3 with graphene which significantly enhanced the photocatalytic performance for methyl orange degradation.26 However, compared to 2D graphene, growing BiFeO3 on 1D MWCNTs is a challenging task because BiFeO3 photocatalysts with a relatively larger size are difficult to coat or deposit on the MWCNT surface as well. Therefore, a tailor-made method is expected to control their size, structure, and morphology.29–31
Herein, we report one-step calcination to directly fabricate magnetic MWCNT/BiFeO3 composites using citric acid as a complexant in 2-methoxyethanol solution. The experimental results show that the as-prepared photocatalysts can serve triply as a physical and photochemical degrader for organic molecules of high concentration under visible light. Prior to irradiation, an adsorption experiment was carried out using mechanical separation using an external magnetic field. Followed by physical absorption, photocatalytic degradation was further carried out under visible light. In a short period of 2 h, RhB as a contamination model with a concentration of 10−3 M can be completely removed.
2. Experimental
2.1 Materials
Bismuth nitrate (Bi(NO3)3·5H2O) and iron nitrate (Fe(NO3)3·9H2O) were purchased from Sigma Aldrich. 2-Methoxyethanol (C3H8O2), poly(ethylene glycol), ammonia solution (NH3·H2O, 3 mol L−1) and citric acid were obtained from Shanghai Ansin Chemical Co. Ltd, Japan. Rhodamine B (RhB) was supplied by Shanghai Yiji Dye Chemicals.
2.2 Synthesis of MWCNT/BiFeO3 photocatalysts
Acid-functional MWCNTs. The MWCNTs were first purified and functionalized using concentrated nitric acid. Typically, 500 mg of MWCNTs were dispersed in 150 mL of concentrated nitric acid and refluxed at 140 °C for 8 h. Then, the functional MWCNTs were filtrated and rinsed with water several times until pH = 7, and were finally dried for further use at 50 °C in an oven.
Synthesis of MWCNT/BiFeO3 photocatalysts. 1 mM Bi(NO3)3·5H2O and 1 mM Fe(NO3)3·9H2O were firstly dissolved in 100 mL C3H8O2 under vigorous stirring. The pH of the solution was controlled to be 4.5 by slowly adding nitric acid. 1 mM citric acid as a complexant and 0.005 g poly(ethylene glycol) as a dispersant were then added to the above solution. The mixture was further stirred for 2 h at 50 °C to obtain the sol, then, different amounts of MWCNTs were added into the above sol and kept at 80 °C for 4 days to form a xerogel. The final xerogel was ground into a fine powder and calcined at t = 800 °C for 2 h in N2. N-Doped TiO2 was synthesized by annealing anatase TiO2 (∼40 nm) under NH3 conditions for 2 h.
2.3 Photoelectrochemical measurements
Photocurrent responses were carried out using a three-electrode system with a potentiostat (CH Instruments, CHI 660), and Pt as a counter electrode, and Ag@AgCl as a reference electrode in 1 M Na2SO4 solution at 0.5 V potential bias under a solar simulator with a Xe lamp through a UV-cutoff filter (>420 nm). The electrodes were prepared by dropping 20 μL of aqueous slurry that consisted of 50 mg of prepared sample, 5 μL of acetylacetone, and 5 μL of Nafion in 200 μL of H2O on a cleaned ITO glass substrate and annealing at 300 °C for 2 h in air.
2.4 Analysis instruments
Diffuse reflectance UV-visible absorption spectra were carried out using a spectrophotometer (Shimadzu UV-2401PC) with an integrating sphere attachment. BaSO4 was used as the reference. XRD patterns were recorded with a diffractometer (Riguka, Japan, RINT 2500 V) using Cu Kα radiation. A N2 adsorption isotherm was measured at 77 K using a BEL sorp Analyzer (BEL, Japan). Raman spectra were produced using a Horiba Jobin Yvon LabRAM using a 100× objective lens with a 532 nm laser excitation. XPS analysis was obtained using an ESCALAB-220I-XL (THERMO-ELECTRON, VG Company) device. Transmission electron microscopy (TEM) observations were performed on a JEOL, JEM-2100F (Japan) electron microscope with an accelerating voltage of 200 kV. The photoluminescence of powder samples was recorded at room temperature using a fluorescence spectrometer (Shimadzu, RF-5410PC).
2.5 Photocatalytic test
For the photodegradation of organic dyes, a decomposition reaction of a 50 mL RhB (1 mM) aqueous solution was carried out. A 30 mg powdered sample was dispersed in the RhB solution under ultrasonication for 1 min, and then was separated using an external magnetic field. The remaining RhB solution was added to 30 mg of refreshed sample under the irradiation of a 300 W Xe lamp with a 420 nm filter with a distance of 100 mm from the solution in a darkness box. The samples were then withdrawn regularly from the reactor and the dispersed powder was separated using a magnet. The clean transparent solution was analyzed using UV-vis spectroscopy (Shimadzu 2550). The concentration of RhB in the solution was determined as a function of irradiation time from the absorbance region at a functional wavelength.
3. Results and discussion
Fig. 1a presents the XRD patterns of pure BiFeO3 and BiFeO3/MWCNT photocatalysts. It reveals that the BiFeO3 nanoparticles are highly crystallized and exhibit a single-phase perovskite structure in both BiFeO3 and BiFeO3/MWCNT photocatalysts. The crystal phase of BiFeO3 is determined to be rhombohedral,32 which is obviously different from the tetragonal structure of BiFeO3 films.33
 |
| Fig. 1 (a) XRD patterns of BiFeO3 and BiFeO3/MWCNT, and (b) Raman shift of MWCNT and BiFeO3/MWCNT. | |
The crystal structure of BiFeO3 and the presence of MWCNTs are further investigated using Raman spectra, as shown in Fig. 1b. The broad peaks below 250 cm−1 can be assigned to A1 (1TO), A1 (2TO), and A1 (3TO) modes for the rhombohedral BiFeO3 system, which is consistent with what has been previously reported.34 Moreover, a larger number of bands between 500 and 600 cm−1 are found, which may reflect a space group with a lower symmetry.35 Furthermore, two peaks of D and G corresponding to sp3 and sp2 carbon are found in BiFeO3/MWCNT, indicating the co-existence of BiFeO3 and MWCNTs in the BiFeO3/MWCNT photocatalysts.
The morphology of BiFeO3/MWCNT is examined using transmission electron microscopy (TEM) and field emission scanning electron microscopy (FESEM). As shown in Fig. 2a, the diameter of the MWCNTs is around 10–20 nm. The uniform growth of BiFeO3 can be observed distinctly in Fig. 2b, where the BiFeO3 grows homogeneously in the MWCNT matrix. Fig. 2c and d are typical TEM images that show the detailed structure of BiFeO3/MWCNT. It can clearly be seen that BiFeO3 particles with an average size of larger than 100 nm are distributed in the MWCNT matrix.
 |
| Fig. 2 (a) TEM image of MWCNT, (b) FE-SEM images of BiFeO3/MWCNT, and (c) and (d) TEM images of BiFeO3/MWCNT. | |
It was well-known that a strong interface could maximize charge transport and separation, while suppressing the recombination of the electron–hole pairs in hybrids. Hence, the chemical bonding between BiFeO3 and MWCNTs could play an important role in promoting their photocatalytic activity. The chemical bonding between BiFeO3 particles and MWCNTs is characterized using XPS analysis. The XPS signals from the high resolution C 1s and O 1s spectra are shown in Fig. 3a and b, respectively. Curve fittings have determined the exact peak location of C 1s in C–C, C–O, C
O and COOH bonds, which originate from the primary oxidation of MWCNTs. The XPS spectrum of O 1s in the BiFeO3/MWCNT is then deconvoluted by four peaks of COOH, C
O, Bi–O, and Fe–O bonds, as shown in Fig. 3b. The two binding energies in BiFeO3 usually exhibit 530.5 eV for Bi–O and 529.9 eV for Fe–O bonds.28 Interestingly, the two binding energies obtained for the Bi–O position and Fe–O position are 531.7 eV and 530.1 eV, respectively, in the BiFeO3/MWCNT, and the red-shift of several electron volts might imply the existence of C–O–Fe or C–O–Bi bonds. The Bi 4f and Fe 2p XPS spectra in Fig. 3c and d further indicate the formation of a BiFeO3 crystal, where the peaks at 159.3 and 164.5 eV are respectively assigned to the Bi 4f7/2 and Bi 4f5/2 states for Bi3+, and both of the peaks at 711.1 and 724.4 eV correspond to Fe3+ in BiFeO3/MWCNT.
 |
| Fig. 3 XPS spectra of BiFeO3/MWCNT. (a) C 1s XPS spectrum, (b) O 1s XPS spectrum, (c) Bi 4f XPS spectrum, and (d) Fe 2p XPS spectrum. | |
Before the photocatalytic activity characterization, the optical absorption of the as-prepared BiFeO3 particles was investigated which is relevant to their band gap. As shown in Fig. 4a, the absorption edge of the bare BiFeO3 is about 600 nm, while that of 10 wt% BiFeO3/MWCNT is red-shifted to longer than 700 nm. Fig. 4b shows the spectra transformation of the bare BiFeO3 from the diffuse reflection spectra according to the Kubelka–Munk (K–M) theory.36 The band gap for the bare BiFeO3 is determined to be 2.15 eV, which is similar to a previous result.37 A much smaller band gap of 1.7 eV is obtained for the 10 wt% BiFeO3/MWCNT, which is agreement with a red shift in the absorption edge of BiFeO3/MWCNT as compared to the bare BiFeO3. Analogous band gap narrowing of BiFeO3 is also found in the case of BiFeO3/graphene25,28,38 which could be attributed to the chemical bonding between BiFeO3 and the specific sites of carbon.
 |
| Fig. 4 (a) UV-vis DRS spectra of BiFeO3 and BiFeO3/MWCNT. (b) Plot of the transformed Kubelka–Munk function versus the energy of light. | |
As mentioned above, the magnetic properties of BiFeO3 make it a good absorbent for organic contamination removal using magnetic separation techniques. As shown in Fig. 5, the characteristic UV-vis absorption peak of rhodamine B is located at 553 nm, and when the amount of 10 wt% BiFeO3/MWCNT is established to be 1 g L−1, over 41% of rhodamine B can be removed within 10 min at room temperature. Adsorption efficiencies express the relationship between the MWCNT content in BiFeO3/MWCNT and the concentration changes of adsorbate in the bulk solution at a given temperature under equilibrium conditions, as shown in Fig. 5b. The BiFeO3/MWCNT with 30% BiFeO3/MWCNT content possesses the maximum adsorption efficiency for RhB. The increased adsorption efficiency with increasing MWCNT content up to 30 wt% could be ascribed to their rising BET surface area (Table 1), which is beneficial for RhB uptake, as is well-known. The decreased adsorption efficiency of the BiFeO3/MWCNT photocatalysts with 50 wt% content can be ascribed to excess MWCNTs that lead to difficult magnetic separation due to weakly magnetic BiFeO3/MWCNT.
 |
| Fig. 5 (a) UV-Vis spectra of RhB solution before and after adsorption by 10 wt% BiFeO3/MWCNT. The inset shows the adsorption process of BiFeO3/MWCNT for RhB. (b) Adsorption efficiencies of RhB over BiFeO3/MWCNT with different MWCNT content. | |
Table 1 BET surface area and degradation ratio of various samples
Samples |
Surface area (m2 g−1) |
Degradation ratio (min−1) |
MWCNT |
212 |
|
BiFeO3 |
8.9 |
0.00466 |
5 wt% BiFeO3/MWCNT |
25.6 |
0.0096 |
10 wt% BiFeO3/MWCNT |
47.8 |
0.02441 |
20 wt% BiFeO3/MWCNT |
65.4 |
0.00738 |
30 wt% BiFeO3/MWCNT |
84.6 |
0.00536 |
50 wt% BiFeO3/MWCNT |
102.9 |
0.00208 |
The photocatalytic activities of the BiFeO3/MWCNT photocatalysts with different MWCNT content for the degradation of RhB were performed under visible-light irradiation. Fig. 6a shows the RhB concentration decrease against irradiation time over BiFeO3 and various BiFeO3/MWCNT composites. The initial RhB concentration is based on each adsorption efficiency. Under irradiation for 120 min, the RhB solution is completely bleached over 10% BiFeO3/MWCNT. Fig. 6b further shows that the photodegradation rates of RhB decreased in the following order: 10 wt% BiFeO3/MWCNT > 5 wt% BiFeO3/MWCNT > 20 wt% BiFeO3/MWCNT > BiFeO3 > 30 wt% BiFeO3/MWCNT > 50 wt% BiFeO3/MWCNT. Clearly, when the weight addition ratio of MWCNTs is increased to 20%, the activity is lower than the 10% BiFeO3/MWCNT, although it is higher than BiFeO3. A further increase of the weight addition ratio of MWCNTs will lead to a significant decrease of photocatalytic activity. The results demonstrated that 10 wt% BiFeO3/MWCNT could be the optimal condition for organic contamination removal, considering its sustainability. The pseudo-first-order kinetics of photocatalytic degradation ratio for these BiFeO3/MWCNT composites are shown in Fig. 6c, and the equation could be expressed as follows:
−ln C/C0 = kT |
where
C and
C0 are the concentration and initial concentration of RhB, respectively,
T is the irradiation time and
k is the degradation rate constant. By plotting ln(
C/
C0) as a function of time, the
k (min
−1) values are obtained from the slopes of the lines of best fit and listed in
Table 1. The RhB removal stability of the optimized 10 wt% BiFeO
3/MWCNT composite is evaluated by repeating the performance four times, as shown in
Fig. 6d. It can be clearly seen that the removal efficiency only reveals a slight fading after four continuous times, indicating a high practicability for organic contamination removal.
 |
| Fig. 6 (a) Photocatalytic removal of RhB over BiFeO3/MWCNT with different MWCNT content under visible light irradiation. (b) Photocatalytic degradation rates of BiFeO3/MWCNT with different MWCNT content under visible light irradiation. (c) The kinetics of photocatalytic degradation of RhB over BiFeO3/MWCNT with different MWCNT content under visible light irradiation. (d) Durability study of 10 wt% BiFeO3/MWCNT used as both photocatalyst and absorbent. | |
To demonstrate the efficient removal of high concentrations of RhB from solution, the photocatalytic performance of the optimized 10 wt% BiFeO3/MWCNT composite is compared with N doped TiO2. As shown in Fig. 7a, the overall removal of 1 mM RhB over 10% BiFeO3/MWCNT could be carried out within 130 min, which includes 10 of min physical adsorption and 120 of min photocatalytic degradation. However, the overall removal of 1 mM RhB over N doped TiO2 needs to take more than 310 min under the same conditions as the process for the 10% BiFeO3/MWCNT. The results indicate that BiFeO3/MWCNT as an advanced photocatalyst could efficiently remove highly concentrated RhB through the means of physical adsorption and photocatalytic degradation.
 |
| Fig. 7 (a) Comparison of removal of RhB through both physical adsorption and photocatalytic degradation between 10 wt% BiFeO3/MWCNT and N-doped TiO2. (b) Transient photocurrent responses of the 10 wt% BiFeO3/MWCNT and N-doped TiO2 in 1 M Na2SO4 aqueous solution under visible-light irradiation at 0.5 V vs. Ag/AgCl. | |
To give further evidence to support the above comparison, the transient photocurrent responses were recorded for several on–off cycles of irradiation. Fig. 7b shows photocurrent curves for the aforementioned two samples. It is clear that the sensitive photocurrent response against light turns on/off actions which indicate that most of the photogenerated electrons are transported to the FTO and generate photocurrent under visible-light irradiation. Notably, the 10% BiFeO3/MWCNT electrode exhibits a higher photocurrent density than N-doped TiO2, suggesting a higher charge carrier generation and transport under visible light, corresponding to high photocatalytic activity.
Except for the limited effects on adsorption and light harvesting capacities, the dominant contribution of MWCNTs in BiFeO3/MWCNT composites to photoactivity is determined to be charge separation. A schematic illustration for the mechanism of enhanced photoactivity of the BiFeO3/MWCNT nanocomposites is shown in Fig. 8. Considering the electron affinity of BiFeO3 is ∼3.3 eV,39 and the bandgap of BiFeO3 is 2.15 eV, the CB and VB edges of BiFeO3 could be established at −0.9 and 1.25 eV, respectively. Generally, the work function of CNT is located at −0.30 eV vs. NHE.40 Therefore, under light illumination, electrons (e−) are excited from the VB to the CB, leaving holes (h+) in the VB. Normally, these e−/h+ pairs tend to recombine rapidly along with only a fraction taking part in the photocatalytic reaction, resulting in a low reactivity. However, when BiFeO3 particles are chemically bonded to the surface of MWCNTs, these photo-generated electrons could be easily transferred to MWCNTs, resulting in efficient hole–electron separation. These separated electrons and holes are directly trapped by some surface adsorbates or generate highly reactive radical species.
 |
| Fig. 8 Proposed mechanism for the enhanced electron transfer in the BiFeO3/MWCNT composites. | |
4. Conclusion
In this study, we successfully grafted magnetically responsive BiFeO3 particle into multiwalled carbon nanotube matrix (BiFeO3/MWCNT) composites with different loadings (5–50 wt%) of MWCNTs by annealing a MWCNT and Bi–Fe complex mixture at 800 °C. Moreover, chemical bonds are formed at the interface between MWCNTs and BiFeO3. It is shown that a 10 wt% MWCNT loading significantly influences the textural properties including physical adsorption and photodegradation of highly concentrated RhB (1 mM). The dominant factor in enhancing the photocatalytic performance of the BiFeO3/MWCNT composite is identified as the interfacial electron transfer and separation. It clearly indicates that only these BiFeO3/MWCNT composites with appropriate CNT loadings (10 wt%) have a much higher efficiency for RhB removal as compared with N-doped TiO2.
Acknowledgements
This work was supported by the National Natural Science Foundation of China (NSFC) (Grants no. 21271010) and the Shanghai Municipal Education Commission (No. 15ZZ088 and No. 14JC1402500) and the Science and Technology Commission of Shanghai Municipality (No. 14DZ2261000).
References
- Y. H. Ng, I. V. Lightcap, K. Goodwin, M. Matsumura, P. V. Kamat and J. Phys, Chem. Lett., 2010, 1, 2222–2227 CAS.
- K. Woan, G. Pyrgiotakis and W. Sigmund, Adv. Mater., 2009, 21, 2233–2239 CrossRef CAS.
- K. Zhang, L. Y. Wang, X. W Sheng, M. Ma, M. S. Jung, W. J. Kim, H. Y. Lee and J. H. Park, Adv. Energy Mater., 2016 DOI:10.1002/aenm.201502352.
- M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Chem. Rev., 1995, 95, 69–96 CrossRef CAS.
- H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, J. Zou, H. M. Cheng and G. Q. Lu, J. Am. Chem. Soc., 2009, 131, 4078–4083 CrossRef CAS PubMed.
- A. J. Bard and M. A. Fox, Acc. Chem. Res., 1995, 28, 141–145 CrossRef CAS.
- K. Zhang, L. Y. Wang, J. K. Kim, M. Ma, G. Veerappan, C. L. Lee, K. J. Kong, H. Y. Lee and J. H. Park, Energy Environ. Sci., 2016, 9, 499–503 CAS.
- M. S. Zhu, P. L. Chen, W. H. Ma, B. Lei and M. H. Liu, ACS Appl. Mater. Interfaces, 2012, 4, 6386–6392 CAS.
- C. Y. Zhai, M. S. Zhu, Y. T. Lu, F. F. Ren, C. Q. Wang, Y. K. Du and P. Yang, Phys. Chem. Chem. Phys., 2014, 16, 14800–14807 RSC.
- C. H. Lu, R. Y. Chen, X. Wu, M. F. Fan, Y. H. Liu, Z. G. Le, S. J. Jiang and S. Q. Song, Appl. Surf. Sci., 2016, 360, 1016–1022 CrossRef CAS.
- S. Q. Song, B. Cheng, N. S. Wu, A. Y. Meng, S. W. Cao and J. G. Yu, Appl. Catal., B, 2016, 181, 71–78 CrossRef CAS.
- M. S. Zhu, C. Y. Zhai, L. Q. Qiu, C. Lu, A. S. Paton, Y. K. Du and M. C. Goh, ACS Sustainable Chem. Eng., 2015, 3, 3123–3129 CrossRef CAS.
- Y. L. Min, G. Q. He, Q. J. Xu and Y. C. Chen, J. Mater. Chem. A, 2014, 2, 2578–2584 CAS.
- B. B. Ding, H. Y. Peng, H. S. Qian, L. Zheng and S. H. Yu, Adv. Mater. Interfaces, 2016, 3(3), 1500649 Search PubMed.
- S. N. Guo, Y. Zhu, Y. Y. Yan, Y. L. Min, J. C. Fan and Q. J. Xu, Appl. Catal., B, 2016, 185, 315–321 CrossRef CAS.
- Z. G. Zou, J. H. Ye, K. Sayama and H. Arakawa, Nature, 2001, 414, 625–627 CrossRef CAS PubMed.
- J. H. Tang, Z. G. Zou and J. H. Ye, Chem. Mater., 2004, 16, 1644–1649 CrossRef CAS.
- M. Oshikiri, J. H. Ye, Z. G. Zou and G. Kido, J. Chem. Phys., 2002, 117, 7313–7318 CrossRef CAS.
- K. Zhang, N. Heo, X. Shi and J. H. Park, J. Phys. Chem. C, 2013, 117, 24023–24032 CAS.
- J. H. Ye, Z. G. Zou, H. Arakawa, M. Oshikiri, M. Shimoda, A. Matsushita and T. Shishido, J. Photochem. Photobiol., A, 2002, 148, 79–83 CrossRef CAS.
- F. Gao, X. Y. Chen, K. B. Yin, S. Dong, Z. F. Ren, F. Yuan, T. Yu, Z. G. Zou and J. M. Liu, Adv. Mater., 2007, 19, 2889–2892 CrossRef CAS.
- Y. N. Huo, M. Miao, Y. Zhang, J. Zhu and H. X. Li, Chem. Commun., 2011, 47, 2089–2091 RSC.
- S. M. Sun, W. Z. Wang, L. Zhang and M. Shang, J. Phys. Chem. C, 2009, 113, 12826–12831 CAS.
- S. Li, Y. H. Lin, B. P. Zhang, Y. Wang and C. W. Nan, J. Phys. Chem. C, 2010, 114, 2903–2908 CAS.
- Z. X. Li, Y. Shen, C. Yang, Y. C. Lei, Y. H. Guan, Y. H. Lin, D. B. Liu and C. W. Nan, J. Mater. Chem. A, 2013, 1, 823–829 CAS.
- J. F. Dai, T. Xian, L. J. Di and H. Yang, J. Nanomater., 2013, 2013, 1–5 Search PubMed.
- P. Li, Q. Chen, Y. Y. Lin, G. Chang and Y. B. He, J. Alloys Compd., 2016, 672, 497–504 CrossRef CAS.
- Z. X. Li, Y. Shen, Y. H. Guan, Y. B. Hu, Y. H. Lin and C. W. Nan, J. Mater. Chem. A, 2014, 2, 1967–1973 CAS.
- L. W. Zhang, H. B. Fu, C. Zhang and Y. F. Zhu, J. Solid State Chem., 2006, 179, 804–811 CrossRef CAS.
- S. S. Dunkle and K. S. Suslick, J. Phys. Chem. C, 2009, 113, 10341–10345 CAS.
- Z. J. Zhang, W. Z. Wang, M. Shang and W. Z. Yin, J. Hazard. Mater., 2010, 177, 1013–1018 CrossRef CAS PubMed.
- S. T. Zhang, M. H. Lu, D. Wu, Y. F. Chen and N. B. Ming, Appl. Phys. Lett., 2005, 87, 262907 CrossRef.
- J. Wang, J. B. Neaton, H. Zheng, V. Nagarajan, S. B. Ogale, B. Liu, D. Viehland, V. Vaithyanathan, D. G. Schlom, U. V. Waghmare, N. A. Spaldin, K. M. Rabe, M. Wuttig and R. Ramesh, Science, 2003, 299, 1719–1722 CrossRef CAS PubMed.
- Y. Wang and C. W. Nan, J. Appl. Phys., 2008, 103, 114104 CrossRef.
- H. Bea, M. Bibes, S. Petit, J. Kreisel and A. Barthelemy, Philos. Mag. Lett., 2007, 87, 165–174 CrossRef CAS.
- P. Kubelka and F. Munk, Z. Tech. Phys., 1931, 12, 593 Search PubMed.
- G. Q. Tan, Y. Q. Zheng, H. Y. Miao, A. Xia and H. J. Ren, J. Am. Ceram. Soc., 2012, 95, 280–289 CrossRef CAS.
- F. Li, J. F. Shen, N. Li and M. X. Ye, Mater. Lett., 2013, 91, 42–44 CrossRef.
- S. J. Clark and J. Robertson, Appl. Phys. Lett., 2007, 90, 132903 CrossRef.
- B. Chai, T. Peng, X. Zhang, J. Mao and K. Li, Dalton Trans., 2013, 42, 3402–3409 RSC.
|
This journal is © The Royal Society of Chemistry 2016 |
Click here to see how this site uses Cookies. View our privacy policy here.