Luminescence properties of Ca2Si5N8:Eu2+ prepared by gas-pressed sintering using BaF2 as flux and cation substitution

Chuang Wang, Zhengyan Zhao, Xicheng Wang, Yanyan Li, Quansheng Wu and Yuhua Wang*
Lanzhou University, Research Institute of Environmental Materials, Department of Materials Science, Lanzhou, China. E-mail: wyh@lzu.edu.cn

Received 19th May 2014 , Accepted 15th October 2014

First published on 15th October 2014


Abstract

Eu2+ doped Ca2Si5N8 phosphors were successfully prepared by gas-pressed sintering. The red-shift of the emission band from 608 nm to the longer wavelength 622 nm of the Ca2Si5N8:Eu2+ phosphor under blue excitation has been achieved, and a large enhancement in the emission intensity has been obtained by using BaF2. XRD data revealed that the lattice of Ca2Si5N8:Eu2+ was expanded with Ba2+ ion doping. XPS results suggested that there were more Eu2+ ions incorporated into the lattice of Ba2+ doped samples than those of the undoped samples. The doping effect of Ba2+ ions has been discussed in detail.


1. Introduction

Light emitting diode (LED)-based solid-state lighting has recently received worldwide attention, owing to the characteristics of high efficiency, simple structure, and long life.1 The need for new phosphors arose in the past decade with the invention of the near-UV to blue emitting GaN-based LEDs by Nakamura in 1991 and the development of high-power LEDs (HP-LEDs) in the same spectral region.2 It is well known that the spectral properties of rare-earth ions with 5d–4f transitions (e.g., Eu2+, Ce3+) strongly depend on the surrounding environment (e.g., symmetry, covalence, coordination, bond length, site size, crystal-field strength, etc.), due to the fact that the 5d excited state is not shielded from the crystal field by the 5s2 and 5p6 electrons.3–5 These phosphors can efficiently absorb in the near UV to blue spectral range and emit light in the visible range. And the small Stokes shift leads to high conversion efficiency. For a LED phosphor to be applied in commercial products, several criteria have to be met such as high efficiency in light conversion, high thermal quenching temperature, and the possibility to adjust the color point by means of varying the chemical composition. Host lattices with a high degree of covalence and/or a large crystal field splitting at the site for which Eu2+ substitute can lead to efficient visible emission while absorbing light in the near UV to blue range of the electromagnetic spectrum. Presently Y3Al5O12:Ce3+ (YAG:Ce3+) is applied in phosphor converted LEDs (pc-LEDs) as luminescent converter.6,7 The most common commercial white LEDs (WLEDs) are combination of the blue-emitting InGaN chip and the yellow YAG:Ce3+ phosphor. However, they have a low Colour Rendering Index (CRI) because of their lack of emission in the red region, and high color temperature due to the deficiency of emission in the visible spectrum. An alternative way to overcome this weakness is the incorporated with red phosphors. In this respect, a promising new class of LED phosphor materials is the nitridosilicates. The N3− in this lattice is a soft Lewis base, which results in a high covalence. This shifts the energy of the 4f–5d absorption and emission for Eu2+ ions to sufficiently low energies.8,9 In addition, nitridosilicates are known to be highly stable with oxidation and hydrolysis.10 At present the commercial red nitrides phosphors are CaAlSiN3:Eu2+ (ref. 11 and 12) and Sr2Si5N8:Eu2+.13–15 However, the sintering process of CaAlSiN3:Eu2+ phosphor needs critical preparation conditions (higher temperature, higher N2 pressure, and air sensitive starting powders). For the latter, Piao et al.16 reported on the carbothermal reduction and nitridation method to synthesize Sr2Si5N8:Eu2+ phosphor. In this method, residual carbon is inevitably incorporated into the phosphor, which reduces its intensities of absorption and emission. Thus these nitrides phosphors cannot meet the requirements of orange-red phosphors at present. So there is a still need to discover novel orange-red phosphors. Currently, there is a increasing interest in nitridosilicates Ca2Si5N8:Eu2+ due to the potential application in warm white LEDs.17,18 And the preparation conditions are easier than the other commercial nitrides phosphors, such as the lower sintering temperature and pressure as well as the cheaper raw materials. Recently the researches on Ca2Si5N8:Eu2+ mainly concentrate on the different doping concentrations of Eu2+ to tune the emission color and increase the emission intensity. But there is no report on the cations substitution in Ca2Si5N8 host to adjust the luminescence property. BaF2 is a potential flux.19,20 So we focus on the Ca2Si5N8:Eu2+ by using a BaF2 as flux and cations substitution.

In this paper we report the possibility to tune the emission color of Ca2Si5N8:Eu2+ by incorporating BaF2, to determine the most promising way to design new compositions that can serve as efficient phosphors in pc-LEDs with a warmer color. The luminescence and thermal quenching properties have been estimated. And the mechanism for the emission increasing and wavelength shift after BaF2 doping is discussed.

2. Experimental section

2.1 Materials and synthesis

A series of nitridosilicate phosphors, Ca2Si5N8:Eu2+ with different concentration of BaF2 were prepared by gas-pressed sintering. Stoichiometric amounts of powder BaF2 (AR), Ca3N2 (Aldrich, >95.0%), Si3N4 (Aldrich, 99.5%), and Eu2O3 (Aldrich, 99.99%) were ground in an agate mortar for 30 min in a glove box to form a homogeneous mixture. The concentrations of both moisture and oxygen in the glovebox were <1 ppm. Thereafter, the powder mixture was transferred into a BN crucible and heated at 1500 °C for 4 h under high-purity nitrogen (99.9995%) atmosphere at a pressure of 0.2 MPa. The sintered products were ground again, yielding crystalline powder.

2.2 Characterization

All measurements were made on finely ground powder. The phase purity of samples were analyzed by X-ray diffraction (XRD) using a Rigaku D/Max-2400 X-ray diffractometer with Ni-filtered CuKα radiation. Photoluminescence (PL) and PL excitation (PLE) spectra were measured at room temperature using an FLS-920T fluorescence spectrophotometer equipped with a 450 W Xe light source and double excitation monochromators. High temperature luminescence intensity measurements were carried out by using an aluminum plaque with cartridge heaters; the temperature was measured by thermocouples inside the plaque and controlled by a standard TAP-02 high temperature fluorescence controller. X-ray photoelectron spectroscopy (XPS) measurements were performed on an ESCALAB250xi high-performance electron spectrometer using a monochromatized AlKα excitation source ( = 1486.6 eV).

3. Results and discussion

Fig. 1 shows the XRD patterns of the Ca2Si5N8:Eu2+ phosphors as a function of Eu2+ concentration. The detailed analysis of the XRD patterns of samples (Ca1−xEux)2Si5N8 (x = 0.05, 0.1, 0.15, 0.2 and 0.25) shows that their phase compositions depend on the concentration of Eu2+. When 0 ≤ x ≤ 0.15, the samples are almost the single phase of (Ca1−xEux)2Si5N8 with a small amount of α-Si3N4 as impurity. A further increase of Eu2+ concentration leads to the formation of α-Si3N4 and EuSiO3, both as impurities. These observations indicate that the solubility of Eu2+ in Ca2Si5N8 is very limited, probably, to a range of 0–0.15, i.e. 0 ≤ x ≤ 0.15, which is in consistent well with the result reported by Li et al.21 Such a limitation may be due to two reasons. One is the difference in the crystal structure that Ca2Si5N8 is monoclinic while Eu2Si5N8 is orthorhombic. The other is the difference between the ionic radii: the radius of Eu2+ (1.17 Å) (1 Å = 0.1 nm) is obviously larger than that of Ca2+ (1.00 Å).22
image file: c4ra04683h-f1.tif
Fig. 1 XRD patterns of the Ca2Si5N8:Eu2+ phosphors as a function of Eu2+ concentration.

Fig. 2(a) shows the excitation spectrum (λem = 617 nm) of Ca1.85Eu0.15Si5N8. The Ca1.85Eu0.15Si5N8 phosphor exhibited a typical broad excitation band resulting from the crystal field splitting of the 5d orbital due to the 4f7-ground state to the 4f65d-excited state of the Eu2+ ion electronic transitions.23 Fig. 2(b) shows the emission spectra of the Ca2−xEuxSi5N8 phosphors synthesized at 1500 °C for 4 h excited at 460 nm. The relative emission peak originated from the transitions of the 5d to the 4f states. As the Eu2+ doping concentration increases, the relative emission intensity increases continuously. The highest emission intensity is observed for the 0.15 mol of Eu2+ sample. However, when the Eu2+ concentration exceeds 0.15 mol, there was a sudden decrease in the emission intensity due to concentration quenching.24 As the Eu2+ contents increase, the distance between the Eu2+ ions becomes smaller, which leads to the probability of energy transfer among Eu2+ ions.25 When the Eu2+ concentration increases, the emission band shifts to the red side. This may be ascribed to the lattice distortion caused by Eu2+ ions introducing the mismatch between the small Ca2+ and large Eu2+ ionic radius in the lattice.26


image file: c4ra04683h-f2.tif
Fig. 2 PLE spectrum (a) (monitored at 617 nm) and PL spectra (b) (excited at 460 nm) of the Ca2Si5N8:Eu2+ phosphor with different Eu2+ contents.

Fig. 3 shows the XRD patterns of Ca1.95Eu0.05Si5N8 with different weight of BaF2. When BaF2 is doped, the sample is the single phase of (Ca1−xEux)2Si5N8. The positions of the peaks move to lower angles and the volumes of lattice parameters show smooth evolution as the BaF2 content increases, which means that the higher the Ba2+ content is, the larger the lattice parameters are (seen in Fig. S1 ESI) and the Ba2+ should occupy the Ca2+ position. The crystallinity has been improved with the addition of BaF2. When 8 wt% BaF2 is added, the crystallinity of the sample reaches the highest. And the crystallinity begins to decline when the content of BaF2 exceeds 8 wt%.


image file: c4ra04683h-f3.tif
Fig. 3 XRD patterns of the Ca2Si5N8:Eu2+ phosphors with different concentrations of BaF2.

Fig. 4 shows the experimental, calculated, and difference results of Rietveld refinement XRD patterns of Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 at room temperature. The crystal structure of Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 was analyzed by the Materials Studio program on the basis of the XRD data. The pattern factor Rp, and the weighted pattern factor Rwp, are 10.33% and 14.22%, respectively. The XRD patterns of Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 obtained herein indicate that single phase is formed. The Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 synthesized crystallized as a monoclinic structure with the space group of Cc. The refined structure parameters of BaF2 doped Ca2Si5N8:Eu2+ are given in Table 1.


image file: c4ra04683h-f4.tif
Fig. 4 Rietveld refinement results of the XRD patterns of the Ca2Si5N8:Eu2+ with 8 wt% BaF2, including the experimental and calculated intensities as well as differences in intensity between experimental and calculated data.
Table 1 The refined structure parameters of BaF2 doped Ca2Si5N8:Eu2+
Atom Site x/a y/b z/c
Ca1 0.95 −0.00658 0.75426 0.05207
Ba 0.05 −0.00658 0.75426 0.05207
Ca2 1 0.61050 0.73056 0.26069
Si1 1 0.05412 0.79009 0.41769
Si2 1 0.75221 0.19935 0.34537
Si3 1 0.76388 0.51462 0.11491
Si4 1 0.35762 0.21081 0.42527
Si5 1 0.85405 0.02307 0.17204
N1 1 0.94559 0.55880 0.44382
N2 1 0.12258 0.12972 1.08940
N3 1 0.81030 0.25662 0.23980
N4 1 0.79107 0.85548 0.15102
N5 1 0.98494 0.98304 0.27838
N6 1 0.86102 0.17452 1.06034
N7 1 0.62438 0.04133 0.36236
N8 1 0.79600 0.49423 0.41610


When compared the two images of (a) Ca1.95Eu0.05Si5N8 and (b) Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 in Fig. 5, we can find that the addition of the BaF2 in the host is helpful for enhancing the crystallization degree and decreasing surface defects. This indicates that the Ca1.95Eu0.05Si5N8 phosphor has a good dispersion, a regular shape, and the particle size of the synthesized powder was about 6–12 μm. The corresponding EDX spectra analysis (Fig. 5(c) and (d)) and elemental mappings (Fig. 6(a) and (b)) indicates that the products have a chemical composition of Ca, Si, O and N and Ca, Ba, Si, O and N. And the Ba2+ is incorporated into the Ca1.95Eu0.05Si5N8.


image file: c4ra04683h-f5.tif
Fig. 5 SEM and EDX spectra of the (a and c) Ca1.95Eu0.05Si5N8 and (b and d) Ca1.95Eu0.05Si5N8 with 8 wt% BaF2.

image file: c4ra04683h-f6.tif
Fig. 6 Elemental mappings of Ca1.95Eu0.05Si5N8 and Ca1.95Eu0.05Si5N8 with 8 wt% BaF2.

Fig. 7 shows the excitation and emission spectra of Ca1.95Eu0.05Si5N8. It is obvious that all the spectral features of the as-synthesized samples are similar. The excitation spectra consist of three broad bands peaking at about 295, 397 and 467 nm, which mainly arise from the 4f65d1 multiplets of Eu2+ excitation states. And the remarkable enhancement of the emission intensity is observed with increasing the content of BaF2. After incorporating 8 wt% BaF2, the emission intensity reaches twice than that of the sample without BaF2. In addition, it is noticeable that the emission peak of the Eu2+ shifts to longer wavelength (608 nm to 622 nm) with increasing the concentration of BaF2. This would be beneficial to the color point tuning. The excitation, emission, centre of gravity and Stokes shift crystal filed splitting of Ca1.95Eu0.05Si5N8 with different weight of BaF2 are listed in Table 2. The enhancement can be explained by the fluxing agent (BaF2). The redshift could be explained by the fact that the Ca2Si5N8:Eu2+ structure is preserved while a part of the Ca2+ ions are replaced by the larger Ba2+ ions. To accommodate these larger cations, the distance between Ca2+ (or Eu2+) and the anions could not increase or even become slightly smaller, thus leading to the increase of the crystal field splitting, and then causing a red shift of the emission.27


image file: c4ra04683h-f7.tif
Fig. 7 PL/PLE spectra of Ca1.95Eu0.05Si5N8 with different weight of BaF2.
Table 2 Excitation, emission, centre of gravity and Stokes shift crystal filed splitting of Ca1.95Eu0.05Si5N8 with different weight of BaF2
Samples λex (nm) λem (nm) Center of gravity (cm−1) Stocks shift (cm−1)
Without BaF2 295, 397, 467 608 26[thin space (1/6-em)]830 4966
2% BaF2 295, 397, 467 613 26[thin space (1/6-em)]830 5100
4% BaF2 295, 397, 467 615 26[thin space (1/6-em)]830 5153
6% BaF2 295, 397, 467 618 26[thin space (1/6-em)]830 5232
8% BaF2 295, 397, 467 622 26[thin space (1/6-em)]830 5336


Table 3 XPS quantitative results of Ca1.95Eu0.05Si5N8 with and without BaF2
Peak Position BE (eV) Atomic conc.% Mass conc.%
N 402 15.44 10.23
O 536 21.34 16.16
N 402 18.81 12.68
O 536 20.56 15.84


Fig. 8 shows the XPS of Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 and without BaF2. The lattice parameters are greatly affected by the occupation of Eu2+ and Ba2+ ions in the critical structure, which depends on the difference in electronegativity and ionic radii compared with the replaced ions. The two ions have a similar possibility to replace Ca2+ ions and be incorporated into the structure. However, it is well known that the vacancy formation caused by charge imbalance and lattice strain can self-limit the inclusion of guest ions into a host lattice.28 Thus there is a propensity for the ions to migrate to less strained surface sites, rather than incorporate in the crystal lattice, which can be confirmed by XPS data. In Fig. 6, the peak at about 135.6 eV attributed to Eu4d is assigned to Eu2O3 which was formed on the surface of the sample. That means more Eu2+ ions are incorporated into the lattice, lead to emission intensity increasing and red shift of the emission. Another reason is that the N/O ratio is higher in the Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 sample. The nitrogen ion (N3−) has a higher effective charge compared with the oxygen ion (O2−), and the electronegativity of nitrogen (3.04) is smaller than that of oxygen (3.50). Therefore, coordinating with nitrogen would cause a stronger nephelauxetic effect (covalence), the centre of gravity of the 5d states of the activator ions shift to longer wavelength, and the crystal-field splitting larger than that in a similar oxygen environment,29,30 which leads to the red shift of the emission (Table 3).


image file: c4ra04683h-f8.tif
Fig. 8 XPS survey spectrum of Ca1.95Eu0.05Si5N8 with and without BaF2.

Fig. 9(a) shows the crystal structure of Ca2Si5N8 viewed along [010] and Fig. 9(b) depicts the proposed model of substitution of Eu2+ and Ba2+ for Ca2+. Furthermore, a random ion displacement model can be used to clarify the modification of the lattice. This allows the use of an analysis similar to Vegard's law, which is an empirical law that relates the statistical substitution of a guest ion into the host lattice with the experimentally observed degree of lattice change with increasing defect ion concentration. Statistical substitution into a lattice site is predicted to lead to a lattice contraction for smaller ions and a lattice expansion for larger ions. When there are only Eu2+ ions doped in the structure, the cell lattice will be expanded, since the radius of Eu2+ ions is larger than that of Ca2+ ions. A strain may arise in the lattice around the Eu2+ ions, and may limit the stability of the Eu2+ ions that had been incorporated into the lattice. Then, when Ba2+ was doped into the structure, the Ba2+ ions with larger radius than that of Eu2+ could make the lattice expand. So more Eu2+ would incorporate into the lattice because of the larger lattice expended by Ba2+. This means that the structure doped with Ba2+ ions can make more Eu2+ ions incorporate into the lattice.


image file: c4ra04683h-f9.tif
Fig. 9 (a) Crystal structure of Ca2Si5N8 viewed along [010]. (b) The proposed model of substitution of Eu2+ and Ba2+ for Ca2+.

Fig. 10 shows the temperature dependence of the integrated emission intensity for Ca1.95Eu0.05Si5N8 with and without 8 wt% BaF2, which shows an identical thermal stability. It is believed that thermal ionization is responsible for quenching of the luminescence of Eu2+ at high temperatures in Ca2Si5N8 host,17 because the excited 5d electrons are easily ionized by the absorption of thermal energy and entrance into the bottom of the conduction band of the host through the top of the Eu2+ excitation levels. At 200 °C, the integral emission intensity of the both phosphors is about 30% of that measured at room temperature.


image file: c4ra04683h-f10.tif
Fig. 10 Temperature dependent integrated emission intensities of Ca1.95Eu0.05Si5N8 with and without 8 wt% BaF2. (The inset show the emission spectra with increasing temperature).

Fig. 11 represents the Commission International de I'Eclairage (CIE) chromaticity coordinates for Ca1.95Eu0.05Si5N8 with different amounts of BaF2 (0–8 wt%). With increasing the content of Ba2+, the chromaticity coordinates (x, y) vary systematically from (0.556, 0.437) to (0.591, 0.407), corresponding to color points of the samples change gradually from orange-yellow to orange-red. Therefore, it is expected that the white light with good rendering could be obtained when the tunable emission phosphors Ca1.95Eu0.05Si5N8 with different amounts of BaF2 (0–8 wt%) for white LEDs.


image file: c4ra04683h-f11.tif
Fig. 11 CIE chromaticity coordinates of Ca1.95Eu0.05Si5N8 with different BaF2 (0–8 wt%).

4. Conclusions

A remarkable orange-red nitridosilicate phosphor Ca1.95Eu0.05Si5N8 with 8 wt% BaF2 was synthesized by gas-pressed sintering at 1500 °C using the raw materials BaF2, Ca3N2, Si3N4, and Eu2O3. This method promises the use of inexpensive, commercially available and powder handling at ambient pressure, thus offering a simple, efficient, and high-yield way to obtain orange-red emitting phosphors. The color of the emission can be tuned with adding BaF2 into the Ca2Si5N8:Eu2+ host lattice. We have demonstrated that adding BaF2 can make better single phase, emission peak shift to longer wavelength (608 nm to 622 nm) and significantly enhance the emission intensities of Ca2Si5N8:Eu2+ phosphors. This is mainly due to the fact that the lattice structure of Ca2Si5N8:Eu2+ phosphors can be modified and the solubility of the Eu2+ ions can be increased by adding BaF2. More Eu2+ ions would incorporate the lattice because of the larger lattice expended by Ba2+. This novel Ca2Si5N8:Eu2+ with 8 wt% BaF2 phosphor is expected to be useful for phosphor converted white LEDs.

Acknowledgements

This work is supported by Specialized Research Fund for the Doctoral Program of Higher Education (no. 20120211130003) and the National Natural Science Funds of China (Grant no. 51372105).

Notes and references

  1. E. F. Schubert and J. K. Kim, Science., 2005, 308, 1274–1278 CrossRef CAS PubMed.
  2. S. Nakamura, Jpn. J. Appl. Phys., 1991, 30(10A), L1705 CrossRef.
  3. G. Blasse and B. C. Grabmaier, Luminescent Materials, Springer, 1994, vol. 44 Search PubMed.
  4. J. W. H. van Krevel, H. T. Hintzen, R. Metselaar and A. Meijerink, J. Alloys Compd., 1998, 268, 272–277 CrossRef CAS.
  5. R. J. Xie, N. Hirosaki and M. Mitomo, J. Electroceram., 2008, 21, 370–373 CrossRef CAS.
  6. P. Schlotter, J. Baur, C. Hielscher, M. Kunzer, H. Obloh, R. Schmidt and J. Schneider, J. Mater. Sci. Eng. B, 1999, 59, 390–396 CrossRef.
  7. Y. Hu, W. Zhuang, H. Ye, S. Zhang, Y. Fang and X. Huang, J. Lumin., 2005, 111, 139–145 CrossRef CAS PubMed.
  8. J. Ballato, J. S. Lewis III and P. Holloway, Mater. Res. Bull., 1999, 24(9), 51 CAS.
  9. Z. J. Zhang, O. M. ten Kate, A. Delsing, E. van der Kolk, P. H. L. Notten and H. T. Hintzen, J. Mater. Chem., 2012, 22, 9813–9820 RSC.
  10. C. Hecht and W. Schnick, Chem. Mater., 2009, 21(8), 1595–1601 CrossRef CAS.
  11. X. Piao, K. Machida, T. Horikawa, H. Hanzawa, Y. Shimomura and N. Kijima, Chem. Mater., 2007, 19, 4592–4599 CrossRef CAS.
  12. Y. Li, N. Hirosaki, R. Xie, T. Takeda and M. Mitomo, Chem. Mater., 2008, 20, 6704–6714 CrossRef CAS.
  13. X. Piao and K.-i. Machida, J. Lumin., 2010, 13, 08–12 CrossRef PubMed.
  14. H. Nersisyan, H. Il Won and C. W. Won, Chem. Commun., 2011, 47, 11897–11899 RSC.
  15. Y. Q. Li and H. T. Hintzen, J. Solid State Chem., 2008, 181, 515–524 CrossRef CAS PubMed.
  16. X. Piao, K.-i. Machida, T. Horikawa and H. Hanzawa, J. Electrochem. Soc., 2008, 155(1), 17–22 CrossRef PubMed.
  17. X.-D. Wei and Q.-L. Liu, Chin. Phys. B, 2009, 18, 3555–3563 CrossRef CAS.
  18. B. Lei, Y. Shimomura and H. Yamamoto, J. Electrochem. Soc., 2010, 157, J196–J201 CrossRef CAS PubMed.
  19. S. Ho Lee, H. Young Koo, D. Soo Jung, J. Man Han and Y. Chan Kang, Opt. Mater., 2009, 31, 870–875 CrossRef PubMed.
  20. S. Ho Lee, D. Soo Jung, J. Man Han, H. Young Koo and Y. Chan Kang, J. Alloys Compd., 2009, 477, 776–779 CrossRef PubMed.
  21. Y. Q. Li, G. de With and H. T. Hintzen, J. Alloys Compd., 2006, 417, 273 CrossRef CAS PubMed.
  22. R. D. Shannon, Acta Crystallogr., 1976, A32, 751–767 CrossRef CAS.
  23. Y. Gu, Q. Zhang, Y. Li, H. Wang and R. J. Xie, Mater. Lett., 2009, 63, 1448–1450 CrossRef CAS PubMed.
  24. D. L. Dexter and J. H. Schulman, J. Chem. Phys., 1954, 22, 1063 CrossRef CAS PubMed.
  25. L. G. Van Uitert, J. Electrochem. Soc., 1967, 114, 1048–1053 CrossRef CAS PubMed.
  26. H. Zhang, T. Horikawa, H. Hanzawa, A. Hamaguchi and K.-I. Machida, J. Electrochem. Soc., 2007, 154(2), J59–J61 CrossRef CAS PubMed.
  27. V. Bachmann and A. Meijerink, Chem. Mater., 2009, 21, 316–325 CrossRef CAS.
  28. Y. Gu, Q. Zhang, Y. Li, H. Wang and R. J. Xie, Mater. Lett., 2009, 63, 1448–1450 CrossRef CAS PubMed.
  29. Z. Y. Zhao, Z. G. Yang, Y. R. Shi, C. Wang, B. Liu, G. Zhu and Y. H. Wang, J. Mater. Chem. C, 2013, 1, 1407–1412 RSC.
  30. C. W. Yeh, W. T. Chen, R. S. Liu, S. F. Hu, H. S. Sheu, J. M. Chen and H. T. Hintzen, J. Am. Chem. Soc., 2012, 134, 14108–14117 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra04683h

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.