Optical temperature sensing properties of Yb3+–Er3+ co-doped NaLnTiO4 (Ln = Gd, Y) up-conversion phosphors

Dong Hea, Chongfeng Guo*a, Sha Jiangb, Niumiao Zhanga, Changkui Duan*b, Min Yinb and Ting Lia
aNational Key Laboratory of Photoelectric Technology, Functional Materials (Culture Base) in Shaanxi Province, National Photoelectric Technology, Functional Materials & Application of Science and Technology International Cooperation Base, Institute of Photonics & Photon-Technology, Northwest University, Xi'an, 710069, China. E-mail: guocf@nwu.edu.cn; Fax: +86-29-88302661; Tel: +86-29-88302661
bSchool of Physical Science, University of Science and Technology of China, Hefei, 230026, China

Received 4th October 2014 , Accepted 24th November 2014

First published on 27th November 2014


Abstract

Yb3+–Er3+ ion co-doped NaLnTiO4 (Ln = Y, Gd) up-conversion (UC) phosphors were successfully synthesized by a sol–gel method. The phase purity and the structure of the samples were characterized by powder X-ray diffraction (XRD), and the optimal compositions were also determined according to their UC emission intensities. The samples emit orange light and their UC spectra were recorded with excitation by a laser diode with a 980 nm wavelength. The UC luminescence intensity could be enhanced greatly after introducing sensitizer Yb3+ ions, and the energy transfer (ET) from Yb3+ to Er3+ plays a vital role. The UC mechanism and processes responsible for the emissions were investigated and found to involve two-photon absorption. The lifetime of green emission in Er3+ singly doped and Yb3+–Er3+ co-doped samples were measured to prove the existence of ET. The temperature dependence of the fluorescence intensity ratios (FIR) for the two green UC emission bands peaked at 530 and 550 nm was studied in the range of 300–510 K under excitation by a 980 diode laser with about 4 W cm−2 power density, and the maximum sensitivity was approximately 0.0045 K−1 at 510 K for NaYTiO4 and 480 K for NaGdTiO4. This indicates that Yb3+–Er3+ ion co-doped NaLnTiO4 (Ln = Y, Gd) phosphors are potential candidates for optical temperature sensors.


1. Introduction

Temperature is a fundamental and significant physical parameter, which could be accurately measured by many methods. Conventional temperature measurements are based on the principle of liquid and metal expansion, which is realized by heat flow to an invasive probe. However, these contact methods could not be used in many cases, such as the electrical transformer in power stations, oil refineries and intracellular temperature, which promote the development of non-contract thermometry technique. Recently, the non-contract temperature sensing based on rare earth (RE) ions activated luminescent material has been given more attention.1–7 In particular, the fluorescence intensity ratio (FIR) technique is regarded as the most promising, which involves the measurement of photoluminescence intensities from two thermally coupled energy levels (TCLs) of RE ion. FIR technique could offer excellent measurement accuracy because it is independent of spectrum losses and fluctuations in the excitation intensity. Now, optical temperature sensors based on RE activated up-conversion (UC) phosphor have been extensively investigated.8–12 Among these RE ions, Er3+ is the most popular activator because its two thermally coupled levels (TCLs) 2H11/2 and 4S3/2 and characteristic green UC emission from TCLs. However, only low luminescence efficiency is obtained in Er3+ solely doped phosphor under the 980 nm laser excitation due to the low absorption. The Er3+ and Yb3+ are usually used in couples, which could improve the luminescence efficiency of UC phosphor remarkably due to the high and broad absorption (ranging from 850–1050 nm) of Yb3+ and efficiently transfer its energy to Er3+.13–17

The layered perovskite titanates complex oxides ALnTiO4 (A = Li, Na, K; Ln = rare earth) have been proven to serve as excellent phosphor hosts due to their typical two-dimensional structures and good chemical and physical stabilities.18–20 They usually show higher critical concentration than those of other conventional inorganic phosphors because the energy transfer is restricted to quasi-two-dimensional sub-lattice in present compounds.21 Recently, a series of Eu3+ doped red emitting phosphors ALnTiO4 (A = Na, K; Ln = La, Gd, Y):Eu3+ have been prepared by sol–gel method in our group, results indicate that they show high luminescence efficiencies and could be used as red components in white light-emitting diodes (w-LEDs) and field emission displays (FEDs).22,23 However, very little research on UC phosphors using ALnTiO4 (A = Li, Na, K; Ln = rare earth) as hosts has been carried out.24

In this paper, UC phosphors based on Yb3+ and Er3+ doped NaLnTiO4 (Ln = Y, Gd) were prepared by sol–gel method and their UC luminescence properties were investigated. The UC emission mechanisms were evaluated, and the energy transfer processes between Yb3+ and Er3+ were proved through the measurements of lifetime. Additionally, their thermometry behaviors have also been illustrated by FIR technique.

2. Experimental

2.1 Synthesis of samples

The Yb3+ sensitized Er3+ doped NaLnTiO4 (Ln = Y, Gd) phosphors with composition NaLn1−xyYbxEryTiO4 were prepared using a modified sol–gel method. In present systems, Yb3+ and Er3+ ions are expected to occupy the sites of Ln3+ due to their identical valence and similar ionic radius. Here, the starting materials include analytical reagent (AR) Na2CO3, rare earth oxides Y2O3, Gd2O3, Yb2O3, Er2O3 with high purity (99.99%, Shanghai Yuelong Nonferrous Metals Ltd.) and analytical reagent tetrabutyl titanate Ti(OC4H9)4, which are used as for the sources of Na, Y (or Gd), Yb, Er and Ti, respectively. In addition, anhydrous ethanol and acetic acid played the role of solvent and hydrolysis inhibitor in this experiment. According to the chemical formula of the UC phosphors NaLn1−xyYbxEryTiO4: (a) x = 0.1, y = 0.01, 0.03, 0.05, 0.07, 0.09, and (b) y = 0.05, x = 0.06, 0.08, 0.10, 0.12, 0.14, 0.16, 0.18, 0.20, a stoichiometric amount of oxides Ln2O3 (Ln = Y, Gd, Yb and Er) and Na2CO3 (excess 30% as flux) were first dissolved in dilute HNO3(AR) under continuous stirring, and the excess HNO3 was evaporated at high temperature. Then, the required amount of deionized water was added to get transparent solution A after fully stirring. The calculated volume of Ti(OC4H9)4, ethanol and acetic acid were mixed to form the transparent solution B. Subsequently, solution B was transferred to solution A drop by drop under constant stirring to get the final transparent solution. The final transparent solution was kept at 70 °C for 24 h until an ivory white resin formed. The resin was further dried at 120 °C for 24 h to get a dried gel. The dried gel was ground and preheated in a furnace at 500 °C for 5 h, and then sintered at required temperature at 900 °C for 2 h to obtain final samples.

2.2 Characterization

The crystal structures were identified by powder X-ray diffractions (XRD) using a Rigaku-Dmax 3C powder diffractometer (Rigaku Corp, Tokyo, Japan) with Cu-Ka radiation (λ = 1.5406 Å). UC luminescence spectra of the phosphors at room temperature (RT) were recorded using a Hitachi F-4600 spectrophotometer (Hitachi high technologies corporation, Tokyo, Japan) with an external power-controllable 980 nm semiconductor laser (Hi-Tech Optoelectronics Company, Beijing, China) as excitation source. The lifetimes for 4S3/24I15/2 transitions of Er3+ at 547 nm at RT were carried out using an 980 nm optical parametric oscillator (OPO) pulsed laser as excitation source, and the signals were detected by a Tektronix digital oscilloscope (TDS 3052). In order to investigate the temperature dependence of the UC emission, the sample was placed in a temperature-controlled copper cylinder, and its temperature was increased from RT to 240 °C. The UC spectra of sample at various temperatures were obtained using a Jobin-Yvon HRD-1 double monochromator equipped with a Hamamatsu R928 Photomultiplier under the excitation of a 980 nm diode laser with 164 mW and the excitation power density was about 4W cm−2. The signal was analyzed by an EG&G 7265 DSP Lock-in amplifier and stored into computer memories.

3. Results and discussion

3.1 Crystal structure and phase composition

The structures and compositions of the obtained products were characterized by XRD. Fig. 1 shows the XRD patterns of NaY1−xyYbxEryTiO4 and NaGd0.81Yb0.14Er0.05TiO4 phosphors. XRD patterns of NaY0.90−yYb0.10EryTiO4 and NaY0.95−xYbxEr0.05TiO4 samples as a function of Er3+ concentrations y and Yb3+ concentrations x are shown in Fig. 1a, it is observed that all diffraction peaks of samples can readily be identified as the characteristic peaks of pure orthorhombic phase NaYTiO4 with JCPDS standard card no. 50-0022. Results indicate that no noticeable diffraction peaks from any impurity phases are found within the whole range of our experiments, which illuminates that the parameters to synthesis of the present samples are fit. In addition, it is found that the positions of the strongest diffraction peak (113) gradually shift to high angle, which due to the smaller radius of Yb3+ (0.86 Å) or Er3+ (0.88 Å) ions than that of Y3+ (0.89 Å).25 As a representative, XRD pattern of NaGd0.81Yb0.14Er0.05TiO4 phosphor is shown in Fig. 1b with the standard NaGdTiO4 (JCPDS 86-0830) profile, all of their diffraction peaks are matched well. Above results prove that Yb3+ or Er3+ ions occupy the sites of Y3+ and do not alter the host structure significantly.
image file: c4ra11771a-f1.tif
Fig. 1 XRD patterns of NaLnTiO4 (Ln = Y, Gd) phosphors: (a) NaY1−xYbxEryTiO4 phosphors and (b) NaGd0.81Yb0.14Er0.05TiO4 phosphors annealed at 900 °C for 2 h in air.

3.2 Effects of do-pant contents on up-conversion emission

It is well known that the UC emission intensity greatly depends on the concentrations of the doped ions, thus we investigated the effect of dopant contents on NaY1−xyYbxEryTiO4: xYb3+, yEr3+ phosphors with the continue wave (CW) excitation at 980 nm. Room temperature UC spectra of samples as function of Er3+ and Yb3+ contents are displayed in Fig. 2a and b, respectively. The obtained emission spectra of Er3+/Yb3+ codoped NaYTiO4 UC phosphors consist of two green emission (530 and 550 nm) bands and a red emission (668 nm) band, which are assigned to 2H11/24I15/2, 4S3/24I15/2 and 4F9/24I15/2 the characteristic inner shell transitions of Er3+ from the excited levels to the ground state of 4f electronic configuration, respectively. The normalized integrated intensities for green (from 510 to 570 nm) and red (from 640 to 690 nm) UC emissions were also calculated and presented in the insets of Fig. 2a and b with variable Er3+ or Yb3+ concentrations for the two series of samples, respectively. From the inset of Fig. 2a, it is found that the UC luminescence intensities for both the green and red emissions increase first with increasing Er3+ contents and then decrease with the continual increase of Er3+ concentration after reaching the maximum at y = 0.05 due to the concentration quenching, which is caused by non-radiative energy transfer and the cross relaxation between Er3+.26 Fig. 2b shows the UC spectra of samples NaY1−x−0.05YbxEr0.05TiO4: xYb3+, 0.05Er3+ with invariable Er3+ concentration y = 0.05 and variable Yb3+ concentrations ranging from 0 to 0.20, which shows similar trend as Fig. 2a. Seen from Fig. 2b, it is observed that the UC emission of sample NaY0.95Er0.05TiO4 without Yb3+ is too weak to be checked due the weak absorption of Er3+ for 980 nm laser, and increases remarkably with the growth of Yb3+ dosage due to the efficient energy transfer from Yb3+ to Er3+ for the larger absorption cross section of Yb3+ at 980 nm and excellent energy levels match between Yb3+ and Er3+.27 Results indicate that the optimal concentration of Yb3+ is x = 0.14, and the UC emission intensity descends gradually as the concentration of Yb3+ beyond 0.14. The main reasons for this may come from the energy back transfer (EBT) Er3+ (4S3/2) + Yb3+ (2F7/2) → Er3+ (4I13/2) + Yb3+ (2F5/2).28,29 The higher Yb3+concentration, the higher ratio of EBT from Er3+ to Yb3+ to ET from Yb3+ to Er3+.24
image file: c4ra11771a-f2.tif
Fig. 2 UC luminescence spectra of NaY1−xyYbxEryTiO4: xYb3+, yEr3+ (a) x = 0.10; y = 0.01, 0.03, 0.05, 0.07, 0.09 and (b) y = 0.05; x = 0, 0.06, 0.08, 0.10, 0.12, 0.14, 0.16, 0.18, 0.20 under excitation at 980 nm.

3.3 Energy scheme and mechanism of UC emission

In order to better understand the possible UC processes and the mechanisms of the green and red emission in Yb3+/Er3+ co-doped NaLnTiO4 (Ln = Y, Gd) phosphors, the energy level diagram for Er3+ and Yb3+, feasible excitation pathways, emission process and UC mechanisms are demonstrated in Fig. 3. The possible up-conversion populating processes of the Er3+ and Yb3+ co-doped samples under the excitation of 980 nm are shown in Fig. 3a. As drawn in the schematic, the energy level 2F5/2 of Yb3+ is close to that of 4I11/2 of Er3+, thus the energy transfer from Yb3+ to Er3+ is high efficient. Under the excitation of 980 nm LD, Yb3+ first absorbs an infrared photon from the ground state transition 2F7/2 to the excited state 2F5/2. The excitation wavelength (980 nm) is in resonance with the transition of 2F7/22F5/2 of Yb3+ and thus easy to be absorbed by this ion. For the green emission, the peaks at 530 and 550 nm come from the excited states 2H11/2/4S3/2 to ground state 4I15/2 transitions of Er3+, and the 4S3/2 level could be populated by non-radiative (NR) relaxation from the upper 2H11/2 level populated through NR process from 4F7/2 level of Er3+. The population 4F7/2 level may follow three process: Er3+ (4I11/2) + a photon (980 nm) → Er3+ (4F7/2) (excited state absorption: ESA), Er3+ (4I11/2) + Yb3+ (2F5/2) → Er3+ (4F7/2) + Yb3+ (2F7/2) (energy transfer, ET),and Er3+ (4I11/2) + Er3+ (4I11/2) → Er3+ (4F7/2) + Er3+ (4I15/2) (cross relaxation: CR). Here, the intermediary level 4I11/2 of Er3+ play an vital role in these processes, and population process of this energy includes the energy transfer (ET) process Er3+ (4I15/2) + Yb3+ (2F5/2) → Er3+ (4I11/2) + Yb3+ (2F7/2) and the ground state absorption (GSA) process from 4I15/2 to 4I11/2. However, the probability for the GSA of Er3+ is low due to the narrow absorption cross section, which illuminates that the ET process dominated the UC process.30 For the red emission assigned to 4F9/24I15/2 transitions of Er3+, the population of 4F9/2 could involve following these processes: Er3+(4I13/2) + a photon (980 nm) → Er3+(4F9/2) (ESA), Yb3+(2F5/2) + Er3+(4I13/2) → Yb3+(2F7/2) + Er3+(4F9/2) (ET), and the NR from 4S3/2 to 4F9/2. The long-lived intermediary 4I13/2 level is populated by NR relaxation from 4I11/2 level.
image file: c4ra11771a-f3.tif
Fig. 3 (a) Energy level diagram of the Yb3+, Er3+ ions and the proposed UC mechanisms in Er3+–Yb3+ co-doped NaYTiO4 under 980 nm excitation; (b) and (c) dependence of UC luminescence intensity of NaY0.81Yb.0.14Er0.05TiO4 and NaGd0.81Yb0.14Er0.05TiO4 upon the pumping power, respectively.

For a multi-photon UC process, the UC luminescence intensity (I, integrated intensity) depends on the nth the pump laser powder (P), i.e., IPn, here n denotes the number of absorbed infrared (IR) photons to produce an UC emission photon31 and it is important for the determination of UC mechanism. It is known that which could be assigned to the slope of the straight line that obtained through the plot of ln[thin space (1/6-em)]I versus ln[thin space (1/6-em)]P. Fig. 3b and c shows the experimental data and fitting curves reflecting the relationship between the integrated UC intensities and pump power for the samples NaY0.81Yb0.14Er0.05TiO4 and NaGd0.81Yb0.14Er0.05TiO4 with optimal composition, respectively. As shown in Fig. 3b and c, it could find that the slopes of the both curves for the green peak are 2.05 and 1.99 for Ln = Y and Gd, and the slopes are 1.87 and 1.83 for the red emission, which are very close to 2.0 within the margin of error. Results indicate that the green 4S3/24I15/2 and red 4F9/24I15/2 transition in Yb3+/Er3+ co-doped NaLnTiO4 (Ln = Y, Gd) are two-photon absorption process.

3.4 Lifetime measurements

According to above mentioned possible UC processes, the enhanced UC intensities with the addition of Yb3+ are attributed to the efficient energy transfer from Yb3+ to Er3+, and ET plays a dominant role in this process. In order to prove the existance of the ET between Yb3+ and Er3+, the emission decay curves of the green (4S3/24I15/2, 550 nm) for the samples NaY0.95−xYbxEr0.05TiO4: 0.05Er3+, xYb3+ (x = 0, 0.08.) with and without sensitizer Yb3+ were shown in Fig. 4 after 980 nm pulsed laser excitation, in which the decay profiles of the 4S3/24I15/2 (green) transitions are normalized. The lifetime could be calculated according to eqn (1):32
 
image file: c4ra11771a-t1.tif(1)
where I(t) is fluorescence intensity at time t. The decay curve can be well fitted with quadratic index equation, and the average lifetime can obtained through the following equation:33
 
image file: c4ra11771a-t2.tif(2)

image file: c4ra11771a-f4.tif
Fig. 4 The temporal evolution of the green 4S3/24I15/2 emissions in NaY0.95−xYbxEr0.05TiO4: 0.05Er3+, xYb3+ (x = 0, 0.08).

The corresponding lifetimes were listed in Fig. 4, it could be found that the lifetime for Yb3+–Er3+ co-doped NaY0.87Yb0.08Er0.05TiO4 is larger (84 μs) than that (36 μs) of Er3+ solely doped sample NaY0.95Er0.05TiO4, which confirms the presence of ET from Yb3+ to Er3+ in present system.

3.5 Temperature sensing properties

As already mentioned, UC bands centered at 530 and 550 nm arise from 2H11/24I15/2 and 4S3/24I15/2 transitions. The energy gap between the levels 2H11/2 and 4S3/2 of Er3+ is around 750 cm−1, the state of 2H11/2 may also be populated from 4S3/2 by thermal excitation and the UC emission intensity ratio of emission band at 530 to 550 could change with the variable temperature, therefore it is could be used as optically temperature sensor for the present UC phosphors. In order to investigate the temperature sensing properties of synthesized phosphors, the green UC emission spectra for samples NaLnTiO4: 0.05Er3+, 0.14Yb3+ (Ln = Y, Gd) at different temperature (from 300 to 510 K) were recorded and displayed in Fig. 5a and b, in which the spectra are normalized to the most intense emission peak at about 551 nm. In present cases, the excitation power density was estimated to be around 4W cm−2, which is low enough to neglect the heating effect from the pumping source. It is found that no remarkable shift in emission wavelength for two samples while their fluorescence intensity ratio (FIR) of UC emission from 2H11/24I15/2 to 4S3/24I15/2 (550 nm) increases with the rise of temperature. The relative population of the thermally coupled energy levels follows the Boltzmann distribution and the FIR of two emissions can be written as follows:34,35
 
image file: c4ra11771a-t3.tif(3)
where IH and IS are intensities (the integrated areas below the emission curves) for the upper (2H11/24I15/2) and lower (4S3/24I15/2) thermally coupled levels transitions, respectively. N, g, σ, ω are the number of ions, the degeneracy, the emission cross-section, the angular frequency of fluorescence transitions from the 2H11/2 and 4S3/2 levels to the 4I15/2 level, respectively. K is the Boltzmann constant, and ΔE is the energy gap between the 2H11/2 and 4S3/2 levels, T is absolute temperature and C is proportionality constant.

image file: c4ra11771a-f5.tif
Fig. 5 Temperature dependence of the green UC luminescence spectra of NaYTiO4: 0.14Yb3+, 0.05Er3+ (a) and NaGdTiO4: 0.14Yb3+, 0.05Er3+ (b) under 980 nm excitation (the spectra are normalized to the most intense emission peak at about 551 nm).

Fig. 6a and b present the behaviors of FIR (IH/IS) for samples NaYTiO4: 0.14Yb3+, 0.05Er3+ and NaGdTiO4: 0.14Yb3+, 0.05Er3+ as a function of temperature, in which the fitting results analysized using eqn (3) are also included. It is found that the monolog of experimental FIR data are fitted as a straight line on the inverse of temperature from Fig. 6a. The slopes of two lines are about 1079 and 1061, and the energy gap ΔE could be calculated to about 783 and 765 cm−1 in NaYTiO4 and NaGdTiO4, which is close to previous results in the range of 750–800 cm−1.36,37 The dependences of FIR on temperature are showed in Fig. 6b, the values of the coefficient C is about 8.55 and 8.85 by the best fit curves.


image file: c4ra11771a-f6.tif
Fig. 6 Temperature depentent FIR (a), (b) and relative sensitivities (c) for NaYTiO4:Er3+/Yb3+ (left column) and NaGdTiO4:Er3+/Yb3+ (right column) phosphors.

The sensor sensitivity is an important reference parameter, which is defined as the rate of the FIR (R) changes with temperature and expressed using the following formula:38

 
image file: c4ra11771a-t4.tif(4)

Fig. 6c shows the sensitivity (S) as a function of absolute temperature. With the increase of temperature, the sensitivity of Er3+–Yb3+ co-doped NaYTiO4 reached the maximum about 0.0045 K−1 at 510 K, but the sensitivity for Er3+–Yb3+ co-doped NaGdTiO4 keep a constant after reach at maximum 0.0045 K−1 at temperature of 480 K. It indicates that Yb3+–Er3+ ions co-doped NaLnTiO4 (Ln = Y, Gd) phosphors could be used as potential candidates for optical temperature sensors based on the FIR technique.

4. Conclusions

The Yb3+–Er3+ ions co-doped NaLnTiO4 (Ln = Y, Gd) up-conversion (UC) phosphors were prepared by sol–gel method, and the optimal preparation parameters for NaLnTiO4 (Ln = Y, Gd) were also determined as 0.05Er3+, 0.14Yb3+ at 900 °C. Under 980 nm excitation, samples emit orange light and their UC spectra are composed of prominent green emission centered at 530, 550 and red emission peaked at 668 nm originating from 2H11/24I15/2, 4S3/24I15/2 and 4F9/24I15/2 transitions of Er3+ ion, respectively. The lifetimes of green emission centered at 550 nm, possible UC processes and mechanism investigations indicate that emission are two-photon processes and the efficient energy transfer from Yb3+ to Er3+ plays an important role in UC process. The dependences of FIR for the samples NaLnTiO4: 0.05Er3+, 0.14Yb3+ (Ln = Y, Gd) with optimal compositions on temperature were measured in the range of 300–510 K, and the sensitivities of samples reach the same maximum 0.0045 K−1 for NaYTiO4 Er3+–Yb3+ at 510 K and NaGdTiO4:Er3+–Yb3+ at 480 K, respectively. The results illuminate that the present UC phosphors could be used as promising candidates for optical temperature sensors.

Acknowledgements

This work was supported by the high-level talent project of Northwest University, National Natural Science Foundation of China (no. 11274251,11274299, 11374291), Ph.D. Programs Foundation of Ministry of Education of China (20136101110017), Technology Foundation for Selected Overseas Chinese Scholar, Ministry of Personnel of China (excellent), Natural Science Foundation of Shaanxi Province (no. 2014JM1004) and Foundation of Key Laboratory of Photoelectric Technology in Shaanxi Province (12JS094).

References

  1. Z. Boruc, M. Kaczkan, B. Fetlinski, S. Turczynski and M. Malinowski, Opt. Lett., 2012, 37, 5214–5216 CrossRef CAS PubMed.
  2. C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millán, V. S. Amaral, F. Palacio and L. D. Carlos, Nanoscale, 2013, 5, 7572–7580 RSC.
  3. C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millán, V. S. Amaral, F. Palacio and L. D. Carlos, New J. Chem., 2011, 35, 1177–1183 RSC.
  4. S. Chattopadhyay, P. Sen, J. T. Andrews and P. K. Sen, J. Appl. Phys., 2012, 111, 034310 CrossRef PubMed.
  5. F. Vetrone, R. Naccache, A. Zamarrón, A. J. Fuente, F. Sanz-Rodrıguez, L. M. Maestro, E. M. Rodriguez, D. Jaque, J. G Solé and J. A. Capobianco, ACS Nano, 2010, 4, 3254–3258 CrossRef CAS PubMed.
  6. J. M. Yang, H. Yang and L. Lin, ACS Nano, 2011, 5, 5067–5071 CrossRef CAS PubMed.
  7. C. D. S. Brites, P. P. Lima, N. J. O. Silva, A. Millán, V. S. Amaral, F. Palacio and L. D. Carlos, Adv. Mater., 2010, 22, 4499–4504 CrossRef CAS PubMed.
  8. C. Joshia, A. Dwivedib and S. B. Rai, Spectrochim. Acta, Part A, 2014, 129, 451–456 CrossRef PubMed.
  9. D. Li, Y. Wang, X. Zhang, K. Yang, L. Liu and Y. Song, Opt. Commun., 2012, 285, 1925–1928 CrossRef CAS PubMed.
  10. W. Xu, X. Gao, L. Zheng, Z. Zhang and W. Cao, Sens. Actuators, B, 2012, 173, 250–253 CrossRef CAS PubMed.
  11. B. Dong, B. Cao, Y. He, Z. Liu, Z. Li and Z. Feng, Adv. Mater., 2012, 24, 1987–1993 CrossRef CAS PubMed.
  12. N. Rakov and G. S. Maciel, Opt. Lett., 2014, 39, 3767–3769 CrossRef CAS PubMed.
  13. S. Zhou, K. Deng, X. Wei, G. Jiang, C. Duan, Y. Chen and M. Yin, Opt. Commun., 2013, 291, 138–142 CrossRef CAS PubMed.
  14. H. Zheng, B. Chen, H. Yu, J. Zhanga, J. Suna, X. Li, M. Sun, B. Tian, S. Fu, H. Zhong, B. Dong, R. Hua and H. Xi, J. Colloid Interface Sci., 2014, 420, 27–34 CrossRef CAS PubMed.
  15. M. Quintanilla, E. Cantelar, F. Cussó, M. Villegas and A. C. Caballero, Appl. Phys. Express, 2011, 4, 022601 CrossRef.
  16. Y. Yang, C. Mi, F. Yu, X. Su, C. Guo, G. Li, J. Zhang, L. Liu, Y. Liu and X. Li, Ceram. Int., 2014, 40, 9875–9880 CrossRef CAS PubMed.
  17. N. Rakova and G. S. Maciel, Sens. Actuators, B, 2012, 164, 96–100 CrossRef PubMed.
  18. G. Blasse and A. Bril, J. Chem. Phys., 1968, 48, 3652–3656 CrossRef CAS PubMed.
  19. A. E. Lavat and E. J. Baran, J. Alloy Compd., 2006, 419, 334–336 CrossRef CAS PubMed.
  20. H. Zhong, X. Li, R. Shen, J. Zhang, J. Sun, H. Zhong, L. Cheng, Y. Tian and B. Chen, J. Alloy Compd., 2012, 517, 170 CrossRef CAS PubMed.
  21. K. Toda, Y. Kameo, M. Ohta and M. Sato, J. Alloy Compd., 1995, 218, 228–232 CrossRef CAS.
  22. N. Zhang, C. Guo and H. Jing, RSC Adv., 2013, 3, 7495–7502 RSC.
  23. N. Zhang, C. Guo, J. Zheng, X. Su and J. Zhao, J. Mater. Chem. C, 2014, 2, 3988–3994 RSC.
  24. Y. Jiang, R. Shen, X. Li, J. Zhang, H. Zhong, Y. Tian, J. Sun, L. Cheng, H. Zhong and B. Chen, Ceram. Int., 2012, 38, 5045–5051 CrossRef CAS PubMed.
  25. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751–767 CrossRef.
  26. J. Li, J. Sun, J. Liu, X. Li, J. Zhang, Y. Tian, S. Fu, L. Cheng, H. Zhong, H. Xia and B. Chen, Mater. Res. Bull., 2013, 48, 2159–2165 CrossRef CAS PubMed.
  27. T. Li, C. Guo, Y. Wu, L. Li and J. H. Jeong, J. Alloy Compd., 2012, 540, 107–112 CrossRef CAS PubMed.
  28. I. Etchart, A. Huignard, M. Bérard, M. N. Nordin, I. Hernández, R. J. Curry, W. P. Gillin and A. K. Cheetham, J. Mater. Chem., 2010, 20, 3989–3994 RSC.
  29. G. Chen, G. Somesfalean, Y. Liu, Z. Zhang, Q. Sun and F. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 195204 CrossRef.
  30. T. Li, C. Guo and L. Li, Opt. Express, 2013, 21, 18281–18289 CrossRef PubMed.
  31. J. C. Boyer, F. Vetrone, L. A. Cuccia and J. A. Capobianco, J. Am. Chem. Soc., 2006, 128, 7444–7445 CrossRef CAS PubMed.
  32. H. Jing, C. Guo, G. Zhang, X. Su, Z. Yang and J. H. Jeong, J. Mater. Chem., 2012, 22, 13612–13618 RSC.
  33. T. Fujii, K. Kodaira, O. Kawauchi, N. T. H. Yamashita and M. Anpo, J. Phys. Chem. B, 1997, 101, 10631–10637 CrossRef CAS.
  34. M. D. Shinn, W. A. Sibley, M. G. Drexhage and R. N. Brown, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 27, 6635–6648 CrossRef CAS.
  35. L. Li, C. Guo, S. Jiang, D. K. Agrawal and T. Li, RSC Adv., 2014, 4, 6391–6396 RSC.
  36. X. Wang, X. Kong, Y. Yu, Y. Sun and H. Zhang, J. Phys. Chem. C, 2007, 111, 15119–15124 CAS.
  37. S. A. Wade, S. F. Collins and G. W. Baxter, J. Appl. Phys., 2003, 94, 4743–4756 CrossRef CAS PubMed.
  38. V. K. Rai, A. Pandey and R. Dey, J. Appl. Phys., 2013, 113, 083104 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2015