Enhanced red emission in Ca0.5La(MoO4)2:Eu3+ phosphors for UV LEDs

Yurong Shi*a, Bitao Liub, Bo Liua, Chunyang Lia and Zhenling Wanga
aThe Key Laboratory of Rare Earth Functional Materials and Applications, Zhoukou Normal University, Zhoukou 466001, P. R. China. E-mail: shiyr09@sina.com
bDepartment of Research Center for Materials Interdisciplinary Science, Chongqing University of Arts and Sciences, Chongqing, China

Received 13th October 2015 , Accepted 30th October 2015

First published on 30th October 2015


Abstract

La3+ was introduced into CaMoO4 to form a novel compound with the chemical formula Ca0.5La(MoO4)2 and Eu3+ activated the new compound was synthesized by a solid-state reaction method. An enhanced red emission at 615 nm originating from the 5D07F2 transition of Eu3+ was reported in the Ca0.5La(MoO4)2:Eu3+ phosphor. XRD patterns indicate a structural similarity to CaMoO4. The red emission intensity of the Ca0.5La(MoO4)2:Eu3+ phosphor is 5.92 times higher than that of CaMoO4:Eu3+. Photoluminescent spectra, diffuse reflection spectra and lifetime measurements indicate that the red emission enhancement is due to the increase of the 7F05L6 absorption intensity of Eu3+ in the distorted crystal field of the Ca0.5La(MoO4)2:Eu3+ phosphor. Enhanced red emission, stronger 7F05L6 absorption intensity, and high color purity indicate that the Ca0.5La(MoO4)2:Eu3+ phosphor is a promising red phosphor for UV LEDs. The luminescence properties of the Bi3+ co-doped Ca0.5La(MoO4)2:Eu3+ phosphor was also investigated. Although Bi3+ co-doped into the Ca0.5La(MoO4)2:Eu3+ phosphor broadens PLE spectra in the UV range and makes it more suitable for UV pumped LEDs, it decreases the distortion and induces a new quenching path, and thus makes the emission intensity decrease.


1. Introduction

With the development of photoelectric technology and materials science, white-light emitting diodes (w-LED) have been regarded as a fourth generation lighting source because of their important benefits including energy saving, safety, reliability, long operation time and environmentally friendly. Among the many ways of achieving white light, phosphor-converted (pc) white LEDs are the main method due to their high efficiency, low cost and easy preparation.1 Usually, there are two methods to obtain white light in pc-LEDs. One is using a blue LED chip combined with orange or yellow-red phosphors.2,3 The other is using a (near) ultraviolet (NUV/UV) LED chip combined with blue-green-red phosphors.4,5 The yellow, blue and green phosphors are commercially available and could meet the requirements. However, the red phosphors have many drawbacks, such as unstable,6,7 high cost,8,9 hard to prepare, re-absorption and low efficiency.10 Lacking of red phosphors limits the scope of w-LEDs applications, such as light sources in offices, schools, medicals, hospitals and hotels. No matter what kind of method to achieving white light, exploring red phosphors are necessary. Thus, it is important to obtain red phosphors with high quantum efficiency and good color saturation for blue and UV LEDs.

Eu3+ activated materials are of particular interest because the lowest excited level 5D0 of the 4f6 configuration is situated below the 4f55d configuration for Eu3+.11 Eu3+ activated luminescence materials mainly show sharp line at red region peaked around 612 nm. Though Eu3+ shows high 7F05L6 transitions at about 395 nm in most hosts, the transition is parity-forbidden, so it is a narrow line and cannot absorb the excitation energy efficiently, especially for absorption LED chip's emission.

Among of many inorganic hosts, molybdate compounds have attracted many attentions owing to their peculiar spectroscopic properties and potential application as laser and luminescent hosts.12,13 Molybdate (MoO42−) has a tetrahedral symmetry (Td) with the central Mo6+ coordinated by four O2− ions. Molybdate phosphors have broad and intense absorption bands due to charge transfer (CT) from O2− to Mo6+ in the UV region,14,15 which may increase the absorption for UV LED chip. Besides, molybdates are chemically stable, which is better than oxysulfide and sulfide red-emitting phosphors, such as Y2O2S:Eu3+,6 CaS:Eu2+.7 Molybdates are also easier synthesized with low cost than nitride and oxynitride, such as Sr2Si5N8:Eu2+,8 CaSiAlN3:Eu2+,9 which require critical preparation conditions like high temperature, high pressure and expensive raw materials, and so on.

Eu3+ activated molybdates have been widely prepared as red phosphors for UV LEDs.16–18 The most popular molybdate host is CaMoO4. Eu3+ activated CaMoO4 can be excited by UV light and emit highly red light. However, there is charge mismatch between Eu3+ and Ca2+ in CaMoO4, causing lower solubility saturation of activator and luminescence intensity. Thus, it is better to use charge compensators to enhance the intensity. For example, Li+, Na+ or K+ was included into CaMoO4 host, and the emission intensity was greatly enhanced.19 Besides, Bi3+ was also used to enhance the red emission by increasing transition probability from the ground state to 5L6 and 5D2 states of Eu3+, as the addition of Bi3+ broke the symmetry of CaMoO4 and distorted its crystal structure.11

In this paper, La3+ is introduced into CaMoO4 to form a novel molybdate host as Ca0.5La(MoO4)2. On the one hand, La3+ can provide a proper crystal site for Eu3+ doping, on the other hand, the addition of La3+ will distort the crystal and decrease the symmetry. Thus, enhanced red emission can be expected in Ca0.5La(MoO4)2:Eu3+ phosphor. Besides, the effect of Bi3+ co-doped Ca0.5La(MoO4)2:Eu3+ phosphor is also studied.

2. Experimental

Ca0.5La1−xEux(MoO4)2 (0.1 ≤ x ≤ 1) (Ca0.5La(MoO4)2:xEu3+) and Ca0.5La0.4−yBiyEu0.6(MoO4)2 (0 ≤ y ≤ 0.2) (Ca0.5La(MoO4)2:0.6Eu3+, yBi3+) phosphors were synthesized by solid state reaction. Starting materials were CaCO3, La2O3, Eu2O3, MoO3 and Bi(NO3)3 with a purity of A. R. The stoichiometric raw materials were ground in an agate mortar and heated to 950 °C in air for 4 h. Then the as-synthesized samples were slowly cooled to room temperature. CaMoO4:0.24Eu3+ (best doping concentration11,20) was synthesized using CaCO3, Eu2O3 and (NH4)6Mo7O24·4H2O as raw materials and calculated at 800 °C in air for 6 h. The commercial Y2O3:Eu3+ (com-Y2O3:Eu3+) red phosphor (GS-Y-R) was purchased from Gansu Rare Earth New Materials Co., Ltd.

The phase structure of the obtained samples was characterized by a Bruker D8 Advance X-ray diffraction (XRD) with Cu Kα radiation (λ = 0.15406 nm). The accelerating voltage and emission current were 40 kV and 40 mA, respectively. The excitation (PLE)/emission (PL) spectra and lifetime were recorded using a FLS920P Edinburgh Analytical Instrument apparatus equipped with a 450 W xenon lamp and a μF900H high-energy micro-second flash lamp as the excitation sources. The diffuse reflection spectra of as-synthesized samples were obtained by an UV-vis spectrophotometer (PE lambda 950) using BaSO4 as a standard reference in the measurements. All of the measurements were done at room temperature.

3. Results and discussion

3.1 X-ray diffraction analysis of Ca0.5La(MoO4)2:xEu3+ phosphors

Fig. 1 shows XRD patterns of Ca0.5La(MoO4)2:xEu3+ and synthesized CaMoO4:0.24Eu3+ phosphors. All XRD profiles are well matched with the PDF card no. 29-0351 (CaMoO4). No detectable impurity phase is observed in the obtained samples even in Ca0.5Eu(MoO4)2 (x = 1). This indicates that all obtained samples have a scheelite (CaMoO4) structure and all of them are single phased with the space group of I41/a(88). In Fig. 1a, we observe the highest intensity peak around 28° (2θ) marked in green. It is obviously observed in the expanded version of the XRD patterns in Fig. 1b. The peak of Ca0.5La(MoO4)2:0.2Eu3+ shifts to smaller 2θ value compared with that of CaMoO4:0.24Eu3+ due to larger ionic radii of La3+ (r = 1.16 Å when the coordination number (CN) = 8) than that of Ca2+ (r = 1.12 Å when CN = 8).21 Based on the above phenomena and analysis, it is expected that the new scheelite structure is most probably formed by one Ca2+, two La3+ and one Ca2+ vacancy (VCa) in Ca0.5La(MoO4)2 substitution four Ca2+ sites in CaMoO4, described by:
Ca2+ + 2La3+ + VCa → 4Ca2+

image file: c5ra21208a-f1.tif
Fig. 1 XRD patterns of Ca0.5La(MoO4)2:xEu3+ and synthesized CaMoO4:0.24Eu3+ phosphors.

In addition, Fig. 1b shows that as the content of Eu3+ increase the peaks shift to higher 2θ value. This is due to the smaller ionic radii of Eu3+ (r = 1.07 Å when CN = 8) than La3+ (r = 1.16 Å when CN = 8),21 which proves that Eu3+ ions have been successfully occupied the La3+ sites in the Ca0.5La(MoO4)2 host lattice.

3.2 Luminescence properties of Ca0.5La(MoO4)2:xEu3+ phosphors

Fig. 2a shows a typical PLE spectra of Ca0.5La(MoO4)2:Eu3+ phosphor by monitoring at 615 nm and 612 nm. Both excitation spectra show the same profile including two broad bands range in 233–353 nm and several peaks ranged from 360–434 nm, which can be assigned to the charge transfer bands (CTB) of O–Eu22,23 and O–Mo,24,25 and the characteristic intra-configurational transition 4f–4f transitions of Eu3+,22 respectively. Similar profiles indicate that the emission peaks centered at 615 nm and 612 nm belong to the same 4f–4f transition of Eu3+. In addition, the PLE spectra indicate that Ca0.5La(MoO4)2:Eu3+ phosphor can be efficiently excited by UV-LED chips. Fig. 2b gives the PL spectra of Ca0.5La(MoO4)2:Eu3+ phosphor under different excitation wavelength. There is no perceptible difference between three PL spectra except the emission intensity. All PL spectra consist of a number of sharp lines ranging from 577 to 710 nm, associated with the transitions 5D07FJ (J = 1, 2, 3, 4) of Eu3+. The strongest peak is at 615 nm (5D07F2, electric dipole transition), which is much higher than that of 591 nm (5D07F1, magnetic dipole transition).26 The electric dipole transition is allowed in Ca0.5La(MoO4)2:Eu3+ phosphor indicating good CIE (Commission Internationale de I'Eclairage) chromaticity coordinates. Besides, there is no host emission in Ca0.5La(MoO4)2, as shown in the insert of Fig. 2b. Under 395 nm excitation, the emission intensity was the highest. Thus, we use 395 nm to excite all synthesized phosphors for the following.
image file: c5ra21208a-f2.tif
Fig. 2 Typical PLE (a), PL (b) spectra and the x value dependent on the emission intensity change (c) of Ca0.5La(MoO4)2:xEu3+ phosphors, (d) log plot for the emission intensity at 395 nm per activator ions as a function of the activator concentration.

A series of Ca0.5La(MoO4)2:xEu3+ phosphors were synthesized and the luminescence properties were measured as shown in Fig. 2c. Under 395 nm excitation, with increasing Eu3+ contents, the intensity of Eu3+ emission increases rapidly and reaches a maximum at x = 0.6 and then slowly decreases. Luminescence quenching is relation with energy transfer. Energy transfer is generally associated with multipolar interactions, radiation reabsorption, or exchange interaction.27–30 Among them, multipolar interactions are usually prevalent which have several types, such as dipole–dipole (d–d), dipole–quadrupole (d–q), and quadrupole–quadrupole (q–q) interactions. Exchange interaction is generally limited to interactions between RE ions in nearest or next nearest neighbour. If migration is rapid compared to direct transfer, quenching tends to be proportional to quenching-ions concentration.27,28,31–34 And the exchange interaction is generally for the energy transfer of forbidden transition.35 For a better understanding of energy transfer in the Ca0.5La(MoO4)2:xEu3+ phosphors, the relationship of emission intensity and activator concentration is discussed. As the reports of L. G. van Uitert and Ozawa, the type of energy transfer can be determined from the change in the emission intensity from the emitting level.27,36 The emission intensity (I) per activator ion follows the equation:

 
image file: c5ra21208a-t1.tif(1)
where x is the activator concentration, I/x is the emission intensity (I) per activator concentration (x), and K and β are constants for the same excitation condition for a given host crystal. According to eqn (1), θ = 3 for exchange interaction, namely the energy transfer among the nearest-neighbour ions, while θ = 6, 8, 10 for d–d, d–q, q–q interactions, respectively. Fig. 2d plots the dependence of log(I/x) on log(x) and a slope of (−θ/3) is obtained to be −0.885. Then the value of θ can be determined to be 2.66, close to 3, which means that the quenching is directly proportional to the ion concentration. The result indicates that the energy transfer among nearest-neighbour ions is the main mechanism of concentration quenching for the Eu3+-site emission centers in the Ca0.5La(MoO4)2:xEu3+ phosphors.

3.3 Bi3+ co-doped Ca0.5La(MoO4)2:0.6Eu3+ phosphors

Usually, Bi3+ was utilized to enhance the red emission of Eu3+ in lots of luminescence materials.11,37,38 In this paper, Bi3+ co-doped Ca0.5La(MoO4)2:0.6Eu3+ phosphors were also synthesized, and their XRD patterns are all consistent with PDF card no. 29-0351, shown in Fig. 3. This suggests that all Ca0.5La(MoO4)2:0.6Eu3+, yBi3+ (0 ≤ y ≤ 0.2) phosphors are still single phase. The ionic radii of Bi3+ (r = 1.17 Å when CN = 8) are close to that of La3+ (r = 1.16 Å when CN = 8) and Ca2+ (r = 1.12 Å when CN = 8), therefore we believe that Bi3+ may occupy the La3+ sites or VCa in Ca0.5La(MoO4)2.
image file: c5ra21208a-f3.tif
Fig. 3 XRD patterns of Ca0.5La(MoO4)2:0.6Eu3+, yBi3+ phosphors.

Fig. 4 shows the PLE spectra of Ca0.5La(MoO4)2:0.6Eu3+, yBi3+ phosphors. The profiles of the PLE spectra are similar to Fig. 2a, to clearly understand the effect of introducing Bi3+, the excitation spectra were normalized with 5D07F2 transition and were given in Fig. 4a. A quite broader shoulder band appears at the longer-wavelength side of the O–Eu and O–Mo CTBs. And by increasing the content of Bi3+, the broad band intensity increases. The broad shoulder band should be ascribed to the absorption of Bi3+.39 Broaden PLE spectra in UV range make Ca0.5La(MoO4)2:Eu3+ phosphor more suitable for UV pumped LEDs. The shape of PL spectra of Ca0.5La(MoO4)2:0.6Eu3+, yBi3+ phosphors are similar, while the intensity raises with increasing Bi3+ concentration until its concentration exceeds 0.03, then the intensity decrease, shown in Fig. 4b.


image file: c5ra21208a-f4.tif
Fig. 4 PLE spectra monitored by 615 nm (a) and the y value dependent on the emission intensity change under 395 nm excitation (b) of Ca0.5La(MoO4)2:0.6Eu3+, yBi3+ phosphors.

3.4 The analysis of the abnormal phenomenon

PLE/PL spectra and lifetimes of 5D0 state of Eu3+ in CaMoO4:0.24Eu3+, Ca0.5La(MoO4)2:0.6Eu3+ and Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ phosphors were measured and shown in Fig. 5a–c. As all the transitions include 5D07FJ (J = 1, 2, 3, 4) are contributed to the characteristic intra-configurational transition 4f–4f transitions of Eu3+. And in Ca0.5La(MoO4)2:Eu3+ phosphor, all of them are enhanced compared with CaMoO4:0.24Eu3+ and com-Y2O3:Eu3+ phosphors. In order to facilitate analysis, only the 5D07F2 transition (615 nm) is discussed. Fig. 5a shows the PLE spectra of three phosphors normalized by O–Eu CTB. It is clearly seen, the 7F05D2 intensity in Ca0.5La(MoO4)2:0.6Eu3+ is about 1.3 times stronger than that of Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+, and 5.9 times stronger than that of CaMoO4:0.24Eu3+. In Fig. 5b, the 5D07F2 intensity of Ca0.5La(MoO4)2:0.6Eu3+ is about 1.3 times stronger than that of Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+, and 5.92 times stronger than that of CaMoO4:0.24Eu3+, similar with PLE spectra. It is clearly seen that the red emission is greatly enhanced by La3+ introduced into CaMoO4 host. Besides, Bi3+ co-doped into Ca0.5La(MoO4)2:Eu3+ phosphor decreases the emission intensity. This is different with previous reports.
image file: c5ra21208a-f5.tif
Fig. 5 PLE (a), PL (b) spectra, luminescence decay traces recorded upon excitation at 395 nm (c) and diffuse reflection spectra (d) of CaMoO4:0.24Eu3+, Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ and Ca0.5La(MoO4)2:0.6Eu3+ phosphors.

PL decay curves of three phosphors upon excitation at 395 nm (Fig. 5c) can be well fitted into a single exponential function as:

 
I(t) = I0[thin space (1/6-em)]exp(−t/τ) (2)
where I0 is the initial intensity at t = 0. The lifetime (τ) of three phosphors is determined and given in Fig. 5c. The lifetime of 5D0 reduces following this order CaMoO4:0.24Eu3+ > Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ > Ca0.5La(MoO4)2:0.6Eu3+, indicating that the quantum efficiency of 7F05D2 emission is hardly affected by La3+ or Bi3+ addition. Therefore, according to Fig. 5, it hints that the red emission enhancement of Ca0.5La(MoO4)2:0.6Eu3+ phosphor should result from the increase of the 7F05L6 absorption intensity.

Fig. 5d shows diffuse reflection spectra of three phosphors. In Fig. 5d, the absorption peaks of intra-4f transitions of Eu3+ at 395 nm and 465 nm are presented. The absorbance increases from following order, CaMoO4:0.24Eu3+ > Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ > Ca0.5La(MoO4)2:0.6Eu3+, which was the same as PLE, PL and decay curves. This is direct evidence that the red emission enhancement is caused mainly by the increase of absorption strength of 7F05L6 absorption.

In view of the same phase structure in CaMoO4:0.24Eu3+, Ca0.5La(MoO4)2:0.6Eu3+ and Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ phosphors, combined with above analysis, the change of the 7F05L6 absorption intensity of Eu3+ should be caused by crystal symmetry destroying. For CaMoO4:0.24Eu3+ phosphor, the central Mo6+ metal ion is coordinated with four oxygen atoms in tetrahedral symmetry, and Ca2+ ion, with eight oxygen atoms from a different tetrahedron, shown in Fig. 6a. For Ca0.5La(MoO4)2:0.6Eu3+ phosphor, as described in Fig. 1, is probably formed by one Ca2+, two La3+ and one Ca2+ vacancy (VCa) substitution four Ca2+ sites in CaMoO4. As the radius of La3+ (r = 1.16 Å when CN = 8) is larger than that of Ca2+ (Ca2+ (r = 1.12 Å when CN = 8)),21 the [Ca–LaO8] lattice will get slightly expand compared with [Ca–CaO8]; in the meanwhile, the [La–VCaO8] lattice must shrink greatly, seen in Fig. 6b. These distort the crystal field around Eu3+, and make the symmetry decrease. As the difference between La3+ radius and Ca2+ radius is very little, it is easily understand that the [La–VCaO8] lattice shrink is the main reason for crystal symmetry destroying. It is therefore speculated that more opposite parity components, namely, 4f55d states, mixed into the 4f6 transitional levels of Eu3+. This will cause enhancement of the intra 4f6 transition probability of the 7F05L6 of Eu3+. While for Bi3+ co-doped Ca0.5La(MoO4)2:0.6Eu3+ phosphor, as the ionic radii of Bi3+ (r = 1.17 Å, when CN = 8) and are close to that of Ca2+ (r = 1.12 Å when CN = 8) and La3+ (r = 1.16 Å when CN = 8).21 If Bi3+ occupies La3+ sites, the larger radius of Bi3+ will decrease the [Bi–VCaO8] lattice shrinkage, shown in Fig. 6c. If Bi3+ occupies VCa sites, was shown in Fig. 6d, it will restore the shrinkage. Namely, no matter which site is substituted, the crystal symmetry will be restored. And thus the transition probability will decrease. Meanwhile, if Bi3+ occupies VCa site, due to short distance, the energy of Eu3+ will easily transfer to Bi3+ on higher Eu3+ content. All above reason will cause the red emission intensity decrease compared with Ca0.5La(MoO4)2:0.6Eu3+ phosphor, but higher than CaMoO4:0.24Eu3+ phosphor.


image file: c5ra21208a-f6.tif
Fig. 6 Crystal structure of CaMoO4:0.24Eu3+ (a), Ca0.5La(MoO4)2:0.6Eu3+ (b), Ca0.5La(MoO4)2:0.6Eu3+, 0.03Bi3+ (c and d) phosphors.

The emission intensity of the optimized Ca0.5La(MoO4)2:0.6Eu3+ phosphor is compared with the com-Y2O3:Eu3+ phosphor and CaMoO4:0.24Eu3+ shown in Fig. 7, colors are filled with photos of each phosphors under 365 nm light. The results show that the intensity of synthesized phosphor is 2.67 and 5.92 times than that of the com-Y2O3:Eu3+ phosphor and CaMoO4:0.24Eu3+, respectively. The CIE chromaticity coordinates of Ca0.5La(MoO4)2:0.6Eu3+ phosphor is also calculated chromaticity coordinates (0.667, 0.330) compared with (0.629, 0.340) for com-Y2O3:Eu3+ phosphor. The CIE chromaticity coordinates are on the margin of the NTSC (National Television Standard Committee) standard values (0.67, 0.33).40 Enhanced red emission, stronger 7F05L6 absorption intensity, and high color purity of Ca0.5La(MoO4)2:Eu3+ phosphor indicate that it is a promising red phosphor for white LEDs.


image file: c5ra21208a-f7.tif
Fig. 7 The PL intensity of Ca0.5La(MoO4)2:0.6Eu3+, com-Y2O3:Eu3+ and CaMoO4:0.24Eu3+ phosphors, colors are filled with photos of each phosphors under 365 nm light.

4. Conclusion

Novel Ca0.5La(MoO4)2 molybdate host and red emitting Ca0.5La(MoO4)2:Eu3+ phosphors were synthesized by solid state reaction. The structure of Ca0.5La(MoO4)2 molybdate host is similar to CaMoO4 with space group of I41/a(88). Eu3+ activated this molybdate shows intense absorption in the UV region and exhibits intense red emission. The excitation spectra indicate novel molybdate phosphor is suitable for UV pumped LEDs. For optimized Ca0.5La(MoO4)2:Eu3+ phosphor, the intensity is 2.67 and 5.92 times than that of the com-Y2O3:Eu3+ phosphor and CaMoO4:0.24Eu3+, respectively. The red emission enhancement is due to the increase of the 7F05L6 absorption intensity of Eu3+ in the distorted crystal field in Ca0.5La(MoO4)2:Eu3+ phosphor. The CIE chromaticity coordinates (0.667, 0.330) of the synthesized phosphor are on the margin of the NTSC standard values. Enhanced emission and high color purity indicate that Ca0.5La(MoO4)2:Eu3+ phosphor could be a good candidate for the red component of UV white LEDs. The introduction of Bi3+ into optimized phosphor can broaden excitation spectra in the UV range, but decrease the emission intensity slightly due to restoring the distortion and providing new quenching path.

Acknowledgements

Project supported by Program for Scientific Research Foundation of the Higher Education Institutions of He'nan Province (15A430054), High level personnel fund of Zhoukou Normal University (ZKNU2014106), Open project of Zhoukou Normal University (K201501), Innovative Research Team (in Science and Technology) in University of Henan Province (14IRTSTHN009), Excellent Youth Foundation of He'nan Scientific Committee (134100510018). Henan Province Key Discipline of Applied Chemistry (201218692), and Innovation Scientists and Technicians Troop Construction Projects of Henan Province (2013259).

References

  1. Z. Ci, Q. Sun, S. Qin, M. Sun, X. Jiang, X. Zhang and Y. Wang, Phys. Chem. Chem. Phys., 2014, 16, 11597 RSC.
  2. Y. Shi, G. Zhu, M. Mikami, Y. Shimomura and Y. Wang, Dalton Trans., 2015, 44, 1775 RSC.
  3. J. R. Oh, S. H. Cho, Y. H. Lee and Y. R. Do, Opt. Express, 2009, 17, 7450 CrossRef CAS PubMed.
  4. L. Yi, J. Zhang, Z. Qiu, W. Zhou, L. Yu and S. Lian, RSC Adv., 2015, 5, 67125 RSC.
  5. J. K. Park, K. J. Choi, C. H. Kim, H. D. Park and S. Y. Choi, Electrochem. Solid-State Lett., 2004, 7, H15 CrossRef CAS.
  6. S. Neeraj, N. Kijima and A. K. Cheetham, Chem. Phys. Lett., 2004, 387, 2 CrossRef CAS.
  7. J. Lin, Y. Huang, Y. Bando, C. Tang and D. Golberg, Chem. Commun., 2009, 66313 Search PubMed.
  8. R. J. Xie, N. Hirosaki, T. Suehiro, F. F. Xu and M. Mitomo, Chem. Mater., 2006, 18, 5578 CrossRef CAS.
  9. X. Piao, K. Machida, T. Horikawa, H. Hanzawa, Y. Shimomura and N. Kijima, Chem. Mater., 2007, 19, 4592 CrossRef CAS.
  10. Y. Cao, G. Zhu and Y. Wang, RSC Adv., 2015, 5, 65710 RSC.
  11. S. X. Yan, J. H. Zhang, X. Zhang, S. Z. Lu, X. G. Ren, Z. G. Nie and X. J. Wang, J. Phys. Chem. C, 2007, 111, 13256 CAS.
  12. Y. Zhou and B. Yan, CrystEngComm, 2013, 15, 5694 RSC.
  13. Y. C. Chang, C. H. Liang, S. A. Yan and Y. S. Chang, J. Phys. Chem. C, 2010, 114, 3645 CAS.
  14. A. Xie, X. M. Yuan, S. J. Hai, J. J. Wang, F. X. Wang and L. Li, J. Phys. D: Appl. Phys., 2009, 42, 105107 CrossRef.
  15. H. Y. Li, H. M. Noh, B. K. Moon, B. C. Choi, J. H. Jeong, K. Jang, H. S. Lee and S. S. Yi, Inorg. Chem., 2013, 52, 11210 CrossRef CAS PubMed.
  16. C. L. Zhao, Y. S. Hu, W. D. Zhuang, X. W. Huang and T. He, J. Rare Earths, 2009, 27, 758 CrossRef.
  17. H. Y. Li, H. K. Yang, B. K. Moon, B. C. Choi, J. H. Jeong, K. Jang, H. S. Lee and S. S. Yi, Inorg. Chem., 2011, 50, 12522 CrossRef CAS PubMed.
  18. X. Q. Liu, L. Li, H. M. Noh, B. K. Moon, B. C. Choi, S. K. Neeraj and J. H. Jeong, Dalton Trans., 2014, 43, 8814 RSC.
  19. J. Liu, H. Lian and C. Shi, Opt. Mater., 2007, 29, 1591 CrossRef CAS.
  20. Y. Hu, W. Zhuang, H. Ye, D. Wang, S. Zhang and X. Huang, J. Alloys Compd., 2005, 390, 226 CrossRef CAS.
  21. R. D. Shannon, Acta Crystallogr., Sect. A: Cryst. Phys., Diffr., Theor. Gen. Crystallogr., 1976, 32, 751 CrossRef.
  22. G. Blasse and A. Bril, J. Inorg. Nucl. Chem., 1967, 29, 2231 CrossRef CAS.
  23. V. R. Bandi, M. Jayasimhadri, J. Jeong, K. Jang, H. S. Lee, S. S. Yi and J. H. Jeong, J. Phys. D: Appl. Phys., 2010, 43, 395103 CrossRef.
  24. A. K. Parchur, R. S. Ningthoujam, S. B. Rai, G. S. Okram, R. A. Singh, M. Tyagi, S. C. Gadkari, R. Tewari and R. K. Vatsa, Dalton Trans., 2011, 40, 7595 RSC.
  25. Z. Lu and T. Wanjun, Ceram. Int., 2012, 38, 837 CrossRef.
  26. M. Thomas, P. P. Rao, S. P. K. Mahesh, L. S. Kumari and P. Koshy, Phys. Status Solidi A, 2011, 208, 2170 CrossRef CAS.
  27. L. G. van Uitert, J. Electrochem. Soc., 1967, 114, 1048 CrossRef CAS.
  28. L. G. van Uitert, E. F. Dearborn and J. J. Rubin, J. Chem. Phys., 1967, 47, 547 CrossRef CAS.
  29. H. Liu, Y. Hao, H. Wang, J. Zhao, P. Huang and B. Xu, J. Lumin., 2011, 131, 2422 CrossRef CAS.
  30. S. Dutta, S. Som and S. K. Sharma, Dalton Trans., 2013, 42, 9654 RSC.
  31. L. G. van Uitert, E. F. Dearborn and J. J. Rubin, J. Chem. Phys., 1966, 45, 1578 CrossRef.
  32. L. G. van Uitert, E. F. Dearborn and H. M. Marcos, Appl. Phys. Lett., 1966, 9, 255 CrossRef CAS.
  33. L. G. van Uitert, E. F. Dearborn and J. J. Rubin, J. Chem. Phys., 1967, 46, 3551 CrossRef CAS.
  34. L. G. van Uitert, E. F. Dearborn and J. J. Rubin, J. Chem. Phys., 1967, 46, 420 CrossRef CAS.
  35. X. Zhang, L. Zhou and M. Gong, Opt. Mater., 2013, 35, 993 CrossRef.
  36. L. Ozawa and P. M. Jaffe, J. Electrochem. Soc., 1971, 118, 1678 CrossRef CAS.
  37. L. L. Han, Y. H. Wang, J. Zhang and Y. Z. Wang, Mater. Chem. Phys., 2013, 139, 87 CrossRef CAS.
  38. L. Chen, K. J. Chen, C. C. Lin, C. Chu, S. F. Hu, M. H. Lee and R. S. Liu, J. Comb. Chem., 2010, 12, 587 CrossRef CAS PubMed.
  39. A. Xie, X. M. Yuan, Y. Shi, F. X. Wang and J. J. Wang, J. Am. Ceram. Soc., 2009, 92, 2254 CrossRef CAS.
  40. Z. L. Wang, H. B. Liang, J. Wang and M. L. Gong, Appl. Phys. Lett., 2006, 89, 071921 CrossRef.

This journal is © The Royal Society of Chemistry 2015