Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Synthesis and photoluminescence characteristics of the LiGd3(MoO4)5:Eu3+ red phosphor with high color purity and brightness

Open Access Open Access

Abstract

Red phosphors LiGd3(MoO4)5 doped with various Eu3+ concentrations were synthesized by the solid state reaction method. Diffuse reflection spectra, photoluminescence spectra, and temperature-dependent luminescence spectra of the phosphors were measured. Due to the weak concentration quenching effect, the optimal Eu3+ concentration is up to 90 at.% in the LiGd3(MoO4)5:Eu3+, which shows higher color purity and brightness than the commercial Y2O2S:6.3at.%Eu3+ red phosphor excited at 395 nm. When the temperature was increased to 450 K, the red emission intensity of the phosphor decreased to 58% of that at 300 K. The results show that the LiGd3(MoO4)5:90at.%Eu3+ may be a promising red phosphor for white light-emitting diodes.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

White light-emitting diodes (WLEDs) have attracted attention due to the advantages of compactness, high brightness, long operating lifetime, environmental friendliness as well as good physicochemical stability [1]. Presently, the most prevalent method of fabricating WLEDs is to combine commercial blue LED chips with yellow phosphors such as the well-known Ce3+-doped yttrium aluminum garnet (YAG:Ce3+) [2]. However, this method has some inherent drawbacks, such as poor color rendering index and high color temperature [3]. Therefore, an alternative strategy based on a combination of a near ultraviolet (NUV) LED chip (380-410 nm) with red, green, and blue phosphors has been proposed to achieve a warm white light [4]. Compared with the blue and green phosphors, the commonly used Y2O2S:Eu3+ red phosphor shows lower physicochemical stability [5].

Recently, Eu3+-doped tungstate and molybdate phosphors have attracted much attention due to the advantages of easy-synthesis, high physicochemical stability, and intense red light emission [6, 7]. In order to strengthen the absorption of Eu3+ in the NUV region, some tungstates and molybdates that can be high Eu3+-doped have been investigated, such as Na5Y(MoO4)4 and Li3Ba2La3(MoO4)8 [8, 9]. Especially, it has been found that the concentration quenching effect in some defect scheelite structures is weak, such as the optimal Eu3+ concentrations in the M2Gd4(MoO4)7 (M = Li and Na) phosphors can be up to 85at.% and 70at.%, respectively [10]. LiGd3(MoO4)5 crystal also has a similar defect scheelite structure. Compared with the scheelite CaWO4, the Ca2+ sites in the LiGd3(MoO4)5 are not fully occupied by Li+ and Gd3+, and 20% of them are vacancies [11].

In this work, LiGd3(MoO4)5:Eu3+ red phosphors with different Eu3+ concentrations were synthesized by the solid state reaction method. The effects of Eu3+ concentration on the photoluminescence (PL) properties of the phosphors were investigated in detail. Meanwhile, the thermal stability, chromatic property, and quantum yield of the phosphors are reported.

2. Experimental

2.1 Preparations

A series of red phosphors LiGd3(1-x)Eu3x(MoO4)5 (LGM:xEu3+, x = 0, 10, 20, …, 100at.%) were synthesized by the solid-state reaction method. The starting materials Li2CO3 (Aladdin, A.R.), Gd2O3 (Aladdin, 99.99%), Eu2O3 (Aladdin, 99.99%) and MoO3 (Aladdin, A.R.) were weighted according to the stoichiometric ratio. The raw materials were mixed and ground in an agate mortar thoroughly. Then, the homogenous mixtures were transferred to an alumina crucible and had been sintered in air at 500 °C for 3 h in a muffle furnace. The pre-samples were re-ground and had been calcined at 850 °C for 5 h again. After cooling to room temperature naturally, the as-prepared phosphors were ground into powders for measurements.

2.2 Characterizations

XRD measurements were carried out on a MiniFlex 600 powder diffractometer with Cu Kα radiation (1.5405 Å) operating at 40 kV and 15 mA. The scanning speed for phase determination was 0.2 °/min. The morphology and elemental analysis of the samples were inspected using a field emission scanning electron microscope (FESEM, SU-8010) equipped with an energy-dispersive spectrometer (EDS). Diffuse refection spectra (DRS) of the powder samples were measured by a spectrophotometer (Lambda900, Perkin-Elmer) and using BaSO4 as a standard reference. Photoluminescence spectra, decay curves, and temperature-dependent luminescence spectra were recorded by a spectrometer (Edinburgh Instruments, FLS980) equipped with both continuous (450 W) and pulsed xenon lamps as the excited sources. Quantum yield (QY) was measured by a spectrometer (Edinburgh Instruments, FLS920) using a barium sulfate coated integrating sphere attached to the spectrometer.

3. Results and discussion

3.1 Phase composition

The recorded XRD patterns of the LGM:xEu3+ (x = 0, 10, 20, …, 100 at.%) phosphors are similar to each other, therefore, only three of them are shown in Fig. 1 for the sake of brevity. The main intense diffraction peaks of the samples can be well indexed by the tetragonal scheelite structure with space group I41/a as shown in Fig. 1(a). Besides, it can be found from the magnified patterns in Fig. 1(b) that there are some weak extra peaks that cannot be indexed by the scheelite structure. Similar peaks, which were also found in the recorded XRD patterns of Na0.2Eu0.6MoO4, have been attributed to the cation-vacancy ordering and the formation of incommensurately modulated structure caused by the influence of a mass of vacancies in the crystal [12, 13]. The concrete distribution of the cations and vacancies has been solved in Ref [12]. Furthermore, it can be seen from Fig. 1(c) that the location of the diffraction peak corresponding to (112) slightly shifts to lower 2θ angel with the increment of Eu3+ concentration. According to the Bragg equation λ = 2dsinθ, the shift of peak location can be attributed to the slightly larger radius of Eu3+ (r = 1.066 Å, CN = 8) than that of Gd3+ (r = 1.053 Å, CN = 8) [14].

 figure: Fig. 1

Fig. 1 (a) XRD patterns of the LGM:xEu3+ (x = 0, 50, and 90 at.%) phosphors, and the corresponding magnified XRD curves in the ranges of 10-28° (b) and 28-29.5° (c).

Download Full Size | PDF

The morphology and elemental compositions of the LGM:90at.%Eu3+ phosphor sample are shown in Figs. 2(a) and 2(b), respectively. The morphology in Fig. 2(a) shows that the sizes of the irregular particles are mainly ranging from 2 to 5 μm. As presented in Fig. 2(b), a small area of sample was selected for the EDS measurement. The result confirms the presence of Mo, Eu, and Gd elements and no other impurity peaks. However, the light Li element is hard to be identified by the EDS measurement. The atomic ratios of Eu/(Gd + Eu) and (Gd + Eu)/Mo in the sample were calculated to be about 90% and 58%, respectively, which are almost consistent with the chemical formula of Li(Gd0.1Eu0.9)3(MoO4)5. To further investigate the spatial distribution of elements, the EDS elemental mapping analysis corresponding to the particle imaged in Fig. 3(a) was carried out, as shown in Fig. 3(b), (c) and (d). The pictures show that Mo, Eu, and Gd uniformly distribute and the as-prepared phosphor is homogeneous phase.

 figure: Fig. 2

Fig. 2 (a) SEM micrograph and (b) EDS data of the LGM:90at.%Eu3+ phosphor.

Download Full Size | PDF

 figure: Fig. 3

Fig. 3 The mapping EDS images of Mo (b), Eu (c) and Gd (d) corresponding to the particle shown in (a).

Download Full Size | PDF

3.2 Reflection and photoluminescence properties of Eu3+ doped LiGd3(MoO4)5

Figure 4(a) shows the room-temperature diffuse reflection spectra of the LGM:xEu3+ (x = 0, 50, and 90 at.%) phosphors in 200-800 nm. The host LGM has a broad absorption band from 250 to 400 nm corresponding to the charge-transfer state (CTS) band of Mo6+-O2- [15]. When Eu3+ ions are doped into the LGM, the CTS of Eu3+-O2- appears and the absorption edge of Mo6+-O2- CTS band shifts to a longer wavelength, which may be caused by the increment of covalency of Mo-O bond due to the incorporation of Eu3+ [5, 16]. Due to the larger electronegativity of Eu (1.18 [17]) than that of Gd (0.9 [18]), the electronegativity difference between Eu and O is smaller than that of Gd and O and the covalency of Eu-O bond is stronger than that of Gd-O bond. The stronger covalency of Ln-O bond in Ln–O–Mo will weaken the bond ionicity and increase the bond covalency of Mo–O bond. So the incorporation of Eu3+ results in the increment of covalency of Mo-O bond and the location of CTS of Mo6+-O2- shifts to lower energy. Meanwhile, several narrow absorption bands can be observed in the visible region, which are originated from the characteristic absorption of Eu3+ intra-4f transitions. Obviously, with the increment of Eu3+ concentration, the characteristic absorption of Eu3+ is prominently enhanced, and the Mo6+-O2- CTS leads to a stronger and broader absorption band in NUV region.

 figure: Fig. 4

Fig. 4 (a) Room-temperature diffuse reflection spectra of the LGM, LGM:50at.%Eu3+ and LGM:90at.%Eu3+ phosphors in 200-800 nm, (b) Room-temperature excitation spectra of the LGM:10at.%Eu3+ and LGM:90at.%Eu3+phosphors.

Download Full Size | PDF

Figure 4(b) shows the room-temperature excitation spectra of the LGM:10at.%Eu3+ and LGM:90at.%Eu3+ phosphors monitored at 615 nm corresponding to the 5D07F2 transition. The excitation spectrum can be divided into two regions. One part is a broad band from 250 to 350 nm, which is mainly attributed to Eu3+-O2- CTS band due to the intensity increases with increasing Eu3+ concentration [19]. The charge transfer absorption of Mo6+-O2- existed in diffuse reflection spectra in NUV region may not make contribution to the 5D07F2 emission [10]. The other part includes sharp lines in 350-550 nm corresponding to the intra-4f transitions of Eu3+. Among these lines, the strongest exciting peak is located at 395 nm and assigned to the 7F05L6 transition, which implies that the LGM:Eu3+ phosphor might be used as red phosphor excited by the NUV LED chips.

Figure 5(a) shows the room-temperature emission spectra of the LGM:xEu3+ phosphors under excitation at 395 nm. The profile of emission spectrum is not changed with the Eu3+ concentration in the phosphors. In the emission spectrum, the host emission corresponding to Mo6+-O2- CTS is not observed, which may be due to that thermal quenching occurred at room temperature [19]. The strongest emission peak of the phosphor at 615 nm is attributed to the 5D07F2 transition of Eu3+. The emissions originated from the 5D07F1 (592 nm), 5D07F3 (654 nm) and 5D07F4 (701 nm) transitions are much weaker. The 5D07F2 is a hypersensitive transition and strongly influenced by the coordination environment of Eu3+, while the 5D07F1 transition is insensitive to that [20]. Therefore, the emission intensity ratio R between the 5D07F2 and 5D07F1 transitions can be used as an index to measure the site symmetry of Eu3+ ions, and a larger R value means that the Eu3+ site in a host has a lower local symmetry [21]. In the LGM:xEu3+ phosphors, the intensity of the 5D07F2 is much stronger than that of the 5D07F1, and the R values are almost kept as a constant of 13.8 for different Eu3+ concentrations, as shown in Fig. 5(b). The value is smaller than that of the Li2Gd4(MoO4)7:85at.%Eu3+ (16 [10]), but larger than those of the Li3Ba2La3(MoO4)8:90at.%Eu3+ (8.82 [8]) and KBaEu(MoO4)3 (12.4 [22]) phosphors. It indicates that the LGM:xEu3+ phosphor has a low Eu3+ site symmetry and high red color purity [22].

 figure: Fig. 5

Fig. 5 (a) Room-temperature emission spectra of the LGM:xEu3+ (x = 10, 50, 90 at.%) phosphors excited at 395 nm. and the enlarged region in the inset. (b) the dependence of integrated emission intensity of the 5D07F2 transition and the (5D07F2)/(5D07F1) emission ratio R on the Eu3+ concentration.

Download Full Size | PDF

Figure 5(b) shows the integrated emission intensity of the 5D07F2 transition as a function of Eu3+ concentration. The emission intensity is increased with the increment of Eu3+ concentration and reached the maximum at x = 90 at.% (~68 × 1020 cm−3). The optimal Eu3+ concentration in the LGM phosphors is higher than those in the M2Gd4(MoO4)7 (~62 × 1020 and ~51 × 1020 cm−3 for M = Li and Na, respectively [10]), KBaEu(MoO4)3 (~37 × 1020 cm−3 [22]), Li3Ba2La3(WO4)8 (~40 × 1020 cm−3 [8]) and LiGd(MoO4)2 (~19 × 1020 cm−3 [23]) phosphors. The result shows that the concentration quenching effect in the LGM:Eu3+ phosphor is very weak, which may be caused by its defect scheelite structure. The weak concentration quenching has also been found in some phosphors with the incommensurately modulated structure, such as Na2Gd4(MoO4)7:Eu3+ [24, 25] and Na0.2Eu0.6MoO4 [12]. In the LGM:Eu3+ phosphor, Gd3+ ions may tend to occupy the adjacent Ca2+ sites in the scheelite structure and then gather together to form Gd-aggregates with longer distance between each other, as in the Na0.2Eu0.6MoO4 phosphor [12]. Therefore, the cluster structure of Gd3+ in the LGM, which is similar to the layer structure of La3+ in Li3Ba2La3(WO4)8 [8] and the chain structure of Eu3+ in KBaEu(MoO4)3 [22], makes the interaction between doping Eu3+ mainly occur inside the aggregates, and then reduces the energy migration to the quenching centers such as crystal defects or trace ions. On account of the fact that the cross-relaxation between Eu3+ ions is very weak [26], the concentration quenching should be mainly attributed to the energy migration from Eu3+ luminescent centers to the quenching centers [27]. Consequently, the concentration quenching effect is weak in the LGM:Eu3+ phosphor.

The excitation and emission spectra of the LGM:90at.%Eu3+ and the commercial Y2O2S:6.3at.%Eu3+ red phosphors recorded in the same experimental conditions are compared in Figs. 6(a) and 6(b), respectively. The Y2O2S:6.3at.%Eu3+ phosphor was bought from Jiangmen Kanhoo Industry Co., Ltd. It can be seen from Fig. 6(a) that the Y2O2S:6.3at.%Eu3+ phosphor shows a stronger and broader excitation band from 250 to 380 nm corresponding to the CTS of Eu3+-O2- and Eu3+-S2- [28], but the LGM:90at.%Eu3+ phosphor has a more efficient absorption in NUV region. Under excitation at 395 nm, the integrated emission intensity of the 5D07F2 transition for the LGM:90at.%Eu3+ phosphor is approximate 2.5 times of that of the Y2O2S:6.3at.%Eu3+ phosphor. Furthermore, the emission intensity ratio R between the 5D07F2 and 5D07F1 transitions for the Y2O2S:6.3at.%Eu3+ was calculated to be 7.9, and the LGM:90at.%Eu3+ has larger R value and narrower emission band corresponding to the 5D07F2 transition than the Y2O2S:6.3at.%Eu3+ phosphor, which can make the LGM:90at.%Eu3+ phosphor have higher color purity. The above results indicate that compared with the commercial Y2O2S:6.3at.%Eu3+ red phosphor, the LGM:90at.%Eu3+ is an excellent red one with higher luminescence intensity and color purity excited by the 395 nm LEDs. The LEDs with emission wavelength around 400 nm has the highest external quantum efficiency and is safer for human eyes among the LED chips with emission wavelength from 250 to 400 nm [29, 30]. It must be pointed out that the full width at half maximum (FWHM) of the exciting band around 395 nm for the LGM:90at.%Eu3+ phosphor is only 1.5 nm, therefore, On account of utilization ratio of excitation source, the LGM:Eu3+ phosphor may be more suitable for being excited by laser diode (LD) at 395 nm with narrower bandwidth. Efficient white light has been achieved by combining a NUV LD with red, green, and blue phosphors [31].

 figure: Fig. 6

Fig. 6 (a) Excitation and (b) emission spectra of the LGM:90at.%Eu3+ and commercial Y2O2S:6.3at.%Eu3+ red phosphors.

Download Full Size | PDF

Figure 7 shows the fluorescence decay curves of the LGM:xEu3+ phosphors excited at 395 nm and the emission related to the 5D07F2 transition at 615 nm was monitored. All curves show a single exponential decay and the fluorescence lifetime τ of the 5D0 level can be obtained by

I=I0exp(t/τ)
where I represents the emission intensity at time t after the cut-off of the excitation light and I0 is the initial emission intensity at t = 0. The τ values of the LGM:xEu3+ phosphors are nearly kept at 420 μs for different Eu3+ concentrations as shown in Fig. 7. In general, with the increment of Eu3+ concentration, the nonradiative process is intensified, the fluorescence lifetime is reduced, and the effect of concentration quenching occurs. However the cluster structure in LGM can cut off the energy migration from luminescent centers to the quenching centers and remain the fluorescence lifetime constant. So the concentration quenching effect is weak in the LGM:Eu3+ phosphor.

 figure: Fig. 7

Fig. 7 Fluorescence decay curves of the LGM:xEu3+ (x = 10, 30, 50, 70, 80, 90, and 100 at.%) phosphors.

Download Full Size | PDF

3.3 Thermal quenching property of LGM:90at.%Eu3+

The temperature dependent emission spectra of the LGM:90at.%Eu3+ phosphor under excitation at 395 nm are shown in Fig. 8. The emission intensity of the phosphor obviously declines with the temperature increment from 300 K to 550 K. As the temperature reaches to 450 K, the emission intensity declines to 58% of the initial value at 300 K. For Eu3+-doped tungstate and molybdate phosphors [8, 22, 32], the thermal quenching of fluorescence emitted from the 5D0 level is usually owed to nonradiative decay via the charge transfer state that has a crossover point with the 5D0 and 7FJ levels, as shown in Fig. 9(a) [33].

 figure: Fig. 8

Fig. 8 Temperature dependence of fluorescence for the LGM:90at.%Eu3+ phosphor under excitation at 395 nm. The inset shows the temperature dependence of emission intensity of the 5D07F2 transition under excitation at 395 nm and 465 nm, respectively.

Download Full Size | PDF

 figure: Fig. 9

Fig. 9 (a) Pathway for the thermal quenching of the 5D0 state through a CTS band, (b) The dependence of ln(I0/I-1) on 1/kT for the LGM:90at.%Eu3+ phosphor.

Download Full Size | PDF

The activation energy ∆E is the energy separation from the bottom of the 5D0 level to the 5D0-charge transfer state crossover, and can be calculated by the relationship of the emission intensity and temperature [34],

ln(I0/I1)=lnAΔE/kT
where I and I0 represent the emission intensity of the LGM:90at.%Eu3+ phosphor at the experimental and room temperature, respectively. A is a constant for a certain host, and k is Boltzmann’s constant. Figure 9(b) shows the linear relation of ln(I0/I-1) and 1/kT, and the ∆E was calculated to be about 0.24 eV. The value is similar to those of the reported red phosphors such as Li3Ba2La3(WO4)8:90at.%Eu3+ (0.24 eV [8]), Gd2Mo3O12:25at.%Eu3+ (0.24 eV [35]), LiLaMgWO6:30at.%Eu3+ (0.225 eV [36]) and Sr9Eu2W4O24 (0.22 eV [37]). Furthermore, the thermal quenching property at 465 nm excitation has been investigated. In this excitation range, excitation should solely occur via Eu3+ without interference of CTS absorption. From the inset of Fig. 8, it can be found that the LGM:90at.%Eu3+ phosphor excited at 395 nm shows higher thermal stability than that at 465 nm. So the CTS of Mo6+-O2- has little negative influence on the thermal stability at 395 nm excitation.

3.4 CIE chromaticity coordinates and quantum efficiency of LGM:90at.%Eu3+

The color chromaticity is regarded as a critical parameter for evaluating the performance of WLED phosphors. The corresponding Commission International del’Eclairage (CIE) chromaticity coordinate of the LGM:90at.%Eu3+ phosphor at room temperature is about (0.6705, 0.3292), as shown in Fig. 10, which is close to the standard red chromaticity (0.67, 0.33) for the National Television Standard Committee (NTSC) system. With the increment of temperature, the CIE chromaticity coordinate for the LGM:90at.%Eu3+ phosphor only changes slightly as shown in the inset of Fig. 10. According to Ref [38], the color purity describes the saturation and degree of vividness of color, and the closer to spectral locus the chromaticity coordinate is, the higher the color purity is. From Fig. 10, it can be clear that the LGM:90at.%Eu3+ phosphor indeed shows higher color purity than the Y2O2S:6.3at.%Eu3+ phosphor.

 figure: Fig. 10

Fig. 10 CIE chromaticity coordinate of the LGM:90at.%Eu3+ phosphor at 395 nm excitation.

Download Full Size | PDF

The QY of the LGM:90at.%Eu3+ phosphor under excitation at 395 nm is 41%, which is higher than those of the KBaEu(MoO4)3 (16% at λex = 394 nm [22]) and Na2Gd4(MoO4)7:70at.%Eu3+ (19% at λex = 396 nm, 22.6% at λex = 465 nm [10]), but lower than those of the Y2O2S:6.3at.%Eu3+ (76% at λex = 317 nm, 77% at λex = 395 nm, measured in our lab), and Li3Ba2La3(WO4)8:90at.%Eu3+ (99% at λex = 395 nm [8]) phosphors. The QY of the LGM:90at.%Eu3+ phosphor at 465 nm excitation is 87%, which is much higher than that at 395 nm excitation. Because there is no CTS interference at 465 nm excitation, so the low QY of the LGM:90at.%Eu3+ phosphor is likely due to that the charge transfer absorption of Mo6+-O2- around 395 nm only little make contribution to the 5D07F2 emission. By partial substitution of W6+ for Mo6+, the concentration of the MoO4 group is diluted. the distance between MoO4 group becomes longer, and the energy transition is blocked by MoO4, which may result in more energy transferred from the host to Eu3+ ions [39].

4. Conclusion

A series of LGM:xEu3+ red phosphors have been synthesized by the solid state reaction method. The variations of fluorescence spectra and lifetimes with the Eu3+ concentration show that the concentration quenching effect is weak and the optimal Eu3+ concentration can be up to 90at.% in the phosphors. The emission intensity of the LGM:90at.%Eu3+ phosphor is 2.5 times of that of the Y2O2S:6.3%Eu3+ commercial red phosphor under excitation at 395 nm. The LGM:90at.%Eu3+ phosphor has a high color purity with CIE chromaticity coordinates of (0.6705, 0.3292), which is very close to the NTSC system standard for red chromaticity. Therefore, the LGM:90at.%Eu3+ phosphor has potential applications for the NUV excited WLEDs, especially, laser-driving white lighting.

Funding

National Natural Science Foundation of China (51772290); Strategic Priority Research Program of the Chinese Academy of Sciences (XDB20000000).

References and links

1. S. Pimputkar, J. S. Speck, S. P. DenBaars, and S. Nakamura, “Prospects for LED lighting,” Nat. Photonics 3(4), 179–180 (2009). [CrossRef]  

2. P. Pust, P. J. Schmidt, and W. Schnick, “A revolution in lighting,” Nat. Mater. 14(5), 454–458 (2015). [CrossRef]   [PubMed]  

3. F. Baur, F. Glocker, and T. Juestel, “Photoluminescence and energy transfer rates and efficiencies in Eu3+ activated Tb2Mo3O12,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(9), 2054–2064 (2015). [CrossRef]  

4. J. McKittrick, M. E. Hannah, A. Piquette, J. K. Han, J. I. Choi, M. Anc, M. Galvez, H. Lugauer, J. B. Talbot, and K. C. Mishra, “Phosphor selection considerations for near-UV LED solid state lighting,” Ecs J. Solid State Sci. Technol. 2(2), R3119–R3131 (2012). [CrossRef]  

5. S. Neeraj, N. Kijima, and A. K. Cheetham, “Novel red phosphors for solid-state lighting: the system NaM(WO4)2-x(MoO4)x:Eu3+ (M=Gd, Y, Bi),” Chem. Phys. Lett. 387(1-3), 2–6 (2004). [CrossRef]  

6. C. Qin, Y. Huang, G. Chen, L. Shi, X. Qiao, J. Gan, and H. J. Seo, “Luminescence properties of a red phosphor europium tungsten oxide Eu2WO6,” Mater. Lett. 63(13-14), 1162–1164 (2009). [CrossRef]  

7. C. Litterscheid, S. Krueger, M. Euler, A. Dreizler, C. Wickleder, and B. Albert, “Solid solution between lithium-rich yttrium and europium molybdate as new efficient red-emitting phosphors,” J. Mater. Chem. C Mater. Opt. Electron. Devices 4(3), 596–602 (2016). [CrossRef]  

8. J. Hu, X. Gong, J. Huang, Y. Chen, Y. Lin, Z. Luo, and Y. Huang, “Near ultraviolet excited Eu3+ doped Li3Ba2La3(WO4)8 red phosphors for white light emitting diodes,” Opt. Mater. Express 6(1), 181–190 (2016). [CrossRef]  

9. C. Guo, F. Gao, Y. Xu, L. Liang, F. G. Shi, and B. Yan, “Efficient red phosphors Na5Ln(MoO4)4:Eu3+ (Ln = La, Gd and Y) for white LEDs,” J. Phys. D Appl. Phys. 42(9), 095407 (2009). [CrossRef]  

10. C. Zhao, X. Yin, F. Huang, and Y. Hang, “Synthesis and photoluminescence properties of the high-brightness Eu3+-doped M2Gd4(MoO4)7 (M=Li, Na) red phosphors,” J. Solid State Chem. 184(12), 3190–3194 (2011). [CrossRef]  

11. R. K. Pandey, “Growth of LiGd3(MoO4)5 single-crystals,” J. Cryst. Growth 48(3), 355–358 (1980). [CrossRef]  

12. A. Arakcheeva, D. Logvinovich, G. Chapuis, V. Morozov, S. V. Eliseeva, J.-C. G. Bünzli, and P. Pattison, “The luminescence of NaxEu3+(2−x)/3MoO4 scheelites depends on the number of Eu-clusters occurring in their incommensurately modulated structure,” Chem. Sci. (Camb.) 3(2), 384–390 (2012). [CrossRef]  

13. A. Arakcheeva and G. Chapuis, “Capabilities and limitations of a (3 + d)-dimensional incommensurately modulated structure as a model for the derivation of an extended family of compounds: example of the scheelite-like structures,” Acta Crystallogr. B 64(Pt 1), 12–25 (2008). [CrossRef]   [PubMed]  

14. R. D. Shannon, “Revised effective ionic-radii and systematic studies of interatomic distances in halides and chalcogenides,” Acta Crystallogr. A 32(5), 751–767 (1976). [CrossRef]  

15. P. S. Dutta and A. Khanna, “Eu3+ activated molybdate and tungstate based red phosphors with charge transfer band in blue region,” ECS J. Solid State Sci. Technol. 2(2), R3153–R3167 (2012). [CrossRef]  

16. S. Qi, H. Xie, Y. Huang, S. I. Kim, and H. J. Seo, “A narrow red-emitting phosphor of NaLa4 Mo3O15 F: Eu3+ with broad excitation band extending in blue wavelength region,” Opt. Mater. Express 4(2), 190–197 (2014). [CrossRef]  

17. R. L. Frost, K. L. Erickson, M. L. Weier, A. R. McKinnon, P. A. Williams, and P. Leverett, “Use of infrared spectroscopy for the determination of electronegativity of rare earth elements,” Appl. Spectrosc. 58(7), 811–815 (2004). [CrossRef]   [PubMed]  

18. X. Y. Wu, H. P. You, H. T. Cui, X. Q. Zeng, G. Y. Hong, C. H. Kim, C. H. Pyun, B. Y. Yu, and C. H. Park, “Vacuum ultraviolet optical properties of (La,Gd)PO4: RE3+ (RE = Eu, Tb),” Mater. Res. Bull. 37(9), 1531–1538 (2002). [CrossRef]  

19. F. Baur and T. Juestel, “New red-emitting phosphor La2Zr3(MoO4)9:Eu3+ and the influence of host absorption on its luminescence efficiency,” Aust. J. Chem. 68, 1727–1734 (2015). [CrossRef]  

20. W. M. Yen, S. Shionoya, and H. Yamamoto, Phosphor Handbook (CRC Press, Boca Raton, 2007).

21. B. S. Tsai, Y. H. Chang, and Y. C. Chen, “Synthesis and luminescent properties of MgIn2-xGaxO4: Eu3+ phosphors,” Electrochem. Solid St. 8(7), H55–H57 (2005). [CrossRef]  

22. G. Q. Wang, X. H. Gong, Y. J. Chen, J. H. Huang, Y. F. Lin, Z. D. Luo, and Y. D. Huang, “Novel red phosphors KBaEu(XO4)3 (X = Mo, W) show high color purity and high thermostability from a disordered chained structure,” Dalton Trans. 46(20), 6776–6784 (2017). [CrossRef]   [PubMed]  

23. L. Yi, X. He, L. Zhou, F. Gong, R. Wang, and J. Sun, “A potential red phosphor LiGd(MoO4)2:Eu3+ for light-emitting diode application,” J. Lumin. 130(6), 1113–1117 (2010). [CrossRef]  

24. V. Morozov, A. Arakcheeva, B. Redkin, V. Sinitsyn, S. Khasanov, E. Kudrenko, M. Raskina, O. Lebedev, and G. Van Tendeloo, “Na(2/7)Gd(4/7)MoO4: a modulated scheelite-type structure and conductivity properties,” Inorg. Chem. 51(9), 5313–5324 (2012). [CrossRef]   [PubMed]  

25. V. A. Morozov, B. I. Lazoryak, S. Z. Shmurak, A. P. Kiselev, O. I. Lebedev, N. Gauquelin, J. Verbeeck, J. Hadermann, and G. Van Tendeloo, “Influence of the structure on the properties of NaxEuy(MoO4)z red phosphors,” Chem. Mater. 26(10), 3238–3248 (2014). [CrossRef]  

26. K. Binnemans, “Interpretation of europium (III) spectra,” Coord. Chem. Rev. 295, 1–45 (2015). [CrossRef]  

27. D. Dexter and J. H. Schulman, “Theory of concentration quenching in inorganic phosphors,” J. Chem. Phys. 22(6), 1063–1070 (1954). [CrossRef]  

28. K. R. Reddy, K. Annapurna, and S. Buddhudu, “Fluorescence spectra of Eu3+:Ln2O2S (Ln=Y, La, Gd) powder phosphors,” Mater. Res. Bull. 31(11), 1355–1359 (1996). [CrossRef]  

29. M. Kneissl, D. W. Treat, M. Teepe, N. Miyashita, and N. M. Johnson, “Advances in InAlGaN laser diode technology towards the development of UV optical sources,” Proc. SPIE 4995, 103–107 (2003). [CrossRef]  

30. T. Taguchi, “Present status and future prospect of system and design in white LED lighting technologies,” Proc. SPIE 5530, 7–16 (2004). [CrossRef]  

31. K. A. Denault, M. Cantore, S. Nakamura, S. P. DenBaars, and R. Seshadri, “Efficient and stable laser-driven white lighting,” AIP Adv. 3(7), 072107 (2013). [CrossRef]  

32. Y. C. Chang, C. H. Liang, S. A. Yan, and Y. S. Chang, “Synthesis and photoluminescence characteristics of high color purity and brightness Li3Ba2Gd3(MoO4)8:Eu3+ red phosphors,” J. Phys. Chem. C 114(8), 3645–3652 (2010). [CrossRef]  

33. W. H. Fonger and C. W. Struck, “Eu+3 5D resonance quenching to charge-transfer states in Y2O2S, La2O2S, and LaOCl,” J. Chem. Phys. 52(12), 6364–6372 (1970). [CrossRef]  

34. D. Wen, J. Feng, J. Li, J. Shi, M. Wu, and Q. Su, “K2Ln(PO4)(WO4):Tb3+, Eu3+ (Ln = Y, Gd and Lu) phosphors: highly efficient pure red and tuneable emission for white light-emitting diodes,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(9), 2107–2114 (2015). [CrossRef]  

35. F. Meng, X. Zhang, S. I. Kim, Y. M. Yu, and H. J. Seo, “Luminescence properties of Eu3+ in gadolinium molybdate β′-Gd2Mo3O12 phosphors,” Optik (Stuttg.) 125(14), 3578–3582 (2014). [CrossRef]  

36. L. Li, W. Chang, W. Chen, Z. Feng, C. Zhao, P. Jiang, Y. Wang, X. Zhou, and A. Suchocki, “Double perovskite LiLaMgWO6:Eu3+ novel red-emitting phosphors for solid state lighting: synthesis, structure and photoluminescent properties,” Ceram. Int. 43(2), 2720–2729 (2017). [CrossRef]  

37. C. Qin, Y. Huang, and H. J. Seo, “The thermal stability and structural site-distribution of Eu3+ ions in the red-emitting phosphors Ca9Eu2W4O24 and Sr9Eu2W4O24,” J. Alloys Compd. 534, 86–92 (2012). [CrossRef]  

38. P. Du, Y. Guo, S. H. Lee, and J. S. Yu, “Broad near-ultraviolet and blue excitation band induced dazzling red emissions in Eu3+-activated Gd2MoO6 phosphors for white light-emitting diodes,” Rsc Adv. 7(6), 3170–3178 (2017). [CrossRef]  

39. J. Huang, B. Hou, H. Ling, J. Liu, and X. Yu, “Crystal structure, electronic structure, and photoluminescence properties of La3BW1-xMoxO9:Eu3+ red phosphor,” Inorg. Chem. 53(18), 9541–9547 (2014). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (10)

Fig. 1
Fig. 1 (a) XRD patterns of the LGM:xEu3+ (x = 0, 50, and 90 at.%) phosphors, and the corresponding magnified XRD curves in the ranges of 10-28° (b) and 28-29.5° (c).
Fig. 2
Fig. 2 (a) SEM micrograph and (b) EDS data of the LGM:90at.%Eu3+ phosphor.
Fig. 3
Fig. 3 The mapping EDS images of Mo (b), Eu (c) and Gd (d) corresponding to the particle shown in (a).
Fig. 4
Fig. 4 (a) Room-temperature diffuse reflection spectra of the LGM, LGM:50at.%Eu3+ and LGM:90at.%Eu3+ phosphors in 200-800 nm, (b) Room-temperature excitation spectra of the LGM:10at.%Eu3+ and LGM:90at.%Eu3+phosphors.
Fig. 5
Fig. 5 (a) Room-temperature emission spectra of the LGM:xEu3+ (x = 10, 50, 90 at.%) phosphors excited at 395 nm. and the enlarged region in the inset. (b) the dependence of integrated emission intensity of the 5D07F2 transition and the (5D07F2)/(5D07F1) emission ratio R on the Eu3+ concentration.
Fig. 6
Fig. 6 (a) Excitation and (b) emission spectra of the LGM:90at.%Eu3+ and commercial Y2O2S:6.3at.%Eu3+ red phosphors.
Fig. 7
Fig. 7 Fluorescence decay curves of the LGM:xEu3+ (x = 10, 30, 50, 70, 80, 90, and 100 at.%) phosphors.
Fig. 8
Fig. 8 Temperature dependence of fluorescence for the LGM:90at.%Eu3+ phosphor under excitation at 395 nm. The inset shows the temperature dependence of emission intensity of the 5D07F2 transition under excitation at 395 nm and 465 nm, respectively.
Fig. 9
Fig. 9 (a) Pathway for the thermal quenching of the 5D0 state through a CTS band, (b) The dependence of ln(I0/I-1) on 1/kT for the LGM:90at.%Eu3+ phosphor.
Fig. 10
Fig. 10 CIE chromaticity coordinate of the LGM:90at.%Eu3+ phosphor at 395 nm excitation.

Equations (2)

Equations on this page are rendered with MathJax. Learn more.

I= I 0 exp(t/τ)
ln( I 0 /I 1)=lnA ΔE/ kT
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.