A novel borate phosphor Lu5Ba6B9O27:Ce3+ codoped with Sr2+/Tb3+ for NUV-white light emitting diode application

Chenggang Ma a, Hailiang Chen b, Min Luo a, Fuyun Duan b, Yun Ding a, Yihang Han c, Tianxiang Zheng a, Xun Yang d and Yu Xiao *a
aCollege of Science, Nanjing Forestry University, Nanjing 210037, P. R. China. E-mail: xiaoyulaser@yeah.net
bShandong INOV Polyurethane CO., Ltd, Zibo 255000, P. R. China
cCollege of Chemistry and Molecular Engineering, Peking University, Beijing 100871, P. R. China
dCollege of Materials Science and Engineering, Nanjing Forestry University, Nanjing 210037, P. R. China

Received 26th June 2024 , Accepted 17th July 2024

First published on 27th July 2024


Abstract

At present, there are still challenges in developing highly efficient and thermally stable phosphors for ultraviolet/near ultraviolet-white light emitting diodes (UV/NUV-WLEDs). Herein, we use traditional high-temperature solid-state reactions to prepare blue-emitting phosphors Lu5−xBa6B9O27:xCe3+ (0.1% ≤ x ≤ 3.0%) and Lu4.975Ba6−ySryB9O27:0.5%Ce3+ (0% ≤ y ≤ 20%), and green-emitting phosphors Lu4.975−zBa6B9O27:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%), abbreviated as LBB:xCe3+, LBB:0.5%Ce3+,ySr2+ and LBB:0.5%Ce3+,zTb3+, respectively. Upon 340 nm excitation, LBB:Ce3+ exhibits an asymmetric blue emission ranging from 360 nm to 480 nm. Furthermore, the emission intensity of LBB:0.5%Ce3+ increased 4.8-fold without a spectral shift through the partial substitution of Sr2+ for Ba2+. Through constructing the Ce3+ → Tb3+ energy transfer in the LBB structure, the temperature-dependent integral emission intensity at 483 K improved from only 25% to 68% of the intensity at 303 K due to the existence of fast energy transfer from Ce3+ to Tb3+. The related results indicate that LBB:xCe3+, LBB:0.5%Ce3+,ySr2+ and LBB:0.5%Ce3+,zTb3+ phosphors can be used for UV/NUV-WLEDs.


1. Introduction

Phosphor-converted white light emitting diodes (pc-WLEDs) have become the next-generation solid-state lighting because of their advantages such as environmental friendliness, high efficiency, and low cost, among other advantages.1–4 Currently, the commercialized pc-WLEDs combine blue LEDs with the yellow phosphor (Y,Gd)3Al5O12:Ce3+ (YAG:Ce3+). However, these pc-WLED devices still present several shortcomings, such as a lower color rendering index (Ra < 80), a high correlated color temperature (CCT > 5000 K), and harmful blue light.5,6 Alternatively, using ultraviolet (UV) or near-ultraviolet (NUV) LEDs to excite red-, green-, and blue-emitting phosphors may solve these shortcomings, since their white light is entirely generated by phosphors that contain the entire visible range with a flat spectral distribution.7,8 Nevertheless, several commercial phosphors for UV/NUV-WLEDs also have drawbacks, such as the harsh preparation conditions of K2SiF6:Mn4+ (red),9 strong thermal quenching of (Ba,Sr)2SiO4:Eu2+ (green),10 and easy aging of BaMgAl10O17:Eu2+ (blue).11 The phosphors determine the final performance of pc-WLEDs. Thus, it is still essential to develop efficient and thermally stable phosphors for UV/NUV-WLEDs.

The performance of phosphors mainly depends on two factors: the activators, which are usually Ce3+/Eu2+ due to their strong absorption from the spin-allowed transition 4f → 5d, and the host, which generally should have a large bandgap with structural rigidity.7,8 Commercial LED phosphors not only require high efficiency and excellent thermal stability, but also must be low cost and environmentally friendly. Therefore, borates are always considered to be an optimal host due to their low synthesis temperature, high UV transparency, and expected chemical stability.12 In recent years, several highly efficient borate phosphors (quantum efficiency, QE > 50%) have been designed by doping Ce3+ into the RE2O3-BaO-B2O3 (RE = Y, Lu) phase system; for instance, Y5Ba2B5O17:Ce3+ (λex = 365 nm; λem = 443 nm; FWHM ≈ 100 nm),11 Lu5Ba2B5O17:Ce3+ (λex = 340 nm; λem = 447 nm; FWHM = 104 nm),13 Y2Ba3B6O15:Ce3+ (λex = 340 nm; λem = 446 nm; FWHM = 70 nm),14 and Lu2Ba3B6O15:Ce3+ (λex = 473 nm; λem = 446 nm; FWHM = 68 nm).15,16 Recently, S. K. Filatov et al. developed a novel Lu5Ba6B9O27 structure through the RE2O3-BaO-B2O3 phase system.17 Thus, doping Ce3+ ions into a Lu5Ba6B9O27 structure to obtain novel phosphors is highly anticipated for LED applications.

The unshielded 5d energy levels of Ce3+/Eu2+ ions (i.e., the centroid shift εc and the crystal field splitting εcfs) are sensitive to local crystal fields. Typically, alterations in the local crystal fields can be achieved through cation substitution, which can also cause a shift in the PL spectra.7,8 In the LBB structure, a solid solution may be formed by cation substitution upon replacing Ba2+ (r = 1.35 Å; CN = 6) with Sr2+ (r = 1.18 Å; CN = 6). In most cases, as Sr2+ is gradually substituted for Ba2+, a red-shift in the PL spectrum occurs as the crystal field splitting becomes stronger, such as the emission peaks of (Ba1−xSrx)9Sc2Si6O24:Eu2+ tuned from 506 nm to 526 nm,18 the emission spectra of (Ba1−xSrx)9Sc2Si6O24:Ce3+ red-shifted by 20 nm,19 and the emission of SrxBa2−xSiO4:Eu2+ with a red-shift of ∼65 nm.20 In contrast, the emission peaks of (Ba,Sr)3Lu(PO4)3:Eu2+ show a significant blue-shift from 506 nm to 479 nm due to the neighboring-cation effect upon increasing the ratio of Sr/Ba.21 In addition, the emission color also can be adjusted by constructing Ce3+ → Tb3+ energy transfer in the host. Importantly, previous studies have shown that a fast energy transfer from Ce3+ to Tb3+ in the RE2O3-BaO-B2O3 system results in the improvement of luminescence thermal stability.22–24 Therefore, it is also quite possible to construct a Ce3+ → Tb3+ energy transfer with a fast process in the LBB host.

To the best of our knowledge, no relevant research has yet focused on the Ce3+/Tb3+-activated Lu5Ba6B9O27 (LBB) structure. Herein, we prepare a series of blue phosphors Lu5−xBa6B9O27:xCe3+ (0.1% ≤ x ≤ 3.0%) and Lu4.975Ba6−yB9O27:0.5%Ce3+,ySr (0% ≤ y ≤ 20%), and green phosphors Lu4.975−zBa6B9O27:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%) abbreviated as LBB:xCe3+, LBB:0.5%Ce3+,ySr2+ and LBB:0.5%Ce3+,zTb3+, respectively. The emission intensity of LBB:0.5% increases more than four-fold without a spectral shift and its internal QE increases from 20% to 95% when Ba2+ is partially replaced with Sr2+. Then, we improved the thermal stability of the emission by constructing energy transfer from Ce3+ → Tb3+ in the LBB structure. All results indicate that LBB:xCe3+, LBB:0.5%Ce3+,ySr2+ and LBB:0.5%Ce3+,zTb3+ phosphors can be used for UV/NUV-WLEDs.

2. Experimental and methodology

2.1 Materials and synthesis

The phosphors LBB:xCe3+ (0.1% ≤ x ≤ 3.0%), LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%), and LBB:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%) were synthesized by traditional high-temperature solid-state reactions. The starting materials Ba2CO3 (A.R.), Sr2CO3 (A.R.), H3BO3 (A.R.), CeO2 (4N), and Tb4O7 (4N) were mixed and ground according to the given stoichiometric ratio. The weighed powders were mixed in an agate mortar and placed in an alumina crucible, and then fired at 450 °C for 4 h in air. Finally, the mixtures were heated at 1200 °C for 8 h in a reducing atmosphere (10% H2–90% N2).

2.2 Measurements and characterization

The X-ray diffraction (XRD) patterns were collected from a multifunctional horizontal X-ray diffractometer (Ultima IV, Rigaku, Japan) with Cu Kα radiation (λ = 1.5418 Å). The high-resolution transmission electron microscopy (HRTEM) images and component elemental maps were obtained using an FEI Tecnai G2 S-Twin transmission electron microscope with a field emission gun operating at 200 kV. X-ray photoelectron spectroscopy (XPS) spectra were measured with an XPS spectrometer (Thermo Kalpha). Photoluminescence (PL) and photoluminescence excitation (PLE) spectra were recorded on an Edinburg FLS-980 fluorescence spectrophotometer. A pulsed laser from an optical parametric oscillator and the electric signal detected using a Tektronix digital oscilloscope TDS 3052 were used to measure the fluorescence decay. The quantum efficiency (QE) was obtained using an absolute PL quantum yield measurement system (Quantaurus-QY Plus C13534-12, Hamamatsu Photonics).

3. Results and discussion

3.1 Optimizing the doping concentration of Ce3+ in the LBB structure

The LBB structure belongs to the space group C2/c and consists of edge- and corner-sharing Lu-, Ba-, and B-centered polyhedra, which form a three-dimensional net, as depicted in Fig. 1(a) and S1. The [LuO6] octahedra are connected to each other by their vertices. The B atoms are surrounded by three O atoms, which form [BO3] triangles. There are three configurations of Ba atoms surrounded by O atoms, and they form distorted polyhedra [Ba1O9], [Ba2O8], and [Ba2O7]. All the Ba–O polyhedra are connected to each other by their common edges and to the [BO3] triangles and [LuO6] octahedra by their common vertices and edges. As shown in Fig. 1(b) and S2, the XRD patterns of LBB:xCe3+ (0.1% ≤ x ≤ 3.0%) phosphors match well with the standard card published by S. K. Filatov, et al.,17 indicating that Ce3+ ions were completely dissolved in the LBB structure.
image file: d4dt01843e-f1.tif
Fig. 1 Crystal structure of LBB (a); XRD patterns of LBB:xCe3+ (x = 0.1% and 3%) phosphors (b).

Fig. 2(a) shows the PLE and PL spectra of LBB:0.1%Ce3+. The PLE spectrum consists of two bands that range from 250 nm to 380 nm in the ultraviolet (UV) region, which are ascribed to the allowed transitions 5d–4f of Ce3+ ions. Upon 340 nm excitation, LBB:0.01Ce3+ exhibits an asymmetric blue emission band that can be well fit with two Gaussian bands (dashed) peaking at 383 nm (26[thin space (1/6-em)]109 cm−1) and 411 nm (24[thin space (1/6-em)]330 cm−1). Their energy difference is about 2000 cm−1, being consistent with the energy separation between the sub-states 2F7/2 and 2F5/2 of Ce3+. Fig. 2(b) shows the PL spectra depending on the Ce3+ concentration of LBB:xCe3+ (0.1% ≤ x ≤ 3.0%). The maximal PL intensity occurs around 0.5% Ce3+, and then the PL intensity gradually decreases with the increase of Ce3+ content due to the well-known concentration quenching effect.25


image file: d4dt01843e-f2.tif
Fig. 2 PLE and PL spectra of LBB:0.1Ce3+ (a); PL spectra (b) and fluorescence decay curves (c) depending on the Ce3+ concentration of LBB:xCe3+ (0.1% ≤ x ≤ 3.0%); linear fitting of log(I/x) versus log(x) for LBB:xCe3+ (0.5% ≤ x ≤ 3.0%) (d).

Here, the critical distance (Rc) is defined as the average distance between Ce3+ ions when luminescence quenching occurs. Rc can be estimated from geometrical consideration by following eqn (1):26

 
image file: d4dt01843e-t1.tif(1)
where V represents the volume of the unit cell, Xc represents the critical concentration of Ce3+ for concentration quenching, and N represents the number of total Ce3+ sites in the unit cell. In this study, V = 2566.9 Å3, N = 20, and Xc = 0.05 for LBB:Ce3+. Thus, the Rc of Ce3+ ions was calculated to be 36.6 Å3.

The fluorescence decay curves of LBB:xCe3+ (0.1% ≤ x ≤ 3.0%) are depicted in Fig. 2(c). The fluorescence lifetime of LBB:0.1%Ce3+ is calculated to be 23.2 ns, probably reflecting intrinsic decay of Ce3+ in the LBB structure. The lifetime of Ce3+ gradually decreases with increasing Ce3+ content (%), which is probably caused by the enhanced nonradiative rates and energy transfer among Ce3+ ions in the LBB structure. The energy transfer mechanism of Ce3+ in the LBB structure is governed by the Dexter theory.27 The activator (Ce3+) concentration satisfies eqn (2):28,29

 
image file: d4dt01843e-t2.tif(2)
where I is the PL intensity, y is the Ce3+ concentration larger than Xc = 0.05, k and β are constants, and θ is an indication of the type of electric multipolar interaction. The values of θ are 6, 8, and 10, standing for the energy transfer mechanism of electric dipole–dipole, dipole–quadrupole, and quadrupole–quadrupole interactions, respectively. As shown in Fig. 2(d), the relationship of log(I/x) versus log(x) is well fit by a linear function with a slope –(θ/3) of 1.7. Thus, the value of θ is determined to be 5.1, implying that the concentration quenching of Ce3+ ions in the LBB structure mainly results from the electric dipole–dipole interactions since the θ = 5.1 approximates to 6.

3.2 Substituting Sr2+ for Ba2+ in LBB:Ce3+ increases the emission intensity by 4.8-fold

The XRD patterns of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) phosphors are shown in Fig. 3(a). With an increase in the Sr2+ content (x), the diffraction peaks gradually shift to higher angles as shown in the magnified XRD patterns (29°–31°), which indicates that Sr2+ replaces Ba2+ ions in the LBB structure with contraction of the unit cell. The analysis results of HRTEM and XPS are presented in Fig. 3(b) and S3–S5. The HRTEM image shows a clear lattice fringe, which indicates that the samples possess fine crystallization (Fig. S3). The elemental mapping and XPS results (Fig. S4 and S5) indicate that the LBB:0.5%Ce3+,20%Sr2+ compound contains strontium (Sr), barium (Ba), lutetium (Lu), cerium (Ce), boron (B) and oxygen (O) with a homogeneous distribution.
image file: d4dt01843e-f3.tif
Fig. 3 The XRD patterns of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) phosphors, with magnification around 29°–31° (a); elemental mapping images of LBB:0.5%Ce3+,20%Sr2+ (b).

Fig. 4 shows the PLE, PL spectra, and decay curves of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) and luminescence thermal stability of LBB:0.5%Ce3+,20%Sr2+. In general, Sr2+ is substituted for Ba2+ to form a solid solution resulting in a red-shift or blue-shift of PL spectra, due to the variation in crystal field splitting (εcfs) and centroid shift (εc) of the activators (Ce3+/Eu2+). However, in the LBB structure, as Sr2+ ions gradually replace Ba2+ ions, there is no obvious shift in the PLE or PL spectra, but the emission intensity of LBB:0.5%Ce3+,20%Sr2+ increased by 4.8-fold compared with that of LBB:0.5%Ce3+. Fig. 4(c) exhibits the decay curves of the phosphor LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%), which are monitored at 410 nm and upon pulse excitation at 340 nm. The lifetime slightly increased from 23.5 ns (y = 0) to 25.3 ns (y = 20%), probably due to the decrease in the nonradiative transition probability with the increase in Sr2+ content (y). We consider that the structural rigidity is probably enhanced by the substitution of Sr2+ for Ba2+ in the LBB structure, which results in an increase in the lifetime.30 To verify the hypothesis, we measured the QEs of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%). The IQE, absorption efficiency (AE), and EQE values are calculated using eqn (3)–(5):31,32

 
image file: d4dt01843e-t3.tif(3)
 
image file: d4dt01843e-t4.tif(4)
 
ηEQE = ηIQE × εAE,(5)
where Ls refers to integration of the emission spectrum of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) and ER and ES refer to spectra of excitation light with and without the samples in the integrated sphere, respectively. As shown in Fig. 5, the internal QE increased from 20% (y = 0%) to 95% (y = 20%), but there is no significant change in the absorption efficiency (AE) values. In addition, the luminescence thermal stability is another key parameter for LED phosphors. Thus, we measured the PL spectra of the LBB:0.5%Ce3+,20%Sr2+ phosphor and the corresponding relative integrated emission intensities at different temperatures. As depicted in Fig. 4(d), increasing the temperature from 303 K to 483 K causes small changes in the profile of the spectra; however, the luminescence intensity at 483 K is only 25% of the intensity at room temperature (303 K). To better understand luminescence thermal stability of LBB:0.5%Ce3+,20%Sr2+, the activation energy was calculated using the Arrhenius equation:33
 
image file: d4dt01843e-t5.tif(6)
where I0 is the PL intensity at 0 K, I(T) is the PL intensity at a given temperature T, A is a constant, Ea is the activation energy for thermal quenching, and kB is the Boltzmann constant. The experimental data are well fit using eqn (6). The value of Ea is determined to be 0.31 eV for LBB:0.5%Ce3+,20%Sr2+ as shown in Fig. 4(d).


image file: d4dt01843e-f4.tif
Fig. 4 PLE and PL spectra (a and b); decay curves of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) phosphors (c); the temperature dependent PL spectra and relative integrated PL intensities of the LBB:0.5%Ce3+,20%Sr2+ phosphor (d).

image file: d4dt01843e-f5.tif
Fig. 5 Emission spectra of LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%) for QE measurements in the absence and presence of the phosphor under 340 nm excitation (a–d).

3.3 Tb3+-doping into LBB:Ce3+ phosphors improves the luminescence thermal stability

Constructing the energy transfer from Ce3+ to Tb3+ is one of the potential solutions to improve the poor thermal stability of LBB:Ce3+ phosphors.22–24,34 Thus, we prepared a series of LBB:0.5%Ce3+,ZTb3+ (0% ≤ z ≤ 16%) phosphors. LBB:0.5%Ce3+,ZTb3+ were characterized by XRD and no obvious impurity was detected (see Fig. 6(a) and S6). As shown in Fig. 6(b, c) and S7–S9, the elemental mapping, HRTEM images, and XPS results indicate that the dopants were homogeneously dissolved in the LBB host with fine crystallization.
image file: d4dt01843e-f6.tif
Fig. 6 XRD patterns of LBB:0.5%Ce3+,zTb3+ (z = 1%, 8%, and 16%) (a); elemental mapping images (b); and the XPS spectrum of the LBB:0.5%Ce3+,16%Tb3+ phosphor (c).

The PLE and PL spectra of LBB:0.5%Ce3+, LBB:8%Tb3+, and LBB:0.5%Ce3+,8%Tb3+ phosphors are depicted in Fig. 7(a). Clearly, there is a spectral overlap between the PL spectrum of LBB:0.5%Ce3+ and the PLE spectrum corresponding to the 7F65D3 absorption of LBB:8%Tb3+, indicating the possibility of energy transfer from Ce3+ to Tb3+. The energy transfer was confirmed by the luminescence properties of LBB:0.5%Ce3+,8%Tb3+. The PLE spectrum of Tb3+ emission at 543 nm exhibits the characteristic 4f–5d transition band of the Ce3+ ion. Moreover, upon Ce3+ excitation at 340 nm, strong Tb3+ emission peaks at 488, 543, 585 and 625 nm corresponding to the 5D47FJ (J = 6, 5, 4, and 3) transitions appear in the PL spectrum. To investigate the effect of energy transfer, the spectra with their corresponding CIE diagram and the fluorescence decay curves of LBB:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%) phosphors are depicted in Fig. 7(b–d) and S10. With the increase in Tb3+ content (z), the Ce3+ emission continuously decreases with the enhancement of the Tb3+ emission. Thus, the color tuning from blue to green occurs and is shown in the Commission Internationale de L'Eclairage (CIE) chromaticity diagram (Fig. 7(c)). In the LBB structure, there is also a fast energy transfer from Ce3+ to Tb3+, which can be confirmed by the Ce3+ lifetime of LBB:0.5%Ce3+,zTb3+. In general, the energy transfer efficiency (ηET) of Ce3+ → Tb3+ can be calculated as follows:35–38

 
image file: d4dt01843e-t6.tif(7)
where IS0 and IS denote the luminescence intensity of Ce3+ in the absence and in the presence of Tb3+, respectively. The energy transfer efficiencies listed in Table. S1 gradually increase with the increase of Tb3+ concentration in LBB:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%). As the doping concentration (z) of Tb3+ reaches 16%, the energy transfer efficiency (ηET) of Ce3+ → Tb3+ is determined to be 77%. We also use Ce3+ fluorescence lifetimes to calculate the energy transfer efficiency (ηET):
 
image file: d4dt01843e-t7.tif(8)
where τS and τS0 are the lifetimes of the sensitizer (Ce3+) with and without the activator (Tb3+), respectively. However, the Ce3+ lifetime only decreased from 23.5 ns in LBB:0.5%Ce3+ to 14.2 ns in LBB:0.5%Ce3+,16%Tb3+. Thus, the energy transfer efficiency (ηET) of Ce3+ to Tb3+ in LBB:0.5%Ce3+,16%Tb3+ is calculated to be 40%. The reduction of emission intensity of Ce3+ is larger than that of lifetime, indicating the existence of fast energy transfer among some Ce3+ ions in LBB:Ce3+,Tb3+. Such a small decrease implies that there is a fast decay of Ce3+ that is too rapid to detect (see Fig. S11). A similar fast energy transfer may occur in the nearest Ce3+–Tb3+ pairs, which has been demonstrated in previous work.22–24,34,38


image file: d4dt01843e-f7.tif
Fig. 7 PLE and PL spectra of LBB:0.5%Ce3+, LBB:8%Tb3+, and LBB:0.5%Ce3+,8%Tb3+ (a); PL spectra (b); emission color (c); and Ce3+ lifetimes of LBB:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%) (d).

As depicted in Fig. 8(a), the energy transfer rate (WET) is closely related to the distance between Ce3+ and Tb3+. Thus, we use f (WET) representing the distribution function of the energy transfer rate WET, and the energy transfer efficiency can be written as follows:24

 
image file: d4dt01843e-t8.tif(9)


image file: d4dt01843e-f8.tif
Fig. 8 The energy transfer model for Ce3+ → Tb3+ (a and b); temperature-dependent PL spectra (c) and relative integrated PL intensities (d) of LBB:0.5%Ce3+,16%Tb3+; schematic diagram of the preparation of a pc-WLED (e); emission spectrum of a NUV-chip (370 nm)-based pc-WLED using BaMgAl10O17:Eu2+ (blue), LBB:Ce3+,Tb3+ (green), and K2SiF6:Mn4+ (red). The inset shows an image of the WLED package (f).

From eqn (9), we can find that there is a competition among energy transfer (WET), intrinsic decay (γ), and thermal activation (ΔE) after exciting Ce3+ ions. For simplicity, as depicted in Fig. 8(b), the process of energy transfer from Ce3+ to Tb3+ can be roughly classified into two groups in terms of energy transfer rates (the fast and the slow). The slow energy transfer (process ①) cannot compete with the thermal activation (ΔE), but the fast energy transfer (process ②) of Ce3+ → Tb3+ is so rapid that it cannot be detected in the fluorescence decay pattern (Fig. 7(d) and S11). Owing to the fast energy transfer, there is a significant improvement in the luminescence thermal stability of LBB:Ce3+,Tb3+ due to the intrinsic properties of f–f transitions of Tb3+ with a large energy gap (about 15[thin space (1/6-em)]000 cm−1) between the emitting level 5D4 and its next lower level 7F1.22–24,39 The luminescence intensity of LBB:0.5%Ce3+,16%Tb3+ at 483 K is improved to 68% of the intensity at room temperature (303 K), as shown in Fig. 8(c and d). Finally, we use LBB:0.5%Ce3+,16%Tb3+ as a green phosphor and the commercial phosphors BaMgAl10O17:Eu2+ (BAM:Eu2+; blue) and K2Si4F6:Mn4+ (KSF:Mn4+; red) to fabricate a NUV-WLED. As depicted in Fig. 8(e and f), the RGB phosphors are fully mixed with epoxy resin and covered on a NUV-LED chip (365 nm). Then, the prototype NUV-WLED should be solidified at 200 °C for 20 min. The correlated color temperature (CCT) and color rendering index (Ra) of the as-fabricated NUV-WLED are found to be Ra = 82 and CCT = 4082 K, indicating the potential of LBB:Ce3+,Tb3+ phosphors for LED application.

4. Conclusion

In summary, we prepared a series of blue phosphors LBB:xCe3+ (0.1% ≤ x ≤ 3.0%) and LBB:0.5%Ce3+,ySr2+ (0% ≤ y ≤ 20%), and green phosphors LBB:0.5%Ce3+,zTb3+ (0% ≤ z ≤ 16%) via a high-temperature solid-state reaction. In the LBB structure, the luminescence intensity increased by 4.8-fold with no spectral shift upon the substitution of Ba2+ with Sr2+. The emission intensity of LBB:0.5%Ce3+ at 483 K is only 25% of the intensity at 303 K, which improves to 68% when the Ce3+ → Tb3+ energy transfer is designed in the LBB structure. The doping amount of Sr2+ and Tb3+ ions should not be too large and they cannot be simultaneously doped into the LBB structure, otherwise the purity of the crystal structure will be lost. In the future, we will explore ways to stabilize the LBB structure and establish structure–property relationships between the local crystal structure and PL properties using characterization methods such as Rietveld refinement, synchrotron radiation, and low-temperature spectroscopy, which may provide new ideas for the development of efficient and thermally stable phosphors for LED applications.

Data availability

The data used to support the findings of this study are available from the corresponding author upon request.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was partially supported by the National Natural Science Foundation of China (Grant No. 12104231), the Shandong Postdoctoral Fund (Grant No. SDCX-ZG-202203064), and the Innovation Training Program for College Students in Nanjing Forestry University (Grant No. 202410298184Y).

References

  1. P. Dang, G. Li, X. Yun, Q. Zhang, D. Liu, H. Lian, M. Shang and J. Lin, Light: Sci. Appl., 2021, 10, 29 CrossRef CAS PubMed .
  2. H. Jin, N. Fu, C. Wang, C. Qi, Z. Liu, D. Wang, L. Guan, F. Wang and X. Li, Dalton Trans., 2023, 52, 787–795 RSC .
  3. X. Zou, X. Wang, H. Zhang, Y. Kang, X. Yang, X. Zhang, M. S. Molokeev and B. Lei, Chem. Eng. J., 2022, 428, 132003 CrossRef CAS .
  4. S. Wei, Z. Lyu, Z. Lu, P. Luo, L. Zhou, D. Sun, T. Tan, S. Shen and H. You, Chem. Mater., 2023, 35, 7125–7132 CrossRef CAS .
  5. P. D. Baheti, R. A. Talewar and S. V. Moharil, Mater. Lett., 2023, 348, 134727 CrossRef CAS .
  6. Z. Huang, Z. Lyu, D. Sun, S. Shen, Z. Lu, L. Zhou, S. Wei and H. You, Adv. Opt. Mater., 2023, 11(24), 2301172 CrossRef CAS .
  7. G. Li, Y. Tian, Y. Zhao and J. Lin, Chem. Soc. Rev., 2015, 44, 8688–8713 RSC .
  8. S. Wang, Z. Song and Q. Liu, J. Mater. Chem. C, 2023, 11, 48–96 RSC .
  9. H. Zhu, C. C. Lin, W. Luo, S. Shu, Z. Liu, Y. Liu, J. Kong, E. Ma, Y. Cao, R.-S. Liu and X. Chen, Nat. Commun., 2014, 5, 4312 CrossRef CAS PubMed .
  10. Q. Shao, H. Lin, Y. Dong and J. Jiang, J. Lumin., 2014, 151, 165–169 CrossRef CAS .
  11. M. Hermus, P.-C. Phan and J. Brgoch, Chem. Mater., 2016, 28, 1121–1127 CrossRef CAS .
  12. S. Verma, K. Verma, D. Kumar, B. Chaudhary, S. Som, V. Sharma, V. Kumar and H. C. Swart, Phys. B, 2018, 535, 106–113 CrossRef CAS .
  13. M. Hermus, P.-C. Phan, A. C. Duke and J. Brgoch, Chem. Mater., 2017, 29, 5267–5275 CrossRef CAS .
  14. A. C. Duke, S. Hariyani and J. Brgoch, Chem. Mater., 2018, 30, 2668–2675 CrossRef CAS .
  15. B. Li, G. Annadurai, L. Sun, J. Liang, S. Wang, Q. Sun and X. Huang, Opt. Lett., 2018, 43, 5138–5141 CrossRef CAS PubMed .
  16. Z. Sun, Z. Zhu, Z. Guo, Z.-c. Wu, Z. Yang, T. Zhang and X. Zhang, Ceram. Int., 2019, 45, 7143–7150 CrossRef CAS .
  17. S. K. Filatov, Y. P. Biryukov, R. S. Bubnova and A. P. Shablinskii, Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater., 2019, 75, 697–703 CrossRef CAS PubMed .
  18. L. Bian, T. L. Zhou, J. J. Yang, Z. Song and Q. L. Liu, J. Lumin., 2012, 132, 2541–2545 CrossRef CAS .
  19. J. Brgoch, C. K. H. Borg, K. A. Denault, S. P. DenBaars and R. Seshadri, Solid State Sci., 2013, 18, 149–154 CrossRef CAS .
  20. K. A. Denault, J. Brgoch, M. W. Gaultois, A. Mikhailovsky, R. Petry, H. Winkler, S. P. DenBaars and R. Seshadri, Chem. Mater., 2014, 26, 2275–2282 CrossRef CAS .
  21. Z. Wang, Z. Xia, M. S. Molokeev, V. V. Atuchin and Q. Liu, Dalton Trans., 2014, 43, 16800–16804 RSC .
  22. D. Wu, W. Xiao, L. Zhang, X. Zhang, Z. Hao, G.-H. Pan, Y. Luo and J. Zhang, J. Mater. Chem. C, 2017, 5, 11910–11919 RSC .
  23. Y. Xiao, Z. Hao, L. Zhang, W. Xiao, D. Wu, X. Zhang, G.-H. Pan, Y. Luo and J. Zhang, Inorg. Chem., 2017, 56, 4538–4544 CrossRef PubMed .
  24. Y. Xiao, Z. Hao, L. Zhang, X. Zhang, G.-H. Pan, H. Wu, H. Wu, Y. Luo and J. Zhang, J. Mater. Chem. C, 2018, 6, 5984–5991 RSC .
  25. D. L. Dexter and J. H. Schulman, J. Chem. Phys., 1954, 22, 1063–1070 CrossRef CAS .
  26. G. Blasse, Phys. Lett., 1968, 28, 444–445 CrossRef CAS .
  27. D. L. Dexter, J. Chem. Phys., 1953, 21, 836 CrossRef CAS .
  28. L. G. Van Uitert, J. Electrochem. Soc., 1967, 114, 1048–1053 CrossRef CAS .
  29. L. G. Van Uitert and L. F. Johnson, J. Chem. Phys., 1966, 44, 3514–3522 CrossRef CAS .
  30. X. Wu, X. Ji, Z. Wang, Z. Qiu, L. Yu, W. Zhou, J. Zhang and S. Lian, J. Alloys Compd., 2021, 855, 157520 CrossRef CAS .
  31. H. Lin, D. Q. Chen, Y. L. Yu, R. Zhang and Y. S. Wang, Appl. Phys. Lett., 2013, 103, 091902 CrossRef .
  32. X. Zhang, L. Huang, F. Pan, M. Wu, J. Wang, Y. Chen and Q. Su, ACS Appl. Mater. Interfaces, 2014, 6, 2709–2717 CrossRef CAS PubMed .
  33. K. J. Laidler, J. Chem. Educ., 1984, 61, 494 CrossRef CAS .
  34. W. G. Xiao, X. F. Liu, J. H. Zhang and J. R. Qiu, Adv. Opt. Mater., 2019, 7, 1801677 CrossRef .
  35. D. R. Tallant, M. P. Miller and J. C. Wright, J. Chem. Phys., 1976, 65, 510–521 CrossRef CAS .
  36. M. P. Miller and J. C. Wright, J. Chem. Phys., 1979, 71, 324–338 CrossRef CAS .
  37. P. I. Paulose, G. Jose, V. Thomas, N. V. Unnikrishnan and M. K. R. Warrier, J. Phys. Chem. Solids, 2003, 64, 841–846 CrossRef CAS .
  38. J. H. Zhang, Z. D. Hao, J. Li, X. Zhang, Y. S. Luo and G. H. Pan, Light: Sci. Appl., 2015, 4, e239 CrossRef CAS .
  39. Y. Xiao and Z. Hao, J. Mater. Chem. C, 2020, 8, 1151–1152 RSC .

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4dt01843e

This journal is © The Royal Society of Chemistry 2024