DOI:
10.1039/D3QI00685A
(Research Article)
Inorg. Chem. Front., 2023,
10, 4230-4240
A multi-centre activated single-phase white light phosphor with high efficiency for near-UV based WLEDs†
Received
13th April 2023
, Accepted 15th June 2023
First published on 21st June 2023
Abstract
Single-phase white light phosphors for phosphor-converted white light-emitting diodes (pc-WLEDs) have become a hot spot and attracted considerable attention. However, the unsatisfactory luminous efficiency and colour rendering index of this kind of device hinder their large-scale applications in the field of lighting. Herein, a novel multi-centre activated single-phase white light phosphor, Gd2Sr3B4O12:Ce3+,Tb3+,Sm3+, has been designed and prepared. The crystal structures of the host and the doped samples are characterized with XRD and Rietveld refinements. The band gap of the host calculated with the density functional theory is consistent with the value obtained from the UV-Vis reflection spectrum. Benefiting from the energy transfer paths of Ce3+–Tb3+, Ce3+–Sm3+ and Tb3+–Sm3+, which are quantitatively evaluated with photoluminescence spectra and decay curves, the intensities of the characteristic emission peaks of the three dopants are found to be closely interlocked and the emission colour of the phosphors can be precisely tuned by adjusting the relative ratio of the three ions. Under the excitation of near-UV light, the two phosphors of Gd2Sr3B4O12:0.03Ce3+,0.18Tb3+,0.05Sm3+ and Gd2Sr3B4O12:0.03Ce3+,0.22Tb3+,0.07Sm3+ emit white light with CIE coordinates of (0.32, 0.35) and (0.34, 0.34), and the quantum efficiencies are 46.8% and 34.9%, respectively. The temperature-dependent photoluminescence spectra of the phosphors show that the quenching temperature (T50%) is as high as 450 K. Finally, WLED devices with maximal colour rendering index of 87.5 are obtained by fabricating the two single-phase white light phosphors with 365 nm LED chips. The balanced and preferable comprehensive performances of Gd2Sr3B4O12:Ce3+,Tb3+,Sm3+ demonstrate that the as-prepared single-phase white light phosphor can be a promising candidate for fabricating near-UV chip-based WLEDs.
1. Introduction
Recently, solid-state lighting has become rapidly growing research field owing to extensive worry about environmental and energy issues.1,2 Phosphor-converted white light-emitting diodes (pc-WLEDs) are considered to be a promising solid-state lighting source for their emerging advantages of high efficiency, energy saving, environmental friendliness, etc.3–5 Two approaches are employed to develop commercialized WLEDs. The first method is fabricating blue InGaN chips with yellow phosphors, typically Y3Al5O12:Ce3+. However, the colour rendering index (Ra) of this kind of WLEDs is strongly restricted by the absence of cyan and red regions in emission spectra.6–8 The second strategy is coating the mixture of multi-colour phosphors on near-UV LED chips.9 The Ra of this kind of WLEDs is much higher than that of the former, but the reduction of luminous efficiency caused by the strong reabsorption in blue light region limits its application.10,11 Therefore, producing white light in a single matrix has gained much attention.11
Generally, there are three routes to generate white light in single-phase phosphors activated with rare-earth ions.10,12 The first is doping single rare-earth ions into certain hosts, such as narrow-band emitting BiF3:Tb3+, CaIn2O4:Eu3+ and Ca9NaZn(PO4)7:Dy3+, or broadband emitting Ba3Lu4O9:Ce3+ and (Rb,K)2CaPO4F:Eu2+.13–17 Nevertheless, the f–f forbidden transitions of Tb3+, Eu3+ and Dy3+ make the quantum efficiency of the former too low to meet the requirements of WLED applications, for example, the internal quantum efficiencies of CaIn2O4:Eu3+, Ca9NaZn(PO4)7:Dy3+ and BiF3:Tb3+ are 10%, 19.2% and 3.44%, respectively. Although the quantum efficiency of the phosphors adopting f–d allowed transitions is significantly increased, the application of such phosphors in high-quality WLED devices is still restricted by the imbalance of emission spectra.18 White light can also be generated by the co-excitation of two kinds of rare earth ions, as in Sr2LaGaO5:Dy3+,Sm3+ and LaOCl:Tb3+,Eu3+.19,20 However, since such single-phase white light phosphors have strict requirements on the co-excitation wavelength, the quality of white light will be significantly reduced when the excitation wavelength is changed.10,18 The third route is co-doping two or three kinds of ions with energy transfer processes into certain hosts. The processes make the luminescence of the activators be controlled by the sensitizers, but the slight variation of excitation wavelength does not significantly change the quality of white light. So, this route for generating white light has been widely concerned.10,18
Nowadays, many single-phase white light phosphors via energy transfer have been reported. Zheng et al. prepared the white light-emitting phosphor, Ba9La2Si6O24:Bi3+,Eu3+, with satisfying thermal stability (T50% = 498 K), and fabricated a high colour rendering index (Ra = 87) WLED device with the phosphor and 365 nm chip.21 SrBPO5:Ce3+,Dy3+ phosphor prepared by Vinodkumar et al. exhibits relatively high thermal stability and colour stability.43 Ca3YAl3B4O15:Ce3+,Tb3+,Sm3+ prepared by Khan et al. shows both high Ra and thermal stability.18 Though single-phase white light phosphors via energy transfer have made great progress in improving luminescent properties, few of them can achieve satisfying balanced overall performances, including but not limited to quantum efficiency, Ra and thermal stability. Therefore, it still needs to be addressed to meet the demands of high-quality WLEDs.
Ce3+ ion usually emit broad blue light originating from the 4f–5d transition, which makes it a suitable sensitizer for single-phase white light phosphors. Tb3+ and Sm3+ ions are essential green and red activators in traditional phosphors. However, because of the f–f parity-forbidden transitions, the luminescence efficiencies of Tb3+ and Sm3+ are relatively low, which strongly restricts their applications. Taking advantages of the broadband emission of Ce3+ and the energy transfer from Ce3+ to Tb3+ or Sm3+, the emission efficiencies of Tb3+ or Sm3+ may be greatly improved.18,22 The choice of host materials is also important for researchers to obtain phosphors with satisfying luminescent properties. Borates are considered to be favourite hosts for their excellent stability, low synthetic temperature, high ultraviolet transparency and wide bandgap.22,23 As a new type of rare-earth and alkali-halide-based double borates, RE2M3B4O12 have been investigated for their excellent optical properties.23–25 According to the structure models of Y2Sr3B4O12 and La2Sr3B4O12, the structure of Gd2Sr3B4O12 belongs to Pnma. There are two distinct cationic sites in Gd2Sr3B4O12, among which the [BO3] planar triangles fill in and form a network structure.23
Herein, the novel phosphor, Gd2Sr3B4O12:Ce3+,Tb3+,Sm3+, were designed and prepared to obtain white light emission by controlling the energy transfer processes of Ce3+–Tb3+, Ce3+–Sm3+ and Tb3+–Sm3+. The photoluminescent properties of singly-, doubly- and triply-doped phosphors were studied in detail to evaluate the three paths of energy transfer. By finely adjusting the relative ratio of the three ions to control the energy transfer processes, white light with CIE chromaticity coordinates of (0.32, 0.35) and (0.34, 0.34) can be obtained in Gd2Sr3B4O12:0.03Ce3+,0.18Tb3+,0.05Sm3+ and Gd2Sr3B4O12:0.03Ce3+,0.22Tb3+,0.07Sm3+, respectively. The quantum efficiency and the quenching temperature (T50%) are as high as 46.8% and 450 K. A WLED device with maximal Ra of 87.5 is obtained by fabricating the as-prepared single-phase white light phosphor with 365 nm LED chip. Based on the balanced and preferable comprehensive performances, Gd2Sr3B4O12:Ce3+,Tb3+,Sm3+ can be a candidate for single-phase white light phosphors being potential application in WLEDs.
2. Experimental section
2.1. Sample preparation
All samples were prepared by conventional high temperature solid-state reaction. Raw materials of Sr2CO3 (A.R.), Gd2O3 (99.99%), CeO2 (99.99%), Sm2O3 (99.9%), and Tb4O7 (99.99%) were purchased from Aladdin, and H3BO3 (A.R.) was purchased from Guangzhou Chemical Reagent Factory. All the raw materials were stoichiometrically weighed, and 5% excess H3BO3 were mixed with other reactants. After properly ground, the mixture was placed into an alumina crucible and preheated at 1073 K for 4 h. After cooling to room temperature, the mixture was ground again and heated up to 1373 K for 4 h under weak reducing atmosphere. Finally, the as-synthesized Gd2Sr3B4O12 (abbreviated to GSBO) and Gd2(1−x−y−z)Sr3B4O12:xCe3+,yTb3+,zSm3+ samples were crushed into powders and collected for further characterization. Two WLED devices were fabricated by coating GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ and GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ phosphors with 365 nm LED chips, respectively.
2.2. Measurements and characterization
The powder X-ray diffraction measurements were examined on a Rigaku D-Max 2200 X-ray diffraction system with Cu Kα radiation working at 40 kV and 26 mA. The refinement of XRD data was conducted by using JANA 2006 software. The microstructure and the element mapping were investigated by an FEI Quanta scanning with an energy-dispersive spectrometer. The photoluminescence emission (PL), excitation (PLE) spectra were measured with an FLS-980-combined time resolved and steady state fluorescence spectrometer (Edinburgh) with Xe/μs lamps. And fluorescence decay curves were obtained with the same fluorescence spectrometer equipped with a μs lamp (μF2, Edinburgh) and an EPLED-340 pulsed laser (central wavelength 345.4 nm, bandwidth 13.4 nm, pulse width 806.9 ps, Edinburgh). Temperature-dependent PL spectra with the temperature range from 300 K to 500 K were collected on the same instrument with an Oxford temperature controller. And the quantum efficiencies were obtained by a Hamamatsu C9920-03G absolute photoluminescence quantum efficiencies measurement system. The WLED devices were driven by LED300E programmable test power for LEDs under different forward bias current, and the electroluminescence (EL) spectra of WLED devices were measured with a HAAS-2000 high accuracy array spectroradiometer equipped with an integrating sphere (Hangzhou Everfine Photo-E-Info Co., Ltd).
2.3. Computational methods
The band structure and the density of state (DOS) of GSBO were carried out by first-principle calculations based on the density functional theory (DFT) with the Vienna Ab initio Software Package (VASP).26 The interaction between ions and electrons was defined by the projector augments wave (PAW) pseudopotentials.27 The Perdew–Burke–Ernzerhof (PBE) functional was used to perform the geometry optimizations.28 The plant wave cutoff energy was set to 520 eV. The k-mesh of 3 × 2 × 3 Monkhorst–Pack (M–P) grid was used for geometry optimizations and 4 × 3 × 4 M–P grid was used during self-consistent relaxation. The convergence criterion for iteration was set to 10−5 eV per atom and the maximal force was set to 0.001 eV Å−1.
3. Results and discussion
3.1. Crystal structure of GSBO and GSBO:Ce3+,Tb3+,Sm3+
The crystal structure of Gd2Sr3B4O12 (GSBO) is isostructural with Y2Sr3B4O12, which belongs to orthorhombic structure with space group of Pnma (No. 62).23 The powder XRD patterns of the host and the co-doped samples depicted in Fig. 1(a) show that all the patterns can match well with the simulated one. According to the Bragg equation, the patterns will shift to smaller angles, when the smaller ions are replaced by the larger ions. As shown in the enlarged XRD patterns between 30° and 32°, the patterns of GSBO:0.03Ce3+,0.05Sm3+, GSBO:0.03Ce3+,0.10Tb3+ and GSBO:0.10Tb3+,0.05Sm3+ are shift to smaller angles. However, the pattern of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ is almost in the same position with the host, which should be caused by the substitution of Gd3+ by a large amount of Tb3+ in this triply-doped sample. In order to further analyse the phase purity, the Rietveld refinement was performed by using Y2Sr3B4O12 (ICSD#421574) as the starting model, and the fitting results are displayed in Tables S1 and S2, Fig. S1 (ESI†) and Fig. 1(b).29 The reliability factors (Rwp and Rp) of both GSBO and GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ are less than 5%, suggesting that there is no distinct impurity phase and the doping of RE3+ will not cause any significant phase change. Cell parameters of the doped sample are a = 7.4244 Å, b = 16.0184 Å, c = 8.7620 Å, and V = 1042.04 Å3, respectively. According to the fitting results, the unit cell volume of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ is smaller than that of GSBO matrix, which confirms the result of XRD of each samples. The crystal structure of GSBO along a-axis and the coordination environments of cations in the host are exhibited in Fig. 1(c) and (d). There are three different cation sites in the unit cell, named Sr1, Sr2 and Gd. Sr1 and Sr2 are eight-fold coordinated with O atoms and occupy 8d and 4c Wyckoff site, respectively. And Gd also locates at 8d Wyckoff site with 7-fold coordination. The ionic radii of Gd3+, Ce3+, Sm3+ and Tb3+ with coordination number (CN) of 7 are 1.00 Å, 1.07 Å, 1.02 Å and 0.98 Å, respectively.
 |
| Fig. 1 (a) The XRD patterns of GSBO, GSBO:0.03Ce3+,0.05Sm3+, GSBO:0.03Ce3+,0.10Tb3+, GSBO:0.10Tb3+,0.05Sm3+, and GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, (b) the Rietveld refinement profile of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, (c) the crystal structure of GSBO, and (d) the coordination environments of cations. | |
The similarity of ionic radii and valence states of RE3+ ions show that the doping ions are more likely to replace Gd3+ in the matrix. The possibility of cation substitution can be estimated by the empirical equation of Hume–Rothery, and the difference (Dr) between dopants and possible replaced cations should be less than 30%. The values of Dr can be estimated by the following equation:30
|  | (1) |
in which,
rD(CN) is the ionic radius of dopants and
rR(CN) refers to the ionic radius of replaced ions. The values of
Dr between the dopants (Ce
3+, Sm
3+ and Tb
3+) and the possible replaced ions (Gd
3+ or Sr
2+) are shown in
Table 1. As shown in the table, the radius difference percentages between the dopants and Gd
3+ are significantly smaller than those between the dopants and Sr
2+, which further indicates that Ce
3+, Tb
3+, Sm
3+ will mainly replace Gd
3+ rather than Sr
2+ in GSBO.
Table 1 The radius difference percentage between replaced ions and dopants
Elements |
Ionic radii |
CN |
D
r
|
Ce3+ |
Sm3+ |
Tb3+ |
Sr2+ |
1.26 Å |
8 |
10.24% |
14.37% |
17.46% |
Gd3+ |
1.00 Å |
7 |
7.00% |
2.00% |
2.00% |
Fig. 2 shows the scan electron microscope (SEM) image and the element mappings of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ sample. The smooth surface of the as-prepared sample with an average particle size of about 60 μm suggests good crystallization. The element mappings show the uniform distribution of all the metal elements, which confirms that all rare earth ions successfully enter the host lattice. The atomic percent of the metal elements in the as-prepared sample, the calculated ratio and the stoichiometric ratio of RE3+ in GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ are listed in Table S3.† And the mappings indicate that all the metal elements are homogeneously distributed. Besides, the ratio of RE3+ calculated with EDS results is consistent with the stoichiometric ratio of RE3+ in the sample, also proofing the homogeneous distribution of rare earth ions.
 |
| Fig. 2 SEM image of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ and the element mappings of all the metal elements. | |
Fig. 3 shows the UV-Vis reflection spectra of the GSBO host, GSBO:0.05Ce3+, and GSBO:0.03Ce3+,0.14Tb3+,0.07Sm3+. There is a sharp peak at 275 nm in the reflection spectrum of the host, which should come from the absorption of Gd3+ in the host lattice. An obvious broad absorption ranging from 225 nm to 400 nm and some sharp peaks at 402 nm, 416 nm and 474 nm can be observed in the reflection spectra of Ce3+-doped and triply-doped samples, which are attributed to the f–d transition of Ce3+ and the f–f transitions of Sm3+, respectively. Because of the overlapping of the absorption of Ce3+ (ranging from 300 nm to 400 nm) and Tb3+ (ranging from 310 nm to 390 nm), the f–f transitions of Tb3+ are not obvious in the reflection spectrum of GSBO:0.03Ce3+,0.14Tb3+,0.07Sm3+. The optical band gap (Eg) of the host can be calculated by using the Tauc and Kubelka−Munk equations:31,32
|  | (2) |
|  | (3) |
where
α is the absorption coefficient,
R stands for the diffuse reflectance of the sample,
hν represents the photon energy, and
A is the proportionality constant. The exponent
n refers to the nature of optical band gap,
e.g.,
n = 1/2 represents indirect bandgap semiconductor. As shown in the inset of
Fig. 3a,
Eg of GSBO is estimated to be 5.78 eV by extrapolating the tail of the curve. From the calculated band structure of GSBO as displayed in
Fig. 3b, the band gap value is estimated to be 4.59 eV. Since the band gap calculation adopting generalized gradient approximation (GGA) function always underestimates bandgap, the calculated value will be smaller than optical band gap.
33,34 The total and projected electron density of states (TDOS/PDOS) of GSBO are presented in Fig. S2.
† According to the calculation, the conduction band of GSBO is mainly contributed to Gd-s and Sr-d orbitals and the valence band is composed of O-p orbitals.
 |
| Fig. 3 (a) The UV-Vis reflection spectra of GSBO, GSBO:0.05Ce3+ and GSBO:0.03Ce3+,0.14Tb3+,0.07Sm3+ (inset with the optical band gap of GSBO) and (b) the band structure of the host. | |
3.2. Photoluminescent properties of singly-doped GSBO phosphors
Fig. 4 depicts the photoluminescence emission (PL) and excitation (PLE) spectra of GSBO:Ce3+ samples. As presented in Fig. 4a, GSBO:0.03Ce3+ shows a broad excitation band ranging from 300 nm to 400 nm, which is derived from the transition of 4f ground state to 5d excited state of Ce3+. Under 350 nm excitation, the sample shows a characteristic blue emission ranging from 360 nm to 600 nm with a peak at 425 nm and the full width at half maximum (FWHM) of 103 nm. The Gaussian fitting result is shown in Fig. S3,† the PL spectrum of GSBO:0.03Ce3+ can be well fitted with two sub-bands, which belongs to the emission from 5d excited state to the ground states of Ce3+ (2F5/2 and 2F7/2). The concentration-dependent PL spectra in Fig. 4b show that the emission intensity reaches the maximal value when x = 0.03 and decreases with increasing the doping concentration of Ce3+ because of the concentration quenching of luminescence. The decay curves of Ce3+-doped samples in Fig. 4c can be fitted by the following equation:19,34 |  | (4) |
where It is the luminescent intensity at time t, and τ is the lifetime. Since the change of the doping concentration is small, the lifetimes show little variation from 30.94 ns to 29.01 ns (details in Table S5, ESI†).
 |
| Fig. 4 (a) The PL and PLE spectra of GSBO:0.03Ce3+, (b) the concentration-dependent PL spectra of GSBO:xCe3+ (inset with the dependence of the emission intensity with the doping concentration of Ce3+), (c) the decay curves of GSBO:xCe3+ (x = 0.01–0.05), and (d) the I–H fitting of GSBO:0.05Ce3+. | |
In order to further illustrate the concentration quenching mechanism, the Inokuti–Hirayama (I–H) model with the following equation was employed:34,35
|  | (5) |
where
I0 represents the luminescent intensity at
t = 0, and
τ0 is the lifetime of donors without acceptors. Since the energy migration should be a unique energy transfer where donors and acceptors are the same ions, we chose the lifetime of the least quenched samples (GSBO:0.01Ce
3+) as
τ0 to fit the decay curve of GSBO:0.05Ce
3+.
Q stands for the macroscopic parameter, and
S = 6, 8, 10 represents the dominant of dipole–dipole (D–D), dipole–quadrupole (D–Q) and quadrupole–quadrupole (Q–Q) interactions, respectively. Here, the fitting results of each value of
S are presented in
Fig. 4d. Because of the value of
R2 for
S = 6, 8, 10 is too close to each other, I–H fitting for quenched samples were conducted (Fig. S4
†). According to the fitting results, D–D interaction should be the dominant energy migration mechanism in Ce
3+ singly-doped samples.
The PLE, PL and concentration-dependent PL spectra of GSBO:Tb3+ are displayed in Fig. 5a and b. When monitoring at 542 nm, several sharp peaks located within 300 nm to 380 nm can be observed, which are derived from the f–f transitions of Tb3+. Under 377 nm excitation, four sharp peaks located at 488 nm, 542 nm, 583 nm, and 623 nm are ascribed to 5D4 to 7FJ (J = 6, 5, 4, 3) transitions, respectively. The luminescence of GSBO:Tb3+ shows concentration quenching when the doping concentration of Tb3+ is about 0.50. Fig. 5c and d exhibit the luminescent spectra of Sm3+-doped samples. Similar to Tb3+ singly-doped samples, the PLE spectrum of Sm3+ doped sample contains a series of narrow peaks ranging from 270 nm to 500 nm. Upon 402 nm excitation, Sm3+-doped sample shows orange-red emission which is attributed to the transitions of 4G5/2 to 6HJ (J = 5/2, 7/2, 9/2, 11/2). The lifetimes of Tb3+ and Sm3+ singly-doped samples are listed in Tables S6 and S7 (ESI†), both of which show a monotonic decreasing trend with the increasing of the concentration of Tb3+ and Sm3+. And by fitting the decay curve of the most quenched Tb3+ and Sm3+ singly-doped samples, their dominant energy migration mechanism should be D–D interaction (Fig. S5†).
 |
| Fig. 5 (a) The PL and PLE spectra of GSBO:0.50Tb3+, (b) the concentration-dependent PL spectra of GSBO:yTb3+ (inset is about the dependence of the emission intensity with the doping concentration of Tb3+), (c) the PL and PLE spectra of GSBO:0.03Sm3+, and (d) the concentration-dependent PL spectra of GSBO:zSm3+ (inset is about the dependence of the emission intensity with the doping concentration of Sm3+). | |
3.3. Energy transfer between dopants in doubly-doped GSBO phosphors
Because the emission intensities of each doping ion in single-phase white light phosphors are usually interrelated by energy transfer, a minor variation of doping concentration of RE3+ will cause a significant change of the colour of emission. It is important to identify the energy transfer between doped ions for achieving white light in a single-phase phosphor. Hence, the doubly-doped phosphors of GSBO:Ce3+,Tb3+, GSBO:Ce3+,Sm3+ and GSBO:Tb3+,Sm3+ were prepared and their PL spectra were measured and are presented in Fig. 6. In order to fully exclude the influence of co-activation of dopants, the excitation wavelengths of each doubly-doped phosphor should be carefully chosen (Fig. S6†). As depicted in Fig. 6a, when excited with 330 nm, GSBO:0.03Ce3+,yTb3+ phosphors show the characteristic emissions both from Ce3+ (broadband ranged at 400 nm–525 nm) and from Tb3+ (sharp peaks within 450 nm–650 nm). And with increasing y value, the emission from Ce3+ decreases constantly, while the emission from Tb3+ shows an increasing trend, which indicates the energy transfer from Ce3+ to Tb3+ takes place in GSBO:Ce3+,Tb3+. The decrease in the blue/green (B/G) ratio with increasing the doping concentration of Tb3+ as shown in Fig. 6b suggests the emission colour of GSBO:0.03Ce3+,yTb3+ phosphors can be tuned from blue to green. The decay curves of the phosphors are presented in the inset of Fig. 6c. The calculated results of with eqn (4) show that τ of Ce3+ at 425 nm decreases constantly with increasing the doping concentration of Tb3+ (Table S8†). The critical distance (RC) between each dopants are calculated in ESI and the results are listed in Table S4.† Because the RC values for both Ce3+–Tb3+, Ce3+–Sm3+ and Tb3+–Sm3+ are larger than 5 Å, it indicates that the dominant interaction mechanism should be electric multipolar interaction. By fitting the decay curve of GSBO:0.03Ce3+,0.20Tb3+ with eqn (5), the energy transfer mechanism can be easily determined as D–D interaction.
 |
| Fig. 6 (a) The PL spectra of GSBO:0.03Ce3+,yTb3+, (b) the B/G ratio of GSBO:0.03Ce3+,yTb3+, (c) the I–H fitting of GSBO:0.03Ce3+,0.20Tb3+, (d) the PL spectra of GSBO:0.03Ce3+,zSm3+, (e) the B/R ratio of GSBO:0.03Ce3+,zSm3+, (f) the I–H fitting of GSBO:0.03Ce3+,0.09Sm3+, (g) the PL spectra of GSBO:0.50Tb3+,zSm3+, (h) the G/R ratio of GSBO:0.50Tb3+,zSm3+, and (i) the I–H fitting of GSBO:0.50Tb3+,0.09Sm3+. | |
The changing trends observed in PL spectra (Fig. 6d and g), the blue/red (B/R) ratio (Fig. 6e), the green/red (G/R) ratio (Fig. 6h), τ of Ce3+ at 425 nm (Fig. 6f and Table S9†), and τ of Tb3+ at 542 nm (Fig. 6i and Table S10†) for GSBO:0.03Ce3+,zSm3+ and GSBO:0.50Tb3+,zSm3+ are similar to those for GSBO:0.03Ce3+,yTb3+. To exhibit the colour variation directly, the CIE chromaticity diagram for doubly-doped samples is presented in Fig. S7.† With increasing the concentration of acceptors, the CIE coordinates of the doubly-doped samples gradually deviate from those of the singly-doped samples. The mechanisms of energy transfer from Ce3+ to Sm3+ and from Tb3+ to Sm3+ are also dominated by D–D interaction.
3.4. Photoluminescent and thermal quenching properties of GSBO:Ce3+,Tb3+,Sm3+ phosphors
Based on understanding the mechanisms of energy transfer in singly- and doubly-doped phosphors, it is easy to speculate the energy transfer processes in GSBO:Ce3+,Tb3+,Sm3+ phosphors (as shown in Fig. 7a). Due to the existence of energy transfer pathways between any two RE3+ ions, it could be hard to infer the colour variation when the doping concentrations of Tb3+ and Sm3+ change at the same time. In order to obtain white light emission in the GSBO matrix, Tb3+ ions could serve as the energy transfer channel to tune the colour. Hence, triply-doped samples were synthesized by fixing the doping concentration of Ce3+ and Sm3+. The PL spectra of GSBO:0.03Ce3+,yTb3+,0.05Sm3+ and GSBO:0.03Ce3+,yTb3+,0.07Sm3+ are depicted in Fig. 7b and c. Under 350 nm excitation, these triply-doped samples exhibit the characteristic emission of Ce3+, Tb3+ and Sm3+. Fig. 7d illustrates the variation of their emission intensities. The emission from Tb3+ shows an upward trend and reaches the maximum at y = 0.14, while the emission from Ce3+ decreases constantly and the emission from Sm3+ increases monotonically. The decay curves of triply-doped phosphors at 425 nm were measured and are shown in Fig. S8 (ESI†). The lifetime of Ce3+ is largely reduced by the co-doping of Tb3+ and Sm3+, dropping from 30.41 ns to 22.51 ns or 20.81 ns (Tables S11 and S12, ESI†).
 |
| Fig. 7 (a) The schematic energy-level diagram of Ce3+, Tb3+ and Sm3+ and the energy transfer processes in GSBO:Ce3+,Tb3+,Sm3+, (b) and (c) the concentration-dependent PL spectra of GSBO:0.03Ce3+,yTb3+,zSm3+ (y = 0.02–0.22 and z = 0.05, 0.07), and (d) the emission peaks variation of GSBO:0.03Ce3+,yTb3+,zSm3+ (y = 0.02–0.22 and z = 0.05, 0.07) in intensities. | |
As shown in Fig. 8a and c, the emission intensity from each dopant is closely interlocked in the triply-coped phosphors. Hence, by adjusting the doping concentrations of Tb3+ and Sm3+, the CIE coordinates can be precisely tuned. In order to directly present the emission colour of as-synthesized phosphors under near-UV excitation, the correlated CIE coordinates of GSBO:0.03Ce3+,yTb3+,0.05Sm3+ and GSBO:0.03Ce3+,yTb3+,0.07Sm3+ phosphors are exhibited in Fig. 8b and d. White light can be achieved in GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ and GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ samples with CIE coordinates of (0.32, 0.35) and (0.34, 0.34), respectively, which is quite close to standard white light (0.33, 0.33). Apart from the PL properties mentioned above, the internal quantum efficiency (IQE) is also a crucial factor to evaluate the performances of phosphors. The IQE is defined by the equation below:36,37
|  | (6) |
where
Ls is the emission spectra of phosphors,
Er and
Es are the reflections of excitation with and without samples. As shown in Fig. S9,
† the measured
ηint values of GSBO:0.03Ce
3+,0.18Tb
3+,0.05Sm
3+ and GSBO:0.03Ce
3+,0.22Tb
3+,0.07Sm
3+ are 46.8% and 34.9%, respectively.
 |
| Fig. 8 (a) The R/G/B ratio variation of GSBO:0.03Ce3+,yTb3+,0.05Sm3+ (y = 0.02–0.22), (b) the CIE chromaticity diagram of the corresponding samples, (c) the R/G/B ratio variation of GSBO:0.03Ce3+,yTb3+,0.07Sm3+ (y = 0.02–0.22) and (d) the CIE chromaticity diagram of the corresponding samples. | |
To further analyse the thermal quenching behaviours of the as-synthesized phosphors, the temperature-dependent PL spectra with the temperature range of 300 K–500 K were measured. As shown in Fig. 9a, the emission intensity of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ shows a downward tendency owing to the thermal quenching. The normalized integrated intensity of the temperature-dependent PL spectra is depicted in Fig. 9b. The integrated intensity at 423 K can be maintained at more than 60% of that at room temperature and the quenching temperature (T50%) is as high as 450 K. The emission intensity variations of Ce3+ (at 425 nm), Tb3+ (542 nm) and Sm3+ (at 599 nm) are presented in Fig. 9c. The emissions from Ce3+, Tb3+ and Sm3+ constantly decrease with increasing temperature. Since the f–f transition remains insensitive to the variation of temperature compared to f–d transition, the thermal quenching rate of Ce3+ is faster than Tb3+ and Sm3+. Fig. 9d exhibits the thermal activation energy calculation of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ sample with the Arrhenius equation:38,39
|  | (7) |
 |
| Fig. 9 (a) The temperature-dependent PL spectra of GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, (b) the trend of integrated intensity of the whole spectra, (c) the trend of Ce3+ (at 425 nm), Tb3+ (542 nm) and Sm3+ (at 599 nm) and (d) the Arrhenius fitting of the integral intensities of the temperature-dependent PL spectra. | |
By fitting with
vs.
, the thermal activation energy can be obtained to be 0.2226 eV.
3.5. The electroluminescence properties of WLED devices
Finally, WLED devices were fabricated by coating the phosphors, GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ and GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, with 365 nm LED chips. Fig. 10 and Fig. S8† display the EL spectra of the as-fabricated devices with an inset of the images of corresponding packed WLED devices. As shown in Fig. 10a and Fig. S10a,† both GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ and GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ can be effectively excited by the n-UV LED chip and emit bright white light, the Ra of the two WLED devices is 87.5 and 77.3, respectively. Fig. 10b and Fig. S10b† display the EL spectra of the as-fabricated WLED devices under various currents. It can be seen that the shape and peak position of the EL spectra have little variation. However, the EL intensity slightly decreases after the current exceeds 150 mA, which should be caused by the temperature rise under high current.11 Because the application prospect of phosphors cannot be evaluated with any single property, IQE, Ra and T50% were chosen to compare our work with some other recent studies. The comprehensive comparison of the as-synthesized phosphors with other single-phase white light phosphors is listed in Table 2. It can be seen that the performances of as-prepared phosphors are balanced and the comprehensive performance is preferable. Hence, the comparison result shows that the as-synthesized phosphors could have a potential application in WLEDs.
 |
| Fig. 10 (a) The EL spectrum of the WLED fabricated by GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ with 365 nm LED chip and (b) the EL spectra of the as-fabricated WLED under various forward bias currents. | |
Table 2 The comparison of IQE, maximal Ra and T50% of some single-phase white light phosphors with GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ and GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+
Phosphors |
λ
ex (nm) |
IQE (%) |
Max. Ra |
T
50% (K) |
Ref. |
Gd2Sr3B4O12:0.03Ce3+,0.18Tb3+,0.05Sm3+ |
350 |
46.8 |
77.3 |
450 |
This work |
Gd2Sr3B4O12:0.03Ce3+,0.22Tb3+,0.07Sm3+ |
350 |
34.9 |
87.5 |
— |
This work |
Rb2CaP2O7:0.005Ce3+,0.01Eu2+ |
310 |
74.36 |
88.4 |
∼430 |
40
|
NaSrBO3:0.01Ce3+,0.08Tb3+,0.06Sm3+ |
360 |
48.2 |
80.1 |
480 |
22
|
Sr2LaGaO5:0.02Bi3+,0.05Eu3+ |
370 |
∼47.3 |
78.5 |
350 |
41
|
Sr3Ce0.55La0.45(PO4)3:0.05Eu3+ |
365 |
37 |
90 |
402 |
42
|
Ca3Y0.65Al3B4O15:0.03Ce3+,0.20Tb3+,0.12Sm3+ |
345 |
34.84 |
84.4 |
>500 |
18
|
SrBPO5:0.01Ce3+,0.012Dy3+ |
296 |
28.26 |
— |
>423 |
43
|
SrBaLaGaO5:0.02Bi3+,0.05Eu3+ |
370 |
∼24.9 |
83.1 |
∼375 |
41
|
GdBO3:0.02Ce3+,0.12Tb3+,0.015Eu3+ |
360 |
20.2 |
— |
— |
10
|
Ba2LaGaO5:0.02Bi3+,0.05Eu3+ |
370 |
∼14.7 |
92.2 |
∼325 |
41
|
Ca0.99In2O4:0.01Eu3+ |
250 |
10 |
— |
— |
14
|
Mg2Gd8(SiO4)6O2:0.03Ce3+,0.01Mn2+ |
365 |
7.93 |
87.4 |
373 |
44
|
Li3Gd2.34Te2O12:0.15Bi3+,0.07Pr3+ |
297 |
5 |
— |
<300 |
45
|
NaSr0.9425VO4:0.05Dy3+,0.0075La3+ |
347 |
3.26 |
— |
— |
46
|
Bi4BPO10:0.07Dy3+ |
395 |
— |
40 |
— |
47
|
Sr2LaGaO5:0.03Dy3+,0.07Sm3+ |
320 |
— |
91.4 |
400 |
19
|
4. Conclusions
In conclusion, Ce3+-, Tb3+- and Sm3+-doped Gd2Sr3B4O12 phosphors have been successfully synthesized by the conventional high temperature solid-state reaction. The XRD and Rietveld refinements of GSBO host and the doped samples have shown that the substitution of Ce3+, Tb3+ and Sm3+ ions for Gd3+ in the host has no any significant effect on the phase purity. The band structure of the host has been calculated with the first-principle calculations based on the density functional theory, and the band gap is calculated to be 4.59 eV, which is close to the value obtained from the UV-Vis reflection spectrum. The Inokuti–Hirayama model has been used to fit the photoluminescence spectra of the three singly-doped phosphors and the fitted results have indicated that the dipole–dipole interaction is the dominant mechanism of energy transfer in the three phosphors.
The photoluminescence spectra and the decay curves of GSBO:Ce3+,Tb3+, GSBO:Ce3+,Sm3+ and GSBO:Tb3+,Sm3+ phosphors have shown that there are energy transfer processes of Ce3+–Tb3+, Ce3+–Sm3+ and Tb3+–Sm3+, and the mechanisms of all the energy transfer processes are dominated by dipole–dipole interaction. The intensities of the characteristic emission peaks of the three dopants in GSBO:Ce3+,Tb3+,Sm3+ have been found to be closely interlocked and the emission colour of the phosphors can be precisely tuned by adjusting the relative ratio of the three ions. Bright white light with CIE coordinates of (0.32, 0.35) and (0.34, 0.34) have been obtained in GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+ and GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+, respectively. Under 350 nm excitation, the internal quantum efficiency up to 46.8% has been achieved in GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, which is relatively high in RE3+ doped single-phase white light phosphors. The temperature-dependent photoluminescence spectra of the phosphors have shown that the quenching temperature (T50%) is as high as 450 K. Two WLED devices with high Ra, 87.5 for GSBO:0.03Ce3+,0.22Tb3+,0.07Sm3+ and 77.3 for GSBO:0.03Ce3+,0.18Tb3+,0.05Sm3+, have been fabricated by coating the prepared single-phase white light phosphors with near-UV chips. The balanced and preferable comprehensive luminescent properties manifest that Gd2Sr3B4O12:Ce3+,Tb3+,Sm3+ is an outstanding single-phase white light phosphor activated with multi-centre via energy transfer, which has a potential application in near-UV pumped WLEDs for lighting.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
This work was financially supported by grants from the National Natural Science Foundation of China (NSFC No. 51972347 and 21771195) and the joint projects of NSFC with Guangdong Province (No. U22A20135 and U1301242).
References
- L. Wang, R.-J. Xie, T. Suehiro, T. Takeda and N. Hirosaki, Down-conversion nitride materials for solid state lighting: recent advances and perspectives, Chem. Rev., 2018, 118, 1951–2009 CrossRef CAS PubMed.
- M.-H. Fang, Z. Bao, W.-T. Huang and R.-S. Liu, Evolutionary generation of phosphor materials and their progress in future applications for light-emitting diodes, Chem. Rev., 2022, 122, 11474–11513 CrossRef CAS PubMed.
- Y.-C. Lin, M. Karlsson and M. Bettinelli, Inorganic phosphor materials for lighting, Top. Curr. Chem., 2016, 374, 21 CrossRef PubMed.
- S. Gu, M. Xia, C. Zhou, Z. Kong, M. S. Molokeev, L. Liu, W.-Y. Wong and Z. Zhou, Red shift properties, crystal field theory and nephelauxetic effect on Mn4+-doped SrMgAl10−yGayO17 red phosphor for plant growth LED light, Chem. Eng. J., 2020, 396, 125208 CrossRef CAS.
- X. Zhao, Y. Wang, Z. Pan, Y. Lu, J. Li and M. Wu, Synthesis of Sm3+-doped YGa1.5Al1.5(BO3)4 phosphor via mechanical activation-assisted solid-state reaction, Dalton Trans., 2023, 52, 4808–4818 RSC.
- T. Hu, Y. Gao, M. S. Molokeev, Z. Xia and Q. Zhang, Eu2+ stabilized at octahedrally coordinated Ln3+ site enabling red emission in Sr3LnAl2O7.5 (Ln = Y or Lu) phosphors, Adv. Opt. Mater., 2021, 9, 2100077 CrossRef CAS.
- J. Qiao, L. Ning, M. S. Molokeev, Y.-C. Chuang, Q. Zhang, K. R. Poeppelmeier and Z. Xia, Site-selective occupancy of Eu2+ toward blue-light-excited red emission in a Rb3YSi2O7:Eu phosphor, Angew. Chem., Int. Ed., 2019, 58, 11521–11526 CrossRef CAS PubMed.
- J. Ding, M. Kuang, S. Liu, Z. Zhang, K. Huang, J. Huo, H. Ni, Q. Zhang and J. Li, Sensitization of Mn2+ luminescence via efficient energy transfer to suit the application of high color rendering WLEDs, Dalton Trans., 2022, 51, 9501–9510 RSC.
- Y. Zhuo, S. Hariyani, J. Zhong and J. Brgoch, Creating a green-emitting phosphor through selective rare-earth site preference in NaBaB9O15:Eu2+, Chem. Mater., 2021, 33, 3304–3311 CrossRef CAS.
- J. Li, Q. Liang, J.-Y. Hong, J. Yan, L. Dolgov, Y. Meng, Y. Xu, J. Shi and M. Wu, White light emission and enhanced color stability in a single-component host, ACS Appl. Mater. Interfaces, 2018, 10, 18066–18072 CrossRef CAS PubMed.
- D. Zhang, X. Zhang, B. Zheng, Q. Sun, Z. Zheng, Z. Shi, Y. Song and H. Zou, Li+ ion induced full visible emission in single Eu2+-doped white emitting phosphor: Eu2+ site preference analysis, luminescence properties, and WLED applications, Adv. Opt. Mater., 2021, 9, 2100337 CrossRef CAS.
- M. Shang, C. Li and J. Lin, How to produce white light in a single-phase host?, Chem. Soc. Rev., 2014, 43, 1372–1386 RSC.
- D. Lin and P. Du, Room temperature synthesis of Tb3+-activated BiF3 green-emitting nanoparticles with high thermal stability for near-ultraviolet white light-emitting didoes, Appl. Phys. A: Mater. Sci. Process., 2021, 127, 65 CrossRef CAS.
- X. Liu, C. Lin and J. Lin, White light emission from Eu3+ in CaIn2O4 host lattices, Appl. Phys. Lett., 2001, 90, 081904 CrossRef.
- D. Zhu, M. Liao, Z. Mu and F. Wu, Preparation and luminescence properties of Ca9NaZn(PO4)7:Dy3+ single-phase white light-emitting phosphor, J. Electron. Mater., 2018, 47, 4840–4844 CrossRef CAS.
- S. Ma, J. Zhang, J. Chen, Z. Yang, S. Wang and Z. Ye, Synthesis, crystal structure and photoluminescence properties of novel Ba3Lu4O9:Ce3+ orange-red phosphors for white light emitting diodes, J. Alloys Compd., 2020, 819, 153047 CrossRef CAS.
- D. Wu, C. Shi, J. Zhou, Y. Gao, Y. Huang, J. Ding and Q. Wu, Full-visible-spectrum lighting enabled by site-selective occupation in the high efficient and thermal stable (Rb, K)2CaPO4F: Eu2+ solid-solution phosphors, Chem. Eng. J., 2022, 430, 133062 CrossRef CAS.
- W. U. Khan, L. Zhou, X. Li, W. Zhou, D. Khan, S.-I. Niaz and M. Wu, Single phase white LED phosphor Ca3YAl3B4O15:Ce3+,Tb3+,Sm3+ with superior performance: color-tunable and energy transfer study, Chem. Eng. J., 2021, 410, 128455 CrossRef.
- Z. Zhang, J. Li, N. Yang, Q. Liang, Y. Xu, S. Fu, J. Yan, J. Zhou, J. Shi and M. Wu, A novel multi-center activated single-component white light-emitting phosphor for deep UV chip-based high color-rendering WLEDs, Chem. Eng. J., 2020, 390, 124601 CrossRef CAS.
- G. Li, Z. Hou, C. Peng, W. Wang, Z. Cheng, C. Li, H. Lian and J. Lin, Electrospinning derived one-dimensional LaOCl: Ln3+ (Ln = Eu/Sm, Tb, Tm) nanofibers, nanotubes and microbelts with multicolor-tunable emission properties, Adv. Funct. Mater., 2010, 20, 3446–3456 CrossRef CAS.
- Z. Zheng, P. Xie, P. Yi, J. Zhao, L. Liu and F. Yi, Achieving white-light/tunable emissions via the controllable energy transfer in the single Ba9La2Si6O24:Eu3+,Bi3+ phosphor for UV converted white LEDs, J. Lumin., 2020, 221, 117052 CrossRef CAS.
- M. Xin, D. Tu, H. Zhu, W. Luo, Z. Liu, P. Huang, R. Li, Y. Cao and X. Chen, Single-composition white-emitting NaSrBO3:Ce3+,Sm3+,Tb3+ phosphors for NUV light-emitting diodes, J. Mater. Chem. C, 2015, 3, 7286–7293 RSC.
- Y. Zhang, Z. Lin and G. Wang, Synthesis, growth, structure and characterization of the new laser host crystal Sr3Y2(BO3)4, Laser Phys. Lett., 2013, 10, 075806 CrossRef CAS.
- B. V. Mill, A. M. Tkachuk, E. L. Belokoneva, G. I. Ershova, D. I. Mironov and I. K. Razumov, Spectroscopic studies of Ln2Ca3B4O12:Nd3+ (Ln = Y, La, Gd) crystals, J. Alloys Compd., 1998, 275–277, 291–294 CrossRef CAS.
- B. Wei, Z. Hu, Z. Lin, L. Zhang and G. Wang, Growth and spectral properties of Er3+/Yb3+-codoped Ca3Y2(BO3)4 crystal, J. Cryst. Growth, 2004, 273, 190–194 CrossRef CAS.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169–11186 CrossRef CAS PubMed.
- P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized gradient approximation made simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- V. Petříček, M. Dušek and L. Palatinus, Crystallographic computing system JANA2006: General features, Z. Kristallogr. - Cryst. Mater., 2014, 229, 345–352 CrossRef.
- N. Yang, Z. Zhang, L. Zou, J. Chen, H. Ni, P. Chen, J. Shi and Y. Tong, A novel red-emitting phosphor with an unusual concentration quenching effect for near-UV-based WLEDs, Inorg. Chem. Front., 2022, 9, 6358–6368 RSC.
- Y. Wei, G. Xing, K. Liu, G. Li, P. Dang, S. Liang, M. Liu, Z. Cheng, D. Jin and J. Lin, New strategy for designing orangish-red-emitting phosphor via oxygen-vacancy-induced electronic localization, Light: Sci. Appl., 2019, 8, 15 CrossRef PubMed.
- X. Li, B. Milićević, M. D. Dramićanin, X. Jing, Q. Tang, J. Shi and M. Wu, Eu3+-activated Sr3ZnTa2O9 single-component white light phosphors: emission intensity enhancement and color rendering improvement, J. Mater. Chem. C, 2019, 7, 2596–2603 RSC.
- M. D. Dramicanin, L. Marciniak, S. Kuzman, W. Piotrowski, Z. Ristic, J. Perisa, I. Evans, J. Mitric, V. Dordevic, N. Romcevic, M. G. Brik and C. G. Ma, Mn5+-activated Ca6Ba(PO4)4O near-infrared phosphor and its application in luminescence thermometry, Light: Sci. Appl., 2022, 11, 279 CrossRef CAS PubMed.
- R. Shi, B. Li, C. Liu and H. Liang, On doping Eu3+ in Sr0.99La1.01Zn0.99O3.495: The photoluminescence, population pathway, de-excitation mechanism, and decay dynamics, J. Phys. Chem. C, 2016, 120, 19365–19374 CrossRef CAS.
- R. Shi, J. Xu, G. Liu, X. Zhang, W. Zhou, F. Pan, Y. Huang, Y. Tao and H. Liang, Spectroscopy and luminescence dynamics of Ce3+ and Sm3+ in LiYSiO4, J. Phys. Chem. C, 2016, 120, 4529–4537 CrossRef CAS.
- L. Cao, W. Li, B. Devakumar, N. Ma, X. Huang and A. F. Lee, Full-spectrum white light-emitting diodes enabled by an efficient broadband green-emitting CaY2ZrScAl3O12:Ce3+ garnet phosphor, ACS Appl. Mater. Interfaces, 2022, 14, 5643–5652 CrossRef CAS PubMed.
- S. Zhao, S. Liao, R. Shi, J. Zhang, Y. Han and S. Lian, Tuning emission color of Eu2+-activated phosphor through phase segregation, Chem. Eng. J., 2023, 452, 139640 CrossRef CAS.
- Z. Zhang, J. Yan, Q. Zhang, G. Tian, W. Jiang, J. Huo, H. Ni, L. Li and J. Li, Enlarging sensitivity of fluorescence intensity ratio-type thermometers by the interruption of the energy transfer from a sensitizer to an activator, Inorg. Chem., 2022, 61, 16484–16492 CrossRef CAS PubMed.
- Y. Chen, F. Liu, Z. Zhang, J. Hong, M. S. Molokeev, I. A. Bobrikov, J. Shi, J. Zhou and M. Wu, A novel Mn4+-activated fluoride red phosphor Cs30(Nb2O2F9)9(OH)3·H2O: Mn4+ with good waterproof stability for WLEDs, J. Mater. Chem. C, 2022, 10, 7049–7057 RSC.
- Y. Yin, M. Zhu, Y. Zhang, C. Hu, Z. Wang, Y. Che, J. Lu, R. Sun, J. Zhao, B. Teng, J. Li, S. Sun and D. Zhong, Achievement of high-efficiency sunlight-like emission in Rb2CaP2O7: Ce3+, Eu2+ phosphors, J. Lumin., 2022, 242, 118547 CrossRef CAS.
- D. Liu, P. Dang, X. Yun, G. Li, H. Lian and J. Lin, Luminescence color tuning and energy transfer properties in (Sr,Ba)2LaGaO5:Bi3+,Eu3+ solid solution phosphors: Realization of single-phased white emission for WLEDs, J. Mater. Chem. C, 2019, 7, 13536–13547 RSC.
- P. Dai, Q. Wang, M. Xiang, T.-M. Chen, X. Zhang, Y.-W. Chiang, T.-S. Chan and X. Wang, Composition-driven anionic disorder-order transformations triggered single-Eu2+-converted high-color-rendering white-light phosphors, Chem. Eng. J., 2020, 380, 122508 CrossRef CAS.
- P. Vinodkumar, S. Panda, G. Jaiganesh, R. K. Padhi, U. Madhusoodanan and B. S. Panigrahi, SrBPO5: Ce3+, Dy3+ - A cold white-light emitting phosphor, Spectrochim. Acta, Part A, 2021, 253, 119560 CrossRef CAS PubMed.
- B. Zheng, X. Zhang, D. Zhang, F. Wang, Z. Zheng, X. Yang, Q. Yang, Y. Song, B. Zou and H. Zou, Ultra-wideband phosphor Mg2Gd8(SiO4)6O2:Ce3+,Mn2+: Energy transfer and pressure-driven color tuning for potential applications in LEDs and pressure sensors, Chem. Eng. J., 2022, 427, 131897 CrossRef CAS.
- A. Bindhu, A. S. Priya, J. I. Naseemabeevi and S. Ganesanpotti, Deciphering crystal structure and photophysical response of Bi3+ and Pr3+ co-doped Li3Gd3Te2O12 for lighting and ratiometric temperature sensing, J. Alloys Compd., 2022, 893, 162246 CrossRef CAS.
- M. Chen, K. Qiu, P. Zhang, W. Zhang and Q. Yin, Enhancement of NaSrVO4:Dy3+-white-phosphor photoluminescence via La3+ co-doping, Ceram. Int., 2019, 45, 22547–22552 CrossRef.
- Y. Zheng, H. Li, T. Yang, J. Li, S. Yang and J. Zhu, Bi4BPO10:Dy3+: A single-phase white-emitting phosphor for light-emitting diodes, Mater. Today Chem., 2022, 26, 101199 CrossRef CAS.
|
This journal is © the Partner Organisations 2023 |
Click here to see how this site uses Cookies. View our privacy policy here.