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

Luminescence properties of Tb-doped Ba3Y2O5Cl2

Open Access Open Access

Abstract

Herein, we report the synthesis and luminescence properties of Tb-doped layered perovskite oxychloride, Ba3Y2O5Cl2. Similar to Ba3Y2O5Cl2:Eu3+, single-phase Ba3Y2O5Cl2:Tb3+ was successfully synthesized through a solid-state reaction. The luminescence properties were studied and compared with those of Y2O3:Tb3+. There was only one excitation peak due to the 4f-5d transition of Tb3+, which was different from that of Y2O3:Tb3+ that comprises two peaks. Green emissions from Tb3+ 4f−4f transitions were observed for both the samples. The luminescence properties at different doping concentrations were investigated. Concentration quenching at 8% was found, which is rather high, owing to the specially separated layered structure. The decay times of luminescence were also studied, which are consistent with the concentration quenching. In addition, energy diagram for the luminescence mechanism were also proposed. Considering the photoluminescence spectra and high concentration quenching, this compound is a potential mother compound for optical materials, especially for the green phosphors.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

In recent years, mixed-anion compounds have attracted significant attention owing to their various functionalities, which originate from their unique structural and electronic properties [14]. These properties are induced by the incorporation of multiple anions with different ionic radii, electronegativities, oxidation states, and polarizabilities. One of the most important aspects of studies based on mixed-anion compounds is to control and understand the arrangement of anions that determine their electronic structures, for example, the indirect bandgap characteristics of Sr2ScO3Cl originating from ScO5 pyramids [5] and two-dimensional quantum antiferromagnetism in Sr2CuO2Cl2 with trans-configuration of Cl ions in the CuO4Cl2 octahedra [6].

Ruddlesden–Popper (RP)-type layered perovskite with a general chemical formula An+1BnX3n+1 (where A = electropositive cation, B = transition metal, X = anion, and n = perovskite block number) is an interesting stage for studying the relationship between the anion order and physical and chemical properties [79]. The ideal RP-type structure contained more than one different anion site. For example, for the n = 1 phase, the apical and equatorial anion sites of the octahedra are surrounded by AB5 and A2B4 polyhedra, respectively, which differ from six equivalent sites with A4B2 coordination in the ideal cubic perovskite system ABX3. Therefore, if there are large differences in these anions, this phase can form ordered arrangements.

In our previous study, we reported the synthesis, crystal structure, and optical properties of a new RP-type layered perovskite oxychloride, Ba3Y2O5Cl2 [4]. The crystal structure is shown in Fig. 1. The Y atom in this compound shifted toward the apical oxygen site owing to the asymmetric coordination of O and Cl. The direct bandgap of the compound is ∼5.4 eV. In addition, we investigated the luminescence properties with Eu doping. As Y3+ ions with a distorted octahedron have an asymmetric coordination environment, both magnetic and electric dipole transitions were observed by doping Eu3+ to Y3+ sites, which gives characteristic orange luminescence. This compound also exhibited a high internal quantum yield (IQY) (>70%). Considering the characteristic luminescence and high IQY, this compound is a potential host lattice for luminescent materials.

 figure: Fig. 1.

Fig. 1. Crystal structure of Ba3Y2O5Cl2. Gray, blue, red, and green balls represent Ba, Y, O and Cl ions, respectively. The blue pyramids show polyhedral of YO5.

Download Full Size | PDF

It is well known that the Tb3+ ion shows green emission when doped into wide-gap materials [10,11]. In addition, it emits some sharp lines in a wide visible range, and the most intense line at approximately 545 nm is frequently utilized in green lighting. Luminescence characteristics, such as peak positions, peak shapes, and decay time, are all influenced by the crystal field and site symmetry. These characteristics can also be observed when perovskites and/or layered perovskites are used as host lattices [12,13]. For Tb3+-doped LaScO3 perovskite, Tb3+ at B sites shows green luminescence; in contrast, Tb3+ at A sites shows blue-violet and green luminescence. The decay time of luminescence from the B sites was systematically longer than that from the A sites [14].

For phosphors with a perovskite-related matrix, it is common to locate the luminescence center to the A site, whereas doping at the B site is relatively unusual [15]. In order to accommodate lanthanide ions at B sites, not only B ions but also A ions are required to be large [16], similar to the case for Ba3Y2O5Cl2. In addition, owing to its special structural and electronic properties, it is interesting to investigate the luminescence of Tb3+-doped Ba3Y2O5Cl2.

In this study, we synthesized Ba3Y2O5Cl2 with different Tb3+ doping concentrations. The luminescence properties were studied and compared with those of the Y2O3: Tb3+ phosphors. We also proposed the luminescence mechanisms and concentration quenching.

2. Experimental procedure

2.1 Synthesis

Polycrystalline samples of Ba3Y2O5Cl2 were synthesized by a conventional solid-state reaction with stoichiometric amounts of BaCO3 (3N), BaCl2 (3N), and Y2O3 (3N), according to our previous report [4]. Powder mixtures were pelletized and heated one or two times at 1173 K in air for 24 h with intermittent grindings.

Polycrystalline samples of Ba(Y1-xTbx)2O5Cl2 (Ba3Y2O5Cl2:Tb3+, x = 0-0.10) were synthesized by the same method with stoichiometric amounts of BaCO3 (3N), BaCl2 (3N), Y2O3 (3N) and the corresponding amount of Tb4O7 (3N) and heating at 1248 K under 5% H2/Ar for 12 h. Y2O3 and Tb4O7 were used as starting materials and fired at 1773 K in air for 24 h, to prepare Y2O3:Tb3+ powder for comparison.

2.2 X-ray diffraction

The sintered samples were measured by powder X-ray diffraction (PXRD) with an Ultima-IV diffractometer (Rigaku). The data were collected in the range 5° ≤ 2θ ≤ 80° with a step size of 0.02° using Cu radiation. The lattice parameters were refined using PXRD patterns with a Si internal standard.

2.3 Optical properties

Photoluminescence spectra were collected using a spectrofluorometer (JASCO, FP-8500). All the optical measurements were performed at room temperature.

3. Results and discussion

Ba3Y2O5Cl2:Tb3+ samples were successfully prepared by solid-state synthesis, as reported previously. Figure 1 shows the crystal structure of Ba3Y2O5Cl2. Figure 2(a) shows the XRD patterns of Tb-doped Ba3Y2O5Cl2 samples. Almost a single phase was formed, except for a small amount of Y2O3 impurity. Considering that there is no Tb4O7 impurity, we can speculate that Tb ions were successfully doped into the matrix. The lattice parameters for different Tb concentrations are plotted in the inset of Fig. 2(b). The lattice parameters of undoped sample were a = 4.3971(8) Å and c = 24.848(5) Å [4]. Since ionic radius of Y3+ (0.9 Å, coordination number (C.N.) 6) and Tb3+ (0.923 Å, C.N. 6) are similar [17], Tb3+ doping does not affect the lattice constants too much, except for a slight increase in the a-axis. For the c-axis, the lattice constants show large disturbances, especially for the steep drop of 10%- doped sample. This is consistent with the crystal structure of this compound. Y-Cl is weakly bonded compared to Y-O; therefore, with increasing Tb concentration, doping expands only the a-axis but not the c-axis [4,12]. The unit cell volumes (Vcell) calculated from the lattice parameters with different Tb concentrations are plotted in Fig. 2(b). We can observe that the general trend increases with increasing Tb concentration. The increased cell volume is consistent with related rare earth ions doing at B site rather than A site [1214], which is in accord with the specific photoluminescent property originates from the B site ions that we will discuss later.

 figure: Fig. 2.

Fig. 2. (a) XRD patterns and (b) unit cell volumes of Ba3Y2O5Cl2:xTb3+ (x = 0, 1, 2,3, 4, 6, 8, and 10%). Inset is lattice parameters of Ba3Y2O5Cl2: xTb3+ (x = 0-10%).

Download Full Size | PDF

Figure 3 shows the excitation and emission spectra of Ba3Y2O5Cl2:1% Tb3+ and Y2O3:1% Tb3+ at room temperature. The excitation spectrum was recorded by monitoring with bright green emission at 544 nm owing to the 4f8 −4f75d1 transition of the Tb3+ ions. Both of the samples show excitation peak characteristics of the 4f−5d transition of Tb3+. We also observed several weak peaks in the range of 300–500 nm, which were caused by the spin and orbital forbidden intra-4f transitions (see inset in Fig. 3) [18]. For the Ba3Y2O5Cl2:Tb3+ sample, the broad peak at 238 nm is the host lattice excitation peak, which is consistent with the optical band gap of the Ba3Y2O5Cl2 host. Another broad peak at 290 nm can be ascribed to the 4f8 -4f75d1 transitions of Tb3+ ions. In the case of the 4f→5d transitions of Y2O3:Tb3+, there are two peaks at ∼278 and 304 nm, whereas in Ba3Y2O5Cl2:Tb3+, there is only a peak at ∼290 nm. The shape of 4f→5d band varied, indicating different host lattice structures. In Y2O3, there are two different sites of Y3+, whereas in Ba3Y2O5Cl2, there is only one site [4].

 figure: Fig. 3.

Fig. 3. Photoluminescence of Ba3Y2O5Cl2:1% Tb3+(up) and Y2O3:1% Tb3+ (down) measured at room temperature. The dashed and solid lines show the excitation spectrum and emission line, respectively (inset: expanded spectrum of excitation spectrum of Ba3Y2O5Cl2:1% Tb3+).

Download Full Size | PDF

In the emission spectrum, when Ba3Y2O5Cl2:Tb3+ and Y2O3:Tb3+ were excited at wavelengths of 290 and 304 nm, respectively, photoluminescence from Tb3+ 4f−4f transitions were observed in both samples. The spectra show emission lines peaking at 487-498, 539-560, 585-597, and 616-635 nm due to transitions of 5D47F6, 5D47F5, 5D47F4, and 5D47F3 levels from Tb3+, respectively. The most intense emission at 544 nm stems from the 5D47F5 transition, which is a green emission. These results are consistent with the results for Tb3+ ions in other hosts, such as Y3Al5O12 and Lu2O3 [1921]. The 5D37FJ emissions were not observed in our experiments. It is noteworthy that the highest peak at 544 nm splits into two peaks for both the samples. Ba3Y2O5Cl2:Tb3+ splits to 544 and 554 nm for Y2O3:Tb3+ at 544 nm and 552 nm, respectively. Further, we observe that the intensity ratio of split peaks is different from that of Y2O3:Tb3+. This type of peak split is due to the effect of crystal field energy. The split width for Ba3Y2O5Cl2:Tb3+ was 10 nm, while that for Y2O3:Tb3+ was 8 nm. This larger split width suggests that the crystal field energy of Ba3Y2O5Cl2:Tb3+ is larger than that of Y2O3:Tb3+ [11]. For Ba3Y2O5Cl2, the average bond length of Y-O is 2.2038 Å, whereas that for Y2O3 is 2.2842 Å, which is much larger than that for Ba3Y2O5Cl2 [4]. The closer the distance of Y and O ions, the stronger the repulsion of the O ions and the ${d_{{x^2} - {y^2}}}$ $\textrm{and}\; {d_{{z^2}}}\; $ orbitals of Y ions, causing the larger crystal field energy.

Figure 4 shows the PLE and PL spectra with different Tb concentrations. With an increase in Tb concentration, the intensity increases gradually, reaching a maximum value when the Tb3+ concentration is up to 8%. When the doping concentration is beyond 8%, the intensity decreases significantly. For increasing dopant concentrations, higher light outputs are obtained due to the larger number of luminescent centers. Up to a critical point, the intensity decreases. In addition, the spectral shape and locations of the peaks do not vary with the doping concentrations of Tb3+ ions for Ba3Y2O5Cl2:Tb3+. From the normalized PLE spectra, as shown in the inset of Fig. 4(a), we can observe a slight red shift of the f-d excitation band for increasing Tb concentrations. The Tb3+ has [Xe]4f8 electronic configuration, the excitation of Tb3+ originates from 4f8 to 4f75d1 transition. Because of the shielding by the outer 5s and 5p orbitals, the 4f electron is insensitive to the crystal field. However, the 5d electron interacts strongly with the crystal lattice, resulting in strong phonon coupling and large crystal field effects on the excited states [22]. In addition, we observe that the excitation bands of f-d excitation become broader with the increment of Tb3+ concentration. It’s known that the crystal field splitting energy of 4d element is larger than that of 3d element. Therefore, this broader bands can be assigned to the larger splitting of energy levels as a result of the larger crystal field energy caused by Tb3+ doping.

 figure: Fig. 4.

Fig. 4. The room temperature (a) PLE spectra and (b) PL spectra of Ba3Y2O5Cl2:xTb3+ (x = 1, 2, 3, 4, 6, 8, and 10%) phosphors (λex= 290 nm). Inset of (a) shows the normalized PLE spectra, inset of (b) shows the expanded spectra of the emission spectra.

Download Full Size | PDF

Figure 5(a) shows the photograph of Ba3Y2O5Cl2:xTb3+ under excitation by UV lamp. Figure 5(b) shows the normalized intensities of the strongest emission peak (544 nm) with different Tb concentrations. Compared with simple perovskite host materials, this quenching concentration is very high 12; this can be attributed to the layered structure of Ba3Y2O5Cl2. Tb3+ occupies the center sites of the YO5 polyhedron, and Ba-Cl layers can effectively isolate each Tb3+. The relatively large spatial distance between Tb3+ ions in the layers results in a reduction in nonradiative energy between sites, which serves to suppress concentration quenching effect [23]. This can be further demonstrated by Density Functional Theory (DFT) calculation result in our previous study 4. When x = 8%, the luminescence center is just the right saturation, as the emission and excitation intensity are at their strongest values, which are higher than those at other concentrations. When x >8%, Tb3+ ions position close to each other, and the probability of interaction between adjacent two Tb3+ ions increases; this, in turn, increases the new energy loss, thereby resulting in dislocation and cross-relaxation phenomenon significantly weakening the luminescence properties of phosphors [24]. With layered structure separating active ions, quenching can be hindered [25]. Therefore, the quenching concentrations of this compound are relatively higher.

 figure: Fig. 5.

Fig. 5. (a) Ba3Y2O5Cl2:xTb3+ under excitation by UV lamp (lem.  250 nm); and (b) relation between PL intensity and Tb3+ concentration.

Download Full Size | PDF

The emission intensity decreased because of the energy transfer between the Tb3+ ions. In general, there are two main kinds of energy transfer: one is exchange interaction, wherein the typical critical distance is no more than 5 Å, and the other is multipolar interaction, wherein the critical distance is longer than 10 Å [19,20]. The distance between adjacent ions when concentration quenching occurs is called the critical distance (Rc). Rc can be calculated using the following formula:

$${R_c} = 2{\left[ {\frac{{3V}}{{4\pi {x_c}N}}} \right]^{1/3}}$$
where xc is the critical concentration, N is the number of available sites for the dopant in the unit cell, and V is the volume of the unit cell [21]. The critical concentration in this study was 8%, V was 482.21 Å3 and N was 10. The critical distance was determined to be 10.48 Å. Because the distance is much larger than 5 Å, we can draw a conclusion that the exchange interaction plays little role in energy transfer. Therefore, energy transfer is considered as multipolar interaction. There are three forms in electric multipolar interaction, named dipole-dipole (d-d), dipole-quadrupole (d-q) and quadrupole-quadrupole (q-q) interactions. Based on the Dexter theory about multipolar interactions [2629], the emission intensity (I) of each activator ion can be obtained as follows:
$$\frac{I}{x} = k{[{1 + \beta {x^{Q/3}}} ]^{ - 1}}$$
where I is the luminescence emission intensity, x is the activator concentration, and k and b are constants for the same excitation conditions for a given host. For dipole-dipole, dipole-quadrupole, or quadrupole-quadrupole interactions, Q is 6, 8, and 10, respectively. In addition, if Q is less than 6, it can be attributed to the charge transfer mechanism. When the activator ion content reaches the concentration quenching, according to Reisfeld’s approximation [28], formula (2) about interaction type between sensitizer and sensitizer or between sensitizer and activator can be simplified to formula (3):
$$\log \left( {\frac{I}{x}} \right) = C - \frac{Q}{3}\textrm{log}\; x$$
The nature of energy transfer can be determined by plotting a graph between log(I/x) and log x, as shown in Fig. 6. The slope of fitted curve is approximately −1.85; thus, the estimated Q value is 5.55, which is closer to 6. Therefore, we can draw a conclusion that the principal concentration quenching mechanism of Ba3Y2O5Cl2: Tb3+ phosphors is dipole-dipole interaction [27,30].

 figure: Fig. 6.

Fig. 6. Relationship of log (I/x) vs. log (x)for Ba3Y2O5Cl2:Tb3+ phosphor.

Download Full Size | PDF

The luminescence lifetime at the 5D47F5 transition in Ba3Y2O5Cl2:8%Tb3+ is shown in Fig. 7(a). The decay curve I(t) was fitted using the following equation:

$$I(t )= I(0 )\textrm{exp}\left( { - \frac{t}{\tau }} \right)$$
where I (0) is the initial intensity and τ is the decay constant. The decay curve fitted well to the one-component exponential decay; the decay constant τ was evaluated as 1.27 ms. Figure 7(b) shows the lifetimes of the luminescence at 5D47F5 transition with different concentrations of Tb3+. Before 8%, the lifetimes remain almost constant; after 8%, it decreases significantly. This trend is consistent with the concentration quenching of luminescence at 8%.

 figure: Fig. 7.

Fig. 7. (a) Luminescence decay of Ba3Y2O5Cl2:8% Tb3+; (b) lifetimes of Ba3Y2O5Cl2:xTb3+ (x = 0.01–0.1). The excitation and monitoring wavelength are 290 and 544 nm, respectively.

Download Full Size | PDF

Figure 8 shows the proposed luminescence mechanisms in Ba3Y2O5Cl2:Tb3+. Tb3+ can be excited through 4f8 → 4f75d1 transitions. Empty Tb 5d orbitals are expanded more than 4f orbitals and easily influenced by the crystal field. Due to the crystal field energy of Ba3Y2O5Cl2, the empty Tb 5d orbitals have larger splitting. The bottom of Tb3+ 5d orbitals shift near 5D3 regimen, whereas they shift below the 5D3 excited states by relaxation just after excitation, that is 5D3 energy levels in Ba3Y2O5Cl2:Tb3+ lie above the band gap or within the conduction band. Therefore, it can give rise only to green luminescence originated from 5D4 to 7FJ, and no energy transfer occurs from the 5d orbitals to 5D3 excited states, thus 5D3 to 7FJ emission cannot be observed.

 figure: Fig. 8.

Fig. 8. Schematic luminescence mechanisms in Ba3Y2O5Cl2:Tb3+ on 4f–5d excitation.

Download Full Size | PDF

4. Conclusion

Herein, we successfully synthesized Tb-doped Ba3Y2O5Cl2 via a solid-state reaction. Green emission from the Tb3+ 4f−4f transitions was observed. Peak splitting of green 5D4 luminescence was observed. The width of split peaks of Ba3Y2O5Cl2 is larger than that of Y2O3 owing to the stronger crystal field energy of Ba3Y2O5Cl2. In addition, the photoluminescence at different concentrations was investigated. Concentration quenching at 8% was found, which is much higher than that in simple perovskite host materials. This concentration quenching is due to the dipole-dipole interaction between neighboring Tb3+ ions. Further, luminescence mechanisms have been proposed. Layered perovskite structure, Ba3Y2O5Cl2, can be host lattice for different luminescence centers, and also hinder luminescence quenching, therefore, it is promising to be used in optical devices.

Funding

Japan Society for the Promotion of Science (JP16H06439, JP16H06440), 18J01627, 20K15032); High-level Introduction of Talents for Scientific Research Start-up Funds in China Pharmaceutical University ( 3150050052).

Acknowledgments

The authors of this work gratefully appreciate the financial support provided by Japan Society for the Promotion of Science (JSPS) Grant-in-Aid for Scientific Research on Innovative Areas Mixed Anion”, Grant-in-Aid, and High-level Introduction of Talents for Scientific Research Start-up Funds from China Pharmaceutical University. We thank Dr. Huiwen Lin for discussions and feedback on the manuscript.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. H. Kageyama, K. Hayashi, K. Maeda, J. P. Attfield, Z. Hiroi, J. M. Rondinelli, and K. R. Poeppelmeier, “Expanding frontiers in materials chemistry and physics with multiple anions,” Nat. Commun. 9(1), 772 (2018). [CrossRef]  

2. M. Kobayashi, T. Yasunaga, H. Sato, H. Kato, K. Fujii, M. Yashima, and M. Kakihana, “Synthesis, structure, and photoluminescence of a novel oxynitride BaYSi2O5N activated by Eu2+ and Ce3+,” Chem. Lett. 46(6), 795–797 (2017). [CrossRef]  

3. Y. Masubuchi, M. Tadaki, and S. Kikkawa, “Synthesis of the perovskite SrTaO2N using C3N4 for both reduction and nitridation,” Chem. Lett. 47(1), 31–33 (2018). [CrossRef]  

4. Y. Iwasa, Y. Su, Y. Tsuchiya, M. Tatsuda, K. Kishio, T. Yanagida, F. Takada, T. Nishio, Y. Tsujimoto, K. Fujii, M. Yashima, and H. Ogino, “Synthesis, structure, and luminescence properties of layered oxychloride Ba3Y2O5Cl2,” J. Mater. Chem. C 8(48), 17162–17168 (2020). [CrossRef]  

5. Y. Su, Y. Tsujimoto, K. Fujii, M. Tatsuta, K. Oka, M. Yashima, H. Ogino, and K. Yamaura, “Synthesis, crystal structure, and optical properties of layered perovskite scandium oxychlorides: Sr2ScO3Cl, Sr3Sc2O5Cl2, and Ba3Sc2O5Cl2,” Inorg. Chem. 57(9), 5615–5623 (2018). [CrossRef]  

6. L. L. Miller, X. L. Wang, S. X. Wang, C. Stassis, D. C. Johnston, J. Faber, and C. K. Loong, “Synthesis, structure, and properties of Sr2Cu2O2Cl2,” Phys. Rev. B 41(4), 1921–1925 (1990). [CrossRef]  

7. N. Tarasova, A. Galisheva, I. Animitsa, D. Korona, and K. Davletbaev, “Novel proton-conducting layered perovskite based on BaLaInO4 with two different cations in B-sublattice: Synthesis, hydration, ionic (O2−, H+) conductivity,” Int. J. Hydrog. Energy 47(44), 18972–18982 (2022). [CrossRef]  

8. A. Kumar, K. Kumari, M. K. Sharma, A. Vij, S. Kumar, S. H. Huh, and B. H. Koo, “Chemically inducing room temperature spin-crossover in double layered magnetic refrigerants Pr1.4 + xSr1.6-xMn2O7 (0.0 ≤ x ≤ 0.5),” J. Mater. Sci. Technol. 124, 232–242 (2022). [CrossRef]  

9. S. T. Li and T. Birol, “Free-Carrier-Induced Ferroelectricity in Layered Perovskites,” Phys. Rev. Lett. 127(8), 087601 (2021). [CrossRef]  

10. R. E. Muenchausen, L. G. Jacobsohn, B. L. Bennett, E. A. McKigney, J. F. Smith, J. A. Valdez, and D. W. Cooke, “Effects of Tb doping on the photoluminescence of Y2O3: Tb nanophosphors,” J. Lumin. 126(2), 838–842 (2007). [CrossRef]  

11. T. F. Ma, H. F. Li, S. Zhang, W. Z. Sun, Z. Q. Cheng, R. Pang, J. Feng, L. H. Jiang, D. Li, and C. Y. Li, “Study of a color-tunable long afterglow phosphor Gd1.5Y1.5Ga3Al2O12: Tb3+: luminescence properties and mechanism,” RSC Adv. 10(47), 28049–28058 (2020). [CrossRef]  

12. K. Ueda, S. Tanaka, R. Yamamoto, Y. Shimizu, T. Honma, F. Massuyeau, and S. Jobic, “Site dependence of Tb3+ luminescence in double perovskite-Type alkaline earthlanthanum tantalates,” J. Phys. Chem. C 124(1), 854–860 (2020). [CrossRef]  

13. K. Ueda, S. Tanaka, T. Yoshino, Y. Shimizu, and T. Honma, “Site-selective doping and site-sensitive photoluminescence of Eu3+ and Tb3+ in perovskite-yype LaLuO3,” Inorg. Chem. 58(16), 10890–10897 (2019). [CrossRef]  

14. K. Ueda, T. Aoki, T. Shimizu, Y. Massuyeau, and S. Jobic, “Luminescence and location of Gd3+ or Tb3+ ions in perovskite-type LaScO3,” Inorg. Chem. 57(15), 8718–8721 (2018). [CrossRef]  

15. D. Y. Huang, Y. Wei, P. P. Dang, X. Xiao, H. Z. Lian, and J. Lin, “Tunable color emission in LaScO3:Bi3+,Tb3+,Eu3+ phosphor,” J. Am. Ceram. Soc. 103(5), 3273–3285 (2020). [CrossRef]  

16. Y. Shimizu and K. Ueda, “Phase formation and UV luminescence of Gd3+ doped perovskite-type YScO3,” J. Solid State Chem. 242, 170–174 (2016). [CrossRef]  

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

18. M. B. Xie, W. K. Ruan, and J. Wang, “Energy transfer, tunable photoluminescence of Sr3Lu(PO4)3: Tb3+, Eu3+ phosphors,” J. Lumin. 224, 117278 (2020). [CrossRef]  

19. M. Upasani, “Synthesis of Y3Al5O12: Tb & Y3Al5O12: Tb, Si phosphor by combustion synthesis: Comparative investigations on the structural and spectral properties,” Opt. Mater. 64, 70–74 (2017). [CrossRef]  

20. G. Blasse and B. Grabmaier, Luminescent Materials (Springer–Verlag, 1994).

21. T. P. Joanna, N. Janne, H. Jorma, and Z. Eugeniusz, “Thermoluminescence and kinetics of persistent luminescence of vacuum-sintered Tb3+-doped and Tb3+, Ca2+-codoped Lu2O3 materials,” Chem. Mater. 20(6), 2252–2261 (2008). [CrossRef]  

22. L. Chen, X. Chen, F. Liu, H. Chen, H. Wang, E. Zhao, Y. Jiang, T. S. Chan, C. H. Wang, W. Zhang, Y. Wang, and S. Chen, “Charge deformation and orbital hybridization: intrinsic mechanisms on tunable chromaticity of Y3Al5O12:Ce3+ luminescence by doping Gd3+ for warm white LEDs,” Sci. Rep. 5(1), 11514 (2015). [CrossRef]  

23. L. Li, X. H. Tang, Z. J. Wu, Y. F. Zheng, S. Jiang, X. Tang, G. T. Xiang, and X. J. Zhou, “Simultaneously tuning emission color and realizing optical thermometry via efficient Tb3+→Eu3+ energy transfer in whitlockite-type phosphate multifunctional phosphors,” J. Alloys Compd. 780, 266–275 (2019). [CrossRef]  

24. C. Yang, Q. S. Liu, D. X. Huang, X. Li, X. Y. Zhang, Z. H. Bai, X. C. Wang, and X. Y. Mi, “Study on the mechanism of improved luminescence in SrB2Si2O8: Ce3+/Tb3+ phosphor,” J. Lumin. 214, 116541 (2019). [CrossRef]  

25. G. Blasse, “Luminescence of inorganic solids: from isolated centres to concentrated systems,” Solid State Chem. 18(2), 79–171 (1988). [CrossRef]  

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

27. R. Reisfeld, E. Greenberg, and R. Velapody, “Luminescence quantum efficiency of Gd and Tb in borate glasses and the mechanism of energy transfer between them,” J. Chem. Phys. 56(4), 1698–1705 (1972). [CrossRef]  

28. S. Tamboli, P.D. Bhoyar, and S.J. Dhoble, “Analysis of the electron-vibrational interaction in the 5d states of Eu2+ ions in LiSrPO4 host matrix,” J. Lumin. 214, 116564 (2019). [CrossRef]  

29. M. X. Ma, D. C. Zhu, and M. J. Tu, “The effect of Eu2+ doping concentration on luminescence properties of BaAl2Si2O8: Eu2+ blue phosphor,” Acta Phys. Sin. 58(8), 5826 (2009). [CrossRef]  

30. R. Vijayakumar, B. Devakumar, and X. Y. Huang, “Energy transfer induced color-tunable emissions from Ba2Gd5B5O17: Ce3+/Tb3+ borate phosphors for white LEDs,” J. Lumin. 229, 117685 (2021). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (8)

Fig. 1.
Fig. 1. Crystal structure of Ba3Y2O5Cl2. Gray, blue, red, and green balls represent Ba, Y, O and Cl ions, respectively. The blue pyramids show polyhedral of YO5.
Fig. 2.
Fig. 2. (a) XRD patterns and (b) unit cell volumes of Ba3Y2O5Cl2:xTb3+ (x = 0, 1, 2,3, 4, 6, 8, and 10%). Inset is lattice parameters of Ba3Y2O5Cl2: xTb3+ (x = 0-10%).
Fig. 3.
Fig. 3. Photoluminescence of Ba3Y2O5Cl2:1% Tb3+(up) and Y2O3:1% Tb3+ (down) measured at room temperature. The dashed and solid lines show the excitation spectrum and emission line, respectively (inset: expanded spectrum of excitation spectrum of Ba3Y2O5Cl2:1% Tb3+).
Fig. 4.
Fig. 4. The room temperature (a) PLE spectra and (b) PL spectra of Ba3Y2O5Cl2:xTb3+ (x = 1, 2, 3, 4, 6, 8, and 10%) phosphors (λex= 290 nm). Inset of (a) shows the normalized PLE spectra, inset of (b) shows the expanded spectra of the emission spectra.
Fig. 5.
Fig. 5. (a) Ba3Y2O5Cl2:xTb3+ under excitation by UV lamp (lem.  250 nm); and (b) relation between PL intensity and Tb3+ concentration.
Fig. 6.
Fig. 6. Relationship of log (I/x) vs. log (x)for Ba3Y2O5Cl2:Tb3+ phosphor.
Fig. 7.
Fig. 7. (a) Luminescence decay of Ba3Y2O5Cl2:8% Tb3+; (b) lifetimes of Ba3Y2O5Cl2:xTb3+ (x = 0.01–0.1). The excitation and monitoring wavelength are 290 and 544 nm, respectively.
Fig. 8.
Fig. 8. Schematic luminescence mechanisms in Ba3Y2O5Cl2:Tb3+ on 4f–5d excitation.

Equations (4)

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

R c = 2 [ 3 V 4 π x c N ] 1 / 3
I x = k [ 1 + β x Q / 3 ] 1
log ( I x ) = C Q 3 log x
I ( t ) = I ( 0 ) exp ( t τ )
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.