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

Eu3+ doped transparent germanate glass ceramic scintillators containing LaF3 nanocrystals for X-ray detection

Open Access Open Access

Abstract

Eu3+ doped transparent germanate glass ceramics containing LaF3 nanocrystals have been prepared by the conventional melt quenching technique with subsequent heat treatment. XRD, TEM and HRTEM were used to verify the formation of LaF3 nanocrystals. Emission spectra, fluorescence decay and X-ray excited luminescence (XEL) were employed to elucidate the optical properties of Eu3+ doped germanate glasses and glass ceramics. The maximum integrated XEL intensity of the glass ceramic is about 20% of that of the commercial Bi4Ge3O12 (BGO) scintillating crystal. The results indicate that Eu3+ doped germanate glass ceramic could be a promising scintillating material used in X-ray detection field.

© 2019 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

In the past decades, trivalent and divalent rare earth ions doped luminescent materials have attracted great attention due to their promising applications in white light-emitting-diodes (W-LED), temperature sensor, laser, scintillator, and so forth [1–7]. However, the luminescence efficiency of rare earth ions has a great relationship with the phonon energy of the host material [8]. The lower phonon energy of the host is, the higher luminescence efficiency could be. Compared with oxide glass matrix, fluoride host has lower phonon energy, but its poor chemical and mechanical stabilities restrict its practical applications [9]. Therefore, much attention has been devoted to transparent glass ceramics (GCs) which are embedded with fluoride nanocrystals in the oxide glass matrix. The oxyfluoride GC has comparatively low phonon energy because it contains fluoride nanocrystals, and also has high chemical and mechanical stability due to the oxide glass matrix. Consequently, rare earth ions doped transparent GCs containing various kinds of fluoride nanocrystals have been investigated. Compared with CaF2 [10], BaF2 [11–13], SrF2 [14,15], PbF2 [16], and other difluorides, lanthanide trifluorides nanocrystals are much suitable for rare earth ions doping because the same valence of lanthanide trifluorides can be easily substituted by active rare earth ions [8,17–19].

Rare earth ions doped scintillating materials have been extensively studied due to their applications in high-energy physics, nuclear radiation detection, medicine imaging and other X-ray detection fields [20–23]. Compared with single crystal and ceramic scintillator, rare earth ions (especially Ce3+ and Tb3+) doped glass scintillating materials have been widely investigated owing to their low cost, easy preparation, and easy to be made into various shapes [22,24,25]. However, the low luminescent efficiency of glass scintillators due to the amorphous structure of the glass hinders their practical applications. Therefore, it is important to improve their luminescence efficiency. Thus, rare earth ions doped scintillating GCs have become promising candidates [20,23,25–28] due to the advantages of glass scintillating materials compared with scintillating crystal, as well as the superiority of high luminescence efficiency compared with scintillating glass.

As one of the most widely studied rare earth ion, Eu3+ was often studied for its intense red emission (5D07FJ transition) and probe function for the rare earth ion site structure (intensity ratio of 5D07F2 and 5D07F1 transitions) [6,7,17,29]. Moreover, it also could exhibit intense red emission around 610 nm under the excitation of X-ray. Thus Eu3+ doped glass have been researched for its potential application as scintillating materials in the recent years. Most of those works are focused on the borosilicate and borogermanate glass system [30,34,35]. However, due to the amorphous structure of those glasses, the luminescence efficiency of them are not so high. Replacing glasses by glass ceramics may be a effective way to improve the luminescence efficiency [26]. But there is rare research on Eu3+ doped GC as a scintillating material.

In the present work, Eu3+ doped transparent germanate GCs containing LaF3 nanocrystals were successfully prepared. The incorporation of Eu3+ into LaF3 was evidenced by XRD, PL spectra and decay time measurements. The photoluminescence and the scintillating properties of Eu3+ doped GCs were investigated by PL, PLE and XEL spectra.

2. Experimental

Eu3+ doped germanate glasses with nominal composition of 58GeO2-8Al2O3-10Na2O-10LiF-(14-x)LaF3-xEuF3 (x = 0.1, 1, 2, 3, 4 and 5, in mol%) were prepared by a conventional melt quenching technique. About 20 g raw materials of GeO2 (99.999%), Al2O3 (99.99%), Na2CO3 (99.99%), LiF (99.99%), LaF3 (99.99%) and EuF3 (99.99%) were mixed homogeneously in an agate mortar for each batch, and then melt in a covered alumina crucible at 1470 °C for 45 min in an electric furnace in air. The melt was poured onto a preheated stainless steel plate to obtain the bulk glass. In order to release the inner stress, the obtained bulk glass was annealed at 500 °C for 3 h, then cooled to room temperature naturally. Transparent glass ceramics were formed after heat treatment on 3 pieces of 4 mol% Eu3+ doped germanate glass (the precursor glass, denoted PG) at 620 °C for 3 h, 640 °C for 3 h and 640 °C for 6 h, which were named as GC620°C-3h, GC640°C-3h and GC640°C-6h, respectively. All of the PG and GC samples were cut and polished for further optical measurements with the size of 10 mm × 10 mm × 2 mm.

The glass transition temperature (Tg) and the crystallization temperature (Tx) of the glass were characterized by differential thermal analysis (DTA) on powdered samples using a NetzschSTA9/C differential scanning calorimeter at a heating rate of 10 K/min. The X-ray diffraction (XRD) patterns were carried out on powdered samples using an X-ray diffraction (Bruker D2 PHASER) with Cu-Kα radiation and 2θ data were collected between 10° and 80° with a 0.1° step size. Transmission electron microscope (TEM) and high-resolution TEM (HRTEM) were used in order to study the microstructure of GCs by a transmission electron microscope (FEI TF20). The transmittance spectra were recorded on a spectrophotometer (Shimadzu UV-3600) in the range of 300-750 nm. The photoluminescence (PL) spectra, photoluminescence excitation (PLE) spectra and fluorescence decay were performed by a fluorescence spectrophotometer (Jobin-Yvon Fluorolog3) equipped with Xe lamp and a Spectra LED as external excitation light source. X-ray excitation spectra (XEL) were recorded by a spectrometer (Ocean Optical QE65000) with a X-ray tube (Copper target, 80 kV, 1.5 mA). All of the structural and optical measurements were performed at room temperature.

3. Results and discussions

The DTA curve of powdered 4 mol% Eu3+ doped oxyfluoride germanate glass sample is shown in Fig. 1. The glass transition temperature (Tg) is 528 °C and two exothermic peak with the value at 645 °C (Tx1) and 805 °C (Tx2) are observed. XRD analysis shows the first peak is due to the precipitation of LaF3 nanocrystal and the latter peak is due to the glass crystallization. The thermal stability factor [ΔT = (Tx-Tg)] of this glass is 117 °C, which indicates the oxyfluoride germanate glass has a good thermal property [1]. Based on these results, two temperatures, 620 °C and 640 °C, near the LaF3 crystallization temperature were chosen as the crystallization temperature to form transparent glass ceramics.

 figure: Fig. 1

Fig. 1 DTA curve of 4 mol% Eu3+ doped PG.

Download Full Size | PDF

The XRD patterns of PG, GC620°C-3h, GC640°C-3h and GC640°C-6h are displayed in Fig. 2(a). PG sample shows amorphous structure characterized by the broad hump without any crystalline diffraction peaks of its XRD pattern. On the contrary, after heating the PG at 620 °C and 640 °C for 3h or 6h, sharp diffraction peaks occur in GC samples, and can be easily assigned to the hexagonal LaF3 (JCPDS No.32-0483) without any other phase. The intensities of these peaks become stronger with the increasing of heat treatment temperature and heat treatment time, which indicates the gradual growth of nanocrystals. The mean size of LaF3 nanocrystal can be calculated from the peak widths by Scherrer equation [16,36]:

 figure: Fig. 2

Fig. 2 (a) XRD patterns; (b) partial enlargement of XRD patterns in the range of 27° to 29°; (c) experimental and calculated XRD patterns of the Rietveld refinement of GC640°C-6h; (d) Transmittance spectra; (e) Photographs of PG, GC620°C-3h, GC640°C-3h and GC640°C-6h (All samples are 2 mm thick).

Download Full Size | PDF

D=Kλβcosθ

Herein, D is the crystal size, K = 0.89 is a dimensionless shape factor, λ = 0.15406 nm is the wavelength of X-ray, β is the full-width at half maximum (FWHM) of the diffraction peak (derived from XRD pattern) and θ is the angle of diffraction peak, respectively. The calculated sizes of LaF3 nanocrystals are 24.4 nm, 29.3 nm and 32.1 nm for GC620°C-3h, GC640°C-3h and GC640°C-6h, respectively. The results demonstrate that the mean size of LaF3 nanocrystal can be easily tuned by changing the conditions of heat treatment. As shown in Fig. 2(b), the diffraction peak of (111) shifts to larger Bragg angle (other diffraction peaks have the same tendency to change), which is due to the incorporation of Eu3+ (with ionic radius 0.095 nm) into LaF3 nanocrystals by substituting La3+ (with ionic radius 1.032 nm).

To get the detailed crystal structure information on the obtained GC sample, the experimental, calculated, different XRD profiles and Bragg positions for the Rietveld refinement of GC640°C-6h are presented in Fig. 2(c). According to the refinement data, the GC640°C-6h crystallizes in a hexagonal LaF3 phase with a P-3c1 space group. Additionally, the lattice parameters were determined to be a = b = 7.19 Å, c = 7.35 Å, α = β = 90°, γ = 120°, V = 328.8 Å3, Z = 6, and the refinement finally converged to Rwp = 5.53%, Rp = 4.30% and GOF = 2.18. The results reveal that the actual structure agrees well with the initial structure.

Figure 2(d) shows the transmittance spectra of PG, GC620°C-3h, GC640°C-3h, and GC640°C-6h in the range of 300 nm to 750 nm. It is obviously noted that all the PG and GC samples have a transparency over 71% in the range of 450 nm to 700 nm. In order to display the transparency of the GCs intuitively, the digital photographs of PG and GC samples are shown in Fig. 2(e). Owing to the nanosized LaF3 crystal is much smaller than the wavelength of visible light, so the obtained GC samples still maintain a high transparency. In addition, the high transparency of GC samples indicates the successful prevention of the collapse of the GeO2 glassy network after heat treatment and demonstrates the promising prospect of optical application [9].

TEM bright field image and HRTEM image of GC640°C-6h sample are presented in Fig. 3(a) and (b). As shown in Fig. 3(a), nearly spherical nanoparticles (with dark appearance) were distributed among glass matrix. The LaF3 crystals scatter more electrons due to larger density than glass matrix and therefore appear dark [9]. The size of these nanocrystals as observed in TEM image is ~30 nm, which matches well with the values evaluated from XRD pattern by Scherrer equation. The small crystal size of near 30 nm avoids light scattering and result in the high transparency as show in Fig. 2(e). As verified by Fig. 3(b), the distance of lattice fringes is about 0.317 nm, which is in good agreement with the (111) plane in hexagonal LaF3 crystal. These results are the further evidences for the precipitation of hexagonal LaF3 in the glass matrix.

 figure: Fig. 3

Fig. 3 TEM image (a) and HRTEM image (b) of GC640°C-6h.

Download Full Size | PDF

Figure 4(a) depicts the emission spectra of precursor germanate glasses with different concentration of Eu3+ under the excitation wavelength of 393 nm. As shown in the figure, the emission spectra are composed of four dominant emissions with peaks at 582 nm, 595 nm, 616 nm, 654 nm and 701 nm in the range of 575 nm to 750 nm, which can be readily assigned to 5D07F0, 7F1, 7F2, 7F3 and 7F4 transitions of Eu3+ ion, respectively. The emission at 616 nm (5D07F2 transition) is the predominant and has the strongest intensity among them. Moreover, the emission intensity increases monotonously with the increasing of Eu3+ concentration until it reaches 4 mol%, then decreases due to the concentration quenching. The results indicate the optimum concentration of Eu3+ is 4 mol%. Thus, herein we choose the glass which is doped with 4 mol% Eu3+ for further heat treatment to obtain transparent GCs.

 figure: Fig. 4

Fig. 4 (a) Emission spectra of different Eu3+ doped PGs under excitation at 393 nm; (b) Emission spectra of 0.1% mol Eu3+ doped PG and GC620°C-3h under excitation at 393 nm.

Download Full Size | PDF

To reveal the surrounding structural environment of Eu3+ in germanate glass and GCs, the emission spectra of PG and GC doped with 0.1 mol% Eu3+ were recorded under the excitation wavelength of 393 nm, as exhibited in Fig. 4(b). Clearly, the emission spectra of the PG and the GC samples are quite different. The emission intensities of 5D1, 2, 3 to 7FJ (J = 0, 1, 2, 3, 4) transitions of Eu3+ in the GC sample are much stronger than that in the PG sample. For Eu3+ ions, the higher excited states 5D1, 2, 3 will be quenched through the multi-phonon relaxation owing to the high phonon energy of host matrix [31]. So, the reason for the difference in the emission spectra of PG and GC samples can be attributed to the incorporation of Eu3+ ions into LaF3 nanocrystals of lower phonon energy (~300 cm−1) after heat treatment.

The spectra of 5D0 level is also an effective tool for detecting the location of Eu3+ in glass ceramic [31]. Due to the intensity of the electric dipolar 5D07F2 is depended on the Eu3+ ligand field while the magnetic dipolar 5D07F1 is not, therefore the ratio (R) of the peak intensities of 5D07F2 and 5D07F1 transitions is an indicator of the structural environment change of Eu3+. Herein, the R of the PG sample is 2.79 while GC sample is 0.93. What is more, it can be seen clearly that Stark split occurs at 5D07F2 and 5D07F4 transition in GC sample. And a weak emission peak appears at 749 nm in the GC sample, which is due to the 5D07F5 of the Eu3+ ion [32,33]. All these results confirm that after the crystallization of PG, Eu3+ entered into the LaF3 nanocrystal phase.

Figure 5(a) shows the excitation spectra of 4 mol% Eu3+ doped PG and GC samples measured by monitoring 617 nm emission (5D07F2 transition of Eu3+). The excitation spectra of PG and GC samples consist of a number sharp peaks corresponding to 7F05D4 (361 nm), 7F05L7 (381 nm), 7F05L6 (393 nm), 7F05D3 (414 nm), 7F05D2 (463 nm), and 7F05D1 (531 nm) transitions of Eu3+ ions, respectively. In addition, the intensity of each excitation peak increases with the increasing of heating temperature and heating time, which is similar to the emission spectra.

 figure: Fig. 5

Fig. 5 (a) Excitation spectra monitored at 617 nm emission; (b) Emission spectra under excitation at 393nm; (c) Fluorescence decay curves for 5D07F0 (617 nm) transition under excitation at 393 nm; (d) XEL spectra of PG, GC620°C-3h, GC640°C-3h, GC640°C-6h and BGO.

Download Full Size | PDF

The emission spectra of 4 mol% Eu3+ doped PG and GC samples under the excitation wavelength of 393 nm light are displayed in Fig. 5(b). Compared with the PG sample, the intensities of emission peaks of GC samples are increased after crystallization. And the intensity ratios (R) of the peak intensities of 5D07F2 and 5D07F1 transitions is about 2.68, 1.19, 1.12 and 1.07 for PG, GC620°C-3h, GC640°C-3h and GC640°C-6h, respectively. Impressively, the emission of the 5D1, 2, 37FJ (J = 1,2,3,4,5) and 5D07F5 transition of Eu3+ disappeared in these GC samples which is due to the concentration quenching of Eu3+ in LaF3 nanocrystals [6]. However, despite reaching the quenching concentration, Stark split still occurs at the emission peaks of 617 nm (5D07F2) and 701 nm (5D07F4).

The fluorescence decay curves for 617 nm (5D07F2) of PG and GCs samples under the excitation wavelength of 393 nm are shown in Fig. 5(c). All the curves are well fitted to a double-exponential decay function [14],

I=A1exp(tτ1)+A2exp(tτ2)

where I stand for the intensity of luminescence, A1 and A2 are fitting constants, t is the time, τ1 and τ2 are fast and slow lifetimes components, respectively. According to the above parameters, the average lifetime (τ*) of Eu3+ can be evaluated by:

τ*=A1τ12+A2τ22A1τ1+A2τ2

According to Eq. (3), the average lifetime could be calculated to be 2.14, 2.59, 2.73 and 2.77 ms of PG, GC620°C-3h, GC640°C-3h and GC640°C-6h, respectively. It is obvious that the lifetime of GCs have a slight increase (from 2.14 to 2.77 ms) after crystallization of PG sample. In GC sample, Eu3+ entered into LaF3 crystalline phase with a low phonon energy environment, so reduce the non-radiative relaxation and enhance the lifetime.

With the aim of exploring the potential application of these PG and GC samples in X-ray detection, the XEL spectra of the prepared samples and BGO scintillating crystal were recorded under the excitation of X-ray (80 kV, 1.5 mA) and exhibited in Fig. 5(d). The emission bands with peaks centered at 589, 614, 651 and 700 nm can be readily assigned to 5D07F1, 7F2, 7F3, 7F4 transitions of Eu3+ ion, respectively. Compared with the emission spectra that under ultraviolet light (393 nm) excitation, the emission peaks of XEL spectra have a slight move to the short wavelength, which is attributed to the different excitation mechanism between ultraviolet and X-ray. Under X-ray excitation, the energy of X-ray has a direct interaction with electrons and holes in the matrix material, producing secondary electrons. Then, the energy of secondary electrons is transferred to the luminescent centers (Eu3+). Finally, the excited luminescent centers relax and emit the characteristic scintillating emission. On the contrary, Eu3+ were excited directly by ultraviolet light [34,35]. It is worthy to note that, the integrated XEL intensities (the integral of the emission peaks) of PG, GC620°C-3h, GC640°C-3h and GC640°C-6h samples in the range of 300 to 800 nm are 12%, 13%, 18% and 20% of that of the BGO crystal, respectively. It is clearly that the emission intensity is enhanced after crystallization. And the integrated XEL intensities are much higher than that of the reported in the literature [35].

4 Conclusions

In summary, Eu3+ doped transparent GC containing LaF3 nanocrystals have been successfully prepared by melt-quenching technique with subsequent heat treatment. The XRD and TEM results show that hexagonal LaF3 were precipitated in glass matrix. The transmittance spectra show that all the GC samples still remain a high transparency over 71% in Vis light region. Characteristic emission spectra of Eu3+ in GC show that Eu3+ were incorporated into LaF3 nanocrystals and the emission intensity enhanced after crystallization. Additionally, the integrated XEL intensity of GC640°C-6h is about 20% of that of the BGO crystal scintillator. The results indicate that Eu3+ doped transparent germanate GC containing LaF3 nanocrystals could be a promising scintillating material for X-ray detection.

Funding

Zhejiang Provincial Natural Science Foundation of China (Grant no. LY19E020004, LR15F050003); National Natural Science Foundation of China (Grant no. 51472225).

References

1. Y. Tian, R. Xu, L. Zhang, L. Hu, and J. Zhang, “1.8 µm mission of highly thulium doped fluorophosphate glasses,” J. Appl. Phys. 108(8), 083504 (2010). [CrossRef]  

2. M. Ding, D. Chen, Z. Wan, Y. Zhou, J. Zhong, J. Xi, and Z. Ji, “Achieving efficient Tb3+ dual-mode luminescence via Gd-sublattice-mediated energy migration in a NaGdF4 core–shell nanoarchitecture,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(21), 5372–5376 (2015). [CrossRef]  

3. D. Chen, W. Xu, Y. Zhou, J. Zhong, and S. Li, “Color tunable dual-phase transparent glass ceramics for warm white light-emitting diodes,” J. Mater. Chem. C Mater. Opt. Electron. Devices 5(3), 738–746 (2017). [CrossRef]  

4. R. Ye, G. Jia, D. Deng, Y. Hua, Z. Cui, S. Zhao, L. Huang, H. Wang, C. Li, and S. Xu, “Controllable Synthesis and Tunable Colors of α- and β-Zn2SiO4:Mn2+ Nanocrystals for UV and Blue Chip Excited White LEDs,” J. Phys. Chem. C 115(21), 10851–10858 (2011). [CrossRef]  

5. D. Chen, Z. Wan, and Y. Zhou, “Dual-phase nano-glass-ceramics for optical thermometry,” Sens. Actuators B Chem. 226, 14–23 (2016). [CrossRef]  

6. X. Li, X. Chen, S. Yuan, S. Liu, C. Wang, and D. Chen, “Eu3+-Doped glass ceramics containing NaTbF4 nanocrystals: controllable glass crystallization, Tb3+-bridged energy transfer and tunable luminescence,” J. Mater. Chem. C Mater. Opt. Electron. Devices 5(39), 10201–10210 (2017). [CrossRef]  

7. J. He, Y. Wang, Y. Liu, K. Wang, R. Li, J. Fan, S. Xu, and L. Zhang, “Tailoring the Luminescence of Europium Ions in Mesoporous AlPO4 Monolithic Glass,” J. Phys. Chem. C 117(42), 21916–21922 (2013). [CrossRef]  

8. M. A. Ali, X. Liu, and J. Qiu, “A comparative investigation on upconversion luminescence in glass–ceramics containing LaF3 and CaF2 nanocrystals,” J. Mater. Sci. Mater. Electron. 29(10), 8701–8709 (2018). [CrossRef]  

9. C. Lin, C. Bocker, and C. Rüssel, “Nanocrystallization in Oxyfluoride Glasses Controlled by Amorphous Phase Separation,” Nano Lett. 15(10), 6764–6769 (2015). [CrossRef]   [PubMed]  

10. L. Huang, T. Yamashita, R. Jose, Y. Arai, T. Suzuki, and Y. Ohishi, “Intense ultraviolet emission from Tb3+ and Yb3+ codoped glass ceramic containing CaF2 nanocrystals,” Appl. Phys. Lett. 90(13), 131116 (2007). [CrossRef]  

11. L. Huang, S. Jia, Y. Li, S. Zhao, D. Deng, H. Wang, G. Jia, Y. Hua, and S. Xu, “Enhanced emissions in Tb3+-doped oxyfluoride scintillating glass ceramics containing BaF2 nanocrystals,” Nucl. Instrum. Methods. Phys. Res. Sect. A 788, 111–115 (2015). [CrossRef]  

12. L. Q. Xvsheng, F. Xianping, and W. Minquam, “Local vibration around rare earth ions in alkaline earth fluorosilicate transparent glass and glass ceramics using Eu3+ probe,” J. Rare Earths 26(6), 883–888 (2008). [CrossRef]  

13. C. Bocker, S. Bhattacharyya, T. Höche, and C. Rüssel, “Size distribution of BaF2 nanocrystallites in transparent glass ceramics,” Acta Mater. 57(20), 5956–5963 (2009). [CrossRef]  

14. Q. Luo, X. Qiao, X. Fan, and X. Zhang, “Luminescence Properties of Eu2+-Doped Glass Ceramics Containing SrF2 Nanocrystals,” J. Am. Ceram. Soc. 93(9), 2684–2688 (2010). [CrossRef]  

15. Q. Luo, X. Qiao, X. Fan, S. Liu, H. Yang, and X. Zhang, “Reduction and luminescence of europium ions in glass ceramics containing SrF2 nanocrystals,” J. Non-Cryst. Solids 354(40-41), 4691–4694 (2008). [CrossRef]  

16. L. A. Bueno, A. S. Gouveia-Neto, E. B. da Costa, Y. Messaddeq, and S. J. L. Ribeiro, “Structural and spectroscopic study of oxyfluoride glasses and glass-ceramics using europium ion as a structural probe,” J. Phys. Condens. Matter 20(14), 145201 (2008). [CrossRef]  

17. Y. W. Daqin Chen, Y. Yu, and P. Huang, “Structure and Optical Spectroscopy of Eu-Doped Glass Ceramics Containing GdF3 Nanocrystals,” J. Phys. Chem. C 112(48), 18943–18947 (2008). [CrossRef]  

18. Y. Y. Daqin Chen, Y. Wang, P. Huang, and F. Weng, “Cooperative Energy Transfer Up-Conversion and Quantum Cutting Down-Conversion in Yb3+:TbF3 Nanocrystals Embedded Glass Ceramics,” J. Phys. Chem. C 113(16), 6406–6410 (2009). [CrossRef]  

19. D. Chen, S. Liu, X. Li, S. Yuan, and P. Huang, “Upconverting luminescence based dual-modal temperature sensing for Yb3+/Er3+/Tm3+:YF3 nanocrystals embedded glass ceramic,” J. Eur. Ceram. Soc. 37(15), 4939–4945 (2017). [CrossRef]  

20. J. Cao, X. Wang, X. Li, Y. Wei, L. Chen, and H. Guo, “Enhanced emissions in Tb3+-doped oxyfluoride scintillating glass ceramics containing KLu2F7 nano-crystals,” J. Lumin. 170, 207–211 (2016). [CrossRef]  

21. M. B. Beckert, S. Gallego, Y. Ding, E. Elder, and J. H. Nadler, “Medical imaging scintillators from glass-ceramics using mixed rare-earth halides,” Opt. Mater. 60, 513–520 (2016). [CrossRef]  

22. X. Y. Sun, Z. P. Ye, Y. T. Wu, P. Gao, R. H. Mao, Z. J. Zhang, J. T. Zhao, and G. Villalobos, “A Simple and Highly Efficient Method for Synthesis of Ce3+-Activated Borogermanate Scintillating Glasses in Air,” J. Am. Ceram. Soc. 97(11), 3388–3391 (2014). [CrossRef]  

23. J. Cao, W. Chen, D. Xu, X. Li, R. Wei, L. Chen, X. Sun, and H. Guo, “Transparent glass ceramics containing Lu6O5F8:Tb3+ nano-crystals: Enhanced photoluminescence and X-ray excited luminescence,” J. Am. Ceram. Soc. 101(4), 1585–1591 (2018). [CrossRef]  

24. X. Y. Sun, Y. Yuan, Z. H. Xiao, X. G. Yu, W. F. Wang, D. G. Jiang, G. T. Zheng, Z. Kang, and S. Kasap, “Optical Investigation of Ce3+-Activated Borogermanate Glass Induced by Substitution of BaF2 for BaO,” J. Am. Ceram. Soc. 98(12), 3655–3658 (2015). [CrossRef]  

25. J. Fu, M. Kobayashi, S. Sugimoto, and J. M. Parker, “Scintillation from Eu2+in Nanocrystallized Glass,” J. Am. Ceram. Soc. 92(9), 2119–2121 (2009). [CrossRef]  

26. J. Cao, L. Chen, W. Chen, D. Xu, X. Sun, and H. Guo, “Enhanced emissions in self-crystallized oxyfluoride scintillating glass ceramics containing KTb2F7 nanocrystals,” Opt. Mater. Express 6(7), 2201 (2016). [CrossRef]  

27. J. Cao, W. Chen, L. Chen, X. Sun, and H. Guo, “Synthesis and characterization of BaLuF5:Tb3+ oxyfluoride glass ceramics as nanocomposite scintillator for X-ray imaging,” Ceram. Int. 42(15), 17834–17838 (2016). [CrossRef]  

28. G. Lee, N. Savage, B. Wagner, Y. Zhang, B. Jacobs, H. Menkara, C. Summers, and Z. Kang, “Synthesis and Luminescence Properties of Transparent Nanocrystalline GdF3:Tb Glass-Ceramic Scintillator,” J. Lumin. 147, 363–366 (2014). [CrossRef]   [PubMed]  

29. B. Zhu, S. Zhang, S. Zhou, N. Jiang, and J. Qiu, “Enhanced upconversion luminescence of transparent Eu3+-doped glass-ceramics containing nonlinear optical microcrystals,” Opt. Lett. 32(6), 653–655 (2007). [CrossRef]   [PubMed]  

30. J. Fu, M. Kobayashi, S. Sugimoto, and J. M. Parker, “Eu3+-activated heavy scintillating glasses,” Mater. Res. Bull. 43(6), 1502–1508 (2008). [CrossRef]  

31. D. Chen, Y. Yu, P. Huang, H. Lin, Z. Shan, and Y. Wang, “Color-tunable luminescence of Eu3+ in LaF3 embedded nanocomposite for light emitting diode,” Acta Mater. 58(8), 3035–3041 (2010). [CrossRef]  

32. Y. Dwivedi and S. B. Rai, “Optical properties of Eu3+ in oxyfluoroborate glass and its nanocrystalline glass,” Opt. Mater. 31(1), 87–93 (2008). [CrossRef]  

33. J. E. Muñoz-Santiuste, U. R. Rodríguez-Mendoza, J. González-Platas, and V. Lavín, “Structural study of the Eu3+ environments in fluorozirconate glasses: role of the temperature-induced and the pressure-induced phase transition processes in the development of a rare earth’s local structure model,” J. Chem. Phys. 130(15), 154501 (2009). [CrossRef]   [PubMed]  

34. X. Y. Sun, D. G. Jiang, Y. Z. Sun, X. Zhang, Q. L. Hu, Y. Huang, and Y. Tao, “Eu3+-activated B2O3–GeO2–RE2O3 (RE=Y3+, La3+ and Gd3+) borogermanate scintillating glasses,” J. Non-Cryst. Solids 389, 72–77 (2014). [CrossRef]  

35. X. Y. Sun, X. Zhang, H. H. Chen, Q. L. Hu, W. F. Wang, Z. J. Zhang, and J. T. Zhao, “Investigation on the luminescent properties of Eu3+-activated dense oxyfluoride borogermanate scintillating glasses,” J. Non-Cryst. Solids 404, 162–166 (2014). [CrossRef]  

36. U. Holzwarth and N. Gibson, “The Scherrer equation versus the ‘Debye-Scherrer equation’,” Nat. Nanotechnol. 6(9), 534 (2011). [CrossRef]   [PubMed]  

Cited By

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

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 DTA curve of 4 mol% Eu3+ doped PG.
Fig. 2
Fig. 2 (a) XRD patterns; (b) partial enlargement of XRD patterns in the range of 27° to 29°; (c) experimental and calculated XRD patterns of the Rietveld refinement of GC640°C-6h; (d) Transmittance spectra; (e) Photographs of PG, GC620°C-3h, GC640°C-3h and GC640°C-6h (All samples are 2 mm thick).
Fig. 3
Fig. 3 TEM image (a) and HRTEM image (b) of GC640°C-6h.
Fig. 4
Fig. 4 (a) Emission spectra of different Eu3+ doped PGs under excitation at 393 nm; (b) Emission spectra of 0.1% mol Eu3+ doped PG and GC620°C-3h under excitation at 393 nm.
Fig. 5
Fig. 5 (a) Excitation spectra monitored at 617 nm emission; (b) Emission spectra under excitation at 393nm; (c) Fluorescence decay curves for 5D07F0 (617 nm) transition under excitation at 393 nm; (d) XEL spectra of PG, GC620°C-3h, GC640°C-3h, GC640°C-6h and BGO.

Equations (3)

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

D= Kλ βcosθ
I= A 1 exp( t τ 1 )+ A 2 exp( t τ 2 )
τ * = A 1 τ 1 2 + A 2 τ 2 2 A 1 τ 1 + A 2 τ 2
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.