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

Enhanced emissions in self-crystallized oxyfluoride scintillating glass ceramics containing KTb2F7 nanocrystals

Open Access Open Access

Abstract

Self-crystallized KTb2F7 oxy-fluoride glass ceramics (GC) were successfully manufactured via the traditional melt-quenching method. KTb2F7 nanocrystals were already formed after melt-quenching, which is beneficial to the realization of controllable glass crystallization to some degree for affording desirable nano-crystal size and activator partition. Their microstructural and optical properties were systemically investigated by X-ray diffraction (XRD), transmission electron microscopy (TEM), high-resolution TEM (HRTEM), absorption spectra, photoluminescence (PL), luminescence lifetime measurements and X-ray excited luminescence (XEL). Both PL and XEL of GC samples are highly enhanced because more nanocrystals formed and grew up after heat-treatment. Our investigation suggests that transparent KTb2F7 glass ceramics may present potential application in X-ray scintillator for X-ray imaging. And our strategy that takes active ions as host may contribute to designing other oxy-fluoride GC by using active ions as host.

© 2016 Optical Society of America

1. Introduction

In the past decades, scintillating materials have been drawing spectacular attention on account of their potential applications in the field of high energy engineering, industrial and medical imaging [1–3]. Owing to the much concern of nuclear security recently [4], great interest has been sparked in alternative scintillating materials systems ranging from new mono-crystal compositions to novel composite systems such as phosphor-based polymer materials, transparent ceramics and scintillating glasses [4,5]. However, there are several restrictions for the applications of these scintillating materials.

As for mono-crystals, its growth via Czochralski or Bridgman techniques is extremely complex and time-consuming, and always leads to the restricted cylindrical geometries [6]. Moreover, the techniques are only applicative for the materials that work well in such techniques. As for phosphor-based polymer materials, at first, the easily agglomerated phosphor particles encapsulated in the polymer matrix causes strong light scattering, which limits spatial resolution [7,8]. At second, the aged polymer matrix and mismatch of refractive indices between phosphor particles and polymer matrix reduce transparency, which also limits spatial resolution and hinders luminescent efficiency. As for transparent ceramics, there are still lots of challenges such as complicated and time-consuming procedures, limited compositions, tendency to scatter light, segregation of doping agents, etc [9]. The intensively explored scintillating glasses doped with rare earth ions are promising alternatives to single crystal and ceramic scintillators for practical application in high-energy physics, industrial and medical imaging fields owing to the advantages of transparency, low-cost, large-volume production and easy shaping of elements [2,7]. However, they suffer from low light yield resulting from high phonon energy, hindering their performance for practical applications [10,11].

Transparent oxyfluoride glass ceramics (GC), one of the most promising luminescent materials, are characterized by the homogeneous precipitation of fluoride nanocrystals in the oxide glass matrix [12–14]. When rare earth (RE) ions are introduced to the precursor glasses, these ions are preferentially enrich into the crystallized nanocrystals with low phonon energy after thermal treatment [13]. The low phonon energy of these precipitated nanocrystals can reduce the multi-phonon non-radiative relaxation rate, and enormously enhance the emission efficiency of RE ions [15]. In this case, the excellent optical properties of RE ions in nanocrystals can be utilized in glass matrix, with the advantages of glass preserved, such as good forming ability, high chemical, photochemical and mechanical stability [16–18]. However, to our best knowledge, reports concerning on oxyfluoride GC as potential scintillating materials are rare [19,20]. Hence, it is of paramount importance to search for oxyfluoride GC used as potential alternative scintillating materials.

Such GC can be achieved through controlled ucleation and crystallization from precursor glasses by appropriate heat-treatment procedures [21]. If the nanocrystals are formed, even the active ions have been participated into formed nanocrystals before thermal treatment, it is more convenient and energy-saving to acquire GC. Recently, Chen et al reported the fabrication and tunable upconversion luminescence in self-crystallized Er3+:K(Y1-xYbx)3F10 [21] and Na(Lu1-xYbx)F4 GC [13]. In such system, nanocrystals are already formed via self-crystallization during melt-quenching, which facilitates the incorporation of more RE ions into the precipitated nanocrystals through further heat-treatment at relatively low temperature and prevents overgrowth of nanocrystals. That is, enabling high light yield and ensuring high transparency of derived GC at the same time.

As we all know, the green emission of Tb3+ matches the spectral sensitivity of conventional Charge Coupled Device (CCD) detectors very well so as to be widely applied in scintillating materials. But it is usually difficult for Tb3+ ions to participate into formed nanocrystals in GC compared with other RE ions such as Er3+ and Ho3+, making the enhancement of emissions in Tb3+-doped GC neglectable [18,22]. To overcome this problem, herein, self-crystallized KTb2F7 oxy-fluoride GC were successfully fabricated and systematically investigated. KTb2F7 nanocrystals were already formed before heat-treatment. By taking Tb3+ as host in KTb2F7 GC, greatly enhanced emissions for both PL and XEL are observed, resulting from the crystallization of KTb2F7 nanocrystals with low phonon energy. Our results indicate that KTb2F7 GC present potential applications in X-ray imaging such as industrial radiography and X-ray intensifying screen considering the good transparency, enhanced light yield of Tb3+ in KTb2F7 glass ceramics.

2. Experimental

Glass samples with nominal composition of 53SiO2-10K2CO3-15Al2O3-18KF-4TbF3 (in mol %) were prepared by melt-quenching method in air atmosphere. SiO2, Al2O3, K2CO3, KF (A.R., all from Sinopharm Chemical Reagent Co., Ltd., China) and high purity EuF3, TbF3 (99.99%, from AnSheng Inorganic Materials Co., Ltd., China) were used as starting materials. The well ground stoichiometric chemicals were put into a covered alumina crucible and melted at 1550 °C for 1 h in air atmosphere. The melt was poured onto a stainless-steel plate and then pressed by another plate to form precursor glasses (labeled as PG). Subsequently, PG glasses were heat-treated for 2 h at 660 and 680 °C, which were labeled as GC660 and GC680, respectively. All samples were polished optically for further characterization.

To confirm the crystallization of all samples, X-ray diffraction (XRD) patterns were performed on a Philips X’Pert PRO SUPER X-ray diffraction apparatus with CuKa radiation (λ = 0.154056 nm) over the angular range 10°≤ ≤80° in a step size of 0.0167°. Transmittance spectra were measured on a U-3900 ultraviolet-visible (UV-VIS) spectrophotometer (Hitachi). The microstructure of GC was analyzed by a JEM-2010 transmission electron microscopy (TEM) (JOEL Ltd., Tokyo, Japan). Excitation, emission spectra and lifetime measurement were recorded on an Edinburgh FLS920 spectrofluorometer equipped with a continuous wave 450 W Xe lamp and a microsecond flashlamp (μF900) as excitation sources. X-ray tube operated under 50 kV and 3 mA was used as the source for the measurement of XEL spectra on FLS920. All the measurements were performed at room temperature.

3. Results and discussions

Figure 1(a) presents the XRD patterns of PG and GC samples. The distinct diffraction peaks in the XRD pattern of PG directly confirm the formation of nanocrystals before heat-treatment. With further heat-treatment, the characteristic diffraction peaks corresponding to monoclinic KTb2F7 nanocrystals (JCPDS No. 32-0849) become increasingly obvious, which indicates the growth of KTb2F7 nanocrystals after heat-treatment. The mean crystalline size D of KTb2F7 nanocrystals can be evaluated by Scherer’s equation [16],

D=kλ/βcosθ
where k = 0.89, θ is the Bragg angle of the X-ray diffraction peak, β represents the corrected half width of diffraction peak andλ( = 0.154056 nm) is the wavelength of Cu Kα radiation. The mean crystalline size is estimated to be 26, 28 and 33 nm for PG, GC660 and GC680, respectively. The manufacture of KTb2F7 GC via self-crystallization favors the incorporation of active ions into the precipitated nanocrystals through further heat-treatment at relatively low temperature and prevents overgrowth of nanocrystals, which enables high light yield and ensures high transparency of derived GC at the same time [11].

 figure: Fig. 1

Fig. 1 (a) XRD patterns of PG, GC660 and GC680, and the reference data of JCPDS card No. 32-0849 for KTb2F7; (b) Transmittance spectra; (c) TEM image of PG, the inset is corresponding SAED patterns; (d) HRTEM image of PG. (e) TEM image of GC680, the inset is corresponding SAED patterns; (f) HRTEM image of GC680.

Download Full Size | PDF

As described in the UV-VIS transmittance spectra of PG and GC given in Fig. 1(b), all samples maintain a fair transmittance up to 70% in VIS region owing to the much smaller size of precipitated KTb2F7 nanocrystals compared to the wavelength of VIS light, which will be favorable for their practical applications [15]. The absorption bands at 317, 341, 351, 368, 376 and 484 nm are attributed to the transitions of Tb3+ from 7F6 ground state to 5H7, 5L7, 5L9, 5D2, 5D3 and 5D4 excited states, respectively [20].

To probe the microstructures of PG and GC samples, TEM and high-resolution TEM (HRTEM) images of PG and GC680 sample were performed (displayed in Fig. 1(c)-1(f)). TEM bright-field micrographs for both PG and GC680 clearly show that uniform spherical nanocrystals are lying on the gray background corresponding to the glass matrix. Both the size of nanocrystals and degree of crystallinity for GC increased compared to that of PG sample after proper heat-treatment, indicating the growth of KTb2F7 nanocrystals in GC. The mean size of KTb2F7 nanocrystals is about 30 and 36 nm in diameter for PG and GC680, respectively, which are similiar to the size estimated by Scherer’s equation. The polycrystalline diffraction feature of the precipitated nanocrystals is confirmed by the corresponding selected area electron diffraction (SAED) patterns (inset of Fig. 1(c) and Fig. 1(e)). The HRTEM images in Fig. 1(d) and 1(f) for PG and GC680 clearly reveal well-defined lattice structure and the relevant interplanar space d value measured by two separated crystal planes is about 0.330 and 0.586 nm for PG and GC680, which could be attributed to (111) and (101) crystal plane of KTb2F7 (d(111) = 0.328 nm and d(101) = 0.590 nm), respectively.

The excitation and emission spectra of PG, GC660 and GC680 are depicted in Fig. 2(a) and 2(b) respectively. By monitoring the 5D47F5 transition of Tb3+ emission at 543 nm, a series of characteristic excitation peaks centered at 283, 302, 317, 340, 351, 368, 376 and 484 nm are observed and assigned to 7F65I8, 5H6, 5H7, 5L7, 5L9, 5D2, 5D3 and 5D4 transitions of Tb3+, respectively [14]. With increasing the temperature of thermal treatment, remarkable enhancement of the excitation intensity occurred for GC660 and GC680 compared to that of PG.

 figure: Fig. 2

Fig. 2 (a) Excitation spectra (λem = 543 nm) and (b) emission spectra (λex = 376 nm) of Tb3+-doped PG, GC660 and GC680 samples. (c) Decay curves for 5D47F5 transition (543 nm) of Tb3+ in PG, GC660 and GC680 (λex = 376 nm). (d) XEL spectra of KTb2F7 PG,GC660 and GC680 samples under the excitation of X-ray (50 kV, 3 mA), the inset is corresponding enhanced factor of PL spectra and XEL spectra of GC660 and GC680 samples compared with PG sample.

Download Full Size | PDF

As displayed in Fig. 2(b), under the excitation of 376 nm (7F65D3), four characteristic emission bands of Tb3+ are observed in PG, GC660 and GC680. The emission bands centered at 488, 542, 583 and 621 nm are assigned to 5D47FJ (J = 6, 5, 4 and 3) of Tb3+, respectively [12]. A great enhancement of the emission intensity is also observed in emission spectra after heat-treatment and the integrated emission intensity is enhanced by 1.44 and 2.75 times for GC660 and GC680, respectively. More KTb2F7 nanocrystals are formed after further heat-treatment, leading to further decrease of phonon energy and reduction of structure defects [11]. As a result, the non-radiative transition rate is further reduced and the emission intensity is enhanced [23].

Figure 2(c) exhibits the decay curves of 543 nm emission (5D47F5) of Tb3+ in PG, GC660 and GC680 excited by 376 nm light. The average fluorescence lifetime τ can be evaluated by the following equation [16],

τ=I(t)tdt/I(t)dt
where the I(t) represents the fluorescence intensityIat time t. The obtained lifetimes of 5D4 level of Tb3+ in PG, GC660 and GC680 are 3.28, 2.33 and 2.32 ms, respectively. The gradually shortened lifetimes are contradicted to early reported results in RE-doped GC [19–22]. Usually, the formation of fluoride nanocrystals with low phonon energy will result in prolonged lifetimes of RE ions [19,23]. Nevertheless, in such self-crystallized KTb2F7 GC, taking Tb3+ as host will predominantly result in shorter Tb3+-Tb3+ distance especially with the increase of thermal temperature, which leads to decreased lifetime [11].

In order to investigate the potential application of such KTb2F7 GC in X-ray scintillator, the XEL spectra of PG and GC samples excited by X-ray (50 kV, 3 mA) were performed, as shown in Fig. 2(d). Under the excitation of X-ray, four intense emission bands centered at 489, 543, 587, and 620 nm are observed and they can be ascribed to 5D47FJ (J = 6, 5, 4 and 3) of Tb3+, respectively [14]. Compared to PG, the integrated XEL emission intensity is enhanced by 1.36 and 2.15 times for GC660 and GC680, respectively. Similarly, the astonishing enhancement in the XEL spectra is also caused by the crystallization of KTb2F7 nanocrystals with low phonon energy, which reduces the probability of non-radiative transition [19].

As exhibited in the inset of Fig. 2(d), the enhanced factor for XEL and PL are different. This is because the interaction mechanism of ultraviolet light excitation differs remarkably from that of X-ray excitation [11]. Under ultraviolet excitation, energy is deposited by direct excitation into the Tb3+ ions [10]. However, scintillating process is more complicated. Usually, three phases are involved in one scintillating process: (1) Energy conversion: initial energy release with formation of “hot” electrons and holes. (2) Transfer to luminescent centers: formation of excitonic states and groups of excited luminescent centers. (3) Light emission: relaxation of excited luminescent centers and emission of scintillating light [7,24].

4. Conclusions

Transparent self-crystallized KTb2F7 GC were successfully manufactured. The average size of KTb2F7 nanocrystals was estimated to be around 36 nm. By taking Tb3+ as host, both PL and XEL of KTb2F7 GC are highly enhanced after further heat-treatment due to the formation of KTb2F7 nanocrystals with low phonon energy, which suggests that such KTb2F7 GC present potential application in X-ray scintillator for X-ray imaging and our strategy that takes active ions as host may contributes to designing other oxy-fluoride GC using active ions as host.

Acknowledgments

This work was supported by National Natural Science Foundation of China (Nos. 11374269 and 11165010), the Natural Science Foundation of Jiangxi Province (20152ACB21017), Zhejiang Provincial Science and Technology Key Innovation Team (No. 2011R50012) and Zhejiang Provincial Key Laboratory (No. 2013E10022).

References and links

1. M. Nikl, J. A. Mares, E. Mihokova, K. Nitsch, N. Solovieva, V. Babin, A. Krasnikov, S. Zazubovich, M. Martini, A. Vedda, P. Fabeni, G. P. Pazzi, and S. Baccaro, “Radio- and thermoluminescence and energy transfer processes in Ce3+(Tb3+)-doped phosphate scintillating glasses,” Radiat. Meas. 33(5), 593–596 (2001). [CrossRef]  

2. X. Y. Sun, Z. P. Ye, Y. T. Wu, Z. H. Xiao, P. Gao, R. H. Mao, Z. J. Zhang, and J. T. Zhao, “Role of minor quantity of Si3N4 addition on the optical properties of Ce3+-activated borogermanate scintillating glass,” Opt. Mater. Express 5(6), 1381–1388 (2015). [CrossRef]  

3. I. H. Campbell and B. K. Crone, “Quantum-dot/organic semiconductor composites for radiation detection,” Adv. Mater. 18(1), 77–79 (2006). [CrossRef]  

4. B. D. Milbrath, A. J. Peurrung, M. Bliss, and W. J. Weber, “Radiation detector materials: An overview,” J. Mater. Res. 23(10), 2561–2581 (2008). [CrossRef]  

5. G. V. M. Williams, A. Bittar, C. Dotzler, A. Beaudin, C. Varoy, and C. Dunford, “Glass-ceramics and epoxy-composites for radiation imaging,” Radiat. Meas. 42(4-5), 899–902 (2007). [CrossRef]  

6. U. W. Arndt, “X-ray television area detectors,” Nucl. Instrum. Methods Phys. Res. 201(1), 13–20 (1982). [CrossRef]  

7. X. Y. Sun, Q. M. Yang, P. Gao, H. S. Wu, and P. Xie, “Luminescence, energy transfer properties of Tb3+/Gd3+-coactivated oxyfluoride borogermanate scintillating glasses,” J. Lumin. 165, 40–45 (2015). [CrossRef]  

8. 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. A 788, 111–115 (2015). [CrossRef]  

9. S. F. Wang, J. Zhang, D. W. Luo, F. Gu, D. Y. Tang, Z. L. Dong, G. E. B. Tan, W. X. Que, T. S. Zhang, S. Li, and L. B. Kong, “Transparent ceramics: Processing, materials and applications,” Prog. Solid State Chem. 41(1-2), 20–54 (2013). [CrossRef]  

10. C. Zuo, Z. Zhou, L. Zhu, A. Xiao, Y. Chen, X. Zhang, Y. Zhuang, X. Li, and Q. Ge, “Luminescence properties of Tb3+-doped borosilicate scintillating glass under UV excitation,” Spectrochim. Acta A Mol. Biomol. Spectrosc. 147, 324–327 (2015). [CrossRef]   [PubMed]  

11. J. Cao, X. Li, Z. Wang, Y. Wei, L. Chen, and H. Guo, “Optical thermometry based on up-conversion luminescence behavior of self-crystallized K3YF6:Er3+ glass ceramics,” Sens. Actuators B Chem. 224, 507–513 (2016). [CrossRef]  

12. C. Lin, C. Liu, Z. Zhao, L. Li, C. Bocker, and C. Rüssel, “Broadband near-IR emission from cubic perovskite KZnF3:Ni2+ nanocrystals embedded glass-ceramics,” Opt. Lett. 40(22), 5263–5266 (2015). [CrossRef]   [PubMed]  

13. G. Gao and L. Wondraczek, “Near-infrared downconversion in Pr3+/Yb3+ co-doped boro-aluminosilicate glasses and LaBO3 glass ceramics,” Opt. Mater. Express 3(5), 633–644 (2013). [CrossRef]  

14. P. Li, M. Peng, L. Wondraczek, Y. Zhao, and B. Viana, “Red to near infrared ultralong lasting luminescence from Mn2+-doped sodium gallium aluminum germanate glasses and (Al,Ga)-albite glass-ceramics,” J. Mater. Chem. C Mater. Opt. Electron. Devices 3(14), 3406–3415 (2015). [CrossRef]  

15. G. Wu, S. Fan, Y. Zhang, G. Chai, Z. Ma, M. Peng, J. Qiu, and G. Dong, “2.7 μm emission in Er3+:CaF2 nanocrystals embedded oxyfluoride glass ceramics,” Opt. Lett. 38(16), 3071–3074 (2013). [CrossRef]   [PubMed]  

16. X. Liu, Y. Wei, R. Wei, J. Yang, and H. Guo, “Elaboration, Structure, and Luminescence of Eu3+-Doped BaLuF5-Based Transparent Glass-Ceramics,” J. Am. Ceram. Soc. 96(3), 798–800 (2013). [CrossRef]  

17. M. Brooke Barta, J. H. Nadler, Z. Kang, B. K. Wagner, R. Rosson, and B. Kahn, “Effect of host glass matrix on structural and optical behavior of glass–ceramic nanocomposite scintillators,” Opt. Mater. 36(2), 287–293 (2013). [CrossRef]  

18. 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]  

19. 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]  

20. X. Sun and S. Huang, “Tb3+-activated SiO2–Al2O3–CaO–CaF2 oxyfluoride scintillating glass ceramics,” Nucl. Instrum. Methods Phys. Res. A 621(1-3), 322–325 (2010). [CrossRef]  

21. D. Chen, Z. Wan, and Y. Zhou, “Optical spectroscopy of Cr³⁺-doped transparent nano-glass ceramics for lifetime-based temperature sensing,” Opt. Lett. 40(15), 3607–3610 (2015). [CrossRef]   [PubMed]  

22. S. N. Zhang, J. H. Huang, Y. J. Chen, X. H. Gong, Y. F. Lin, Z. D. Luo, and Y. D. Huang, “Site-selective excitation and emission of Eu3+-doped transparent glass ceramic containing Ca5(PO4)3F nanocrystals,” Opt. Mater. Express 3(6), 868–874 (2013). [CrossRef]  

23. Y. L. Wei, X. M. Li, and H. Guo, “Enhanced upconversion in novel KLu2F7:Er3+ transparent oxyfluoride glass-ceramics,” Opt. Mater. Express 4(7), 1367–1372 (2014). [CrossRef]  

24. P. Lecoq, A. Anenkow, A. Gektin, M. Korzhik, and C. Pedrini, Inorganic Scintillators for Detector Systems (Springer Verlag Berlin Heidelberg, 2006), Chap. 1.

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 (2)

Fig. 1
Fig. 1 (a) XRD patterns of PG, GC660 and GC680, and the reference data of JCPDS card No. 32-0849 for KTb2F7; (b) Transmittance spectra; (c) TEM image of PG, the inset is corresponding SAED patterns; (d) HRTEM image of PG. (e) TEM image of GC680, the inset is corresponding SAED patterns; (f) HRTEM image of GC680.
Fig. 2
Fig. 2 (a) Excitation spectra (λem = 543 nm) and (b) emission spectra (λex = 376 nm) of Tb3+-doped PG, GC660 and GC680 samples. (c) Decay curves for 5D47F5 transition (543 nm) of Tb3+ in PG, GC660 and GC680 (λex = 376 nm). (d) XEL spectra of KTb2F7 PG,GC660 and GC680 samples under the excitation of X-ray (50 kV, 3 mA), the inset is corresponding enhanced factor of PL spectra and XEL spectra of GC660 and GC680 samples compared with PG sample.

Equations (2)

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

D = k λ / β cos θ
τ = I ( t ) t d t / I ( t ) d t
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.