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

Trap-controlled mechanoluminescence in Pr3+-activated M2Nb2O7 (M = Sr, Ca) isomorphic perovskites

Open Access Open Access

Abstract

Isomorphic materials tend to exhibit differentiated performance due to the small differences between the constituent atoms and the crystal structures. Inspired by the isomorphic strategy, herein we developed a novel trap-controlled mechanoluminescent (TC-ML) material, Pr3+-activated Sr2Nb2O7, that shows mechanoluminescence (ML) under the mechanical actions of compression and friction. Compared to the isomorphic Ca2Nb2O7:Pr3+ which requires a long-time delay to achieve high-contrast ML due to the strong long-lasting afterglow, Sr2Nb2O7:Pr3+ enables the obtaining of high-contrast ML in a short-time delay because of the weak short-lasting afterglow, which greatly simplifies the imaging process to facilitate the further application of ML. The investigations of structural and optical characteristics between Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ reveal that the unique ML of Sr2Nb2O7:Pr3+ should result from its lower trap concentration and shallower trap depth relative to Ca2Nb2O7:Pr3+. Our results are expected to deepen our understanding on the behavioral diversity of isomorphic phosphors and thereby broaden the horizon of designing and regulating TC-ML materials.

© 2018 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Mechanical-responsive luminescent materials have attracted special attention as sensing system owing to the remote observability of light emission [1–7]. One of the typical interests is trap-controlled mechanoluminescence (TC-ML), which is a form of reproducible, quantitative mechano-optical conversion resulting from trap-mediated storage and release of excitation energy [8–10]. This kind of reproducible ML enables ones to directly observe the stress concentration of load-bearing objects through imaging the distribution of ML, showing broad prospects in fields as diverse as stress sensors, structural healthcare, bio-imaging and lighting displays [11–17]. The first report of the robust TC-ML dates back to 1999 when Chao-Nan Xu and associates discovered the intense, reproducible elasticoluminescence from ZnS:Mn2+ and SrAl2O4:Eu2+ [18,19]. Xu’s studies are driving a renewed interest in ML research, and are creating opportunities to develop new and improved TC-ML materials to solve real-world problems. It is known that TC-ML materials developed during the past 20-years are universally activated by the doped luminescent ions, such as ZnS:Mn2+/Cu/Cu,Mn2+ [18,20], SrAl2O4:Eu2+/Ce3+/Eu2+,Dy3+/Eu2+,Er3+ [19,21,22], CaZnOS:Mn2+/Cu/Er3+/Sm3+ [23–26], ZrO2:Ti4+ [27], and CaZr(PO4)2:Eu2+ [28]. Among the developed ML materials, there exists a series of isomorphic materials which show a similarity of crystal form including the same chemical formulation and space group, and containing similar atoms which have corresponding chemical properties and similar sizes. The typical instances are Eu2+-activated CaAl2Si2O8 and SrAl2Si2O8 (space group P1) [29], Eu2+-activated Sr2MgSi2O7 and Ca2MgSi2O7 (space group P4¯21m) [30], and Mn2+-activated ZnGa2O4, MgGa2O4 and ZnAl2O4 (space group Fd3¯m) [31,32]. Interestingly, these isomorphic materials activated by the same luminescent ions show differentiated ML properties (e.g. intensity and wavelength) due to the slight discrepancy in unit cell dimensions, thereby supplying selectivity for diverse application needs. This clue hints us to develop new TC-ML materials and regulate their characteristics via constructing isomorphic crystals of the existing TC-ML materials.

Recently, we created a series of novel TC-ML materials through doping Pr3+ activators to introduce luminescent centers and carrier traps in piezoelectric calcium niobates [8]. We showed that Ca2Nb2O7:Pr3+ could be used as a sensitive optical probe for real-time observation of dynamic stress concentration because of the intense emission and high reproducibility of ML. However, Ca2Nb2O7:Pr3+ emits a bright long-lasting afterglow after UV excitation. Such strong afterglow noise would mask the ML signal, reducing the sharpness of ML images. In order to achieve high-contrast ML images, Ca2Nb2O7:Pr3+ requires a long delay time to fade the strong noise of afterglow. Obviously, long-time delay complicates the imaging process, thereby proposing the requirement of reducing delay time. Inspired by the isomorphic strategy, in this work, we report a new TC-ML material Sr2Nb2O7:Pr3+ displaying intense ML under the mechanical actions of compression and friction. More importantly, Sr2Nb2O7:Pr3+ shows a weak and short-lasting afterglow and enables us to achieve high-contrast ML in a short-time delay, which greatly simplifies the imaging process to meet application needs. Furthermore, we investigate the unit cell dimensions, trap distribution and optical bandgap of Sr2Nb2O7 and Ca2Nb2O7, and finally propose a possible mechanism underlying their different properties of afterglow and ML.

2. Experimental

2.1 Material synthesis

Pr3+-activated strontium niobate and calcium niobate with compositions of Sr2-xPrxNb2O7 (0 ≤ x ≤ 0.03, abbreviated as Sr2Nb2O7:Pr3+ (0-3 mol%)) and Ca2-xPrxNb2O7 (0 ≤ x ≤ 0.02, abbreviated as Ca2Nb2O7:Pr3+ (0-2 mol%)) were synthesized by solid-state reaction among CaCO3 (Aladdin, 99.99%), SrCO3 (Aladdin, 99.99%), Nb2O5 (Sinopharm, 99.99%), and Pr6O11 (Alfa Aesar, 99.99%). Stoichiometric raw materials were ground and pre-calcined at 900 °C for 4 h in air, and re-ground. The powders were finally sintered at 1400 °C for 4 h in air, and pulverized after natural cooling.

2.2 Physical characterization

Powder X-ray diffraction (XRD) patterns were collected on a Rigaku SmartLab X-ray diffractometer (Cu Kα, λ = 1.54056 Å) with a continuous scanning rate of 5° min−1 for phase identification and a step scanning rate of 8 s step−1 (step size: 0.01°) for Rietveld analysis. The refinement of crystal structure was performed using the General Structure Analysis System program. Diffuse reflectance spectra were recorded by a UV/vis/NIR spectrophotometer (V570, Jasco). Spectra and decay curves of afterglow were measured using a fluorescence spectrometer (F4600, Hitachi). Thermoluminescence (ThL) was measured using a ThL meter (FJ427A1, Beijing Nuclear Instrument Factory). To characterize ML triggered by compression and friction, cylinder-shaped composite disks (25 mm in diameter and 15 mm in thickness) were prepared by mixing the as-synthesized particles with an optical epoxy resin. Compressive loading (0-1000 N) at a deformation rate of 3 min min−1 was applied on the composite disks using a lab-built electronic universal testing machine. Mechanical friction was applied by an in-house made friction machine with a metal tip of 1 mm diameter (load of 10 N, friction rate of 120 round min−1). Afterglow decay and ML response of composite disks were recorded using a lab-made photon-counting system. Spectra of ML were captured with a photon multi-channel analyser system (QE65000, Ocean Optics). Optical photographs of afterglow and ML were recorded using a Canon 7D camera. The imaging parameters are manual/ISO 6400/0.5 s. Before the measurements of afterglow, ML and ThL, if not otherwise specified, the samples were irradiated by a hand-held UV lamp (8 W, 308 nm) for 1 min to charge traps. All measurements except ThL were performed at room temperature.

3. Results and discussion

The XRD results confirm that the as-synthesized Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ both crystallize in phase-pure orthorhombic structure with space group Pna21 [33,34]. Figures 1(a) and 1(b) show the Rietveld refined XRD patterns of Sr2Nb2O7 and Ca2Nb2O7 host lattices, respectively. In comparison with Ca2Nb2O7, Sr2Nb2O7 shows a lattice expansion due to larger radius of Sr2+ than Ca2+. The refined lattice parameters are a = 26.7803 Å, b = 7.9101 Å, c = 5.7015 Å for Sr2Nb2O7, and a = 26.4559 Å, b = 7.6901 Å, c = 5.4964 Å for Ca2Nb2O7, as summarized in Table 1. Figures 1(c) and 1(d) show the crystal structures and Sr/Ca-O coordination environments of Sr2Nb2O7 and Ca2Nb2O7, respectively. Sr2Nb2O7 and Ca2Nb2O7 have extremely similar perovskite-derivative structures, in which the perovskite slabs consisting of vertex sharing NbO6 octahedra are separated by additional O atoms. Sr2+ and Ca2+ ions both possess two coordination numbers (CN = 8 and 12), but the average bond distances of Sr-O (2.7913 Å for CN = 8 and 2.9069 Å for CN = 12) are longer than those of Ca-O (2.6394 Å for CN = 8 and 2.7275 Å for CN = 12). The average bond distance of Nb-O (2.085 Å) in Sr2Nb2O7 is also a little longer than that (2.000 Å) in Ca2Nb2O7. The differences in these structural parameters together result in Ca2Nb2O7 exhibiting a more distorted crystal structure than Sr2Nb2O7 [Figs. 1(c) and 1(d)].

 figure: Fig. 1

Fig. 1 Rietveld refinement of XRD profiles, crystal structures and Sr/Ca-O coordination environment (including bond distances): (a) and (c) for Sr2Nb2O7; (b) and (d) for Ca2Nb2O7.

Download Full Size | PDF

Tables Icon

Table 1. Refinement, crystallographic, and structural parameters of Sr2Nb2O7 and Ca2Nb2O7

Figure 2 shows the diffuse reflectance spectra of Sr2Nb2O7 and Ca2Nb2O7. According to the Kubelka-Munk function [35] and the Tauc relation [36], the optical band gaps are estimated [insets of Figs. 2(a) and (b)], i.e., 3.97 eV for Sr2Nb2O7 and 4.17 eV for Ca2Nb2O7, which are really close to the previously reported values of 3.92 eV [37] and 4.18 eV [38], respectively. After the introduction of Pr3+ ions, the absorption edges of both compounds show an insignificant shift, indicating a negligible change of bandgaps. It indicates that the bandgap of Sr2Nb2O7:Pr3+ is ca. 0.2 eV narrower than that of Ca2Nb2O7:Pr3+.

 figure: Fig. 2

Fig. 2 Diffuse reflectance spectra of (a) Sr2Nb2O7:Pr3+ (0 and 0.4 mol%) and (b) Ca2Nb2O7:Pr3+ (0 and 0.2 mol%).

Download Full Size | PDF

Figure 3 shows the characteristics of red-emitting long-lasting afterglow from Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+. The spectra of afterglow indicate that the red emission of Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ both mainly originates from the 1D2-3H4 transition of Pr3+ activators [upper insets of Figs. 3(a) and 3(b)]. Furthermore, the concentration of Pr3+ ions strongly influences the intensity and duration of afterglow, suggesting that the Pr3+ ions, in addition to serving as luminescent centers, could also regulate trap distribution. Overall, the afterglow of the Sr2Nb2O7:Pr3+ series is much weaker and shorter than that of the Ca2Nb2O7:Pr3+ series. Take their optimization components for comparison, after irradiation of a hand-held UV lamp (308 nm, 8 W) for 1 min, the afterglow of Sr2Nb2O7:Pr3+ (0.4 mol%) powders lasts less than 2 min [lower insets of Fig. 3(a)], whereas that of Ca2Nb2O7:Pr3+ (0.2 mol%) powders can be observed by naked eyes more than 20 min [lower insets of Fig. 3(b)].

 figure: Fig. 3

Fig. 3 Dependence of afterglow decay curves on Pr3+ concentration: (a) Sr2Nb2O7:Pr3+ (0.2-3 mol%) (315 nm irradiation, 1 min) and (b) Ca2Nb2O7:Pr3+ (0.1-2 mol%) (295 nm irradiation, 1 min). The upper insets show the spectra of afterglow. The lower inset show the photographs of afterglow captured at different delay times.

Download Full Size | PDF

In a further set of experiments, we incorporated the Sr2Nb2O7:Pr3+ (0.4 mol%) and Ca2Nb2O7:Pr3+ (0.2 mol%) powders into phosphor/resin composition disks, and executed compression- and friction-triggered ML tests under different delay times after UV irradiation [Figs. 4 and 5]. We firstly measured the compression-triggered ML of two disks without any delay after irradiation. For the Sr2Nb2O7:Pr3+ (0.4 mol%) disk showing a weak afterglow background [Fig. 4(a-i)], we can clearly observe the fusiform distribution of ML [Fig. 4(a-ii)]. The profile and location of ML spectra (not shown here) are consistent with those of afterglow spectra [insets of Fig. 3] [8,10,28], indicating that ML also mainly originates from the 1D2-3H4 transition of Pr3+. However, the compression-triggered ML from Ca2Nb2O7:Pr3+ (0.2 mol%) disk recorded under the same compression condition is unrecognized [Fig. 4(b-ii)] due to the cover of strong afterglow noise [Fig. 4(b-i)]. In order to eliminate the negative effect of afterglow, we set the irradiated disks in dark for some time, i.e., 90 s delay for Sr2Nb2O7:Pr3+ (0.4 mol%) disk [Fig. 4(a-iii)] and 900 s delay for Ca2Nb2O7:Pr3+ (0.2 mol%) disk [Fig. 4(b-iii)], which fades the afterglow background to an unobservable level for naked eyes. After such a delay, compression-triggered ML from Sr2Nb2O7:Pr3+ (0.4 mol%) and Ca2Nb2O7:Pr3+ (0.2 mol%) disks all present a linear growth with increasing compressive load, and at peak compression ML reach a close intensity [Figs. 4(a) and 4(b)] and a similar image contrast [Figs. 4(a-iv) and 4(b-iv)]. Figure 5(a) lists the photographs of compression-triggered ML of two disks captured at different delay times. It is evident that Sr2Nb2O7:Pr3+ (0.4 mol%) disk shows high-contrast images of compression-triggered ML after a short delay time (≤ 90 s) due to a weak and short afterglow, whereas Ca2Nb2O7:Pr3+ disk showing a strong and long afterglow takes a long-time delay (≥ 600 s) to obtain a similar image of compression-triggered ML. Moreover, as the delay time increases, the intensity of compression-triggered ML shows a similar attenuation as that of afterglow, indicating that afterglow and ML originate from the evacuation process of carriers from the same traps. In addition to the characteristic of compression-triggered ML, the Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ disks also show friction-triggered ML [Fig. 5(b)], which is reported for the first time. Similar with compression-triggered ML, the Sr2Nb2O7:Pr3+ (0.4 mol%) disk enables to capture a high-contrast image of friction-triggered ML within a shorter delay compared to the Ca2Nb2O7:Pr3+ (0.2 mol%) disk. These results show that the short-delay, high-contrast ML of Sr2Nb2O7:Pr3+ has a more attractive prospect in the practical application of stress imaging.

 figure: Fig. 4

Fig. 4 Afterglow decay and compression-triggered ML response during applying dynamic triangle loading on phosphor/resin composition disks: (a) Sr2Nb2O7:Pr3+ (0.4 mol%) and (b) Ca2Nb2O7:Pr3+ (0.2 mol%). Insets show photographs of afterglow (i, iii) and ML (ii, iv) captured at typical delay times.

Download Full Size | PDF

 figure: Fig. 5

Fig. 5 (a) Photographs of compression-triggered ML (captured at the peak compression of 1000 N) after different delay times: Sr2Nb2O7:Pr3+ (0.4 mol%) (upper) and Ca2Nb2O7:Pr3+ (0.2 mol%) (lower). (b) Friction exertion device (left), and photographs of afterglow and friction-triggered ML recorded at different delay times: Sr2Nb2O7:Pr3+ (0.4 mol%) (top right) and Ca2Nb2O7:Pr3+ (0.2 mol%) (bottom right).

Download Full Size | PDF

In order to understand the differences of trap-controlled afterglow and ML described above, we characterized ThL to clarify the distribution of traps in Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ [Fig. 6]. Both Sr and Ca compounds show ThL after Pr3+ doping, indicating the formation of traps by Pr3+ dopants [Figs. 6(a) and 6(c)]. It is known that the intensity and location of ThL curves reveal two important parameters of trap levels, i.e., the former reflects information of trap density and the latter that of trap depth [39]. The concentration of Pr3+ activators significantly changes the trap density of both compounds. However, the increase of Pr3+ concentration shifts the ThL profiles of Sr2Nb2O7:Pr3+ to higher temperature, while hardly shifts that of Ca2Nb2O7:Pr3+. According to the discrepancy of ion radius between Ca2+, Sr2+ and Pr3+ [40], we argue that the non-shift in Ca2Nb2O7:Pr3+ should be ascribed to the extremely close radii of Pr3+ and Ca2+, while the shift in Sr2Nb2O7:Pr3+ might result from the less radius of Pr3+ than that of Sr2+. Despite the underlying reason still needs further study, it is reasonable to infer that more Pr3+ substitution for Sr2+ would induce more lattice disturbance, thereby shifting the ThL curves. Furthermore, with the increase of Pr3+ concentration, the variation trend of ThL integrated intensity [insets of Figs. 6(a) and 6(c)] is consistent with that of afterglow intensity [Figs. 3(a) and 3(b)] in two compounds, indicating a strong dependence of afterglow intensity on the trap density. Typically, the intensity of ThL curve of Sr2Nb2O7:Pr3+ (0.4 mol%) is much less than that of Ca2Nb2O7:Pr3+ (0.2 mol%) [Figs. 6(a) and 6(c)]. Accordingly, the much weaker afterglow of Sr2Nb2O7:Pr3+ [lower insets of Fig. 3] is ascribed to its greatly lower trap density than that of Ca2Nb2O7:Pr3+. The lower trap density in Sr2Nb2O7:Pr3+ may result from the more non-radiative lattice defects due to the larger size difference of Sr2+ relative to Pr3+. On the other hand, each ThL curve can be further deconvoluted into two components which should result from two kinds of traps (Trap 1 and Trap 2) in materials. Figures 6(b) and 6(d) shows the Gaussian deconvolution of ThL curves of Sr2Nb2O7:Pr3+ (0.4 mol%) and Ca2Nb2O7:Pr3+ (0.2 mol%), respectively. According to the above structural analysis, these two traps should originate from the substitution of Pr3+ for Sr2+/Ca2+ (i.e., electron traps [PrSr/Ca]o) in two different coordination environments (CN = 8 and 12) [Fig. 1(c)]. We estimate the depths of two traps by the Hoogenstraaten’s method [41], i.e., 0.46 eV (Trap 1) and 0.59 eV (Trap 2) in Sr2Nb2O7:Pr3+ (0.4 mol%), while 0.52 eV (Trap 1) and 0.63 eV (Trap 2) in Ca2Nb2O7:Pr3+ (0.2 mol%). The correspondingly shallower depths in Sr2Nb2O7:Pr3+ results in the shorter duration of afterglow [Fig. 3(a)], conversely in Ca2Nb2O7:Pr3+ [Fig. 3(b)]. The shallower trap depth in Sr2Nb2O7:Pr3+ may be related to the narrower bandgap of Sr2Nb2O7 [Fig. 2]. A narrow bandgap narrows the trap level to the bottom of the conduction band, making the trap shallow.

 figure: Fig. 6

Fig. 6 ThL curves and estimated trap depths of Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+. (a) and (c) Dependence of ThL on Pr3+ concentration in Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+, respectively. Insets show the relative integral intensity of ThL as a function of Pr3+ concentration. (b) and (d) Gaussian deconvolution of ThL curves and trap depths estimated by Hoogenstraaten plots in Sr2Nb2O7:Pr3+ (0.4 mol%) and Ca2Nb2O7:Pr3+ (0.2 mol%), respectively.

Download Full Size | PDF

According to the above analyses, we propose a possible process of ML to describe the differences of TC-ML between Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+, as shown in Fig. 7. Upon the excitation of UV light, the pumped electrons of Pr3+ activators would be captured by traps. Under the stimulus of ambient thermal energy, the trapped electrons are slowly released back to the excited state of Pr3+ through conduction band or tunneling effect [8]. The population of 3P0 level of Pr3+ would be non-radiatively relaxed to 1D2 level via the intervalence charge transfer state of Pr-O-Nb [42,43], resulting in a red-emitting afterglow corresponding to 1D2-3H4 transition. Due to the lower trap concentration and shallower depth, Sr2Nb2O7:Pr3+ displays a weaker and shorter afterglow than Ca2Nb2O7:Pr3+. Under the mechanical action, the captured electrons from the same traps would be detrapped much faster to produce a red-emitting ML as the emission process of afterglow. In view of the simultaneous input of ambient energy, ML is accordingly accompanied by afterglow. Therefore, Sr2Nb2O7:Pr3+ requires a shorter delay time to achieve a high-contrast ML image due to weak and short afterglow, whereas Ca2Nb2O7:Pr3+ conversely.

 figure: Fig. 7

Fig. 7 Schematic illustration of bandgap and trap-distribution co-tailored ML of Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ (CB: conduction band, VB: valence band).

Download Full Size | PDF

4. Conclusion

A novel TC-ML material, Sr2Nb2O7:Pr3+, has been developed that displays intense red-emitting ML under the mechanical actions of compression and friction. Compared with its isomorphism Ca2Nb2O7:Pr3+, Sr2Nb2O7:Pr3+ shows a weaker and shorter-lasting afterglow. This capability of Sr2Nb2O7:Pr3+ enables us to achieve high-contrast ML over a short-time delay, which greatly simplifies the imaging process to facilitate the further application of ML. The investigation of crystal structure, optical bandgap and trap distribution indicates that the short-delay, high-contrast ML of Sr2Nb2O7:Pr3+ should result from its lower trap concentration and shallower trap depth than Ca2Nb2O7:Pr3+. When one considers a library of isomorphic candidates for host lattices, we expect that a growing amount of novel TC-ML materials will become obtainable for the fundamental research and technological utilization.

Funding

National Natural Science Foundation of China (11774189, 11647168, 51673103 and 11465010); Basic Scientific Fund for National Public Research Institutes of China (2018Y02); Program of Science and Technology in Qingdao City (17-1-1-85-jch); China Postdoctoral Science Foundation (2016M592128); Natural Science Foundation of Shandong Province (2016ZRB019MU); Fundamental Research Funds for the Central Universities (2232018D-39).

References and links

1. Y. Sagara and T. Kato, “Mechanically induced luminescence changes in molecular assemblies,” Nat. Chem. 1(8), 605–610 (2009). [CrossRef]   [PubMed]  

2. M. Jin, T. Seki, and H. Ito, “Mechano-responsive luminescence via crystal-to-crystal phase transitions between chiral and non-chiral space groups,” J. Am. Chem. Soc. 139(22), 7452–7455 (2017). [CrossRef]   [PubMed]  

3. Y. Sagara, K. Kubo, T. Nakamura, N. Tamaoki, and C. Weder, “Temperature-dependent mechanochromic behavior of mechanoresponsive luminescent compounds,” Chem. Mater. 29(3), 1273–1278 (2017). [CrossRef]  

4. X. Li, R. Liang, J. Tao, Z. Peng, Q. Xu, X. Han, X. Wang, C. Wang, J. Zhu, C. Pan, and Z. L. Wang, “Flexible light emission diode arrays made of transferred Si microwires-ZnO nanofilm with piezo-phototronic effect enhanced lighting,” ACS Nano 11(4), 3883–3889 (2017). [CrossRef]   [PubMed]  

5. Y. Chen, Y. Zhang, D. Karnaushenko, L. Chen, J. Hao, F. Ding, and O. G. Schmidt, “Addressable and color-tunable piezophotonic light-emitting stripes,” Adv. Mater. 29(19), 1605165 (2017). [CrossRef]   [PubMed]  

6. D. Tu, C. N. Xu, A. Yoshida, M. Fujihala, J. Hirotsu, and X. G. Zheng, “LiNbO3:Pr3+: a multipiezo material with simultaneous piezoelectricity and sensitive piezoluminescence,” Adv. Mater. 29(22), 1606914 (2017). [CrossRef]   [PubMed]  

7. J. C. Zhang, L. Z. Zhao, Y. Z. Long, H. D. Zhang, B. Sun, W. P. Han, X. Yan, and X. Wang, “Color manipulation of intense multiluminescence from CaZnOS:Mn2+ by Mn2+ concentration effect,” Chem. Mater. 27(21), 7481–7489 (2015). [CrossRef]  

8. J. C. Zhang, Y. Z. Long, X. Yan, X. Wang, and F. Wang, “Creating recoverable mechanoluminescence in piezoelectric calcium niobates through Pr3+ doping,” Chem. Mater. 28(11), 4052–4057 (2016). [CrossRef]  

9. C. Pan, J. C. Zhang, M. Zhang, X. Yan, Y. Z. Long, and X. Wang, “Intrinsic oxygen vacancies mediated multi-mechano-responsive piezoluminescence in undoped zinc calcium oxysulfide,” Appl. Phys. Lett. 110(23), 233904 (2017). [CrossRef]  

10. X. H. Fan, J. C. Zhang, M. Zhang, C. Pan, X. Yan, W. P. Han, H. D. Zhang, Y. Z. Long, and X. Wang, “Piezoluminescence from ferroelectric Ca3Ti2O7:Pr3+ long-persistent phosphor,” Opt. Express 25(13), 14238–14246 (2017). [CrossRef]   [PubMed]  

11. N. Terasaki and C. N. Xu, “Historical-log recording system for crack opening and growth based on mechanoluminescent flexible sensor,” IEEE Sens. J. 13(10), 3999–4004 (2013). [CrossRef]  

12. N. Terasaki, H. Yamada, and C. N. Xu, “Ultrasonic wave induced mechanoluminescence and its application for photocatalysis as ubiquitous light source,” Catal. Today 201, 203–208 (2013). [CrossRef]  

13. Y. Fujio, C. N. Xu, Y. Terasawa, Y. Sakata, J. Yamabe, N. Ueno, N. Terasaki, A. Yoshida, S. Watanabe, and Y. Murakami, “Sheet sensor using SrAl2O4:Eu mechanoluminescent material for visualizing inner crack of highpressure hydrogen vessel,” Int. J. Hydrogen Energy 41(2), 1333–1340 (2016). [CrossRef]  

14. K. S. Sohn, S. Timilsina, S. P. Singh, J. W. Lee, and J. S. Kim, “A mechanoluminescent ZnS:Cu/rhodamine/SiO2/PDMS and piezoresistive CNT/PDMS hybrid sensor: red-light emission and a standardized strain quantification,” ACS Appl. Mater. Interfaces 8(50), 34777–34783 (2016). [CrossRef]   [PubMed]  

15. S. M. Jeong, S. Song, K. I. Joo, J. Kim, S. H. Hwang, J. Jeong, and H. Kim, “Bright, wind-driven white mechanoluminescence from zinc sulphide microparticles embedded in a polydimethylsiloxane elastomer,” Energy Environ. Sci. 7(10), 3338–3346 (2014). [CrossRef]  

16. S. M. Jeong, S. Song, H. J. Seo, W. M. Choi, S. H. Hwang, S. G. Lee, and S. K. Lim, “Battery-free, human-motion-powered light-emitting fabric: Mechanoluminescent textile,” Adv. Sust. Syst. 1(12), 1700126 (2017). [CrossRef]  

17. M. C. Wong, L. Chen, G. Bai, L. B. Huang, and J. Hao, “Temporal and remote tuning of piezophotonic-effect-induced luminescence and color gamut via modulating magnetic field,” Adv. Mater. 29(43), 1701945 (2017). [CrossRef]   [PubMed]  

18. C. N. Xu, T. Watanabe, M. Akiyama, and X. G. Zheng, “Artificial skin to sense mechanical stress by visible light emission,” Appl. Phys. Lett. 74(9), 1236–1238 (1999). [CrossRef]  

19. C. N. Xu, T. Watanabe, M. Akiyama, and X. G. Zheng, “Direct view of stress distribution in solid by mechanoluminescence,” Appl. Phys. Lett. 74(17), 2414–2416 (1999). [CrossRef]  

20. S. M. Jeong, S. Song, S. K. Lee, and N. Y. Ha, “Color manipulation of mechanoluminescence from stress-activated composite films,” Adv. Mater. 25(43), 6194–6200 (2013). [CrossRef]   [PubMed]  

21. H. Zhang, H. Yamada, N. Terasaki, and C. N. Xu, “Ultraviolet mechanoluminescence from SrAl2O4:Ce and SrAl2O4:Ce,Ho,” Appl. Phys. Lett. 91(8), 081905 (2007). [CrossRef]  

22. Y. Terasawa, C. N. Xu, H. Yamada, and M. Kubo, “Near infra-red mechanoluminescence from strontium aluminate doped with rare-earth ions,” IOP Conf. Series Mater. Sci. Eng. 18(21), 212013 (2011). [CrossRef]  

23. J. C. Zhang, C. N. Xu, S. Kamimura, Y. Terasawa, H. Yamada, and X. Wang, “An intense elastico-mechanoluminescence material CaZnOS:Mn2+ for sensing and imaging multiple mechanical stresses,” Opt. Express 21(11), 12976–12986 (2013). [CrossRef]   [PubMed]  

24. D. Tu, C. N. Xu, Y. Fujio, and A. Yoshida, “Mechanism of mechanical quenching and mechanoluminescence in phosphorescent CaZnOS:Cu,” Light Sci. Appl. 4(11), e356 (2015). [CrossRef]  

25. H. Zhang, D. Peng, W. Wang, L. Dong, and C. Pan, “Mechanically induced light emission and infrared-laser-induced upconversion in the Er-doped CaZnOS multifunctional piezoelectric semiconductor for optical pressure and temperature sensing,” J. Phys. Chem. C 119(50), 28136–28142 (2015). [CrossRef]  

26. W. Wang, D. Peng, H. Zhang, X. Yang, and C. Pan, “Mechanically induced strong red emission in samarium ions doped piezoelectric semiconductor CaZnOS for dynamic pressure sensing and imaging,” Opt. Commun. 395, 24–28 (2017). [CrossRef]  

27. M. Akiyama, C. N. Xu, and K. Nonaka, “Intense visible light emission from stress-activated ZrO2:Ti,” Appl. Phys. Lett. 81(3), 457–459 (2002). [CrossRef]  

28. J. C. Zhang, C. N. Xu, and Y. Z. Long, “Elastico-mechanoluminescence in CaZr(PO4)2:Eu2+ with multiple trap levels,” Opt. Express 21(11), 13699–13709 (2013). [CrossRef]   [PubMed]  

29. L. Zhang, C. N. Xu, H. Yamada, and N. Bu, “Enhancement of mechanoluminescence in CaAl2Si2O8:Eu2+ by partial Sr2+ substitution for Ca2+,” J. Electrochem. Soc. 157(3), J50–J53 (2010). [CrossRef]  

30. H. Zhang, N. Terasaki, H. Yamada, and C. N. Xu, “Mechanoluminescence of europium-doped SrAMgSi2O7 (A = Ca, Sr, Ba),” Jpn. J. Appl. Phys. 48, 04C109 (2009).

31. H. Matsui, C. N. Xu, M. Akiyama, and T. Watanabe, “Strong mechanoluminescence from UV-irradiated spinels of ZnGa2O4:Mn and MgGa2O4:Mn,” Jpn. J. Appl. Phys. 39(1), 6582–6586 (2000). [CrossRef]  

32. H. Matsui, C. N. Xu, and H. Tateyama, “Stress-stimulated luminescence from ZnAl2O4:Mn,” Appl. Phys. Lett. 78(8), 1068–1070 (2001). [CrossRef]  

33. K. Scheunemann and H. K. Müller-Buschbaum, “Zur kristallstruktur von Sr2Nb2O7,” J. Inorg. Mater. 37(7–8), 1679–1680 (1975).

34. K. Scheunemann and H. K. Müller-Buschbaum, “Zur kristallstruktur von Ca2Nb2O7,” J. Inorg. Mater. 36(9), 1965–1970 (1974).

35. G. Kortüm, W. Braun, and D. C. G. Herzog, “Principles and techniques of diffuse-reflectance spectroscopy,” Angew. Chem. Int. Ed. Engl. 2(7), 333–341 (1963). [CrossRef]  

36. J. I. Pankove, “Optical Processes in Semiconductors,” p. 34, (Dover Publications, 1971).

37. J. Nisar, B. Pathak, and R. Ahuja, “Screened hybrid density functional study on Sr2Nb2O7 for visible light photocatalysis,” Appl. Phys. Lett. 100(18), 181903 (2012). [CrossRef]  

38. L. Zhang, H. Fu, C. Zhang, and Y. Zhu, “Effects of Ta5+ substitution on the structure and photocatalytic behavior of the Ca2Nb2O7 photocatalyst,” J. Phys. Chem. C 112(8), 3126–3133 (2008). [CrossRef]  

39. S. W. McKeever, “Thermoluminescence of solids,” (Cambridge University Press, 1988).

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

41. W. Hoogenstraaten, “Electron traps in zinc-sulfide phosphors,” Philips Res. Rep. 13, 515–693 (1958).

42. P. Boutinaud, E. Cavalli, and M. Bettinelli, “Emission quenching induced by intervalence charge transfer in Pr3+- or Tb3+-doped YNbO4 and CaNb2O6,” J. Phys. Condens. Matter 19(38), 255–263 (2007). [CrossRef]  

43. Y. Gao, F. Huang, H. Lin, J. Zhou, J. Xu, and Y. Wang, “A novel optical thermometry strategy based on diverse thermal response from two intervalence charge transfer states,” Adv. Funct. Mater. 26(18), 3139–3145 (2016). [CrossRef]  

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 (7)

Fig. 1
Fig. 1 Rietveld refinement of XRD profiles, crystal structures and Sr/Ca-O coordination environment (including bond distances): (a) and (c) for Sr2Nb2O7; (b) and (d) for Ca2Nb2O7.
Fig. 2
Fig. 2 Diffuse reflectance spectra of (a) Sr2Nb2O7:Pr3+ (0 and 0.4 mol%) and (b) Ca2Nb2O7:Pr3+ (0 and 0.2 mol%).
Fig. 3
Fig. 3 Dependence of afterglow decay curves on Pr3+ concentration: (a) Sr2Nb2O7:Pr3+ (0.2-3 mol%) (315 nm irradiation, 1 min) and (b) Ca2Nb2O7:Pr3+ (0.1-2 mol%) (295 nm irradiation, 1 min). The upper insets show the spectra of afterglow. The lower inset show the photographs of afterglow captured at different delay times.
Fig. 4
Fig. 4 Afterglow decay and compression-triggered ML response during applying dynamic triangle loading on phosphor/resin composition disks: (a) Sr2Nb2O7:Pr3+ (0.4 mol%) and (b) Ca2Nb2O7:Pr3+ (0.2 mol%). Insets show photographs of afterglow (i, iii) and ML (ii, iv) captured at typical delay times.
Fig. 5
Fig. 5 (a) Photographs of compression-triggered ML (captured at the peak compression of 1000 N) after different delay times: Sr2Nb2O7:Pr3+ (0.4 mol%) (upper) and Ca2Nb2O7:Pr3+ (0.2 mol%) (lower). (b) Friction exertion device (left), and photographs of afterglow and friction-triggered ML recorded at different delay times: Sr2Nb2O7:Pr3+ (0.4 mol%) (top right) and Ca2Nb2O7:Pr3+ (0.2 mol%) (bottom right).
Fig. 6
Fig. 6 ThL curves and estimated trap depths of Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+. (a) and (c) Dependence of ThL on Pr3+ concentration in Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+, respectively. Insets show the relative integral intensity of ThL as a function of Pr3+ concentration. (b) and (d) Gaussian deconvolution of ThL curves and trap depths estimated by Hoogenstraaten plots in Sr2Nb2O7:Pr3+ (0.4 mol%) and Ca2Nb2O7:Pr3+ (0.2 mol%), respectively.
Fig. 7
Fig. 7 Schematic illustration of bandgap and trap-distribution co-tailored ML of Sr2Nb2O7:Pr3+ and Ca2Nb2O7:Pr3+ (CB: conduction band, VB: valence band).

Tables (1)

Tables Icon

Table 1 Refinement, crystallographic, and structural parameters of Sr2Nb2O7 and Ca2Nb2O7

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.