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

Enhancing near infrared persistent luminescence from Cr3+-activated zinc gallogermanate powders through Ca2+ doping

Open Access Open Access

Abstract

Near infrared (NIR) long-persistent luminescence phosphors are promising for applications ranging from night-vision surveillance to in vivo bioimaging. Yet, the luminescence brightness and afterglow period remain insufficient for the reported persistent phosphors with both the activator content optimized and the host material defined. Here, we show that the emission profile of the emerging NIR persistent phosphors of Cr3+-activated spinel zinc gallogermanate (emission at 650–850 nm from the 2E→4A2 transition of Cr3+) can be improved through the incorporation of non-luminescent, divalent calcium (Ca2+) into the host lattice. We found that a slight introduction of 3% Ca2+ ions into the formulated afterglow material was able to enhance its persistent luminescence intensity (recorded after 300s stoppage of the excitation light) by about 15 fold. This was possibly ascribed to the engineering of tetrahedral trapping defects (spinel inversion) surrounding the emitting Cr3+ ions at the octahedral sites and the reduction of luminescence quenching centers in the crystal, enacted by the calcium doping. The simple performance-enhancing route described here has an immediate implication for other visible and NIR persistent phosphors engaged in a plethora of photonic and biomedical applications.

© 2017 Optical Society of America

1. Introduction

Persistent luminescence is a phenomenon whereby luminescence can last for a long time ranging from hours to days after the stoppage of the excitation light [1–4]. Since its discovery in 17th century, phosphors of long-persistent luminescence have gained a great deal of interest due to their use as night-vision materials in numerous important fields such as, security sign, emergency route sign, traffic signage, and medical diagnostics [5–9]. Persistent luminescence phosphors typically involve two kinds of active centers (i.e., emitters and traps) in an appropriate host lattice to create afterglow luminescence. Emitters are the doped metallic ions capable of emitting radiation at designated wavelength, while the traps (typically defined by the host lattice) store the excitation energy and then gradually transfer it to the emitters owing to thermal or other physical stimulations. Persevere investigations of the combination of the inorganic host lattices and the doped metallic emitting ions, have resulted in successful development and marketing of multicolor persistent phosphors in the visible range [10, 11]. The representative ones include CaAl2O4:Eu2+/Nd3+ (blue), SrAl2O4:Eu2+/Dy3+ (green), and Y2O2S:Eu3+, Mg2+, Ti4+ (red) phosphors of three primary colors [10, 12, 13]. These phosphors are able to emit strong and long-lasting (>10 h) luminescence under excitation with sunlight or room light within minutes. However, the development of bright long-lasting persistent phosphors in the near infrared (NIR) range remains rather limited.

Recently, a set of spinel gallate materials doped with trivalent chromium (Cr3+) (e.g., La3Ga5GeO14:Cr3+, and ZnGa2O4:Cr3+) are emerging as long-lasting NIR persistent substance [14–18]. The emission wavelength is about 700 nm originating from the 2E→4A2 transition of Cr3+ emitters, while the afterglow time is over hours that are mainly determined by the type and distribution of formed traps during the growth of the spinel crystal. Among them, zinc gallogermanate phosphors with an optimized formula of Zn3Ga2Ge2O10:0.5%Cr3+ are of great interest, which are reported to exhibit persistent luminescence with the longest afterglow time over 360 h, observed with the aid of a night-vision monocular in a dark room [17]. This compound also possesses the AB2O4 spinel structure, whereas Zn2+ occupies the tetrahederal sites, Ga3+ occupies the octahedral sites, and Ge4+and Cr3+ replace Ga3+ to occupy the octahedral sites [3, 19]. Though the brightness of NIR long-lasting luminescence can be optimized through optimizing the content of the composition (Zn1+xGa2−2xGexO4, 0≤x≤0.5) or through substitution of the occupants at the octahedral sites with trivalent tin (Sn3+) [3], there lack routes to further improve the optical properties of the established spinel compound with a defined formula.

In this work, we show that the brightness of persistent luminescence from the established nominal Zn3Ga2Ge2O10:0.5%Cr3+ materials can be improved further through divalent calcium (Ca2+) doping. A slight substitution of zinc ions with Ca2+ (ca. 3%) at the tetrahedral sites can improve the persistent luminescence brightness by 15 fold after 5 min stoppage of the excitation light.

2. Method

Material Synthesis. The powder samples of Zn3Ga2Ge2O10:0.5%Cr3+, x%Ca2+ (x = 0, 1, 3, 5) and zin-deficient ZGGO:0.5%Cr,d3%Zn (with 3% zinc deficient) were synthesized using a conventional solid-state reaction method. In a typical procedure, stoichiometric ZnO (Aladdin, 99.99%), Ga2O3 (Aladdin, 99.99%), GeO2 (Aladdin, 99.99%), Cr2O3 (Aladdin, 99.95%), CaO (Aladdin, 99.9%) were first mixed thoroughly, and ground to form a homogeneous fine powder in an agate mortar. Subsequently, the mixed powders were transferred to a corundum crucible, and then heated at 900 °C in a muffle furnace for 2 hours. After this, the prefired material was ground again to a fine powder suitable for sintering. Finally, the powder samples were sintered at 1100 °C for 2 hours in air.

Characterization Methods. The morphology of the resulting powders was characterized by a field emission scanning electron microscope (SEM) (Zeiss Supra 55) at an acceleration voltage of 15 kV. The powder X-ray diffraction (XRD) patterns were recorded by a Siemens D500 diffractometer, using Cu Kα radiation (λ = 0.15418 nm). The 2θ angle of the XRD spectra was recorded at a scanning rate of 5°/min. Photoluminescence excitation spectra, emission spectra, afterglow spectra, and afterglow decay curves were all recorded on FLS980 spectrometer (Edinburgh instruments). Thermoluminesence spectra was acquired using a home-made setup with a temperature-controlled sample holder and a monochromator for afterglow luminescence detection. To acquire the excitation spectra, the powders were excited by light in the wavelength range of 330-500 nm while monitoring the emission intensity at 696 nm. Emission spectra were recorded under the excitation at 410 nm. Light excitation was provided by the 450 W ozone-free xenon arc lamp with optical filtering defined by the excitation monochormator of FLS 980. After charging the samples with ultraviolet light at 320 nm for 5 mins, photographic images of persistent luminescence were recorded by iPhone 6s (Apple Inc.) at varying time points after the stoppage of excitation light.

3. Results and discussions

To investigate the influence of Ca2+ doping on the morphology change of the resulting particles, we acquired scanning electron microscopic (SEM) images of Zn3Ga2Ge2O10:0.5%Cr3+, x%Ca2+ (x = 0, 1, 3, 5) powder samples prepared under the identical condition (Fig. 1). As one can see in this figure, the size and morphology of the sample without Ca2+ doping are diversified, with a large amount of small particles with size of 200-700 nm, and also rod-shaped particles with length over 2 µm. Along with the increase of Ca2+ doping content, the resulting shape becomes irregular, and the resulting size grows significantly into an average size of 2-4 µm for Ca2+ 1%, and over 4-5 µm for both Ca2+ 3 and 5%, respectively. Note that sub-micrometer sized particles observed for the sample with null Ca2+ doping almost disappear for all the samples with Ca2+ doping. Moreover, the escalated Ca2+ content tends to evoke the growth of the resulting particles into one single crystal due to the increased grain size and the smeared particle boundaries. The improved crystallization generally can help reduce the luminescence quenching centers, thus favoring higher photoluminescence and persistent luminescence intensity.

 figure: Fig. 1

Fig. 1 Scanning electron microscopic (SEM) images of the resulting Zn3Ga2Ge2O10:0.5%Cr3+, x%Ca2+ (x = 0, 1, 3, 5) powder samples. Scale bar, 2 µm.

Download Full Size | PDF

We further acquired the x-ray diffraction (XRD) patterns of Zn3Ga2Ge2O10:0.5%Cr3+, x% Ca2+ (x = 0, 1, 3, 5) powder samples to investigate the impact of calcium doping on the crystallographic structure (Fig. 2). A dominant spinel phase was observed for the resulting sample of Zn3Ga2Ge2O10:0.5%Cr3+ (null Ca2+ doping), but accompanied with impurity crystal phases of Zn2GeO4 (Fig. 2(b)) and GeO2 (Fig. 2(c)). Whereas Ca2+-doped samples all have identical peak positions that agree well with the standard PDF 38-1240 pattern of spinel ZnGa2O4. No impurity diffraction peaks were observed. This observation implies that the doped Ca2+ ion substitutes the Zn2+ ion at the tetrahedral sites to enter the host lattice of a spinel structure. This is reasonable as the ionic radii of Ca2+ (114 pm) and Zn2+ (88 pm) are similar, and no charge compensation is needed for such substitution. However, the ionic radii difference will create impact on the existing structure disorder (the number and distribution of antisite defect) and the local environment of Cr3+-occupied octahedral site, thus influencing their luminescent properties. The well-defined peaks in Fig. 2 are indicative of the high crystallinity of the resulting sample of Zn3Ga2Ge2O10:0.5%Cr3+, x% Ca2+ (x = 0, 1, 3, 5). Moreover, the disappearance of impurity XRD peaks in Ca2+-doped sample suggests that Ca2+ is able to help form samples with improved crystallinity, in good agreement with SEM result in Fig. 1.

 figure: Fig. 2

Fig. 2 X-ray diffraction (XRD) patterns of Zn3Ga2Ge2O10:0.5%Cr3+ codoped with Ca2+ of (d) null, (e) 1%, (f) 3%, and (g) 5%, in reference to the standard ones of (a) ZnGa2O4 (PDF 38-1240), (b) Zn2GeO4 (PDF 11-0687), and (c) GeO2 (PDF 65-8052)

Download Full Size | PDF

Figure 3(a) shows the luminescence spectra of the samples of Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) and Zn2.91Ga1.99Ge2O9.91: 0.5%Cr3+ under excitation at 410 nm. A broadband emission peaked at 698 nm (extending from 650 to 850 nm) is observed for all the samples, which originates from the 2E→4A2 transition of Cr3+ emitter. This is in marked contrast to the reported luminescence spectra of Cr-doped spinel persistent phosphors (ZnGa2O4:Cr3+), which consists of zero phonon (ZP) lines (687.0 nm), N2 line (695 nm, structure-dependent line) and phonon side bands (PSB) (708 nm, 715 nm and 722 nm for Stokes PSB, and 670 nm and 680 nm for Anti-Stokes PSB) [16]. It has been established that the ZP lines (687 nm) and PSB lines are characteristic of Cr3+ occupying an octahedral site (D3h symmetry) of ideal spinel environment, while the N2 line (695 nm) corresponds to an environment distorted by an antisite defect located in the first cationic neighbors of Cr3+ (inversion spinel structure) [16]. The observation of the intense peak at 698 nm (i.e., the N2 line) from our samples implies that the luminescence primarily comes from the Cr3+ ions located at antisite-defect-distorted octahedral sites. It is striking that the luminescence intensity increases by more than two-fold along with an escalation of Ca2+ content from 0 to 3%. However, the luminescence intensity drops significantly with a further increase of Ca2+ concentration to 5%. A small amount of zinc deficiency in the lattice was reported to be able to enhance the brightness and lengthen afterglow time of persistent luminescence from a spinel structure [20, 21]. We, therefore, prepared the Zn3Ga2Ge2O10:0.5%Cr3+, d3%Zn sample with 3% zinc deficiency, and include its spectrum in Fig. 3. The observation of decreased luminescence from the Zn3Ga2Ge2O10:0.5%Cr3+, d3%Zn than from the Zn3Ga2Ge2O10:0.5%Cr3+ sample rules out the mechanisms that the observed luminescence enhancement comes from the reduction of Zn content in the host lattice due to the replacement of Zn2+ by Ca2+. Indeed, when normalizing the luminescence spectra to the PSB peak at 716 nm (the R line), the intensity of N2 line at 695 nm in reference to the R line at 716 nm increases with Ca2+ content, and reaches an apex at 3% Ca2+. This demonstrates the introduction of Ca2+ is able to increase the antisite defects in the host lattice that are essential for persistent luminescence. Note that the spectra for samples with Ca2+ 1% and 5% are similar in Fig. 3 (b), but quite different in intensity in Fig. 3(a). This difference implies that the Ca2+ dopant is able to enhance the energy trapping sites, but its high concentration can introduce quenching effects. Moreover, the Ca2+-induced alternation of luminescence intensity in Fig. 3(a) corresponds well to the change of the excitation spectra in Fig. 3(c). The excitation spectrum covers a spectral region from 330 to 500 nm, which consists of two main excitation bands originating from the 4A24T1 (te2) transition (330 nm), and the 4A24T1 (t2e) transition (410 nm) of Cr3+, respectively. Taken together, the results in Fig. 3 indicate that Ca2+ renders photoluminescence enhancement through engineering the structural disorder of the host lattice.

 figure: Fig. 3

Fig. 3 (a) Photoluminescence, (b) normalized photoluminescence (normalized to the peak at 716 nm), and (c) excitation spectra of the samples Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) and Zn2.91Ga1.99Ge2O9.91: 0.5%Cr3+. The emission intensity at 698 nm was monitored to acquire the excitation spectra of Fig. 3 (c).

Download Full Size | PDF

We also investigated the doping influence of Ca2+ on the persistent luminescence from the Zn3Ga2Ge2O10:0.5%Cr3+ sample, and presented afterglow decay curves of all samples in Fig. 4(a). To acquire the decay profiles, the samples were irradiated with 410 nm light for 60 s, and then recorded luminescence at 698 nm for a period of 300 s post the stoppage of light stimulus. As one can see, the persistent luminescence intensity of Zn3Ga2Ge2O10:0.5%Cr3+ sample was significantly improved due to the introduction of Ca2+ ions into the host lattice. The observed afterglow spectra (Fig. 4(c)) is similar to the photoluminescence spectra in Fig. 3(a). However, unlike the observation of 4-fold lower luminescence intensity for Ca2+ 5% sample than Ca2+ 0% sample, the Ca2+ 5% sample exhibits about 7-fold higher persistent luminescence than the Ca2+ 0% sample at 300 s post the stoppage of the light excitation (Fig. 4(b)). This confirms that the Ca2+ doping is able to engineer the structural disorder of the spinel host lattice. Moreover, the sample doped with 3% Ca2+ concentration exhibits the highest persistent luminescence intensity for the whole period of recording. After the cease of excitation for 300s, the persistence luminescence intensity of this sample is calculated to be about 15 fold higher than the powder Zn3Ga2Ge2O10:0.5%Cr3+ sample with null Ca2+ doping, which was reported to emit observed afterglow luminescence over 360 h with the aid of a night-vision monocular in a dark room [17]. We have also measured thermoluminesence (TL) spectra for the samples of Zn3Ga2Ge2O10: 0.5%Cr3+ codoped with null and 3% Ca2+ (Fig. 4(d)). The TL spectrum for the sample with 3% Ca2+ is about 14.5-fold higher than the one with null Ca2+, agreeing well with the result of Fig. 3(b). Moreover, the TL spectrum for 3%Ca2+-doped sample is extremely broad (over a range of 25-250 °C), indicating that the traps distribute over a wide range of energies (thus, a wide range of depths). This broad spectrum is in marked contrast to the narrow one for the Ca2+-free sample with a range of 25-150 °C, further supporting that Ca2+ doping can enhance the trapping sites of the lattice, especially those deep traps with high energies. Moreover, since light irradiation was performed at 400 nm that has energy only able to populate the 4T1(t2e) state instead of the 4T1(te2) close to the bandgap, the TL spectra substantiate the occurrence of quantum tunneling effect between the 4T1(t2e) state and the deep trap sites.

 figure: Fig. 4

Fig. 4 (a) Afterglow decay curves of the samples Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) and Zn2.91Ga1.99Ge2O9.91: 0.5%Cr3+. The samples were charged with 410 nm light for 60s, and then the time-lapsed emission intensity at 698 nm was recorded. (b) The persistent luminescence peak intensity at 698 nm for all the samples after 300 s stoppage of excitation light at 410 nm. (c) Afterglow spectra of Zn3Ga1.99Ge2O10:0.5%Cr3+, 3%Ca2+ (charged with 410 nm for 60s, recorded after 10 s stoppage of light excitation). (d) Thermoluminescence curves monitored at 695 nm over 25–250 °C for Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0 and 3%). The samples were irradiated with 400 nm light for 5 min, and the curves were measured at 2 min after the stoppage of light irradiation.

Download Full Size | PDF

The significantly enhanced persistent luminescence due to Ca2+ doping can also be clearly observed from photographic images of afterglow luminescence (Fig. 5). The persistent luminescence from 0% Ca2+ sample can barely be seen for 10 s, while the 3% Ca2+ sample is able to exhibit clear and discernible persistent luminescence (by iphone 6S) for up to 80 s. It has been demonstrated that divalent magnesium (Mg2+) doping can replace Zn2+ ions in the spinel structure (ZnGa2O4:0.5%Cr3+), and the increase of Mg2+/Zn2+ ratio can promote the formation of antisite defects [22–25]. In analogy, the introduction of Ca2+ ions in the spinel structure of Zn3Ga2Ge2O10:0.5%Cr3+ also favors the formation of antisite defects (the occupation of tetrahedral Zn2+ site with Ga3+ or Ge4+ ions), i.e., the trapping centers, which are beneficial for long-lasting luminescence creation. Moreover, the enhanced structural disorder at the tetrahedral sites also render more distorted octahedral sites, enhancing the number of persistent luminescence emitting centers of Cr3+.

 figure: Fig. 5

Fig. 5 Photographic images for persistent luminescence from the samples of Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) post the stoppage of light excitation at 320 nm at defined time points. Prior to taking these photographs, the samples were exposed to light at 320 nm for 2 mins.

Download Full Size | PDF

The mechanisms to generate NIR persistent luminescence can possibly come from two processes (Fig. 6): (i) The electrons of Cr3+ ions occupying the octahedral sites of spinel crystals can be excited from the ground state 4A2 to the 4T1(te2) state (process1) under high energy UV irradiation. At this state, the excited electron can transmit to the conduct band (CB) and then get trapped by either shallow (process 4) or deep (process 6) traps. When stopping excitation, the electrons will escape from the shallow traps and go to 4T1(te2) via CB band (process 5). Non-radiative transition (process 2) will result in the population of the 2E state, which subsequently releases the trapped energy through a radiative decay to the ground state 4A2, producing NIR luminescence at 698 nm (process 3). (ii) The deep traps can also release the electrons captured by process 6 through quantum tunneling to the 4T1(t2e) and 2E states (process 7). This tunneling process does not need to go through the CB and thus proceeds at a slow rate. Nonradiative decay can depopulate the electrons at the 4T1(t2e) sate to the 2E state (process 8), from which NIR persistent luminescence is produced. Note that akin to UV light, visible light excitation at 410 nm (corresponding to the 4A24T1(t2e) transition of Cr3+) is also able to elicit NIR persistent luminescence from our Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0, 1%, 3%, 5%) samples (Fig. 4). This is possibly because the 410 nm light excitation is able to excite electrons to the 4T1(t2e) state, from which the energy can be transferred through quantum tunneling to the deep traps for storing (a reverse process to 7). Followed by the proceeding of process 7, NIR persistent luminescence will be produced through process 3.

 figure: Fig. 6

Fig. 6 A schematic of the energy transfer mechanism of the persistent luminescence material Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 1%, 3%, 5%)

Download Full Size | PDF

The NIR persistent luminescence intensity and afterglow period are in close relation to the number and distribution of energy trapping centers in the spinel structure. Moreover, both parameters are also affected by the number of luminescence quenching centers existed in the crystal of the resulting powder. Taken together of observed experimental results, the introduction of Ca2+ into the Zn3Ga2Ge2O10:0.5%Cr3+ crystal might produce two beneficial effects on persistent luminescence: (i) The increase of the resulting particle size but without producing additional crystallographic phases, as confirmed by SEM (Fig. 1) and XRD results (Fig. 2). This will reduce the number of luminescence quenching centers in the crystal, which thus enhance the photoluminescence and persistent luminescence intensity. (ii) The substitution of Zn2+ with Ca2+ at the tetrahedral site in the Zn3Ga2Ge2O10:0.5%Cr3+ crystal promotes the formation of antisite defects (the occupation of tetrahedral Zn2+ site with Ga3+ or Ge4+ ions) (Figs. 3 and 4). The antisite defects functionalize as centers to trap electrons and/or holes, preventing from fast recombination of each other. Moreover, the substitution can tailor the local crystal environment surrounding the Cr3+ ions occupying the octadehedral site. The Ca2+-enhanced structural disorder of a spinel crystal results in the production of enhanced long-lasting persistent luminescence.

3. Conclusion

In this work, we systematically investigated the doping effect of divalent calcium (Ca2+) on the NIR persistent phosphors of Cr3+-activated spinel zinc gallogermanate (emission at 650–850 nm). An introduction of 3% Ca2+ ions into the formulated phosphors was able to enhance its NIR persistent luminescence intensity by about 15 fold (observed 300s posted stoppage of light excitation). The Ca2+ doping entails the size increase and the crystallization improvement of the resulting phosphor powder, which can reduce luminescence-quenching centers in the crystal. Moreover, spectral and structural analyses indicate that Ca2+ doping promotes the formation tetrahedral antisite trapping defects surrounding the emitting Cr3+ ions at the octahedral sites, thus enhancing the NIR persistent luminescence profiles. The described simple route here provides a new paradigm to enhance the performance of persistent phosphors engaged in a number of important technological applications.

Funding

National Natural Science Foundation of China (51672061); the Fundamental Research Funds for the Central Universities, China (AUGA5710052614, AUGA8880100415, and HIT.BRETIV.201503).

Acknowledgments

We thank Dr. Zhiguo Zhang at Department of Physics, Harbin Institute of Technology for helping measure the thermoluminesence spectra.

References and links

1. K. Van den Eeckhout, P. F. Smet, and D. Poelman, “Persistent luminescence in Eu2+-doped compounds: a review,” Materials (Basel) 3(4), 2536–2566 (2010). [CrossRef]  

2. K. Van den Eeckhout, D. Poelman, and P. F. Smet, “Persistent luminescence in non-Eu2+-doped compounds: a review,” Materials (Basel) 6(7), 2789–2818 (2013). [CrossRef]  

3. M. Allix, S. Chenu, E. Veron, T. Poumeyrol, E. A. Kouadri-Boudjelthia, S. Alahrache, F. Porcher, D. Massiot, and F. Fayon, “Considerable improvement of long-persistent luminescence in germanium and tin substituted ZnGa2O4,” Chem. Mater. 25(9), 1600–1606 (2013). [CrossRef]  

4. L. Li, Y. H. Wang, H. Li, H. J. Huang, and H. Zhao, “Suppression of photocatalysis and long-lasting luminescence in ZnGa2O4 by Cr3+ doping,” RSC Advances 5(70), 57193–57200 (2015). [CrossRef]  

5. Q. le Masne de Chermont, C. Chanéac, J. Seguin, F. Pellé, S. Maîtrejean, J. P. Jolivet, D. Gourier, M. Bessodes, and D. Scherman, “Nanoprobes with near-infrared persistent luminescence for in vivo imaging,” Proc. Natl. Acad. Sci. U.S.A. 104(22), 9266–9271 (2007). [CrossRef]   [PubMed]  

6. T. Maldiney, A. Lecointre, B. Viana, A. Bessière, M. Bessodes, D. Gourier, C. Richard, and D. Scherman, “Controlling electron trap depth to enhance optical properties of persistent luminescence nanoparticles for in vivo imaging,” J. Am. Chem. Soc. 133(30), 11810–11815 (2011). [CrossRef]   [PubMed]  

7. T. Maldiney, A. Bessière, J. Seguin, E. Teston, S. K. Sharma, B. Viana, A. J. J. Bos, P. Dorenbos, M. Bessodes, D. Gourier, D. Scherman, and C. Richard, “The in vivo activation of persistent nanophosphors for optical imaging of vascularization, tumours and grafted cells,” Nat. Mater. 13(4), 418–426 (2014). [CrossRef]   [PubMed]  

8. H. F. Brito, J. Holsa, T. Laamanen, M. Lastusaari, M. Malkamaki, and L. C. V. Rodrigues, “Persistent luminescence mechanisms: human imagination at work,” Opt. Mater. Express 2(4), 371–381 (2012). [CrossRef]  

9. R. Weissleder and V. Ntziachristos, “Shedding light onto live molecular targets,” Nat. Med. 9(1), 123–128 (2003). [CrossRef]   [PubMed]  

10. X. X. Wang, Z. T. Zhang, Z. L. Tang, and Y. H. Lin, “Characterization and properties of a red and orange Y2O2S-based long afterglow phosphor,” Mater. Chem. Phys. 80(1), 1–5 (2003). [CrossRef]  

11. S. Ye, J. H. Zhang, X. Zhang, S. Z. Lu, X. G. Ren, and X. J. Wang, “Mn2+ activated red phosphorescence in BaMg2Si2O7: Mn2+, Eu2+, Dy3+ through persistent energy transfer,” J. Appl. Phys. 101(6), 063545 (2007). [CrossRef]  

12. H. Yamamoto and T. Matsuzawa, “Mechanism of long phosphorescence of SrAl2O4:Eu2+, Dy3+ and CaAl2O4:Eu2+, Nd3+,” J. Lumin. 72–4, 287–289 (1997). [CrossRef]  

13. T. Matsuzawa, Y. Aoki, N. Takeuchi, and Y. Murayama, “New long phosphorescent phosphor with high brightness, SrAl2O4:Eu2+,Dy3+,” J. Electrochem. Soc. 143(8), 2670–2673 (1996). [CrossRef]  

14. A. A. Kaminskii, A. P. Shkadarevich, B. V. Mill, V. G. Koptev, A. V. Butashin, and A. A. Demidovich, “Tunable Stimulated-Emission of Cr3+ ions and generation frequency self-multiplication effect in acentric crystals of Ca-gallocermanate structure,” Inorg. Mater. 24, 579–581 (1988).

15. D. Jia, L. A. Lewis, and X. J. Wang, “Cr3+-doped lanthanum gallogermanate phosphors with long persistent IR emission,” Electrochem Solid St 13(4), J32–J34 (2010). [CrossRef]  

16. A. Bessière, S. Jacquart, K. Priolkar, A. Lecointre, B. Viana, and D. Gourier, “ZnGa2O4:Cr3+: a new red long-lasting phosphor with high brightness,” Opt. Express 19(11), 10131–10137 (2011). [CrossRef]   [PubMed]  

17. Z. Pan, Y. Y. Lu, and F. Liu, “Sunlight-activated long-persistent luminescence in the near-infrared from Cr3+-doped zinc gallogermanates,” Nat. Mater. 11(1), 58–63 (2011). [CrossRef]   [PubMed]  

18. H. Takasaki, S. Tanabe, and T. Hanada, “Long-lasting afterglow characteristics of Eu, Dy codoped SrO-Al2O3 phosphor,” J. Ceram. Soc. Jpn. 104(1208), 322–326 (1996). [CrossRef]  

19. I. K. Jeong, H. L. Park, and S. I. Mho, “Two self-activated optical centers of blue emission in zinc gallate,” Solid State Commun. 105(3), 179–183 (1998). [CrossRef]  

20. D. R. Li, Y. H. Wang, K. Xu, H. Zhao, and Z. F. Hu, “Persistent luminescent and photocatalytic properties of ZnxGa2O3+x (0.8 <= x <= 1) phosphors,” RSC Advances 5(27), 20972–20975 (2015). [CrossRef]  

21. J. W. Moon, H. S. Moon, E. S. Oh, H. I. Kang, J. S. Kim, H. L. Park, and T. W. Kim, “Dependence of the structural and the optical properties of ZnGa2O4 phosphors on the mixture molar ratio of ZnO and Ga2O3,” Int. J. Inorg. Mater. 3(6), 575–578 (2001). [CrossRef]  

22. N. Basavaraju, S. Sharma, A. Bessiere, B. Viana, D. Gourier, and K. R. Priolkar, “Red persistent luminescence in MgGa2O4: Cr3+; a new phosphor for in vivo imaging,” J. Phys. D Appl. Phys. 46(37), 375401 (2013). [CrossRef]  

23. N. Basavaraju, K. R. Priolkar, D. Gourier, S. K. Sharma, A. Bessière, and B. Viana, “The importance of inversion disorder in the visible light induced persistent luminescence in Cr3+ doped AB2O4 (A = Zn or Mg and B = Ga or Al),” Phys. Chem. Chem. Phys. 17(3), 1790–1799 (2015). [CrossRef]   [PubMed]  

24. N. Basavaraju, K. R. Priolkar, D. Gourier, A. Bessière, and B. Viana, “Order and disorder around Cr3+ in chromium doped persistent luminescent AB(2)O(4) spinels,” Phys. Chem. Chem. Phys. 17(16), 10993–10999 (2015). [CrossRef]   [PubMed]  

25. N. Basavaraju, K. R. Priolkar, A. Bessière, S. K. Sharma, D. Gourier, L. Binet, B. Viana, and S. Emura, “Controlling disorder in the ZnGa2O4:Cr3+ persistent phosphor by Mg2+ substitution,” Phys. Chem. Chem. Phys. 19(2), 1369–1377 (2017). [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 (6)

Fig. 1
Fig. 1 Scanning electron microscopic (SEM) images of the resulting Zn3Ga2Ge2O10:0.5%Cr3+, x%Ca2+ (x = 0, 1, 3, 5) powder samples. Scale bar, 2 µm.
Fig. 2
Fig. 2 X-ray diffraction (XRD) patterns of Zn3Ga2Ge2O10:0.5%Cr3+ codoped with Ca2+ of (d) null, (e) 1%, (f) 3%, and (g) 5%, in reference to the standard ones of (a) ZnGa2O4 (PDF 38-1240), (b) Zn2GeO4 (PDF 11-0687), and (c) GeO2 (PDF 65-8052)
Fig. 3
Fig. 3 (a) Photoluminescence, (b) normalized photoluminescence (normalized to the peak at 716 nm), and (c) excitation spectra of the samples Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) and Zn2.91Ga1.99Ge2O9.91: 0.5%Cr3+. The emission intensity at 698 nm was monitored to acquire the excitation spectra of Fig. 3 (c).
Fig. 4
Fig. 4 (a) Afterglow decay curves of the samples Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) and Zn2.91Ga1.99Ge2O9.91: 0.5%Cr3+. The samples were charged with 410 nm light for 60s, and then the time-lapsed emission intensity at 698 nm was recorded. (b) The persistent luminescence peak intensity at 698 nm for all the samples after 300 s stoppage of excitation light at 410 nm. (c) Afterglow spectra of Zn3Ga1.99Ge2O10:0.5%Cr3+, 3%Ca2+ (charged with 410 nm for 60s, recorded after 10 s stoppage of light excitation). (d) Thermoluminescence curves monitored at 695 nm over 25–250 °C for Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0 and 3%). The samples were irradiated with 400 nm light for 5 min, and the curves were measured at 2 min after the stoppage of light irradiation.
Fig. 5
Fig. 5 Photographic images for persistent luminescence from the samples of Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 0,1%, 3%, 5%) post the stoppage of light excitation at 320 nm at defined time points. Prior to taking these photographs, the samples were exposed to light at 320 nm for 2 mins.
Fig. 6
Fig. 6 A schematic of the energy transfer mechanism of the persistent luminescence material Zn3(1-x)Ga1.99Ge2O10: 0.5%Cr3+, xCa2+ (x = 1%, 3%, 5%)
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.