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

Green-emissive Mn-activated nanocrystallized glass with willemite-type Zn2GeO4

Open Access Open Access

Abstract

We fabricated Mn2+-activated nanocrystallized glasses (NCG) with willemite-type Zn2GeO4 from a zincogermanate system glass added by various MnO-concentrations, and examined their photoluminescence (PL) properties. The Mn2+-activated NCGs indicated bright green-emission based on the 4T16A1 transition of the Mn2+, indicating exclusive Mn2+-occupation of Zn site in the Zn2GeO4 nanocrystals. In addition, green long-lasting photoluminescence (LLP) was also observed in the NCGs with low MnO-concentration. Relation between the MnO-concentration and the LLP property was also considered.

©2011 Optical Society of America

1. Introduction

Synthesis and architecture of nanostructured semiconductive crystal have been vigorously investigated due to its extensive application in photonics [15]. Particularly, aiming to produce advanced-photonic component, researchers have made a great effort to fabricate the nanocrystallized glass (NCG) consisting of semiconductive nanocrystals showing various functions, e.g., photoluminescence (PL) and photocatalysis [6,7].

Willemite-type wide-gap-semiconductive oxide, Zn2GeO4, has recently attracted much attention because of its various applicabilities and excellent properties: Since Sato et al. reported an excellent photocatalytic property of the Zn2GeO4 ceramics for water-splitting [8], different nanostructured Zn2GeO4 crystals have been synthesized, which enable to convert CO2 into a hydrocarbon fuel and to apply an anode for Li-ion battery [9,10]. In addition, Mn2+-activated Zn2GeO4 thin film has been largely expected for its display-application since brilliant green electroluminescence was demonstrated [11,12]. Furthermore, Mn2+: Zn2GeO4 nanocrystal is a great candidate for spectral shifter for photovoltaic device (i.e., solar cell), which enable to convert ultraviolet (UV) light to visible light, to improve the conversion efficiency of the cell [13].

Recently, present authors’ group have fabricated transparent NCG with Zn2GeO4 in Li2O-ZnO-GeO2 system, and have demonstrated an bluish PL and anomalous long-lasting photoluminescence (LLP) in the NCG without any emissive dopants [13,14]. The LLP is attributed to presence of an excess of interstitial Zn-defects (Zni ), which are introduced during crystallization process, in the Zn2GeO4 nanocrystals [14]. Several transition-metal ions with non-fulfilled 3d-shell state possess various PL features based on d-d transition, depending on coordination environment. Particularly, Mn2+ (3d 5) is an emissive center able to occupy an equivalence Zn site because ionic radius of Mn2+ is close to that of Zn2+. Elucidation of the Mn2+-activation effect on PL properties in the precursor and resulting NCG is significant for sophisticated glass-ceramic application, such as photoemission and spectral-sifter devices. Therefore, the aim of this study is to fabricate the NCGs with willemite-type Zn2GeO4 activated by different Mn2+-concentrations, and to discuss the activation effects on their PL and LLP properties.

2. Experimental

We employed a zincogermanate glass, which reveals nanocrystallization of willemite-type Zn2GeO4 along with Li2Ge4O9 phase, i.e., 15Li2O-15ZnO-70GeO2 glass, as a base glass [14,15]. The base glasses added by various amount of MnO, i.e., 15Li2O-15ZnO-70GeO2-xMnO (x = 0−1.0) were synthesized as the precursor by a conventional melt-quenching technique (melting condition: 1300°C for 30 min by Pt crucible in air) using commercial powders of reagent grade Li2CO3, ZnO, GeO2, and MnO (purity: ≥ 99.9%). As a preliminary study, we checked the thermal property in the precursors with a differential thermal analysis at a heating rate of 10 K/min, and their glass-transition and crystallization-peak temperatures were situated within T g = 476−483°C and T p = 569−572°C, respectively, suggesting no significant effect of MnO-addition on crystallization behavior. Indeed, the precursors heat-treated at each T p for 1 h showed the crystallization of Zn2GeO4 phase along with Li2Ge4O9 phase, which was confirmed by means of a powder X-ray diffraction analysis, irrespective of the MnO-concentration (x), as previously reported [14,15]. The precursors were annealed at each T g for 1 h to reduce internal stress and then cut into several pieces with an appropriate size. The pieces were heat-treated at the T p for 1 h to obtain the NCG samples.

Field-emission transmission electron microscope (FE-TEM) was utilized for observation of nanometric texture in the NCG. Visual confirmation of PL and afterglow of the precursor and NCG samples were carried out using UV lamps with wavelength of 312 nm. PL and PL excitation (PLE) spectra were studied by a spectrofluorometer with a xenon lamp as an excitation source. Diffuse-reflectance spectra were measured using a UV-visible spectrometer with an attached integrating sphere. PL and LLP properties and quantum yield (QY) were examined referring to [14,16]. For the LLP spectral and decay curve measurements in this study, monitoring of the LLP intensity was started after exposing the sample to excitation source for 30 s. All of experimental measurements in this study were done at room temperature.

3. Results and discussion

In Fig. 1 , we show the result of fabrication of Mn2+-activated precursor glasses and resulting samples subjected to the heat-treatment. As described in experimental section, the MnO-added precursors crystallized the willemite-type Zn2GeO4 along with the Li2Ge4O9 phase, maintaining optical transparency, irrespective of the x [Fig. 1(a)]. For example, the sample with x = 0.3 (as representative) exhibited densely-crystallized texture consisting of crystallites (~50-100 nm) [Fig. 1(b)]. Nanocrystallization of the Zn2GeO4 phase together with Li2Ge4O9 phase was also confirmed by comparison of the electron diffraction pattern to the ICDD data of No. 11-0687 (Zn2GeO4; red rings) and No. 37-1363 (Li2Ge4O9; blue rings). Although the precursors with x = 0.05−0.5 were colorless, those with higher x values (0.8 and 0.1) possessed a dull purple coloration, implying the presence of Mn3+ [17]. In addition, reflectance difference spectrum between the precursors of x = 0 (i.e., base glass) and that of x = 1.0 (i.e., ΔR) revealed broad bands around 3.5 eV and 2.5 eV [Fig. 1(c)]. The former and latter bands could be attributed to the 6A14E, 4T2 transitions of Mn2+ and the 5E→5T2 transition of Mn3+, which is probably occurred by oxidation of Mn2+ of the added MnO, respectively [17,18].

 figure: Fig. 1

Fig. 1 (a) Precursor (glass) samples and the samples subjected to heat-treatment at T p for 1 h (NCG). (b) TEM image of the heat-treated sample with x = 0.3 and a typical selected area electron diffraction (ED) pattern. Blue and red circles correspond to ED rings of the Li2Ge4O9 and Zn2GeO4 phases. White bar corresponds to 200 nm. (c) Reflectance spectra of the precursors with x = 0 and 1.0, and the difference spectrum, ΔR.

Download Full Size | PDF

In Fig. 2 , we show the contour plot of PL spectra in the NCGs. In case of the non-activated NCG (x = 0), an expanse of broad emission was observed in the energy range of ~2.2-3.0 eV and ~3.6-5.1 eV for the PL and PL excitation, respectively [Fig. 2(a)]. This vast and asymmetric emission is due to that the PL observed in the NCG with x = 0 is as a consequence of plural emissive bands/phases, i.e., 2.30 eV and 2.63 eV for the Zn2GeO4 phase, and 2.91 eV for Li2Ge4O9 phase [14]. On the other hand, stronger/sharper emission band around 2.30 eV (~540 nm in visible green region) could be singly detected in the Mn2+-activated NCG (x = 0.3) [Fig. 2(b)]. The emission band was almost identical to that of Mn2+-activated Zn2GeO4 thin-film/nanoparticle so far [12,13]. In addition, dependence of the band position on the excitation energy could not be observed. These clearly indicate that the origin of green emission is extrinsic, i.e., 4T16A1 transition of the Mn2+ with 3d 5-configuration, leading to localized PL. Although this transition is originally a Laporte forbidden, the transition is possible if the Mn2+ occupies a tetrahedral site with Td symmetry. Since the ionic radius of Zn2+ is as large as that of Mn2+ under tetrahedral coordination state [19], the green emission generated from the activated NCG suggests that the Mn2+ exclusively occupies a four-coordinated Zn site in the Zn2GeO4 structure, which consisting of tetrahedral ZnO4 and GeO4 units. Thus, we conclude that the willimite-type Zn2GeO4 nanocrystals in the NCG are activated by the Mn2+.

 figure: Fig. 2

Fig. 2 Contour plots of PL in (a) non-activated (x = 0) and (b) Mn2+-activated NCGs (x = 0.3). The PL spectra, which were measured under excitation at the peak energy of PL excitation spectra (~4.27 eV and ~3.81 eV for the non-activated and activated NCGs, respectively), are also indicated as white curves. Red bands across the plots correspond to secondary wave of excitation.

Download Full Size | PDF

Figure 3 depicts the visual PL features in the precursor and NCG samples under UV-light irradiation at 312 nm. In the precursors, we can see a visible orange/red PL, which is also ascribed to the 4T16A1 transition [Fig. 3(a)], because the activated Mn2+ in the based glass is also situated in oxygen-coordinated environment [17]. After the nanocrystallization, the emission color totally converted into a vivid green in the whole samples. This can be interpreted as coordination/crystal field of the Zn sites in the Zn2GeO4 nanocrystal being stronger than that in the precursor (i.e., glassy sate), in accordance with Tanabe-Sugano diagram [20]. In addition, the maximum of QY in the precursor and NCG samples were estimated to be ~16% and ~19%, respectively (i.e., x = 0.4). As increasing the x, QY of both the precursor and NCG samples slightly increased up to x = 0.4, and subsequently decreased [Fig. 3(b)]. This may be due to the so-called concentration-quenching [21]. In addition, the quenching in the precursor glasses was more significant than that in the NCGs. Since clustering of emissive ions in glasses has fatal effect on their PL properties [22,23], it is deduced that the nanocrystallization suppresses the clustering of Mn2+ in the NCG. Furthermore, as seen in bottom of Fig. 3(a), in the Mn2+-activated NCGs (x = 0.05 and 0.1), green LLP could be confirmed for several minutes by naked eyes in dark after the UV-irradiation was stopped.

 figure: Fig. 3

Fig. 3 (a) Photographs of precursor and NCG samples under UV-irradiation, and of the NCGs after the UV-exposure was stopped (10 s later). (b) QY of the precursor and NCG samples. The QYs were measured under excitation at the peak energies of PL excitation spectra, which were situated within ~3.76-4.00 eV. As representative, chromaticity coordinates of the orange and green emission in the sample with x = 0.3 are (0.545, 0.386) and (0.351, 0.609), respectively.

Download Full Size | PDF

Figure 4 depicts the LLP features of the Mn2+-activated NCGs. In previous study, non-activated NCG possesses a LLP band consisting of broad bands at ~2.3 eV and ~2.6 eV, which are attributed to the intrinsic defects of Zn2GeO4 nanocrystals [13]. With respect to the Mn2+-activated NCG (x = 0.05, as representative), a sharp LLP band at ~2.3 eV (green solid curve) could be obtained by the deconvolution using Gaussian functions, in addition to broad bands at ~2.3 eV and ~2.6 eV (dashed curves) [Fig. 4(a)]. Particularly, the sharp band at ~2.3 eV could be superimposed on the PL band shown in Fig. 2(b). In PL phenomena based on the donor-acceptors recombination, the decay curves are expressed by the equation I(t) ∝ (1 + at) n, where a and n are parameters, and particularly the n value depends on recombination process [24]. The equation is also applicable to the LLP phenomenon [16,25]. In the both non-activated and Mn2+-activated NCGs, the decay curves for the LLP intensities, which were monitored at ~2.3 eV, were well expressed by this equation, and the slopes (or n value) of the activated NCG agreed with that of non-activated NCG (n ~0.6) [Fig. 4(b)]. According to Takahashi et al., the Zni of 15Li2O-15ZnO-70GeO2 glass-ceramics (i.e., non-activated sample), which is produced through nanocrystallization of Zn2GeO4 in the glassy phase, acts as an electron trap for the LLP [14,26]. Thus, taking the LLP features concerning the emission and excitation into account, we strongly suggest that the 4T16A1 transition and Zni are responsible for the emissive and electron-trapping centers, respectively, in the green LLP of the Mn2+-activated NCG. The green LLP phenomenon in Mn2+-activated Zn2GeO4 nanocrystals is peculiar to the glass-ceramic route since there is no report about such LLP in the activated Zn2GeO4 synthesized through different route, i.e, solid-state reaction and spattering techniques [1113,27].

 figure: Fig. 4

Fig. 4 LLP properties in the Mn2+-activated NCG: (a) LLP emission (closed circles) and excitation (open circles) spectra in the NCG with x = 0.05. LLP excitation spectrum of the non-activated NCG (triangles) is also included (cited from [13]). Emission and excitation photon energies for the LLP observations were ~2.31 eV and ~4.13 eV, respectively. These spectra correspond to the LLP after 10 s. (b) Decay curves of LLP in the NCGs with x = 0, 0.05, 0.3, and 1.0. Monitored photon energy was ~2.31 eV. Solid lines correspond to the fitting curves by the equation (see text). Afterglow lifetimes (time at which the initial LLP intensity becomes 1/10) were ~0.3 s, ~0.08 s, ~10−2 s, and ~10−2 s for the x = 0, 0.05, 0.3, and 1.0, respectively.

Download Full Size | PDF

In terms of the LLP excitation spectrum, the excitation band was deconvoluted into large band at ~4.1 eV and medium band at ~4.6 eV. he former band could be superimposed on the LLP excitation band of the non-activated NCG (plotted by triangles, cited from [14], indicating that origin of the latter band is extrinsic. For instance, Takeshita et al. attributed a PL excitation band within ~4.59−5.39 eV to the transition related to the charge-transfer to Mn2+ [13]. In addition, the green-LLP performance considerably declined as increasing the x [Fig. 4(a)], implying that the introduction of excess of the MnO leads to decrease the number of trapping-site or to incapacitate the Zni for electron-trapping. At this moment, we tentatively assume that the decline of green-LLP is associated with presence of the Mn3+, which is increased with the MnO-addition in the precursor glass [Figs. 1(a) and 1(c)].

4. Summary

We presented the green-emissive Mn2+-activated NCGs with willemite-type semiconductive Zn2GeO4. The NCGs showed green-emission based on d-d transition of the Mn2+, which is incorporated at Zn sites of Zn2GeO4 nanocrystals, at room temperature. In addition, the NCGs with low Mn-concentration revealed the visible green LLP. To date, several phosphors have been considered to apply them to spectral shifter material [2830]. We think that following features are required for the advanced spectral shifter phosphor, i.e., i) sun-energy preservation based on LLP process for night generation and ii) rare-earth (RE) free from a view point of RE-resource conservation. Since the NCG in this study actually meets the i) and ii), it also has a potential for the energy/environmental applications.

Acknowledgments

This work was supported by the Ministry of Education, Culture, Sports, Science, and Technology of the Japanese Government. The authors would like to thank Dr. Takamichi Miyazaki and Mr. Kenichiro Iwasaki of Tohoku University for significant contributions to this study.

References and links

1. R. N. Bhargava, D. Gallagher, X. Hong, and A. Nurmikko, “Optical properties of manganese-doped nanocrystals of ZnS,” Phys. Rev. Lett. 72(3), 416–419 (1994). [CrossRef]   [PubMed]  

2. T. Nagata, P. Ahmet, Y. Sakuma, T. Sekiguchi, and T. Chikyow, “GaN nanostructure fabrication by focused-ion-beam-assisted chemical vapor deposition,” Appl. Phys. Lett. 87(1), 013103 (2005). [CrossRef]  

3. R. Nakamura, J. G. Lee, D. Tokozakura, H. Mori, and H. Nakajima, “Formation of hollow ZnO through low temperature oxidation of Zn nanoparticles,” Mater. Lett. 61(4-5), 1060–1063 (2007). [CrossRef]  

4. A. Venkatachalam, B. Klein, J. H. Ryou, S. C. Shen, R. D. Dupuis, and P. D. Yoder, “Design strategies for InGaN-based green lasers,” IEEE J. Quantum Electron. 46(2), 238–245 (2010). [CrossRef]  

5. M. Zhu, S. You, T. Detchprohm, T. Paskova, E. A. Preble, and C. Wetzel, “Various misfit dislocations in green and yellow GaInN/GaN light emitting diodes,” Phys. Status Solidi 207(6), 1305–1308 (2010). [CrossRef]  

6. C. Liu, Y. K. Kwon, and J. Heo, “Near-infrared photoluminescence of PbS QDs precipitated in the glass matrix,” J. Ceram. Soc. Jpn. 116(1358), 1071–1074 (2008). [CrossRef]  

7. R. Nagai, T. Honma, and T. Komatsu, “Laser patterning of ZnO crystals on the surface of borosilicate glass,” J. Am. Ceram. Soc. 93(3), 658–661 (2010). [CrossRef]  

8. J. Sato, H. Kobayashi, K. Ikarashi, N. Saito, H. Nishiyama, and Y. Inoue, “Photocatalytic activity for water decomposition of RuO2-dispersed Zn2GeO4 with d10 configuration,” J. Phys. Chem. B 108(14), 4369–4375 (2004). [CrossRef]  

9. Q. Liu, Y. Zhou, J. Kou, X. Chen, Z. Tian, J. Gao, S. Yan, and Z. Zou, “High-yield synthesis of ultralong and ultrathin Zn2GeO4 nanoribbons toward improved photocatalytic reduction of CO2 into renewable hydrocarbon fuel,” J. Am. Chem. Soc. 132(41), 14385–14387 (2010). [CrossRef]   [PubMed]  

10. J. K. Feng, M. O. Lai, and L. Lu, “Zn2GeO4 nanorods synthesized by low-temperature hydrothermal growth for high-capacity anode of lithium battery,” Electrochem. Commun. 13(3), 287–289 (2011). [CrossRef]  

11. J. P. Bender, J. F. Wager, J. Kissick, B. L. Clark, and D. A. Keszler, “Zn2GeO4:Mn alternating-current thin-film electroluminescent devices,” J. Lumin. 99(4), 311–324 (2002). [CrossRef]  

12. K. H. Yoon and J. H. Kim, “Structure analysis and luminescent study of thin film zinc germanate doped with manganese,” Thin Solid Films 519(5), 1583–1586 (2010). [CrossRef]  

13. S. Takeshita, J. Honda, T. Isobe, T. Sawayama, and S. Niikura, “Size-tunable solvothermal synthesis of Zn2GeO4:Mn2+ nanophosphor in water/diethylene glycol system,” Cryst. Growth Des. 10(10), 4494–4500 (2010). [CrossRef]  

14. Y. Takahashi, M. Ando, K. Iwasaki, H. Masai, and T. Fujiwara, “Defect activation in willemite-type Zn2GeO4 by nanocrystallization,” Appl. Phys. Lett. 97(7), 071906 (2010). [CrossRef]  

15. M. Ando, Y. Takahashi, K. Iwasaki, H. Masai, R. Ihara, and T. Fujiwara, “Fabrication and photoluminescent property of transparent nanocrystallized-glass in Li2O-ZnO-GeO2 system,” IOP Conf. Ser.: Mater. Sci. Eng. (to be published).

16. K. Iwasaki, Y. Takahashi, H. Masai, and T. Fujiwara, “Blue photoluminescence, greenish-blue afterglow and their Ti-concentration dependence in rare-earth free bazirite-type BaZn1-xTixSi3O9,” Opt. Express 17(20), 18054–18062 (2009). [CrossRef]   [PubMed]  

17. M. Kawano, H. Takebe, and M. Kuwabara, “Compositional dependence of the luminescence properties of Mn2+-doped metaphosphate glasses,” Opt. Mater. 32(2), 277–280 (2009). [CrossRef]  

18. N. Da, M. Peng, S. Krolikowski, and L. Wondraczek, “Intense red photoluminescence from Mn2+-doped (Na+; Zn2+) sulfophosphate glasses and glass ceramics as LED converters,” Opt. Express 18(3), 2549–2557 (2010). [CrossRef]   [PubMed]  

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

20. Y. Tanabe and S. Sugano, “On the absorption spectra of complex ions II,” J. Phys. Soc. Jpn. 9(5), 766–779 (1954). [CrossRef]  

21. G. Gao, R. Meszaros, M. Peng, and L. Wondraczek, “Broadband UV-to-green photoconversion in V-doped lithium zinc silicate glasses and glass ceramics,” Opt. Express 19(S3), A312–A318 (2011). [CrossRef]  

22. F. Auzel and P. Goldner, “Towards rare-earth clustering control in doped glasses,” Opt. Mater. 16(1-2), 93–103 (2001). [CrossRef]  

23. A. Chiasera, G. Alombert-Goget, M. Ferrari, S. Berneschi, S. Pelli, B. Boulard, and C. D. Arfuso, “Rare earth-activated glass-ceramic in planar format,” Opt. Eng. 50(7), 071105 (2011). [CrossRef]  

24. D. Ehrentraut, H. Sato, Y. Kagamitani, A. Yoshikawa, T. Fukuda, J. Pejchal, K. Polak, M. Nikl, H. Odaka, K. Hatanaka, and H. Fukumura, “Fabrication and luminescence properties of single-crystalline, homoepitaxial zinc oxide film doped with tri- and tetravalent cations prepared by liquid phase epitaxy,” J. Mater. Chem. 16(33), 3369–3374 (2006). [CrossRef]  

25. P. Zhang, Z. Hong, M. Wang, X. Fang, G. Qian, and Z. Wang, “Luminescence characterization of a new long afterglow phosphor of single Ti-doped Y2O2S,” J. Lumin. 113(1-2), 89–93 (2005). [CrossRef]  

26. Y. Takahashi, M. Ando, R. Ihara, and T. Fujiwara, “Formation of Zn defects in willemite-type Zn2GeO4 during supercooled liquid−crystal phase transition,” Appl. Phys. Lett. 98(22), 221907 (2011). [CrossRef]  

27. T. M. Chen, Y. H. Lee, L. Y. Luo, and N. S. Wang, “Luminescence and time-resolved fluorescence decay of Mn2+-activated Zn2GeO4 phosphors under ultraviolet excitation,” Chinese J. Lumin. 26, 183–188 (2005).

28. T. Maruyama, A. Enomoto, and K. Shirasawa, “Solar cell module colored with fluorescent plate,” Sol. Energy Mater. Sol. Cells 64(3), 269–278 (2000). [CrossRef]  

29. P. Chung, H. H. Chung, and P. H. Holloway, “Phosphor coatings to enhance Si photovoltaic cell performance,” J. Vac. Sci. Technol. A 25(1), 61–66 (2007). [CrossRef]  

30. A. F. Khan, D. Haranath, R. Yadav, S. Singh, S. Chawla, and V. Dutta, “Controlled surface distribution and luminescence of YVO4:Eu3+ nanophosphor layer,” Appl. Phys. Lett. 93(7), 073103 (2008). [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 (4)

Fig. 1
Fig. 1 (a) Precursor (glass) samples and the samples subjected to heat-treatment at T p for 1 h (NCG). (b) TEM image of the heat-treated sample with x = 0.3 and a typical selected area electron diffraction (ED) pattern. Blue and red circles correspond to ED rings of the Li2Ge4O9 and Zn2GeO4 phases. White bar corresponds to 200 nm. (c) Reflectance spectra of the precursors with x = 0 and 1.0, and the difference spectrum, ΔR.
Fig. 2
Fig. 2 Contour plots of PL in (a) non-activated (x = 0) and (b) Mn2+-activated NCGs (x = 0.3). The PL spectra, which were measured under excitation at the peak energy of PL excitation spectra (~4.27 eV and ~3.81 eV for the non-activated and activated NCGs, respectively), are also indicated as white curves. Red bands across the plots correspond to secondary wave of excitation.
Fig. 3
Fig. 3 (a) Photographs of precursor and NCG samples under UV-irradiation, and of the NCGs after the UV-exposure was stopped (10 s later). (b) QY of the precursor and NCG samples. The QYs were measured under excitation at the peak energies of PL excitation spectra, which were situated within ~3.76-4.00 eV. As representative, chromaticity coordinates of the orange and green emission in the sample with x = 0.3 are (0.545, 0.386) and (0.351, 0.609), respectively.
Fig. 4
Fig. 4 LLP properties in the Mn2+-activated NCG: (a) LLP emission (closed circles) and excitation (open circles) spectra in the NCG with x = 0.05. LLP excitation spectrum of the non-activated NCG (triangles) is also included (cited from [13]). Emission and excitation photon energies for the LLP observations were ~2.31 eV and ~4.13 eV, respectively. These spectra correspond to the LLP after 10 s. (b) Decay curves of LLP in the NCGs with x = 0, 0.05, 0.3, and 1.0. Monitored photon energy was ~2.31 eV. Solid lines correspond to the fitting curves by the equation (see text). Afterglow lifetimes (time at which the initial LLP intensity becomes 1/10) were ~0.3 s, ~0.08 s, ~10−2 s, and ~10−2 s for the x = 0, 0.05, 0.3, and 1.0, respectively.
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.