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

Tunable broadband near-infrared luminescence in glass realized by defect-engineering

Open Access Open Access

Abstract

Tunable broadband near-infrared (NIR)-luminescent materials play a crucial role as light sources and tunable fiber lasers in modern technologies such as high-capacity telecommunication, imaging, and remote sensing. Despite considerable effort in studying the luminescent materials doped with rare-earth or transition metal ions, it is still challenging to achieve tunable broadband emission in photonic materials, especially in glasses, for active-fiber applications. In the present work, such NIR emission is achieved by modifying oxygen-deficient structural defects (i.e., singly ionized oxygen vacancies ($\textrm{V}_\textrm{O}^ \bullet $) in tellurium (Te)-doped germanate glass). The local glass chemistry around Te is controlled by engineering singly ionized oxygen vacancies ($\textrm{V}_\textrm{O}^ \bullet $) in alkali-alumino-germanate glass. This enables fine-tuning of the configurations and chemical states of Te centers over a wide range of chemical states, from ionic states to neutrally charged clusters and to positively charged clusters, resulting in various intriguing luminescent behaviors (e.g., wavelength-tunable emission, great emission enhancement, bandwidth extension).

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Tunable broadband luminescent materials have been of great interest owing to their promising applications in fiber-optic communication, imaging, and remote sensing [13]. With the ever-growing demand for high-capacity telecommunication, it is necessary to develop superior luminescent materials with broadband, efficient, and tunable NIR luminescence, for the next generation of communication systems [3,4]. Extensive studies of the luminescent materials have been concentrating on the glass systems doped with rare-earth (e.g., Pr3+, Nd3+, Tm3+, and Er3+) or transition-metal ions (e.g., Cr3+, Ni2+, Bi°). However, the emission bandwidth in rare-earth (RE) doped glasses is still not satisfying because the f-f transitions of RE ions lead to narrow absorption and emission bands [5]. Furthermore, the f-f transitions spectral bandwidth is relatively independent of glass composition and crystal field strength. Compared to the RE-doped glasses, transition-metal doped glasses exhibit broader NIR emission [68]. Despite great progress in developing a new class of gain media for optical amplifiers, there are still some critical issues that remain unsolved. For example, in Bi-doped glasses, the mechanism of the luminescence of active emission centers has not been really revealed, and this hinders tailoring light absorption and luminescence [9]. For Ni-doped glasses, the polycrystals, in which Ni2+ ions are octahedrally coordinated to their ligands, are required to realize NIR emission [10]. The scattering loss from the glass-crystal interface results in the degradation of spectral performance of fiber amplifiers [11].

As an alternative, Te-doped glasses with broadband NIR luminescence have been developed [12,13]. Unlike the above-mentioned single active ion doped glasses, the broadband luminescence of Te-doped glasses arises from Te clusters (e.g., Te2-, Te4, and Te42+) [1416]. The formation process of clusters involves both the reduction of cations to the atoms and the incorporation of multi-components into a composite entity in an ordered manner, and this process strongly depends on the surrounding ligand field [17]. Moreover, Te can exist as various types of species (from single ions to ions in cluster) and in different valence states (from positive to negative). The conversions can easily occur among different species or among different valence states [1821]. Thus, Te NIR emission (originating from Te clusters) is rather sensitive to the local chemistry (e.g., redox states, defects and stoichiometric deviation) around Te. Both the configuration and valence states of Te clusters are controlled by the local chemistry, and the former factors affect the formation of Te clusters and the optical performances in glasses.

Several approaches have been proposed to adjust local glass chemistry, e.g., by introducing oxygen or hydrogen into glass melts, or by high-energy radiation) [22,23]. High-energy radiation could induce the formation of defects in glasses, leading to an increase in optical loss [24]. It is difficult to control the degree of reduction or oxidation of glass melts just by applying gases since the redox reaction is a spontaneous and rapid process.

In addition to the aforementioned reduction methods, oxygen-deficient structural defects (e.g., E’ centers, i.e., unpaired electrons at network forming cations bonded to three oxygens [25,26]) can affect the local glass chemistry, and thereby modify the physical properties of glasses [27]. This is a powerful tool for adjusting the local glass chemistry, and hence for tuning the optical performances of Te-doped glasses. Here the three key questions arise: how can the oxygen-deficient structural defects be created in Te-doped glasses? What determines the number of structural defects? Is structural defect engineering an effective approach to realize the tunable broadband luminescence in Te-doped glasses? In this study, we answered these questions by taking the potassium-alumino-germanate glass as an example. The introduction of Al2O3 into the [GeO4] tetrahedral network leads to an unstable structural state with higher potential energy in the glass host, and thereby generating structural defects [28,29]. Also, the present study gives insight into the relationship between the structural defects and the optical performances of Te-doped germanate glass, as well as into the mechanism of Te NIR luminescence.

2. Experimental section

2.1 Sample preparation

Glass samples with molar composition of (91.9-x) GeO2-8Al2O3-xK2O-0.1TeO2 (with x = 5, 10, 15, 20, 25, 30, 35, 40) were fabricated by melt-quenching method in air without any annealing. High pure GeO2 and TeO2 (99.99 mol%) and analytical pure reagents Al(OH)3 and K2CO3 were used as raw materials to avoid glass-melt contamination. A batch of ∼30 g for each glass composition was weighed and uniformly mixed in an agate mortar and then melted in high purity corundum crucibles at the temperature range from 1500 to 1580 °C for 30 min. The amount of volatiles is considered to be fixed. The melt was subsequently cast onto a cold brass plate and immediately pressed with another plate to increase the cooling rate and promote glass formation. The obtained bulk glass was then cut, ground, and finely polished for subsequent optical measurement. The thickness of samples is almost the same.

2.2 Material characterization

Structural analyses of all samples were conducted by using a Renishaw InVia Raman spectrometer with a 532 nm laser source at an output power of ∼5 mW for excitation. Structural defect in germanate glasses was detected by electron paramagnetic resonance (EPR) spectra using an X-band spectrometer (ELEXSYS, Bruker A300) with a microwave power of 0.5 mW, and modulation amplitude of 3 G. The transmittance spectra of samples were measured using a Perkin Elmer Lambda-900 Ultraviolet-Visible-Near-infrared (UV-Vis-NIR) spectrophotometer in a spectral range from 200 to 3200 nm. Static excitation spectra and emission spectra were taken using an Edinburgh FLS920 spectrofluorometer equipped with a liquid nitrogen-cooled photomultiplier (Hamamatsu R5509-72). A 450 W xenon lamp was used as excitation sources. NIR luminescence spectra under 808 and 980 nm laser diode (LD) excitation were recorded through a Zolix Omni λ3007 spectrometer equipped with an InGaAs photodetector and an SR830 Stanford Research lock-in amplifier.

3. Results and discussion

Figure 1(a) shows the Raman spectra of the studied glasses, reflecting the evolution of microscopic structure with varying K2O content. As K2O content increases, the strongest band at 420 cm-1 is gradually shadowed by the band at 515 cm-1. The former band originates from the symmetric stretching vibration of Ge-O-Ge bonds in six-GeO4-membered rings (rings compose of six GeO4 tetrahedra), while the latter one arises from that in superstructural units such as four- or three-GeO4-membered rings, respectively [30]. In parallel, the bands around 973 and 860 cm-1, which are induced by the asymmetric stretching of the Ge-O-Ge bonds, shift toward lower frequency [30,31]. It is also seen that a band around 830 cm-1 and a shoulder band around 770 cm-1 develop with increasing K2O content, which are attributed to the asymmetric stretching vibrations of Ge-O- bonds in Q3 and Q2 tetrahedra (where Qn is a tetrahedron with n bridging oxygen), respectively [31]. This trend indicates that the glass network is depolymerized, and a decrease in the size of topological cage (i.e., dissociation of six-GeO4-membered rings into smaller rings). Above 30 mol% K2O, Ge-O- bonds in Q2 tetrahedra lowers network connectivity. In addition, the decrease of network connectivity with K2O content is unfavorable for the formation of structural defects, i.e., positively charged oxygen vacancies, since Ge-O- bonds are charge-balanced and stabilized by K+ cations. In contrast, if there is no K2O present in the structural network, the Al2O3 will act as modifiers and thereby non-bridging Ge-O- bonds will form, which are charge-balanced by Al3+. Yet, such network is energetically unstable, and this is confirmed by the fact that the glass consisting of network-forming oxide and alumina has extremely low glass-forming ability [29]. Thus, oxygen vacancies may easily occur in the unstable network. The presence of these structural defects (i.e., singly ionized oxygen vacancies) is confirmed by an intense EPR peak at g = 1.91 in the K2O free sample (Fig. 1(b)) [32]. It is also seen that the EPR peak gradually diminishes with increasing K2O content to 40 mol%, indicating a decrease and even disappearance of oxygen vacancies, and therefore a less concentration of singly ionized oxygen vacancies (..). Owing to the fact that NIR emission of Te clusters is very sensitive to variations in local chemistry around Te, the topology and structural defect engineering is a powerful approach to control Te cluster configurations and chemical states, and thereby to tune their luminescence behavior in the glass system.

 figure: Fig. 1.

Fig. 1. Structural analyses of the germanate glasses with various K2O contents. (a) Raman spectra; (b) EPR spectra.

Download Full Size | PDF

The inset of Fig. 2(a) shows photographs of the samples with various K2O contents. As K2O content increases from 5 to 20 mol%, glass color changes from reddish-brown to dark-green. When further K2O is added, the dark-green becomes light-green, finally colorless (see the 35 mol% K2O containing sample). Such a trend is shown by the transmission spectra in Fig. 2(a) for the samples with K2O= 5, 10, 15, 20, 25, 30, 35, 40 mol%. As the K2O content increases from 0 to 20 mol%, the strongest absorption peak near 960 nm, originating from nonluminous Te4 (NL-Te4) clusters [33], gets weaker, while the absorption band near 640 nm appears, which is attributed D2h-symmetric Te4 clusters (Fig. 2(b)) [34,35]. This indicates that the NL-Te4 clusters transform into the D2h-symmetric Te4 clusters with increasing K2O content. As K2O content increases to 30 mol%, a new absorption peak near 500 nm emerges, which is accompanied by a decrease of the absorption band near 640 nm, indicating the presence of a new luminescence center. With increasing K2O content from 0 to 30 mol%, the absorption bands at 500, 640, and 960 nm slightly shift towards longer wavelengths (Fig. 2(c)). When K2O increases to 35 mol%, no peak can be observed.

 figure: Fig. 2.

Fig. 2. Control of the configurations and chemical state of Te clusters by varying K2O content. (a) Transmission spectra, Inset: Photographs of glass samples; b) Varying absorption intensity; (c) Varying position of the absorption peak; (d) Excitation spectra monitored at 980 nm of Te-doped glasses with various K2O content compared with 20 mol% K2O-containing glass without Te doping.

Download Full Size | PDF

As shown in Fig. 2(d), no excitation peak can be observed in the 20 mol% K2O-containing glass without Te doping, suggesting that these absorption bands are attributed to Te centers. With an increase of K2O content from 0 to 20 mol%, the excitation band at 330 nm, which has been attributed to C1-symmetric Te5, becomes evidently weaker, whereas the excitation bands at 435 and 645 nm, which are assigned to D2h-symmetric Te4 [34], are enhanced. This suggests that C1-symmetric Te5 is transformed into D2h-symmetric Te4. Further increasing K2O content to 30 mol% leads to weakening of excitation bands at 435 and 645 nm, and the appearance of a new excitation band at 515 nm, indicating that D2h-symmetric Te4 is converted into a new luminous center. When K2O content rises to 35 mol%, all excitation bands weaken and even disappear at 40 mol%.

The strong dependence of the emission intensity on K2O content is consistent with that of excitation band intensity (Fig. 3). The absorption edge of glass matrix without TeO2 is far below that of the Te-doped glass [34]. The absorption edge is beyond 330 and 435 nm excitation band because short wavelength light (below the absorption edge) is absorbed by the Te clusters. Thus, these excitation lights can penetrate the glass matrix and excite Te cluster for emission. Upon 330 nm excitation, the emission peak appears at 855 nm, with an FWHM of 410 nm, in free K2O sample (Fig. 3(a)). With the increase of K2O content, the emission intensity decreases and its FWHM becomes narrower, down to around 384 nm. Upon 435 nm excitation, the emission at 925 nm increases rapidly with K2O content from 15 to 20 mol%, and then decreases with further increasing K2O (i.e., reaching the maximum in sample with 20 mol% K2O) (Figs. 3(b) and (e)). As the K2O content increases from 15 to 35 mol%, the emission peak shifts from 925 to 968 nm and its FWHM broadens from 288 to 322 nm. Upon 645 nm excitation, the variation of luminescence behavior with K2O content is similar to those of the luminescence excited by 435 nm. Specifically, the strongest emission appears in the 20 mol% K2O containing sample, and the peak position shifts from 920 to 968 nm and its FWHM broadens from 288 to 340 nm with increasing K2O content from 15 to 35 mol% (Figs. 3(c) and (e)). Upon 515 nm excitation, as the K2O content increases from 15 to 35 mol%, the emission peak shows a large redshift from 920 to 1026 nm and its FWHM significantly broadens from 288 to 396 nm, while the intensity of luminescence reaches the maximum when the K2O content is increased to 30 mol% (Figs. 3(d) and (e)). Both the redshift and the increased FWHM might result from the emergence of new active Te centers. It is because the emission excited by 515 nm with large FWHM is located longer wavelength. When the excitation center appears, the emission, which is excited by the near wavelength that overlaps with this new excitation band, will shift to longer wavelength and become wider.

 figure: Fig. 3.

Fig. 3. Tuning of the luminescent behavior of germanate glasses by varying K2O content. (a) Emission spectra excited by the 330 nm light; (b) 435 nm; (c) 515 nm; (d) 645 nm; (e) K2O content dependence of the emission peak position, intensity, and full width at half maximum (FWHM); (f) Emission spectra obtained by the 808 nm LD excitation; (g) K2O content dependences of both the peak position and the FWHM for the emission excited by 808 nm LD; (h) Emission spectra obtained by the 980 nm LD excitation.

Download Full Size | PDF

For practical application in fiber laser, Fig. 3(f) illustrates the NIR luminescence spectra of glass samples with different K2O contents upon 808 nm LD excitation. A super-broadband luminescence at 1160 nm with an FWHM of ∼310 nm can be observed in samples with K2O content of 0∼10 mol%. As K2O content increases from 0 to 10 mol%, the C1-symmetric Te5 fraction decrease, and therefore the emission intensity drops (Figs. 3(f) and (g)). This indicates that the luminescence located at 1160 nm is related to active C1-symmetric Te5. With increasing K2O content from 10 to 20 mol%, the emission intensity is greatly enhanced, because of the conversion of C1-symmetric Te5 into D2h-symmetric Te4. In parallel, the emission peak shifts from 1160 to 985 nm. Such a blue-shift indicates an increase in the contribution of the D2h-symmetric Te4 clusters to the emission at 980 nm. Accompanied by the blue-shift of the emission, the FWHM becomes narrowed from 310 to 258 nm, since the contribution of C1-symmetric Te5 becomes less. From 20 to 30 mol% K2O, the emission intensity increases, whereas from 30 to 35 mol% the luminescence decreases and finally vanishes at 40 mol%. When increasing K2O from 20 to 35 mol%, the emission peak shifts from 985 to 1073 nm and its FWHM broadens from 258 to 363 nm. These phenomena are attributed to the appearance of a new cluster center (discussed below).

Upon 980 nm LD excitation, the ultra-broadband emission at 1215 nm with an FWHM of ∼250 nm can be observed in glass samples with K2O=15 to 35 mol% (Fig. 3(h)). From 15 to 20 mol% K2O, the emission intensity is enhanced by four times. A further increase of K2O content, however, leads to a decrease in the emission intensity until NIR luminescence disappears for 40 mol% K2O.

Based on the above results, we propose a scheme to describe the possible mechanism for the stabilization and conversion of active Te cluster centers by engineering structural defects (i.e., singly ionized oxygen vacancies $\textrm{V}_\textrm{O}^ \bullet $) and network topology (Fig. 4). TeO2 as the raw material offers the source to form Te clusters (Fig. 4(a)). In the aluminogermanate glass, 4 oxygen vacancies ($\textrm{V}_\textrm{O}^ \bullet $) give electrons to Te4+ to yield one elemental tellurium (i.e., Te4+ + 4$\textrm{V}_\textrm{O}^ \bullet $= Te°). The six-GeO4-member rings acting as topological cages can accommodate Te atoms to form Te5 emission centers and NL-Te4 (Fig. 4(b)). When K2O content is increased to 20 mol%, the decrease in $\textrm{V}_\textrm{O}^ \bullet $ content weakens the reduction effect. On the other hand, when K2O content is increased, the absorption edge shifts to the lower wavelength signify a stabilization of oxidation states (Fig. 2(a)). The stabilization of oxidation states is favored by increasing optical basicity leading to higher valence states of multivalent metal ions [36,37], and thus less multivalent Te4+ towards Te atoms and aggregate into clusters. Moreover, the six-GeO4-membered rings transform into smaller rings, leading to smaller topological cages, and hence promoting the conversion of larger tellurium clusters (C1-symmetric Te5) into smaller D2h-symmetric Te4 clusters (Fig. 4(c)). This conversion enhances Te NIR emission not only because of the formation of the D2h-symmetric Te4 clusters, but also due to the isolation of the Te clusters from each other. The latter effect weakens the interaction between those clusters, and hence lowers the non-radition transitions (Figs. 3(b), (c), and f). With further increasing K2O content, $\textrm{V}_\textrm{O}^ \bullet $ rapidly decreases and even disappears, thereby suppressing the reduction effect, and hence, leading to the conversion of D2h-symmetric Te4 into higher valence state (Fig. 4(d)). This may result in the enhancement of the NIR emission excited by 515 nm (Fig. 3(d)). Therefore, the emission may originate from high valence D2h-symmetric Te4+ or Te43+, which correspond to the excitation bands at 508 and 518 nm (according to TDDFT calculation), respectively [15]. This agrees with the excitation at 515 nm in our experiment. Above 35 mol% K2O, the network connectivity is drastically lowered, and hence, the luminous Te clusters have a low probability to be confined in the topological cages, leading to the disappearance of Te NIR emission (Figs. 2(d), 3(f) and 3 h). The presence of the new active emission center induced by $\textrm{V}_\textrm{O}^ \bullet $ allows both for extending the tunable emission range and for improving luminescence efficiency.

 figure: Fig. 4.

Fig. 4. Structural defect engineering of the configurations and chemical states of Te clusters. Schematic for the stabilization and conversion of active Te cluster centers with the evolution of structural defect and topology network.

Download Full Size | PDF

4. Conclusions

This work demonstrated the unique correlation between the configuration and chemical state of dopant Te and oxygen-deficient structural defects (i.e., singly ionized oxygen vacancies) in germanate glass. The results made it possible to manipulate local chemistry around Te through engineering structural defects of glass by varying network modifier K2O content. By controlling the local glass chemistry of the glass, the configuration and chemical state of Te centers are finely tuned over a wide range of valence states, from single ions and aggregated clusters to positively charged clusters. The defect engineering could not only improve the optical performances of luminescent materials with efficient emission, but also give new functionalities (e.g., extension in optical amplification bandwidth and improvement of fiber laser wavelength-tunability). Furthermore, the new approach could be applicable to other multivalent dopants in disordered systems in order to tailor a variety of material properties such as optical, magnetic and electrical characteristics.

Funding

K. C. Wong Magna Fund in Ningbo University.; Major Basic Research Cultivation Project of Natural Science Foundation of Guangdong Province (2018B03038009); National Natural Science Foundation of China (61775110).

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

References

1. P. Dragic, T. Hawkins, P. Foy, S. Morris, and J. Ballato, “Sapphire-derived all-glass optical fibres,” Nat. Photonics 6(9), 627–633 (2012). [CrossRef]  

2. M. E. Fermann and I. Hartl, “Ultrafast fibre lasers,” Nat. Photonics 7(11), 868–874 (2013). [CrossRef]  

3. E. M. Dianov, S. Semjonov, and I. A. Bufetov, “New generation of optical fibres,” Quantum Electron. 46(1), 1–10 (2016). [CrossRef]  

4. I. Bufetov and E. Dianov, “Bi-doped fiber lasers,” Laser Phys. Lett. 6(7), 487–504 (2009). [CrossRef]  

5. S. Kang, T. Qiao, X. Huang, C. Yang, X. Liu, J. Qiu, Z. Yang, and G. Dong, “Enhanced CW Lasing and Q-Switched Pulse Generation Enabled by Tm3+-Doped Glass Ceramic Fibers,” Adv. Opt. Mater. 9(3), 2001774 (2021). [CrossRef]  

6. Y. Zhao, L. Wondraczek, A. Mermet, M. Peng, Q. Zhang, and J. Qiu, “Homogeneity of bismuth-distribution in bismuth-doped alkali germanate laser glasses towards superbroad fiber amplifiers,” Opt. Express 23(9), 12423–12433 (2015). [CrossRef]  

7. Y. Yu, Z. Fang, C. Ma, H. Inoue, G. Yang, S. Zheng, D. Chen, Z. Yang, A. Masuno, J. Orava, S. Zhou, and J. Qiu, “Mesoscale engineering of photonic glass for tunable luminescence,” NPG Asia Mater. 8(10), e318 (2016). [CrossRef]  

8. J. Chen, Z. Shi, S. Zhou, Z. Fang, S. Lv, H. Yu, J. Hao, H. Zhang, J. Wang, and J. Qiu, “Local Chemistry Engineering in Doped Photonic Glass for Optical Pulse Generation,” Adv. Opt. Mater. 7(6), 1801413 (2019). [CrossRef]  

9. A. Romanov, E. Haula, D. Shashkin, D. Vtyurina, and V. Korchak, “On the origin of near-IR luminescence in SiO2 glass with bismuth as the single dopant. Formation of the photoluminescent univalent bismuth silanolate by SiO2 surface modification,” J. Lumin. 183, 233–237 (2017). [CrossRef]  

10. S. Zhou, N. Jiang, H. Dong, H. Zeng, J. Hao, and J. Qiu, “Size-induced crystal field parameter change and tunable infrared luminescence in Ni2+-doped high-gallium nanocrystals embedded glass ceramics,” Nanotechnology 19(1), 015702 (2008). [CrossRef]  

11. S. Kang, T. Ouyang, D. Yang, Q. Pan, J. Qiu, and G. Dong, “Enhanced 2 µm Mid-Infrared Laser Output from Tm3+-Activated Glass Ceramic Microcavities,” Laser Photonics Rev. 14(5), 1900396 (2020). [CrossRef]  

12. L. Tan, S. Kang, Z. Pan, Y. Zhang, Y. Yue, S. Xu, M. Peng, and L. Wondraczek, “Topo-chemical tailoring of tellurium quantum dot precipitation from supercooled polyphosphates for broadband optical amplification,” Adv. Opt. Mater. 4(10), 1624–1634 (2016). [CrossRef]  

13. P. Punpai, S. Morimoto, S. Khonthon, Y. Arai, T. Suzuki, and Y. Ohishi, “Effect of carbon addition and TeO2 concentration on NIR luminescent characteristics of TeO2-doped soda-lime-silicate glasses,” J. Non-Cryst. Solids 354(52-54), 5529–5532 (2008). [CrossRef]  

14. S. Khonthon, S. Morimoto, Y. Arai, and Y. Ohishi, “Luminescence characteristics of Te-and Bi-doped glasses and glass-ceramics,” J. Ceram. Soc. Jpn. 115(1340), 259–263 (2007). [CrossRef]  

15. H.-T. Sun, Y. Sakka, N. Shirahata, M. Fujii, and T. Yonezawa, “Near-infrared photoluminescence from molecular crystals containing tellurium,” J. Mater. Chem. 22(47), 24792–24797 (2012). [CrossRef]  

16. L. Tan, J. C. Mauro, S. Xu, Z. Yang, and M. Peng, “Unusual thermal response of tellurium near-infrared luminescence in phosphate laser glass,” Opt. Lett. 43(19), 4823–4826 (2018). [CrossRef]  

17. Q.-M. Wang, Y.-M. Lin, and K.-G. Liu, “Role of Anions Associated with the Formation and Properties of Silver Clusters,” Acc. Chem. Res. 48(6), 1570–1579 (2015). [CrossRef]  

18. J. Beck, “New Forms and Functions of Tellurium: From Polycations to Metal Halide Tellurides,” Angew. Chem. Int. Ed. 33(2), 163–172 (1994). [CrossRef]  

19. B. C. Pan, “Geometric structures, electronic properties, and vibrational frequencies of small tellurium clusters,” Phys. Rev. B 65(8), 085407 (2002). [CrossRef]  

20. J. Beck, “Rings, cages and chains - The rich structural chemistry of the polycations of the chalcogens,” Coordin. Chem. Rev. 163, 55–70 (1997). [CrossRef]  

21. B. V. Korzun, A. A. Fadzeyeva, A. V. Mudryi, and S. Schorr, “Optical absorption and photoluminescence of CuAlTe2,” Phys. Status Solidi C 3(8), 2626–2629 (2006). [CrossRef]  

22. Y. Huang, Z. Lu, and L. Zheng, “Study of SiCl4/H2/O2 chemical kinetics and its application to fused silica glass synthesis,” Combust. Sci. Technol. 190(10), 1861–1885 (2018). [CrossRef]  

23. Q. Q. Chen, H. Song, F. T. Zhang, H. Zhang, Y. Z. Yu, Z. Chen, R. F. Wei, Y. Dai, and J. R. Qiu, “A strategy for fabrication of controllable 3D pattern containing clusters and nanoparticles inside a solid material,” Nanoscale 9(26), 9083–9088 (2017). [CrossRef]  

24. K. M. Davis, K. Miura, N. Sugimoto, and K. Hirao, “Writing waveguides in glass with a femtosecond laser,” Opt. Lett. 21(21), 1729–1731 (1996). [CrossRef]  

25. S. Girard, A. Alessi, N. Richard, L. Martin-Samos, V. De Michele, L. Giacomazzi, S. Agnello, D. Di Francesca, A. Morana, and B. Winkler, “Overview of radiation induced point defects in silica-based optical fibers,” Rev. Phys. 4, 100032 (2019). [CrossRef]  

26. L. Skuja, M. Hirano, H. Hosono, and K. Kajihara, “Defects in oxide glasses,” Phys. Stat. Sol. C 2, 15–24 (2005). [CrossRef]  

27. J. Qiu, K. Miyauchi, Y. Kawamoto, N. Kitamura, J. Qiu, and K. Hirao, “Long-lasting phosphorescence in Sn2+-Cu2+ codoped silicate glass and its high-pressure treatment effect,” Appl. Phys. Lett. 81(3), 394–396 (2002). [CrossRef]  

28. Y. Yue and C. A. Angell, “Clarifying the glass-transition behaviour of water by comparison with hyperquenched inorganic glasses,” Nature 427(6976), 717–720 (2004). [CrossRef]  

29. Y. F. Zhang, D. H. Zhao, and Y. Z. Yue, “Phase transitions and glass transition in a hyperquenched silica-alumina glass,” J. Am. Ceram. Soc. 100(8), 3434–3439 (2017). [CrossRef]  

30. H. Luo, J. Cao, X. Li, X. Wang, and M. Peng, “Topological tailoring of structure and defects to enhance red to near-infrared afterglow from Mn2+-doped germanate photonic glasses,” J. Mater. Chem. C 6(43), 11525–11535 (2018). [CrossRef]  

31. Y. Zhao, M. Peng, A. Mermet, J. Zheng, and J. Qiu, “Precise frequency shift of NIR luminescence from bismuth-doped Ta2O5-GeO2 glass via composition modulation,” J. Mater. Chem. C 2(37), 7830–7835 (2014). [CrossRef]  

32. M. Epifani, J. D. Prades, E. Comini, E. Pellicer, M. Avella, P. Siciliano, G. Faglia, A. Cirera, R. Scotti, and F. Morazzoni, “The role of surface oxygen vacancies in the NO2 sensing properties of SnO2 nanocrystals,” J. Phys. Chem. C 112(49), 19540–19546 (2008). [CrossRef]  

33. P. Hassanzadeh, C. Thompson, and L. Andrews, “Absorption spectra of tellurium clusters in solid argon,” J. Chem. Phys. 96(21), 8246–8249 (1992). [CrossRef]  

34. L. Tan, L. Huang, C. He, J. C. Mauro, M. Peng, X.-B. Yang, and Y. Yue, “Tailoring Cluster Configurations Enables Tunable Broad-Band Luminescence in Glass,” Chem. Mater. 32(19), 8653–8661 (2020). [CrossRef]  

35. L. Tan, L. Huang, and M. Peng, “D2h-Symmetric Tetratellurium Clusters in Silicate Glass as a Broadband NIR Light Source for Spectroscopy Applications,” ACS Appl. Mater. Interfaces 12(46), 51628–51636 (2020). [CrossRef]  

36. J. A. Duffy, “Redox equilibria in glass,” J. Non-Cryst. Solids 196, 45–50 (1996). [CrossRef]  

37. Z. Song, C. Li, Y. Li, Z. Yang, D. Zhou, Z. Yin, X. Wang, Q. Wang, T. M. Hau, Z. Zhao, Y. Yang, X. Yu, and J. Qiu, “The influence of alkali ions size on the superbroadband NIR emission from bismuth-doped alkali aluminoborophosphsilicate glasses,” Opt. Mater. 35(1), 61–64 (2012). [CrossRef]  

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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. Structural analyses of the germanate glasses with various K2O contents. (a) Raman spectra; (b) EPR spectra.
Fig. 2.
Fig. 2. Control of the configurations and chemical state of Te clusters by varying K2O content. (a) Transmission spectra, Inset: Photographs of glass samples; b) Varying absorption intensity; (c) Varying position of the absorption peak; (d) Excitation spectra monitored at 980 nm of Te-doped glasses with various K2O content compared with 20 mol% K2O-containing glass without Te doping.
Fig. 3.
Fig. 3. Tuning of the luminescent behavior of germanate glasses by varying K2O content. (a) Emission spectra excited by the 330 nm light; (b) 435 nm; (c) 515 nm; (d) 645 nm; (e) K2O content dependence of the emission peak position, intensity, and full width at half maximum (FWHM); (f) Emission spectra obtained by the 808 nm LD excitation; (g) K2O content dependences of both the peak position and the FWHM for the emission excited by 808 nm LD; (h) Emission spectra obtained by the 980 nm LD excitation.
Fig. 4.
Fig. 4. Structural defect engineering of the configurations and chemical states of Te clusters. Schematic for the stabilization and conversion of active Te cluster centers with the evolution of structural defect and topology network.
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.