Next Article in Journal
Evaluation of the Microstructure and Mechanical Properties of a New Modified Cast and Laser-Melted AA7075 Alloy
Previous Article in Journal
Femtosecond Laser Fabrication of Stable Hydrophilic and Anti-Corrosive Steel Surfaces
Previous Article in Special Issue
Terahertz Time Domain Spectroscopy of Transformer Insulation Paper after Thermal Aging Intervals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Spectroscopic Properties of Erbium-Doped Oxyfluoride Phospho-Tellurite Glass and Transparent Glass-Ceramic Containing BaF2 Nanocrystals

1
Faculty of Materials Science and Ceramics, AGH University of Science and Technology, Av. Mickiewicza 30, 30059 Krakow, Poland
2
Faculty of Electrical Engineering, Bialystok University of Technology, Wiejska Street 45D, 15351 Bialystok, Poland
3
Faculty of Mechanical Engineering, Bialystok University of Technology, Wiejska Street 45C, 15351 Bialystok, Poland
4
Institute of Chemistry, University of Silesia, Szkolna 9, 40007 Katowice, Poland
*
Author to whom correspondence should be addressed.
Materials 2019, 12(20), 3429; https://doi.org/10.3390/ma12203429
Submission received: 21 August 2019 / Revised: 15 October 2019 / Accepted: 17 October 2019 / Published: 20 October 2019
(This article belongs to the Special Issue Cutting-Edge Research in Nano-Optics)

Abstract

:
The ErF3-doped oxyfluoride phospho-tellurite glasses in the (40-x) TeO2-10P2O5-45 (BaF2-ZnF2) -5Na2O-xErF3 system (where x = 0.25, 0.50, 0.75, 1.00, and 1.25 mol%) have been prepared by the conventional melt-quenching method. The effect of erbium trifluoride addition on thermal, structure, and spectroscopic properties of oxyfluoride phospho-tellurite precursor glass was studied by differential scanning calorimetry (DSC), Fourier-transform infrared (FTIR), and Raman spectroscopy as well as emission measurements, respectively. The DSC curves were used to investigate characteristic temperatures and thermal stability of the precursor glass doped with varying content of ErF3. FTIR and Raman spectra were introduced to characterize the evolution of structure and phonon energy of the glasses. It was found that the addition of ErF3 up to 1.25 mol% into the chemical composition of phospho-tellurite precursor glass enhanced 2.7 µm emission and upconversion. By controlled heat-treatment process of the host glass doped with the highest content of erbium trifluoride (1.25 mol%), transparent erbium-doped phospho-tellurite glass-ceramic (GC) was obtained. X-ray diffraction analysis confirmed the presence of BaF2 nanocrystals with the average 16 nm diameter in a glass matrix. Moreover, MIR, NIR, and UC emissions of the glass-ceramic were discussed in detail and compared to the spectroscopic properties of the glass doped with 1.25 mol% of ErF3 (the base glass).

1. Introduction

Since Wang reported for the first time in the literature the method of obtaining transparent oxyfluoride glass-ceramic doped with Er3+ and Yb3+ ions [1], the dynamic development of research on transparent rare-earth-doped (RED) oxyfluoride glass-ceramics (GC) with fluoride nanocrystals has been noted [2,3,4,5]. As a result of controlled heat treatment of the precursor glass in the SiO2-Al2O3-PbF2-CdF2-YbF3-ErF3 system, Wang and Ohwaki obtained PbxCd1-xF2 nanocrystals. The authors proved that the presence of a fluoride phase in the glass matrix increased the efficiency of upconversion with the participation of Er3+ and Yb3+ co-dopants [1].
Many advantages of fluoride crystals, such as the transparency from ultraviolet to infrared region, a wide range of the energy band gap (greater than 10 eV) and the low phonon energy (300–500 cm−1) make them active centers for the optical rare-earth ions in the glass-ceramic materials [2,3,4,5,6].
The glass-ceramics can be prepared by controlled crystallization of the glasses using different processing methods [7], but still, the common method to obtain GC is crystallization by a controlled heat treatment process. The RED glass-ceramics with nanocrystals in the glass matrix can be obtained as a result of controlled heat treatment above the glass transition temperature of oxide [8,9,10] and oxyfluoride glasses [11,12]. In this method, the rare-earth ions are incorporated into the nanocrystals by the diffusion-controlled process, that is dependent on the temperature. According to the literature [11,12], to obtain the optimum dopant concentration in the nanocrystals and thus the highest luminescence efficiency, the precise control of the rare-earth concentration and crystallization process is required.
Transparent glass-ceramic is a very attractive material for mid-infrared (MIR) application because an active crystal phase in the glass matrix can be obtained [13,14,15]. Among the rare-earth ions used as optically active dopants in the glasses and GC, erbium ion can be recommended as a candidate for 3 µm emission due to the transition of 4I11/24I13/2 [13,14]. The tellurite glasses show the broadband transmission window (0.35–6 um), the high linear and nonlinear refractive indices, the low phonon energy (~750 cm−1) and the high rare-earth solubility. By the addition of suitable metal fluorides with low phonon energy, oxyfluoride tellurite glasses are expected to reduce the possibility of non-radiative transition, which can provide efficient emission and ensure a broad bandwidth [16,17].
Since the discovery of transparent GC in the K2O-Nb2O5-TeO2 system by Shioya et al. [18], tellurite GC has a special scientific interest. These materials combine the high mechanical stability and the high solubility of rare-earth ions of tellurite glass and have a low-phonon-energy environment of the precipitated nanocrystals. The low melt (~900 °C) and crystallization temperatures of the tellurite glasses allowed to easily control the crystal size of glass-ceramic. These advantages make tellurite glass-ceramic as a suitable glass matrix for MIR materials [19].
Not only oxyfluoride tellurite glass-ceramic but also oxyfluoride phosphate glass-ceramic is of particular interest as the hosts for luminescent rare-earth ions. This has been confirmed among others by publications of Petit et al. [20,21,22,23]. Adding P2O5 glass former to the tellurite glass improves its thermal stability and tensile strength. This is due to the network-former mixing effect in the tellurite glasses [24]. The addition of fluorine compounds into the phospho-tellurite glasses can provide important advantages compared to the pure fluoride and pure oxide glasses [24]. Oxyfluoride phospho-tellurite glasses combine the optical properties of fluoride and tellurite glasses with better mechanical and thermal stability of phosphate glasses [25].
Unfortunately, only a few studies have been conducted on oxyfluoride tellurite glass-ceramics [26,27,28]. Rajesh and Camargo [26] proposed oxyfluoride tellurite glasses and glass-ceramics containing NaYF4 nanocrystals doped with variable concentration of Nd3+ in the TeO2-ZnO-YF3-NaF-xNd2O3 (x = 0.2, 0.5, 1.0, 1.5, and 2.0 mol%) system for 1.06 µm emission. Transparent oxyfluoride glass-ceramics with SrF2 nanocrystals have been prepared in the system of TeO2-BaO-Bi2O3-SrF2-2RE2O3 (RE = Eu, Dy) [27]. Transparent oxyfluoride tellurite glasses with the composition of TeO2-SiO2-AlF3-CaO-NaF-yHoF3 (y = 0.5, 1, 1.5, 2.0, 3.0) were synthesized by Hou et al. [28]. The luminescence intensity of the glass and transparent glass-ceramics with β-NaCaAlF6 nanocrystals at 1.53 μm dropped monotonously with the increase of Ho3+ ions concentration under 980 nm LD excitation [28].
Definitely more reports in the literature on oxyfluoride glass-ceramic refers to the silicate glasses. Glass-ceramics prepared by the heat-treated glass from SiO2-Na2CO3-Al2O3-NaF-LuF3-Yb2O3-Er2O3 system was characterized by strong green upconversion fluorescence due to Er3+ doping in Na5Lu9F32 crystallites [29]. The novel nanostructured NdF3 glass-ceramic was obtained by Zhang et al. upon heating of the precursor glass with the composition of 50SiO2-22Al2O3-10NdF3-18NaF (in mol%) as the bandpass color filter for wide-color-gamut white LED [30]. Another example of glass-ceramics obtained from the precursor silicate glass can be Tb3+-doped transparent oxyfluoride glass-ceramics containing LiYF4 [31]. According to the authors, this material may be a novel scintillator applied for X-ray detection for the slow event [31].
The reports on the transparent oxyfluoride nano-glass-ceramics obtained by the annealing process of the silicate glasses can be also found in numerous publications of Durán et al. [11,12,32,33,34]. The results obtained by Duran et al. confirmed the incorporation of rare-earth in the nanocrystals (fluoride compounds) and explained their higher emission efficiency due to the lower phonon energy of the fluoride crystal lattice, which reduces the multi-phonon relaxation rates.
The present work deals with the role of ErF3 on the thermal, structure, and near-, mid-infrared as well as the upconversion emission of host oxyfluoride phospho-tellurite glass. To our best knowledge, no work is reported about oxyfluoride tellurite glasses containing 45 mol% of fluorine compounds in the host glass. In our study, we proposed the host oxyfluoride phospho-tellurite glass with ca. 46 mol% of fluoride compounds. After obtaining the chemically and thermally stable glass, we prepared the glass-ceramic with BaF2 nanocrystals by controlling the heat treatment. No reports of the transparent Er3+-doped BaF2 glass-ceramic in the oxyfluoride phospho-tellurite system exist in the literature. Comparing to the precursor glass, we achieved an enhanced 2.7 µm emission and upconversion from erbium-doped glass-ceramic upon excitation with 976 nm laser diode due to the incorporation of erbium ions into the BaF2 nanocrystals.

2. Experimental Procedure

Oxyfluoride phospho-tellurite glasses with the molar composition (40-x)TeO2-10P2O5-45(BaF2-ZnF2)-5Na2O-xErF3 system (where x = 0.25, 0.50, 0.75, 1.00, and 1.25 mol%), denoted as TP0.25ErF3, TP0.50ErF3, TP0.75ErF3, TP1.00ErF3, TP1.25ErF3, respectively, were prepared by conventional melt-quenching method. High purity (99.99%) raw materials (TeO2, P2O5, BaF2, ZnF2, Na2CO3, and ErF3) were used and each batch of 10 g was well-mixed in an agate mortar, and then melted in a covered platinum crucible at 950 °C for 90 min in ambient air. The melt was cast onto a preheated stainless steel, next annealed at 310 °C for 10 h, and then cooled down to room temperature. All the obtained samples were transparent. The glass samples were cut to the 5 mm × 5 mm size and polished to optical quality before measurements with a 2 mm thickness. Crystallized material was obtained for the addition of 1.5 mol% ErF3. Therefore, the glass with 1.25 mol% ErF3 content has been chosen to obtain a transparent glass-ceramic sample. Based on the recorded DSC curve of the TP1.25ErF3 glass, it was found that the onset crystallization temperature was 390 °C and the significant exothermic crystal peak corresponded to 456 °C. According to this data, the heat treatment temperature of TP1.25ErF3 glass sample was determined from 395 to 460 °C. In order to obtain the BaF2 fluoride crystalline phase in a nanometric scale, the temperature of the controlled heat treatment process of TP1.25ErF3 glass was set at 400 °C for 3 h.
The sample with 1.25 mol% of erbium trifluoride was heat-treated for 3 h at 400 °C to obtain the glass-ceramic.
X-ray diffraction studies were carried out on the X’Pert Pro X-ray diffractometer supplied by PANalytical (Almelo, Netherlands) with Cu Kα1 radiation (λ = 1.54056 Å) in the 2θ range of 10–90°. The step size, time per step, and scan speed were as follows: 0.017°, 184.79 s, and 0.011°/s. The X-ray tube was operated at 40 kV and 40 mA and a scintillation detector (Almelo, Netherlands) were used to measure the intensity of the scattered X-rays. Qualitative identification of the phase composition in the glass-ceramic sample was performed with reference to the ICDD PDF-2 database. From the peak width of the X-ray pattern of the glass-ceramic sample, the crystalline size of crystals can be calculated on the basis of Scherrer‘s equation.
The thermal characteristic temperatures such as the glass transition (Tg), the crystallization (Tc), and melting (Tm) temperatures were measured by using the Jupiter DTA STA 449 F3 thermal analyzer (NETZSCH Thermal Analysis, Selb, Germany), operating in the heat flux DSC mode at a heating rate of 10 °C/min under synthetic air atmosphere. Measurements were carried out with an uncertainty of ±1 °C.
The FTIR spectra of the glasses were obtained with the Fourier spectrometer (Bruker Optics-Vertex70V, Rheinstetten, Germany). The measurements were done using the KBr pellet technique. Absorption spectra were recorded at 128 scans and the resolution of 4 cm−1.
Raman spectra of all samples were obtained using a LabRAM HR spectrometer (HORIBA Jobin Yvon, Palaiseau, France) using the excitation wavelength of 532 nm. The diffraction grating was 1800 lines/mm. The spectra were recorded in several points with the standard spot of about 1 μm.
Raman and FTIR spectra have been normalized and then deconvoluted using Fityk software (0.9.8 software, open-source (GPL2+)). The coefficient of determination (R square) of all the deconvoluted FTIR and Raman spectra was 0.99. The standard deviation of the position and full width at half maximum (FWHM) of each of the component bands was ± 4 cm−1.
The mid-infrared spectra were obtained with the Acton Spectra Pro 2300i monochromator (Princeton Instruments, Trenton, NJ, USA) in the spectral range of 2550–2850 nm using high power laser diode (λexc = 976 nm) as a pump source and PbS detector supported by lock-in-amplifier SR510 (Stanford Research Systems, Sunnyvale, CA, USA). The NIR luminescence was performed using the Acton Spectra Pro 2300i monochromator with InGaAs detector in the range of 1.4–1.7 μm. Both measurements were carried in the transmission mode, where the laser beam was focused on the surface of the glass sample and the luminescence signal was collected on the entrance of monochromator. The upconversion luminescence spectra of the glasses and glass-ceramic were measured in a range of 500–700 nm using the Stelarnet GreenWave monochromator (Stellarnet Inc., Tampa, FL, USA) and the laser diode (λexc = 976 nm). In this part of the experiment the luminescence signal was collected by transmitting optical fiber with NA = 0.5 and 400 µm diameter. The fiber end was located perpendicular to the excitation laser beam. All the measurements were carried out at room temperature. In order to eliminate the strong laser radiation, spectral filters FEL 1250 (Thorlabs, NJ, USA) for NIR and MIR spectra measurements and “heat glass” in the case of UC luminescence were used.

3. Results

3.1. Studies of Erbium-Doped Oxyfluoride Phospho-Tellurite Glass

3.1.1. X-ray Diffraction Analysis

Figure 1 presents X-ray diffraction patterns of the erbium-doped phospho-tellurite glass with the varying content of ErF3 (from 0.25 to 1.25 mol%). The observed diffraction patterns showed the amorphous character of all the samples. The diffractograms confirmed the absence of any crystalline phase, only broad and intense broad humps in the 20–35° range of the two theta angles are observed. X-ray diffraction analysis of the studied glasses shows that all oxyfluoride phospho-tellurite glasses with different contents of ErF3 exhibits similar patterns and the mentioned hump does not become broader with the increasing erbium fluoride content. It suggests the absence of the evolution to a lower degree of the order of the local structure [35].

3.1.2. Thermal Properties of Oxyfluoride Erbium-Doped Phospho-Tellurite Glass

Thermal stability factor of the glass ΔT, defined as a resistance to crystallization during heating, is an important parameter to consider in the manufacture and technological applications of the glass. Thermal stability factor is measured as the difference between the crystallization Tc and the glass transition Tg temperatures. If the value of ΔT is higher than 100 °C, it is assumed that the glass can be considered as thermally stable [36].
Tellurite glasses combine both useful technological properties such as low melting temperatures and good thermal stability [36,37]. The phospho-tellurite glasses formed chemically stable glasses over wide compositional ranges [38,39,40,41,42]. According to the results of Rinke et al., the addition of TeO2 to sodium phosphate glass caused the increase in the glass transition temperature from 284 to 327 °C [43]. The increase in the glass transition temperature with the addition of P2O5 into the chemical composition of tellurite glass was also observed by Nandi and Jose [44]. Phosphorus pentoxide (P2O5) also plays a role as the stabilizing component, influencing the structure and the crystallization behavior in fluoride systems [45,46,47]. The influence of P2O5 on the thermal properties of fluoroindate glasses activated by Pr3+ ions was reported by Pisarska et al. [47]. In this work, authors obtained thermally stable and transparent non-crystalline oxyfluoride glasses with high ΔT factors for low 4–12 mol% of P2O5 concentration. The stability parameter ΔT of the mentioned fluoroindate glasses increased from 114 to 155 °C with an increase in P2O5 concentration [47].
The effect of erbium doping concentration on the thermal properties of tellurite glasses was studied in the literature. These data relate mainly to oxide glasses and report that the increasing amount of erbium ions increases the characteristic temperatures and thermal stability of tellurite glasses [48,49,50]. The transition temperature of the TZNE (TeO2-ZnO-Na2O-Er2O3) glass increased by 13 °C with the increasing Er2O3 content from 0.5 to 2.5 mol% [48]. According to the literature [49], doping the TZN (TeO2-ZnO-Na2O) glass up to 2 mol% of Er2O3 showed the increase of the transition temperature Tg and the thermal stability ΔT. For the above-mentioned concentration of erbium ions, the authors noted a decrease in the thermal stability of the analyzed glass [49]. However, the introduction of Er2O3 from 0.5 to 2.8 mol% resulted in increase of the Tg value and the thermal stability of the TeO2-Li2O-ZnO-Nb2O5-Er2O3 glass (TLZNE) [50].
In order to analyze the effect of ErF3 concentration on the thermal properties of the precursor glass (TP glass), DSC curves of oxyfluoride erbium-doped phospho-tellurite glass samples were measured and shown in Figure 2. Based on obtained DSC curves the characteristic temperatures such as the glass transition Tg (onset), the crystallization Tc (in the maximum), and the melting Tm temperatures were obtained. Based on these parameters the stability factor ΔT was calculated. The results of thermal parameters for different ErF3 concentrations into the chemical composition of the TP glass are presented in Table 1. It can be observed that the glass transition temperature Tg increases from 313 ± 1 to 329 ± 1 °C with the increase of ErF3 concentration into the chemical composition of precursor glass. The same trend was observed in the literature [51]. Jha et al. related the increase of transition temperature Tg by the increase of ErF3 to the strong bonding of erbium ions with non-bringing oxygens, which led to the increase of rigidity of the glass network [51].
All DSC curves showed the existence of exothermic peak/peaks in the maximum in the 430–485 °C range (Figure 2, Table 1), which indicates that partial crystallization has taken place in the oxyfluoride erbium-doped phospho-tellurite glasses. In the DCS curve of glass doping with 0.25 mol% of ErF3 (TP0.25ErF3 glass), the maximum of crystallization peak is located at around 438 ± 1 °C. With the increase of the erbium fluoride concentration to 0.5 mol%, the position of mentioned peak shifted to a higher temperature (458 ± 1 °C) and additionally, the second crystallization peak appeared (at 415 ± 1 °C) – glass TP0.50ErF3. The exothermic peak in the DSC curve of TP0.75ErF3 glass split into two peaks, at 451 ± 1 and 483 ± 1 °C. A wide exothermic peak is visible in the DSC thermal curve of TP1.00ErF3 glass. Related to the DSC curve of TP0.75ErF3 glass, it can be deduced that two exothermic peaks overlapped, therefore the crystallization temperature of TP1.00ErF3 glass has been determined to the maximum of exothermic effect and is 466 ± 1 °C.
For 1.25 mol% of ErF3 glass, two crystallization peaks were observed in the maximum at 420 ± 1 and 456 ± 1 °C, respectively. It is worth noting that the crystallization peak in the DSC curves of TP0.75ErF3, TP1.00ErF3, and TP1.25ErF3 glasses has a greater value of full width at half maximum (FWHM) compared to the FWHM of the exothermic peak at 458 ± 1 °C (TP0.50ErF3) and at 438 ± 1 °C (TP0.25ErF3). The melting temperature Tm seems to be not affected by the increase of ErF3 (Figure 2 and Table 1).
As can be seen in Table 1, oxyfluoride phospho-tellurite TP0.25ErF3, TP0.75ErF3, TP1.00ErF3, and TP1.25ErF3 glasses have a thermal stability > 100 °C (ΔT = 109–142 ± 1 °C), except TP0.50ErF3 glass, which has a smaller value < 100 °C (ΔT = 95 ± 1 °C). Therefore, TP0.25ErF3, TP0.75ErF3, TP1.00ErF3, TP1.25ErF3 glasses should be suitable for optical fiber drawing [52,53,54]. In comparison with other oxyfluoride tellurite glasses, the value of ΔT (case of 1 mol% of ErF3) is larger than that of TeO2-ZnF2-ZnO-Er2O3 glasses (ΔT = 98–126 °C) [55], TeO2-ZnO-La2O3-Tm2O3-Yb2O3 (ΔT = 126–135 °C) [56], TeO2-GeO2-InO2/3-ZnO-KF (ΔT = 129 °C) [57], and TeO2-ZnO-7ZnF2 (ΔT = 120 °C) [51].

3.1.3. Structural Studies of Oxyfluoride Erbium-Doped Phospho-Tellurite Glass

In order to investigate the evolution of erbium-doped phospho-tellurite glass structure, FTIR and Raman measurement were carried out on the precursor glass TP0.25ErF3 doped with varying concentrations of ErF3.
Figure 3 shows the normalized to the band at ~1030 cm−1 FTIR spectra (in the 1300–500 cm−1 range) of the precursor oxyfluoride phospho-tellurite glass with varying ErF3 concentrations. As shown in Figure 3, all spectra presented the typical bands characteristic for tellurite glasses, i.e., at 610, 710, 770 cm−1 [58,59,60] and phosphate glasses, i.e., the bands at 500 cm−1 and in the 1300–900 cm−1 range [61,62,63,64,65,66,67].
On the basis of Figure 3 it could be concluded that the increase in ErF3 does not result in the incorporation of erbium ions in the glass network, because all FTIR spectra are very similar. Most likely, erbium ions act as the network modifier [51]. In order to better understand the influence of erbium trifluoride addition on the oxyfluoride phospho-tellurite glass structure, the deconvolution of the FTIR spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 samples were carried out. Figure 4, Figure 5 and Figure 6 present a deconvolution of the FTIR spectra of samples with the selected value of ErF3. Figure 7 presents a plot of bands position and their integral intensity (area under the curve in %) for deconvoluted TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 spectra. The parameters of deconvoluted spectra and bands assignment to the appropriate vibrations are shown in Table 2 and Table 3, respectively [68,69,70,71,72,73,74,75,76,77].
Bands A at ~545 cm−1 and B at ~566 cm−1 correspond to the deformation vibration of δO–P–O and δO–P–O bonds in Q2 units [63]. The bands C and D at ~606 and ~680 cm−1 are due to the stretching vibrations of Te–O bonds in trigonal bipyramidal units TeO4 (tbp) [69,70]. B and E at ~740 cm−1 relates to the stretching vibrations of trigonal pyramidal units TeO3 (tp) or TeO3+1 polyhedra [70]. B and F at ~780 cm−1 is assigned to the vibration of the continuous network composed of TeO4 and Te–O stretching vibration of TeO3+1 polyhedra [69] or symmetric P–O=P bonds in Q1 units [71]. B and G at ~800 cm−1 is ascribed to the asymmetric stretching vibrations of TeO3 (tp) units or TeO3+1 polyhedra [69].
In the FTIR spectra of phosphate glasses, there are bands associated with Q3–Q0 groups (where Q is the number of bridging bonds) [72,73,74,75,76,77]. B and H at ~870 cm−1 relates to the asymmetric stretching vibrations of Q1 units [72,73]. B and I at ~920 cm−1 is ascribed to the stretching vibrations of P–O–P linked with metaphosphate chain and P–F groups in Q2 units [74,75]. The bands at ~970, ~1020, ~1077/1094/1102 and ~1135/1158/1178 cm−1 are due to the asymmetric stretching vibration of P–O bonds in Q0 units [68], the symmetric stretching vibrations of PO3 groups in Q1 units [74], the asymmetric stretching vibrations of PO2−3 groups in Q1 units [76], and the asymmetric stretching vibrations of non-bridging oxygen in Q2 units [77], respectively.
As can be seen in Figure 4, Figure 5, Figure 6 and Figure 7, Table 2 and Table 3, the sum of the bands intensity at ~1020 and ~1070–1110 cm−1 (both related to Q1 units) increased from 27% to 41%, while the integral intensity of the bands at ~970 cm−1 (related to Q0 units) and at 1130–1180 cm−1 (related to Q2 units) decreased (from 10% to 4% and 15% to 4%, respectively) with increasing ErF3 concentration. This clearly indicates that the addition of ErF3 results in an increase of Q1 units at the expense of Q0 and Q2 units, according to the following equation: Q2 + Q0 ↔ 2Q1 [78]. The P–(O, F)–P bonds are broken and the Er–(O, F)–P (NBO, non-bridging oxygen) bonds are formed as the number of erbium ions increased, which means that the phosphate chain is becoming shorter [79,80].
As seen in Figure 4, Figure 5, Figure 6 and Figure 7, the position of the bands at ~1077 and ~1135 cm−1 shifted to the higher wavenumbers with the increasing ErF3 content in the glass. This might be due to the replacement of fluoride for oxide ions leading to the increase in the bond strength. This confirms the formation of F–P–F bonds [23].
The cut-off frequency of vibrational modes, related to the maximum phonon energy of the analyzed glass network, occurs at ~1180 cm−1 for 1.25 mol% ErF3 concentration (Figure 6). This value is lower than found in the literature for phospho-tellurite glasses [43,44,61].
Figure 8 shows normalized Raman spectra of glasses doped with various erbium trifluoride content in the 1300–500 cm−1 range. The Raman spectra of all samples (Figure 8) have bands that are attributed to the bond vibration occurring in the tellurium (300–800 cm−1) [81,82,83,84,85] as well as phosphate (900–1300 cm−1) [86,87,88,89,90] glass network. Additionally for Raman spectra, it was necessary to deconvolution selected spectra for better understanding the structure of glasses (Figure 9, Figure 10 and Figure 11). Figure 12 presents a plot of Raman bands position and their integral intensity for deconvoluted TP0.25ErF3, TP0.75ErF3 and TP1.25ErF3 spectra. The parameters of deconvoluted Raman spectra and bands assignment to appropriate vibrations are shown in Table 4 and Table 5, respectively.
According to the literature data, the Raman bands in the 360–580 cm−1 range can be attributed to the bending vibration of Te–(O, F)–Te or O,F–Te–O,F bands of [Te(O, F)4] trigonal bipyramidal units [81,82]. The band at around 669/653/645 cm−1can be assigned to the stretching variation of Te–O, F bonds in [Te(O, F)4] units [83]. The band at around 714–704 cm−1is assigned to the Te(O, F)4 tbp units [84]. The band located at 780 cm−1 can be ascribed to Te–O stretching vibration in [TeO3] trigonal pyramids or symmetric stretching vibration in [TeO3+1] units [85].
As reported in the literature [87], the band at around 870–880 cm−1 is due to symmetric stretching vibration of the P–F bonds. As can be seen in Figure 9, Figure 10, Figure 11 and Figure 12 and in Table 4 and Table 5, the integral intensity of this band increased with the addition of erbium trifluoride. This may be due to the replacement of fluorine for oxygen ions leading to the increase in bond strength [88].
All Raman spectra of glasses modified with various ErF3 content do not have bands above 1300 cm−1 (Figure 9), the presence of which is attributed to the Q3 units. This means that there are no Q3 units in the structure of the analyzed glasses [89].
In all deconvoluted Raman spectra of TeP0.25ErF3, TeP0.75ErF3, and TeP1.25ErF3 glasses there is a band at about 950 cm−1, which is attributed to symmetric PO4 stretch on Q0 tetrahedra [25,90]. A decrease in the intensity of this band from 16% ± 3% (TeP0.25ErF3 glass) to 5% ± 3% (TeP1.25ErF3 glass) indicates a reduction in the number of Q0 units as the ErF3 content increases (Table 4 and Table 5). The band at about 1020–1040 cm−1 (Figure 9, Figure 10, Figure 11 and Figure 12) can be attributed to the stretching vibrations of P-O bridging bonds in Q1 units [89]. The integral intensity of this band increased from 2% ± 1% (TeP0.25ErF3 glass) to 5% ± 1% (TeP0.25ErF3 glass) (Table 4 and Table 5). This indicates an increase in Q1 units with the addition of erbium trifluoride. As can be seen in Figure 9, Figure 10, Figure 11 and Figure 12, with the increasing amount of ErF3 there was a decrease in the intensity of the integral band at about 1090–1150 cm−1 (from 3% ± 0.5% for TeP0.25ErF3 glass to 1% ± 0.5% for TeP1.25ErF3 glass), attributed to symmetrical stretching vibrations of non-bridging PO2 bonds of Q2 units [23].
In summary, it can be concluded that with the increase in the content of Er3+ and F ions in the structure of the glasses, the integral intensity of the bands associated with Q0 (band at ~950 cm−1) and Q2 (band at ~1090–1150 cm−1) units decreased at the expense of an increase in the integral intensity of the band associated with Q1 units (band at ~1020–1040 cm−1)—Figure 8, Figure 9, Figure 10, Figure 11 and Figure 12, Table 4 and Table 5. This confirms the conclusions from the interpretation of the FTIR spectra. Fluoride ions are embedded in the glass structure (F replace O anions), while erbium ions depolymerized the phosphate network and phosphate chain is becoming shorter [23,25,89,90].

3.1.4. Spectroscopic Properties of Oxyfluoride Phospho-Tellurite Glass Doped Er3+

In general, the luminescent properties of RE ions in the inorganic glasses depend on the chemical composition of the host glass, the activator concentration (lanthanide ions as optically active dopants), and the excitation power (pumping system). These factors and in particular the oxyfluoride environment (phonon energy) of the active ion significantly affect the emission efficiency of excited states of Er3+ ions in the glasses [91,92,93,94]. It is also discussed in our investigated system glass and GC TP system.
MIR (mid-infrared) emission spectra of oxyfluoride Er3+-doped phospho-tellurite glasses in the 2400–2900 nm range recorded under 976 nm LD excitation at room temperature, is shown in Figure 13. The emission peak at 2725 nm corresponding to 4I11/24I13/2 transition is observed. The changes in the intensity of this peak are characterized for all concentrations of erbium ions in oxyfluoride phosphate-tellurite glass (inset of Figure 13). The 2725 nm emission intensity increased with the increase of ErF3 concentration (Figure 13). A similar effect was observed in tellurite glass [95].
The luminescence spectra of Er3+-doped phospho-tellurite glass recorded at room temperature in the wavelength of 1400–1700 nm under the excitation of 976 nm, is shown in Figure 14. The NIR intensity of emission line at 1532 nm occurred by 4I13/24I15/2 transition increased with the increase of ErF3 concentration up to 1.25 mol% in precursor glass—the inset of Figure 14.
Figure 15 shows the emission spectra of erbium-doped phospho-tellurite glasses obtained in the visible regions under excitation at 976 nm. The observed emission bands are attributed to 2H11/24I15/2 (528 nm), 4S3/24I15/2 (551 nm), and weak emission band at 668 nm to 4F9/24I15/2 transitions of Er3+, respectively. As can be seen in Figure 15, glasses have a strong green emission at 528 and 551 nm and a weak red emission at 668 nm. The green upconversion emission corresponding to 2H11/2,4S3/24I15/2 transition is dominant. This phenomenon can be explained as follows: Two dominant mechanisms are involved in the upconversion process: (1) the excited state absorption (ESA) and (2) energy transfer upconversion (ETU) (Figure 16). In the case of low doping level (0.25 mol% of ErF3 in the precursor glass), erbium ions are scarcely distributed in the glassy phase. The 2H11/2 and 4S3/2 levels, corresponding to the green emission at 525 and 545 nm, are populated by the multiphonon relaxation from the 4F7/2 level, thus the 525 nm emission intensity is reduced and stronger emission signals at 545 nm than that at 525 nm were obtained. For high Er3+ concentration (more than 0.5 mol% of the ErF3) is greatly promoting the ETU1 process of excited Er3+ (4I11/2) from one to another and ultimately excited to 4F7/2 level. Similar behavior was observed in oxyfluoride glass-ceramics containing LaF3 nanocrystals [96].
The inset of Figure 15 shows the relation between the UC luminescence intensity IUP as a function of the pumping power IIR. It is well known that the relation expressed as I   UP   I IR m , where m determines the number of photons of the optical pump used in the conversion of excitation process which occurs in a given structure of the energy levels of RE elements [97].
The mechanism of the upconversion process was determined from the log-log dependence of the emission intensity on the excitation power and presented for the precursor glass doped with the highest content of erbium trifluoride (Figure 17). The slopes of 1.32 and 1.24 for green and red transitions of erbium ions indicates that the 2-photon mechanism was involved in the upconversion luminescence process (Figure 17) [96,98].

3.2. Studies of Glass-Ceramic

3.2.1. X-ray Diffraction Analysis of Transparent Glass-Ceramic

The TP1.25ErF3 glass was heat-treated to induce crystallization and form glass-ceramic. To get a transparent glass-ceramic, the heat treatment temperature was selected to be 400 °C. The sample of TP1.25ErF3 glass was heat-treated at 400 °C for 3 h. The X-ray diffraction pattern of TP1.25ErF3 glass heat-treated at 400 °C is presented in Figure 18. As shown in Figure 1, the precursor phospho-tellurite glass doped with 1.25 mol% of ErF3 is amorphous with no diffraction peaks. After the heat-treated process the diffraction peaks are clearly observed, next to the broad and intense broad hump in the 20–35° range of two theta angle (Figure 18). These peaks are assigned in 100% to barium difluoride (BaF2) cubic phase (JCPDF: 00-001-0533). The size of precipitated crystals in the obtained glass-ceramic was calculated on the basis of Scherrer‘s formula and evaluated to be about 16 nm [99]. The value of the calculated lattice parameter of BaF2 crystals is 0.5899 ± 0.0002 nm and was smaller than the JCPDF value of 0.6187 nm. The similar shrinkage of barium difluoride crystals lattice was observed in the SiO2-ZnF2-BaF2-ErF3-YbF3 glass [99] and can be ascribed to the entrance of erbium ions into BaF2 nanocrystals because the radius of erbium ions is smaller than that of barium ions [100,101,102].

3.2.2. Spectroscopic Properties of the Transparent Glass-Ceramic

Rare-earth doped oxyfluoride glass-ceramics have research interest for attracted combined advantages of low phonon energy of fluoride and very good mechanical and chemical properties of the oxide matrix [103,104]. In the literature many investigations can be found, which have been done on the rare earth ions doped glass-ceramic containing lead or cadmium fluoride nanocrystals [105,106,107,108]. However, cadmium fluoride CdF2 and lead fluoride PbF2 are poisonous raw materials, which limited their application. This is reason that the rare earth doped transparent glass-ceramic containing nanocrystals without the toxic ingredients such as Pb or Cd is still evolving [109,110,111,112,113]. Much effort has been devoted to the search for novel transparent glass-ceramic. The glass-ceramic with LaF3 nanocrystals was reported for the first time by Dejneka as non-toxic glass-ceramic [114]. Next, researchers studied systemically the preparation of the rare-earth-doped glass-ceramic containing lanthanide trifluoride nanocrystals [115,116,117,118,119,120,121,122,123,124].
The rare-earth-doped fluorides such as CaF2, BaF2, or SrF2 in the various glass host have been extensively investigated [125,126,127,128,129,130]. These alkaline-earth fluorides have the same fluorite structure as the β-PbF2. In the alkaline-earth fluorides glass-ceramic, the divalent alkaline-earths cations (Ca2+, Ba2+, or Sr2+) may be substituted by trivalent rare-earth cations [131].
The optical properties of BaF2 make it a universal optical material. Optical properties of BaF2 are related to the structural and electronic properties of barium difluoride such as very large bandgap (11 eV), the low phonon energy (319 cm−1) and the large optical transparency from ultraviolet (UV) to far-infrared (FIR) from 0.2 to 14.0 μm [132,133].
In the research, it may be found that rare-earth ions like Er3+ are the luminescence centers in BaF2 nanocrystals in the glass-ceramic [101,134,135,136]. Chen et al. obtained transparent glass-ceramics containing BaF2 nanocrystals doped with Er3+, prepared by the sol-gel route. The authors recorded the efficient upconversion emissions around 545, 565, and 655 nm under 980 nm excitation due to the lower phonon energy environment of Er3+ ions in glass-ceramic [85]. Transparent glass-ceramic with erbium-doped BaF2 nanocrystals in the fluoroborate system was synthesized for Shinozaki et al. [134]. In this paper, the authors presented that crystallization enhanced the luminescence intensity 30 times compared to the precursor glass [134]. The enhanced upconversion luminescence of the Er3+ ions in transparent oxyfluoride GCs containing BaF2 nanocrystals in the SiO2-ZnF2-BaF2-ErF3 system has been also investigated by Qiao et al. [135]. The erbium-doped germano-gallate oxyfluoride glass-ceramics containing BaF2 nanocrystals with the high transmittance in the mid-infrared region were prepared by Zhao et al. [136]. Unfortunately, to the best of our knowledge, no reports of the transparent Er3+-doped BaF2 glass-ceramic in the oxyfluoride phospho-tellurite exist in the literature. An issue which this report is intended to address.
Figure 19 shows the MIR emission spectra of the TP1.25ErF3 glass and glass-ceramic under 980 nm LD excitation recorded at room temperature. Enhanced emission at a wavelength of 2725 nm was observed in glass-ceramic. The influence of the heat-treatment process on the intensity of MIR luminescence shows that for 3 h of the annealing of the TP1.25ErF3 glass sample, the 30% enhancement was observed for the glass-ceramic (Figure 19).
The higher NIR luminescence intensity exhibited (22% enhancement) when the TP1.25ErF3 glass sample was annealed for 3 h (Figure 20).
In the case of visible emission of Er3+ ions, the glass-ceramic was characterized by the higher intensity at 551, 528, and 668 nm (Figure 21) in comparison to the TP1.25ErF3 glass (Figure 16). It is worth noting that the shape of UC luminescence in the green bands is narrower for the GC sample than the parent glass. Thus, it is confirmed that the structural changes in the vicinity of erbium ions.
The intensity of MIR, NIR, and upconversion emission spectra of glass-ceramic were stronger compared to the intensity emission spectra of the TP1.25ErF3 glass sample. This is most likely due to incorporated erbium ions in the crystalline environment of BaF2 nanocrystals [129,130,131].

4. Discussion

Rare-earth doped oxyfluoride glass-ceramics have been extensively investigated due to potential applications, i.e., in solid laser and fiber amplifiers. Oxyfluoride glass-ceramics are more appropriate for practical applications compared to oxide and fluoride glasses, because the glass-ceramics have the lower phonon energy than the oxide glasses and excellent chemical durability and mechanical strength compared to the fluoride glasses.
Considering the above, authors of this paper first investigated precursor oxyfluoride phospho-tellurite glass in the TeO2-P2O5-BaF2-ZnF2-Na2O system. Addition of the second glass former (P2O5) was aimed at improving the thermal stability and the tensile strength of tellurite glass. Next, oxyfluoride phospho-tellurite precursor glass was doped with erbium trifluoride (ErF3). In this system, thermal and chemical stable glass with 0.25, 0.50, 0.75. 1.00, and 1.25 mol% of ErF3 was obtained. Above 1.25 mol% of ErF3 (for 1.5 mol% of ErF3), we obtained non-transparent material. This showed that the agglomeration of Er3+ ions has occurred.
The thermal, structural, and optical studies were performed for precursor TeO2-P2O5-BaF2-ZnF2-Na2O glass with varying ErF3 concentrations (up to 1.25 mol%). The thermal characteristic of erbium-doped oxyfluoride phospho-tellurite glass provided the information of characteristic temperatures. The glass transition temperature Tg increased from 313 to 329 °C with the increase of ErF3 concentration into the chemical composition of precursor glass. This increase in the transition temperature indicates a more strongly bound network in the glass doped with ErF3. A stronger ionic cross-linking between modifier cations (Er3+) and non-bridging oxygens (NBO) has taken place, which led to the increase of rigidity of the glass network. The existence of clear exothermic peak/peaks in the 430–485 °C range indicated that crystallization occurred in the erbium-doped oxyfluoride phospho-tellurite glasses. The crystalline phase, obtained after 1 h annealing in the maximum of each crystallization peak was only cubic barium difluoride (BaF2). These data were used to synthesis and characterization of transparent glass-ceramics with barium difluoride nanocrystals. Moreover, the oxyfluoride phospho-tellurite glass with various concentration of erbium fluoride can be considered as a good thermal stability. These glasses should be suitable for optical fiber drawing.
FTIR and Raman spectra analysis confirmed the presence of tellurite and phosphate units in the precursor glass doped with ErF3. The replacement of fluoride for oxide leading to an increase in the bond strength confirms the formation of F–P–F bonds when TeO2 is replaced by ErF3. Comparison of deconvoluted spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses indicated that the erbium ions were modifier ions and depolymerized phosphate network. It was found, that the addition of ErF3 resulted in the increase of Q1 phosphate units at the expense of Q0 and Q2 phosphate units. The P–(O, F)–P bonds were broken and the Er–(O, F)–P (NBO, non-bridging oxygen) bonds were formed as the amount of ErF3 increased up to 1.25 mol%. The phosphate chains were becoming shorter. A possible mechanism that explains the structural modifications by Er3+ in the erbium-doped oxyfluoride phospho-tellurite glass can be related to the electronegativity of the structural units in a network of glass. According to Rada and Culea [137] the presence of multiple oxides/fluorides in the glass increases the tendency of network formers to attract oxygen/fluoride ions. This is due to the competition between the cations themselves and this preference. The units, which have higher electronegativity value pick up oxygen or fluoride ions and get modifiers. For erbium trifluoride, the erbium ions are firstly inserted in the trivalent state and they are considered as the modifiers ions due to the strong affinity of Er3+ toward groups containing negative charged non-bridging oxygens. Both tellurium and phosphorus cations in the glass matrix attract fluoride anions, which yield a competition between them. Since phosphorus pentoxide has a higher electronegativity value than tellurium dioxide, P2O5 picked up fluoride ion. Hereby, for 1.25 mol% of ErF3, erbium ions participated in the glass network as the modifier ions for Q2 phosphate units [137].
The spectroscopic study showed, that the emission peak at around 2725 nm corresponding to 4I11/24I13/2 transition was observed in the mid-infrared emission spectra of all doped glasses. The near-infrared emission line at 1533 nm occurred by 4I13/24I15/2 transition. In the emission spectra in the visible regions of erbium-doped phospho-tellurite glasses bands, attributed to 2H11/24I15/2 (528 nm), 4S3/24I15/2 (551 nm), and 4F9/24I15/2 (668 nm) transitions of Er3+ were observed, respectively. The glasses have a strong green emission at 528 and 551 nm and a weak red emission at 668 nm. The green upconversion emission corresponding to 2H11/2,4S3/24I15/2 transition is dominant. The mechanism of the upconversion process was determined from the log-log dependence of emission intensity on the excitation power and indicated that the 2-photon mechanism was involved in the upconversion luminescence process. Finally, the intensity of mid-, near-infrared emission spectra and upconversion of erbium-doped oxyfluoride glass exhibits an increased tendency with the increment of Er3+ concentration until 1.25 mol%. The intensity of the emission peaks has been enhanced with increasing the mol% ErF3 in the precursor glass. It suggests that the observed spectra arisen due to the presence of the erbium trifluoride in the glass and proportional to its concentration.
By annealing of the precursor oxyfluoride glass doping with the highest value of ErF3 (1.25 mol%) for 3 h in 400 °C we obtained oxyfluoride transparent phospho-tellurite glass-ceramic with Er3+-doped barium difluoride nanocrystals with the size of 16 nm. Compared to the literature [55,138,139], in our transparent glass-ceramic, single nanocrystals (100% of BaF2) appeared after the heat treatment.
To check if the Er3+ ions are incorporated into the BaF2 nanocrystals, the mid- and near-infrared emission spectra of the glass-ceramics and upconversion were measured under excitation at 976 nm in room temperature. The glass-ceramic sample has a stronger 2725 nm emission intensity than the base glass. The 30% enhancement of the intensity band at 2725 nm was observed for glass-ceramic. The near-infrared emission of the glass-ceramic that corresponds to Er3+: 4I13/24I15/2 was 22% enhanced compared to the glass. Under 980 nm LD pumping, the green upconversion intensity of Er3+ in the glass-ceramic was observed much stronger than that in the glass. The cubic site of Er3+ and a low vibration frequency of its BaF2 environment resulted in the enhancement of spectroscopic properties of the glass-ceramic compared to the oxyfluoride phospho-tellurite glass with 1.25 mol% of ErF3.
The data presented in this paper first report of the transparent Er3+-doped BaF2 glass-ceramic in the oxyfluoride phospho-tellurite system. The following step will be the investigation of luminescence kinetic of obtained glass-ceramic. It might provide useful information for further development of this material in both RE optimization and composition modifications leading to photonics applications.

5. Conclusions

In summary, Er3+-doped oxyfluoride phospho-tellurite glasses were prepared in the 10P2O5-45(BaF2-ZnF2)-5Na2O-xErF3 system by adding x = 0.25, 0.50, 0.75, 1.00, and 1.25 mol% of ErF3 and characterized for their thermal, structure, and spectroscopic properties. An increase in the glass transition temperature of glasses with the addition of increase of ErF3 into the chemical composition of precursor glass indicated a stronger ionic cross-linking between Er3+ and non-bridging oxygens (NBO) atom. FTIR and Raman analysis confirmed that the addition of ErF3 resulted in an increase of Q1 at the expense of Q0 and Q2 phosphate units. The erbium ions were modifier ions and depolymerized phosphate network. The intensity of the mid-, near-infrared and upconversion emission peaks showed the increase with the increasing mol% of ErF3 in the precursor glass. Transparent oxyfluoride phospho-tellurite glass-ceramics containing cubic BaF2 nanocrystals were synthesized by heat treatment of the base glass containing 1.25 mol% ErF3. The X-ray diffraction (XRD) analysis revealed the formation of 16 nm barium difluoride nanocrystals in the oxyfluoride phospho-tellurite glass-ceramics. Preferential incorporation of erbium ions into the BaF2 nanocrystals was confirmed by the infrared and upconversion emission spectra. The low vibration frequency of erbium-doped BaF2 nanocrystals resulted in the enhancement of spectroscopic properties of the phospho-tellurite glass-ceramic compared to the base glass.

Author Contributions

Conceptualization, M.L. and D.D.; investigation, M.L., J.Z., M.K., P.M., A.B., G.M., M.K.; data curation, M.L.; writing—original draft preparation, M.L.; writing—review and editing, D.D.; project administration, D.D.; funding acquisition, D.D.; formal analysis, J.P., W.A.P.

Funding

The research activity was granted by the National Science Centre, Poland No. 2016/23/B/ST8/00706.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Y.; Ohwaki, J. New transparent vitroceramics codoped with Er3+ and Yb3+ for efficient frequency up-conversion. Appl. Phys. Lett. 1993, 63, 3268–3270. [Google Scholar] [CrossRef]
  2. Guo, Y.; Zeng, H.; Jiang, Y.; Qi, G.; Chen, G.; Chen, J.; Sun, L. Tunable up-conversion emission in Er3+/Yb3+ co-doped oxyfluoride glass ceramics containing NaYF4 nanocrystals by the incorporation of Li+ ions. J. Lumin. 2019, 214, 116524. [Google Scholar] [CrossRef]
  3. Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Zubko, M.; Leiątko, J.; Pisarski, W.A. Structure and luminescent properties of oxyfluoride glass-ceramics with YF3: Eu3+ nanocrystals derived by sol-gel method. J. Eur. Ceram. Soc. 2019, 39, 5010–5017. [Google Scholar] [CrossRef]
  4. Kesavulu, C.R.; Dwaraka Viswanath, C.S.; Karki, S.; Aryal, P.; Kim, H.J.; Jayasankar, C.K. Enhanced visible emissions of Pr3+-doped oxyfluoride transparent glass-ceramics containing SrF2 nanocrystals. Ceram. Int. 2018, 44, 1737–1743. [Google Scholar] [CrossRef]
  5. Ren, P.; Yang, Y.; Zhou, D.; Li, Z.; Qiu, J. Effect of lithium halide on glass network structure and up-conversion luminescence in Er3+ co-doped oxyfluoride glass ceramics containing NaGdF4 nanocrystals. Opt. Mater. 2017, 72, 330–333. [Google Scholar] [CrossRef]
  6. Chen, W.; Cao, J.; Hu, F.; Wei, R.; Chen, L.; Guo, H. Sr2GdF7: Tm3+/Yb3+ glass ceramic: A highly sensitive optical thermometer based on FIR technique. J. Alloys Compd. 2018, 735, 2544–2550. [Google Scholar] [CrossRef]
  7. Deubener, J.; Allix, M.; Davis, M.J.; Duran, A.; Höche, T.; Honma, T.; Komatsu, T.; Krüger, S.; Mitra, I.; Müller, R.; et al. Updated definition of glass-ceramics. J. Non-Cryst. Solids 2018, 501, 3–10. [Google Scholar] [CrossRef]
  8. Zur, L.; Armellini, C.; Belmokhtar, S.; Bouajaj, A.; Cattaruzza, E.; Chiappini, A.; Coccetti, F.; Ferrari, M.; Gonella, F.; Righini, G.C.; et al. Comparison between glass and glass-ceramic silica-hafnia matrices on the down-conversion efficiency of Tb3+/Yb3+ rare earth ions. Opt. Mater. 2019, 87, 102–106. [Google Scholar] [CrossRef]
  9. Bouajaj, A.; Belmokhtar, S.; Britel, M.R.; Armellini, C.; Boulard, B.; Belluomo, F.; Di Stefano, A.; Polizzi, S.; Lukowiak, A.; Ferrari, M.; et al. Tb3+/Yb3+ codoped silica-hafnia glass and glass-ceramic waveguides to improve the efficiency of photovoltaic solar cells. Opt. Mater. 2016, 52, 62–68. [Google Scholar] [CrossRef]
  10. Enrichi, F.; Belmokhtar, S.; Benedetti, A.; Bouajaj, A.; Cattaruzza, E.; Coccetti, F.; Colusso, E.; Ferrari, M.; Ghamgosar, P.; Gonella, F.; et al. Ag nanoaggregates as efficient broadband sensitizers for Tb3+ ions in silica-zirconia ion-exchanged sol-gel glasses and glass-ceramics. Opt. Mater. 2018, 84, 668–674. [Google Scholar] [CrossRef]
  11. Gorni, G.; Balda, R.; Fernández, J.; Pascual, L.; Durán, A.; Pascual, M.J. Effect of the heat treatment on the spectroscopic properties of Er3+-Yb3+-doped transparent oxyfluoride nano-glass-ceramics. J. Lumin. 2018, 193, 51–60. [Google Scholar] [CrossRef]
  12. Gorni, G.; Velázquez, J.J.; Mather, G.C.; Durán, A.; Chen, G.; Sundararajan, M.; Balda, B.; Fernández, J.; Pascual, M.J. Selective excitation in transparent oxyfluoride glass-ceramics doped with Nd3+. J. Eur. Ceram. Soc. 2017, 37, 1695–1706. [Google Scholar] [CrossRef]
  13. Liu, Y.; Liu, X.; Wang, W.; Yu, T.; Zhang, Q. Intense 2.7 μm mid-infrared emission of Er3+ in oxyfluoride glass ceramic containing NaYF4 nanocrystals. Mater. Res. Bull. 2016, 76, 305–310. [Google Scholar] [CrossRef]
  14. Sun, Y.; Yang, Q.; Wang, H.; Zhang, Q.; Shao, Y. Preparation and mid-infrared 2.7 µm luminescence property of high content Er3+-doped (Y0.9La0.1)2O3 transparent ceramics pumped at 980 nm. Ceram. Int. 2018, 44, 1812–1816. [Google Scholar] [CrossRef]
  15. Zhao, Z.; Liu, C.; Xia, M.; Wang, J.; Han, J.; Xie, J.; Zhao, X. Effects of Y3+/Er3+ ratio on the 2.7 μm emission of Er3+ ions in oxyfluoride glass-ceramics. Opt. Mater. 2016, 54, 89–93. [Google Scholar] [CrossRef]
  16. Tan, D.; Sharafudeen, K.N.; Yue, Y.; Qiu, J. Femtosecond laser induced microstructure and luminescence changes in oxyfluoride tellurite glasses. Prog. Mater. Sci. 2016, 76, 154–228. [Google Scholar] [CrossRef]
  17. Feng, Z.; Yang, S.; Xia, H.; Wang, C.; Jiang, D.; Zhang, J.; Gu, X.; Zhang, Y.; Chen, B.; Jiang, H. Investigation of 2.0 μm emission in Tm3+ and Ho3+ co-doped oxyfluoride tellurite glass. Mater. Res. Bull. 2016, 76, 279–283. [Google Scholar] [CrossRef]
  18. Shioya, K.; Komatsu, T.; Kim, H.G.; Sato, R.; Matusita, K. Optical properties of transparent glass-ceramic in K2O-Nb2O5-TeO2 glasses. J. Non-Cryst. Solids 1995, 189, 16–24. [Google Scholar] [CrossRef]
  19. Xing, Z.J.; Liu, Q.X.; Gao, S.; Zhang, Y.; Liao, M.S. ~3 μm fluorescence behavior of Ho3+ doped transparent tellurite glass ceramics. J. Lumin. 2019, 215, 116562. [Google Scholar] [CrossRef]
  20. Nommeots-Nomm, A.; Boetti, N.G.; Salminen, T.; Massera, J.; Hokka, M.; Petit, L. Luminescence of Er3+ doped oxyfluoride phosphate glasses and glass-ceramics. J. Alloys Compd. 2018, 751, 224–230. [Google Scholar] [CrossRef]
  21. Szczodra, A.; Kuusela, L.; Norrbo, I.; Mardoukhi, A.; Hokka, M.; Lastusaari, M.; Petit, L. Successful preparation of fluorine containing glasses with persistent luminescence using the direct doping method. J. Alloys Compd. 2019, 767, 1260–1264. [Google Scholar] [CrossRef]
  22. Sen, R.; Boetti, N.N.G.; Hokka, M.; Petir, L. Optical, structural and luminescence properties of oxyfluoride phosphate glasses and glass-ceramics doped with Yb3+. J. Non-Cryst. Solids 2019, 1, 100003. [Google Scholar] [CrossRef]
  23. Cui, S.; Massera, J.; Lastusaari, M.; Hupa, L.; Petit, L. Novel oxyfluorophosphate glasses and glass-ceramics. J. Non-Cryst. Solids 2016, 445, 40–44. [Google Scholar] [CrossRef]
  24. Ennouri, M.; Jlassi, I.; Habib, E.; Bernard, G. Improvement of spectroscopic properties and luminescence of Er3+ions in phospho-tellurite glass ceramics by formation of ErPO4 nanocrystals. J. Lumin. 2019, 216, 116753. [Google Scholar] [CrossRef]
  25. Möncke, D.; Eckert, H. Review on the structural analysis of fluoride-phosphate and fluoro-phosphate glasses. J. Non-Cryst. Solids 2019, 3, 100026. [Google Scholar] [CrossRef]
  26. Rajesh, D.; de Camargo, A.S.S. Nd3+ doped new oxyfluoro tellurite glasses and glass ceramics containing NaYF4 nano crystals—1.06 µm emission analysis. J. Lumin. 2019, 207, 469–476. [Google Scholar] [CrossRef]
  27. Walas, M.; Lisowska, M.; Lewandowski, T.; Becerro, A.I.; Łapiński, M.; Synak, A.; Sadowski, W.; Kościelska, B. From structure to luminescence investigation of oxyfluoride transparent glasses and glass-ceramics doped with Eu3+/Dy3+ ions. J. Alloys Compd. 2019, 806, 1410–1418. [Google Scholar] [CrossRef]
  28. Hou, Z.X.; Xue, Z.L.; Li, F.; Wang, M.H.; Hu, X.D.; Wang, S.H. Luminescence and up-conversion mechanism of Er3+/Ho3+ co-doped oxyfluoride tellurite glasses and glass–ceramics. J. Alloys Compd. 2013, 577, 523–527. [Google Scholar] [CrossRef]
  29. Deng, Y.; Niu, C. Up-conversion Luminescence Properties of Er3+/Yb3+ Co-Doped Oxyfluoride Glass Ceramic. J. Lumin. 2019, 209, 39–44. [Google Scholar] [CrossRef]
  30. Zhang, L.; Lin, L.; Xiang, X.; Cheng, Y.; Hua, C.; Wang, C.; Lin, S.; Xu, J.; Wang, Y. Nanostructured NdF3 glass ceramic: An efficient bandpass color filter for wide-color-gamut white LED. J. Eur. Ceram. Soc. 2019, 39, 2155–2160. [Google Scholar] [CrossRef]
  31. Gu, Z.; Chen, C.; Zhang, Y. Enhanced luminescence in Tb3+-doped glass-ceramic scintillators containing LiYF4 nanocrystals. Vacuum 2019, 169, 1018832. [Google Scholar] [CrossRef]
  32. Gorni, G.; Pascual, M.J.; Caballero, A.; Velázquez, J.J.; Mosa, M.; Castro, Y.; Durán, A. Crystallization mechanism in sol-gel oxyfluoride glass-ceramics. J. Non-Cryst. Solids 2018, 501, 145–152. [Google Scholar] [CrossRef]
  33. Velázquez, J.J.; Balda, R.; Fernández, J.; Gorni, G.; Pascual, L.; Chen, G.; Sundararajan, M.; Durán, A.; Pascual, M.J. Transparent oxyfluoride glass-ceramics with NaGdF4 nanocrystals doped with Pr3+ and Pr3+-Yb3+. J. Lumin. 2018, 193, 61–69. [Google Scholar] [CrossRef]
  34. Velázquez, J.J.; Mosa, J.; Gorni, G.; Balda, R.; Fernández, J.; Durán, A.; Castro, Y. Novel sol-gel SiO2-NaGdF4 transparent nano-glass-ceramics. J. Non-Cryst. Solids 2019, 520, 119447. [Google Scholar] [CrossRef]
  35. Dehelean, A.; Popa, A.; Rada, S.; Suciu, R.C.; Stan, M.; Culea, E. Spectroscopic investigation of new manganese tellurite glasses synthesized by sol-gel method. J. Alloys Compd. 2019, 801, 181–187. [Google Scholar] [CrossRef]
  36. Tang, W.; Tian, T.; Li, B.; Xu, Y.; Liu, Q.; Zhang, J.; Xu, S. Effect of introduction of TiO2 and GeO2 oxides on thermal stability and 2 μm luminescence properties of tellurite glasses. Ceram. Int. 2019, 45, 16411–16416. [Google Scholar] [CrossRef]
  37. Kut’in, A.M.; Plekhovich, A.D.; Balueva, K.V.; Motorin, S.E.; Dorofeev, V.V. Thermal properties of high purity zinc-tellurite glasses for fiber-optics. Thermochim. Acta 2019, 673, 192–197. [Google Scholar] [CrossRef]
  38. Tuan, T.H.; Asano, K.; Duan, Z.; Liao, M.; Suzuki, T.A.; Ohishi, Y. Novel tellurite-phosphate glass for hybrid microstructured optical fibers. Proc. SPIE 2012, 12, 2598–2601. [Google Scholar]
  39. Elkholy, M.M. Thermoluminescence dosimetry of rare-earth doped tellurite phosphate glasses. Mater. Chem. Phys. 2003, 77, 321–330. [Google Scholar] [CrossRef]
  40. Cheng, T.; Sakai, Y.; Suzuki, T.; Ohishi, Y. Fabrication and characterization of an all-solid tellurite—Phosphate photonic bandgap fiber. Opt. Lett. 2015, 40, 2088–2090. [Google Scholar] [CrossRef]
  41. Duan, Z.; Tong, H.T.; Liao, M.; Cheng, T.; Erwan, M.; Suzuki, T.; Ohishi, Y. New phospho-tellurite glasses with optimization of transition temperature and refractive index for hybrid microstructured optical fibers. Opt. Mater. 2013, 35, 2473–2479. [Google Scholar] [CrossRef]
  42. Suzuki, T.; Shiosaka, T.W.; Miyoshi, S.; Ohishi, Y. Computational and Raman studies of phospho-tellurite glasses as ultrabroad Raman gain media. J. Non-Cryst. Solids 2011, 357, 2702–2707. [Google Scholar] [CrossRef]
  43. Rinke, M.T.; Zhang, L.; Eckert, H. Structural integration of tellurite oxide into mixed-network-former glasses: Connectivity distribution in the system NaPO3-TeO2. Chem. Phys. Chem. 2007, 8, 1988–1998. [Google Scholar] [CrossRef] [PubMed]
  44. Nandi, P.; Jose, G. Spectroscopic properties of Er3+ doped phospho-tellurite glasses. Physica B 2006, 381, 66–72. [Google Scholar] [CrossRef]
  45. Nazabal, V.; Poulain, M.; Olivier, M.; Pirasteh, P.; Camy, P.; Doualan, J.L.; Guy, S.; Djouama, T.; Boutarfaia, A.; Adam, J.L. Fluoride and oxyfluoride glasses for optical applications. J. Fluor. Chem. 2012, 134, 18–23. [Google Scholar] [CrossRef]
  46. Yasui, T.; Hagihara, H.; Inoue, H. The effect of addition of oxides on the crystallization behavior of aluminum fluoride-based glasses. J. Non-Cryst. Solids 1993, 140, 130–133. [Google Scholar] [CrossRef]
  47. Pisarska, J.; Kaczmarczyk, B.; Mazurak, B.; Żelechower, M.; Goryczka, T.; Pisarski, W.A. Influence of P2O5 concentration on structural, thermal and optical behavior of Pr-activated fluoroindate glass. Physica B 2007, 388, 331–336. [Google Scholar] [CrossRef]
  48. Elkhoshkhany, N.; Abbas, R.; Khamis, N.M. Effect of heat treatment on erbium-doped tellurite glass. Mater. Chem. Phys. 2019, 221, 467–476. [Google Scholar] [CrossRef]
  49. Jlassi, I.; Elhouichet, H.; Ferid, M. Thermal and optical properties of tellurite glasses doped erbium. J. Mater. Sci. 2011, 46, 806–812. [Google Scholar] [CrossRef]
  50. Elkhoshkhany, N.; Syala, E.; Yousef, E.S. Kinetics characterization and visible photoluminescence spectroscopy of an erbium-doped tellurite glass. Res. Phys. 2019, 4, 102370. [Google Scholar] [CrossRef]
  51. Nazabal, V.; Todoroki, S.; Nukui, A.; Matsumoto, T.; Suehara, S.; Hondo, T.; Araki, T.; Inoue, S.; Rivero, C.; Cardinal, T. Oxyfluoride tellurite glasses doped with erbium: Thermal analysis, structural organization and spectral properties. J. Non-Cryst. Solids 2003, 325, 85–102. [Google Scholar] [CrossRef]
  52. Elkhoshkhany, N.; Syala, E. Detailed study about the thermal behavior and kinetics characterization of an oxyfluoride tellurite glass. J. Non-Cryst. Solids 2018, 486, 19–26. [Google Scholar] [CrossRef]
  53. Boetti, N.G.; Lousteau, J.; Chiasera, A.; Ferrari, M.; Mura, E.; Scarpignato, G.C.; Abrate, S.; Milanese, D. Thermal stability and spectroscopic properties of erbium-doped niobic-tungsten–tellurite glasses for laser and amplifier devices. J. Lumin. 2012, 132, 1265–1269. [Google Scholar] [CrossRef]
  54. Gebavi, H.; Taccheo, S.; Milanese, D. The enhanced two micron emission in thulium doped tellurite glasses. Opt. Mater. 2013, 35, 1792–1796. [Google Scholar] [CrossRef]
  55. Chillcce, E.F.; Mazali, I.O.; Alves, O.L.; Barbosa, L.C. Optical and physical properties of Er3+-doped oxy-fluoride tellurite glasses. Opt. Mater. 2011, 33, 389–396. [Google Scholar] [CrossRef]
  56. Wu, T.; Cheng, Y.; Zhong, H.; Peng, H.; Hu, J. Thermal Stability and Spectral Properties of Tm3+-Yb3+CO-Doped Tellurite Glasses. Glass Phys. Chem. 2018, 44, 163–169. [Google Scholar] [CrossRef]
  57. Cho, D.H.; Choi, Y.G.; Kim, K.H. Energy transfer from Tm3+: 3F4 to Dy3+: 6H11/2 in oxyfluoride tellurite glasses. Chem. Phys. Lett. 2000, 322, 263–266. [Google Scholar] [CrossRef]
  58. Bürger, H.; Kneipp, K.; Hobert, H.; Vogel, W.; Kozhukharov, V.; Neov, S. Glass formation, properties and structure of glasses in the TeO2-ZnO system. J. Non-Cryst. Solids 1992, 151, 134–142. [Google Scholar] [CrossRef]
  59. Nazabal, V.; Todoroki, S.; Inoue, S.; Matsumoto, T.; Suehara, S.; Hondo, T.; Araki, T.; Cardinal, T. Spectral properties of Er3+ doped oxyfluoride tellurite glasses. J. Non-Cryst. Solids 2003, 326, 359–363. [Google Scholar] [CrossRef]
  60. Çelikbilek Ersundu, M.; Ersundu, A.E. Structure and crystallization kinetics of lithium tellurite glasses. J. Non-Cryst. Solids 2016, 453, 150–157. [Google Scholar] [CrossRef]
  61. Linganna, K.; Narro-García, R.; Desirena, D.; De la Rosa, E.; Basavapoornima, C.H.; Venkatramu, V.; Jayasankar, C.K. Effect of P2O5 addition on structural and luminescence properties of Nd3+-doped tellurite glasses. J. Alloys Compd. 2016, 684, 322–327. [Google Scholar] [CrossRef]
  62. Li, H.; Liu, S.; Zhang, T.; Wu, H.; Guo, S. The evolution of the network structure in tin-fluoro-phosphate glass with increasing temperature. J. Non-Cryst. Solids 2018, 492, 84–93. [Google Scholar] [CrossRef]
  63. Sitarz, M. Influence of modifying cations on the structure and texture of silicate-phosphate glasses. J. Mol. Struct. 2008, 887, 237–248. [Google Scholar] [CrossRef]
  64. Sitarz, M.; Handke, M.; Fojud, Z.; Jurga, S. Spectroscopic studies of glassy phospho-silicate materials. J. Mol. Struct. 2005, 744, 621–626. [Google Scholar] [CrossRef]
  65. Szumera, M.; Wacławska, I.; Sitarz, M.; Mozgawa, W. Spectroscopic study of biologically active glasses. J. Mol. Struct. 2005, 744, 609–614. [Google Scholar] [CrossRef]
  66. Sitarz, M.; Rokita, M.; Handke, M.; Galuskin, E.W. Structural studies of the NaCaPO4-SiO2 sol-gel derived materials. J. Mol. Struct. 2003, 651, 489–498. [Google Scholar] [CrossRef]
  67. Leśniak, M.; Szal, R.; Starzyk, B.; Gajek, M.; Kochanowicz, M.; Żmojda, J.; Dorosz, J.; Sitarz, M.; Dorosz, D. Influence of barium oxide on glass-forming ability and glass stability of the tellurite–phosphate oxide glasses. J. Therm. Anal. Calorim. 2019. [Google Scholar] [CrossRef]
  68. Djouama, T.; Poulain, M.; Bureau, B.; Lebullenger, R. Structural investigation of fluorophosphates glasses by 19F, 31P MAS NMR and IR spectroscopy. J. Non-Cryst. Solids 2015, 414, 16–20. [Google Scholar] [CrossRef]
  69. Elkhoshkhany, N.; Mohamed, H.M. UV–Vis-NIR spectroscopy, structural and thermal properties of novel oxyhalide tellurite glasses with composition TeO2-B2O3-SrCl2-LiF-Bi2O3 for optical application. Res. Phys. 2019, 13, 102222. [Google Scholar] [CrossRef]
  70. El-Mallawany, R.A.H. Tellurite Glasses Handbook; CRC Press: Washington, DC, USA, 2002. [Google Scholar]
  71. Edathazhe, A.B.; Shashikala, H.D. Effect of BaO addition on the structural and mechanical properties of soda lime phosphate glasses. Mat. Chem. Phys. 2016, 184, 146–154. [Google Scholar] [CrossRef]
  72. Kıbrıslı, O.; Ersundu, A.E.; Ersundu, M. Çelikbilek. Dy3+ doped tellurite glasses for solid-state lighting: An investigation through physical, thermal, structural and optical spectroscopy studies. J. Non-Cryst. Solids 2019, 513, 125–136. [Google Scholar] [CrossRef]
  73. Sene, F.F.; Martinelli, J.R.; Gomes, L. Synthesis and characterization of niobium phosphate glasses containing barium and potassium. J. Non-Cryst. Solids 2004, 348, 30–37. [Google Scholar] [CrossRef]
  74. Liu, W.; Sanz, J.; Pecharromán, C.; Sobrados, I.; Lopez-Esteban, S.; Torrecillas, R.; Wang, D.Y.; Moya, J.S.; Cabai, B. Synthesis, characterization and applications of low temperature melting glasses belonging to P2O5-CaO-Na2O system. Ceram. Int. 2019, 45, 12234–12242. [Google Scholar] [CrossRef]
  75. Babu, S.; Seshadri, M.; Reddy, P.V.; Ratnakaram, Y.C. Spectroscopic and laser properties of Er3+ doped fluoro-phosphate glasses as promising candidates for broadband optical fiber lasers and amplifiers. Mater. Res. Bull. 2015, 70, 935–944. [Google Scholar] [CrossRef]
  76. Konidakis, I.; Varsamis, C.P.E.; Kamitsos, E.I.; Möncke, D.; Ehrt, D. Structure and properties of mixed strontium—Manganese metaphosphate glasses. J. Phys. Chem. 2010, 114, 9125–9138. [Google Scholar] [CrossRef]
  77. Sołtys, M.; Pisarska, J.; Leśniak, M.; Sitarz, M.; Pisarski, W.A. Structural and spectroscopic properties of lead phosphate glasses doubly doped with Tb3+ and Eu3+ ions. J. Mol. Struct. 2018, 1163, 418–427. [Google Scholar] [CrossRef]
  78. Moguš-Milanković, A.; Pavić, L.; Reis, S.T.; Day, D.E.; Ivanda, M. Structural and electrical properties of Li2O–ZnO–P2O5 glasses. J. Non-Cryst. Solids 2010, 356, 715–719. [Google Scholar]
  79. Lai, Y.; Liang, X.; Yin, G.; Yang, S.; Wang, J.; Zhu, H.; Yu, H. Infrared spectra of iron phosphate glasses with gadolinium oxide. J. Mol. Struct. 2011, 1004, 188–192. [Google Scholar] [CrossRef]
  80. Shi, Q.; Yue, Y.; Qu, Y.; Liu, S.; Khater, G.A.; Zhang, L.; Zhao, J.; Kang, J. Structure and chemical durability of calcium iron phosphate glasses doped with La2O3 and CeO2. J. Non-Cryst. Solids 2019, 516, 50–55. [Google Scholar] [CrossRef]
  81. Liao, G.; Chem, Q.; Xing, J.; Gebavi, H.; Milanese, D.; Fokine, M.; Ferraris, M. Preparation and characterization of new fluorotellurite glasses for photonics application. J. Non-Cryst. Solids 2009, 355, 447–452. [Google Scholar] [CrossRef]
  82. Kaur, A.; Khanna, A.; Aleksandrov, L.I. Structural, thermal, optical and photo-luminescent properties of barium tellurite glasses doped with rare-earth ions. J. Non-Cryst. Solids 2017, 476, 67–74. [Google Scholar] [CrossRef]
  83. Sidebottom, D.L.; Hruschka, M.; Potter, G.B.; Brow, R.K. Increased radiative lifetime of rare earth-doped zinc oxyhalide tellurite glasses. Appl. Phys. Lett. 1997, 71, 1963–1965. [Google Scholar] [CrossRef]
  84. Hou, G.; Zhang, C.H.; Fu, W.; Li, G.; Xia, J.; Ping, Y. Broadband mid-infrared 2.0 μm and 4.1 μm emission in Ho3+/Yb3+ co-doped tellurite-germanate glasses. J. Lumin. 2020, 217, 116769. [Google Scholar] [CrossRef]
  85. Ozdanowa, J.; Ticha, H.; Tichy, L. Remark on the optical gap in ZnO–Bi2O3–TeO2 glasses. J. Non-Cryst. Solids 2007, 353, 2799–2802. [Google Scholar] [CrossRef]
  86. Jastrzębski, W.; Sitarz, M.; Rokita, M.; Bułat, K. Infrared spectroscopy of different phosphates structures. Spectrochim. Acta Part A 2011, 79, 722–727. [Google Scholar] [CrossRef]
  87. Santos, L.F.; Almeida, R.M.; Tikhomirov, V.K.; Jha, A. Raman spectra and structure of fluoroaluminophosphate glasses. J. Non-Cryst. Solids 2001, 284, 43–48. [Google Scholar] [CrossRef]
  88. Rao, G.V.; Shashikala, H.D. Structural, optical and mechanical properties of ternary CaO-CaF2-P2O5 glasses. J. Adv. Ceram. 2014, 3, 109–116. [Google Scholar]
  89. Stoch, P.; Stoch, A.; Ciecinska, M.; Krakowiak, I.; Sitarz, M. Structure of phosphate and iron-phosphate glasses by DFT calculations and FTIR/Raman spectroscopy. J. Non-Cryst. Solids 2016, 450, 48–60. [Google Scholar] [CrossRef]
  90. Ehrt, E. Structure and properties of fluoride phosphate glasses. SPIE 1992, 1761, 213–222. [Google Scholar]
  91. Mariselvam, K.; Arun Kumar, R.; Rajeswara Rao, V. Concentration-dependence and luminescence studies of erbium doped barium lithium fluoroborate glasses. Opt. Laser Technol. 2019, 118, 37–43. [Google Scholar] [CrossRef]
  92. Ahmed, E.M.; Youssif, M.I.; Elzelaky, A.A. Structural, thermal and photoemission properties of erbium doped phosphate glass. Ceram. Int. 2019. [Google Scholar] [CrossRef]
  93. Azlan, M.N.; Halimah, M.K.; Suriani, A.B.; Azlina, Y.; Umar, S.A.; El-Mallawany, R. Up-conversion properties of erbium nanoparticles doped tellurite glasses for high efficient laser glass. Opt. Commun. 2019, 448, 82–88. [Google Scholar] [CrossRef]
  94. Marcondes, L.M.; Evangelista, R.O.; Gonçalves, R.R.; de Camargo, A.S.; Manzani, D.; Nalin, M.; Castro Cassanjes, F.; Poirier, G.Y. Er3+-doped niobium alkali germanate glasses and glass-ceramics: NIR and visible luminescence properties. J. Non-Cryst. Solids 2019, 521, 119492. [Google Scholar] [CrossRef]
  95. Guo, J.; Liu, X.; Duan, G.; Yang, Y.; Zhao, G.; Huang, F.; Bai, G.; Zhang, J. Optimization by energy transfer process of 2.7 µm emission in highly Er3+-doped tungsten-tellurite glasses. Infrared Phys. Technol. 2019, 99, 49–54. [Google Scholar] [CrossRef]
  96. Hu, Z.; Ma, E.; Wang, Y.; Chen, D. Fluorescence property investigations on Er3+-doped oxyfluoride glass ceramics containing LaF3 nanocrystals. Mat. Chem. Phys. 2006, 100, 308–312. [Google Scholar] [CrossRef]
  97. Zmojda, J.; Kochanowicz, M.; Miluski, P.; Righini, G.C.; Ferrari, M.; Dorosz, D. Investigation of upconversion luminescence in Yb3+/Tm3+/Ho3+ triply doped antimony-germanate glass and double-clad optical fiber. Opt. Mater. 2016, 58, 279–284. [Google Scholar] [CrossRef]
  98. Auzel, F. Up-conversion and Anti-Stokes Processes with f and d Ions in Solids. Chem. Rev. 2004, 104, 139–174. [Google Scholar] [CrossRef]
  99. Qiao, X.; Fan, X.; Wang, M.; Zhang, X. Spectroscopic properties of Er3+-Yb3+ co-doped glass ceramics containing BaF2 nanocrystals. J. Non-Cryst. Solids 2008, 354, 3273–3277. [Google Scholar] [CrossRef]
  100. Li, H.; Kuang, X.Y.; Mao, A.J.; Li, C.G. Studies of EPR spectra and defect structure for Er3+ ions in BaF2 and SrF2 crystals. Spectrochim. Acta Part A 2013, 102, 169–174. [Google Scholar] [CrossRef]
  101. Chen, D.; Wang, Y.; Yu, Y.; Ma, E.; Zhou, L. Microstructure and luminescence of transparent glass ceramic containing Er3+: BaF2 nano-crystals. J. Solid State Chem. 2006, 179, 532–537. [Google Scholar] [CrossRef]
  102. Richman, I. Longitudinal optical phonons in CaF2, SrF2, and BaF2. J. Chem. Phys. 1964, 41, 2836–2837. [Google Scholar] [CrossRef]
  103. Meejitpaisan, P.; Insiripong, S.; Kedkaew, C.; Kim, H.J.; Kaewkhao, J. Radioluminescence and optical studies of gadolinium calcium phosphate oxyfluoride glasses doped with Sm3+. Radiat. Phys. Chem. 2017, 137, 62–67. [Google Scholar] [CrossRef]
  104. Feng, L.; Bian, L.; Ren, W.; Zhang, X.; Li, H. Cooperative up-conversion of Tb3+/Yb3+-codoped oxyfluoride glasses. Mater. Res. Bull. 2017, 89, 263–266. [Google Scholar] [CrossRef]
  105. Li, X.; Zhang, P.; Zhu, S.; Yin, H.; Li, Z.; Chen, Z.; Zheng, Y.; Yu, J. Enhanced 2.75 µm emissions of Er3+ via Eu3+ deactivation in PbF2 crystal. J. Lumin. 2019, 210, 164–168. [Google Scholar] [CrossRef]
  106. Wang, Y.; Jiang, C.; Zhang, Z.; Zhu, S.; Zheng, Y.; Li, Z.; Hanh, Y.; Chen, Z. Bandwidth enhancement of ∼3 μm emission and energy transfer mechanism in Yb3+/Ho3+/Dy3+co-doped PbF2 crystal. J. Lumim. 2019, 212, 160–165. [Google Scholar] [CrossRef]
  107. Zhang, P.; Chen, Z.; Hang, Y.; Li, Z.; Yin, H.; Zhu, S.; Fu, S. 2 μm emission of PbF2 single crystal co-doped with Ho3+/Yb3+ ions. Infrared Phys. Technol. 2017, 82, 178–182. [Google Scholar] [CrossRef]
  108. Kurosawa, S.; Yokota, Y.; Yanagida, T.; Yoshikawa, A. Optical and scintillation property of Ce, Ho and Eu-doped PbF2. Radiat. Meas. 2013, 55, 120–123. [Google Scholar] [CrossRef]
  109. Okada, G.; Shinozaki, K.; Komatsu, T.; Kawano, N.; Kawaguchi, N.; Yanagida, T. Tb3+-doped BaF2-Al2O3-B2O3 glass and glass-ceramic for radiation measurements. J. Non-Cryst. Solids 2018, 501, 111–115. [Google Scholar] [CrossRef]
  110. Qiang, Y.; Pan, Z.; Liang, M.; Xu, J.; Ye, X.; Xia, L.; You, W.; Fu, J.; Ming, H. Highly transparent and color-adjustable Eu2+doped SrO-SiO2-Al2O3 multilayered glass ceramic prepared by controlling crystallization from glass. J. Eur. Ceram. Soc. 2019, 39, 3856–3866. [Google Scholar] [CrossRef]
  111. Weng, F.; Chen, D.; Wang, Y.; Yu, Y.; Huang, P.; Lin, H. Energy transfer and up-conversion luminescence in Er3+/Yb3+ co-doped transparent glass ceramic containing YF3 nano-crystals. Ceram. Int. 2009, 35, 2619–2623. [Google Scholar] [CrossRef]
  112. Zhao, M.; Zhang, H.; Zou, X.; Jia, W.; Su, C. Effect of microstructure on up-conversion luminescent of Tb3+/Yb3+ co-doped phosphate glass and glass-ceramic. Mater. Lett. 2019, 243, 73–76. [Google Scholar] [CrossRef]
  113. Panyata, S.; Eitssayeam, S.; Tunkasiri, T.; Munpakdee, A.; Pengpat, K. Crystallization kinetic of Er3+-doped BiO1.5-GeO2-BO1.5 glass-ceramic. Ceram. Int. 2018, 44, S46–S49. [Google Scholar] [CrossRef]
  114. Dejneka, M.J. The luminescence and structure of novel transparent oxyfluoride glass-ceramics. J. Non-Cryst. Solids 1998, 239, 149–155. [Google Scholar] [CrossRef]
  115. Wang, X.J.; Huang, S.H.; Reeves, R.; Wells, W.; Dejneka, M.J.; Meltzer, R.S.; Yen, W.M. Studies of the spectroscopic properties of Pr3+doped LaF3 nanocrystals/glass. J. Lumin. 2001, 94, 229–233. [Google Scholar] [CrossRef]
  116. Biswas, A.; Maciel, G.S.; Friend, C.S.; Prasad, P.N. Up-conversion properties of a transparent Er3+–Yb3+ co-doped LaF3–SiO2 glass-ceramics prepared by sol–gel method. J. Non-Cryst. Solids 2003, 316, 393–397. [Google Scholar] [CrossRef]
  117. Wang, J.; Qiao, X.; Fan, X.; Wang, M. Up-conversion and near-infrared emission of Er3+ doped transparent glass ceramics containing LaF3 nanocrystals. Physica B 2004, 353, 242–247. [Google Scholar] [CrossRef]
  118. Wang, J.; Qiao, X.; Fan, X.; Wang, M. Preparation and Luminescence of Er3+ Doped Oxyfluoride Glass Ceramics Containing LaF3Nanocrystals. J. Rare Earths 2006, 24, 67–71. [Google Scholar] [CrossRef]
  119. Chen, D.; Yu, Y.; Huang, P.; Lin, H.; Shan, Z.; Wang, Y. Color-tunable luminescence of Eu3+ in LaF3 embedded nanocomposite for light emitting diode. Acta Mater. 2010, 58, 3035–3041. [Google Scholar] [CrossRef]
  120. Velázquez, J.J.; Rodríguez, V.D.; Yanes, A.C.; del-Castillo, J.; Méndez-Ramos, J. Down-shifting in Ce3+–Tb3+ co-doped SiO2–LaF3 nano-glass—Ceramics for photon conversion in solar cells. Opt. Mater. 2012, 34, 1994–1997. [Google Scholar] [CrossRef]
  121. Zhang, C.; Zhao, S.; Deng, D.; Huang, H.; Tian, Y.; Xu, S. Influence of LaF3 on the crystallization and luminescence of Eu3+-doped oxyfluoride glass ceramics. Ceram. Int. 2014, 40, 2737–2740. [Google Scholar] [CrossRef]
  122. Gao, W.; Dong, J.; Liu, J.; Yan, X. Highly efficient red up-conversion fluorescence emission in Yb3+/Ho3+/Ce3+ codoped LaF3nanocrystals. J. Lumin. 2016, 179, 562–567. [Google Scholar] [CrossRef]
  123. De, L.; Secco, H.; Ferreira, F.F.; Péres, L.O. Simple preparation of fluorescent composite films based on cerium and europium doped LaF3nanoparticles. J. Solid State Chem. 2018, 259, 43–47. [Google Scholar]
  124. Secu, C.E.; Matei, E.; Negrila, C.; Secu, M. The influence of the nanocrystals size and surface on the Yb/Er doped LaF3 luminescence properties. J. Alloys Compd. 2019, 791, 1098–1104. [Google Scholar] [CrossRef]
  125. Shinozaki, K.; Noji, A.; Honma, T.; Komatsu, T. Morphology and photoluminescence properties of Er3+-doped CaF2 nanocrystals patterned by laser irradiation in oxyfluoride glasses. J. Fluor. Chem. 2013, 145, 81–87. [Google Scholar] [CrossRef]
  126. Leo, L.; Xie, B.; Li, Y.; Zhang, J.; Xu, S. Improvement of the luminescent intensity of Yb/Er: CaF2 nanocrystals by combining Na+-doping and active-core/active-shell structure. J. Lumin. 2017, 190, 462–467. [Google Scholar] [CrossRef]
  127. Yagoub, M.Y.A.; Swart, H.C.; Kroon, R.E.; Coetsee, E. Low temperature photoluminescence study of Ce3+ and Eu2+ ions doped SrF2 nanocrystals. Physica B 2018, 535, 310–313. [Google Scholar] [CrossRef]
  128. Ji, G.; Hong, G.J.; Bae, C.H.; Babu, P.; Lim, K.S. Infrared-laser precipitation of Dy3+-Yb3+codoped SrF2 nanocrystals in glass and up-conversion luminescence. Appl. Surf. Sci. 2019, 487, 412–416. [Google Scholar] [CrossRef]
  129. Zhao, Z.; Liu, C.; Xia, M.; Yin, Q.; Zhao, X.; Han, J. Intense ∼1.2 μm emission from Ho3+/Y3+ ions co-doped oxyfluoride glass-ceramics containing BaF2 nanocrystals. J. Alloys Compd. 2017, 701, 392–398. [Google Scholar] [CrossRef]
  130. Huang, L.; Jia, S.; Li, Y.; Zhao, S.; Deng, D.; Wanh, H.; Jia, G.; Hua, W.; Xu, S. Enhanced emissions in Tb3+-doped oxyfluoride scintillating glass ceramics containing BaF2nanocrystals. Nucl. Instrum. Methods Phys. Res. Sect A 2015, 788, 111–115. [Google Scholar] [CrossRef]
  131. Qiu, J.; Jiao, Q.; Zhou, D.; Yang, Z. Recent progress on up-conversion luminescence enhancement in rare-earth doped transparent glass-ceramics. J. Rare Earths 2016, 34, 341–367. [Google Scholar] [CrossRef]
  132. Bitam, A.; Khiari, S.; Diaf, M.; Boubekri, H.; Boulma, E.; Bensalem, C.; Guerbous, L.; Jouart, J.P. Spectroscopic investigation of Er3+ doped BaF2 single crystal. Opt. Mater. 2018, 82, 104–109. [Google Scholar] [CrossRef]
  133. Cai, C.; Jin, Y.; Yang, Q.; Nie, X.; Liu, Y. Synergistic effect of crystal structure and concentration quenching on photoluminescence of Er3+ doped up-conversion nanocrystals. J. Rare Earths 2016, 34, 963–971. [Google Scholar] [CrossRef]
  134. Shinozaki, K.; Konaka, R.; Akai, T. Synthesis of new transparent borate-based BaF2 nanocrystallized glass by formation of nucleation sites induced by rare earth ions. J. Eur. Ceram. Soc. 2019, 39, 1735–1739. [Google Scholar] [CrossRef]
  135. Qiao, X.; Fan, X.; Wang, M. Luminescence behavior of Er3+ in glass ceramics containing BaF2 nanocrystals. Scr. Mater. 2006, 55, 211–214. [Google Scholar] [CrossRef]
  136. Zhao, Z.; Ai, B.; Liu, C.; Yin, Q.; Xia, M.; Zhao, X.; Jiang, Y. Er3+ Ions-Doped Germano-Gallate Oxyfluoride Glass-Ceramics Containing BaF2 Nanocrystals. J. Am. Ceram. Soc. 2015, 98, 2117–2121. [Google Scholar] [CrossRef]
  137. Rada, S.; Culea, E. FTIR spectroscopic and DFT theoretical study on structure of europium—Phosphate—Tellurate glasses and glass ceramics. J. Mol. Struct. 2009, 929, 141–148. [Google Scholar] [CrossRef]
  138. Yu, C.; Zhang, J.; Wen, L.; Jiang, Z. New transparent Er3+-doped oxyfluoride tellurite glass ceramic with improved near infrared and up-conversion fluorescence properties. Mater. Lett. 2007, 61, 3644–3646. [Google Scholar] [CrossRef]
  139. Chen, Y.; Wu, Z.; Hu, Y.; Wu, T.; Zhou, W. Thermal stability and optical properties of a novel Tm3+ doped fluorotellurite glass. J. Rare Earths 2014, 32, 1154–1161. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction patterns of oxyfluoride phospho-tellurite glass doped with erbium trifluoride (ErF3).
Figure 1. X-ray diffraction patterns of oxyfluoride phospho-tellurite glass doped with erbium trifluoride (ErF3).
Materials 12 03429 g001
Figure 2. DSC curves of oxyfluoride phospho-tellurite glass doped with ErF3.
Figure 2. DSC curves of oxyfluoride phospho-tellurite glass doped with ErF3.
Materials 12 03429 g002
Figure 3. FTIR spectra of precursor glass doped with varying ErF3 content.
Figure 3. FTIR spectra of precursor glass doped with varying ErF3 content.
Materials 12 03429 g003
Figure 4. Deconvoluted FTIR spectra of TP0.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 4. Deconvoluted FTIR spectra of TP0.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g004
Figure 5. Deconvoluted FTIR spectra of TP0.75ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 5. Deconvoluted FTIR spectra of TP0.75ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g005
Figure 6. Deconvoluted FTIR spectra of TP1.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 6. Deconvoluted FTIR spectra of TP1.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g006
Figure 7. Plot of bands position and their area for deconvoluted FTIR spectra of the TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Figure 7. Plot of bands position and their area for deconvoluted FTIR spectra of the TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Materials 12 03429 g007
Figure 8. Raman spectra of precursor glass doped with varying ErF3 content.
Figure 8. Raman spectra of precursor glass doped with varying ErF3 content.
Materials 12 03429 g008
Figure 9. Deconvoluted Raman spectra of TP0.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 9. Deconvoluted Raman spectra of TP0.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g009
Figure 10. Deconvoluted Raman spectra of TP0.75ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 10. Deconvoluted Raman spectra of TP0.75ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g010
Figure 11. Deconvoluted Raman spectra of TP1.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Figure 11. Deconvoluted Raman spectra of TP1.25ErF3 glass. (inset) Deconvoluted and experimental spectra.
Materials 12 03429 g011
Figure 12. Plot of bands position and their area for deconvoluted Raman spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Figure 12. Plot of bands position and their area for deconvoluted Raman spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Materials 12 03429 g012
Figure 13. 2725 nm emission spectra of oxyfluoride phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of 2725 nm emission band as a function of ErF3 concentration.
Figure 13. 2725 nm emission spectra of oxyfluoride phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of 2725 nm emission band as a function of ErF3 concentration.
Materials 12 03429 g013
Figure 14. 1553 nm emission spectra of oxyfluoride phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of 1553 nm emission band as a function of ErF3 concentration.
Figure 14. 1553 nm emission spectra of oxyfluoride phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of 1553 nm emission band as a function of ErF3 concentration.
Materials 12 03429 g014
Figure 15. Upconversion emission of phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of green (528 and 551 nm) and red (668 nm) emission bands as a function of ErF3 concentration.
Figure 15. Upconversion emission of phospho-tellurite glass with varying ErF3 content. (inset) The intensity changes of green (528 and 551 nm) and red (668 nm) emission bands as a function of ErF3 concentration.
Materials 12 03429 g015
Figure 16. The energy scheme.
Figure 16. The energy scheme.
Materials 12 03429 g016
Figure 17. Upconverison luminescence spectra of TP1.25ErF3 glass as a function of pumping power. (inset) The log-log dependency of upconversion emission intensity for green and red radiative transitions.
Figure 17. Upconverison luminescence spectra of TP1.25ErF3 glass as a function of pumping power. (inset) The log-log dependency of upconversion emission intensity for green and red radiative transitions.
Materials 12 03429 g017
Figure 18. XRD patterns of glass-ceramic.
Figure 18. XRD patterns of glass-ceramic.
Materials 12 03429 g018
Figure 19. 2725 nm emission spectra of TP1.25ErF3 glass and glass-ceramic.
Figure 19. 2725 nm emission spectra of TP1.25ErF3 glass and glass-ceramic.
Materials 12 03429 g019
Figure 20. 1553 nm emission spectra of TP1.25ErF3 glass and glass-ceramic.
Figure 20. 1553 nm emission spectra of TP1.25ErF3 glass and glass-ceramic.
Materials 12 03429 g020
Figure 21. Upconversion emission of TP1.25ErF3 glass and glass-ceramic. (inset) Image of fabricated glass and glass-ceramic.
Figure 21. Upconversion emission of TP1.25ErF3 glass and glass-ceramic. (inset) Image of fabricated glass and glass-ceramic.
Materials 12 03429 g021
Table 1. Glass transition (Tg), crystallization (Tc), melting (Tm) temperatures, and thermal stability (ΔT) of erbium-doped phospho-tellurite glass.
Table 1. Glass transition (Tg), crystallization (Tc), melting (Tm) temperatures, and thermal stability (ΔT) of erbium-doped phospho-tellurite glass.
Sample of GlassTg [°C]Tc [°C]Tm [°C]ΔT [°C]
TP0.25ErF3313 ± 1438 ± 1539 ± 1125 ± 1
TP0.50ErF3320 ± 1415/458 ± 1537 ± 195 ± 1
TP0.75ErF3322 ± 1451/483 ± 1517 ± 1129 ± 1
TP1.00ErF3324 ± 1466 ± 1547 ± 1142 ± 1
TP1.25ErF3329 ± 1420/456 ± 1557 ± 1109 ± 1
Table 2. The parameters of bands obtained by deconvolution of the FTIR spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Table 2. The parameters of bands obtained by deconvolution of the FTIR spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
TP0.25ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A5452 ± 115
B5653 ± 120
C6069 ± 235
D67411 ± 350
E7398 ± 241
F7803 ± 124
G8021 ± 116
H8723 ± 128
I9189 ± 241
J97910 ± 230
K102710 ± 236
L107717 ± 356
M113515 ± 263
TP0.75ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A5462 ± 115
B5673 ± 120
C6069 ± 235
D67612 ± 350
E7388 ± 241
F7793 ± 125
G8021 ± 117
H8733 ± 129
I9239 ± 243
J9767 ± 228
K102614 ± 243
L109421 ± 358
M11587 ± 256
TP1.25ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A5463 ± 116
B5673 ± 119
C6069 ± 236
D68111 ± 351
E7407 ± 239
F7793 ± 124
G8011 ± 116
H8733 ± 128
I9249 ± 243
J9744 ± 227
K102415 ± 248
L110227 ± 464
M11784 ± 256
Table 3. Band assignment.
Table 3. Band assignment.
Band [cm−1]Assignment
A
545/546/546
deformation vibration of δO–P–O and δO–P–O bonds in Q2 units [68]
B
565/567/567
C
606/606/607
stretching vibrations of Te–O bonds in TeO4 (tbp) units [69,70]
D
674/676/681
E
739/738/740
stretching vibrations of TeO3 (tp) units or TeO3+1 polyhedra [70]
F
780/779/779
the vibration of the continuous network composed of TeO4 and Te–O stretching vibration of TeO3+1 polyhedra [69] or symmetric P–O = P bonds in Q1 units [71]
G
802/802/801
asymmetric stretching vibrations of TeO3 (tp) units or TeO3+1 polyhedral [69,72]
H
872/873/873
asymmetric stretching vibrations of Q2 units [73]
I
918/923/924
asymmetric stretching vibrations of P–O–P linked with metaphosphate chainand P–F groups in Q2 units [74,75]
J
979/976/974
asymmetric stretching vibrations of P–O bonds in Q0 units [68]
K
1027/1026/1024
symmetric stretching vibrations of PO2−3 groups in Q1 units [74]
L
1077/1094/1102
asymmetric stretching vibrations of PO2−3 groups in Q1 units [76]
M
1135/1158/1178
asymmetric stretching vibrations of non-bridging oxygen in Q2 units [77]
Table 4. The parameters of bands for the deconvoluted Raman spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
Table 4. The parameters of bands for the deconvoluted Raman spectra of TP0.25ErF3, TP0.75ErF3, and TP1.25ErF3 glasses.
TP0.25ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A3737 ± 269
B45911 ± 254
C58016 ± 376
D6697 ± 258
E7117 ± 265
F77921 ± 449
G87216 ± 256
H95310 ± 351
I10292 ± 128
J10913 ± 0.556
TP0.75ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A3686 ± 264
B45912 ± 267
C58115 ± 366
D6536 ± 256
E7046 ± 260
F78322 ± 459
G88319 ± 256
H9578 ± 349
I10384 ± 144
J11122 ± 0.540
TP1.25ErF3 Glass
BandPeak [cm−1]Area [%]FWHM [cm−1]
A3617 ± 260
B44212 ± 247
C54819 ± 386
D6455 ± 265
E7147 ± 249
F78819 ± 457
G87823 ± 261
H9525 ± 344
I10202 ± 165
J11141 ± 0.528
Table 5. Raman band assignment.
Table 5. Raman band assignment.
Band [cm−1]Assignment
A 373/368/361bending vibration of Te–(O, F)–Te or O,F–Te–O,F bands of [Te(O, F)4] trigonal bipyramidal units [81,82]
B 459/459/442
C 580/581/548
D 669/653/645stretching variation of Te–O,F bonds in [Te(O,F)4] units [83]
E 711/704/714Bands assigned to the Te(O,F)4 tbp units [84]
F 779/783/788Te–O stretching vibration in [TeO3] trigonal pyramids or symmetric stretching vibration in [TeO3+1] units [85]
G 872/883/878symmetric stretching vibration of the P–F bonds [86]
H 953/957/952symmetric stretching vibration of PO4 in Q0 units [87]
I 1029/1038/1020stretching vibrations of bridging P–O–P bonds in Q1 units [25,90]
J 1091/1112/1142stretching vibrations of non-bridging bonds PO2 of Q2 units [23,25,90]

Share and Cite

MDPI and ACS Style

Lesniak, M.; Zmojda, J.; Kochanowicz, M.; Miluski, P.; Baranowska, A.; Mach, G.; Kuwik, M.; Pisarska, J.; Pisarski, W.A.; Dorosz, D. Spectroscopic Properties of Erbium-Doped Oxyfluoride Phospho-Tellurite Glass and Transparent Glass-Ceramic Containing BaF2 Nanocrystals. Materials 2019, 12, 3429. https://doi.org/10.3390/ma12203429

AMA Style

Lesniak M, Zmojda J, Kochanowicz M, Miluski P, Baranowska A, Mach G, Kuwik M, Pisarska J, Pisarski WA, Dorosz D. Spectroscopic Properties of Erbium-Doped Oxyfluoride Phospho-Tellurite Glass and Transparent Glass-Ceramic Containing BaF2 Nanocrystals. Materials. 2019; 12(20):3429. https://doi.org/10.3390/ma12203429

Chicago/Turabian Style

Lesniak, Magdalena, Jacek Zmojda, Marcin Kochanowicz, Piotr Miluski, Agata Baranowska, Gabriela Mach, Marta Kuwik, Joanna Pisarska, Wojciech A. Pisarski, and Dominik Dorosz. 2019. "Spectroscopic Properties of Erbium-Doped Oxyfluoride Phospho-Tellurite Glass and Transparent Glass-Ceramic Containing BaF2 Nanocrystals" Materials 12, no. 20: 3429. https://doi.org/10.3390/ma12203429

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop