Next Article in Journal
Effect of Different Ultrasonic Power on the Properties of RHA Steel Welded Joints
Next Article in Special Issue
Near-IR Luminescence of Rare-Earth Ions (Er3+, Pr3+, Ho3+, Tm3+) in Titanate–Germanate Glasses under Excitation of Yb3+
Previous Article in Journal
Shaking Table Test of a Transfer-Purge Chamber in Nuclear Island Structure
Previous Article in Special Issue
Investigation of the TeO2/GeO2 Ratio on the Spectroscopic Properties of Eu3+-Doped Oxide Glasses for Optical Fiber Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Spectroscopic Properties of Inorganic Glasses Doped with Pr3+: A Comparative Study

Institute of Chemistry, University of Silesia, Szkolna 9 Street, 40-007 Katowice, Poland
*
Author to whom correspondence should be addressed.
Materials 2022, 15(3), 767; https://doi.org/10.3390/ma15030767
Submission received: 17 December 2021 / Revised: 16 January 2022 / Accepted: 17 January 2022 / Published: 20 January 2022

Abstract

:
The results presented in this communication concern visible and near-IR emission of Pr3+ ions in selected inorganic glasses, i.e., borate-based glass with Ga2O3 and BaO, lead-phosphate glass with Ga2O3, gallo-germanate glass modified by BaO/BaF2, and multicomponent fluoride glass based on InF3. Glasses present several emission bands at blue, reddish orange, and near-infrared spectral ranges, which correspond to 4f–4f electronic transitions of Pr3+. The profiles of emission bands and their relative intensity ratios depend strongly on glass-host. Visible emission of Pr3+ ions is tuned from red/orange for borate-based glass to nearly white light for multicomponent fluoride glass based on InF3. The positions and spectral linewidths for near-infrared luminescence bands at the optical telecommunication window corresponding to the 1G43H5, 1D21G4, and 3H43F3,3F4 transitions of Pr3+ are dependent on glass-host matrices and excitation wavelengths. Low-phonon fluoride glasses based on InF3 and gallo-germanate glasses with BaO/BaF2 are excellent candidates for broadband near-infrared optical amplifiers. Spectroscopic properties of Pr3+-doped glasses are compared and discussed in relation to potential optical applications.

1. Introduction

Praseodymium-doped inorganic glasses, due to several visible and near-infrared emission transitions, are interesting from the spectroscopic point of view. Systematic studies demonstrate that radiative and non-radiative relaxation from the excited states of Pr3+ ions depend significantly on the glass-host matrices. These aspects for borate [1], phosphate [2], silicate [3], tellurite [4], germanate [5], and other non-oxide glass systems [6,7,8,9,10] are well documented in literature. Emission properties of Pr3+-doped glasses have been examined at visible wavelengths [11,12,13,14,15,16,17] and the near-infrared (NIR) region [18,19,20,21]. Most published papers are related to luminescence spectroscopy of Pr3+ ions in glasses belonging to the heavy metal glass family [22,23,24,25,26,27,28]. Special attention has been paid to Pr3+ ions in silicate glass containing lead [29,30,31]. Further comprehensive investigations indicate that the emission bands associated with electronic transitions of Pr3+ ions are enhanced in the presence of silver [32,33,34,35] or gold [36] nanoparticles embedded into glass matrices.
Here we present comparative studies on selected inorganic glasses containing Pr3+, i.e., borate glass with Ga2O3 and BaO, lead-phosphate glass with Ga2O3, gallo-germanate glass modified by BaO/BaF2, and multicomponent fluoride glass based on InF3. Based on luminescence spectra and their decays, several spectroscopic parameters of Pr3+ ions were determined. Previous investigations illustrated quite well the relationship between the structural modifications of glasses and their emission and spectroscopic properties. For example, several glass-modifiers were introduced to borate glasses doped with Pr3+ ions. Anjaiah et al. [37] studied luminescence properties of Pr3+-doped lithium borate glasses modified by MO (where M = Zn, Ca, Cd). Based on some spectroscopic parameters such as the Judd–Ofelt intensity parameter Ω2 and the bonding parameter δ, it was suggested that the covalent environment for Pr3+ increased in the following direction CdO < CaO < ZnO, and glass modified by CdO becomes a better candidate for thermoluminescence among the three studied Pr3+-doped glass systems. Spectroscopic properties of Pr3+ are also changed during modification of the borate glass composition with lithium oxide and fluoride. Jayasankar and Babu [38] revealed that the radiative lifetimes for the excited states of Pr3+ ions are reduced with decreasing lithium oxide concentration, while their values increase with increasing LiF content. The local structure and some properties of borate glasses are also changed with Li2O [39] and Na2O [40], respectively.
Furthermore, Pr3+-doped borate-based glasses modified by MO (M = Ca, Sr, Ba) have also been studied using emission spectroscopy. The emission bands related to the 1D23H4 transition of Pr3+ ions are slightly shifted to lower wavelengths (nephelauxetic effect), and the 1D2 measured lifetimes are reduced in the presence of glass-modifiers in the direction BaO → SrO → CaO [41]. Further modification of borate glass realized by replacement of BaO by BaF2 results in spectral shift of the reddish orange 1D23H4 transition of Pr3+ ions toward shorter wavelengths [42]. The changes in luminescence decays, profiles of emission bands, and their relative intensity ratios will be stronger for glass-host matrices, including different glass-formers. These effects were examined previously for some glass systems singly doped with Tm3+ [43], Sm3+ [44], Dy3+ [45], Yb3+ [46], and glass co-doped with Yb3+/Er3+ [47]. The intention of our work is to present how kinds of glass-host matrix influence the spectral profiles of luminescence bands of Pr3+ ions and their relative intensity ratios measured in the visible and near-infrared ranges. Based on spectroscopic parameters of Pr3+ ions, the glass-host matrices are selected as promising materials for multicolor visible light sources or broadband near-infrared optical amplifiers.

2. Materials and Methods

Selected inorganic glasses doped with Pr3+ ions were synthesized using traditional high-temperature melt-quenching technique. Their chemical compositions and melting conditions are given in Table 1. For the studied glass samples, the activator concentration was the same (0.1 mol%). Pr3+-doped lead-phosphate glass with Ga2O3 (PPG-Pr) was also selected to study reddish orange emission varying with activator concentration. Samples of PPG-Pr with various Pr3+ concentrations were prepared. They are given in Table 2. The appropriate precursor metal oxides and/or fluorides of high purity (99.99%) were mixed in a Pt crucible and then melted in a special glove-box in an Ar atmosphere. Glass samples with dimension = 10 mm × 10 mm and thickness = 2 mm were obtained.
In the next step, absorption and luminescence measurements were carried out. The UV-VIS-NIR spectrophotometer (Cary 5000, Agilent Technology, Santa Clara, CA, USA) was used to measure absorption spectra. Luminescence spectra and their decays were registered using a VIS/NIR laser system. The laser equipment consisted of a Photon Technology International (PTI) Quanta-Master 40 (QM40) UV/VIS Steady State Spectrofluorometer (Photon Technology International, Birmingham, NJ, USA) coupled with tunable pulsed optical parametric oscillator (OPO), pumped by a third harmonic of a Nd:YAG laser (Opotek Opolette 355 LD, OPOTEK, Carlsband, CA, USA), xenon lamp as a light source, double 200 mm monochromator, multimode UVVIS PMT R928 detector (PTI Model 914), and Hamamatsu H10330B-75 detector (Hamamatsu, Bridgewater, NJ, USA). Resolution for spectra measurements was ±0.2 nm. Decays were measured with an accuracy of ±2 µs. Transmittance spectra were performed on the Nicolet iS50 ATR spectrometer (Thermo Fisher Scientific Instruments, Waltham, MA, USA).

3. Results and Discussion

Five glass-host matrices given in Table 1 and referred to as IZSBGL-Pr, GBFG-Pr, GBG-Pr, PPG-Pr, and BBG-Pr were selected for comparative spectroscopic investigations. It should be noted that all glass samples were obtained under the same experimental conditions in order to compare their spectroscopic properties. It is well known that the conditions of synthesis are more restrictive for pure fluoride glasses in contrast to oxide glass systems. In our case, all samples were prepared in a glove-box under an atmosphere of dry argon (O2, H2O < 0.5 ppm). This procedure is especially important for IZSBGL-Pr, due to fluorine evaporation during the glass synthesis. For that reason, a small amount of ammonium bifluoride (NH4HF2) as a fluorinating agent was also added before melting. Unfortunately, the actual concentration of fluorine ions has not been estimated. The final composition of IZSBGL-Pr may be somewhat different from the nominal starting one due to fluorine losses during the melting process. Previously published works suggest that the fluorine losses could be quite large [48,49,50,51,52]. An another important factor that effectively quenched the luminescence is the concentration of OH- groups, which can be calculated from the transmittance spectrum. Figure 1 shows transmittance spectra measured for glass samples in the 3950–2950 cm−1 frequency region. The absorption band centered at about 3400 cm−1 is ascribed to the vibration of OH- groups.
For fluoride glass based on InF3, the concentration of OH groups is extremely low. The absorption coefficient and content of hydroxyl groups are close to 0.088 cm−1 and 3.82 ppm [53], respectively. The reduced concentration of hydroxyl groups is necessary to obtain pure fluoride glass with relatively high quantum efficiency and to enhance near-IR and mid-IR emission [54]. Further investigations indicate that the band intensity of hydroxyl groups is considerably smaller for mixed oxyfluoride gallo-germanate glass with BaF2 (GBFG-Pr) than oxide glass (GBG-Pr). The residual absorption of OH- groups is reduced drastically in gallo-germanate glass, where BaO was replaced by BaF2 [55]. These aspects are also important for phosphate glasses due to the hygroscopic nature of P2O5. The concentration of hydroxyl groups is usually higher in phosphate glass compared to other oxide glasses. Our recent studies for the lead-phosphate system [56] clearly demonstrated that the intensity of the IR band related to vibration of hydroxyl groups is considerably lower for glass samples synthesized in glove-box than in open air. These phenomena are very important from the optical point of view. Based on our published works [56,57,58,59], different physicochemical properties of the studied Pr3+-doped glasses are also summarized in Table 3.
From the average molecular weight, density, Pr3+ ion concentration, and refractive index exhibited in Table 3, various other radiative parameters were calculated. The three phenomenological intensity parameters Ωt (where t = 2, 4, 6) were calculated by using the appropriate relations from the Judd–Ofelt (J–O) theory. In particular, the J–O intensity parameter Ω2 is attributed to the sensitivity to the local glass structure of the rare earth sites. It is affected by symmetry/asymmetry sites and covalent/ionic bonding character between Pr3+ ions and the nearest surroundings. In other words, the lower values of Ω2 suggest a higher degree of ionic bonding between rare earth ions and their ligands. It is clearly seen that the value of Ω2 is greater for glass GBG-Pr, in contrast to fluoride glass IZSBGL-Pr and oxide glasses assigned to PPG-Pr and BBG-Pr, suggesting a higher degree of covalence between Pr3+ ions and the surrounding ligands. Independently of glass-host, the radiative transition rates obtained from the J–O calculations are significantly higher for the 3P0 state than the lower-lying 1D2 state of Pr3+ ions. Further calculations from the relevant expression η = τmrad × 100% (τm and τrad are measured and radiative lifetime, respectively, calculated from the J–O theory) indicate that the quantum efficiency for the excited state 1D2 (Pr3+) is significantly larger for low-phonon oxide (GBG-Pr) and fluoride (IZSBGL-Pr) glasses, confirming their suitability for near-infrared luminescence applications. Glass transition temperature Tg for the studied glass-host matrices was also determined from DSC curve measurements. The value of Tg is much lower for fluoride glass IZSBGL-Pr compared to other systems. Glass GBG-Pr is characterized by the highest glass transition temperature among the studied glass systems. It is also interesting to note that the value of Tg changes from 620 °C (GBG-Pr) to 599 °C (GBFG-Pr) in gallo-germanate glass where BaO was partially substituted by BaF2 [59]. Furthermore, the energy level diagram for Pr3+ ions schematized in Figure 2 favors several visible and near-infrared emission transitions.
The spectroscopic results for Pr3+ ions in fluoride glass based on InF3 (IZSBGL-Pr), borate glass with Ga2O3 and BaO (BBG-Pr), lead-phosphate glass with Ga2O3 (PPG-Pr), and gallo-germanate glassses modified by BaO/BaF2 (referred to as GBG-Pr and GBFG-Pr) are presented and discussed here.
Figure 3 presents absorption (a,b) and visible emission (c,d) spectra and emission decays (e,f) from 1D2 state of Pr3+ ions in the studied glass systems. Absorption spectra consist of characteristic bands which correspond to transitions originating from ground state 3H4 to the higher-lying excited states of praseodymium ions. The most intense bands centered at 445 nm and 590 nm are related to 3H43P2 and 3H41D2 transitions of Pr3+, respectively. The UV cut-off wavelength, defined as the intersection between the zero baseline and the extrapolation of absorption edge, is located in the 300–350 nm range. In general, the absorption edge is shifted to shorter wavelengths from oxide borate glass BBG-Pr to fluoride glass IZSBGL-Pr. Visible emission spectra were excited at 3P2 state (λexc = 445 nm) and show several characteristic bands of Pr3+ ions. The most intense bands are located in the blue and reddish orange spectral ranges and correspond to 3P03H4, 1D23H4, 3P03H6, and 3P03F2 electronic transitions of Pr3+.
Further analysis demonstrates that the relative integrated intensities of emission bands located in the blue and reddish orange region are completely different and depend strongly on kind of glass-host. Previous studies revealed that fluorescence intensity ratio, referred to as red-to-blue R/B [60] or orange-to-blue O/B [61], decreases with increasing Pr3+ ion concentration. In our case, fluorescence intensity ratio IREDDISH-ORANGE/IBLUE varying with glass-host was also estimated and schematized in Figure 3. This factor is enhanced rapidly from fluoride glass IZSBGL-Pr to borate-based glass BBG-Pr, due to the increase in the non-radiative rates. As a consequence, the 3P0 state is depopulated very quickly, and the excitation energy is transferred non-radiatively to the lower-lying state 1D2 (Pr3+). It can be well explained by the phonon energy of the host (Table 3), which increases from 510 cm−1 (IZSBGL-Pr) to 1400 cm−1 (BBG-Pr). Thus, high-phonon borate glass BBG-Pr is favored to bridge the energy gap between 3P0 and 1D2 states of Pr3+ ions, and reddish orange emission due to 1D23H4 transition is dominant. This was also confirmed by luminescence decay analysis. The multi-phonon relaxation rates of Pr3+ increase with increasing phonon energy from IZSBGL-Pr to BBG-Pr. Owing to higher multi-phonon relaxation rates, the measured luminescence lifetimes of 1D2 (Pr3+) are reduced from fluoride glass IZSBGL-Pr to borate-based glass BBG-Pr. Furthermore, luminescence decays from 1D2 state in all glass samples containing 0.1 mol% Pr3+ ions are mono-exponential. According to the excellent review article published recently by Tanner et al. [62], mono-exponential decay using the Förster expression for WET can be given for electric-dipole type transfer by:
I D ( t ) = I D ( 0 ) exp [ ( 1 τ D + W ET ) t ]  
or the following relation:
I D ( t ) = I D ( 0 ) exp [ ( 1 ( R 0 R ) 6 ) ( t τ D ) ]
where R0 is critical transfer distance (also called Förster radius), R is the average interionic separation, equal to (3/4πN)1/3, and N denotes activator concentration.
The energy transfer and cross-relaxation processes are neglected when the average interionic separation R between Pr3+ ions is greater than the critical transfer distance R0. Our studies clearly indicate that calculated values of R for all studied glass-host matrices containing 0.1 mol% Pr3+ ions are greater than the Förster distances R0 (Table 3). Previous results obtained for Pr3+-doped ZBLAN fluoride glass suggest that the average distance is smaller than the critical transfer distance and the energy transfer process will promote the non-exponential decay from the 1D2 state for activator (Pr3+) content ≥ 0.5 mol% [63].
Among inorganic glass systems, it is also found that the measured 1D2 luminescence lifetime is longer than the 3P0 lifetime of Pr3+ ions. This was confirmed by luminescence decay curve measurements for Pr3+ ions in multicomponent fluoro-phosphate glasses [15], oxyfluoroborate glasses [17], lead germanate glasses [23], and borosilicate glasses [61] as well as tellurite [64] and zinc telluro-fluoroborate [65] glass systems. Luminescence lifetimes for 3P0 and 1D2 states of Pr3+ ions in different glass-host matrices are presented in Table 4. Also, the x and y of CIE chromaticity coordinates for IZSBGL-Pr, GBFG-Pr, GBG-Pr, PPG-Pr, and BBG-Pr systems were calculated from the emission spectra. The results are given in Table 5. They are shown in the chromaticity diagram in Figure 4.
Spectroscopic studies indicate that PPG-Pr and BBG-Pr belong to inorganic glasses emitting reddish orange emission, similar to other lead-free and lead-based [66,67,68,69] glass systems doped with Pr3+ published recently. It is noteworthy that the color of emission is changed from reddish orange (BGB-Pr) to yellowish orange (BGFG-Pr) where BaO was replaced by BaF2 (5 mol%). Based on the CIE diagram, we can conclude that emission can be tuned from red/orange (BBG-Pr) to nearly white light region (IZSBGL-Pr) by changing chromaticity parameters by varying the glass-host matrix.
Our previous investigations suggested that the spectral profiles of emission bands of Pr3+ ions and their relative intensity ratios are changed during modification of glass-host. In the orange-red region, two emission bands due to 1D23H4 (orange) and 3P03H6 (red) transitions of Pr3+ are overlapped, and their intensities depend strongly on the glass-host. This was well evidenced for gallo-germanate glasses modified by BaO/BaF2 [70]. Figure 5 shows reddish orange emission spectra dependent on glass-host matrix and Pr3+ content.
From literature data, it is well known that multi-phonon relaxation (MPR) and cross relaxation (CR) processes play an important role in population or depopulation of the 1D2 state of Pr3+ ions in inorganic glasses. The non-radiative transition rate Wnr due to the MPR process is equal to 2.47 × 104 s1 for glass based on InF3 [71], whereas the value of Wnr for borate glass is approximately 103 times larger than that of the fluoride glass system [72]. The phonon energy of the host increases from IZSBGL-Pr to BBG-Pr. Thus, the excitation energy is transferred more efficiently from the higher-lying 3P0 state to the 1D2 state, and, consequently, reddish orange luminescence corresponding to 1D23H4 transition of Pr3+ in borate-based glass (BBG-Pr) is dominant, as mentioned above. This situation is observed for glasses when the molar concentration of Pr3+ ions is relatively low and its value is close to 0.1 mol%. It is generally accepted that the MPR process from the 3P0 state at lower concentrations (usually below 0.5 mol%) favors reddish orange emission from the 1D2 state to be more dominant [61]. For higher activator concentrations (above 0.5 mol%), the non-radiative energy transfer processes between Pr3+ ions become efficient, and luminescence associated to 1D23H4 transition is successfully quenched through cross-relaxation. The following CR processes, 1D2: 3H41G4: (3F3,3F4) and 1D2: 3H4 → (3F3,3F4): 1G4, are responsible for depopulation of the 1D2 state of Pr3+ [73]. In addition, these aspects have been examined by us. In our case, lead-phosphate glass (PPG-Pr) was selected as an intermediate medium, in which luminescence from both 3P0 and 1D2 states of Pr3+ ions are well observed and the 1D23H4 transition is dominant at low activator concentration. The results are presented in Figure 5b. It is well evidenced that the emission intensity of 1D23H4 transition is reduced, whereas the emission intensities of bands originating from the 3P0 state are enhanced with increasing Pr3+ concentration. These phenomena are associated with Pr3+–Pr3+ interaction increasing and the presence of cross-relaxation processes at higher activator concentration. A similar situation was observed for zinc-telluro-fluoroborate glass examined as a function of Pr3+ ion concentration [66]. Our experimental results evidently suggest that the contribution of the glass-host to the change in the spectral factor IREDDISH-ORANGE/IBLUE and measured lifetime (Figure 3) seems to be dominant when content of Pr3+ is relatively low (0.1 mol%). In this case, the multi-phonon relaxation process makes an important contribution to the excited state relaxation of Pr3+ (Figure 5a). The situation was completely changed when concentration of rare earths was relatively high (Figure 5b). Thus, the contribution of activator content was dominant. This behavior is due to the presence of non-radiative energy transfer processes (such as cross-relaxation), which contribute to quenching of luminescence corresponding to 1D23H4 transition of Pr3+.
Figure 6 presents near-infrared emission spectra of Pr3+ ions in inorganic glasses, which were excited at 445 nm (3P0) and 590 nm (1D2), respectively. In order to compare luminescence linewidth, defined as full width at half maximum (FWHM), the spectra measured in the 1200–1650 nm range were also normalized.
The near-infrared luminescence spectra show several bands which correspond to 1D23F3,3F4, 1G43H5, 1D21G4, and 3H43F3,3F4 transitions of Pr3+, respectively. Their relative integrated emission intensities are changed drastically with glass-host matrices. In particular, luminescence bands located in the so-called telecom window (1200–1650 nm) have been examined in detail. In this spectral range, ultra-broadband near-infrared emission of Pr3+ ions related to 1G43H5p = 1330 nm) and 1D21G4p = 1480 nm) transitions is observed for several inorganic glasses, which is extremely useful for optical fiber amplifiers operating at E-, S-, C-, and L-band [74]. In some cases, a near-infrared emission band centered at about 1600 nm is also visible. This emission band is connected with the 3H43F3,3F4 transition of Pr3+ [75]. Interesting results are observed for fluoride glass IZSBGL-Pr. In contrast to other studied glass systems, the intensities of emission bands of Pr3+ ions in glass IZSBGL-Pr, covering a spectral range from 1200 nm to 1650 nm, depend also on the excitation wavelengths (445 nm/590 nm). When glass IZSBGL-Pr was excited at 445 nm (3P2), the intensities of bands were extremely low, and the near-infrared emission near 1335 nm due to the 1G43H5 transition was dominant. The situation changed when the glass sample was excited at 590 nm (1D2). Thus, the near-infrared emission in glass IZSBGL-Pr is the most intense, and the 1D21G4 transition centered at about 1450 nm is dominant.
Independently of excitation wavelengths, broadband near-infrared emission bands (FWHM above 200 nm) are observed for BBG-Pr, GBG, and GBFG glasses. From the literature, it is well known that near-infrared luminescence properties of glasses containing Pr3+ ions depend strongly on the excitation wavelengths. The blue and orange excitation lines are unusually helpful to examine conversion of blue light into near-infrared radiation and its mechanism. These processes have been observed for some fluoride materials and other low-phonon systems. The experimental results for glass-ceramic materials with CaF2:Pr3+ nanocrystals [76] indicated that a two-step near-infrared quantum cutting (NIR-QC) from blue-excited 3P0 state takes place efficiently, with 1G4 acting as an intermediate state. Blue light excitation leading to efficient population of 1G4 state also influences the relative integrated intensities of emission bands, which correspond to near-infrared transitions originating from both 1D2 and 1G4 states of Pr3+. A tunable amplification band depending on the excitation wavelength used (474 nm/980 nm) has been also observed for Pr3+/Yb3+ co-doped systems, where it is possible to select 1D21G4 or 1G43H5 transition of Pr3+. When the excitation wavelength was changed from 474 nm to 980 nm, the near-infrared luminescence switched from the E–S bands near 1480 nm to the O–E bands centered at 1330 nm in Pr3+/Yb3+ co-doped tellurite tungstate glasses [77].
Finally, some spectroscopic parameters for Pr3+ ions were determined. One of the most important radiative parameters is the peak stimulated emission cross-section σem, which can be calculated using the expression:
σ em = λ p 4 8 π cn 2 Δ λ   A J
where λp is the peak emission wavelength, n—the refractive index, c—the velocity of light, Δλ—the emission linewidth (FWHM), and AJ—the calculated radiative transition rate from the J–O theory. The values of n and AJ are given in Table 3. It is generally accepted that a relatively large value of σem is demanded for an efficient laser transition.
In the next step, the stimulated emission cross-section (σem), the measured emission lifetime (τm), and the emission linewidth (FWHM) were applied to calculate the following parameters: figure of merit FOM (σem × τm) and gain bandwidth (σem × FWHM product). The results for the 1D21G4 transition of Pr3+ ions in the glass systems excited at 590 nm are given in Table 6.
The peak stimulated emission cross-section for PPG-Pr close to σem = 1.28 × 10−20 cm2 is relatively large and comparable to the values 1.14 × 10−20 cm2 [78] and 1.29 × 10−20 cm2 [79] reported previously for similar phosphate-based glasses doped with Pr3+. The smaller values of the stimulated emission cross-section (σem = 0.5 × 10−20 cm2) as well as the gain bandwidth (σem × FWHM = 65 × 10−27 cm3) for fluoride glass IZSBGL-Pr are mainly due to the considerably lower spectral linewidth for 1D21G4 transition of Pr3+. On the other hand, the figure of merit (FOM) for IZSBGL-Pr is the highest among the studied glass systems.
The peak stimulated emission cross-section, the figure of merit (FOM), and the gain bandwidth seem to be considerably smaller for glass BBG-Pr, due to its relatively large non-radiative transition rate. For that reason, high-phonon borate-based glass BBG-Pr is useless for near-infrared luminescence applications. The σem × FWHM product, as an important parameter to achieve broadband and high gain amplification, is quite large for GBG and GBFG glasses (above 200 × 1027 cm3). Their calculated values are comparable to the one (174.6 × 1027 cm3) obtained for the 1D21G4 transition of Pr3+ ions in fluorotellurite glass [80], demonstrating suitability for broadband near-infrared amplifiers.

4. Conclusions

In this work, comparative spectroscopic properties of selected inorganic glasses singly doped with Pr3+ ions are reported. The experimental results were limited to borate-based glass with Ga2O3 and BaO, lead-phosphate glass with Ga2O3, gallo-germanate glass modified by BaO/BaF2, and multicomponent fluoride glass based on InF3. Spectroscopic parameters for Pr3+ ions in glass samples were determined based on absorption/emission spectra measurements and emission decay curve analysis. Emission spectra at visible and near-infrared wavelengths were analyzed based on the energy level diagram of Pr3+. The systematic studies revealed that profiles of emission bands and their relative integrated intensity ratios depend significantly on glass-host matrices. Visible emission of Pr3+ is modulated from red/orange for borate-based glass and lead-phosphate glass with Ga2O3 via yellowish orange for gallo-germanate glass with BaO/BaF2 to nearly white light for fluoride glass based on InF3. The band positions and spectral linewidths for near-infrared luminescence at telecom range associated with the 1G43H5, 1D21G4, and 3H43F3,3F4 transitions of Pr3+ are influenced by the kind of glass matrix and excitation wavelengths. Based on several spectroscopic parameters of Pr3+ ions, it was suggested that low-phonon fluoride glasses based on InF3 and gallo-germanate glasses with BaO/BaF2 are promising materials for optical amplification. The results are compared and discussed in relation to potential applications as multicolor visible light sources or broadband near-infrared optical amplifiers.

Author Contributions

Conceptualization, J.P.; Methodology, M.K. and J.P.; formal analysis, W.A.P.; investigation, M.K., W.A.P. and J.P.; writing—original draft preparation, J.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The research activities were co-financed by the funds granted under the Research Excellence Initiative of the University of Silesia in Katowice.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Babu, P.; Jayasankar, C.K. Spectroscopy of Pr3+ ions in lithium borate and lithium fluoroborate glasses. Phys. B 2001, 301, 326–340. [Google Scholar] [CrossRef]
  2. Liu, Y.; Ren, J.; Tong, Y.; Wang, T.; Xu, W.; Chen, G. Observation of intra- and inter-configurational luminescence of Pr3+-doped strontium phosphate glasses. J. Am. Ceram. Soc. 2012, 95, 41–44. [Google Scholar] [CrossRef]
  3. Niu, L.; Zhou, Y.; Zhu, C.; He, Z.; Meng, X. Pr3+ doped oxyfluoride silicate glasses for LEDs. Ceram. Int. 2019, 45, 4108–4112. [Google Scholar] [CrossRef]
  4. Rai, V.K.; Rai, S.B.; Rai, D.K. Spectroscopic properties of Pr3+ doped in tellurite glass. Spectrochim. Acta A 2005, 62, 302–306. [Google Scholar] [CrossRef]
  5. Pisarski, W.A.; Pisarska, J.; Dorosz, D.; Dorosz, J. Rare earths in lead-free oxyfluoride germanate glasses. Spectrochim. Acta A 2015, 134, 587–591. [Google Scholar] [CrossRef]
  6. Seeber, W.; Downing, E.A.; Hesselink, L.; Fejer, M.M.; Ehrt, D. Pr3+-doped fluoride glasses. J. Non-Cryst. Solids 1995, 189, 218–226. [Google Scholar] [CrossRef]
  7. Binnemans, K.; Verboven, D.; Görller-Walrand, C.; Lucas, J.; Duhamel-Henry, N.; Adam, J.L. Absorption and magnetic circular dichroism spectra of praseodymium doped fluorozirconate (ZBLAN) glass. J. Alloys Compd. 1997, 250, 321–325. [Google Scholar] [CrossRef]
  8. Olivier, M.; Parastesh Pirasteh, P.; Doualan, J.-L.; Camy, P.; Lhermite, H.; Adam, J.-L.; Nazabal, V. Pr3+-doped ZBLA fluoride glasses for visible laser emission. Opt. Mater. 2011, 33, 980–984. [Google Scholar] [CrossRef]
  9. Manzani, D.; Paboeuf, D.; Ribeiro, S.J.L.; Goldner, P.; Bretenaker, F. Orange emission in Pr3+-doped fluoroindate glasses. Opt. Mater. 2013, 35, 383–386. [Google Scholar] [CrossRef] [Green Version]
  10. Ma, C.; Guo, H.; Xu, Y.; Wu, Z.; Li, M.; Jia, X.; Nie, Q. Effect of glass composition on the physical properties and luminescence of Pr3+ ion-doped chalcogenide glasses. J. Am. Ceram. Soc. 2019, 102, 6794–6801. [Google Scholar] [CrossRef]
  11. Sun, Y.; Yu, F.; Liao, M.; Wang, X.; Li, Y.; Hu, L.; Knight, J. Emission properties of Pr3+-doped aluminosilicate glasses at visible wavelengths. J. Lumin. 2020, 220, 117013. [Google Scholar] [CrossRef]
  12. Rao, V.H.; Prasad, P.S.; Babu, K.S. Visible luminescence characteristics of Pr3+ ions in TeO2-Sb2O3-WO3 glasses. Opt. Mater. 2020, 101, 109740. [Google Scholar] [CrossRef]
  13. Zhang, L.; Xia, Y.; Shen, X.; Wei, W. Concentration dependence of visible luminescence from Pr3+-doped phosphate glasses. Spectrochim. Acta A 2019, 206, 454–459. [Google Scholar] [CrossRef]
  14. Ramteke, D.D.; Swart, H.C.; Gedam, R.S. Spectroscopic properties of Pr3+ ions embedded in lithium borate glasses. Phys. B 2016, 480, 111–115. [Google Scholar] [CrossRef]
  15. Babu, S.; Rajput, P.; Ratnakaram, Y.C. Compositional-dependent properties of Pr3+-doped multicomponent fluoro-phosphate glasses for visible applications: A photoluminescence study. J. Mater. Sci. 2016, 51, 8037–8054. [Google Scholar] [CrossRef]
  16. Zhang, L.; Dong, G.; Peng, M.; Qiu, J. Comparative investigation on the spectroscopic properties of Pr3+-doped boro-phosphate, boro-germo-silicate and tellurite glasses. Spectrochim. Acta A 2012, 93, 223–227. [Google Scholar] [CrossRef] [PubMed]
  17. Mahato, K.K.; Rai, D.K.; Rai, S.B. Optical studies of Pr3+ doped oxyfluoroborate glass. Phys. Stat. Sol. A 1999, 174, 277–289. [Google Scholar] [CrossRef]
  18. Shaw, L.B.; Harbison, B.B.; Cole, B.; Sanghera, J.S.; Aggarwal, I.D. Spectroscopy of the IR transitions in Pr3+ doped heavy metal selenide glasses. Opt. Express 1997, 1, 87–96. [Google Scholar] [CrossRef]
  19. Choi, Y.G.; Park, B.J.; Kim, K.H.; Heo, J. Crossrelaxations between and multiphonon relaxation of near-infrared excited states of Pr3+ ions in selenide glasses. Chem. Phys. Lett. 2003, 368, 625–629. [Google Scholar] [CrossRef]
  20. Oswald, J.; Kuldova, K.; Frumarova, B.; Frumar, M. Near and mid-infrared luminescence of new chalcohalide glasses doped with Pr3+ ions. Mater. Sci. Eng. B 2008, 146, 107–109. [Google Scholar] [CrossRef]
  21. Sourkova, P.; Frumarova, B.; Frumar, M.; Nemec, P.; Kincl, M.; Nazabal, V.; Moizan, V.; Doualan, J.-L.; Moncorge, R. Spectroscopy of infrared transitions of Pr3+ ions in Ga-Ge-Sb-Se glasses. J. Lumin. 2009, 129, 1148–1153. [Google Scholar] [CrossRef]
  22. Nachimuthu, P.; Vithal, M.; Jagannathan, R. Absorption and emission spectral properties of Pr3+, Nd3+, and Eu3+ ions in heavy-metal oxide glasses. J. Am. Ceram. Soc. 2000, 83, 597–604. [Google Scholar] [CrossRef]
  23. Balda, R.; Fernandez, J.; De Pablos, A.; Fdez-Navarro, J.M. Spectroscopic properties of Pr3+ ions in lead germanate glass. J. Phys. Condens. Matter 1999, 11, 7411–7421. [Google Scholar] [CrossRef]
  24. Balda, R.; Saez de Ocariz, I.; Fernandez, J.; Fdez-Navarro, J.M.; Arriandiaga, M.A. Spectroscopy and orange-blue frequency upconversion in Pr3+-doped GeO2-PbO-Nb2O5 glass. J. Phys. Condens. Matter 2000, 12, 10623–10632. [Google Scholar] [CrossRef]
  25. Burtan, B.; Cisowski, J.; Mazurak, Z.; Jarzabek, B.; Czaja, M.; Reben, M.; Grelowska, I. Concentration-dependent spectroscopic properties of Pr3+ ions in TeO2-WO3-PbO-La2O3 glass. J. Non-Cryst. Solids 2014, 400, 21–26. [Google Scholar] [CrossRef]
  26. Mitra, S.; Jana, S. Intense orange emission in Pr3+ doped lead phosphate glass. J. Phys. Chem. Solids 2015, 85, 245–253. [Google Scholar] [CrossRef]
  27. Sharma, R.; Rao, A.S.; Deopa, N.; Venkateswarlu, M.; Jayasimhadri, M.; Haranath, D.; Vijaya Prakash, G. Spectroscopic study of Pr3+ ions doped zinc lead tungsten tellurite glasses for visible photonic device applications. Opt. Mater. 2018, 78, 457–464. [Google Scholar] [CrossRef]
  28. Basavapoornima, C.; Kesavulu, C.R.; Maheswari, T.; Pecharapa, W.; Depuru, S.R.; Jayasankar, C.K. Spectral characteristics of Pr3+-doped lead based phosphate glasses for optical display device applications. J. Lumin. 2020, 228, 117585. [Google Scholar] [CrossRef]
  29. Bhargavi, K.; Sudarsan, V.; Brik, M.G.; Rao, M.S.; Gandhi, Y.; Rao, P.N.; Veeraiah, N. Influence of Al declustering on the photoluminescent properties of Pr3+ ions in PbO-SiO2 glasses. J. Non-Cryst. Solids 2013, 362, 201–206. [Google Scholar] [CrossRef]
  30. Bhargavi, K.; Sanyal, B.; Rao, M.S.; Kumar, V.R.; Gandhi, Y.; Baskaran, G.S.; Veeraiah, N. γ-Ray induced thermoluminescence characteristics of the PbO-Al2O3-SiO2:Pr3+ glass system. J. Lumin. 2015, 161, 417–421. [Google Scholar] [CrossRef]
  31. Suresh, B.; Purnachand, N.; Zhydachevskii, Y.; Brik, M.G.; Reddy, M.S.; Suchocki, A.; Piasecki, M.; Veeraiah, N. Influence of Bi3+ ions on the amplification of 1.3 μm emission of Pr3+ ions in lead silicate glasses for the applications in second telecom window communications. J. Lumin. 2017, 182, 312–322. [Google Scholar] [CrossRef]
  32. Naranjo, L.P.; De Araújo, C.B.; Malta, O.L.; Santa Cruz, P.A.; Kassab, L.R.P. Enhancement of Pr3+ luminescence in PbO-GeO2 glasses containing silver nanoparticles. Appl. Phys. Lett. 2005, 87, 241914. [Google Scholar] [CrossRef]
  33. Rai, V.K.; Menezes, L.S. De Araújo, C.B.; Kassab, L.R.P.; Da Silva, D.M.; Kobayashi, R.A. Surface-plasmon-enhanced frequency upconversion in Pr3+ doped tellurium-oxide glasses containing silver nanoparticles. J. Appl. Phys. 2008, 103, 093526. [Google Scholar] [CrossRef] [Green Version]
  34. Hua, C.; Zhao, X.; Pun, E.Y.B.; Lin, H. Pr3+ doped tellurite glasses incorporated with silver nanoparticles for laser illumination. RSC Adv. 2017, 7, 55691–55701. [Google Scholar] [CrossRef] [Green Version]
  35. Rajesh, D.; Amjad, R.J.; Reza Dousti, M.; De Camargo, A.S.S. Enhanced VIS and NIR emissions of Pr3+ ions in TZYN glasses containing silver ions and nanoparticles. J. Alloys Compd. 2017, 695, 607–612. [Google Scholar] [CrossRef]
  36. Herrera, A.; Balzaretti, N.M. Effect of high pressure in the luminescence of Pr3+-doped GeO2-PbO glass containing Au nanoparticles. J. Phys. Chem. C 2018, 122, 27829–27835. [Google Scholar] [CrossRef]
  37. Anjaiah, J.; Laxmikanth, C.; Veeraiah, N.; Kistaiah, P. Luminescence properties of Pr3+ doped Li2O-MO-B2O3 glasses. J. Lumin. 2015, 161, 147–153. [Google Scholar] [CrossRef]
  38. Jayasankar, C.K.; Babu, P. Compositional dependence of optical properties of Pr3+ ions in lithium borate glasses. J. Alloys Compd. 1998, 275–277, 369–373. [Google Scholar] [CrossRef]
  39. Boora, M.; Malik, S.; Kumar, V.; Bala, M.; Arora, S.; Rohilla, S.; Kumar, A.; Dalal, J. Investigation of structural and impedance spectroscopic properties of borate glasses with high Li+ concentration. Solid State Ion. 2021, 368, 115704. [Google Scholar] [CrossRef]
  40. Tijaria, M.; Sharma, Y.; Kumar, V.; Dahiya, S.; Dalal, J. Effect of Na2O on physical, structural and electrical properties of borate glasses. Mater. Today Proc. 2021, 45, 3722–3725. [Google Scholar] [CrossRef]
  41. Janek, J.; Sołtys, M.; Żur, L.; Pietrasik, E.; Pisarska, J.; Pisarski, W.A. Luminescence investigations of rare earth doped lead-free borate glasses modified by MO (M = Ca, Sr, Ba). Mater. Chem. Phys. 2016, 180, 237–243. [Google Scholar] [CrossRef]
  42. Pisarska, J.; Pisarski, W.A.; Dorosz, D.; Dorosz, J. Spectroscopic properties of Pr3+ and Er3+ ions in lead-free borate glasses modified by BaF2. Opt. Mater. 2015, 47, 548–554. [Google Scholar] [CrossRef]
  43. Turri, G.; Sudesh, V.; Richardson, M.; Bass, M.; Toncelli, A.; Tonelli, M. Temperature-dependent spectroscopic properties of Tm3+ in germanate, silica, and phosphate glasses: A comparative study. J. Appl. Phys. 2008, 103, 093104. [Google Scholar] [CrossRef] [Green Version]
  44. Yuliantini, L.; Djamal, M.; Hidayat, R.; Boonin, K.; Yasaka, P.; Kaewnuam, E.; Venkatramu, V.; Kaewkhao, J. Optical and X-ray induced luminescence of Sm3+-doped borotellurite and fluoroborotellurite glasses: A comparative study. J. Lumin. 2019, 213, 19–28. [Google Scholar] [CrossRef]
  45. Zaman, F.; Srisittipokakun, N.; Rooh, G.; Khattak, S.A.; Kaewkhao, J.; Rani, M.; Kim, H.J. Comparative study of Dy3+ doped borate glasses on the basis of luminescence and lasing properties for white-light generation. Opt. Mater. 2021, 119, 111308. [Google Scholar] [CrossRef]
  46. Zhang, L.; Xia, Y.; Shen, X.; Yang, R.; Wei, W. Investigations on the effects of the Stark splitting on the fluorescence behaviors in Yb3+-doped silicate, tellurite, germanate, and phosphate glasses. Opt. Mater. 2018, 75, 1–6. [Google Scholar] [CrossRef]
  47. Desirena, H.; De la Rosa, E.; Romero, V.H.; Castillo, J.F.; Diaz-Torres, L.A.; Oliva, J.R. Comparative study of the spectroscopic properties of Yb3+/Er3+ codoped tellurite glasses modified with R2O (R = Li, Na and K). J. Lumin. 2012, 132, 391–397. [Google Scholar] [CrossRef]
  48. Bueno, L.A.; Messaddeq, Y.; Dias Filho, F.A.; Ribeiro, S.J.L. Study of fluorine losses in oxyfluoride glasses. J. Non-Cryst. Solids 2005, 351, 3804–3808. [Google Scholar] [CrossRef]
  49. Lin, L.; Ren, G.; Chen, M.; Liu, Y.; Yang, Q. Study of fluorine losses and spectroscopic properties of Er3+ doped oxyfluoride silicate glasses and glass ceramics. Opt. Mater. 2009, 31, 1439–1442. [Google Scholar] [CrossRef]
  50. Brauer, D.S.; Mneimne, M.; Hill, R.G. Fluoride-containing bioactive glasses: Fluoride loss during melting and ion release in tris buffer solution. J. Non-Cryst. Solids 2011, 357, 3328–3333. [Google Scholar] [CrossRef]
  51. De Pablos-Martín, A.; Contreras Jaimes, A.T.; Wahl, S.; Meyer, S.; Brauer, D.S. Fluorine loss determination in bioactive glasses by laser-induced breakdown spectroscopy (LIBS). Int. J. Appl. Glass Sci. 2021, 12, 213–221. [Google Scholar] [CrossRef]
  52. Möncke, D.; Da Cruz Barbosa Neto, M.; Bradtmüller, H.; De Souza, G.B.; Rodrigues, A.M.; Elkholy, H.S.; Othman, H.O.; Moulton, B.J.A.; Kamitsos, E.I.; Rodrigues, A.C.M.; et al. NaPO3-AlF3 glasses: Fluorine evaporation during melting and the resulting variations in structure and properties. J. Chem. Technol. Metall. 2018, 53, 1047–1060. [Google Scholar]
  53. Kochanowicz, M.; Zmojda, J.; Baranowska, A.; Kuwik, M.; Starzyk, B.; Lesniak, M.; Miluski, P.; Pisarski, W.A.; Pisarska, J.; Dorosz, J.; et al. Fluoroindate glass co-doped with Yb3+/Ho3+ as a 2.85 µm luminescent source for MID-IR sensing. Sensors 2021, 21, 2155. [Google Scholar] [CrossRef]
  54. Zhou, B.; Wei, T.; Cai, M.; Tian, Y.; Zhou, J.; Deng, D.; Xu, S.; Zhang, J. Analysis on energy transfer process of Ho3+ doped fluoroaluminate glass sensitized by Yb3+ for mid-infrared 2.85 µm emission. J. Quant. Spectrosc. Radiat. Transf. 2014, 149, 41–50. [Google Scholar] [CrossRef]
  55. Pisarska, J.; Kuwik, M.; Górny, A.; Kochanowicz, M.; Zmojda, J.; Dorosz, J.; Dorosz, D.; Sitarz, M.; Pisarski, W.A. Holmium doped barium gallo-germanate glasses for near-infrared luminescence at 2000 nm. J. Lumin. 2019, 215, 116625. [Google Scholar] [CrossRef]
  56. Pisarski, W.A.; Żur, L.; Goryczka, T.; Sołtys, M.; Pisarska, J. Structure and spectroscopy of rare earth—Doped lead phosphate glasses. J. Alloys Compd. 2014, 587, 90–98. [Google Scholar] [CrossRef]
  57. Pisarski, W.A. Spectroscopic analysis of praseodymium and erbium ions in heavy metal fluoride and oxide glasses. J. Mol. Struct. 2005, 744–747, 473–479. [Google Scholar] [CrossRef]
  58. Żur, L.; Janek, J.; Sołtys, M.; Pisarska, J.; Pisarski, W.A. Effect of BaF2 content on luminescence of rare-earth ions in borate and germanate glasses. J. Am. Ceram. Soc. 2016, 99, 2009–2016. [Google Scholar] [CrossRef]
  59. Pisarska, J.; Kowal, M.; Kochanowicz, M.; Zmojda, J.; Dorosz, J.; Dorosz, D.; Pisarski, W.A. Influence of BaF2 and activator concentration on broadband near-infrared luminescence of Pr3+ ions in gallo-germanate glasses. Opt. Express 2016, 24, 2427–2435. [Google Scholar] [CrossRef] [PubMed]
  60. Umamaheswari, D.; Jamalaiah, B.C.; Chengaiah, T.; Kim, I.-G.; Moorthy, L.R. Optical properties of sodium fluoroborate glasses containing Pr3+ ions for red luminescence. Phys. Chem. Glasses Eur. J. Glass Sci. Technol. B 2012, 53, 271–275. [Google Scholar]
  61. Naresh, V.; Ham, B.S. Influence of multiphonon and cross relaxations on 3P0 and 1D2 emission levels of Pr3+ doped borosilicate glasses for broad band signal amplification. J. Alloys Compd. 2016, 664, 321–330. [Google Scholar] [CrossRef]
  62. Tanner, P.A.; Zhou, L.; Duan, C.; Wong, K.-L. Misconceptions in electronic energy transfer: Bridging the gap between chemistry and physics. Chem. Soc. Rev. 2018, 47, 5234–5265. [Google Scholar] [CrossRef] [PubMed]
  63. Yang, H.; Dai, Z.; Li, J.; Tian, Y. Energy transfer and frequency upconversion in Pr3+-doped ZBLAN glass. J. Non-Cryst. Solids 2006, 352, 5469–5474. [Google Scholar] [CrossRef]
  64. Annapurna, K.; Chakrabarti, R.; Buddhudu, S. Absorption and emission spectral analysis of Pr3+: Tellurite glasses. J. Mater. Sci. 2007, 42, 6755–6761. [Google Scholar] [CrossRef]
  65. Suthanthirakumar, P.; Basavapoornima, Ch.; Marimuthu, K. Effect of Pr3+ ions concentration on the spectroscopic properties of zinc telluro-fluoroborate glasses for laser and optical amplifier applications. J. Lumin. 2017, 187, 392–402. [Google Scholar] [CrossRef]
  66. Deopa, N.; Rao, A.S.; Mahamuda, Sk.; Gupta, M.; Jayasimhadri, M.; Haranath, D.; Prakash, G.V. Spectroscopic studies of Pr3+ doped lithium lead alumino borate glasses for visible reddish orange luminescent device applications. J. Alloys Compd. 2017, 708, 911–921. [Google Scholar] [CrossRef]
  67. Venkateswarlu, M.; Prasad, M.V.V.K.S.; Swapna, K.; Mahamuda, Sk.; Rao, A.S.; Babu, A.M.; Haranath, D. Pr3+ doped lead tungsten tellurite glasses for visible red lasers. Ceram. Int. 2014, 40, 6261–6269. [Google Scholar] [CrossRef]
  68. Mahamuda, Sk.; Swapna, K.; Rao, A.S.; Sasikala, T.; Moorthy, L.R. Reddish-orange emission from Pr3+ doped zinc alumino bismuth borate glasses. Phys. B 2013, 428, 36–42. [Google Scholar] [CrossRef]
  69. Devi, C.B.A.; Mahamuda, S.; Swapna, K.; Venkateswarlu, M.; Rao, A.S.; Prakash, G.V. Pr3+ ions doped single alkali and mixed alkali fluoro tungsten tellurite glasses for visible red luminescent devices. J. Non-Cryst. Solids 2018, 498, 345–351. [Google Scholar] [CrossRef]
  70. Pisarska, J.; Pisarski, W.A.; Dorosz, D.; Dorosz, J. Spectral analysis of Pr3+ doped germanate glasses modified by BaO and BaF2. J. Lumin. 2016, 171, 138–142. [Google Scholar] [CrossRef]
  71. Oliveira, A.S.; Gouveia, E.A.; De Araujo, M.T.; Gouveia-Neto, A.S.; De Araujo, C.B.; Messaddeq, Y. Twentyfold blue upconversion emission enhancement through thermal effects in Pr3+/Yb3+-codoped fluoroindate glasses excited at 1.064 µm. J. Appl. Phys. 2000, 87, 4274–4278. [Google Scholar] [CrossRef] [Green Version]
  72. Shojiya, M.; Kawamoto, Y.; Kadono, K. Judd-Ofelt parameters and multiphonon relaxation of Ho3+ ions in ZnCl2-based glass. J. Appl. Phys. 2001, 89, 4944–4950. [Google Scholar] [CrossRef]
  73. Choi, Y.G.; Baik, J.H.; Heo, H.J. Spectroscopic properties of Pr3+: 1D21G4 transition in SiO2-based glasses. Chem. Phys. Lett. 2005, 406, 436–440. [Google Scholar] [CrossRef]
  74. Liu, X.; Chen, B.J.; Pun, E.Y.B.; Lin, H. Ultra-broadband near-infrared emission in praseodymium ion doped germanium tellurite glasses for optical fiber amplifier operating at E-, S-, C-, and L-band. J. Appl. Phys. 2012, 111, 116101. [Google Scholar] [CrossRef]
  75. Choi, Y.G.; Kim, K.H.; Park, B.J.; Heo, H.J. 1.6 μm emission from Pr3+: (3F3, 3F4) → 3H4 transition in Pr3+- and Pr3+/Er3+-doped selenide glasses. Appl. Phys. Lett. 2001, 78, 1249–1251. [Google Scholar] [CrossRef]
  76. Yu, D.C.; Chen, Q.J.; Lin, H.H.; Wang, Y.Z.; Zhang, Q.Y. New insight into two-step near-infrared quantum cutting in Pr3+ singly doped oxyfluoride glass-ceramics. Opt. Mater. Express 2016, 6, 197–205. [Google Scholar] [CrossRef]
  77. Belançon, M.P.; Marconi, J.D.; Ando, M.F.; Barbosa, L.C. Near-IR emission in Pr3+ single doped and tunable near-IR emission in Pr3+/Yb3+ codoped tellurite tungstate glasses for broadband optical amplifiers. Opt. Mater. 2014, 36, 1020–1026. [Google Scholar] [CrossRef]
  78. Shen, L.F.; Chen, B.J.; Lin, H.; Pun, E.Y.B. Praseodymium ion doped phosphate glasses for integrated broadband ion-exchanged waveguide amplifier. J. Alloys Compd. 2015, 622, 1093–1097. [Google Scholar] [CrossRef]
  79. Han, X.; Shen, L.; Pun, E.Y.B.; Ma, T.; Lin, H. Pr3+-doped phosphate glasses for fiber amplifiers operating at 1.38–1.53 µm of the fifth optical telecommunication window. Opt. Mater. 2014, 36, 1203–1208. [Google Scholar] [CrossRef]
  80. Zhou, B.; Tao, L.; Tsang, Y.H.; Jin, W.; Pun, E.Y.B. Superbroadband near-IR photoluminescence from Pr3+-doped fluorotellurite glasses. Opt. Express 2012, 20, 3803–3813. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Transmittance spectra measured for the studied glass samples. IZSBGL-Pr (a), GBG-Pr and GBFG-Pr (b), BBG-Pr (c) and PPG-Pr (d).
Figure 1. Transmittance spectra measured for the studied glass samples. IZSBGL-Pr (a), GBG-Pr and GBFG-Pr (b), BBG-Pr (c) and PPG-Pr (d).
Materials 15 00767 g001
Figure 2. Energy level diagram of praseodymium ions in inorganic glasses.
Figure 2. Energy level diagram of praseodymium ions in inorganic glasses.
Materials 15 00767 g002
Figure 3. Absorption (a,b) and visible emission (c,d) spectra and emission decay curves (e,f) for Pr3+ ions in inorganic glasses.
Figure 3. Absorption (a,b) and visible emission (c,d) spectra and emission decay curves (e,f) for Pr3+ ions in inorganic glasses.
Materials 15 00767 g003
Figure 4. CIE chromaticity coordinates for IZSBGL-Pr (A), BGFG-Pr (B), BGG-Pr (C), PPG-Pr (D), and BBG-Pr (E) glass systems.
Figure 4. CIE chromaticity coordinates for IZSBGL-Pr (A), BGFG-Pr (B), BGG-Pr (C), PPG-Pr (D), and BBG-Pr (E) glass systems.
Materials 15 00767 g004
Figure 5. Reddish orange emission as a function of glass-host (a) and Pr3+ concentration (b).
Figure 5. Reddish orange emission as a function of glass-host (a) and Pr3+ concentration (b).
Materials 15 00767 g005
Figure 6. Near-infrared luminescence spectra of Pr3+ ions in inorganic glasses (a,b). Normalized spectra in 1200–1650 nm range excited at 445 nm (c,d) and 590 nm (e,f) are also indicated.
Figure 6. Near-infrared luminescence spectra of Pr3+ ions in inorganic glasses (a,b). Normalized spectra in 1200–1650 nm range excited at 445 nm (c,d) and 590 nm (e,f) are also indicated.
Materials 15 00767 g006
Table 1. Chemical compositions and melting conditions for inorganic glasses doped with Pr3+.
Table 1. Chemical compositions and melting conditions for inorganic glasses doped with Pr3+.
Glass CodeChemical Composition [mol%]Melting Conditions
IZSBGL-Pr37.9InF3-20ZnF2-20SrF2-16BaF2-4GaF3-2LaF3-0.1PrF3900 °C/60 min
GBFG-Pr60GeO2-25BaO-5BaF2-9.9Ga2O3-0.1Pr2O31200 °C/45 min
GBG-Pr60GeO2-30BaO-9.9Ga2O3-0.1Pr2O31200 °C/45 min
PPG-Pr45PbO-45P2O5-9.9Ga2O3-0.1Pr2O31100 °C/30 min
BBG-Pr60B2O3-30BaO-9.9Ga2O3-0.1Pr2O31250 °C/45 min
Table 2. Chemical compositions and melting conditions for lead-phosphate glass doped with Pr3+.
Table 2. Chemical compositions and melting conditions for lead-phosphate glass doped with Pr3+.
Glass CodeChemical Composition [mol%]Melting Conditions
PPG-0.1Pr45PbO-45P2O5-9.9Ga2O3-0.1Pr2O3 45PbO-45P2O5-9.5Ga2O3-0.5Pr2O3 45PbO-45P2O5-9.0Ga2O3-1.0Pr2O3
45PbO-45P2O5-7.5Ga2O3-2.5Pr2O3
1100 °C/30 min
PPG-0.5Pr1100 °C/30 min
PPG-1.0Pr1100 °C/30 min
PPG-2.5Pr1100 °C/30 min
Table 3. Different physicochemical properties of the studied Pr3+-doped inorganic glasses [56,57,58,59].
Table 3. Different physicochemical properties of the studied Pr3+-doped inorganic glasses [56,57,58,59].
ParametersGlass-Host
IZSBGL-PrGBG-PrPPG-PrBBG-Pr
Average molecular weight (M g mol−1)140.94127.63183.22106.63
Density (d g cm−3)4.384.584.113.19
Pr3+ content (molar %)0.10.10.10.1
Pr3+ concentration (Nx1019 ions cm−3)1.874.312.703.59
Average interionic separation (R Å)18.514.016.314.9
Critical transfer distance (R0 Å)11.36.58.57.8
Refractive index (n)1.481.731.751.61
Glass transition temperature (Tg °C)295620437566
Phonon energy of the host (hω cm−1) 51079011201400
Judd–Ofelt parameters Ωt (10−20 cm2)
Ω22.016.931.812.17
Ω45.2519.6818.339.75
Ω65.108.9515.512.62
Radiative transition rate (AJ s−1)
from 3P0 state (Pr3+)30,200123,05095,25060,450
from 1D2 state (Pr3+)2440893083303370
Quantum efficiency 1D2 Pr3+ (η %)8898505
Table 4. Luminescence lifetimes for 3P0 and 1D2 states of Pr3+ ions in inorganic glasses.
Table 4. Luminescence lifetimes for 3P0 and 1D2 states of Pr3+ ions in inorganic glasses.
Glass-Host Composition [mol%]3P0 [µs]1D2 [µs]Ref.
57ZrF4-34BaF2-4AlF3-4.5LaF3-0.5PrF337-[8]
50SiO2-10Al2O3-2MgO-20CaO-15SrO-3BaO-0.1Pr2O3115-[11]
74.8TeO2-15Sb2O3-10WO3-0.2Pr6O1111.73-[12]
60P2O5-4B2O3-7Al2O3-10K2O-17.95BaO-0.05Pr2O3-173[13]
49.5P2O5-10AlF3-10BaF2-10SrF2-10PbO-10MxOy-0.5Pr6O1110–1114–17[15]
M = Li, Na, K, Zn, Bi
60P2O5-4B2O3-7Al2O3-10K2O-17.9BaO-0.1Pr2O3-137[16]
55SiO2-8B2O3-5Al2O3-14Li2O-2Na2O-10GeO2-5.9Y2O3-0.1Pr2O3-73[16]
75TeO2-20ZnO-5Na2O-0.1Pr2O3-51[16]
69H3BO3-20Li2CO3-10LiF-1Pr2O325.130[17]
60PbO-40GeO2-0.05Pr2O36145[23]
5ZnO-15PbO-20WO3-59TeO2-1Pr6O114.5-[27]
44P2O5-17K2O-9Al2O3-23.9PbO-6Na2O-0.1Pr6O11-66[28]
30PbO-5Bi2O3-64SiO2-1Pr2O369-[31]
25Na2O-5LaF3-10CaF2-10AlF3-49.9B2O3-0.1Pr6O11-51[59]
30Li2CO3-20Al2O3-10B2O3-39.9SiO2-0.1Pr2O385.5108.2[61]
60TeO2-25ZnO-10BaO-4.5La2O3-0.5Pr2O32139[64]
29.95B2O3-30TeO2-16ZnO-10ZnF2-7CaF2-7BaF2-0.05Pr2O34576[65]
10Li2O-10PbO-9.95Al2O3-70B2O3-0.05Pr6O11-165[66]
Table 5. CIE chromaticity coordinates for the studied inorganic glasses doped with Pr3+.
Table 5. CIE chromaticity coordinates for the studied inorganic glasses doped with Pr3+.
Glass CodeCIE Chromaticity Coordinates
(A)IZSBGL-Prx = 0.380; y = 0.327
(B)GBFG-Prx = 0.433; y = 0.370
(C)GBG-Prx = 0.523; y = 0.353
(D)PPG-Prx = 0.582; y = 0.374
(E)BBG-Prx = 0.622; y = 0.366
Table 6. Spectroscopic parameters for Pr3+ ions in the studied inorganic glass systems.
Table 6. Spectroscopic parameters for Pr3+ ions in the studied inorganic glass systems.
Glass CodeSpectroscopic Parameters
σem [10−20 cm2]σem × τm [10−26 cm2s]σem × FWHM [10−27 cm3]
IZSBGL-Pr0.5015465
GBFG-Pr0.98108206
GBG-Pr0.97107201
PPG-Pr1.2877165
BBG-Pr0.37681
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pisarska, J.; Kuwik, M.; Pisarski, W.A. Spectroscopic Properties of Inorganic Glasses Doped with Pr3+: A Comparative Study. Materials 2022, 15, 767. https://doi.org/10.3390/ma15030767

AMA Style

Pisarska J, Kuwik M, Pisarski WA. Spectroscopic Properties of Inorganic Glasses Doped with Pr3+: A Comparative Study. Materials. 2022; 15(3):767. https://doi.org/10.3390/ma15030767

Chicago/Turabian Style

Pisarska, Joanna, Marta Kuwik, and Wojciech A. Pisarski. 2022. "Spectroscopic Properties of Inorganic Glasses Doped with Pr3+: A Comparative Study" Materials 15, no. 3: 767. https://doi.org/10.3390/ma15030767

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