Next Article in Journal
Analysis and Optimization of Mechanical Properties of Laser-Sintered Cellulose/PLA Mixture
Next Article in Special Issue
Transition Metals (Cr3+) and Lanthanides (Eu3+) in Inorganic Glasses with Extremely Different Glass-Formers B2O3 and GeO2
Previous Article in Journal
Optical, Structural, and Dielectric Properties of Composites Based on Thermoplastic Polymers of the Polyolefin and Polyurethane Type and BaTiO3 Nanoparticles
Previous Article in Special Issue
Phonon Sideband Analysis and Near-Infrared Emission in Heavy Metal Oxide Glasses
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Photoluminescence Investigations of Tb3+/Eu3+ Co-Doped Silicate Sol-Gel Glass-Ceramics Containing CaF2 Nanocrystals

by
Natalia Pawlik
1,*,
Barbara Szpikowska-Sroka
1,
Tomasz Goryczka
2,
Joanna Pisarska
1 and
Wojciech A. Pisarski
1
1
Institute of Chemistry, University of Silesia, 40-007 Katowice, Poland
2
Institute of Materials Engineering, University of Silesia, 41-500 Chorzów, Poland
*
Author to whom correspondence should be addressed.
Materials 2021, 14(4), 754; https://doi.org/10.3390/ma14040754
Submission received: 12 January 2021 / Revised: 29 January 2021 / Accepted: 2 February 2021 / Published: 5 February 2021

Abstract

:
In this work, the series of Tb3+/Eu3+ co-doped xerogels and derivative glass-ceramics containing CaF2 nanocrystals were prepared and characterized. The in situ formation of fluoride crystals was verified by an X-ray diffraction technique (XRD) and transmission electron microscopy (TEM). The studies of the Tb3+/Eu3+ energy transfer (ET) process were performed based on excitation and emission spectra along with luminescence decay analysis. According to emission spectra recorded under near-ultraviolet (NUV) excitation (351 nm, 7F65L9 transition of Tb3+), the mutual coexistence of the 5D47FJ (J = 6–3) (Tb3+) and the 5D07FJ (J = 0–4) (Eu3+) luminescence bands was clearly observed. The co-doping also resulted in gradual shortening of a lifetime from the 5D4 state of Tb3+ ions, and the ET efficiencies were varied from ηET = 11.9% (Tb3+:Eu3+ = 1:0.5) to ηET = 22.9% (Tb3+:Eu3+ = 1:2) for xerogels, and from ηET = 25.7% (Tb3+:Eu3+ = 1:0.5) up to ηET = 67.4% (Tb3+:Eu3+ = 1:2) for glass-ceramics. Performed decay analysis from the 5D0 (Eu3+) and the 5D4 (Tb3+) state revealed a correlation with the change in Tb3+–Eu3+ and Eu3+–Eu3+ interionic distances resulting from both the variable Tb3+:Eu3+ molar ratio and their partial segregation in CaF2 nanophase.

1. Introduction

In recent years, the materials doped with rare earths (RE3+) are considered to be indispensable in the development of optoelectronics, offering promising applications in LEDs [1], displays [2], lasers [3], or optical thermometry [4]. The proper adjustment of emission (i.e., generation of appropriate color purity and maintaining suitable luminescence lifetimes) usually requires the involvement of several RE3+ ions incorporated into the same host material [5,6,7,8,9,10,11,12]. Indeed, the mutual interactions between them—depending on the concentration of donor and acceptor as well as the type of host—allow for successful tailoring of the above-mentioned optical parameters.
Among numerous combinations of RE3+ in doubly- or triply-doped materials, the optical system consisting of Tb3+ and Eu3+ is a very promising strategy for the generation of multicolor luminescence, which plays a key role in the development of red-green-blue (RGB) optical materials. The Tb3+/Eu3+ energy transfer has been extensively explored and described in various types of phosphors, e.g., LaBWO6 [13], Tb2MoO3O12 [14], ScPO4 [15], or KAlP2O7 [16]; meanwhile, such studies have not been as common for glass-ceramic materials so far, where Tb3+ and Eu3+ ions could be distributed between the amorphous glassy host and crystal phase characterized by different decay rates. The evidence of Tb3+/Eu3+ ET is stated by the shortening of a lifetime from the 5D4 state of Tb3+ in the presence of acceptor ions (Eu3+). It is reported in the literature that the fluorescence decay becomes quicker with the increment of Eu3+ content, which accelerates the Tb3+/Eu3+ ET and makes it more efficient [17,18,19,20,21,22]. Furthermore, a comparative analysis of energy transfer efficiency, ηET, in precursor glasses and glass-ceramics is also carried out to demonstrate the impact of controlled crystallization on Tb3+/Eu3+ ET. Such studies were performed in excellent work by Chen et al. [19] for 44SiO2-28Al2O3-17NaF-(10-x)YF3-1TbF3-xEuF3 (x = 0, 0.1, 0.25, 0.5, 1) (mol%) glasses and derivative glass-ceramics fabricated at 670 °C. The ηET for a glass containing 0.1 mol% Eu3+ achieved as low value as 1.39%, which finally grew to 30.28% for a glass containing 1 mol% Eu3+. Further, the authors have clearly proven that a crucial role in Tb3+/Eu3+ ET plays is glass crystallization, which results in significant growth in ηET values from 16.63% (for glass-ceramic containing 0.1 mol% Eu3+) up to 47.70% (for glass-ceramic containing 1 mol% Eu3+). Similarly, an impact of controlled ceramization on Tb3+/Eu3+ ET behavior was also studied by Hu et al. [21], who found that ηET increased from 8.7% for glasses with composition 45SiO2-20Al2O3-10CaO-24.04CaF2-0.05TbF3-0.01EuF3 (mol%) up to 14.0% for nano-glass-ceramic produced at 700 °C.
It should be pointed out that the majority of glass-ceramics containing Tb3+ and Eu3+ ions characterized and described in literature was prepared by the conventional melt-quenching method followed by controlled heat-treatment at the specified conditions of time and temperature [17,18,19,21,22,23,24,25,26]. An alternative route to the fabrication of glass-ceramics is the sol-gel technique, which offers quite easy fabrication of materials with complex compositions, which could be difficult to obtain via the melt-quenching technique [27,28,29,30,31]. Moreover, particular research attention has been focused on oxyfluoride glass-ceramics, which possess higher chemical and mechanical stability than fluoride glasses and lower phonon energies than oxide glasses. Among fluorides, the calcium fluoride, CaF2, is an optically isotropic crystal characterized by a broad region of transparency from 0.13 up to 9.5 μm, wide bandgap (12 eV), and relatively low phonon energy (~466 cm−1) [32,33,34]. Those features of CaF2 crystals are urgent to be a suitable medium for optically active rare earths, widely dedicated to fulfilling many sophisticated, active functions for optoelectronics. Indeed, the optical materials based on the CaF2 phase are frequently applied to generate an efficient up- [35] and down-conversion luminescence [36] or white light emission [37]. Therefore, such materials could be successfully predisposed for use in laser technologies [38], bio-imaging [39], or to increase the efficiency of solar cells [40]. Moreover, according to our previous results for sol-gel nano-glass-ceramics containing divalent metals fluorides, MF2 (M = Ca, Sr, Ba), singly-doped with Eu3+ ions, the most efficient segregation of Eu3+ inside fluoride crystal lattice, was reported for the SiO2-CaF2 system [41,42,43]. Indeed, a clear correlation was observed between the average decay time of the 5D0 state and growing difference in ionic radius of Eu3+ and each of individual M2+ cation in the following order: Ca2+ → Sr2+ → Ba2+ (SiO2-CaF2:Eu3+: τavg = 11.92; SiO2-SrF2:Eu3+: τavg = 7.77; SiO2-BaF2:Eu3+: τavg = 4.08 ms). Therefore, due to the efficient and long-lived luminescence in fabricated SiO2-CaF2:Eu3+ nano-glass-ceramics, it seems that this material could be considered as a very promising host to study the Tb3+/Eu3+ ET. Indeed, we reported interesting results for Tb3+/Eu3+ ET for sol-gel systems containing selected trivalent metals fluorides, MF3 (M = Y, La) [44]; hence, we performed such measurements for nano-glass-ceramics with divalent metal fluoride, CaF2. Moreover, to the best of our knowledge, the investigation of Tb3+/Eu3+ ET in oxyfluoride sol-gel glass-ceramics is rarely described in the available literature. Indeed, it was examined only in SiO2-SrF2 [20] and SiO2-BaGdF5 [45] sol-gel nano-glass-ceramics; however, there are no reports on the SiO2-CaF2 system so far. Due to the above reasons, it seems justified to study the Tb3+/Eu3+ ET in silicate sol-gel glass-ceramics containing CaF2 nanocrystals.
In this work, we fabricated and examined the series of sol-gel SiO2-CaF2 nano-glass-ceramics co-doped with Tb3+ and Eu3+ ions with the variable Tb3+:Eu3+ molar ratio (0.05:x, where x = 0.025, 0.05, 0.075, 0.1). The CaF2 phase was obtained via in situ crystallization from Ca(CF3COO)2 at as low a temperature as 350 °C, and its presence was verified using XRD measurements and TEM microscopy. The changes in photoluminescence behavior of fabricated sol-gel materials have been described in association with the variable Tb3+:Eu3+ molar ratio, as well as the structural transformation from amorphous xerogels into nano-glass-ceramics. Based on photoluminescence results, the interactions between Tb3+ and Eu3+ ions were systematically investigated. Indeed, a clear correlation was observed between the R/G ratio and energy transfer efficiency (ηET), as well as decay times of the 5D4 state (Tb3+) as the Tb3+:Eu3+ molar ratio gradually decreased. Additionally, the lifetimes of the 5D0 excited level of Eu3+ ions were also determined. The obtained sol-gel materials exhibited bright multicolor luminescence tuned when the Tb3+:Eu3+ molar ratio was changed.

2. Materials and Methods

The series of xerogels co-doped with Tb3+ and Eu3+ ions were synthesized using the sol-gel method described elsewhere [41,46]. All reagents were taken from Sigma Aldrich Chemical Company (St. Louis, MO, USA). After pre-hydrolysis of the mixture containing TEOS, ethyl alcohol, deionized water, and acetic acid in molar ratio equals 1:4:10:0.5 (90 wt.%), the solutions of Ca(CH3COO)2, Tb(CH3COO)3, and Eu(CH3COO)3 in water and trifluoroacetic acid (TFA) were added dropwise. For each sol-gel sample, a mixture of TFA and acetates was 10 wt.%, and the molar ratio equaled TFA:Ca(CH3COO)2:Tb(CH3COO)3: Eu(CH3COO)3 = 5:1:0.05:x (where x = 0, 0.025, 0.05, 0.075, and 0.1). The obtained sols were dried at 35 °C for seven weeks to form solid xerogels. The following xerogel samples were denoted as follows: XG1Tb0.5Eu (x = 0.025), XG1Tb1Eu (x = 0.05), XG1Tb1.5Eu (x = 0.075), and XG1Tb2Eu (x = 0.1). The glass-ceramic materials were obtained by controlled heat-treatment at 350 °C for 10 h. Such glass-ceramics were denoted as appropriate GC samples: GC1Tb0.5Eu, GC1Tb1Eu, GC1Tb1.5Eu, GC1Tb2Eu. The sol-gel samples singly-doped with Tb3+ ions were also prepared (XG1Tb, GC1Tb) to compare with the luminescence properties of Tb3+/Eu3+ co-doped materials.
The sol-gel network’s vibrational modes were identified using the Nicolet iS50 ATR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) in a frequency region 500–4000 cm−1. The X-ray diffraction analysis of fabricated xerogels and glass-ceramics was performed using an X’Pert Pro diffractometer supplied by PANalytical with CuKα radiation (Almelo, The Netherlands). The microstructure of fabricated glass-ceramics was observed using the JEOL JEM 3010 electron transmission microscope operated at 300 kV (JEOL JEM 3010, Tokyo, Japan). The excitation and emission spectra, as well as luminescence decay curves were recorded on Horiba Jobin Yvon FluoroMax-4 spectrofluorimeter supplied with 150 W Xe lamp (Horiba Jobin Yvon, Longjumeau, France). The spectra were recorded with 0.1 nm resolution, and the decay curves were recorded with 2 μs accuracy. All structural and luminescence measurements were carried out at room temperature.

3. Results and Discussion

3.1. Structural Characterization: XRD, TEM, and IR Spectroscopy

In general, fabricated sol-gel materials’ structural properties strongly determine the local environment around Tb3+ and Eu3+, which are crucial in understanding any changes in their photoluminescence behavior (i.e., emission spectra and decay profiles). The detailed studies in this aspect (the structural evolution from sols, through gels, and xerogels, up to nano-glass-ceramics) were systematically investigated and described for the similar SiO2-LaF3:Eu3+ system in our previous work [47]. Therefore, to explain the luminescence features of fabricated SiO2-CaF2:Tb3+, Eu3+ samples, a brief comment on their structural properties was also presented below.
The performed heat-treatment of xerogels is responsible for both in situ crystallization of CaF2 nanophase (verified by XRD and TEM measurements, Figure 1) and evolution of the silicate sol-gel network (indicated by IR-ATR measurements, Figure 2). Such measurements were performed for representative XG1Tb1Eu and GC1Tb1Eu co-doped samples. As was demonstrated in Figure 1, a broad halo pattern was recorded for xerogel, which indicates an amorphous nature without long-range structural order. The diffraction reflexes characteristic for the CaF2 phase crystallized in Fm3m space group (ICDD, PDF-2 No. 65-0.0535) were observed after controlled ceramization. The broadening of recorded diffraction lines indicates the crystallization of the CaF2 phase in nanoscale, the average crystal size of which was estimated to be 8.1 ± 0.4 nm from the Scherrer formula:
D   =   K λ B cos θ
where D is the crystal size, K is a constant value (in our calculations, we took K = 1), λ is the X-ray wavelength (1.54056 Å, CuKα), B is a half of the diffraction line, and θ is the diffraction angle [48]. Another method that allowed us to determine the crystallites’ size, which is an extension of the Scherrer equation, was the Williamson–Hall method, based on plotting the following dependence for several reflexes derived from the same crystalline phase:
β cos θ =   K λ D   +   4 Δ a a sin θ
in which β is a half of the diffraction line and (Δa/a) refers to the lattice deformation [49]. From the Williamson–Hall plot’s linear fit, the crystallite size was estimated to be 16.6 ± 1.5 nm (the “chi-square” regression coefficient was equaled to 0.97). Since the Williamson–Hall approach considers crystal imperfections and lattice distortion, as well as apparatus factors, this method allowed us to determine the mean crystallite size more reliably than the Scherrer method (indeed, the latter does not consider such distortions of the crystal lattice). Indeed, the size of CaF2 nanocrystals from the TEM image (inset of Figure 1) was in more prominent agreement with the data obtained by the Williamson–Hall method (16.6 ± 1.5 nm) than by the Scherrer equation (8.1 ± 0.4 nm). It should be noted that the average crystal size estimated by the Williamson–Hall method was two-fold larger than from the Scherrer equation. Taking into account the factors related to the prepared sample itself, we suppose that one of the reasons for the discrepancy in the values estimated by various methods, apart from the differences in the ionic radii of Ca2+ (1.00 Å) [50], Tb3+ (1.04 Å), and Eu3+ (1.07 Å) [51], could be related to the charge compensation when trivalent cations were substituting divalent Ca2+ in the CaF2 crystal lattice. Indeed, to balance an excess of the positive charge introduced by the RE3+ ion, the F- anions were distributed in interstitial positions [52]. Such interstitial F- anions induced a local distortion in the crystal lattice due to the repulsion between them. Therefore, we assumed that the charge compensation effect might contribute to the identified difference in crystal size.
Besides the crystallization of the CaF2 nanophase, the track of structural changes inside the sol-gel host is also important to explain Tb3+ and Eu3+ dopant ions’ luminescence properties. The sol-gel systems were dynamic during controlled heat-treatment at 350 °C performed for the next 10 h because it induced evaporation of residual solvents (water and organic liquids) from a microporous structure and polycondensation reaction with the participation of Q2 ([Si(O1/2)2O2]−2), Q3 ([Si(O1/2)3O]), and Q4 units ([Si(O1/2)4]) (in the description “O1/2” is corresponding to each oxygen atom, which is involved in the formation of Si–O–Si bridge; however, “O” is according to a non-bridging oxygen atom; therefore higher value of “n” index is, the less Si–OH unreacted groups are). To examine such structural evolution, the IR-ATR spectra were recorded in the frequency region from 500 to 4000 cm−1 (Figure 2), and the bands were assigned to appropriate vibrational modes based on literature data [53,54,55,56,57,58]. Generally, we distinguished three primary regions classified by the functional groups causing the characteristic vibrations: 3750–2500 cm−1 (OH groups and C–H bond), 1820–1510 cm−1 (C=O bond, Si–OH groups, and adsorbed H2O), and 1275–500 cm−1 (silicate host). For a more in-depth interpretation of oscillations that occurred in the sol-gel network, deconvolution was performed. The peak fitting during the deconvolution was done using a Gauss function with a squared regression coefficient of R2 ≥0.998.
Firstly, we analyzed the origin of deconvoluted bands (denoted as A–E) identified within the 1275–620 cm−1 spectral region. It was established that band A (~1200 cm−1) was corresponding to the TO4 mode of created Si–O–Si siloxane bridges, band B (~1140 cm−1) was correlated with oscillations within Q4 structural units ([Si(O1/2)4]), and band C (~1070 cm−1) could be associated with the TO3 mode of Si–O–Si siloxane bridges. Bands D (~1030 cm−1) and E (~960 cm−1) were related to Q3 and Q2 units’ oscillations, respectively [53,54,56]. According to the literature, bands A and B could also be interpreted as vibrations originated from the C–F bond inside –CF3 groups in trifluoroacetates [55]. It was clearly visible that the intensities of bands A, B, and E decreased during the transformation of xerogels into glass-ceramics. Indeed, since Ca(CF3COO)2 underwent thermal decomposition during controlled ceramization [41], the vibrations from the C–F bond disappeared for glass-ceramics, and bands A and B should have originated only from vibrations inside the Si–O–Si siloxane bridges. Further, a decrease in the intensity of band E clearly pointed to the conversion of Q2 structural units into Qn (n = 3, 4) units as a consequence of polycondensation reaction. The additional weak bands located at the 616 and 561 cm−1 frequency region indicated some cyclic structures inside the sol-gel host [53,56].
Finally, we analyzed the broad band recorded in a frequency region from 3750 up to 2500 cm−1. The deconvolution revealed the presence of three bands originated from different types of OH moieties: Geminal or vicinal Si–OH groups (~3630 cm−1, band F), hydrogen-bonded Si–OH groups (~3450 cm−1, band G), and hydrogen-bonded OH groups originated from residual water and organic compounds (~3200 cm−1, band H). It should be noted that deconvolution also revealed the fourth component band (~2990 cm−1, band I), which corresponded to the vibrations of C–H bonds [53,54,56]. It was observed that an indicated broad band was much more intense for xerogels than for glass-ceramics (indeed, to show the deconvolution better, the band’s intensity was fivefold enlarged). In fact, such a relatively strong band for xerogels was a consequence of “trapping” of water and organic liquids inside pores via hydrogen-bonding created with unreacted Si–OH groups. During ceramization, the band was significantly reduced, which clearly evidenced successful evaporation of water and organic liquids as well as a continuation of polycondensation reaction. The conclusions from the above observations could also be confirmed by the behavior of the infrared signal located at ~1660 cm−1, which was attributed to the vibrations of the C=O bond, Si–OH surface groups, and adsorbed water [55,57,58]. The indicated band was well-observable for xerogels, and it almost completely disappeared for glass-ceramics. According to IR and XRD results, the graphical visualization of sol-gel evolution during performed ceramization at 350 °C, as presented in Figure 3.

3.2. Luminescence Properties of Fabricated Sol-Gel Materials

3.2.1. Determination of Local Symmetry Using Spectroscopy of Eu3+ Ions as Spectral Probes

The emission spectra recorded for fabricated XG1TbxEu xerogels using λexc = 395 nm excitation line were shown in Figure 4. The characteristic 5D07FJ luminescence bands of Eu3+ ions were detected in the reddish-orange light area and their maxima were located at following wavelengths: 578 (J = 0), 592 (J = 1), 615 (J = 2), 647 (J = 3), and 698 nm (J = 4). A gradual increase in their intensity was observed, when the Tb3+:Eu3+ molar ratio systematically changed from 1:0.5 to 1:2. It is clearly visible that among series of recorded bands, the 5D07F2 one was the most prominent line for all xerogels before their controlled ceramization, despite the Tb3+:Eu3+ molar ratio. Indeed, Eu3+ ions were frequently used as spectral probes due to the characteristic nature of their transitions. The 5D07F1 is a magnetic-dipole transition (MD) in nature, the intensity of which is rather independent of the host. In contrast, the 5D07F0.2–4 are electric-dipole transitions (ED) known to be forbidden by the Laporte selection rule and may occur due to mixing the 4f orbitals with the opposite parity at the low-symmetry sites. Among ED transitions, the 5D07F2 one has a hypersensitive nature, and its intensity is easily affected by the local vicinity: It is promoted in low-symmetric frameworks; meanwhile, it is inhibited in more symmetric environments. Hence, we could infer about the symmetry based on the ratio of integrated intensities of the bands mentioned above, which is well-known in literature as the R/O ratio (I(5D07F2)/I(5D07F1)) [59,60]. XRD results confirmed the amorphous nature of xerogels without long-range order, so we expected relatively high asymmetry in the immediate vicinity of Eu3+. Indeed, the calculated R/O ratio values hesitated from 3.50 (XG1Tb2Eu) to 3.91 (XG1Tb0.5Eu). On this occasion, it should also be noted that comparability in calculated R/O ratio values clearly pointed to a chemically similar environment of Eu3+ ions in all synthesized xerogels.
The emission spectra recorded for GC1TbxEu glass-ceramics upon excitation at λexc = 394 nm were presented in Figure 5. Similar as for xerogels, a series of characteristic bands corresponding to the intraconfigurational transitions within 4f6 manifold were registered: 5D07F0 (577), 5D07F1 (592), 5D07F2 (612), 5D07F3 (648) and 5D07F4 (683/689/698 nm). It was clearly observed that the intensity of the emission bands successfully increased with the growing content of Eu3+ ions (as the Tb3+:Eu3+ molar ratio changed from 1:0.5 to 1:2). For each glass-ceramic, the orange emission band corresponding to the 5D07F1 MD transition maintained the greatest intensity and dominated the 5D07F2 ED red band. Generally, an almost six-fold decline of R/O ratio value was observed, which equaled close to 0.64 after controlled ceramization (compared with xerogels for which the R/O ratio value was approximately equal to 3.70). Hence, such a decrease in the R/O ratio value pointed to significant changes in the symmetry in the immediate vicinity of Eu3+ ions, as well as a change in the nature of the bonding character from covalent to more ionic [61]. When the nearest framework of Eu3+ ions was more symmetric (which usually accompanied the migration of Eu3+ from an asymmetric amorphous structure without long-range order to crystalline lattice), the probability of the 5D07F0,2–4 electric-dipole transitions successfully decreased [60]. Indeed, the identified decrease in the R/O ratio value was strictly accompanied by partially entering of optically active ions into precipitated CaF2 fluoride crystal fraction with long-range order. In other words, such a decline was direct evidence that Ca2+ cations from fluoride crystal lattice were successfully substituted by Eu3+ ions. It should also be noted that a Stark splitting characteristic in the crystal-field was not observed because part of the Eu3+ ions was still distributed in an amorphous sol-gel host [60].

3.2.2. Studies of Tb3+/Eu3+ Energy Transfer in Sol-Gel Materials with the Variable Tb3+:Eu3+ Molar Ratio

To select the excitation line appropriate for Tb3+/Eu3+ ET studies, the photoluminescence excitation spectra for XG1Tb1Eu co-doped representative xerogel were presented in Figure 6. The spectra were recorded for λem = 543 and λem = 612 nm emissions of Tb3+ (the 5D47F5 green line), and Eu3+ (the 5D07F2 red line), respectively. The recorded excitation bands of Tb3+ ions were associated to the following f-f intraconfigurational transitions: 7F65L9 (352), 7F65L10 (370), 7F65D3 (379), as well as 7F65D4 (487 nm). Simultaneously, the recorded bands were ascribed to the transitions of Eu3+ ions from the 7F0 ground state into the subsequent higher energy levels: 5D4 (363), 5GJ, 5L7 (from 371 to 390), 5L6 (395), and 5D2 (465 nm). Since the 7F65L9 transition of Tb3+ ions did not coincide with any excitation peak of Eu3+, we decided to select the λexc = 351 nm excitation line to study the Tb3+/Eu3+ ET process.
The photoluminescence spectra recorded for XG1Tb xerogel (upon excitation at λexc = 351 nm wavelength) as well as for the XG1Tb1Eu co-doped representative sample (recorded under excitation at λexc = 395 and λexc = 351 nm lines) are depicted in Figure 7. The spectra recorded for the XG1Tb sample revealed two emission bands in bluish-green spectral scope, i.e., 5D47F6 (488) and 5D47F5 (542 nm) of the predominant intensity, as well as two other emission bands of Tb3+ ions in the yellowish-red area: 5D47F4 (582) and 5D47F3 (620 nm). When the XG1Tb1Eu co-doped sample was excited by the λexc = 395 nm line, only the characteristic 4f6-4f6 emission bands originated from Eu3+ ions (5D07FJ, J = 0–4) were recorded. A tune in the excitation wavelength to λexc = 351 nm also led to the generation of the characteristic emission lines of Tb3+ ions. Such coexistence of emission lines originated from both optically active dopants is due to the energy transfer process from Tb3+ to Eu3+ [17,18,19,20,21,22]. In the case of the band recorded in a red spectral scope, a slight shift was observed of a maximum from 620 (XG1Tb) to 618 nm (XG1Tb1Eu), which was caused by overlapping the weak 5D07F2 band of Eu3+ ions (with a maximum at 615 nm) with the 5D47F3 band of Tb3+ ions (with a maximum at 620 nm). In general, the spectral matching of the donor’s emission (Tb3+) and the acceptor’s excitation (Eu3+) regions was a fundamental condition for energy transfer occurrence [62]. In this way, upon irradiation using the λexc = 351 nm line from NUV spectral region, Tb3+ ions could be successfully pumped into the 5L9 level, and then, the non-radiative de-activation to the 5D4 state took place. The excitation energy from the 5D4 state (Tb3+) could be successfully transferred into the 5D1 or the 5D0 level (Eu3+). Hence, among characteristic emission lines from Tb3+ ions, additional bands originated from Eu3+ can also be recorded.
The emission spectra of Tb3+/Eu3+ co-doped xerogels with the varying Tb3+:Eu3+ molar ratio recorded upon excitation at λexc = 351 nm are presented in Figure 8. A slight decrease was observed in the emission intensity of the 5D47F6 and the 5D47F5 bands of Tb3+ ions when the Tb3+:Eu3+ molar ratio gradually decreases. Hence, the R/G ratio values ((I(5D07F2)(Eu3+)+I(5D47F3)(Tb3+)/I(5D47F5)(Tb3+)) were estimated. For the XG1Tb sample, the R/G ratio was defined as the ratio of integrated intensities of the 5D47F3 red band and the 5D47F5 green emission line [63]. In the case of the XG1TbxEu co-doped samples, an additional contribution of luminescence originated from Eu3+ ions into total red emission should also be taken into account. Hence, the R/G ratio was calculated as (I(5D47F3)(Tb3+) +I(5D07F2)(Eu3+)/I(5D47F5)(Tb3+)) and its increase can be interpreted as a growing share of Eu3+ ions in total generated multicolor luminescence. Indeed, a slight increase in R/G ratio values was determined in the following order: From 0.09 (XG1Tb), through 0.15 (XG1Tb0.5Eu), 0.17 (XG1Tb1Eu), 0.26 (XG1Tb1.5Eu), to 0.30 (XG1Tb2Eu). Such an increment of the R/G ratio suggests more efficient Tb3+/Eu3+ ET when the Eu3+ content gradually grew, since the Tb3+:Eu3+ molar ratio changed from 1:0.5 to 1:2. Nevertheless, such a small increase in the R/G ratio resulted from the relatively large interionic distances between Tb3+ and Eu3+ ions, characteristic for the amorphous xerogel host.
Additionally, an increasing background for prepared sol-gel samples was also observed, especially at wavelengths <540 nm. Such a background was associated with the wide band from the silicate sol-gel host, as was shown by other authors, e.g., Tomina et al. [64], for different types of Eu3+-loaded aminosilica spherical particles. The authors suggest that such a band could result, i.e., from the charge transfer on Si-O bonds or defect from the silicate network. They have proven that such a wide band’s intensity is strictly related to the type of complexes formed by Eu3+ ions with amine ligands connected with the silicate sol-gel host. A similar effect was reported by Kłonkowski et al. [20], who synthesized sol-gel glass-ceramics containing SrF2 singly- and doubly-doped with Tb3+/Eu3+. The broad band’s origin was explained by defects, like dangling bonds, inside silicate sol-gel host.
The emission spectra recorded for GC1Tb (upon excitation at λexc = 351 nm wavelength) and GC1Tb1Eu co-doped the representative sample (recorded under excitation at λexc = 394 nm and λexc = 351 nm lines) are depicted in Figure 9. Similarly as for xerogel, for the GC1Tb sample, the characteristic emission bands corresponding to the transitions from the 5D4 excited level into the 7F6 (488), 7F5 (542), 7F4 (581), and 7F3 (621 nm) lower-lying states were detected. In the case of the GC1Tb1Eu sample, the coexistence of the luminescence lines originating from both rare-earth dopants was clearly observed after excitation at λexc = 351 nm line. Therefore, an appearance of characteristic emission bands coming from Eu3+ ions upon excitation of Tb3+ confirmed the occurrence of Tb3+/Eu3+ ET. It should be particularly pointed out that the intensity of Tb3+ emission strongly decreased, accompanied by significant enhancement of Eu3+ luminescence. Additionally, the maxima of bands recorded the in 570–630 nm spectral scope were shifted from 581 (for GC1Tb sample) to 592 nm (for GC1Tb1Eu sample) for an orange band and from 621 (for GC1Tb sample) up to 612 nm (for GC1Tb1Eu sample) for a red band. Indeed, an enhancement of Eu3+ emission via Tb3+/Eu3+ ET was much more effective for glass-ceramics than for xerogels.
The emission spectra recorded under λexc = 351 nm for GC1TbxEu co-doped glass-ceramics are depicted in Figure 10. Based on the collected data, it was established that intensities of the 5D47FJ (J = 5,6) bands of Tb3+ in the bluish-green spectral scope are strongly dependent on the Tb3+:Eu3+ molar ratio. Indeed, the intensity of the Tb3+ emission was strongly reduced when the concentration of the acceptor gradually increased, and such an effect was simultaneously accompanied by a well-observable increase in the intensity of Eu3+ emission within the reddish-orange spectral scope. The observed correlations in mutual intensities of characteristic emission bands were accompanied by an adequate increase in R/G ratio values from 0.14 (GC1Tb) and 0.80 (GC1Tb0.5Eu), through 1.60 (GC1Tb1Eu), 2.47 (GC1Tb1.5Eu), and up to 3.76 (GC1Tb2Eu). Therefore, the increment in the calculated R/G ratio values was more dynamic for glass-ceramics than for xerogels, for which only a slight increase was reported when the Tb3+:Eu3+ molar ratio decreased. Such a correlation was undoubtedly associated with the decreased interionic distance between Tb3+ and Eu3+ ions due to their successful entering into the CaF2 crystal lattice.
For Tb3+ ions, the G/B ratio analysis defined as I(5D47F5)/I(5D47F6) could also be treated as a useful tool for characterization of the symmetry around Tb3+ dopant ions [65]. Since the 5D47F5 line is a magnetic-dipole in nature and the 5D47F6 transition is an electric-dipole one, the G/B ratio value should have changed when xerogels were transformed into glass-ceramic counterparts. Hence, the G/B ratio values should have been higher in more centrosymmetric sites [66]. Indeed, some changes in emission lines originated from Tb3+ ions could also be observed, similarly as for Eu3+. On the other hand, it should be pointed out that the G/B ratio was not as sensitive a spectroscopic probe as the R/O ratio calculated for Eu3+ optically active ions. The G/B ratio was calculated for samples singly-doped with Tb3+ ions, and the ratio changes from 2.95 (XG1Tb) to 3.80 (GC1Tb). The results were consistent with the data presented by us earlier in our previous work, concentrating on Tb3+-doped sol-gel materials’ photoluminescence behavior [67]. Based on structural changes undergone during controlled ceramization at 350 °C, the identified differences in G/B ratio values were clearly related to the migration of Tb3+ ions from the amorphous silicate sol-gel network into CaF2 nanocrystals formed during controlled heat-treatment.

3.2.3. Effect of Changing in the Tb3+:Eu3+ Molar Ratio on Decay Times of the 5D4 (Tb3+)

The further evaluation of Tb3+/Eu3+ ET in fabricated sol-gel materials was based on the decay analysis of the 5D4 excited state of Tb3+ ions. Firstly, the interpretation of collected data allowed us to establish a clear correlation between the decay profile (mono- or double-exponential) and type of sol-gel material (i.e., xerogel or glass-ceramic). Indeed, the curves recorded for xerogels were well-fitted to a first exponential decay mode described by the following equation:
I t   =   I 0 exp t / τ
where I(t) and I0 are the luminescence intensities at time t and t = 0, respectively, while τ is the luminescence decay time [68]. Factually, in our xerogels, the rare-earths were chemically bonded with OH moieties and CF3COO anions in complex compounds [69]. It should be noted that high vibrational energies characterize such ligands, i.e., >3000 (OH groups) and ~1200, ~1140 cm−1 (CF3COO anions) as was demonstrated in the Structural characterization: XRD, TEM, and IR spectroscopy section (Figure 2). According to the energy gap law, the effective phonons with maximum energy located in a local surrounding of RE3+ ions (ħωmax) generate the strongest effect on decay times [70]. In this case, since OH moieties’ vibrational energy was the highest, they played a major role in the non-radiative depopulation of excited states. For glass-ceramics, the decay curves were well-fitted to a second exponential decay mode, which can be expressed by the equation:
I t / I 0 = A 1 exp t / τ 1 + A 2 exp t / τ 2
where A1 and A2 are amplitudes, while τ1 and τ2 are the decay times of short and long lifetime components, respectively [68]. The double-exponential decay profile, as well as considerable differences in τ1 and τ2 values, allowed us to conclude about the distribution of rare-earths between two chemically distinct surroundings characterized by different phonon energies. In fact, part of the RE3+ ions migrated during the controlled heat-treatment into the CaF2 crystal lattice, and formed inside the amorphous sol-gel network as a new chemical environment with low phonon energy (~466 cm−1). Due to such a low phonon energy of the CaF2 lattice, the multiphonon non-radiative depopulation of excited states was strongly restricted. However, the remainder of rare-earths was located in an amorphous sol-gel host. According to IR-ATR spectra recorded for glass-ceramics (Figure 2), it was observed that an intensity of the broad infrared signal originated from OH moieties was significantly reduced; therefore, a major role in non-radiative relaxation was attributed to Q3 groups (~1030 cm−1). Nevertheless, their phonon energy was greater than that of CaF2 crystal lattice. Such differences in phonon energies in the nearest surrounding of rare-earths determined the variable rates of radiative depopulation of their excited states: In silicate sol-gel host, the lifetimes were shorter (τ1 components), while in the CaF2 crystal lattice, the lifetimes were prolonged (τ2 components). Based on such distinguished lifetime components and their relative contributions to the total radiative decay profile, the average luminescence lifetime could be calculated using the following formula [71]:
τ avg =   A 1 τ 1 2   +   A 2 τ 2 2 A 1 τ 1   +   A 2 τ 2
The luminescence decay curves of the 5D4 state (Tb3+) recorded for XG1Tb, GC1Tb, as well as for individual XG1TbxEu, GC1TbxEu co-doped sol-gel samples are measured and plotted in Figure 11 and Figure 12, respectively. The decay curves were recorded upon λexc = 351 nm excitation and monitoring λem = 541 nm green luminescence of Tb3+ ions. A slight shortening of the decay lifetime was observed for xerogels from 1.18 (XG1Tb) to 1.04, 1.01, 0.96, and 0.91 ms for XG1Tb0.5Eu, XG1Tb1Eu, XG1Tb1.5Eu, and XG1Tb2Eu, respectively. For glass-ceramics, a change in the Tb3+:Eu3+ molar ratio from 1:0.5 to 1:2 resulted in significantly more efficient shortening of the average decay time of the 5D4 state from 4.75 (GC1Tb) to 3.75 (GC1Tb0.5Eu), 2.59 (GC1Tb1Eu), 1.92 (GC1Tb1.5Eu), and 1.55 ms (GC1Tb2Eu) (the individual values of τ1 and τ2 components are depicted in Table 1 and Table 2). Shorter luminescence lifetimes for xerogels than for glass-ceramics (when we compare the samples with the same Tb3+:Eu3+ molar ratio) were caused by the coordination of Tb3+ ions by high-vibrational OH groups, involved in the non-radiative depopulation of the 5D4 level. Their remarkable removal during controlled heat-treatment and partial segregation of Tb3+ inside CaF2 nanocrystals with low-phonon energy allowed the share of non-radiative processes in relaxation to be reduced significantly; hence, the lifetimes from the 5D4 state for glass-ceramics were longer.
Since the shortening in luminescence lifetimes was caused by an energy transfer from Tb3+ towards Eu3+ ions, the ratio of luminescence lifetimes of the 5D4 state of Tb3+ ions in the presence (τ) and the absence of Eu3+ ions (τ0) could be used as a valuable tool to estimate the energy transfer efficiency [72]:
η ET =   1     τ τ 0   ×   100 %
For xerogels, the ηET values increased from 11.86% (XG1Tb0.5Eu) through 14.41% (XG1Tb1Eu), 18.64% (XG1Tb1.5Eu) to 22.88% (XG1Tb2Eu). Compared to xerogels, a prompt increase in ηET values has been noted for glass-ceramic materials, which reached 25.69% (GC1Tb0.5Eu), 45.47% (GC1Tb1Eu), 59.58% (GC1Tb1.5Eu), and 67.37% (GC1Tb2Eu). Hence, it was easily observed that the energy transfer efficiencies estimated for glass-ceramic samples were noticeably higher than for analogous xerogels with the same Tb3+:Eu3+ molar ratio, which was mainly caused by migration of rare-earths into the CaF2 crystal lattice during controlled heat-treatment, where the interionic Tb3+–Eu3+ distances were significantly shorter than in the amorphous sol-gel host. Additionally, those results clearly indicated a correlation between ηET and a change in the Tb3+:Eu3+ molar ratio from 1:0.5 to 1:2 in fabricated sol-gel materials, which was undoubtedly associated with a higher probability that more Eu3+ ions could be located adjacent to Tb3+. Indeed, since the Tb3+/Eu3+ ET is characterized by dipole-dipole interactions [13,16], ET’s probability is proportional to 1/R6 (R is the average distance between Tb3+ and Eu3+). On this occasion, when Tb3+ and Eu3+ dopants were segregated inside CaF2 nanocrystal lattice, the interionic distances of Tb3+–Eu3+ pairs were vastly shortened notably if the Tb3+:Eu3+ molar ratio changed (from 1:0.5 to 1:2).
For better readability, the correlation between the R/G ratio, ηET, and τ(5D4) lifetimes for prepared sol-gel samples are graphically presented in Figure 13 and depicted in Table 1 (for xerogels) and Table 2 (for glass-ceramics). The R/G ratio gradually increased when the Tb3+:Eu3+ molar ratio changed from 1:0.5 to 1:2, pointing to the increasing share of emissions originated from Eu3+ ions, along with gradual growth in ηET values and shortening of the τ(5D4) decay times of Tb3+ ions. For xerogels and glass-ceramic materials, such a relation was due to the increasing content of Eu3+ ions in accordance with Tb3+. The changes in the values of the parameters mentioned above for glass-ceramics materials were much more significant for each change of the Tb3+:Eu3+ molar ratio, which resulted from the preferential segregation of optically active ions into the CaF2 nanophase.

3.2.4. The Luminescence Decay Analysis of the 5D0 State of Eu3+ Ions

The characterization of luminescence properties of fabricated Tb3+/Eu3+ co-doped sol-gel materials was supplemented by decay analysis of the 5D0 level (Eu3+) upon excitation at λexc = 394 wavelength and monitoring λem = 592 nm (Figure 14). The decay times were also depicted in Table 3 (for xerogels) and Table 4 (for glass-ceramics).
For xerogels, the τ(5D0) lifetime values hesitated from 0.37 (XG1Tb0.5Eu), 0.43 (XG1Tb1Eu), 0.44 (XG1Tb1.5Eu), to 0.45 ms (XG1Tb2Eu). The growing content of Eu3+ caused such a slight increase in decay times in prepared xerogels due to changing the Tb3+:Eu3+ molar ratio from 1:0.5 to 1:2. The relatively short luminescence lifetimes were caused by numerous OH groups in the immediate vicinity of Eu3+ ions in the silicate xerogel host. Interestingly, it was found that the τavg(5D0) lifetimes in glass-ceramics exhibited no evident and straightforward correlation with the increasing content of Eu3+ ions as was found for xerogels. Indeed, the partial segregation of optically active dopants in CaF2 nanocrystals was responsible for the effective shortening of average distances between them and may have caused competition between radiative and non-radiative processes. Comparing the individual τavg(5D0) lifetime values when the λexc = 394 nm wavelength was used as an excitation source, it was easy to observe that changing the Tb3+:Eu3+ molar ratio from 1:0.5 to 1:1 promoted the slight lifetime prolongation (τavg(5D0) = 8.40 for GC1Tb0.5Eu and τavg(5D0) = 8.59 ms for GC1Tb1Eu). Meanwhile, a further change in the Tb3+:Eu3+ molar ratio (1:1.5 and 1:2) caused shortening of the calculated average decay time (τavg(5D0) = 7.94 for GC1Tb1.5Eu and τavg(5D0) = 6.96 ms for GC1Tb2Eu). Since the R/O ratio values were almost the same for all fabricated glass-ceramic samples (from 0.59 to 0.64), we could assume that the relative distribution of Eu3+ ions between CaF2 nanocrystals and the amorphous sol-gel host was comparable in any case. Simultaneously, it also meant that the content of Eu3+ ions in precipitated CaF2 nanocrystals should have been proportional to the total concentration of Eu3+ introduced during the performed synthesis. Such a relation of the decrease in τavg(5D0) values, when the Tb3+:Eu3+ molar ratio equaled 1:1.5 and 1:2, could be explained by the cross-relaxation process. In this case, an excited Eu3+ ion made a downward transition (5D25D1 and/or 5D15D0), whereas a coupled unexcited neighboring Eu3+ ion made an upward transition (7F07F4 and/or 7F07F3) [73]. Such non-radiative relaxation depended on the separation between Eu3+ interacting ions; hence, the shortening in the interionic distance inside CaF2 nanocrystals (promoted when the Tb3+:Eu3+ molar ratio exceeds 1:1) would be predominantly responsible for such a decrease in τavg(5D0). To compare, in the case of Tb3+ ions, we excluded an involvement of the cross-relaxation process on luminescence lifetimes of the 5D4 state based on our previous results for SiO2-PbF2:Tb3+ sol-gel glass-ceramics, for which we reported the non-radiative relaxation mechanism when the molar ratio of Tb3+ (in accordance to Pb2+ cations) exceeded 0.6:1 [67]. On the other hand, since fabricated sol-gel samples were Eu3+ low-concentrated, such shortening of the τavg(5D0) decay times when the Tb3+:Eu3+ molar ratio equaled 1:1.5 and 1:2 could be caused by lattice defects, which are well-known as quenching channels [74]. Indeed, a charge compensation induced the formation of vacancies inside the crystal lattice, the number of which would be greater if greater amounts of trivalent dopant ions entered into CaF2 nanocrystals [75]. Hence, the defects could be responsible for effective faster depopulation of the 5D0 state when the content of Eu3+ grows, resulting in shortening of the τavg(5D0) decay times. Moreover, it is interesting to note that the τavg(5D0) lifetime was prolonged when the Tb3+:Eu3+ molar ratio was achieved 1:1, and then reduced when the molar ratio equaled to 1:1.5 and 1:2, whereas the luminescence intensity was still increased. A similar effect was also observed for Eu3+-doped silicate hybrid materials [76]. The luminescence intensity successfully grew from 0.1 mol% up to 1 mol%; however, it was reported that the lifetimes of the 5D0 state gradually reduced from 617 (for 0.1 mol% Eu3+-doped sample) to 275 μs (for 1 mol% Eu3+-doped sample). In the case of our fabricated sol-gel samples, the experimental results from luminescence decay analysis have clearly proven that the variable molar ratio of Tb3+:Eu3+ and controlled crystallization of amorphous xerogels could be responsible for modulating the character of interionic processes.
Based on recorded emission spectra and performed decay analysis of the 5D0 state of Eu3+ ions, the quantum yields, ΦEu, were calculated using Φ = kR/k equation. In this equation, k is the total decay rate constant (k = 1/τ(5D0)), whereas kR is the radiative rate constant. The value of kR was estimated from the following relation [77]:
k R = A MD , 0 n 3 I tot I MD
where Itot is the sum of integrated intensities of the 5D07FJ (J = 0–4) emission bands of Eu3+, IMD is the integrated intensity of the 5D07F1 magnetic-dipole transition, n is the refractive index of the host and AMD,0 denotes the Einstein spontaneous emission coefficient for the 5D07F1 transition and its value for sol-gel systems is equal to 14.65 s−1 [78]. Similar to previous reports for CaF2 thin films [79] and CaF2 nanoparticles produced by the fluorolytic sol-gel process [80], the refractive index of CaF2 nanocrystals was close to n = 1.44. The quantum efficiencies for xerogels were very similar and hesitated from 9.3% to 10.5%. These values were changed drastically during controlled heat-treatment when xerogels were transformed into glass-ceramic systems. The quantum efficiencies achieved the following values: 75.6% (GC1Tb0.5Eu), 76.1% (GC1Tb1Eu), 69.0% (GC1Tb1.5Eu), and 60.0% (GC1Tb2Eu). Our calculations are in good agreement with results reported by Sun et al. [81] for Eu3+-doped CaF2 thin films, for which the highest quantum efficiency was estimated to 64.24%.

4. Conclusions

In summary, the Tb3+/Eu3+ ET was systematically investigated in a series of xerogels and glass-ceramics containing CaF2 nanocrystals and the variable Tb3+:Eu3+ molar ratio. The transformation of amorphous xerogels into glass-ceramics was successfully carried out at as low a temperature as 350 °C. A particular emphasis was placed on determining the correlation between the photoluminescence properties of prepared sol-gel materials and controlled crystallization, as well as the change in the Tb3+:Eu3+ molar ratio. The following points have been established:
  • Using spectroscopy of Eu3+ ions as spectral probes, it was found that optically active dopants were preferably segregated inside the lattice of CaF2 nanocrystals during controlled heat-treatment of initial xerogels. Indeed, the 5D07F1 MD transition occupied the predominant advantage for glass-ceramics, which resulted in an almost six-fold decline in R/O ratio values from approximately 3.70 (for amorphous xerogels) to 0.64 (reported after controlled ceramization);
  • The growing R/G ratio (from 0.09 to 0.30 for xerogels, and from 0.14 to 3.76 for glass-ceramics) was observed when the Tb3+:Eu3+ molar ratio changed from 1:0.5 to 1:2. Notably, in glass-ceramics, the emission of Tb3+ ions visibly gradually weakened, while luminescence of Eu3+ ions occupied the predominant advantage, significantly enhancing the reddish-orange emission;
  • Performed decay analysis revealed an interesting dependence of decay times on change in the Tb3+:Eu3+ molar ratio, as well as partial segregation of Tb3+ and Eu3+ ions inside CaF2 nanocrystals formed during controlled heat-treatment at 350 °C. Indeed, a well-observable gradual shortening in τ(5D4) lifetimes for Tb3+ ions when the Tb3+:Eu3+ molar ratio changed from 1:0.5 to 1:2 was reported for xerogels (from 1.18 to 0.91 ms) and glass-ceramics (from 4.75 to 1.55 ms), and it was accompanied by an adequate increase in ηET (from 11.9% to 22.9% for xerogels and from 25.7% to 67.4% for glass-ceramics). Higher ηET values for the glass-ceramics resulted from a significant reduction in interionic distances between Tb3+ and Eu3+ ions inside the CaF2 crystal lattice;
  • The decay analysis of the 5D0 state (Eu3+) clearly revealed that the partial crystallization induced a remarkable prolongation of τavg(5D0) lifetimes even to 8.59 ms when the Tb3+:Eu3+ molar ratio equals 1:1, however, the further change in Tb3+:Eu3+ caused a slight shortening of decay times (7.94 when Tb3+:Eu3+ = 1:1.5, and 6.96 ms when Tb3+:Eu3+ = 1:2), which indicated a competition between radiative and non-radiative processes.

Author Contributions

Conceptualization, N.P.; methodology, N.P. and B.S.-S.; software, N.P.; validation, N.P.; formal analysis, N.P.; investigation, N.P., T.G., and J.P.; resources, W.A.P.; data curation, N.P.; writing—original draft preparation, N.P.; writing—review and editing, W.A.P.; visualization, N.P.; supervision, N.P.; project administration, N.P.; funding acquisition, W.A.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Centre (Poland), grant number 2016/23/B/ST8/01965.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fischer, S.; Pier, T.; Jüstel, T. On the sensitization of Eu3+ with Ce3+ and Tb3+ by composite structured Ca2LuHf2Al3O12 garnet phosphors for blue LED excitation. Dalton Trans. 2019, 48, 315–323. [Google Scholar] [CrossRef] [PubMed]
  2. Psuja, P.; Hreniak, D.; Strek, W. Rare-Earth Doped Nanocrystalline Phosphors for Field Emission Displays. J. Nanomater. 2007. [Google Scholar] [CrossRef]
  3. Kränkel, C.; Marzahl, D.-T.; Moglia, F.; Huber, G.; Metz, P.W. Out of the blue: Semiconductor laser pumped visible rare-earth doped lasers. Laser Photonics Rev. 2016, 10, 548–568. [Google Scholar] [CrossRef] [Green Version]
  4. Klier, D.T.; Kumke, M.U. Upconversion NaYF4:Yb:Er nanoparticles co-doped with Gd3+ and Nd3+ for thermometry on the nanoscale. RSC Adv. 2015, 5, 67149–67156. [Google Scholar] [CrossRef] [Green Version]
  5. Han, L.; Song, J.; Liu, W.; Xiao, A.; Zhang, L.; Lu, A.; Xiao, Z.; You, W.; Zhang, Q. Preparation and luminescent properties of Tm3+-Dy3+ co-doped phosphate glass for white light-emitting-diode applications. J. Lumin. 2020, 227, 117559. [Google Scholar] [CrossRef]
  6. Bian, W.; Lin, Y.; Wang, T.; Yu, X.; Qiu, J.; Zhou, M.; Luo, H.; Yu, S.F.; Xu, X. Direct Identification of Surface Defects and Their Influence on the Optical Characteristics of Upconversion Nanoparticles. ACS Nano 2018, 12, 3623–3628. [Google Scholar] [CrossRef]
  7. Boyer, J.-C.; Cuccia, L.A.; Capobianco, J.A. Synthesis of Colloidal Upconverting NaYF4:Er3+/Yb3+ and Tm3+/Yb3+ Monodisperse Nanocrystals. Nano Lett. 2007, 7, 847–852. [Google Scholar] [CrossRef]
  8. Saidi, K.; Dammak, M. Crystal structure, optical spectroscopy and energy transfer properties in NaZnPO4:Ce3+,Tb3+ phosphors for UV-based LEDs. RSC Adv. 2020, 10, 21867–21875. [Google Scholar] [CrossRef]
  9. Xu, C.; Guan, H.; Song, Y.; An, Z.; Zhang, X.; Zhou, X.; Shi, Z.; Sheng, Y.; Zou, Y.H. Novel highly efficient single-component multi-peak emitting aluminosilicate phosphors co-activated with Ce3+,Tb3+ and Eu2+: Luminescence properties, tunable color, and thermal properties. Phys. Chem. Chem. Phys. 2018, 20, 1591–1607. [Google Scholar] [CrossRef]
  10. Wang, T.; Yu, H.; Siu, C.K.; Qiu, J.; Xu, X.; Yu, S.F. White-light whispering-gallery-mode lasing from lanthanide-doped upconversion NaYF4 hexagonal microrods. ACS Photonics 2017, 4, 1539–1543. [Google Scholar] [CrossRef]
  11. Hassairi, M.A.; Dammak, M.; Zambou, D.; Chadeyron, G.; Mahiou, R. Red-green-blue upconversion luminescence and energy transfer in Yb3+/Er3+/Tm3+ doped YP5O14 ultraphosphates. J. Lumin. 2017, 181, 393–399. [Google Scholar] [CrossRef]
  12. Chu, Y.; Ren, J.; Zhang, J.; Peng, G.; Yang, J.; Wang, P.; Yuan, L. Ce3+/Yb3+/Er3+ triply doped bismuth borosilicate glass: A potential fiber material for broadband near-infrared fiber amplifiers. Sci. Rep. 2016, 6, 33865. [Google Scholar] [CrossRef]
  13. Li, B.; Huang, X.; Guo, H.; Zeng, Y. Energy transfer and tunable photoluminescence of LaBWO6:Tb3+, Eu3+ phosphors for near-UV white LEDs. Dyes Pigm. 2018, 150, 67–72. [Google Scholar] [CrossRef]
  14. Baur, F.; Glocker, F.; Jüstel, T. Photoluminescence and energy transfer rates and efficiencies in Eu3+ activated Tb2Mo3O12. J. Mater. Chem. C 2015, 3, 2054–2064. [Google Scholar] [CrossRef] [Green Version]
  15. Mi, R.; Chen, J.; Liu, Y.-G.; Fang, M.; Mei, L.; Huang, Z.; Wang, B.; Zhaob, C. Luminescence and energy transfer of a color tunable phosphor: Tb3+ and Eu3+ co-doped ScPO4. RSC Adv. 2016, 6, 28887–28894. [Google Scholar] [CrossRef]
  16. Xu, M.; Ding, Y.; Luo, W.; Wang, L.; Li, S.; Liu, Y. Synthesis, luminescence properties and energy transfer behavior of color-tunable KAlP2O7:Tb3+, Eu3+ phosphors. Opt. Laser Technol. 2020, 121, 105829. [Google Scholar] [CrossRef]
  17. Xia, Y.; Zou, X.; Zhang, H.; Zhao, M.; Chen, X.; Jia, W.; Su, C.; Shao, J. Luminescence and energy transfer studies of Eu3+-Tb3+ co-doped transparent glass ceramics containing BaMoO4 crystallites. J. Alloys Compd. 2019, 774, 540–546. [Google Scholar] [CrossRef]
  18. Yao, L.-Q.; Chen, G.-H.; Yang, T.; Cui, S.-C.; Li, Z.-C.; Yang, Y. Energy transfer, tunable emission and optical thermometry in Tb3+/Eu3+ co-doped transparent NaCaPO4 glass ceramics. Ceram. Int. 2016, 42, 13086–13090. [Google Scholar] [CrossRef]
  19. Chen, D.; Wang, Z.; Zhou, Y.; Huang, P.; Ji, Z. Tb3+/Eu3+: YF3 nanophase embedded glass ceramics: Structural characterization, tunable luminescence and temperature sensing behavior. J. Alloys Compd. 2015, 646, 339–344. [Google Scholar] [CrossRef]
  20. Kłonkowski, A.M.; Wiczk, W.; Ryl, J.; Szczodrowski, K.; Wileńska, D. A white phosphor based on oxyfluoride nano-glass-ceramics co-doped with Eu3+ and Tb3+: Energy transfer study. J. Alloys Compd. 2017, 724, 649–653. [Google Scholar] [CrossRef]
  21. Hu, F.; Zhao, Z.; Chi, F.; Wei, X.; Yin, M. Structural characterization and temperature-dependent luminescence of CaF2:Tb3+/Eu3+ glass ceramics. J. Rare Earth. 2017, 35, 536–541. [Google Scholar] [CrossRef]
  22. Li, X.; Peng, Y.; Wei, X.; Yuan, S.; Zhu, Y.; Chen, D. Energy transfer behaviors and tunable luminescence in Tb3+/Eu3+ codoped oxyfluoride glass ceramics containing cubic/hexagonal NaYF4 nanocrystals. J. Lumin. 2019, 210, 182–188. [Google Scholar] [CrossRef]
  23. Li, X.; Chen, X.; Yuan, S.; Liu, S.; Wang, C.; Chen, D. Eu3+-Doped glass ceramics containing NaTbF4 nanocrystals: Controllable glass crystallization, Tb3+-bridged energy transfer and tunable luminescence. J. Mater. Chem. C 2017, 3, 10201–10210. [Google Scholar] [CrossRef]
  24. Xu, F.; Wei, X.; Jiang, S.; Huang, S.; Qin, Y.; Chen, Y.; Duan, C.-K.; Yin, M. Fabrication and Luminescence Properties of Transparent Glass-Ceramics Containing Eu3+-Doped TbPO4 Nanocrystals. J. Am. Ceram. Soc. 2015, 98, 464–468. [Google Scholar] [CrossRef]
  25. Wang, R.; Zhou, D.; Qiu, J.; Yang, Y.; Wang, C. Color-tunable luminescence in Eu3+/Tb3+ co-doped oxyfluoride glass and transparent glass-ceramics. J. Alloys Compd. 2015, 629, 310–314. [Google Scholar] [CrossRef]
  26. Koseva, I.; Tzvetkov, P.; Ivanov, P.; Yordanova, A.; Nikolov, V. Terbium and europium co-doped NaAlSiO4 nano glass-ceramics for LED application. Optik 2017, 137, 45–50. [Google Scholar] [CrossRef]
  27. Gorni, G.; Velázquez, J.J.; Mosa, J.; Balda, R.; Fernández, J.; Durán, A.; Castro, Y. Transparent glass-ceramics produced by sol-gel: A suitable alternative for photonic materials. Materials 2018, 11, 212. [Google Scholar] [CrossRef] [Green Version]
  28. 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]
  29. Gorni, G.; Velázquez, J.J.; Mosa, J.; Mather, G.C.; Serrano, A.; Vila, M.; Castro, G.R.; Bravo, D.; Balda, R.; Fernández, J.; et al. Transparent sol-gel oxyfluoride glass-ceramics with high crystalline fraction and study of RE incorporation. Nanomaterials 2019, 9, 530. [Google Scholar] [CrossRef] [Green Version]
  30. Danks, A.E.; Hall, S.R.; Schnepp, Z. The evolution of ‘sol-gel’ chemistry as a technique for materials synthesis. Mater. Horiz. 2016, 3, 91–112. [Google Scholar] [CrossRef] [Green Version]
  31. Carta, D.; Pickup, D.M.; Knowles, J.C.; Ahmed, I.; Smith, M.E.; Newport, R.J. A structural study of sol–gel and melt-quenched phosphate-based glasses. J. Non Cryst. Solids 2007, 353, 1759–1765. [Google Scholar] [CrossRef]
  32. Wang, H.; Ye, S.; Liu, T.; Li, S.; Hu, R.; Wang, D. Influence of local phonon energy on quantum efficiency of Tb3+-Yb3+ co-doped glass ceramics containing fluoride nanocrystals. J. Rare Earth. 2014, 33, 201–205. [Google Scholar] [CrossRef]
  33. Deng, W.; Cheng, J.-S. New transparent glass-ceramics containing large grain Eu3+: CaF2 nanocrystals. Mater. Lett. 2012, 73, 112–114. [Google Scholar] [CrossRef]
  34. Ryskin, A.I.; Fedorov, P.P.; Bagraev, N.T.; Lushchik, A.; Vasil’chenko, E.; Angervaks, A.E.; Kudryavtseva, I. Stabilization of high-temperature-disorder of fluorine sublattice by quenching in calcium fluoride crystals. J. Fluor. Chem. 2017, 200, 109–114. [Google Scholar] [CrossRef]
  35. Liu, X.; Li, Y.; Aidilibike, T.; Guo, J.; Di, W.; Qin, W. Pure red upconversion emission from CaF2:Yb3+/Eu3+. J. Lumin. 2017, 185, 247–250. [Google Scholar] [CrossRef]
  36. Kuznetsov, S.V.; Morozov, O.A.; Gorieva, V.G.; Mayakova, M.N.; Marisov, M.A.; Voronov, V.V.; Yapryntsev, A.D.; Ivanov, V.K.; Nizamutdinov, A.S.; Semashko, V.V.; et al. Synthesis and luminescence studies of CaF2:Yb:Pr solid solutions powders for photonics. J. Fluor. Chem. 2018, 211, 70–75. [Google Scholar] [CrossRef]
  37. Adusumalli, V.N.K.B.; Koppisetti, H.V.S.R.M.; Mahalingam, V. Ce3+ sensitized bright white light emission from colloidal Ln3+ doped CaF2 nanocrystals for the development of transparent nanocomposites. J. Mater. Chem. C 2016, 4, 2289–2294. [Google Scholar] [CrossRef]
  38. Akchurin, M.S.; Basiev, T.T.; Demidenko, A.A.; Doroshenko, M.E.; Fedorov, P.P.; Garibin, E.A.; Gusev, P.E.; Kuznetsov, S.V.; Krutov, M.A.; Mironov, I.A.; et al. CaF2:Yb laser ceramics. Opt. Mater. 2013, 35, 444–450. [Google Scholar] [CrossRef]
  39. Liu, G.; Sun, Z.; Fu, Z.; Ma, L.; Wang, X. Temperature sensing and bio-imaging applications based on polyethylene/CaF2 nanoparticles with upconversion fluorescence. Talanta 2017, 169, 181–188. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Cheng, Y.; Wang, Y.; Teng, F.; Dong, H.; Chen, L.; Mu, J.; Sun, Q.; Fan, J.; Hu, X.; Miao, H. Down-conversion emission of Ce3+-Tb3+ co-doped CaF2 hollow spheres and application for solar cells. Mater. Res. Express 2018, 5, 036206. [Google Scholar] [CrossRef]
  41. Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarski, W.A. Sol-Gel Glass-Ceramic Materials Containing CaF2:Eu3+ Fluoride Nanocrystals for Reddish-Orange Photoluminescence Applications. Appl. Sci. 2019, 9, 5490. [Google Scholar] [CrossRef] [Green Version]
  42. Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarski, W.A. Photoluminescence investigation of sol-gel glass-ceramic materials containing SrF2:Eu3+ nanocrystals. J. Alloys Compd. 2019, 810, 151935. [Google Scholar] [CrossRef]
  43. Pawlik, N.; Szpikowska-Sroka, B.; Pisarska, J.; Goryczka, T.; Pisarski, W.A. Reddish-Orange Luminescence from BaF2:Eu3+ Fluoride Nanocrystals Dispersed in Sol-Gel Materials. Materials 2019, 12, 3735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Pawlik, N.; Szpikowska-Sroka, B.; Pisarski, W.A. Energy Transfer Study on Tb3+/Eu3+ Co-Activated Sol-Gel Glass-Ceramic Materials Containing MF3 (M = Y, La) Nanocrystals for NUV Optoelectronic Devices. Materials 2020, 13, 2522. [Google Scholar] [CrossRef]
  45. Yanes, A.C.; del-Castillo, J.; Ortiz, E. Energy transfer and tunable emission in BaGdF5:RE3+ (RE = Ce, Tb, Eu) nano-glass-ceramics. J. Alloys Compd. 2019, 773, 1099–1107. [Google Scholar] [CrossRef]
  46. Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarski, W.A. Spectroscopic Properties of Eu3+ Ions in Sol-Gel Materials Containing Calcium Fluoride Nanocrystals. Phys. Status Solidi B 2019, 257, 1900478. [Google Scholar] [CrossRef]
  47. Pawlik, N.; Szpikowska-Sroka, B.; Pietrasik, E.; Goryczka, T.; Pisarski, W.A. Structural and luminescence properties of silica powders and transparent glass-ceramics containing LaF3:Eu3+ nanocrystals. J. Am. Ceram. Soc. 2018, 101, 4654–4668. [Google Scholar] [CrossRef]
  48. Holder, C.F.; Schaak, R.E. Tutorial on Powder X-ray Diffraction for Characterizing Nanoscale Materials. ACS Nano 2019, 13, 7359–7365. [Google Scholar] [CrossRef] [Green Version]
  49. Zak, A.K.; Majid, W.H.A.; Abrishami, M.E.; Yousefi, R. X-ray analysis of ZnO nanoparticles by Williamson-Hall and size-strain plot methods. Solid State Sci. 2011, 13, 251–256. [Google Scholar] [CrossRef]
  50. Withers, S.H.; Peale, R.E.; Schulte, A.F.; Braunstein, G.; Beck, K.M.; Hess, W.P.; Reeder, R.J. Broad distribution of crystal-field environments for Nd3+ in calcite. Phys. Chem. Miner. 2003, 30, 440–448. [Google Scholar] [CrossRef]
  51. Qin, D.; Tang, W. Energy transfer and multicolor emission in single-phase Na5Ln(WO4)4-z(MoO4)z:Tb3+,Eu3+ (Ln = La, Y, Gd) phosphors. RSC Adv. 2016, 6, 45376–45385. [Google Scholar] [CrossRef]
  52. Pan, Y.; Wang, W.; Zhou, L.; Xu, H.; Xia, Q.; Liu, L.; Liu, X.; Li, L. F--Eu3+ charge transfer energy and local crystal environment in Eu3+ doped calcium fluoride. Ceram. Int. 2017, 43, 13089–13093. [Google Scholar] [CrossRef]
  53. Innocenzi, P. Infrared spectroscopy of sol-gel derived silica-based films: A spectra-microstructure overview. J. Non Cryst. Solids 2003, 316, 309–319. [Google Scholar] [CrossRef]
  54. Osswald, J.; Fehr, K.T. FTIR spectroscopic study on liquid silica solutions and nanoscale particle size determination. J. Mater. Sci. 2006, 41, 1335–1339. [Google Scholar] [CrossRef]
  55. Kreienborg, N.M.; Merten, C. How to treat C-F stretching vibrations? A vibrational CD study on chiral fluorinated molecules. Phys. Chem. Chem. Phys. 2019, 21, 3506–3511. [Google Scholar] [CrossRef]
  56. Aguiar, H.; Serra, J.; González, P.; León, B. Structural study of sol-gel silicate glasses by IR and Raman spectroscopies. J. Non Cryst. Solids 2009, 355, 475–480. [Google Scholar] [CrossRef]
  57. Song, C.; Dong, X. Preparation and Characterization of Tetracomponent ZnO/SiO2/SnO2/TiO2 Composite Nanofibers by Electrospinning. Adv. Chem. Eng. 2012, 2, 108–112. [Google Scholar] [CrossRef] [Green Version]
  58. Li, G.; Zhu, T.; Deng, Z.; Zhang, Y.; Jiao, F.; Zheng, H. Preparation of Cu-SiO2 composite aerogel by ambient drying and the influence of synthesizing conditions on the structure of the aerogel. Sci. Bull. 2011, 56, 685–690. [Google Scholar] [CrossRef] [Green Version]
  59. Gökçe, M.; Şentürk, U.; Uslu, D.K.; Burgaz, G.; Şahin, Y.; Gökçe, A.G. Investigation of europium concentration dependence on the luminescent properties of borogermanate glasses. J. Lumin. 2017, 192, 263–268. [Google Scholar] [CrossRef]
  60. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef] [Green Version]
  61. Pisarski, W.A.; Pisarska, J.; Zur, L.; Goryczka, T. Structural and optical aspects for Eu3+ and Dy3+ ions in heavy metal glasses based on PbO-Ga2O3-XO2 (X = Te, Ge, Si). Opt. Mater. 2013, 35, 1051–1056. [Google Scholar] [CrossRef]
  62. Hurenkamp, J.H.; de Jong, J.J.D.; Browne, W.R.; van Esch, J.H.; Feringa, B.L. Tuning energy transfer in switchable donor-acceptor systems. Org. Biomol. Chem. 2008, 6, 1268–1277. [Google Scholar] [CrossRef] [Green Version]
  63. 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]
  64. Tomina, V.V.; Stolyarchuk, N.V.; Katelnikovas, A.; Misevicius, M.; Kanuchova, M.; Kareiva, A.; Beganskienė, A.; Melnyk, I.V. Preparation and luminescence properties of europium(III)-loaded aminosilica spherical particles. Coll. Surf. A 2021, 608, 125552. [Google Scholar] [CrossRef]
  65. Żur, L.; Pisarska, J.; Pisarski, W.A. Terbium -doped heavy metal glasses for green luminescence. J. Rare Earth. 2011, 29, 1198–1200. [Google Scholar] [CrossRef]
  66. Bredol, M.; Gutzov, S. Effect of germanium-codoping on the luminescence of terbium doped silica xerogel. Opt. Mater. 2002, 20, 233–239. [Google Scholar] [CrossRef]
  67. Szpikowska-Sroka, B.; Pawlik, N.; Goryczka, T.; Bańczyk, M.; Pisarski, W.A. Influence of activator concentration on green-emitting Tb3+-doped materials derived by sol-gel method. J. Lumin. 2017, 188, 400–408. [Google Scholar] [CrossRef]
  68. Parchur, A.K.; Prasad, A.I.; Ansari, A.A.; Rai, S.B.; Ningthoujam, R.S. Luminescence properties of Tb3+-doped CaMoO4 nanoparticles: Annealing effect, polar medium dispersible, polymer film and core-shell formation. Dalton Trans. 2012, 41, 11032–11045. [Google Scholar] [CrossRef] [PubMed]
  69. Secu, C.E.; Bartha, C.; Polosan, S.; Secu, M. Thermally activated conversion of a silicate gel to an oxyfluoride glass ceramic: Optical study using Eu3+ probe ion. J. Lumin. 2014, 146, 539–543. [Google Scholar] [CrossRef]
  70. Van Dijk, J.M.F.; Schuurmans, M.F.H. On the nonradiative and radiative decay rates and a modified exponential energy gap law for 4f-4f transitions in rare-earth ions. J. Chem. Phys. 1983, 78, 5317–5323. [Google Scholar] [CrossRef]
  71. Xia, Z.; Zhuang, J.; Liao, L. Novel red-emitting Ba2Tb(BO3)2Cl:Eu phosphor with efficient energy transfer for potential application in white light-emitting diodes. Inorg. Chem. 2012, 51, 7202–7209. [Google Scholar] [CrossRef]
  72. Gopi, S.; Jose, S.K.; Sreeja, E.; Manasa, P.; Unnikrishnan, N.V.; Cyriac, J.; Biju, P.R. Tunable green to red emission via Tb sensitized energy transfer in Tb/Eu co-doped alkali fluoroborate glass. J. Lumin. 2017, 192, 1288–1294. [Google Scholar] [CrossRef]
  73. Podhorodecki, A.; Banski, M.; Misiewicz, J.; Afzaal, M.; O’Brien, P.; Chad, D.; Wang, X. Multicolor light emitters based on energy exchange between Tb and Eu ions co-doped into ultrasmall β-NaYF4 nanocrystals. J. Mater. Chem. 2012, 22, 5356–5361. [Google Scholar] [CrossRef]
  74. Qin, X.; Shen, L.; Liang, L.; Han, S.; Yi, Z.; Liu, X. Suppression of Defect-Induced Quenching via Chemical Potential Tuning: A Theoretical Solution for Enhancing Lanthanide Luminescence. J. Phys. Chem. C 2019, 123, 11151–11161. [Google Scholar] [CrossRef]
  75. Nicoara, I.; Munteanu, M.; Preda, E.; Stef, M. Some dielectric and optical properties of ErF3-doped CaF2 crystals. J. Cryst. Growth 2008, 310, 2020–2025. [Google Scholar] [CrossRef]
  76. Trejo-García, P.M.; Palomino-Merino, R.; De la Cruz, J.; Espinosa, J.E.; Aceves, R.; Moreno-Barbosa, E.; Moreno, O.P. Luminescent Properties of Eu3+-Doped Hybrid SiO2-PMMA Material for Photonic Applications. Micromachines 2018, 9, 441. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Chen, J.; Meng, Q.; May, P.S.; Berry, M.T.; Lin, C. Sensitization of Eu3+ luminescence in Eu:YPO4 nanocrystals. J. Phys. Chem. C 2013, 117, 5953–5962. [Google Scholar] [CrossRef]
  78. Werts, M.H.V.; Jukes, R.T.F.; Verhoeven, J.W. The emission spectrum and the radiative lifetime of Eu3+ in luminescent lanthanide complexes. Phys. Chem. Chem. Phys. 2002, 4, 1542–1548. [Google Scholar] [CrossRef]
  79. Çetin, N.E.; Korkmaz, Ş.; Elmas, S.; Ekem, N.; Pat, S.; Balbağ, M.Z.; Tarhan, E.; Temel, S.; Özmumcu, M. The structural, optical and morphological properties of CaF2 thin films by using Thermionic Vacuum Arc (TVA). Mater. Lett. 2013, 91, 175–178. [Google Scholar] [CrossRef] [Green Version]
  80. Rehmer, A.; Scheurell, K.; Kemnitz, E. Formation of nanoscopic CaF2 via a fluorolytic sol- gel process for antireflective coatings. J. Mater. Chem. C 2015, 3, 1716–1723. [Google Scholar] [CrossRef] [Green Version]
  81. Sun, J.; Wang, H.; Zhang, Y.; Zheng, Y.; Xu, Z.; Liu, R. Structure and luminescent properties of electrodeposited Eu3+-doped CaF2 thin films. Thin Solid Film 2014, 562, 478–484. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of XG1Tb1Eu and GC1Tb1Eu co-doped sol-gel materials. Inset shows the TEM image of glass-ceramic fabricated at 350 °C.
Figure 1. XRD patterns of XG1Tb1Eu and GC1Tb1Eu co-doped sol-gel materials. Inset shows the TEM image of glass-ceramic fabricated at 350 °C.
Materials 14 00754 g001
Figure 2. IR-ATR spectra recorded for xerogels and glass-ceramics. The deconvolution of bands characteristic for silicate host (1275–875 cm−1) as well as OH moieties and C–H bonds (3750–2500 cm−1) was also presented.
Figure 2. IR-ATR spectra recorded for xerogels and glass-ceramics. The deconvolution of bands characteristic for silicate host (1275–875 cm−1) as well as OH moieties and C–H bonds (3750–2500 cm−1) was also presented.
Materials 14 00754 g002
Figure 3. Graphical representation of structural transformation during controlled ceramization at 350 °C.
Figure 3. Graphical representation of structural transformation during controlled ceramization at 350 °C.
Materials 14 00754 g003
Figure 4. The emission spectra of XG1TbxEu xerogels co-doped with Tb3+/Eu3+ ions recorded at λexc = 395 nm.
Figure 4. The emission spectra of XG1TbxEu xerogels co-doped with Tb3+/Eu3+ ions recorded at λexc = 395 nm.
Materials 14 00754 g004
Figure 5. The emission spectra of GC1TbxEu glass-ceramics co-doped with Tb3+/Eu3+ ions recorded at λexc = 394 nm.
Figure 5. The emission spectra of GC1TbxEu glass-ceramics co-doped with Tb3+/Eu3+ ions recorded at λexc = 394 nm.
Materials 14 00754 g005
Figure 6. Excitation spectra recorded for red (λem = 612) and green (λem = 542 nm) emission lines for XG1Tb1Eu xerogel.
Figure 6. Excitation spectra recorded for red (λem = 612) and green (λem = 542 nm) emission lines for XG1Tb1Eu xerogel.
Materials 14 00754 g006
Figure 7. Emission spectra of: (a) XG1Tb1Euexc = 395), (b) XG1Tbexc = 351), and (c) XG1Tb1Euexc = 351 nm).
Figure 7. Emission spectra of: (a) XG1Tb1Euexc = 395), (b) XG1Tbexc = 351), and (c) XG1Tb1Euexc = 351 nm).
Materials 14 00754 g007
Figure 8. The emission spectra of XG1Tb and XG1TbxEu xerogels recorded under excitation at λexc = 351 nm.
Figure 8. The emission spectra of XG1Tb and XG1TbxEu xerogels recorded under excitation at λexc = 351 nm.
Materials 14 00754 g008
Figure 9. Emission spectra of: (a) GC1Tb1Euexc = 395 nm), (b) GC1Tbexc = 351 nm), and (c) GC1Tb1Euexc = 351 nm).
Figure 9. Emission spectra of: (a) GC1Tb1Euexc = 395 nm), (b) GC1Tbexc = 351 nm), and (c) GC1Tb1Euexc = 351 nm).
Materials 14 00754 g009
Figure 10. The emission spectra of GC1Tb and GC1TbxEu xerogels recorded under excitation at λexc = 351 nm.
Figure 10. The emission spectra of GC1Tb and GC1TbxEu xerogels recorded under excitation at λexc = 351 nm.
Materials 14 00754 g010
Figure 11. Luminescence decay curves of the 5D4 level of Tb3+ ions recorded for (a) XG1Tb, and (b) GC1Tb.
Figure 11. Luminescence decay curves of the 5D4 level of Tb3+ ions recorded for (a) XG1Tb, and (b) GC1Tb.
Materials 14 00754 g011
Figure 12. Luminescence decay curves of the 5D4 (Tb3+) level recorded for individual XG1TbxEu and GC1TbxEu co-doped samples: (a) Xerogels, (b) glass-ceramics. The curves were recorded upon excitation at λexc = 351 nm.
Figure 12. Luminescence decay curves of the 5D4 (Tb3+) level recorded for individual XG1TbxEu and GC1TbxEu co-doped samples: (a) Xerogels, (b) glass-ceramics. The curves were recorded upon excitation at λexc = 351 nm.
Materials 14 00754 g012
Figure 13. The relation between R/G ratio, energy transfer efficiency (ηET), and lifetime of the 5D4 (Tb3+) state for: xerogels (a) and glass-ceramics (b).
Figure 13. The relation between R/G ratio, energy transfer efficiency (ηET), and lifetime of the 5D4 (Tb3+) state for: xerogels (a) and glass-ceramics (b).
Materials 14 00754 g013
Figure 14. Luminescence decay curves of the 5D0 state of Eu3+ ions recorded for: (a) xerogels (λexc = 395) and (b) glass-ceramics (λexc = 394 nm).
Figure 14. Luminescence decay curves of the 5D0 state of Eu3+ ions recorded for: (a) xerogels (λexc = 395) and (b) glass-ceramics (λexc = 394 nm).
Materials 14 00754 g014
Table 1. Measured lifetimes of the 5D4 state (Tb3+), energy transfer efficiencies, and R/G ratio values for prepared xerogels.
Table 1. Measured lifetimes of the 5D4 state (Tb3+), energy transfer efficiencies, and R/G ratio values for prepared xerogels.
Xerogelτ(5D4) (ms)ηET (%)R/G
XG1Tb1.18-0.09
XG1Tb0.5Eu1.0411.90.15
XG1Tb1Eu1.0114.40.17
XG1Tb1.5Eu0.9618.60.26
XG1Tb2Eu0.9122.90.30
Table 2. Measured lifetimes of the 5D4 state (Tb3+), average decay times, energy transfer efficiencies, and R/G ratio values for prepared glass-ceramics.
Table 2. Measured lifetimes of the 5D4 state (Tb3+), average decay times, energy transfer efficiencies, and R/G ratio values for prepared glass-ceramics.
Glass-Ceramicτm(5D4) (ms)τavg(5D4) (ms)ηET (%)R/G
GC1Tb1.40 (τ1)
5.67 (τ2)
4.75-0.14
GC1Tb0.5Eu1.32 (τ1)
4.74 (τ2)
3.7525.70.80
GC1Tb1Eu0.62 (τ1)
2.99 (τ2)
2.5945.51.60
GC1Tb1.5Eu0.37 (τ1)
2.08 (τ2)
1.9259.62.47
GC1Tb2Eu0.20 (τ1)
1.66 (τ2)
1.5567.43.76
Table 3. Measured lifetimes of the 5D0 state (Eu3+) in xerogels (λexc = 395 nm excitation).
Table 3. Measured lifetimes of the 5D0 state (Eu3+) in xerogels (λexc = 395 nm excitation).
Xerogelτ(5D0) (ms)
XG1Tb0.5Eu0.37
XG1Tb1Eu0.43
XG1Tb1.5Eu0.44
XG1Tb2Eu0.45
Table 4. Measured lifetimes and calculated average decay times of the 5D0 state (Eu3+) in fabricated glass-ceramics (λexc = 394 nm excitation).
Table 4. Measured lifetimes and calculated average decay times of the 5D0 state (Eu3+) in fabricated glass-ceramics (λexc = 394 nm excitation).
Glass-Ceramicλexc = 394 nm
τm(5D0) (ms)τavg(5D0) (ms)
GC1Tb0.5Eu0.98 (τ1)
9.04 (τ2)
8.40
GC1Tb1Eu1.33 (τ1)
9.52 (τ2)
8.59
GC1Tb1.5Eu1.34 (τ1)
8.79 (τ2)
7.94
GC1Tb2Eu1.35 (τ1)
7.76 (τ2)
6.96
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarska, J.; Pisarski, W.A. Structural and Photoluminescence Investigations of Tb3+/Eu3+ Co-Doped Silicate Sol-Gel Glass-Ceramics Containing CaF2 Nanocrystals. Materials 2021, 14, 754. https://doi.org/10.3390/ma14040754

AMA Style

Pawlik N, Szpikowska-Sroka B, Goryczka T, Pisarska J, Pisarski WA. Structural and Photoluminescence Investigations of Tb3+/Eu3+ Co-Doped Silicate Sol-Gel Glass-Ceramics Containing CaF2 Nanocrystals. Materials. 2021; 14(4):754. https://doi.org/10.3390/ma14040754

Chicago/Turabian Style

Pawlik, Natalia, Barbara Szpikowska-Sroka, Tomasz Goryczka, Joanna Pisarska, and Wojciech A. Pisarski. 2021. "Structural and Photoluminescence Investigations of Tb3+/Eu3+ Co-Doped Silicate Sol-Gel Glass-Ceramics Containing CaF2 Nanocrystals" Materials 14, no. 4: 754. https://doi.org/10.3390/ma14040754

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