Next Article in Journal
Nanocarriers in Veterinary Medicine: A Challenge for Improving Osteosarcoma Conventional Treatments
Next Article in Special Issue
Effect of Sonication and Ceria Doping on Nanoparticles Fabricated by Laser Marker Ablation of Ti in Water
Previous Article in Journal
Valorization of Vegetable Waste from Leek, Lettuce, and Artichoke to Produce Highly Concentrated Lignocellulose Micro- and Nanofibril Suspensions
Previous Article in Special Issue
Influence of Tartrate Ligand Coordination over Luminescence Properties of Chiral Lanthanide-Based Metal–Organic Frameworks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Photoluminescence Investigations of Dy3+-Doped Silicate Xerogels and SiO2-LaF3 Nano-Glass-Ceramic Materials

1
Institute of Chemistry, University of Silesia, 40-007 Katowice, Poland
2
Institute of Materials Engineering, University of Silesia, 41-500 Chorzów, Poland
*
Authors to whom correspondence should be addressed.
Nanomaterials 2022, 12(24), 4500; https://doi.org/10.3390/nano12244500
Submission received: 14 November 2022 / Revised: 14 December 2022 / Accepted: 15 December 2022 / Published: 19 December 2022

Abstract

:
In this work, the series of Dy3+-doped silicate xerogels were synthesized by sol-gel technique and further processed at 350 °C into SiO2-LaF3:Dy3+ nano-glass-ceramic materials. The X-ray diffraction (XRD) measurements, along with the thermal analysis, indicated that heat-treatment triggered the decomposition of La(TFA)3 inside amorphous sol-gel hosts, resulting in the formation of hexagonal LaF3 phase with average crystal size at about ~10 nm. Based on the photoluminescence results, it was proven that the intensities of blue (4F9/26H15/2), yellow (4F9/26H13/2), and red (4F9/26H11/2) emissions, as well as the calculated yellow-to-blue (Y/B) ratios, are dependent on the nature of fabricated materials, and from fixed La3+:Dy3+ molar ratios. For xerogels, the emission was gradually increased, and the τ(4F9/2) lifetimes were elongated to 42.7 ± 0.3 μs (La3+:Dy3+ = 0.82:0.18), however, for the sample with the lowest La3+:Dy3+ molar ratio (0.70:0.30), the concentration quenching was observed. For SiO2-LaF3:Dy3+ nano-glass-ceramics, the concentration quenching effect was more visible than for xerogels and started from the sample with the highest La3+:Dy3+ molar ratio (0.988:0.012), thus the τ(4F9/2) lifetimes became shorter from 1731.5 ± 5.7 up to 119.8 ± 0.4 μs. The optical results suggest, along with an interpretation of XRD data, that Dy3+ ions were partially entered inside LaF3 phase, resulting in the shortening of Dy3+-Dy3+ inter-ionic distances.

1. Introduction

For the past few decades, optical materials doped with luminescent rare earths (RE3+) have attracted immense attention because of their plenteous application in photonic devices, like displays, lasers, light-emitting diodes (LEDs), and sensors [1,2,3,4]. Among RE3+, the visible luminescence of Dy3+ ions inside blue (488 nm, 4F9/26H15/2 transition), and yellow (570 nm, 4F9/26H13/2 hypersensitive transition) regions makes Dy3+-doped optical materials promising candidates for utilization as white light emitters. For instance, barium silicate glasses doped with Dy3+ are suitable for a white light generation defined by chromaticity coordinates equal to (0.31|0.34), which are lying near the standard point for the while illuminant (0.33|0.33) [5]. Similarly, the chromaticity coordinates for Dy3+-doped lithium zinc borosilicate glasses (e.g., (0.318|0.357) or (0.321|0.347)) were also found to be located inside the white light region, and the calculated correlated color temperatures (CCT) are above 5700 K, which indicates that the glasses emit cool white light [6]. Further, it was reported that the lithium aluminum borate glasses co-doped with Gd3+/Dy3+ ions are able to produce neutral white light (0.363|0.402) with CCT equal to 4556 K [7], but the warm white light was obtained for selected Dy3+-doped glass-ceramics containing Na3Gd(PO4)2 phase [8]. Moreover, since the populations of the 4I15/2 and the 4F9/2 excited levels of Dy3+ ions are governed by the Boltzmann statistics, they are thermally coupled, which makes it possible to apply them in optical thermometry. Factually, Bu et al. [9] found for Dy3+-doped glass-ceramics containing LaF3 crystal phase that the intensities of emissions located at 480 nm (4F9/26H15/2) and 572 nm (4F9/26H13/2) are decreased, but conversely, the luminescence within the blue light scope at 454 nm (4I15/26H15/2) gradually increased as temperature rose from 275 to 550 K. Similarly, Komar et al. [10] stated that the fluorescence intensity ratio between the 4I15/26H15/2 and the 4F9/26H15/2 emission lines for La3Ga5.5Ta0.5O14 crystal could be treated as a temperature-dependent parameter up to 773 K. Further, the materials co-doped with Dy3+ and transition metals could also be used in the field of luminescence thermometry. Indeed, Lin et al. [11] identified that the 2Eg4A2g transition of Mn4+ is very sensitive to temperature fluctuations because of the thermal quenching of the luminescence as a result of strong electron–phonon coupling. Contrary, the authors found that the shielded 4f9-4f9 emissions of Dy3+ ions are relatively negligibly affected by the lattice environment and thus, the ratio between emission intensities of Mn4+ and Dy3+ ions is dependent on temperature.
The spectroscopy of trivalent Dy3+ ions is widely described for various types of glassy hosts fabricated by the conventional melt-quenching technique and the derivative glass-ceramics [12,13,14,15,16], as well as phosphors like Li+-doped CaWO4:Dy3+ [17], BaSrY4O8:Dy3+ [18], or SrLaAlO4:Dy3+ [19], and complexes [20,21,22]. However, the literature concerning optical properties of Dy3+-doped sol-gel materials is far less exhausting. In this field, the studies published by the research group of B. Grobelna focused on Dy3+-doped silicate xerogels containing selected types of tungstates, e.g., CaWO4 [23], and Ln2(WO4)3 (Ln = La or Gd) [24,25]. The characteristic blue (4F9/26H15/2) and yellow (4F9/26H13/2) luminescence of Dy3+ ions were generated via the energy transfer from WO42- using an excitation wavelength from the mid-UV area (λex = 240 nm). Thus, the authors stated that fabricated optical materials could be potentially used in solar cells to enhance conversion efficiency. Additionally, the spectroscopy of Dy3+ ions in sol-gel materials was described for Dy3+/Tb3+ co-doped 90SiO2-10YF3 (mol%) [26] and 95SiO2-5LaF3 (mol%) [27] nano-glass-ceramics, and according to the energy transfer from Dy3+ to Tb3+ those systems are considered as promising candidates for solar cells applications. The synthesis of zirconate xerogels and aerogels containing Dy3+ ions were presented in work [28], and their characterization provided the thermal and structural analysis; however, the luminescence measurements of Dy3+ luminescence were not the aim of the presented studies. Therefore, such a relatively small number of papers devoted to Dy3+ ions spectroscopy in sol-gel materials makes investigating those types of optical materials highly meaningful and necessary.
Among various types of optical materials based on fluorides (e.g., CaF2, SrF2, BaF2, YF3), LaF3 is one of the most frequently and willingly studied host, as evidenced by plentiful works reported in the current literature [29,30,31,32,33,34,35]. LaF3 is characterized by exceptionally low phonon energy (~350 cm−1) arising from the strong ionicity of La3+-F bond, compared with other RE3+-F or Na+-F [36,37]. LaF3 crystallizes as the trigonal/hexagonal phase, but the cubic polymorphic form is also known and reported [38,39]. Moreover, LaF3 is also characterized by good transmission within the range between 0.13 and 11 μm [40,41]. The similarity in ionic radii of La3+ cation and other RE3+ dopants allows for relatively easily substituting them in the parent fluoride crystal lattice to improve the optical properties by suppressing the non-radiative losses of photon energy [42]. Thus, these peculiarities clearly point out the great utility potential of glass-ceramic materials containing LaF3 crystals doped with optically active RE3+ ions. Moreover, in this regard, it should be noted that the fabrication of oxyfluoride glass-ceramic materials using the sol-gel method allows for overcoming the fundamental drawback of the melt-quenching technique, which is often correlated with the high risk of the evaporation of fluorides (even about 30–40 mol%) [43]. The appropriate chemical reactions (i.e., hydrolysis of the metal/semi-metal alkoxide, further condensation, and polycondensation) during sol-gel synthesis are usually performed at room temperature (or slightly elevated); it is used as an alternative route for the fabrication of the oxyfluoride glass-ceramic materials. As a result, this low temperature approach allows for the fabrication of nano-glass–ceramics with a greater fluoride crystal fraction.
In this study, the series of silicate xerogels doped with Dy3+ ions were fabricated using the sol-gel technique, and further processed into SiO2-LaF3:Dy3+ nano-glass-ceramics. The molar ratio of La(CH3COO)3:Dy(CH3COO)3 acetates used during the sol-gel synthesis was changed as follows (1 − x):x, where x = 0.012, 0.03, 0.06, 0.12, 0.18, and 0.3. The thermal analysis and XRD technique were used to verify the structural transformation during performed controlled heat-treatment of precursor silicate xerogels. The impact of La3+:Dy3+ molar ratio, as well as the influence of xerogels’ evolution into nano-glass-ceramics on photoluminescence properties was discussed based on excitation and emission spectra, along with the decay analysis from the 4F9/2 excited state of Dy3+ ions.

2. Materials and Methods

The sol-gel preparation method used to synthesize the series of xerogels doped with Dy3+ was described with details elsewhere [44]. All reagents were taken from Sigma Aldrich Chemical Co. (St. Louis, MO, USA). The subsequent chemical reactions, which undergo during sol-gel evolution, e.g., hydrolysis, condensation, and polycondensation of precursor (tetraethoxysilane, TEOS), were carried out in a solution of ethyl alcohol (EtOH), deionized water, and acetic acid (AcOH), with molar ratio equals to 1:4:10:0.5. In parallel, the appropriate amounts of La(AcO)3 and Dy(AcO)3 acetates were dissolved in water and trifluoroacetic acid (TFA), and the resultant mixtures were added dropwise to TEOS-based solutions. The molar ratio of TFA:Ln3+ (La3+ and Dy3+) was set at 5:1, which finally varied as follows: TFA:La3+:Dy3+ = 5:(1 − x):x, where x = 0.012, 0.03, 0.06, 0.12, 0.18, and 0.3. The as-prepared sols were poured into beakers that were kept sealed until rigid xerogels were formed. The sol-gel evolution from the silicate sols, through wet-gels, up to solid xerogels was performed at 35 °C for the next several weeks, and the following samples were denoted as XG1-XG6. The transformation of xerogels into oxyfluoride nano-glass-ceramics was conducted at 350 °C for 10 h. The fabricated SiO2-LaF3:Dy3+ nano-glass-ceramics were marked in the text as GC1-GC6.
The thermogravimetry and differential scanning calorimetry (TG/DSC) were carried out using a Labsys Evo system with a heating rate of 10 °C/min in argon atmosphere (SETARAM Instrumentation, Caluire, France). The prepared sol-gel materials were characterized by X-ray diffraction (XRD) analysis using an X’Pert Pro diffractometer supplied by PANalytical with CuKα radiation with λ = 1.54056 Å wavelength (Almelo, the Netherlands). The luminescence measurements were performed on a Photon Technology International (PTI) Quanta-Master 40 (QM40) UV/VIS Steady State Spectrofluorometer (Photon Technology International, Birmingham, NJ, USA), supplied with a tunable pulsed optical parametric oscillator (OPO) pumped by the third harmonic of a Nd:YAG laser (Opotek Opolette 355 LD, OPOTEK, Carlsband, CA, USA). The laser system was coupled with a xenon lamp, a double 200 mm monochromator, and a multimode UV/VIS PMT detector. The excitation and emission spectra were recorded with a resolution of 0.5 nm. The luminescence decay curves were recorded by a PTI ASOC-10 (USB-2500) oscilloscope with ±0.1 μs accuracy. All structural and optical measurements were carried out at room temperature.

3. Results and Discussion

3.1. Analysis of TG/DTG and DSC Results for Dy3+-Doped Xerogels

Figure 1 presents the TG/DSC curves recorded for fabricated xerogels in an inert gas atmosphere in a temperature range from 30 to 430 °C. The TG technique involves the measurement of weight losses as a function of temperature; therefore, the TG curves demonstrate the thermal stability of the studies samples. According to the analysis of TG curves (solid lines), there are two distinguishable degradation steps for all xerogels: first, identified within a temperature range from ~55 to ~220 °C, and second, between ~220 and ~380 °C. The indicated degradation steps are marked in Figure 1, and the appropriate temperature ranges for individual xerogels doped with Dy3+ ions are collected in Table 1.
The first degradation step, which occurred in lower temperatures (in a range from ~55 to ~220 °C), is correlated with the evaporation of residual organic solvents (ethyl alcohol, acetic acid, unreacted TFA) and water desorption from the porous silicate sol-gel network, and is observed as a gentle degradation. Actually, xerogels are porous solid materials, with pores that are usually filled by liquids. Despite the vibrations characteristic for the silicate sol-gel network (~1200 cm−1 and below), there are additional bands identified as the vibrations of OH groups (>3000 cm−1), C=O moieties (1650 cm−1), and C–H bonds (~1390 cm−1, ~1460 cm−1), which clearly indicate the presence of water and organic compounds residues in xerogels before any heat-treatment [44]. Indeed, the boiling points of indicated chemical compounds under atmospheric pressure (i.e., 72 °C for TFA, 78 °C for C2H5OH, 100 °C for H2O, and 118 °C for CH3COOH). Hence, according to the 1st step of degradation, the evaporation of the compounds mentioned above is expected in a given temperature range. Supplementarily, the derivative thermogravimetry (DTG) expresses the results of TG by providing the first derivative curve as a function of temperature. DTG is a type of thermal analysis in which the rate of xerogels’ mass changes upon heating is plotted against temperature. Therefore, the temperature at which the maximum of the first DTG peak (dashed lines) occurs indicates the temperature at which the evaporation of water and organic compounds undergo the maximum rate. The temperatures from DTG curves for an individual sol-gel sample are identified in a range from 111 °C (XG6) to 157 °C (XG4) and are summarized in Table 1.
According to the processing of nano-glass-ceramic materials, the 2nd step of thermal degradation is essential because it is directly related to the crystallization of the LaF3 fluoride phase, preceded by La(TFA)3 decomposition. Indeed, this process involved a chemical reaction in which the compounds LaF3, (CF3CO)2O, CO2, and CO are obtained. The investigations of the mechanism of this reaction allowed us to conclude that such thermolysis led to cleavage of C–F bonds inside −CF3 groups from TFA ligand, and the fluorine anions (F) tend to react with La–O bonds, forming LaF3 phase resultantly [45]. From TG analysis (solid lines), the indicated transformation within the structure of prepared sol-gel materials is observable as a significant decrease in the mass. Accordingly, the temperatures at which the maximum of the DTG peaks for an individual sample are identified as approximately 300 °C. The indicated weight losses associated with La(TFA)3 thermal decomposition are estimated at 29.76 (XG1), 27.99 (XG2), 27.97 (XG3), 29.54 (XG4), 26.90 (XG5), and 28.02% (XG6). Moreover, for each fabricated xerogel, a strong exothermic DSC peak (dotted line) in this temperature range is recorded with a maximum near 300 °C. Therefore, such a degradation step is according to the release of energy and mass. The location of DSC peaks is consistent with the maxima of DTG peaks, which correspond to temperatures at which the transformation occurs the most rapidly.
The obtained results are consistent with the literature data, which clearly indicate that the thermal decomposition of metal trifluoroacetates and crystallization of appropriate fluoride phases occur at about 300 °C [44,46,47]. Additionally, based on the TG analysis, it should be noted that the prepared sol-gel samples are characterized by good thermal resistance at temperatures close to 350 °C. According to data collected in Table 1, it could be assumed that co-doping with Dy3+ ions was not influenced the thermal parameters of the prepared silicate xerogels. Based on collected data from TG (DTG) and DSC measurements, the temperature of a heat-treatment process to fabricate nano-glass-ceramics was assessed at 350 °C.

3.2. Structural Characterization of Fabricated Dy3+-Doped Sol-Gel Materials

Figure 2 shows the XRD diffractograms of Dy3+-doped precursor xerogels and samples obtained during controlled heat-treatment at 350 °C. For xerogels, the XRD patterns showed no sharp diffraction lines but only a broad hump with a maximum located near ~22°, which confirmed their amorphous nature devoid of long-range order [48]. Conversely, the sharp XRD lines are well-visible for heat-treated samples, and the lines are attributed to the hexagonal LaF3 phase crystallized in P63cm space group (ICDD card no. 00-008-0461). According to the literature, in the nearest framework around La3+ cations, there are nine F anions with four non-equivalent sites, including 3F1, 3F2, 2F3, and 1F4 [49]. The broadening of the diffraction lines was used to calculate the average diameter (D) of the crystallized LaF3 phase using the Scherrer equation [50]:
D = K λ β hkl cos θ
where K is a shape factor (in our calculations it was taken K = 1), λ is a wavelength of X-ray (0.154056 nm, Kα line of Cu), βhkl is a broadening of the (hkl) diffraction peak at half of the maximum intensity, and θ is a Bragg’s angle. The average crystallite size was estimated from 11.9 ± 0.1 (GC6) to 21.3 ± 0.5 nm (GC1). Additionally, the Williamson–Hall theorem was also used to determine the average size of LaF3 phase [51]:
β hkl cos θ = K λ D + 4 Z sin θ
in which βhkl is a broadening of the (hkl) diffraction line, θ is a diffraction angle, λ is an X-ray wavelength, D is an average crystal size, and Z is an effective strain. The lattice strain and the crystallite size were deduced from the intercept of βcos θ/λ versus sin θ/λ. The average crystal sizes of LaF3 from the Williamson–Hall method are similar for all fabricated SiO2-LaF3:Dy3+ nano-glass-ceramics and were estimated from 8.2 ± 0.1 (GC2) to 10.6 ± 0.1 nm (GC1). As can be seen from the obtained results, there is a noticeable difference in the size of the crystallites obtained by the Scherrer and Williamson–Hall methods. The difference is because the Scherrer method does not consider the share of internal stresses in the half-width of the XRD diffraction line. Contrary, the Williamson–Hall method separates the half-width into parts associated with the average crystallite size and parts related to internal stresses. If there would be no internal stresses in the material, the results of methods are convergent. If there are no internal stresses in the material, the results of the methods are convergent. However, the dysprosium ions caused some internal stress, so the estimated crystallite sizes obtained by these methods are slightly different. In the case of fabricated samples, Dy3+ ions, the inset of Figure 2 displays the high-resolution transmission electron microscope (HR-TEM) image of the prepared GC1 sample. Based on it, it was stated that the size of LaF3 nanocrystals is consistent with the average crystal size estimated from XRD analysis.
Figure 2 also shows an evident shift of (002), (110), and (111) diffraction lines toward higher angles as the content of Dy3+ ions increases in the subsequent samples in the prepared series. The shift in the position of (110) diffraction line (∆θ), compared with pure LaF3 phase, is about from 0.01 to 0.40° for GC2 and GC6 nano-glass-ceramics, respectively. These results indicate that the lattice parameters for the cation-exchanged LaF3:Dy3+ phase are smaller than for the pure fluoride phase without any admixtures of Dy3+ ions. So, because Dy3+ ions have a slightly smaller ionic radius (r = 1.083 Å) compared with La3+ cation (r = 1.216 Å) [52], some lattice distortions and intra-stress occur, as was presented in Table 2. Indeed, a general tendency to a progressive decrease in the cell parameters of fluoride nanocrystals was denoted (from a0 = 7.181(8) Å, c0 = 7.359(4) Å for GC1 up to a0 = 7.077(2) Å, c0 = 7.242(9) Å for GC6) in comparison with that of pure and undoped LaF3 phase (a0 = 7.184 Å, c0 = 7.351 Å). So, since the ionic radius of dopant (Dy3+) and cation from parent fluoride crystal lattice (La3+) are slightly different, the substitution of La3+ by Dy3+ modifies the inter-ionic distances and induces the perturbation in the lattice parameters. It generates stress inside the nanocrystal lattice, and for LaF3:Dy3+ system the compressive strain could be observed [53,54]. The lattice strain derived from the Williamson–Hall formula for fabricated sol-gel samples was estimated from 0.11 ± 0.01% to 0.27 ± 0.01%, indicating some lattice distortion. Interestingly, conversely to the above tendency, a very slight increase in c0 parameter for GC1 sample (c0 = 7.359(4) Å) in comparison with those of the pure LaF3 phase (c0 = 7.351 Å) was observed. It may be correlated with a peculiar property of crystals in the nanoscale, as was also denoted e.g., for CeO2 [55,56], BaF2 [57], or Pt nanoparticles [58].

3.3. Optical Properties of Dy3+-Doped Xerogels

Figure 3 illustrates the excitation spectra for the series of Dy3+-doped xerogels, registered on collecting the yellow emission at λem = 570 nm. Within the near-UV and VIS ranges, the 4f9-4f9 intra-configurational transitions originating from the 6H15/2 ground state of Dy3+ ions to the various excited levels were noted, appropriately labeled as the 6P3/2 (326 nm), 4I9/2 (340 nm), 6P7/2 (352 nm), 4I11/2 (366 nm), 4F7/2 (388 nm), 6G11/2 (427 nm), 4I15/2 (452 nm), as well as 4F9/2 (474 nm). It could be observed that the intensities of individual excitation bands have grown with decreasing La3+:Dy3+ molar ratio as the content of Dy3+ ions increased. On the other hand, since the intensities of excitation bands for XG5 and XG6 samples are comparable, it could be stated that the energy transfer processes between Dy3+ ions started to occur, suggesting the concentration quenching. The emission spectra of Dy3+-doped xerogels are presented in Figure 4. The spectra were recorded upon excitation at λex = 352 nm and show three luminescence bands at 477, 570, and 655 nm, according to the following transitions: 4F9/26H15/2 (blue), 4F9/26H13/2 (yellow), and 4F9/26H11/2 (red), as was also presented in the energy level scheme in Figure 5. For fabricated xerogels, the intensities of recorded bands increased with decreasing in La3+:Dy3+ molar ratio from XG1 to XG5 sample, but for XG6 (with the highest content of Dy3+) the luminescence started to quench, suggesting the occurrence of the energy transfer (ET) process between neighboring Dy3+ ions in the host.
Generally, the relative intensities of the 4F9/26H15/2 (ΔJ = 3, forbidden transition) and the 4F9/26H13/2 emissions (ΔJ = 2, hypersensitive electric–dipole transition) are influenced by the symmetry in the nearest framework around Dy3+ ions [59]. Based on recorded spectra, yellow-to-blue (Y/B) ratios were calculated, and the obtained values were equaled to 2.83, 2.37, 2.27, 2.34, and 2.31 for XG2-6, respectively. For XG1 xerogel sample, the Y/B-ratio was not calculated due to the presence of a broad band in a blue light region with a maximum at λ = 434 nm (not shown in the figure), which coincides with the 4F9/26H15/2 emission of Dy3+ ions. The indicated background is associated with defects inside the amorphous sol-gel host, as was stated in the literature [60]. Indeed, it is attributed to photon recombinations from plentiful defects associated with dangling bonds inside the sol-gel skeleton, and its appearance is independent of the introduced rare-earth dopant, as was proven in our earlier works concentrated on Tb3+ and Eu3+ spectroscopies [46,47]. For the same reason (correlated with overlapping of this broad band with blue emission of Dy3+ ions), the Y/B-ratio for XG2 sample is higher than the values calculated for other XG3-XG6 samples characterized by greater intensities of emission lines from Dy3+ ions. Our experimental results for XG3-XG6 samples indicate that Y/B-ratio values are set at a nearly constant level, despite La3+:Dy3+ molar ratio and Dy3+ content, which suggests no significant changes in the local environment around optically active ions in samples before heat-treatment. In general, such high Y/B-ratio values obtained for precursor xerogels specify a relatively high covalent nature of bonds between Dy3+ and the host [13], and they are comparable with the values declared in the literature for selected amorphous systems depicted in Table 3 [12,59,61,62,63,64,65,66,67,68]. Indeed, similar Y/B-ratio values (above 2) have been reported for 35.7SiO2-25.5B2O3-17BaO-3.4K2O-3.4Al2O3-15BaCl2:0.1–1Dy2O3 [59], and 50B2O3-(25−x)CaO-15Al2O3-10CaF2-xDy2O3 (x = 0.5–5) [61] glassy systems. The data collected in Table 3 clearly indicate the strong correlation between Y/B-ratios and modifications in chemical compositions of glasses and amorphous sol-gel materials.
The further characterization of Dy3+-doped xerogels involved the luminescence decay analysis from the 4F9/2 excited state, and the resultant curves are presented in Figure 6ex = 352 nm, λem = 570 nm). The registered luminescence decay curves followed the second-order exponential nature, and the average lifetimes were calculated using the following formula:
τ avg = A 1 τ 1 2 + A 2 τ 2 2 A 1 τ 1 + A 2 τ 2
where A1 and A2 are residual weighting factors and τ1 and τ2 are decay components. The resultant τn(4F9/2):Dy3+ lifetimes with A1 and A2 parameters are depicted in Table 4.
The average decay times continuously elongate as the content of Dy3+ ions increased in the following order: 26.6 ± 0.7 (XG1), 28.9 ± 0.5 (XG2), 32.2 ± 0.8 (XG3), 40.6 ± 0.3 (XG4), and 42.7 ± 0.3 μs (XG5). However, for the sample with the highest content of Dy3+ ions (XG6), an evident shortening in the τavg(4F9/2):Dy3+ value to 34.3 ± 0.1 μs was denoted, and it clearly corroborates with Dy3+-Dy3+ ET process. Generally, according to the numerous works in the literature, another factor that indicates the occurrence of the ET process among neighboring Dy3+ ions is the non-exponential behavior of the decays [12,14,64,67]. Based on this conception, we should assume that ET began to appear in the sample with the lowest content of Dy3+ (XG1), although its influence on the overall luminescence is negligible (indeed, we could observe the continuous elongation of the decays up to XG5 sample, simultaneously with growing intensities of the emission bands, as was presented in Figure 4). Therefore, for XG1-XG5 luminescence is proportional to the number of centers in an excited state. Further, for the XG6, the Dy3+-Dy3+ inter-ionic distances are the shortest in the series of fabricated xerogels, which makes the participation of ET enough to observe the shortening in the τavg(4F9/2) lifetime value and quenching the emission. The τ(4F9/2):Dy3+ lifetimes reported in the current literature for other amorphous systems, i.e., calcium boroaluminate glasses (510–800 μs) [61] or zinc-alumino-borosilicate glasses (296.5–673.7 μs) [62] are significantly longer compared with the decay times obtained for xerogels in this work. However, we could assume that the observed tendency should be related to the limited content of OH groups in glassy hosts prepared by the melt-quenching technique (in comparison with xerogels), which play a crucial role in quenching of the luminescence originating from Dy3+ ions. Indeed, the τavg(4F9/2):Dy3+ lifetimes for studied silicate xerogels are in the order of microseconds, and such relatively short luminescence lifetimes are strictly correlated with the presence of plentiful OH groups originated from silanol Si-OH moieties as well as residual organic solvents and water, inside a highly porous silicate network [44]. Since the 4F9/24F1/2 energy gap of Dy3+ ions equals only ΔE = ~7000 cm−1 [12], merely two high-energy phonons of OH groups (~3500 cm−1) are required to promote a non-radiative relaxation from the 4F9/2 excited state. As was also presented earlier by us for Eu3+ and Tb3+-doped samples [46], the non-radiative deactivation of the 4F9/2 level could also be partially caused by TFA ligands from RE3+ coordination sphere, containing carbonyl groups (~1665 cm−1; four groups to cover the energy gap) and C–F bonds (~1200 cm−1; six groups to cover the energy gap).

3.4. Luminescence Behavior of Dy3+-Doped Nano-Glass-Ceramics

Figure 7 shows the excitation spectra of SiO2-LaF3:Dy3+ nano-glass-ceramics, recorded by monitoring the characteristic yellow emission at λem = 570 nm. The spectra revealed the eight bands corresponding to the following electronic transitions: 6H15/26P3/2 (326 nm), 6H15/24I9/2 (339 nm), 6H15/26P7/2 (351 nm), 6H15/24I11/2 (364 nm), 6H15/24F7/2 (389 nm), 6H15/26G11/2 (427 nm), 6H15/24I15/2 (453 nm), and 6H15/24F9/2 (472 nm). Conversely to precursor xerogels, the excitation bands’ intensities gradually decrease with decreasing La3+:Dy3+ molar ratios in the subsequent GC1-GC6 samples. Hence, the concentration quenching phenomenon is observed from the nano-glass-ceramic with the lowest content of Dy3+ ions.
The emission spectra of Dy3+-doped nano-glass-ceramics, recorded upon excitation at λex = 351 nm are presented in Figure 8. Similarly, as for xerogels, the characteristic emission bands of Dy3+ were identified at 478 nm (4F9/26H15/2), 570 nm (4F9/26H13/2), and 657 nm (4F9/26H11/2). A progressive decrease in the relative intensities of recorded bands for subsequent nano-glass-ceramics was observed. It could be stated that the concentration quenching has occurred even from the sample with the lowest content of Dy3+ ions (GC1), as was also observed in excitation spectra (Figure 7). Generally, the concentration quenching could be realized through the resonant energy transfer (RET) or possible non-radiative cross-relaxation channels CR1-CR3, as was presented in the energy level scheme in Figure 5. According to these channels, an excited Dy3+ ion (donor, D) makes a downward transition, whereas a coupled unexcited neighbor Dy3+ (acceptor, A) simultaneously makes an appropriate upward transition. The electronic transitions involved in each of the individual channel could be denoted as follows:
RET: 4F9/2 (D) + 6H15/2 (A) → 6H15/2 (D) + 4F9/2 (A),
CR1: 4F9/2 (D) + 6H15/2 (A) → (6H9/2 + 6F11/2) (D) + 4F3/2 (A),
CR2: 4F9/2 (D) + 6H15/2 (A) → 6F5/2 (D) + (6H7/2 + 6F9/2) (A),
CR3: 4F9/2 (D) + 6H15/2 (A) → 6F3/2 (D) + (6H9/2 + 6F11/2) (A).
As was discussed according to the gradual shift of XRD diffraction lines (Figure 2), we could expect that part of Dy3+ ions were entered into LaF3 nanocrystal lattice, which significantly promotes the shortening of the Dy3+-Dy3+ inter-ionic distances. The incorporation of Dy3+ ions inside fluoride nanophase could also be stated based on a decrease in Y/B-ratio values compared to those for precursor xerogels: 2.71 (GC1), 2.53 (GC2), 2.34 (GC3), 2.20 (GC4), 2.06 (GC5), and 1.94 (GC6). Indeed, the denoted alterations in Y/B-ratios indicate that the bonding between Dy3+ ions and the local environment is less covalent in prepared nano-glass-ceramics than in xerogels, but it should be noted that the verified decrease in calculated ratios is slight. Furthermore, compared with Y/B-ratios declared in the literature (Table 5 [25,64,69,70,71,72]), the values calculated for prepared SiO2-LaF3:Dy3+ materials remained relatively high. Thus, we could suppose that calculated Y/B-ratio values are correlated with the presence of a broad band attributed to the photon recombinations from structural defects inside the silicate sol-gel host (still visible even after controlled heat-treatment of xerogels; the maximum of this band was shifted from ~434 nm (before heat-treatment) to ~465 nm (after heat-treatment)), which directly overlaps with the characteristic emission lines of Dy3+ ions within the blue (4F9/26H15/2) and the green (4F9/26H13/2) light spectral scopes. On the other hand, a clear trend could be noticed according to gradually decreasing Y/B-ratios for the subsequent SiO2-LaF3:Dy3+ nano-glass-ceramics as the content of optically active Dy3+ ions grow. It suggests an increasing tendency to accumulate Dy3+ ions in the LaF3 phase, which is also confirmed by the continuous shift of the XRD diffraction lines.
Additionally, the comparison of emission spectra recorded for xerogels and nano-glass-ceramics for individual La3+:Dy3+ molar ratios in samples’ compositions was presented in Figure 9. Based on this comparison, it can be stated that for La3+:Dy3+ molar ratios equal to 0.988:0.012, 0.97:0.03, and 0.94:0.06, the heat-treatment process enhances the intensity of the emission bands originated from Dy3+ ions. The most remarkable difference in the bands’ intensity can be observed when the content of Dy3+ is the lowest in the series of obtained samples (La3+:Dy3+ = 0.988:0.012). The correlation between luminescence intensities and La3+:Dy3+ molar ratio starts to change as the content of Dy3+ ions increases. For La3+:Dy3+ molar ratios equal to 0.88:0.12, 0.82:0.18, and 0.70:0.30, the emission intensities of luminescent bands of Dy3+ ions are greater for xerogels than for glass-ceramic materials. It is caused by the progressing concentration quenching, particularly for the highest content of Dy3+ ions, due to a significant shortening in the inter-ionic Dy3+-Dy3+ distances correlated with the incorporation of Dy3+ ions into the LaF3 fluoride phase.
The luminescence decay curves of the 4F9/2 state of Dy3+ for the series of prepared SiO2-LaF3 nano-glass-ceramics are illustrated in Figure 10. For all GCs, the decay curves follow the second-order exponential nature, which could, according to the distribution of Dy3+ ions, be either between a silicate xerogel host and fluoride nanocrystals with different decay rates, but could also indicate the ET process between neighboring Dy3+ ions in the host. The resultant τm(4F9/2):Dy3+ lifetimes with A1 and A2 parameters are depicted in Table 6. Indeed, for the subsequent Dy3+-doped samples, the progressive shortening of the lifetimes was observed, and the average decay times equaled: 1731.5 ± 5.7 (GC1), 1124.1 ± 2.5 (GC2), 612.2 ± 3.0 (GC3), 232.0 ± 2.3 (GC4), 143.8 ± 1.5 (GC5), and 119.8 ± 0.4 μs (GC6). That denoted tendency to shortening of the decay times clearly indicates the continuous concentration quenching. It should also be noted that the ET is much more noticeable for nano-glass-ceramics than for xerogels (the shortening of the τavg(4F9/2) was reported only for XG6 sample with La3+:Dy3+ molar ratio equals to 0.70:0.30), which is strictly associated with substantial decreasing in Dy3+-Dy3+ distances due to their partial entering into LaF3 nanocrystal lattice. Another noteworthy issue is related to the substantial elongation of the τ(4F9/2):Dy3+ lifetimes for nano-glass-ceramic materials in accordance with precursor xerogels, especially for samples with lower content of Dy3+, and it is associated with the low-phonon energy of LaF3 nanocrystal lattice (350 cm−1 [42]), which provides the low probability of depopulation of the excited states. Indeed, about 20 phonons of such fluoride phase would be needed to cover the energy gap between the 4F9/2 level and the 6F1/2 state of Dy3+ to quench the luminescence. Additionally, the remaining part of Dy3+ ions (which did not accumulate inside the fluoride lattice but are still located inside the amorphous sol-gel host) are surrounded by Q3 [SiO4] groups (1045 cm−1) with lower oscillation energy than OH moieties, which also reduces the probability of the 4F9/2 state depopulation. According to our previous research concentrated on the impact of structure on photoluminescence of RE3+ [46], it has been proven that the proposed thermal treatment conditions (350 °C/10 h) cannot trigger the complete elimination of OH groups from the sol-gel network; nevertheless, their amounts are significantly reduced compared to the xerogels. As a result, OH groups do not have a crucial impact on Dy3+ luminescence quenching. According to the literature data the τ(4F9/2):Dy3+ lifetimes for glass-ceramic materials with PbF2 [70], NaGd(WO4)2 [71], KNbO3 [73], or Ca2Ti2O6 [74] crystal phases do not exceed the value of 1 ms, while for studied SiO2-LaF3:Dy3+ nano-glass-ceramics (with lower contents of Dy3+ ions), longer lifetimes of about ~1.8 ms (GC1, La3+:Dy3+ = 0.988:0.012) and ~1.1 ms (GC2, La3+:Dy3+ = 0.97:0.03) were obtained. For higher contents of Dy3+ ions (when La3+:Dy3+ molar ratio equals 0.94:0.06, 0.88:0.12, and 0.70:0.30), the luminescence lifetimes are comparable with the values declared in the literature for those glass-ceramic systems [70,71,72,73,74].
Finally, it should be also pointed out that photoluminescence quantum yield (PLQY) is one of the essential spectroscopic parameters for RE3+-doped materials to judge their suitability for device fabrication, e.g., as visible light or infrared irradiation emitters. In the paper published by N. Maruyama et al. [75], the quantum yields for Dy3+-doped glass with 40BaO-20TiO2-40SiO2-0.5Dy2O3 and derivative nano-glass-ceramic were evaluated directly from measurements using an integrating sphere. As a result of the crystallization of precursor glasses, the intensities of emission bands according to the 4F9/26HJ (J = 15/2, 13/2, 11/2) transitions of Dy3+ ions significantly increased. As a result, the estimated quantum yield for Dy3+-doped nano-glass-ceramic is close to 15.2%, while for precursor glass it equaled 4.1%. Therefore, the quantum yield for nano-glass-ceramic is nearly 4-fold higher than for glass. Indeed, for Dy3+-doped sol-gel materials described in this work, the sum of the integrated intensities of individual blue (4F9/26H15/2), yellow (4F9/26H13/2), and red (4F9/26H11/2) emissions is at least 4.5-fold higher for SiO2-LaF3:Dy3+ nano-glass-ceramics compared with silicate xerogels before controlled heat-treatment. Nevertheless, it should be noted that this correlation is observed only for samples with low concentrations of Dy3+ ions in sol-gel hosts (with La3+:Dy3+ molar ratios equaled 0.988:0.012 and 0.97:0.03) when concentration quenching for glass-ceramics is inhibited. Thus, we believe that for those of fabricated nano-glass-ceramics, the quantum yield will be higher than for xerogels due to the preferable location of Dy3+ ions inside LaF3 fluoride nanocrystals and effective shortening of Dy3+-Dy3+ inter-ionic distances. These important aspects, according to the evaluation of luminescence quantum yields, will be examined in the future.

4. Conclusions

This paper presents the optical characterization of Dy3+-doped silicate xerogels and nano-glass-ceramics containing LaF3 phase, according to the structural modifications and variable La3+:Dy3+ molar ratios in the samples’ composition. The thermal degradation of La(TFA)3 and its transformation into the fluoride phase was verified by TG/DSC analysis, and XRD measurements confirmed the crystallization of LaF3 in the nanoscale. The luminescence characterization of prepared sol-gel samples involved the registration of excitation and emission spectra, along with the decay analysis from the 4F9/2 excited level of Dy3+. For amorphous xerogels, the concentration quenching occurs from the sample with the lowest proposed La3+:Dy3+ molar ratio (0.70:0.30, XG6), when the Dy3+-Dy3+ distances are the shortest in the series. The considerable differences in τ(4F9/2):Dy3+ lifetimes, the decrease in calculated Y/B-ratio, as well as the results from XRD analysis suggest the partial migration of Dy3+ from amorphous xerogel host into crystallized LaF3 nanophase during heat-treatment. Indeed, it was found that luminescence lifetimes are strongly dependent on Dy3+-Dy3+ inter-ionic distances determined by the content of optically active Dy3+ ions and the nature of prepared sol-gel materials (correlated with the vibrational energies in the immediate vicinity of optically active dopant). The embedding of Dy3+ inside LaF3 phase of prepared glass-ceramics resulted in continuous shortening of the inter-ionic distances, thus, the progressive quenching of the luminescence is observable even from the lowest content of Dy3+ (La3+:Dy3+ = 0.988:0.012, GC1). Simultaneously, the incorporation of Dy3+ into fluoride nanocrystals with low phonon energy resulted in substantial elongation of the τ(4F9/2) lifetimes compared with xerogels. The obtained results suggest that the fabricated Dy3+-doped materials could be predisposed for application as visible light emitters, like color screens or three-dimensional displays.

Author Contributions

Conceptualization, N.P.; methodology, N.P.; software, N.P.; validation, N.P., formal analysis, N.P.; investigation, N.P., T.G., E.P. and J.Ś.; resources, W.A.P.; data curation, N.P.; writing—original draft preparation, N.P.; writing—review and editing, N.P. and 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

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

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Archana, L.S.; Rajendran, D.N. Luminescence of rare earth doped ZnS nanophosphors for the applications in optical displays. Mater. Today Proc. 2021, 41, 461–467. [Google Scholar] [CrossRef]
  2. Frieiro, J.L.; Guillaume, C.; López-Vidrier, J.; Blázquez, O.; González-Torres, S.; Labbé, C.; Hernández, S.; Portier, X.; Garrido, B. Toward RGB LEDs based on rare earth-doped ZnO. Nanotechnology 2020, 31, 465207. [Google Scholar] [CrossRef] [PubMed]
  3. Liu, Z.; Ikesue, A.; Li, J. Research progress and prospects of rare-earth doped sesquioxide laser ceramics. J. Eur. Ceram. Soc. 2021, 41, 3895–3910. [Google Scholar] [CrossRef]
  4. Fois, M.; Cox, T.; Ratcliffe, N.; de Lacy Costello, B. Rare earth doped metal oxide sensor for the multimodal detection of volatile organic compounds (VOCs). Sens. Actuators B 2021, 330, 129264. [Google Scholar] [CrossRef]
  5. Mishra, L.; Sharma, A.; Vishwakarma, A.K.; Jha, K.; Jayasimhadri, M.; Ratnam, B.V.; Jang, K.; Rao, A.S.; Sinha, R.K. White light emission and color tunability of dysprosium doped barium silicate glasses. J. Lumin. 2016, 169, 121–127. [Google Scholar] [CrossRef]
  6. Jaidass, N.; Moorthi, C.K.; Babu, A.M.; Babu, M.R. Luminescence properties of Dy3+ doped lithium zinc borosilicate glasses for photonic application. Heliyon 2018, 4, e00555. [Google Scholar] [CrossRef] [Green Version]
  7. Rittisut, W.; Wantana, N.; Ruangtaweep, Y.; Mool-am-kha, P.; Padchasri, J.; Rujirawat, S.; Manyum, P.; Yimnirun, R.; Kidkhunthod, P.; Prasatkhetragarn, A.; et al. Bright white light emission from (Gd3+/Dy3+) dual doped transparent lithium aluminum borate glasses for W- LED application. Opt. Mater. 2021, 122, 111705. [Google Scholar] [CrossRef]
  8. Wang, L.; Guo, Z.; Wang, S.; Zhang, H.; Lv, H.; Wang, T.; Su, C. Luminescence properties of Dy3+ doped glass ceramics containing Na3Gd(PO4)2. J. Non-Cryst. Solids 2020, 543, 120091. [Google Scholar] [CrossRef]
  9. Bu, Y.Y.; Cheng, S.J.; Wang, X.F.; Yan, X.H. Optical thermometry based on luminescence behavior of Dy3+-doped transparent LaF3 glass ceramics. Appl. Phys. A 2015, 121, 1171–1178. [Google Scholar] [CrossRef]
  10. Komar, J.; Lisiecki, R.; Głowacki, M.; Berkowski, M.; Suszyńska, M.; Ryba-Romanowski, W. Spectroscopic parameters of Dy3+ ions in La3Ga0.5Ta0.5O14 single crystal. J. Lumin. 2020, 220, 116989. [Google Scholar] [CrossRef]
  11. Lin, X.; Zhao, L.; Jiang, B.; Mao, J.; Chi, F.; Wang, P.; Xie, C.; Wei, X.; Chen, Y.; Yin, M. Temperature-dependent luminescence of BaLaMgNbO6:Mn4+,Dy3+ phosphor for dual-mode optical thermometry. Opt. Mater. 2019, 95, 109199. [Google Scholar] [CrossRef]
  12. Murali Krishna, V.; Mahamuda, S.; Talewar, R.A.; Swapna, K.; Venkateswarlu, M.; Rao, A.S. Dy3+ ions doped oxy-fluoro boro tellurite glasses for the prospective optoelectronic device applications. J. Alloys Compd. 2018, 762, 814–826. [Google Scholar] [CrossRef]
  13. Pisarska, J.; Żur, L.; Pisarski, W.A. Optical spectroscopy of Dy3+ ions in heavy metal lead-based glasses and glass-ceramics. J. Mol. Struct. 2011, 993, 160–166. [Google Scholar] [CrossRef]
  14. Babu, P.; Jang, K.H.; Kim, E.S.; Shi, L.; Vijaya, R.; Lavín, V.; Jayasankar, C.K.; Seo, H.J. Optical Properties and energy transfer of Dy3+-doped transparent oxyfluoride glasses and glass-ceramics. J. Non-Cryst. Solids 2010, 356, 236–243. [Google Scholar] [CrossRef]
  15. Mahamuda, S.; Syed, F.; Annapurna Devi, C.B.; Swapna, K.; Prasad, M.V.V.K.S.; Venkateswarlu, M.; Rao, A.S. Spectral characterization of Dy3+ ions doped phosphate glasses for yellow laser applications. J. Non-Cryst. Solids 2021, 555, 120538. [Google Scholar] [CrossRef]
  16. Maheshwari, K.; Rao, A.S. Photoluminescence downshifting studies of thermally stable Dy3+ ions doped phosphate glasses for photonic device applications. Opt. Mater. 2022, 129, 112518. [Google Scholar] [CrossRef]
  17. Wu, H.; Liu, X.; Cao, L.; Zheng, Y.; Jin, W.; Li, L. Effect of lattice distortion induced by Li+ doping on white light CaWO4:Dy3+ phosphors: Phase, photoluminescence and electronic structure. J. Lumin. 2021, 235, 118023. [Google Scholar] [CrossRef]
  18. Sheoran, M.; Sehrawat, P.; Kumari, N.; Khatkar, S.P.; Malik, R.K. Cool white light emanation and photo physical features of combustion derived Dy3+ doped ternary yttrate oxide based nanophosphors for down converted WLEDs. Chem. Phys. Lett. 2021, 773, 138608. [Google Scholar] [CrossRef]
  19. Sehrawat, P.; Khatkar, A.; Boora, P.; Kumar, M.; Malik, R.K.; Khatkar, S.P.; Taxak, V.B. Emanating cool white light emission from novel down-converted SrLaAlO4:Dy3+ nanophosphors for advanced optoelectronic applications. Ceram. Int. 2020, 46, 16274–16284. [Google Scholar] [CrossRef]
  20. Eliseeva, S.V.; Salerno, E.V.; Lopez Bermudez, B.A.; Petoud, S.; Pecoraro, V.L. Dy3+ White Light Emission Can Be Finely Controlled by Tuning the First Coordination Sphere of Ga3+/Dy3+ Metallacrown Complexes. J. Am. Chem. Soc. 2020, 142, 16173–16176. [Google Scholar] [CrossRef]
  21. Anderson, B.R.; Gunawidjaja, R.; Eilers, H. Dy3+-doped yttrium complex molecular crystals for two-color thermometry in heterogeneous materials. J. Lumin. 2017, 188, 238–245. [Google Scholar] [CrossRef] [Green Version]
  22. Silva, I.G.N.; Kai, J.; Felinto, M.C.F.C.; Brito, H.F. White emission phosphors based on Dy3+-doped into anhydrous rare-earth benzenetricarboxylate complexes. Opt. Mater. 2013, 35, 978–982. [Google Scholar] [CrossRef]
  23. Grobelna, B.; Synak, A.; Głowaty, D.; Bojarski, P.; Szczodrowski, K.; Gryczyński, I.; Karczewski, J. Novel inorganic xerogels doped with CaWO4:xDy: Synthesis, characterization and luminescence properties. Mater. Chem. Phys. 2017, 199, 166–172. [Google Scholar] [CrossRef]
  24. Grobelna, B.; Synak, A.; Bojarski, P. The luminescence properties of dysprosium ions in silica xerogel doped with Gd1.6Dy0.4(WO4)3. Opt. Appl. 2012, XLII, 337–344. [Google Scholar]
  25. Grobelna, B.; Synak, A.; Bojarski, P.; Szczodrowski, K.; Kukliński, B.; Raut, S.; Gryczyński, I. Synthesis and luminescence characteristics of Dy3+ ions in silica xerogels doped with Ln2-xDyx(WO4)3. Opt. Mater. 2013, 35, 456–461. [Google Scholar] [CrossRef]
  26. Santana-Alonso, A.; Yanes, A.C.; Méndez-Ramos, J.; del-Castillo, J.; Rodríguez, V.D. Down-shifting by energy transfer in Dy3+-Tb3+ co-doped YF3-based sol-gel nano-glass-ceramics for photovoltaic applications. Opt. Mater. 2011, 33, 587–591. [Google Scholar] [CrossRef]
  27. Velázquez, J.J.; Rodríguez, V.D.; Yanes, A.C.; del-Castillo, J.; Méndez-Ramos, J. Increase in the Tb3+ green emission in SiO2-LaF3 nano-glass-ceramics by codoping with Dy3+ ions. J. Appl. Phys. 2010, 108, 113530. [Google Scholar] [CrossRef]
  28. Torres-Rodriguez, J.; Gutierrez-Cano, V.; Menelaou, M.; Kaštyl, J.; Cihlář, J.; Tkachenko, S.; González, J.A.; Kalmár, J.; Fábián, I.; Lázár, I.; et al. Rare-Earth Zirconate Ln2Zr2O7 (Ln: La, Nd, Gd, and Dy) Powders, Xerogels, and Aerogels: Preparation, Structure, and Properties. Inorg. Chem. 2019, 58, 14467–14477. [Google Scholar] [CrossRef]
  29. Cruz, M.E.; Durán, A.; Balda, R.; Fernández, J.; Mather, G.C.; Castro, Y. A new sol–gel route towards Nd3+-doped SiO2-LaF3 glass-ceramics for photonic applications. Mater. Adv. 2020, 1, 3589–3596. [Google Scholar] [CrossRef]
  30. Cheng, S.; Liu, L.; Yang, Q.; Li, Y.; Zeng, S. In vivo optical bioimaging by using Nd-doped LaF3 luminescent nanorods in the second near-infrared window. J. Rare Earths 2019, 37, 931–936. [Google Scholar] [CrossRef]
  31. Wu, H.; Fei, G.T.; Gao, J.; Huang, J.; Zhang, L.D. Single-Phase Organic−Inorganic Hybrid Nanoparticles for Warm-White Lighting. ACS Appl. Nano Mater. 2022, 5, 9112–9116. [Google Scholar] [CrossRef]
  32. Chen, Z.; Wang, W.; Kang, S.; Cui, W.; Zhang, H.; Yu, G.; Wang, T.; Dong, G.; Jiang, C.; Zhou, S.; et al. Tailorable Upconversion White Light Emission from Pr3+ Single-Doped Glass Ceramics via Simultaneous Dual-Lasers Excitation. Adv. Opt. Mater. 2018, 6, 1700787. [Google Scholar] [CrossRef]
  33. Zhang, H.; Dong, X.; Jiang, L.; Yang, Y.; Cheng, X.; Zhao, H. Comparative analysis of upconversion emission of LaF3:Er/Yb and LaOF:Er/Yb for temperature sensing. J. Mol. Struct. 2020, 1206, 127665. [Google Scholar] [CrossRef]
  34. Peng, Y.; Zhong, J.; Li, X.; Chen, J.; Zhao, J.; Qiao, X.; Chen, D. Controllable competitive nanocrystallization of La3+-based fluorides in aluminosilicate glasses and optical spectroscopy. J. Eur. Ceram. Soc. 2019, 39, 1420–1427. [Google Scholar] [CrossRef]
  35. Ansari, A.A.; Parchur, A.K.; Labis, J.P.; Shar, M.A. Physiochemical characterization of highly biocompatible, and colloidal LaF3:Yb/Er upconversion nanoparticles. Photochem. Photobiol. Sci. 2021, 20, 1195–1208. [Google Scholar] [CrossRef] [PubMed]
  36. Nie, L.; Shen, Y.; Zhang, X.; Wang, X.; Liu, B.; Wang, Y.; Pan, Y.; Xie, X.; Huang, L.; Huang, W. Selective synthesis of LaF3 and NaLaF4 nanocrystals via lanthanide ion doping. J. Mater. Chem. C 2017, 5, 9188–9193. [Google Scholar] [CrossRef]
  37. Chien, H.-W.; Huang, C.-H.; Yang, C.-H.; Wang, T.-L. Synthesis, Optical Properties, and Sensing Applications of LaF3:Yb3+/Er3+/Ho3+/Tm3+ Upconversion Nanoparticles. Nanomaterials 2020, 10, 2477. [Google Scholar] [CrossRef] [PubMed]
  38. Ansari, A.A.; Rai, M.; Rai, S.B. Impact of LaF3 and silica shell formation on the crystal, optical and photo-luminescence properties of LaF3:Ce/Tb nanoparticles. Mater. Chem. Front. 2017, 1, 727–734. [Google Scholar] [CrossRef]
  39. Mitroshenkov, N.V.; Matovnikov, A.V.; Kuznetsov, S.V.; Lazutkina, M.V.; Volchek, A.A.; Konoplin, N.A.; Kornev, B.I.; Novikov, V.V. Low-temperature anomalies of thermodynamic properties of lanthanum trifluoride LaF3 and (SrF2)0.5(LaF3)0.5 multivalent solid solution. J. Alloys Compd. 2022, 894, 162537. [Google Scholar] [CrossRef]
  40. Hu, M.; Zhu, Z.; Wang, Y.; Li, J.; You, Z.; Tu, C. Bulk Crystal Growth, First-Principles Calculations, and Mid-Infrared Spectral Properties of Dy3+ Doped and Dy3+/Nd3+ Codoped LaF3 Single Crystals. Cryst. Growth Des. 2018, 18, 5981–5990. [Google Scholar] [CrossRef]
  41. Hong, J.; Zhang, L.; Hang, Y. Enhanced 2.86 μm emission of Ho3+,Pr3+-codoped LaF3 single crystal. Opt. Mater. Express 2017, 7, 1509–1513. [Google Scholar] [CrossRef]
  42. Li, S.; Zhang, L.; Zhang, P.; Hong, J.; Xu, M.; Yan, T.; Ye, N.; Hang, Y. Spectroscopic characterisations of Dy:LaF3 crystal. Infrared Phys. Technol. 2017, 87, 65–71. [Google Scholar] [CrossRef]
  43. Cruz, M.E.; Castro, Y.; Durán, A. Transparent oxyfluoride glass-ceramics obtained by different sol-gel routes. J. Sol-Gel Sci. Technol. 2022, 102, 523–533. [Google Scholar] [CrossRef]
  44. 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]
  45. Sun, X.; Zhang, Y.W.; Du, Y.P.; Yan, Z.G.; Si, R.; You, L.P.; Yan, C.H. From Trifluoroacetate Complex precursors to Monodisperse Rare-Earth Fluoride and Oxyfluoride Nanocrystals with Diverse Shapes through Controlled Fluorination in Solution Phase. Chem. Eur. J. 2007, 13, 2320–2332. [Google Scholar] [CrossRef] [PubMed]
  46. 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. [Google Scholar] [CrossRef]
  47. Pawlik, N.; Szpikowska-Sroka, B.; Pietrasik, E.; Goryczka, T.; Pisarski, W.A. Photoluminescence and energy transfer in transparent glass-ceramics based on GdF3:RE3+ (RE = Tb, Eu) nanocrystals. J. Rare Earths 2019, 37, 1137–1144. [Google Scholar] [CrossRef]
  48. Khan, A.F.; Yadav, R.; Singh, S.; Dutta, V.; Chawla, S. Eu3+ doped silica xerogel luminescent layer having antireflection and spectrum modifying properties suitable for solar cell applications. Mater. Res. Bull. 2010, 45, 1562–1566. [Google Scholar] [CrossRef]
  49. Pokhrel, M.; Gupta, S.K.; Perez, A.; Modak, B.; Modak, P.; Lewis, L.A.; Mao, Y. Up- and Down-Convertible LaF3:Yb,Er Nanocrystals with a Broad Emission Window from 350 nm to 2.8 μm: Implications for Lighting Applications. Appl. Nano Mater. 2021, 4, 13562–13572. [Google Scholar] [CrossRef]
  50. Lim, D.J.; Marks, N.A.; Rowles, M.R. Universal Scherrer equation for graphene fragments. Carbon 2020, 162, 475–480. [Google Scholar] [CrossRef]
  51. Nath, D.; Singh, F.; Das, R. X-ray diffraction analysis by Williamson-Hall, Halder-Wagner and size-strain plot methods of CdSe nanoparticles—A comparative study. Mater. Chem. Phys. 2020, 239, 122021. [Google Scholar] [CrossRef]
  52. D’Angelo, P.; Zitolo, A.; Migliorati, V.; Chillemi, G.; Duvail, M.; Vitorge, P.; Abadie, S.; Spezia, R. Revised Ionic Radii of Lanthanoid(III) Ions in Aqueous Solution. Inorg. Chem. 2011, 50, 4572–4579. [Google Scholar] [CrossRef] [PubMed]
  53. Sharma, R.K.; Mudring, A.-V.; Ghosh, P. Recent trends in binary and ternary rare-earth fluoride nanophosphors: How structural and physical properties influence optical behavior. J. Lumin. 2017, 189, 44–63. [Google Scholar] [CrossRef]
  54. Sharma, R.K.; Ghora, M.; Chouryal, Y.N.; Ganguly, T.; Acharjee, D.; Mondal, D.J.; Konar, S.; Nigam, S.; Ghosh, P. Multifunctional Lanthanide-Doped Binary Fluorides and Graphene Oxide Nanocomposites Via a Task-Specific Ionic Liquids. ACS Omega 2022, 7, 16906–16916. [Google Scholar] [CrossRef]
  55. Kurian, M.; Kunjachan, C. Investigation of size dependency on lattice strain of nanoceria particles synthesised by wet chemical methods. Int. Nano Lett. 2014, 4, 73–80. [Google Scholar] [CrossRef] [Green Version]
  56. Prieur, D.; Bonani, W.; Popa, K.; Walter, O.; Kriegsman, K.W.; Engelhard, M.H.; Guo, X.; Eloirdi, R.; Gouder, T.; Beck, A.; et al. Size Dependence of Lattice Parameter and Electronic Structure in CeO2 Nanoparticles. Inorg. Chem. 2020, 59, 5760–5767. [Google Scholar] [CrossRef]
  57. Andrade, A.B.; Ferreira, N.S.; Valerio, M.E.G. Particle size effects on structural and optical properties of BaF2 nanoparticles. RSC Adv. 2017, 7, 26839–26848. [Google Scholar] [CrossRef] [Green Version]
  58. Leontyev, I.N.; Kuriganova, A.B.; Leontyev, N.G.; Hennet, L.; Rakhmatullin, A.; Smirnova, N.V.; Dmitriev, V. Size dependence of the lattice parameters of carbon supported platinum nanoparticles: X-ray diffraction analysis and theoretical considerations. RSC Adv. 2014, 4, 35959–35965. [Google Scholar] [CrossRef]
  59. Shasmal, N.; Karmakar, B. White light-emitting Dy3+-doped transparent chloroborosilicate glass: Synthesis and optical properties. J. Asian Ceram. Soc. 2019, 7, 42–52. [Google Scholar] [CrossRef] [Green Version]
  60. 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, 649–658. [Google Scholar] [CrossRef]
  61. Lodi, T.A.; Dantas, N.F.; Gonçalves, T.S.; de Camargo, A.S.S.; Pedrocki, F.; Steimacher, A. Dy3+ doped calcium boroaluminate glasses and Blue Led for smart white light generation. J. Lumin. 2019, 207, 378–385. [Google Scholar] [CrossRef]
  62. Monisha, M.; Mazumder, N.; Lakshminarayana, G.; Mandal, S.; Kamath, S.D. Energy transfer and luminescence study of Dy3+ doped zinc-aluminoborosilicate glasses for white light emission. Ceram. Int. 2021, 47, 598–610. [Google Scholar] [CrossRef]
  63. Bajaj, R.; Prasad, A.; Yeswanth, A.V.S.; Rohilla, P.; Kaur, S.; Rao, A.S. Down-shifting photoluminescence studies of thermally stable Dy3+ ions doped borosilicate glasses for optoelectronic device applications. J. Mater. Sci. Mater. Electron. 2022, 33, 4782–4793. [Google Scholar] [CrossRef]
  64. Babu, P.; Jang, K.H.; Rao, C.S.; Shi, L.; Jayasankar, C.K.; Lavín, V.; Seo, H.J. White light generation in Dy3+-doped oxyfluoride glass and transparent glass-ceramics containing CaF2 nanocrystals. Opt. Express 2011, 19, 1836–1841. [Google Scholar] [CrossRef] [PubMed]
  65. Walas, M.; Lisowska, M.; Lewandowski, T.; Becerro, A.I.; Łapiński, M.; Synak, A.; Sadowski, W.; Kościelska, B. From structure to luminescence investigation of oxyfluoride transparent glasses and glass-ceramics doped with Eu3+/Dy3+ ions. J. Alloys Compd. 2019, 806, 1410–1418. [Google Scholar] [CrossRef]
  66. Górny, A.; Kuwik, M.; Pisarski, W.A.; Pisarska, J. Lead Borate Glasses and Glass-Ceramics Singly Doped with Dy3+ for White LEDs. Materials 2020, 13, 5022. [Google Scholar] [CrossRef]
  67. Deopa, N.; Saini, S.; Kaur, S.; Prasad, A.; Rao, A.S. Spectroscopic investigations on Dy3+ ions doped zinc lead alumino borate glasses for photonic device application. J. Rare Earths 2019, 37, 52–59. [Google Scholar] [CrossRef]
  68. Thomas, V.; Jose, G.; Jose, G.; Biju, P.R.; Rajagopal, S.; Unnikrishnan, N.V. Structural Evolution and Fluorescence Properties of Dy3+: Silica Matrix. J. Sol-Gel Sci. Technol. 2005, 33, 269–274. [Google Scholar] [CrossRef]
  69. Sun, X.; Zhao, S.; Fei, Y.; Huang, L.; Xu, S. Structure and optical properties of Dy3+/Tm3+ co-doped oxyfluoride glass ceramics containing β-NaGdF4 nanocrystals. Opt. Mater. 2014, 38, 92–96. [Google Scholar] [CrossRef]
  70. Nageswara Rao, C.; Vasudeva Rao, P.; Kameswari, R.; Ramesh Raju, R.; Chandana, G.; Samatha, K.; Srinivas Prasad, M.V.V.K.; Venkateswarlu, M.; Naveen, A.; Dhar, G.G. Luminescence investigations on Dy3+ doped CdO-PbF2 phosphate glass-ceramics. J. Mol. Struct. 2021, 1243, 130784. [Google Scholar]
  71. Lv, H.; Wang, S.; Su, C.; Zhang, H.; Guo, Z.; Wang, L.; Wang, T.; Wei, Y. Preparation and luminescent properties of Dy3+-doped transparent glass-ceramics containing NaGd(WO4)2. J. Mater. Sci. Mater. Electron. 2020, 31, 6636–6644. [Google Scholar] [CrossRef]
  72. Wei, Y.; Zhang, H.; Su, C.; Wang, T.; Wang, S.; Lv, H.; Zou, X. Luminescence and preparation of Dy2O3 doped SrCO3-WO3-SiO2 glass ceramics. J. Lumin. 2020, 220, 117021. [Google Scholar] [CrossRef]
  73. Ramachari, D.; Rama Moorthy, L.; Jayasankar, C.K. Energy transfer and photoluminescence properties of Dy3+/Tb3+ co-doped oxyfluorosilicate glass-ceramics for solid-state white lighting. Ceram. Int. 2014, 40, 11115–11121. [Google Scholar] [CrossRef]
  74. Jia, F.; Xu, S.; Zhang, G.; Zhao, T.; Zou, X.; Zhang, H. Effect of Mg2+/Sr2+ addition on luminescence properties of Dy3+ doped glass ceramics contining Ca2Ti2O6. Opt. Mater. 2022, 131, 112715. [Google Scholar] [CrossRef]
  75. Maruyama, N.; Honma, T.; Komatsu, T. Enhanced quantum yield of yellow photoluminescence of Dy3+ ions in nonlinear optical Ba2TiSi2O8 nanocrystals formed in glass. J. Solid State. Chem. 2009, 182, 246–252. [Google Scholar] [CrossRef]
Figure 1. TG, DSC and DTG curves recorded for prepared XG1-XG6 xerogels (presented as solid, dashed, and dotted lines, respectively).
Figure 1. TG, DSC and DTG curves recorded for prepared XG1-XG6 xerogels (presented as solid, dashed, and dotted lines, respectively).
Nanomaterials 12 04500 g001
Figure 2. XRD patterns for the series of fabricated sol-gel samples. The region between 20° and 32° was enlarged to show the impact of Dy3+ content on diffraction lines shifting. The inset shows HR-TEM image of GC1 sample.
Figure 2. XRD patterns for the series of fabricated sol-gel samples. The region between 20° and 32° was enlarged to show the impact of Dy3+ content on diffraction lines shifting. The inset shows HR-TEM image of GC1 sample.
Nanomaterials 12 04500 g002
Figure 3. Photoluminescence excitation spectra (PLE) recorded for the series of fabricated xerogels by monitoring the yellow emission at λem = 570 nm.
Figure 3. Photoluminescence excitation spectra (PLE) recorded for the series of fabricated xerogels by monitoring the yellow emission at λem = 570 nm.
Nanomaterials 12 04500 g003
Figure 4. The photoluminescence emission (PL) spectra recorded for the series of prepared xerogels upon near-UV excitation at λem = 352 nm.
Figure 4. The photoluminescence emission (PL) spectra recorded for the series of prepared xerogels upon near-UV excitation at λem = 352 nm.
Nanomaterials 12 04500 g004
Figure 5. The energy level scheme of Dy3+ along with the cross-relaxation (CR) channels.
Figure 5. The energy level scheme of Dy3+ along with the cross-relaxation (CR) channels.
Nanomaterials 12 04500 g005
Figure 6. Luminescence decay curves recorded for the 4F9/2 state of Dy3+ ions in amorphous silicate xerogels (λex = 352 nm, λem = 570 nm).
Figure 6. Luminescence decay curves recorded for the 4F9/2 state of Dy3+ ions in amorphous silicate xerogels (λex = 352 nm, λem = 570 nm).
Nanomaterials 12 04500 g006
Figure 7. Photoluminescence excitation spectra (PLE) recorded for the series of fabricated SiO2-LaF3:Dy3+ nano-glass-ceramics by monitoring the yellow emission at λem = 570 nm.
Figure 7. Photoluminescence excitation spectra (PLE) recorded for the series of fabricated SiO2-LaF3:Dy3+ nano-glass-ceramics by monitoring the yellow emission at λem = 570 nm.
Nanomaterials 12 04500 g007
Figure 8. The photoluminescence emission (PL) spectra recorded for the series of prepared SiO2-LaF3:Dy3+ nano-glass-ceramics upon near-UV excitation at λem = 351 nm.
Figure 8. The photoluminescence emission (PL) spectra recorded for the series of prepared SiO2-LaF3:Dy3+ nano-glass-ceramics upon near-UV excitation at λem = 351 nm.
Nanomaterials 12 04500 g008
Figure 9. The comparison of emission spectra for xerogels and nano-glass-ceramic materials for individual La3+:Dy3+ molar ratios.
Figure 9. The comparison of emission spectra for xerogels and nano-glass-ceramic materials for individual La3+:Dy3+ molar ratios.
Nanomaterials 12 04500 g009
Figure 10. Luminescence decay curves recorded for the 4F9/2 state of Dy3+ ions for SiO2-LaF3 nano-glass-ceramic materials (λex = 351 nm, λem = 570 nm).
Figure 10. Luminescence decay curves recorded for the 4F9/2 state of Dy3+ ions for SiO2-LaF3 nano-glass-ceramic materials (λex = 351 nm, λem = 570 nm).
Nanomaterials 12 04500 g010
Table 1. The parameters from TG, DTG, and DSC analysis for studied XG1-XG6 silicate xerogels doped with Dy3+ ions.
Table 1. The parameters from TG, DTG, and DSC analysis for studied XG1-XG6 silicate xerogels doped with Dy3+ ions.
SampleThermal Degradation
1st Step2nd Step
Temperature
Range (°C)
Maximum of
DTG Peak
(°C)
Temperature Range (°C)Maximum of DTG Peak (°C)DSC Peak Position (°C)
XG155–223129223–383307307
XG256–222134222–382303302
XG350–224135224–382307307
XG457–226157226–378291298
XG551–217132217–386307307
XG645–201111201–375307309
Table 2. Crystal lattice parameters of LaF3 phase in prepared SiO2-LaF3:Dy3+ nano-glass-ceramics. The asterisk (*) refers to parameters of undoped LaF3 phase according to ICDD card no. 00-008-0461.
Table 2. Crystal lattice parameters of LaF3 phase in prepared SiO2-LaF3:Dy3+ nano-glass-ceramics. The asterisk (*) refers to parameters of undoped LaF3 phase according to ICDD card no. 00-008-0461.
SampleLattice Parameter [Å]Crystallite Size [nm]Lattice Strain [%]
LaF3 (*)Sol-Gel SampleScherrerWilliamson–Hall
GC1a0 = 7.184
c0 = 7.351
a0 = 7.181(8)
c0 = 7.359(4)
21.3 ± 0.510.6 ± 0.10.24 ± 0.01
GC2a0 = 7.172(4)
c0 = 7.351(9)
15.3 ± 0.38.2 ± 0.10.27 ± 0.01
GC3a0 = 7.161(0)
c0 = 7.343(3)
12.7 ± 0.19.0 ± 0.10.16 ± 0.01
GC4a0 = 7.147(7)
c0 = 7.321(9)
12.3 ± 0.19.0 ± 0.10.14 ± 0.01
GC5a0 = 7.139(6)
c0 = 7.311(5)
12.6 ± 0.19.8 ± 0.10.11 ± 0.01
GC6a0 = 7.077(2)
c0 = 7.242(9)
11.9 ± 0.19.0 ± 0.10.13 ± 0.01
Table 3. Y/B-ratios for different types of amorphous optical materials doped with Dy3+ ions.
Table 3. Y/B-ratios for different types of amorphous optical materials doped with Dy3+ ions.
Amorphous MaterialY/B-RatioReference
35.7SiO2-25.5B2O3-17BaO-3.4K2O-3.4Al2O3-15BaCl2
(mol%):0.1–1wt% Dy2O3 2
2.88–2.98[59]
XG2-XG6 12.83[this work]
2.37[this work]
2.27[this work]
2.34[this work]
2.31[this work]
50B2O3-(25 − x)CaO-15Al2O3-10CaF2-xDy2O3
(x = 0.5–5) wt% 2
1.94–2.18[61]
20SiO2-(40 − x)B2O3-10Al2O3-20NaF-10ZnO-xDy2O3
(x = 0.1–2.5) mol% 2
1.66–1.77[62]
35B2O3-20SiO2-(15 − x)Al2O3-15ZnO-15Na2CO3-xDy2O3 (x = 0.1–2.5) mol% 21.61–1.75[63]
45SiO2-20Al2O3-10CaO-24.9CaF2-0.1Dy2O3 mol% 21.52[64]
73TeO2-4BaO-3Bi2O3-18SrF2-2Dy2O3 mol% 21.50[65]
Ba2O3-PbO-Al2O3-WO3-Dy2O3 wt.% 2
(B2O3:PbO molar ratio changed from 2:1 to 1:8)
1.04–1.22[66]
(20 − x)Na2O-5BaF2-5CaF2-60B2O3-10TeO2-xDy2O3
(x = 0.5–2.5) mol% 2
0.86–1.11[12]
15ZnO-5PbO-(20 − x)Al2O3-60B2O3-xDy2O3
(x = 0.1–2.0) mol% 2
0.68–0.78[67]
TEOS-based xerogels 10.51–0.76[68]
1 materials prepared by the sol-gel method. 2 materials prepared by the conventional melt-quenching technique.
Table 4. Decay components (τn), residual weighting factors (An), and average decay times (τavg) of the 4F9/2 state of Dy3+ in fabricated silicate xerogels.
Table 4. Decay components (τn), residual weighting factors (An), and average decay times (τavg) of the 4F9/2 state of Dy3+ in fabricated silicate xerogels.
SampleDecay Components
(μs)
Residual Weighting Factors (%)Average Decay Time, τavg (μs)
τ1τ2A1A2
XG18.0 ± 0.135.8 ± 0.868.7631.2426.6 ± 0.7
XG29.4 ± 0.234.3 ± 0.650.7649.7628.9 ± 0.5
XG310.8 ± 0.339.8 ± 0.956.7943.2132.2 ± 0.8
XG412.7 ± 0.147.8 ± 0.349.1650.8440.6 ± 0.3
XG515.0 ± 0.250.4 ± 0.348.2851.7242.7 ± 0.3
XG613.0 ± 0.140.3 ± 0.146.5253.4834.3 ± 0.1
Table 5. Y/B-ratios for glass-ceramics doped with Dy3+ ions.
Table 5. Y/B-ratios for glass-ceramics doped with Dy3+ ions.
Type of Crystal PhaseY/B-RatioReference
LaF3 1
(350 °C)
2.74[this work]
2.53
2.34
2.20
2.06
1.94
CaF2 2 (650 °C, 700 °C)1.58
1.68
[64]
β-NaGdF4 2 (700 °C)1.51[69]
PbF2 2 (380 °C/2 h)1.18[70]
PbF2 2 (380 °C/5 h)1.05
PbF2 2 (380 °C/10 h)1.22
NaGd(WO4)2 2 (450 °C)1.0–1.1[71]
SrWO4 20.787–0.881[72]
Gd2(WO4)3 1,*0.23–1.37[25]
La2(WO4)3 1,*0.26–1.21[25]
1 materials prepared by sol-gel method. 2 materials prepared by conventional melt-quenching technique. * the crystal phase was not formed by in situ nucleation during controlled heat-treatment.
Table 6. Decay components (τn), residual weighting factors (An), and average decay times (τavg) of the 4F9/2 level of Dy3+ in prepared nano-glass-ceramics containing LaF3 phase.
Table 6. Decay components (τn), residual weighting factors (An), and average decay times (τavg) of the 4F9/2 level of Dy3+ in prepared nano-glass-ceramics containing LaF3 phase.
SampleDecay Components
(μs)
Residual Weighting Factors (%)Average Decay Time, τavg (μs)
τ1τ2A1A2
GC1302.7 ± 2.11920.6 ± 6.845.6454.361731.5 ± 5.7
GC2223.3 ± 0.81317.8 ± 3.055.9244.071124.1 ± 2.5
GC3143.1 ± 0.6782.5 ± 3.466.5133.49612.2 ± 3.0
GC464.2 ± 0.5305.1 ± 2.667.4432.56232.0 ± 2.3
GC551.4 ± 0.4197.6 ± 1.869.1230.88143.8 ± 1.5
GC642.9 ± 0.1180.2 ± 0.576.7623.24119.8 ± 0.4
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.; Goryczka, T.; Pietrasik, E.; Śmiarowska, J.; Pisarski, W.A. Photoluminescence Investigations of Dy3+-Doped Silicate Xerogels and SiO2-LaF3 Nano-Glass-Ceramic Materials. Nanomaterials 2022, 12, 4500. https://doi.org/10.3390/nano12244500

AMA Style

Pawlik N, Goryczka T, Pietrasik E, Śmiarowska J, Pisarski WA. Photoluminescence Investigations of Dy3+-Doped Silicate Xerogels and SiO2-LaF3 Nano-Glass-Ceramic Materials. Nanomaterials. 2022; 12(24):4500. https://doi.org/10.3390/nano12244500

Chicago/Turabian Style

Pawlik, Natalia, Tomasz Goryczka, Ewa Pietrasik, Joanna Śmiarowska, and Wojciech A. Pisarski. 2022. "Photoluminescence Investigations of Dy3+-Doped Silicate Xerogels and SiO2-LaF3 Nano-Glass-Ceramic Materials" Nanomaterials 12, no. 24: 4500. https://doi.org/10.3390/nano12244500

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