Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Unravelling the crystallization mechanism and structural evolution of Yb/Er-doped SiO2-GdF3 nano-glass ceramics

C. E. Secu, C. Bartha and M. Secu*
National Institute of Materials Physics, Bucharest-Magurele, Romania. E-mail: mihail.secu@infim.ro

Received 18th April 2025 , Accepted 1st July 2025

First published on 16th July 2025


Abstract

The crystallization mechanism of Yb/Er-doped GdF3 nanocrystals in silica nano-glass ceramics was analyzed using model-free and model-fitting methods and thermal analysis data in correlation with structural data. The formation of GdF3 nanocrystals occurs at around 300 °C, and their size is temperature dependent, ranging from 14 to 40 nm, depending on the processing temperature. A similar trend is observed for cell volume, where a contraction of up to ≈2.3% (at 600 °C) was assigned to the gradual incorporation of Li and Yb,Er dopants. Model-free analysis showed an increase in activation energy (Ea) and the preexponential factor (log[thin space (1/6-em)]A) up to 175 kJ mol−1 and 14.8 s−1, respectively, until the completion of crystallization. Model-fitting analysis indicated a crystallization process controlled by an autocatalytic-type reaction where a second metastable phase (LiF) acts as a catalyst and facilitates a rapid and self-accelerated crystallization of the main GdF3 nanocrystalline phase. The ceramization process boosted UC luminescence up to values comparable to those of NaYF4:18Yb/2Er.


1. Introduction

A novel class of nanostructured materials is represented by nano-glass-ceramics (or glass ceramics), where partial precipitation of a specific nanocrystalline phase within the amorphous glass is reached; nucleation and growth of nanocrystals occur in the parent glass matrix during subsequent thermal processing of the glass.1 The sol–gel approach represents an easy and controllable method for the preparation of transparent nano-glass ceramics, where optically active nanocrystals are embedded in a glass matrix, and optical transparency is assured by the small size of the nanocrystals.2,3

The optical properties of glass ceramics are widely recognized to be significantly influenced by their microstructure, which is directly related to the crystallization process. While these materials have been widely studied for their properties, only a limited number of studies have investigated their crystallization mechanisms and the relationship between these mechanisms and their optical properties.4,5 For the design of novel and/or improved application-based glass ceramic phosphors, a comprehensive study on the crystallization mechanism is complex but crucial as it can provide valuable insights into their structure and microstructure, as well as into how their structural and optical properties can be controlled through processing. It has been shown that formation of nano-glass ceramics involves a thermally activated chemical decomposition reaction of metal trifluoroacetates precursors, followed by the precipitation and growth of nanocrystals at high temperatures.2–5 Therefore, the nanocrystal formation process indicates thermal behaviors that can be analyzed through complex thermal analysis, which should include not only TG-DSC measurements but also thermokinetic analysis. The crystallization mechanism under non-isothermal conditions has been investigated through DTA or DSC analysis, based on models such as Avrami's theory, along with modified formulations of the Ozawa, Friedman, Matusita, or Kissinger equations.6–8 Non-isothermal crystallization kinetic analysis of the β-PbF2 or YF3 crystalline phase in silicate glass ceramics indicated a diffusion-controlled process of three-dimensional growth with decreasing nucleation.4,5 In contrast, we showed that BaF2 nanocrystal precipitation in a glass matrix is controlled by a homogeneous crystallization mechanism: nucleation centers resulting from thermal decomposition of Ba–trifluoroacetate grow into BaF2 nanocrystals at higher temperatures—a process revealing a distinct crystallization peak at high temperatures.9 Later, the same approach was used for the analysis of the crystallization mechanism of an SiO2-LaF3 xerogel, and a chemical decomposition reaction followed by the fast precipitation of crystals was proposed instead of a diffusion-controlled nucleation and growth process.10

The synthesis and optical/luminescence properties of glass ceramics with rare-earth doped GdF3 nanocrystals have been of keen interest for multicolor emitting phosphor applications11 owing to several advantages such as efficient energy transfer between Gd and RE-ions,12,13 and the low phonons energy of GdF3 (around 300 cm−1)14 that assures reduced multi-phonon relaxation rates, all of which lead to highly efficient luminescence properties. In addition, considering their desirable feature, such as low phonon energy,14 Yb/Er-doped GdF3 nanocrystals have been investigated for up-conversion (UC) luminescence properties;15,16 sequential absorption of two or more low energy photons (in infrared range) leads to light emission in the visible range, i.e. UC luminescence. However, the Yb/Er-doped GdF3 nanocrystals showed a tendency to aggregate and exhibited relatively poor UC properties, with enhancement being observed only after Li doping.15,16 Hence, in order to obtain higher UC luminescence efficiency, it is necessary to choose a proper synthesis method for nanoparticle dispersion along with low phonon energy host for high RE3+ ion luminescence efficiency.

A viable way may be to use a sol–gel chemistry approach useful for a wide range of compositions, assuring uniform distribution of optically active nanocrystals within the sample volume (through a thermally activated reaction) without agglomeration effects and with a high transparency degree due to their small size (tens of nm size).2,3 Investigations into RE3+-doped SiO2-GdF3 glass ceramics have demonstrated the precipitation of GdF3 nanocrystals in a silica matrix.11,13,17 The crystallization process of GdF3 nanocrystals (showing hexagonal or orthorhombic structure) is related to the gadolinium trifluoroacetate Gd(CF3COO)3 decomposition chemical reaction12 and driven by Li ion dopants.17 Our investigations highlighted an autocatalytic process where a second metastable phase (lithium fluoride) acts as a catalyst for the GdF3 crystallization process. Hence, the crystallization mechanism seems to be different compared with metal halide4,5 nanocrystal precipitation in glass crystallization, but it nevertheless remains incompletely understood.

The aim of this study is to extend our previous knowledge on crystallization processes of GdF3 nanocrystals in a silica glass matri17 and the UC luminescence properties using a deep and quantitative approach in order to provide further information that could either complement or confirm it. Model-free and model fitting methods were used to compute the kinetic parameters and propose a reaction model function using differential scanning calorimetry (DSC) data in correlation with data obtained from X-ray diffraction measurements. The absolute quantum yield QY of UC-luminescence of the Yb/Er-doped SiO2-GdF3 is also investigated.

2. Materials and methods

2.1. Sample preparation

For the preparation of the undoped and 4Yb3+/1Er3+-doped 90SiO2-5GdF3 (mol%) xerogels and those co-doped with Li, we used the sol–gel synthesis route17 and reagent-grade tetraethylorthosilicate (TEOS), trifluoroacetic acid (TFA), ethyl-alcohol, metal acetates, acetic acid and deionized water as starting materials. In the first step, TEOS was hydrolyzed under constant stirring with a mixed solution (8.2 mL) of ethanol and water and using glacial acetic acid as catalyst; molar ratio was 1[thin space (1/6-em)]:[thin space (1/6-em)]4[thin space (1/6-em)]:[thin space (1/6-em)]3.5[thin space (1/6-em)]:[thin space (1/6-em)]0.5. A second solution (3 mL) of Yb(CH3COO)3, Er(CH3COO)3, Li(CH3COO), Gd(CH3COO)3 and TFA with a Yb[thin space (1/6-em)]:[thin space (1/6-em)]Er[thin space (1/6-em)]:[thin space (1/6-em)]Li[thin space (1/6-em)]:[thin space (1/6-em)]Gd[thin space (1/6-em)]:[thin space (1/6-em)]F molar ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]0.7[thin space (1/6-em)]:[thin space (1/6-em)]3.6[thin space (1/6-em)]:[thin space (1/6-em)]3.6[thin space (1/6-em)]:[thin space (1/6-em)]89 was prepared, mixed and added dropwise to the first solution. After an additional vigorous stirring for 1 h at room temperature, the mixed solution was aged at room temperature for a week in a Petri dish. The wet gels were dried up to 120 °C during 1 week to form the xerogel, and glass ceramization was achieved after thermal calcination in air for 60 min at 525 °C, and the glass ceramic samples were denoted as GCLiYb/Er.

2.2. Characterization methods

The thermal behaviour of the xerogel was investigated using a Netzsch STA 449 F3 Jupiter simultaneous thermal analyser in the TG-DSC mode. Approximately 25 mg sample was measured in synthetic air (80% N2, 20% O2) at heating rates of 5, 10, 15, 20, and 25 °C min−1 from room temperature to 900 °C. Experiments were conducted in open alumina crucibles at a gas flow rate of 20 mL min−1. Heat flux sensitivity was ±0.001 mW, and temperature accuracy was ±0.01 °C. Non-isothermal kinetic analysis of the DSC data was performed using Netzsch Thermokinetics 3.1 software (version 072010).

The structural characterization of the materials was performed using X-ray diffractometry (XRD) and a BRUKER D8 ADVANCE type X-ray diffractometer, in focusing geometry, equipped with a copper target X-ray tube and LynxEye one-dimensional detector, in the 15°–65° range with a 0.05° step and 2 s integration time. For the analysis of the XRD patterns, we used PowerCell dedicated software.18

Photoluminescence and reflectance spectra were recorded at room temperature using a FluoroMax 4P spectrophotometer and its accessories. Up-conversion (UC) luminescence spectra were recorded under laser light pumping at 980 nm from a laser module (200 mW). For the quantitative measurements of UC luminescence efficiency, we used a 50 mm diameter Thorlabs integrating sphere coupled to a commercial spectrophotometer (Ocean Optics usb2000) (see ESI).

3. Theoretical background

The kinetics of xerogel crystallization were evaluated using model-free and model-fitting approaches. The model-free method allows the determination of activation energy and the preexponential factor without assuming a specific kinetic model for the crystallization process. In this study, the Friedman model was used. This model plots the natural logarithm of the conversion rate (ln(dα/dt)) against 1/T for measurements at different heating rates (β).19 According to theoretical studies,20 the equation that describes the relationship between the reaction rate, reacted fraction, and temperature, independent of the thermal sequence used for experiments, has the following form:
 
image file: d5ma00377f-t1.tif(1)
where α represents the reacted fraction, A is the Arrhenius pre-exponential factor, R is the gas constant, Ea is the activation energy, f(x) is the kinetic model (which is constant in the model-free analysis), and T is the temperature of the process. In logarithmic form, eqn (1) can be written as follows:
 
image file: d5ma00377f-t2.tif(2)

At a constant value of x, eqn (2) can be written as follows:

 
image file: d5ma00377f-t3.tif(3)

According to eqn (3), activation energy (Ea) can be determined from the slope of the plot of image file: d5ma00377f-t4.tif versus the inverse of temperature. The model fitting approach facilitates the identification of the crystallization reaction mechanism by minimizing the differences between experimental and calculated values. Using the kinetic parameters derived from the model-free analysis, the reaction model (f(x)), can be determined. The experimental data are then fitted to simulated curves, and through a statistical comparison of the fit for various models, the most suitable model along with its corresponding set of parameters is selected.21

4. Results and discussion

4.1. Thermal analysis

The TG-DSC curves of the Yb3+/Er3+-doped SiO2-GdF3 xerogel co-doped with Li measured in synthetic air with a heating rate of 10 °C min−1 are shown in Fig. 1(a). The DSC curve shows four peaks of different intensities, accompanied by a mass loss, as shown in the TG curve. The first endothermic peak, with a maximum at a temperature of 92 °C, is associated with a minor weight loss of approximately 10 wt%, which is attributed to the evaporation of water and residual organic fragments.
image file: d5ma00377f-f1.tif
Fig. 1 TG-DSC curves of the Yb3+/Er3+-doped SiO2-GdF3 xerogel co-doped with Li, measured in synthetic air at a heating rate of 10 °C min−1 (a) and the crystallization peak highlighted at five different heating rates: 5, 10, 15, 20, and 25 °C min−1 (b).

The main process revealed by thermal analysis is the thermolysis of metal (M) trifluoroacetate,22 and the reaction is characterized by a distinct exothermic DSC peak around 300 °C and a weight loss of approximately 38 wt%. A possible mechanism for this reaction involves the breaking of the C–F bond in trifluoroacetate ligands during thermal processing, leading to the formation of new M–F bonds and subsequent growth of MF3 nanocrystals.23 The 300 °C peak was assigned to gadolinium trifluoroacetate Gd(CF3COO)3 thermolysis with subsequent GdF3 nanocrystalline phase precipitation11,17 and overlaps with lithium and RE-trifluoroacetate decomposition, which occurs in the same temperature range.22

A much weaker exothermic DSC peak at 550 °C is attributed to the crystallization of the LiGdF4 phase, and the small DSC peak at 810 °C is likely associated with glass melting. At higher heating rates, the crystallization maxima shifted to higher temperatures, and their surface areas increased, as shown in Fig. 1(b). This occurs because faster heating leads to the production of a higher number of crystals within the same time.10

4.2. Structural and morphological characterization

The XRD pattern analysis allows tracking structural modifications during the annealing of the Yb3+/Er3+-doped SiO2-GdF3 xerogel (Fig. 2). The initial dried xerogel annealed up to 250 °C shows a broad background that is characteristic of the amorphous structure of the silica matrix, but as the annealing temperature increases, extra-diffraction peaks assigned to the nanocrystalline GdF3 fluoride phase precipitation start to develop. The occurrence of GdF3 nanocrystals dispersed in the amorphous glassy matrix was confirmed using transmission electron microscopy as well.17 As is known, GdF3 exhibits two crystalline structures, orthorhombic (space group Pnma (no. 62)) and hexagonal (space group P63/mcm (no. 193)), which differ significantly in terms of coordination number, lattice volume, and other parameters. The crystallization of the orthorhombic GdF3 nanocrystals (PDF 012-0788) can be observed in the 225–350 °C temperature range and agrees very well with the strong DSC peak at 300 °C from thermal analysis measurements (Fig. 1). At higher temperatures (600 °C), LiGdF4 phase crystallization is observed, accompanied by weaker SiO2 peaks.
image file: d5ma00377f-f2.tif
Fig. 2 XRD patterns of GCLiYb/Er glass-ceramic samples after annealing at different temperatures; the XRD pattern of orthorhombic GdF3 (PDF 012-0788) is included.

The XRD pattern analysis of the GCLiYb/Er sample annealed at different temperatures according to the TG/DSC curves (Fig. 1) indicates a strong lattice relaxation effect compared with GdF3 (PDF 012-0788), revealed by the shift to higher angles (Fig. 2), along with a steady increase in the nanocrystals size, up to about 40 nm (Table 1). The ionic radii of 8-fold coordinated Yb3+ ions (105.3 pm) and Er3+ ions (100.4 pm) are much smaller than that of 8-fold coordinated Gd3+ ions (119 pm),24 and the contraction effect (about 2.3% at 600 °C) was assigned to the Li+ and Yb3+, Er3+ ion incorporation in the lattice. The contraction effect solely related to Yb/Er doping is about 1.4%,25 and therefore, the additional contraction up to 2.3% is related to Li doping. The contraction effect follows nanocrystal growth, with the unit cell volume decreasing by approximately 1.7% at 300 °C and up to 2.3% at 600 °C, relative to orthorhombic GdF3 (as shown in Table 1). Hence, we suppose that the contraction effect is due to the incorporation of Li and Yb,Er ions into the lattice during the nanocrystal growth, most likely through the Ostwald ripening mechanism, where the growth of large nanocrystals occurs through the coalescence of smaller ones.

Table 1 Results of the X-ray diffraction (XRD) pattern analysis of GCLiYb/Er glass-ceramic samples annealed at different temperatures; the lattice parameters for GdF3 (PDF 012-0788) are included for comparison
Temperature (°C)/lattice parameters a (Å) b (Å) c (Å) Cell volume (Å)3 D (nm)
275 6.471 6.951 4.430 199.2 14
300 6.471 6.932 4.416 198.1 14
350 6.471 6.933 4.423 198.4 16
525 6.471 6.915 4.407 197.2 25
600 6.471 6.918 4.398 196.9 40
Orth-GdF3 (PDF file) 6.571 6.984 4.393 201.6


4.3. Thermokinetic analysis

4.3.1. Model-free thermokinetic analysis. The activation energy and preexponential factor estimated using Friedman analysis are shown in the Fig. 3.
image file: d5ma00377f-f3.tif
Fig. 3 Friedman analysis of the crystallization of the Yb3+/Er3+-doped SiO2-GdF3 xerogel co-doped with Li (a). Dependence of activation energy and the pre-exponential factor on the crystallization fraction according to the Friedman method (b).

Model-free analysis indicates that activation energy (Ea) and the preexponential factor (log[thin space (1/6-em)]A) gradually increase up to a partial area (x) of 0.4, reaching values of approximately 171 kJ mol−1 and 14.2 s−1, respectively. These parameters remain constant until x reaches 0.8, after which a slow increase is observed. At the end of crystallization (x = 1), we obtained an Ea of approximately 175 kJ mol−1 and a log[thin space (1/6-em)]A of about 14.8 s−1. These values are comparable to those reported for YF3–SiO2 glass-ceramics, where Ea values of approximately 129 and 139 kJ mol−1 were obtained using Kissinger's and Chen's methods, respectively.5 Similar values of activation energy for PbF2 crystallization were calculated using the Kissinger method to be about 162 kJ mol−1 and 167 kJ mol−1 for x = 0 and x = 1, respectively.4 In contrast, the Ea value for 80SiO2-90LaF2 glass-ceramics is significantly higher, of about 293 kJ mol−1.10

4.3.2. Model-fitting analysis. The DSC curves recorded at different heating rates were analyzed and fitted to several reaction models26 using multivariate least squares regression with the Thermokinetics program. According to Fig. 4, the best-fitted reaction model for the xerogel is the autocatalytic Prout-Tompkins model (Bna).
image file: d5ma00377f-f4.tif
Fig. 4 Experimental data for the Yb3+/Er3+-doped SiO2-GdF3 xerogel co-doped with Li and the fitting curve of the 300 °C crystallization peak.

Previous investigations have shown the Prout-Tompkins autocatalytic model to be useful for describing a crystallization mechanism in which the crystallization rate is influenced by two key factors—the degree of conversion and the presence of catalytic compounds—both of which contribute to accelerating the crystallization process.26,27 The mathematical representation of this model28 is given by the following equation:

 
image file: d5ma00377f-t5.tif(4)
where n is the order of reaction, and a is the is the autocatalytic constant. The kinetic parameters are listed in Table 2: Fexp. is the experimental form of the reaction model (f(x)), obtained from DSC data, and the model with an Fexp-value of 1 has the least deviation out of all the models compared; Fcrit is the critical value (quantile), obtained from F-distribution at a significance level of 0.05; tcrit is the critical value, obtained from t-distribution, and is defined for a given significance level and degrees of freedom.

Table 2 Kinetic model parameters resulting from non-linear regression analysis
Parameters Values
Log[thin space (1/6-em)]A (s−1) 12.25 ± 0.67
Ea (kJ mol−1) 157.50 ± 2.73
React. Ord. 1 1.574 ± 0.09
Exponent a 0.211 ± 0.006
Fexp 1
Fcrit (0.95) 1.08
Statistical parameters
Correlation coefficient 0.994
Durbin–Watson factor 2.41
Rel. precision 0.00100
t-Critic (0.95; 136) 1.969


According to this model, the amorphous xerogel sample shows rapid self-accelerated crystallization, which occurs through the simultaneous formation of Li-related metastable phases, resulting from the thermal decomposition of metal trifluoroacetates at temperatures around 300 °C. Although these phases are not detected in XRD patterns at this temperature—likely because of their low crystallinity, nanoscale dimensions, or limited concentration29 —they play a crucial catalytic role in the crystallization process. Li+ ion doping is favorable for the formation of metastable phases that can act as diffusion barriers or phase interfaces, preventing uncontrolled crystallization kinetics (e.g., by delaying the crystallization of stable phases, blocking volatile pathways that would intensify autocatalysis, modifying diffusion in the solid state, etc.). By lowering the activation energy, they facilitate the crystallization of the Yb3+/Er3+-doped SiO2-GdF3 phase, thus promoting rapid phase formation. Furthermore, a reaction order value (n) greater than 1 (with n ≈ 1.57) suggests an accelerated nucleation and crystallite growth rate. This leads to a nonlinear increase in the crystallization rate, which can result in the formation of larger crystallites or rapid development of the crystalline phase. However, if not properly controlled, such conditions may also introduce issues, including structural defects or incomplete crystallization. Statistical parameters indicate that the model fits well, showing a strong correlation, no significant autocorrelation in the residuals, high precision, and statistical significance in the parameter estimates. This is supported by the Durbin–Watson statistic, which tests for autocorrelation in the residuals. A value near 2 indicates no significant autocorrelation.29 Since the value of 2.41 is close to 2, it suggests that the residuals are independent, which is a positive indication.

4.4. Up-conversion luminescence properties

We observed that Li-ion co-dopant influence is not limited only to crystallization behavior (see above) but also on structural properties (for higher doping levels25) accompanied by optical property improvements.17,25 Under 980 nm IR light pumping, the GCLiYb/Er glass ceramic sample showed UC luminescence of Er3+ ions, caused by the energy transfer Yb–Er within the GdF3 nanocrystals (Fig. 5).
image file: d5ma00377f-f5.tif
Fig. 5 Normalized UC luminescence spectra recorded for GCLiYb/Er glass-ceramic (after calcination at 525 °C) and β-NaYF4:18Yb/2Er polycrystalline powder recorded under 980 laser light pumping and the possible mechanism with the RE energy levels involved (inset).

The substitutional Li+ ions tailor the crystal field symmetry and alter the environment of Er3+,15,16 leading to an increase in UC luminescence.25 The UC mechanism has been extensively investigated in various host Yb3+/Er3+-doped nanocrystalline host materials, including GdF3,15,16 and oxyfluoride glass ceramics.30–33 It is based on highly efficient IR light absorption by Yb3+ ions at around 1000 nm (2F7/22F5/2 transition). For unsaturated up-conversion processes, the UC luminescence intensity is proportional to the nth power of the incident pump power. The value of n, corresponding to the number of pump photons required for the population of the emitting level,34 can be extracted from a double logarithmic plot, luminescence intensity vs. incident pump intensity, where n is the slope of this dependence (see ESI). The Er3+ emitting levels are feed by the two-photon energy transfer (ET) process, accompanied by multi-phonon and cross-relaxation (Yb3+–Er3+) processes, followed by the characteristic green ((2H11/2, 4S3/2) → 4I15/2) and red (4F9/24I15/2) luminescence (Fig. 5 inset); for “uv-blue” up-conversion luminescence (4G11/2, 2H9/24I15/2), three-photon energy transfer processes are involved.30,33

The quantitative measurement of the absolute quantum yield (QY) of up-conversion is challenging; therefore, we used the method proposed by J. C. Boyer et al.35 For comparison and verification, we measured β-NaYF4:18Yb/2Er polycrystalline powder, known to show the best luminescence property among UC luminescent materials. We obtained a QY of 0.2 ± 0.1% for GCLiYb/Er glass ceramic, and 0.3 ± 0.1% for the β-NaYF4:18Yb/2Er, which matches well with the values reported in the literature;35 QY can increase under higher power density.36 For applications, optical transparency is a very important parameter. In the present case, GCLiYb/Er glass ceramic shows good optical transparency in the blue to red spectral optical range (300–700 nm), as shown by the high reflectance values of about 0.75; the small dips at 377, 485, 519 and 650 nm are assigned to the Er3+ ions characteristic absorptions from the ground state 4I15/2 to 4G11/2,9/2, 4F7/2, 2H11/2 and 4F9/2, respectively (see ESI).

5. Conclusion

Complex thermal analysis, correlated with structural data, showed that the precipitation of GdF3 nanocrystals results from the thermolysis of gadolinium-trifluoroacetate at approximately 300 °C. The crystallization mechanism was found to be autocatalytic, with Li-based metastable phases acting as catalysts by lowering activation energy barriers for nanocrystal nucleation and growth. Thermokinetic analysis also indicated an accelerated rate of crystal nucleation and growth, influencing the material structure and microstructure. This behavior was confirmed by the evolution of optical properties, which showed enhanced UC luminescence, reaching values comparable to those of the NaYF4:18Yb/2Er phosphor while maintaining the optical transparency of the material.

Conflicts of interest

There are no conflicts to declare.

Data availability

The experimental data that support the findings of this study will be available on request.

Acknowledgements

The authors acknowledge funding by the Core Program of the National Institute of Materials Physics (NIMP), granted by the Romanian Ministry of Research, Innovation and Digitization through the Project PC3-PN23080303. The help of Dr Silviu Polosan for the Up-Conversion luminescence set-up and efficiency measurements is gratefully acknowledged.

References

  1. E. D. Zanotto, A bright future for glass-ceramics, Am. Ceram. Soc. Bull., 2010, 89(8), 19–27 CAS.
  2. G. Gorni, J. J. Velázquez, J. Mosa, R. Balda, J. Fernández, A. Durán and Y. Castro, Transparent Glass-Ceramics Produced by Sol–Gel: A Suitable Alternative for Photonic, Materials, 2018, 11(212), 1 Search PubMed.
  3. M. Secu, C. Secu and C. Bartha, Optical Properties of Transparent Rare-Earth Doped Sol-Gel Derived Nano-Glass Ceramics, Materials, 2021, 14(22), 6871 CrossRef CAS PubMed.
  4. W. Luo, Y. Wang, F. Bao, L. Zhou and X. Wang, Crystallization behavior of PbF2–SiO2 based bulk xerogels, J. Non-Cryst. Solids, 2004, 347, 31–38 CrossRef CAS.
  5. W. Luo, Y. Wang, Y. Cheng, F. Bao and L. Zhou, Crystallization and structural evolution of YF3 –SiO2 xerogel, Mater. Sci. Eng. B, 2006, 127, 218–223 CrossRef CAS.
  6. Z. Hu, Y. Wang, F. Bao and W. Luo, Crystallization behavior and microstructure investigations on LaF3 containing oxyfluoride glass ceramics, J. Non-Cryst. Solids, 2005, 351, 722–728 CrossRef CAS.
  7. D. Chen, Y. Wang, Y. Yu and Z. Hu, Crystallization and fluorescence properties of Nd 3+ - doped transparent oxyfluoride glass ceramics, Mater. Sci. Eng. B, 2005, 123, 1–6 CrossRef.
  8. Y. Yu, D. Chen, Y. Cheng, Y. Wang, Z. Hu and F. Bao, Investigation on crystallization and influence of Nd3+ doping of transparent oxyfluoride glass-ceramics, J. Eur. Ceram. Soc., 2006, 26, 2761–2767 CrossRef CAS.
  9. C. E. Secu, C. Bartha, S. Polosan and M. Secu, Thermally activated conversion of a silicate gel to an oxyfluoride glass ceramic: Optical study using Eu+ probe ion, J. Lumin., 2014, 146, 539–543 CrossRef CAS.
  10. G. Gorni, M. J. Pascual, A. Caballero, J. J. Velázquez, J. Mosa, Y. Castro and A. Durán, Crystallization mechanism in sol-gel oxyfluoride glass-ceramics, J. Non-Cryst. Solids, 2018, 501, 145–152 CrossRef CAS.
  11. J. J. Velázquez, J. Mosa, G. Gorni, R. Balda, J. Fernández, L. Pascual, A. Durán and Y. Castro, Transparent SiO2 GdF3 sol–gel nano-glass ceramics for optical applications, J. Sol-Gel Sci. Technol., 2019, 89, 322–332 CrossRef.
  12. H. Guan, Y. Sheng, C. Xu, Y. Dai, X. Xie and H. Zou, Energy transfer and tunable multicolor emission and paramagnetic properties of GdF3:Dy3+,Tb3+,Eu3+ phosphors, Phys. Chem. Chem. Phys., 2016, 18, 19807–19819 RSC.
  13. N. Pawlik, B. Szpikowska-Sroka, E. Pietrasik, T. Goryczka and W. A. Pisarski, Photoluminescence and Energy transfer in transparent glass-ceramics based on GdF3:RE3+ (RE = Tb, Eu) nanocrystals, J. Rare Earths, 2019, 37, 1137–1144 CrossRef CAS.
  14. T. Miyakawa and D. L. Dexter, Cooperative and stepwise excitation of luminescence: trivalent rare-earth ions in Yb3+ sensitized crystals, Phys. Rev. B:Condens. Matter Mater. Phys., 1970, 1, 70–80 CrossRef CAS.
  15. H. Wang and T. Nann, Monodisperse upconversion GdF3:Yb, Er rhombi by microwave-assisted synthesis, Nanoscale Res. Lett., 2011, 6(1), 267 CrossRef PubMed.
  16. W. Yin, L. Zhao, L. Zhou, Z. Gu, X. Liu, G. Tian, S. Jin, L. Yan, W. Ren, G. Xing and Y. Zhao, Enhanced red emission from GdF3:Tb3+,Er3+ upconversion nanocrystals by Li+ doping and their application for bioimaging, Chemistry, 2012, 18(30), 9239–9245 CrossRef CAS.
  17. C. E. Secu, C. Bartha, C. Radu and M. Secu, Crystallization processes of rare-earth doped GdF3 nanocrystals in silicate glass matrix: Dimorphism and photoluminescence properties, Ceram. Int., 2024, 20, 37518–37524 CrossRef.
  18. W. Krause and G. Nolze, PowderCell a program for the representation and manipulation of crystal structures and calculation of the resulting X-ray patterns, J. Appl. Cryst., 1996, 29, 301–303 CrossRef.
  19. J. M. Criado, P. E. Sanchez-Jimenez and L. A. Perez-Maqueda, Critical study of the isoconversional methods of kinetic analysis, J. Therm. Anal. Calorim., 2008, 92, 199–203 CrossRef CAS.
  20. A. Khawam and D. R. Flanagan, Solid-state kinetic models: basics and mathematical fundamentals, J. Phys. Chem. B, 2006, 110, 17315–17328 CrossRef CAS PubMed.
  21. E. Moukhina, Determination of kinetic mechanisms for reactions measured with thermoanalytical instruments, J. Therm. Anal. Calorim., 2012, 109, 1203–1214 CrossRef CAS.
  22. Y. Yoshimura and K. Ohara, Thermochemical studies on the lanthanoid complexes of trifluoroacetic acid, J. Alloys Compd., 2006, 408–412, 573–576 CrossRef CAS.
  23. A. C. Yanes, J. Del-Castillo, J. Méndez-Ramos, V. D. Rodriguez, M. E. Torres and J. Arbiol, Luminescence and structural characterization of transparent nanostructures Eu3+-doped LaF3-SiO2 glass-ceramics prepared by sol-gel method, Opt. Mater., 2007, 29(8), 999–1003 CrossRef CAS.
  24. R. D. Shannon, Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides, Acta Crystallogr., Sect. A, 1976, 32, 751–767 CrossRef.
  25. C. Secu, C. Bartha, C. Radu and M. Secu, M. Up-Conversion Luminescence and Magnetic Properties of Multifunctional Er3+/Yb3+-Doped SiO2-GdF3/LiGdF4 Glass Ceramics, Magnetochemistry, 2023, 9, 11 CrossRef CAS.
  26. S. Vyazovkin and C. A. Wight, Model-free and model-fitting approaches to kinetic analysis of isothermal and nonisothermal data, Thermochim. Acta, 1999, 340–341, 53–68 CrossRef CAS.
  27. J. M. S. P. Shahi, S. K. Srivastava and R. S. Tiwari, A Study of Crystallization Kinetics of an Amorphous Glass-Ceramic System Using the Prout-Tompkins Model, J. Non-Cryst. Solids, 2012, 358(16), 2170–2175 Search PubMed.
  28. K. Prout and F. W. Tompkins, The theory of the crystallization kinetics of metals and alloys, Trans. Faraday Soc., 1950, 46, 791–803 Search PubMed.
  29. J. Durbin and G. S. Watson, Testing for serial correlation in least squares regression. I, Biometrika, 1950, 37(3–4), 409–428 CAS.
  30. J. J. Velázquez, G. Gorni, R. Balda, J. Fernández, L. Pascual, A. Durán and M. J. Pascual, Non-Linear Optical Properties of Er3+–Yb3+-Doped NaGdF4 Nanostructured Glass Ceramics, Nanomaterials, 2020, 10(7), 1425 CrossRef PubMed.
  31. M. Secu and C. E. Secu, Up-conversion luminescence of Er3+/Yb3+ co-doped LiYF4 nanocrystals in sol-gel derived oxyfluoride glass-ceramics, J. Non-Cryst. Solids, 2015, 42678–42682 Search PubMed.
  32. A. de Pablos-Martín, J. Méndez-Ramos, J. del-Castillo, A. Durán, V. D. Rodríguez and M. J. Pascual, Crystallization and up-conversion luminescence properties of Er3+/Yb3+-doped NaYF4-based nano-glass-ceramics, J. Eur. Ceram. Soc., 2015, 35(6), 1831–1840 CrossRef.
  33. S. Georgescu, A. M. Voiculescu, C. Matei, C. E. Secu, R. F. Negrea and M. Secu, Ultraviolet and visible up-conversion luminescence of Er3+/Yb3+ co-doped CaF2 nanocrystals in sol–gel derived glass-ceramics, J. Lumin., 2013, 143, 150–156 CrossRef CAS.
  34. M. Pollnau, D. R. Gamelin, S. R. Lüthi and H. U. Güdel, Power dependence of upconversion luminescence in lanthanide and transition-metal-ion systems, Phys. Rev. B: Condens. Matter Mater. Phys., 2000, 61, 3337 CrossRef CAS.
  35. J. C. Boyer and F. C. J. M. van Veggel, Absolute quantum yield measurements of colloidal NaYF4: Er3+, Yb3+ upconverting nanoparticles, Nanoscale, 2010, 2, 1417–1419 RSC.
  36. D. O. Faulkner, S. Petrov, D. D. Perovic, N. P. Kherani and G. A. Ozin, Absolute quantum yields in NaYF4:Er,Yb upconverters – synthesis temperature and power dependence, J. Mater. Chem., 2012, 22, 24330 RSC.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ma00377f

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.