Next Article in Journal
Impact of Nonlinear Chemical Reaction and Melting Heat Transfer on an MHD Nanofluid Flow over a Thin Needle in Porous Media
Next Article in Special Issue
Ag-Sensitized NIR-Emitting Yb3+-Doped Glass-Ceramics
Previous Article in Journal
Path Planning under Force Control in Robotic Polishing of the Complex Curved Surfaces
Previous Article in Special Issue
Chalcogenide–Tellurite Composite Photonic Crystal Fiber: Extreme Non-Linearity Meets Large Birefringence
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sol-Gel Glass-Ceramic Materials Containing CaF2:Eu3+ Fluoride Nanocrystals for Reddish-Orange Photoluminescence Applications

by
Natalia Pawlik
1,*,
Barbara Szpikowska-Sroka
1,
Tomasz Goryczka
2 and
Wojciech A. Pisarski
1,*
1
Institute of Chemistry, University of Silesia, 40-007 Katowice, Poland
2
Institute of Materials Engineering, University of Silesia, 41-500 Chorzów, Poland
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2019, 9(24), 5490; https://doi.org/10.3390/app9245490
Submission received: 14 November 2019 / Revised: 5 December 2019 / Accepted: 9 December 2019 / Published: 13 December 2019
(This article belongs to the Special Issue Photonic Glass-Ceramics: Fabrication, Properties and Applications)

Abstract

:
CaF2:Eu3+ glass-ceramic sol-gel materials have been examined for reddish-orange photoluminescence applications. The transformation from precursor xerogels to glass-ceramic materials with dispersed fluoride nanocrystals was verified using several experimental methods: differential scanning calorimetry (DSC), thermogravimetric analysis (TG), X-ray diffraction (XRD), transmission electron microscopy (TEM), infrared spectroscopy (IR-ATR), energy dispersive X-ray spectroscopy (EDS) and photoluminescence measurements. Based on luminescence spectra and their decays, the optical behavior of Eu3+ ions in fabricated glass-ceramics were characterized and compared to those of precursor xerogels. In particular, the determined luminescence lifetime of the 5D0 excited state of Eu3+ ions in nanocrystalline CaF2:Eu3+ glass-ceramic materials is significantly prolonged in comparison with prepared xerogels. The integrated intensities of emission bands associated to the 5D07F2 electric-dipole transition (ED) and the 5D07F1 magnetic-dipole transition (MD) are changed drastically during controlled ceramization process of xerogels. This implies the efficient migration of Eu3+ ions from amorphous silicate sol-gel network into low-phonon energy CaF2 nanocrystals.

1. Introduction

Transparent glass-ceramics (GCs) containing fluoride nanocrystals fabricated using high-temperature melt-quenching or low-temperature sol-gel method are suitable modern materials for lasers, optical waveguides, solid-state lighting and numerous photonic applications [1,2,3,4,5]. Depending on chemical composition and technological conditions, nano- or micro-crystals are usually well-formed during temperature-controlled crystallization of precursor glasses or xerogels. The heat-treatment process often introduces transformation from amorphous systems to transparent oxyfluoride glass-ceramic materials and rare earths play the role as optically active ions. The coordination sphere around rare earth ions changes drastically during this structural transformation, giving important contribution to the luminescence characteristics.
Among binary and ternary rare-earth fluoride nanophosphors [6], calcium fluoride CaF2 belongs to the most important and perspective nanoparticles, which could be successfully formed during the heat treatment process. CaF2 fluoride nanocrystals were well formed from Na2O/K2O/CaO/CaF2/Al2O3/SiO2 glass [7] and their mean crystallite sizes were in the range from 8 to 10.4 nm and did not get larger with time or increasing the annealing temperature. However, synthesis, structure and properties of transparent glass-ceramics containing CaF2 nanocrystals depended critically on treating temperature and glass composition, where SiO2 was substituted by CaO/CaF2 [8]. A special attention has been paid to crystallization processes and spectroscopic properties of erbium-doped transparent glass ceramics containing CaF2 nanocrystals [9,10,11]. In this CaF2:Er3+ glass-ceramic system, near-IR luminescence at 1.53 µm [12,13] and up-conversion processes [14,15] of Er3+ ions have been examined in detail. The results for the crystallization processes and fluorescence properties of Nd3+-doped glass-ceramics with CaF2 nanocrystals have been also presented and discussed [16,17]. Recently, several research groups have been focused on glass-ceramic materials containing CaF2 nanocrystals for numerous photonic applications. Seo et al. [18] suggest that CaF2:Dy3+ nano-glass-ceramics are promising for white light generation. Moreover, the oxyfluoride glass-ceramics containing CaF2:Sm3+ crystals have become a new fast erasable dosimetric detector material for micro-beam radiation cancer therapy applications at the Canadian Synchrotron [19].
In this work, oxyfluoride silicate xerogels containing europium ions were heat-treated at 350 °C to obtain transparent glass-ceramics containing crystalline fluoride nano-phase CaF2. Luminescence spectra of Eu3+ ions and their decays in transparent glass-ceramic samples have been studied and compared to precursor silicate xerogels. Our previously published work indicates that sol-gel glass-ceramics containing SrF2:Eu3+ fluoride nanocrystals are attractive materials for reddish-orange luminescence applications [20].

2. Materials and Methods

The reagents used during proposed sol-gel synthesis were from the Aldrich Chemical Company and contained analytical purity. Deionized water was taken from an Elix 3 system (Millipore, Molsheim, France).
The Eu3+-doped xerogels with nominal composition (in molar ratio): TEOS:C2H5OH:H2O:CH3COOH = 1:4:10:0.5 (90 wt.%) CF3COOH:Ca(CH3COO)2:Eu(CH3COO)3 = 5:1:0.05 (10 wt.%) were synthesized. In the presented procedure, the mixtures of tetraethoxysilane (TEOS), ethanol, water and acetic acid were put into round-bottom flasks and stirred for 30 min to perform the hydrolysis reaction. Simultaneously, Ca(CH3COO)2 and Eu(CH3COO)3 were dissolved in water and trifluoroacetic acid (TFA). In the next step, the obtained solutions were introduced into hydrolyzed tetraethoxysilane (TEOS) and mixed for another 60 min. After this time, the sols were dried at 35 °C for seven weeks to form transparent and colorless xerogels. To fabricate the glass-ceramics containing CaF2 nanocrystals, the xerogels were heat-treated in a muffle furnace FCF 5 5SHP (Czylok, Jastrzębie-Zdrój, Poland) at 350 °C. The temperature was raised by 10 °C/min until the direct temperature (350 °C) was achieved and the xerogels were heat-treated for 10 h. Afterwards, resulted glass-ceramic materials were cooled down to room temperature in a closed furnace.
Fabricated samples were characterized by a SETARAM Labsys thermal analyzer (SETARAM Instrumentation, Caluire, France) using the thermogravimetric analysis (TG) and differential scanning calorimetry (DSC) method. The DSC curves were acquired with heating rate of 10 °C/min and the curves were registered within temperature range from 40 °C to 500 °C. To verify the crystallization of the CaF2 phase, the X-ray diffraction (XRD) analysis was carried out using an X’Pert Pro diffractometer supplied by PANalytical (Almelo, The Netherlands) with Cu Kα radiation. The TEM microscopy was also used and the CaF2 nanocrystals were observed via a JEOL JEM 3010 electron transmission microscope (JEOL, Tokyo, Japan) operated at 300 kV. To determine the distribution of chemical elements in studied CaF2:Eu3+ GCs, an energy dispersive X-ray spectroscopy (EDS) analysis was also performed using JEOL microscope. To examine the structural changes within silicate network during controlled heat-treatment, the infrared spectroscopy (IR-ATR) spectra were registered using the Nicolet iS50 ATR spectrometer (Thermo Fisher Scientific, Waltham, MA, USA) in the frequency region 500 cm−1–4000 cm−1. The luminescence measurements were carried out using a Horiba Jobin-Yvon FluoroMax-4 spectrofluorimeter (Horiba Jobin Yvon, Longjumeau, France) equipped with 150 W xenon lamp. The spectral resolution was ±0.1 nm. Decay curves were detected with the accuracy of ±2 μs. All structural and photoluminescence measurements were performed at room temperature.

3. Results and Discussion

3.1. Structural and Thermal Characterization of Sol-Gel Materials

Firstly, the IR-ATR technique was used for identification the chemical bonding and functional groups inside prepared sol-gel materials and the results are presented in Figure 1. To verify the structural changes during transformation of xerogels into CaF2:Eu3+ GCs, the infrared spectra were recorded in the 500 cm−1–4000 cm−1 frequency region. The assignment of recorded infrared signals was carried out based on current literature [21,22]. For xerogels obtained after seven weeks from synthesis, the creation of three-dimensional silicate network was confirmed. Indeed, the presence of infrared (IR) signals near 1193 cm−1, 799 cm−1 (Si-O-Si siloxane bridges), 1134 cm−1, 1044 cm−1 and 963 cm−1 (Qn units in SiO4 tetrahedrons: Q4, Q3 and Q2, respectively) indicates that polycondensation reaction occurred. In addition, we identified the presence of vicinal or geminal Si-OH groups (~3664 cm−1), hydrogen-bonded Si-OH moieties (~3398 cm−1) and hydrogen-bonded OH groups (~3233 cm−1) originated from residual organic solvents and water. The IR signals located at 1664 cm−1 (C = O groups vibrations) as well as 1447 cm−1 (C-H vibrations) also confirmed that the porous sol-gel network was filled by liquids. It should be noted that the infrared peak at ~1664 cm−1 frequency region was also assigned to adsorbed water and Si-OH groups. Moreover, the presence of un-decomposed calcium and europium (III) trifluoroacetates was also confirmed (1193 cm−1, 1134 cm−1).
To evaluate the thermal resistance of precursor xerogels and to identify the crystallization temperature, the TG/DSC analysis was performed in a temperature range from 35 °C up to 500 °C (inset of Figure 1). The thermal degradation profile with two distinguishable stages was recorded in temperature intervals: 35 °C–151 °C (1) and 151 °C–348 °C (2). For prepared xerogels, the first stage was recorded as a gentle degradation. An indicated step could be associated with desorption of ethanol and water as well as acetic acid and the accompanying weight-loss was estimated to be 5.82%. Beyond the residual acetic acid (used as a catalyst), such acid was also produced during the chemical reaction between TFA and acetate salts:
Ca(CH3COO)2 + 2CF3COOH → Ca(CF3COO)2 + 2CH3COOH
Eu(CH3COO)3 + 3CF3COOH → Eu(CF3COO)3 + 3CH3COOH
As was shown, CF3COO anions could effectively coordinate RE3+ cations, therefore, the trifluoroacetate ions were chemically bonded. In fact, TFA was introduced as fluorination agent, which allowed for successful precipitation of CaF2 crystal fraction during the heat-treatment of xerogels. Finally, the next TG step for prepared sol-gel samples could be associated with the thermal decomposition of Ca(CF3COO)2 and the formation of CaF2 nanocrystals. The formation of the CaF2 fluoride phase was realized by a homogeneous nucleation through a controlled fluorination, when the Ca-O bond cleaved and the Ca-F bond was formed [23]. The CaF2 crystal phase was obtained during thermal decomposition of Ca(CF3COO)2 as follows:
Ca(CF3COO)2 → CaF2 + CF3CFO + CO2 + CO
The decomposition of Ca(CF3COO)2 was observed in DSC curve as an exothermic peak with maximum at ~318 °C. Generally, such a second degradation step was observed in recorded TG curves as a relatively high weight-loss, which was estimated to 16.18%. Based on presented TG/DSC results, the T = 350 °C was chosen to perform the controlled ceramization process of precursor xerogels in order to precipitate the CaF2 nanocrystals dispersed in silicate sol-gel host. Furthermore, the resultant sol-gel materials were almost completely thermally stable at the indicated temperature level.
As was proven by infrared measurements performed after controlled heat-treatment process, the significant changes in shape of recorded spectra were well-visible. The recorded infrared signals were identified as vibrations that originated from the silicate framework. It was observed that the broad OH-band significantly decreased in intensity, and indicated the evaporation of volatile components (disappearance of maximum located at 3233 cm−1) as well as progressive polycondensation reaction of silicate network (reduction in intensity of maximum at 3390 cm−1). The weak peak at ~1664 cm−1 was also detected and it was associated to residual Si-OH groups. For heat-treated samples, the shoulders located at ~1193 cm−1 and ~1134 cm−1 were related to Si-O-Si siloxane bridges and SiO4 tetrahedrons in Q4 units, respectively. Otherwise, the vibration modes originated from SiO4 tetrahedrons inside Q3 (1051 cm−1) and Q2 units (951 cm−1) were also identified. Indeed, it growth in density from 1.936 g/cm3 before controlled ceramization up to 2.201 g/cm3 for the heat-treated sample.
The influence of controlled heat-treatment process was also evaluated using XRD measurements and the results are depicted in Figure 2a. For xerogels, a broad halo pattern was recorded and it indicated their amorphous nature without long-range order. The diffraction lines were observed after performed controlled ceramization process at 350 °C per 10 h. The XRD patterns are in good accordance with the diffraction lines of regular CaF2 phase from ICDD (The International Centre for Diffraction Data, PDF-2 No. 65-0535) crystallized in the Fm3m space group. The subsequent diffraction lines at 28.2°, 46.9°, 55.6°, 68.6° and 75.8° were identified as (111), (200), (311), (400) and (311) reflexes of CaF2 phase, respectively. Observed broadening of diffraction lines clearly indicates that CaF2 phase crystallized in nanometric range and the average crystals size was estimated using the Scherrer formula:
D = K λ β cos θ
in which D is related to the crystal size, K is a constant value (for our calculations it was taken K = 1), λ is the X-ray wavelength, β is a half width of analyzed diffraction peak and θ is the diffraction angle. The mean value of CaF2 nanocrystals was equaled to 11.7 nm ±1.0 nm. Due to similar ionic radii of Ca2+ (1.00 Å) [24] and Eu3+ (1.07 Å) [25], Ca2+ cations in CaF2 crystal lattice could be effectively substituted by trivalent Eu3+ ions and therefore, a very slight shift of diffraction lines was identified.
Figure 2b presents TEM image of fabricated glass-ceramics. The size of CaF2 nanocrystals was consistent with average crystal size estimated from Scherrer equation. The CaF2 nanocrystals with comparable size (about 10 nm) were obtained by Zhou et al. [26] in GCs fabricated during ceramization of precursor glasses with composition (45SiO2-25CaF2-20Al2O3-10CaO):4%EuF3 (treatment conditions: 700 °C/4 h). In addition, to verify the quantitative distribution of chemical elements in fabricated glass-ceramics, we carried out the analysis from the selected sample area containing CaF2 nanocrystal using energy dispersive X-ray spectroscopy, EDS. Generally, the content of Ca and F in CaF2 nanocrystal were estimated to 5 wt.% and 6 wt.%, respectively; and the content of Si and O from silicate sol-gel matrix were estimated to 41 wt.% and 48 wt.%, respectively. Since Eu3+ ions were introduced during sol-gel synthesis as a dopant, their concentration was below the quantification limit and determination their quantitative distribution between CaF2 nanocrystals and a silicate host was not possible. Moreover, according to another data presented in literature, we proposed the lowest treatment temperature (350 °C) to the fabrication of CaF2:Eu3+ glass-ceramic materials (550 °C–570 °C [27], 620 °C–680 °C [28], 660 °C [29], 800 °C [30]). Nowadays, the low-temperature processes are preferable and therefore, the formation of CaF2 nanocrystals at 350 °C seems to be highly desirable.
Based on weight-loss during sol-gel transformation from initial liquid sol to xerogel and taking the results from TG/DSC analysis into account, we estimated the amount of CaF2 crystal fraction in the prepared GC sample. In general, it was found that during successive evaporation of volatile components (water and organic compounds used during synthesis) and progressive polycondensation reaction of silicate network, the remaining mass of fabricated xerogel was evaluated on 16.7 wt.% of the initial sol weight. Based on the TG curve shown in Figure 1, the total weight loss during heat-treatment could be estimated to about 22.0 wt.%. The second degradation step was strictly associated with thermal decomposition of Ca(CF3COO)2 and the indicated weight-loss (16.18 wt.%) should be related with evaporation of CF3CFO, CO2 as well as CO. Therefore, based on above results and taking the stoichiometry of thermal degradation reaction of Ca(CF3COO)2 into account, the amount of CaF2 crystal fraction was estimated to about 8 wt.% (simultaneously, the estimated mass of silicate sol-gel host is about 92 wt.%). Moreover, based on performed calculations, it was assumed that thermal degradation should be complete (it should be also noted that the amount of introduced Ca(CH3COO)2 salt into reaction system during sol-gel synthesis was only 2.3 wt.%).

3.2. Luminescence Behavior of Fabricated Sol-Gel Materials

Figure 3 shows the photoluminescence excitation spectra (PLE) of prepared xerogels and CaF2:Eu3+ glass-ceramics. The spectra were monitored at λem = 611 nm wavelength corresponding to the 5D07F2 optical transition of Eu3+ ions. The recorded excitation lines were identified as the intra-configurational electronic transitions inside 4f6 manifold of Eu3+ optically active ions. The bands were assigned to transitions from the 7F0 ground level to the excited states: 5D4 (363 nm), 5GJ, 5L7 (372 nm–389 nm), 5L6 (395 nm–xerogel, 394 nm/398 nm–CaF2:Eu3+ GCs) and 5D2 (465 nm–xerogel, 464 nm–CaF2:Eu3+ GCs). In addition, the weak bands according to the 7F05D3 (416 nm) and 7F05D1 (525 nm) transitions were recorded. Since, the 7F1 level is thermally populated at room temperature, a satellite line corresponding to the 7F15D1 (535 nm) was also observed.
It was observed that for xerogels the 7F05L6 excitation band has only one maximum, meanwhile for prepared CaF2:Eu3+ GC samples such a line is a double peak with two splitted components located at 394 nm and 398 nm. Since Eu3+ ions could be distributed between silicate sol-gel host and CaF2 nanocrystals in fabricated GCs, such splitting into two components could indicate that Eu3+ ions are residing in two different frameworks, as was suggested by A.C. Yanes for sol-gel LaF3:Eu3+ GCs [31]. Thus, 394 nm component could be associated with their location within silicate network, meanwhile 398 nm component could be related with location of Eu3+ in fluoride nanocrystals. Based on paper by J. Pan et al. [32], the excitation spectra of pure CaF2:Eu3+ crystal phase revealed only one component of the 7F05L6 band. Therefore, it may confirm the hypothesis that visible splitting of indicated excitation band could be related with distribution of Eu3+ ions between CaF2 nanocrystals and silicate sol-gel host. Since the most intense excitation line corresponded to the 7F05L6 transition for fabricated sol-gel samples, the appropriate wavelengths were used to perform the emission measurements.
The photoluminescence spectrum (PL) registered for precursor xerogels is shown in Figure 4. Generally, the recorded spectrum consisted from five emission bands assigned to the transitions from the 5D0 excited state into the 7FJ levels: 7F0 (578 nm), 7F1 (591 nm), 7F2 (616 nm), 7F3 (649 nm) as well as 7F4 (697 nm). Since the relative intensities of Eu3+ emission bands were strictly dependent on symmetry in their local framework, the emission spectra gave valuable information about the nearest framework around them in host matrix [32]. It was observed that for silicate xerogels an emission band assigned to the 5D07F2 electric-dipole transition was more intense compared to a line assigned to the 5D07F1 magnetic-dipole transition. The 5D07F2 transition was very sensitive to the symmetry in the local vicinity around Eu3+ ions and it was hypersensitive in nature. Conversely, an intensity of the 5D07F1 band was rather independent on the symmetry in nearest surrounding of Eu3+ ions. Hence, the ratio between the 5D07F2 and the 5D07F1 emission intensities—well-known as R/O—can play the role as a useful tool for estimating the symmetry in which Eu3+ ions are located [33,34]. The R/O-ratio value calculated for prepared xerogels was estimated to be 4.13.
Furthermore, as was shown in the inset of Figure 4, the luminescence decay curve registered for precursor xerogels was well-fitted to the mono-exponential function and the estimated luminescence lifetime of the 5D0 state was estimated to be τ = 0.46 ms. The short luminescence lifetime was strictly related with the structure of fabricated xerogels. The luminescence from the 5D0 state was quenched by numerous OH groups in local framework of Eu3+ ions and such groups originated from silanol Si-OH moieties (3664 cm−1, 3398 cm−1) as well as residual water and organic solvents (3233 cm−1) inside the porous silicate network. Since the 5D07F6 energy gap of Eu3+ ions was equal to ΔE = 12,500 cm−1 [35], only about four OH phonons were required to promote a non-radiative deactivation of the 5D0 excited level. Since CF3COO anions were in the coordination sphere around Eu3+ cations, it should be noted that such non-radiative relaxation from the 5D0 state could be also caused by C = O groups (1664 cm−1, eight phonons) and the C-F bond (1193 cm−1, ten phonons). However, among identified functional groups, we expected that OH moieties could play a major role in non-radiative deactivation. In general, the non-radiative rates of intra-configurational 4fn-4fn transitions of rare earths were exponentially dependent on the energy gap (ΔE) and the phonon energy in their nearest surrounding (ħω). Such correlation is well-known as the energy gap law and it states that an increase in the non-radiative decay rate is assisted by a decreasing number of phonons needed to cover the energy gap, ΔE. Hence, if there is a functional group with high vibrational energy in a local framework of Eu3+ ions, the probability of non-radiative multiphonon relaxation increases. The phonons with maximum energy in a host are also usually called as effective phonons (ħωmax) [36,37]. The influence of different types of functional groups in Eu3+-doped silicate glasses on multiphonon relaxation was discussed in work by A. Herrmann et al. [38]. The rates of multiphonon relaxation, kNR, via OH modes (3750 cm−1, kNR = 1.4·× 10−2 s−1) and Si-O vibrations (1250 cm−1, kNR = 9.0·× 10−13 s−1) clearly indicates that OH groups with maximum phonon energy in studied glasses play a major role in quenching the luminescence from the 5D0 state. Moreover, the luminescence lifetime is not dependent only on phonon energy in the nearest framework around the optically active ion, but also on another factors, like symmetry. In general, the lower the symmetry is in the nearest vicinity of Eu3+ ion, the more allowed are the forbidden f-f transitions. Therefore, taking the symmetry aspect into account, the probability of radiative relaxation is relatively high, which results in a short luminescence lifetime.
As was presented in Figure 5, the emission spectrum of CaF2:Eu3+ GCs also consisted of the characteristic bands of Eu3+ ions corresponding to the intra-configurational 5D07FJ transitions within the 4f6 manifold. Such bands were recorded at the following wavelengths: 577 nm (J = 0), 589 nm (J = 1), 613 nm (J = 2), 648 nm (J = 3), 683 nm/689 nm/699 nm (J = 4). Compared to xerogels, a significant growth in intensity of the orange 5D07F1 band was observed and it was accompanied by eight-fold decrease in the R/O-ratio value (from 4.13 to 0.51). The indicated decline in the R/O-ratio value clearly points to the change in the symmetry in the nearest vicinity around dopant ions and the nature of the bonding character between Eu3+ ions and their nearest surrounding covalent to become more ionic. Obviously, this is related with the partial incorporation of optically active ions into the crystallized CaF2 fluoride phase.
The 7FJ energy levels of Eu3+ ions in crystal lattice could split and the number of individual sublevels depends on the J number and the site symmetry [33]. According to paper by Brown et al. [39], Eu3+ ions are located in C4v symmetry sites in the CaF2 crystal lattice, despite the fact that Ca2+ cations are located in Oh symmetry sites. This effect could be explained by charge compensation, when divalent Ca2+ cations in CaF2 crystal lattice are substituted by trivalent Eu3+ ions. Then, due to prevent the non-equilibrium charge in the crystal lattice, cation vacancies could be also formed and some fluorine anions could occupy the interstitial positions. If the Eu3+ ion is located in the C4v site in the CaF2 crystal lattice, the individual 5D07FJ (J = 0–4) band should split into one (J = 0), two (J = 1), four (J = 2), five (J = 3) and seven (J = 4) components. However, the distinguishing of individual components was quite difficult for fabricated glass-ceramic samples, because Eu3+ ions are distributed both in amorphous silicate sol-gel network as well as CaF2 nanocrystals. Hence, we assumed that the influence of the crystal field is largely masked.
The photoluminescence decay curve of the 5D0 state of Eu3+ ions recorded for fabricated glass-ceramics was presented in the inset of Figure 5. The decay curve was well-fitted to double-exponential function and therefore, it was distinguished two different decay components: fast (τ1 = 1.63 ms) and slow (τ2 = 12.84 ms). The double-exponential decay clearly indicates that two different decay channels are involved in the total decay process from the 5D0 excited state of Eu3+. The contribution of fast and slow components could be determined using fitting constants in the following equation:
τ n , % = A n A n + A m · 100 %
Since the A1 and A2 fitting constants were close to 1.9408·× 106 and 2.7539·× 106, respectively, a percentage contribution of τ1 was equal to 41.34% and the contribution of τ2 was 58.66%. The first of the luminescence lifetime, τ1, was associated with Eu3+ ions located in the silicate sol-gel host consisting of Q3 units of SiO4 tetrahedrons (~1051 cm−1) and residual Si-OH groups (3390 cm−1, 3656 cm−1). In this way, to cover the energy gap in silicate host, four Si-OH groups or twelve Q3 units were needed. The second luminescence lifetime component, τ2, was related with the remaining part of Eu3+ ions which were successfully incorporated into CaF2 nanocrystals during the controlled heat-treatment process. Indeed, due to low-phonon energy of the CaF2 crystal lattice (~466 cm−1 [40]), about 27 Ca-F phonons were needed to cover the 5D07F6 energy gap. The probability of non-radiative relaxation from the 5D0 state was greater within the silicate framework than in the CaF2 crystal lattice due to higher-phonon energies. Moreover, the probability of radiative relaxation was lower in more symmetric CaF2 than in asymmetric sol-gel host. Therefore, the τ2 components were significantly prolonged compared to τ1. Moreover, we suggested that the tendency of the Eu3+ ions to migrate into CaF2 could be promoted by a similar ionic radii of Ca2+ (1.00 Å) and Eu3+ (1.07 Å). It should be also noted that the τ1 luminescence lifetime component of fabricated CaF2:Eu3+ GCs was prolonged compared with the τ lifetime estimated for xerogel. Both of the presented lifetimes were related with Eu3+ ions, which were dispersed inside the silicate sol-gel host. To explain the differences in luminescence lifetime values, we used the results from the IR-ATR spectroscopy. As was evidenced by IR measurements, the huge amounts of OH groups with high vibrational energies (above 3000 cm−1) were identified in the structure of fabricated xerogels. Therefore, such numerous OH groups in the nearest vicinity of Eu3+ ions were mainly responsible for non-radiative quenching of luminescence from the 5D0 excited state, which resulted in a relatively short lifetime. Simultaneously, the infrared measurements carried out for CaF2:Eu3+ glass-ceramic indicated a strong reduction of OH groups, hence, the amount of effective ‘quenchers’ in the nearest surrounding of Eu3+ ions in the silicate sol-gel network was smaller than in xerogels. In consequence, the τ1 luminescence lifetime was prolonged due to smaller amounts of OH quenchers.

4. Conclusions

Transparent glass-ceramic materials containing CaF2 fluoride nanocrystals were prepared by a low-temperature sol-gel route. Thermal decomposition of Ca(CF3COO)2 in xerogels was confirmed by using TG/DSC methods. The formation of CaF2 nanocrystals during the heat-treatment process was confirmed by X-ray diffraction measurements and TEM microscopy. The structural changes in the sol-gel silica-network were verified by the IR-ATR spectroscopy. The systematic investigations demonstrated that the Eu3+:CaF2 fluoride nanocrystals dispersed in sol-gel glass-ceramic materials showed the prolonged luminescence lifetimes of the 5D0 excited state of Eu3+ ions. The significant changes in calculated R/O-ratio values and in luminescence decay profiles clearly indicated the successful entering of Eu3+ ions into CaF2 nanocrystals during the controlled ceramization process at 350 °C. It suggests that sol-gel glass-ceramic materials containing CaF2:Eu3+ fluoride nanocrystals are promising candidates for reddish-orange photoluminescence applications.

Author Contributions

N.P., B.S.-S., T.G. and W.A.P. conceived and designed the experiments; N.P., B.S.-S. and T.G. performed the experiments; N.P. and W.A.P. analyzed the data; N.P., B.S.-S., T.G. and W.A.P. contributed reagents/materials/analysis tools; N.P. and W.A.P. wrote the paper.

Funding

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

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Beall, G.H.; Pinckney, L.R. Nanophase Glass-Ceramics. J. Am. Ceram. Soc. 1999, 82, 5–16. [Google Scholar] [CrossRef]
  2. Ferrari, M.; Righini, G.C. Glass-ceramic materials for guided-wave optics. Int. J. Appl. Glass Sci. 2015, 6, 240–248. [Google Scholar] [CrossRef]
  3. Fujita, S.; Tanabe, S. Glass-ceramics and solid-state lighting. Int. J. Appl. Glass Sci. 2015, 6, 356–363. [Google Scholar] [CrossRef]
  4. Berneschi, S.; Soria, S.; Righini, G.C.; Alombert-Goget, G.; Chiappini, A.; Chiasera, A.; Jestin, Y.; Ferrari, M.; Guddala, S.; Moser, E.; et al. Rare-earth-activated glass-ceramic waveguides. Opt. Mater. 2010, 32, 1644–1647. [Google Scholar] [CrossRef]
  5. Gorni, G.; Velázquez, J.J.; Mosa, J.; Balda, R.; Fernández, J.; Durán, A.; Castro, Y. Transparent glass-ceramics produced by sol-gel: A suitable alternative for photonic materials. Materials 2018, 11, 212. [Google Scholar] [CrossRef] [Green Version]
  6. 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]
  7. Rüssel, C. Nanocrystallization of CaF2 from Na2O/K2O/CaO/CaF2/Al2O3/SiO2 glasses. Chem. Mater. 2005, 17, 5843–5847. [Google Scholar] [CrossRef]
  8. Wang, Z.; Cheng, L. Effect of substitution of SiO2 by CaO/CaF2 on structure and synthesis of transparent glass-ceramics containing CaF2 nanocrystals. J. Mater. Sci. 2015, 50, 4066–4074. [Google Scholar] [CrossRef]
  9. Chen, D.; Wang, Y.; Yu, Y.; Ma, E.; Hu, Z. Spectroscopic properties of Er3+ ions in transparent oxyfluoride glass-ceramics containing CaF2 nano-crystals. J. Phys. Condens. Matter 2005, 17, 6545–6557. [Google Scholar] [CrossRef]
  10. Qiao, X.; Fan, X.; Wang, J.; Wang, M. Luminescence behavior of Er3+ ions in glass-ceramics containing CaF2 nanocrystals. J. Non-Cryst. Solids 2005, 351, 357–363. [Google Scholar] [CrossRef]
  11. Hu, Z.; Wang, Y.; Ma, E.; Bao, F.; Yu, Y.; Chen, D. Crystallization and spectroscopic properties investigations of Er3+ doped transparent glass ceramics containing CaF2. Mater. Res. Bull. 2006, 41, 217–224. [Google Scholar] [CrossRef]
  12. Chen, D.; Wang, Y.; Yu, Y.; Ma, E.; Bao, F.; Hu, Z.; Cheng, Y. Luminescence at 1.53 µm for a new Er3+-doped transparent oxyfluoride glass ceramic. Mater. Res. Bull. 2006, 41, 1112–1117. [Google Scholar] [CrossRef]
  13. Chen, D.; Wang, Y.; Yu, Y.; Ma, E. Improvement of Er3+ emissions in oxyfluoride glass ceramic nano-composite by thermal treatment. J. Solid State Chem. 2006, 179, 1445–1452. [Google Scholar] [CrossRef]
  14. Chen, D.; Wang, Y.; Yu, Y.; Ma, E.; Bao, F.; Hu, Z.; Cheng, Y. Influences of Er3+ content on structure and upconversion emission of oxyfluoride glass ceramics containing CaF2 nanocrystals. Mater. Chem. Phys. 2006, 95, 264–269. [Google Scholar] [CrossRef]
  15. Kishi, Y.; Tanabe, S. Infrared-to-visible upconversion of rare-earth doped glass ceramics containing CaF2 crystals. J. Alloys Compd. 2006, 408–412, 842–844. [Google Scholar] [CrossRef]
  16. Chen, D.; Wang, Y.; Yu, Y.; Hu, Z. Crystallization and fluorescence properties of Nd3+-doped transparent oxyfluoride glass ceramics. Mater. Sci. Eng. B 2005, 123, 1–6. [Google Scholar] [CrossRef]
  17. Chen, D.; Wang, Y.; Yu, Y.; Ma, E.; Liu, F. Fluorescence and Judd-Ofelt analysis of Nd3+ ions in oxyfluoride glass ceramics containing CaF2 nanocrystals. J. Phys. Chem. Solids 2007, 68, 193–200. [Google Scholar] [CrossRef]
  18. 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]
  19. Okada, G.; Ueda, J.; Tanabe, S.; Belev, G.; Wysokinski, T.; Chapman, D.; Tonchev, D.; Kasap, S. Samarium-doped oxyfluoride glass-ceramic as a new fast erasable dosimetric detector material for microbeam radiation cancer therapy applications at the Canadian Synchrotron. J. Am. Ceram. Soc. 2014, 97, 2147–2153. [Google Scholar] [CrossRef]
  20. Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarski, W.A. Photoluminescence investigation of sol-gel glass-ceramic materials containing SrF2:Eu3+ nanocrystals. J. Alloys Compd. 2019, 810, 151935. [Google Scholar] [CrossRef]
  21. Innocenzi, P. Infrared spectroscopy of sol-gel derived silica-based films: A spectra-microstructure overview. J. Non Cryst. Solids 2003, 316, 309–319. [Google Scholar] [CrossRef]
  22. He, S.; Huang, D.; Bi, H.; Li, Z.; Yang, H.; Cheng, X. Synthesis and characterization of silica aerogels dried under ambient pressure bed on water glass. J. Non Cryst. Solids 2015, 410, 58–64. [Google Scholar] [CrossRef]
  23. 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]
  24. Withers, S.H.; Peale, R.E.; Schulte, A.F.; Braunstein, G.; Beck, K.M.; Hess, W.P.; Reeder, R.J. Broad distribution of crystal-field environments for Nd3+ in calcite. Phys. Chem. Miner. 2003, 30, 440–448. [Google Scholar] [CrossRef]
  25. Qin, D.; Tang, W. Energy transfer and multicolor emission in single-phase Na5Ln(WO4)4-z(MoO4)4:Tb3+,Eu3+ (Ln = La, Y, Gd) phosphors. RSC Adv. 2016, 6, 45376–45385. [Google Scholar] [CrossRef]
  26. Zhou, B.E.; E, C.Q.; Bu, Y.Y.; Meng, L.; Yan, X.H.; Wang, X.F. Temperature-controlled downconversion luminescence behavior of Eu3+-doped transparent MF2 (M = Ba, Ca, Sr) glass ceramics. Luminescence 2017, 32, 195–200. [Google Scholar] [CrossRef] [PubMed]
  27. Deng, W.; Cheng, J.-S. New transparent glass-ceramics containing large grain Eu3+:CaF2 nanocrystals. Mater. Lett. 2012, 73, 112–114. [Google Scholar] [CrossRef]
  28. Li, Y.; Zhao, L.; Zhang, Y.; Ma, J. Preparation and luminescence properties of Eu3+ doped oxyfluoride borosilicate glass ceramics. J. Rare Earth 2012, 31, 1195–1198. [Google Scholar] [CrossRef]
  29. Qiao, X.; Luo, Q.; Fan, X.; Wang, M. Local vibration around rare earth ions in alkaline earth fluorosilicate transparent glass and glass ceramics using Eu3+ probe. J. Rare Earth. 2008, 26, 883–888. [Google Scholar] [CrossRef]
  30. Secu, M.; Secu, C.E.; Ghica, C. Eu3+-doped CaF2 nanocrystals in sol-gel derived glass-ceramics. Opt. Mater. 2011, 33, 613–617. [Google Scholar] [CrossRef]
  31. Yanes, A.C.; del-Castillo, J.; Méndez-Ramos, J.; Rodríguez, V.D.; Torres, M.E.; Arbiol, T. Luminescence and structural characterization of transparent nanostructures Eu3+-doped LaF3-SiO2 glass-ceramics prepared by sol-gel method. Opt. Mater. 2007, 29, 999–1003. [Google Scholar] [CrossRef]
  32. Pan, J.; Zhong, K.; Zhang, Z.; Chen, W.; Lin, Y.; Wang, Q.; Li, L.; Yu, Y. Using CaF2:Eu3+ powder as luminescent probe to detect Cr2O72- ions: A new application on the environmental conservation of an old optical material. Opt. Mater. Express 2018, 8, 2782–2794. [Google Scholar] [CrossRef]
  33. Binnemans, K. Interpretation of europium (III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef] [Green Version]
  34. Gökçe, M.; Şentürk, U.; Uslu, D.K.; Burgaz, G.; Şahin, Y.; Gökçe, A.G. Investigation of europium concentration dependence on the luminescent properties of borogermanate glasses. J. Lumin. 2017, 192, 263–268. [Google Scholar] [CrossRef]
  35. Sołtys, M.; Janek, J.; Żur, L.; Pisarska, J.; Pisarski, W.A. Compositional-dependent europium-doped lead phosphate glasses and their spectroscopic properties. Opt. Mater. 2015, 40, 91–96. [Google Scholar] [CrossRef]
  36. Hehlen, M.P.; Cockroft, N.J.; Gosnell, T.R. Spectroscopic properties of Er3+- and Yb3+-doped soda-lime silicate and aluminosilicate glasses. Phys. Rev. B 1997, 56, 9302–9318. [Google Scholar] [CrossRef]
  37. Van Dijk, J.M.F.; Schuurmans, M.F.H. On the nonradiative and radiative decay rates and a modified exponential energy gap law for 4f-4f transitions in rare-earth ions. J. Chem. Phys. 1983, 78, 5317–5323. [Google Scholar] [CrossRef]
  38. Herrmann, A.; Rüssel, C. New Aluminosilicate Glasses as High-Power Laser Materials. Int. J. Appl. Glass Sci. 2015, 6, 210–219. [Google Scholar] [CrossRef]
  39. Brown, M.R.; Roots, K.G.; Williams, J.M. Experiments on Er3+ in SrF2. II. Concentration Dependence of Site Symmetry. J. Chem. Phys. 1969, 50, 891–899. [Google Scholar] [CrossRef]
  40. Ritter, B.; Haida, P.; Fink, F.; Krahl, T.; Gawlitza, K.; Rurack, K.; Scholz, G.; Kemnitz, E. Novel and easy access to highly luminescent Eu and Tb doped ultra-small CaF2, SrF2 and BaF2 nanoparticles—Structure and luminescence. Dalton Trans. 2017, 46, 2925–2936. [Google Scholar] [CrossRef]
Figure 1. Infrared spectroscopy (IR-ATR) spectra of sol-gel materials before and after controlled heat-treatment. Inset shows the thermogravimetric analysis (TG)/differential scanning calorimetry (DSC) curves recorded for prepared xerogels.
Figure 1. Infrared spectroscopy (IR-ATR) spectra of sol-gel materials before and after controlled heat-treatment. Inset shows the thermogravimetric analysis (TG)/differential scanning calorimetry (DSC) curves recorded for prepared xerogels.
Applsci 09 05490 g001
Figure 2. (a) X-ray diffraction (XRD) patterns for xerogel and CaF2:Eu3+ glass-ceramic fabricated during controlled heat-treatment at 350 °C. Inset shows the image of glass-ceramic (GC) sample; (b) transmission electron microscopy (TEM) image of CaF2:Eu3+ nanocrystals dispersed within amorpous sol-gel host.
Figure 2. (a) X-ray diffraction (XRD) patterns for xerogel and CaF2:Eu3+ glass-ceramic fabricated during controlled heat-treatment at 350 °C. Inset shows the image of glass-ceramic (GC) sample; (b) transmission electron microscopy (TEM) image of CaF2:Eu3+ nanocrystals dispersed within amorpous sol-gel host.
Applsci 09 05490 g002
Figure 3. Photoluminescence excitation spectra (PLE) spectra registered for fabricated xerogels and CaF2:Eu3+ GCs.
Figure 3. Photoluminescence excitation spectra (PLE) spectra registered for fabricated xerogels and CaF2:Eu3+ GCs.
Applsci 09 05490 g003
Figure 4. Photoluminescence spectrum (PL) spectra recorded for silicate xerogels. Inset presents the luminescence decay curve of the 5D0 state of Eu3+exc = 395 nm, λem = 591 nm).
Figure 4. Photoluminescence spectrum (PL) spectra recorded for silicate xerogels. Inset presents the luminescence decay curve of the 5D0 state of Eu3+exc = 395 nm, λem = 591 nm).
Applsci 09 05490 g004
Figure 5. PL spectra recorded for fabricated CaF2:Eu3+ GCs. The inset presents the luminescence decay curve of the 5D0 state of Eu3+exc = 394 nm, λem = 589 nm).
Figure 5. PL spectra recorded for fabricated CaF2:Eu3+ GCs. The inset presents the luminescence decay curve of the 5D0 state of Eu3+exc = 394 nm, λem = 589 nm).
Applsci 09 05490 g005

Share and Cite

MDPI and ACS Style

Pawlik, N.; Szpikowska-Sroka, B.; Goryczka, T.; Pisarski, W.A. Sol-Gel Glass-Ceramic Materials Containing CaF2:Eu3+ Fluoride Nanocrystals for Reddish-Orange Photoluminescence Applications. Appl. Sci. 2019, 9, 5490. https://doi.org/10.3390/app9245490

AMA Style

Pawlik N, Szpikowska-Sroka B, Goryczka T, Pisarski WA. Sol-Gel Glass-Ceramic Materials Containing CaF2:Eu3+ Fluoride Nanocrystals for Reddish-Orange Photoluminescence Applications. Applied Sciences. 2019; 9(24):5490. https://doi.org/10.3390/app9245490

Chicago/Turabian Style

Pawlik, Natalia, Barbara Szpikowska-Sroka, Tomasz Goryczka, and Wojciech A. Pisarski. 2019. "Sol-Gel Glass-Ceramic Materials Containing CaF2:Eu3+ Fluoride Nanocrystals for Reddish-Orange Photoluminescence Applications" Applied Sciences 9, no. 24: 5490. https://doi.org/10.3390/app9245490

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