Next Article in Journal
Effect of the Ratio of Protein to Water on the Weak Gel Nonlinear Viscoelastic Behavior of Fish Myofibrillar Protein Paste from Alaska Pollock
Previous Article in Journal
Food Gels Based on Polysaccharide and Protein: Preparation, Formation Mechanisms, and Delivery of Bioactive Substances
Previous Article in Special Issue
Effects of Moisture Content and Heat Treatment on the Viscoelasticity and Gelation of Polyacrylonitrile/Dimethylsulfoxide Solutions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels

by
Alan D. Alcantar Mendoza
1,
Antonieta García Murillo
1,*,
Felipe de J. Carrillo Romo
1 and
José Guzmán Mendoza
2
1
Instituto Politécnico Nacional CIITEC, Azcapotzalco, Mexico City 02250, Mexico
2
Instituto Politécnico Nacional CICATA, Legaria, Mexico City 11500, Mexico
*
Author to whom correspondence should be addressed.
Gels 2024, 10(11), 736; https://doi.org/10.3390/gels10110736
Submission received: 8 October 2024 / Revised: 8 November 2024 / Accepted: 11 November 2024 / Published: 13 November 2024
(This article belongs to the Special Issue Gel-Based Materials: Preparations and Characterization (2nd Edition))

Abstract

:
This study compared the chemical, structural, and luminescent properties of xerogel-based ceramic powders (CPs) with those of a new series of crystallized aerogels (CAs) synthesized by the epoxy-assisted sol–gel process. Materials with different proportions of Eu3+ (2, 5, 8, and 10 mol%) were synthesized in Lu2O3 host matrices, as well as a Eu2O3 matrix for comparative purposes. The products were analyzed by infrared spectroscopy (IR), X-ray diffraction (XRD), scanning electron microscopy (SEM) with energy-dispersive spectroscopy (EDS), transmission electron microscopy (TEM), photoluminescence analysis, and by the Brunauer–Emmett–Teller (BET) technique. The results show a band associated with the M-O bond, located at around 575 cm−1. XRD enabled us to check two ensembles: matrices (Lu2O3 or Eu2O3) and doping (Lu2O3:Eu3+) with appropriate chemical compositions featuring C-type crystal structures and intense reflections by the (222) plane, with an interplanar distance of around 0.3 nm. Also, the porous morphology presented by the materials consisted of interconnected particles that formed three-dimensional networks. Finally, emission bands due to the energy transitions (5DJ, where J = 0, 1, 2, and 3) were caused by the Eu3+ ions. The samples doped at 10 mol% showed orange-pink photoluminescence and had the longest disintegration times and greatest quantum yields with respect to the crystallized Eu2O3 aerogel.

1. Introduction

Materials synthesized from rare-earth (RE) elements, commonly used in the form of oxides (REOs), have a wide diversity of applications directly related to their properties. Among these, they have become indispensable due to their fascinating and highly efficient luminescent properties, which arise from transitions of the f orbitals that are shielded by the external s and p orbitals [1,2,3,4]. That is why they can be found widely applied as phosphors used for lighting in ceramics, display screens, security label inks, lasers, fiber optics, night vision devices, biomarkers, nanothermometers, nanoscopy, and the administration of activated drugs [5,6,7,8,9,10,11]. Several well-known, commercial phosphors have been used to meet the needs of these applications; some of the most popular are Y2O3:Eu3+, Y2O2S:Eu3+, BaMgAl10O17:Eu2+, and MgAl11O19:Ce3+ [12]. Commonly, REO phosphors are synthesized by chemical methods such as hydrothermal, capping, cluster-generation, and electrochemical methods [2,13].
However, authors such as Tshikovi et al. [14] have indicated that the sol–gel process is the most effective method for the preparation of luminescent materials as they have impurities in their structuring. Among the rare-earth trivalent ions (RE3+), Eu3+ and Tb3+ ions are commonly used as luminescence-activating dopants. Moreover, the sol–gel process has established itself as an extremely effective method because of its advantages such as synthesis at lower temperatures and the high homogeneity and purity that can be achieved in processing high-end materials [15,16]. In addition, the epoxy-assisted variant has been widely used in recent years to synthesize the gels of metal oxide compositions; among other benefits, it can be doped with RE ions to modify their properties or induce new ones. In this sense, several studies using this variant can be found, involving the production of porous materials such as REO xerogels or REO aerogels [8,9,17,18,19,20], although it should be mentioned that in these works, the luminescent properties of the materials were marginal or completely disregarded.
Furthermore, aerogels are a type of material that, despite having been discovered more than 90 years ago by Samuel Kistler [21], have not yet been fully exploited [22] due to their high production cost, which is why they are researched and synthesized only in very small quantities in laboratories [23,24]. These materials are made up of a three-dimensional network of interconnected particles, which is extremely fragile due to the huge number of “chains” of pores (micro- and mesopores) that it possesses [25,26], although this same porosity is the basis of its physicochemical properties [27]. With the combination of a macroscopic external structure and a highly specific surface area on the nanoscale, these materials have seen greatly increased use in novel technologies in different areas [28], for example, in aerospace technology [29], construction [30], and energy storage [17]. More recently, their potential use in nano biomedicine has been investigated for its application as a drug carrier [22,24,27,31]. Overall, aerogels are very interesting materials but complicated to manipulate and even more difficult to characterize.
As mentioned, the common way to obtain aerogels is through the sol–gel process [32,33,34,35], where, once the gel is obtained, it must be carefully treated to extract all the liquid phases inside. What makes the difference in the porosity of aerogels and xerogels and the pattern of particle dispersion is the type of drying [24,36,37]. Xerogels are made up of a series of “collapsed” particles in the three-dimensional network of interconnected particles that form the gel. When these materials are dried at room temperature and sintered, the result is an agglomeration of material that can be defined as ceramic powders. To prevent the three-dimensional network from collapsing due to surface tension and capillary forces during drying, supercritical CO2 (scCO2) drying is used as the most effective technique for replacing the liquid phase of the gels with a gaseous phase (which, in this case, is air). As a consequence, scCO2 causes only a slight contraction of the original three-dimensional structure of the gel, keeping the three-dimensional network of interconnected particles intact, or “open” [23,36,38,39,40]; scCO2 is especially effective and is applied in the production of aerogels to obtain better stability and durability [26,27].
In general, especially when it comes to aerogels, there is a limited amount of work dedicated to the investigation of porous lutetium sesquioxide (Lu2O3) materials. However, there is a large amount of research related to photoluminescent materials based on Lu2O3:RE3+ systems [41,42,43,44,45,46,47]. Some of the most notable works in this area are reviewed below.
Jia et al. [48] reported their synthesis of luminescent Lu2O3:Ln3+ (Ln = Eu, Er, Yb) microspheres using a precipitation method. After being calcined at 800 °C for 2 h, Ln3+-doped products were able to index into the cubic Lu2O3 phase. The sample Lu2O3:5%Eu3+ especially produced strong red emissions caused by the excitation of the charge transfer band (CTB) at approximately 245 nm and was subsequently found to be dominated by the 5D0-7F2 transition of Eu3+ ions. The decay curve indicated that the lifetime of the Eu3+ ions was 1.21 ms. Similarly, Gao et al. [49] synthesized microspheres using a precipitation method; these were calcined at different temperatures, and it was determined that at 800 °C, cubic phases were obtained with well-defined diffraction peaks, with a d222 distance of 0.300 nm. The abundant luminescence emitted by the microspheres was dominated by the “red” 5D0-7F2 transition (613 nm), which is described as an electric-dipole-allowed transition that is hypersensitive to the environment. Also, Zhao et al. [4] managed to synthesize “3D hierarchical architectures” of europium-doped lutetium oxide using hydrothermal and calcination processes. The phosphors exhibited polycrystalline cubic-type structures, with the strongest red emissions caused by transitions 5D07FJ (J = 0, 1, 2, 3, and 4) with lifetimes of 1.23, 1.49, and 1.19 ms. Later, in a similar study, Popielarski et al. [50] synthesized ball-like Lu2O3:Eu3+ nanoparticles using the hydrothermal process, which, if carefully observed, could be considered to have a coralliferous morphology. These materials emitted photoluminescence due to the 5D07DJ transitions of the Eu3+ ions at the C2 sites. The decay time when doping the Lu2O3:3%Eu was 1.5 ms. More recently, Tang et al. [51], using a two-step precipitation method, synthesized dispersed spherical particles that were annealed at 800 °C for 2 h, resulting in cubic structures associated with the space group Ia3. The Lu2O3:Eu3+ sample showed luminescence due to Eu3+ ion transitions, and the 5 mol% Eu3+ sample had a decay time of 2.04 ms. Finally, Olvera Salazar et al. [16] conducted one of the few experiments where a wide series of Lu2O3 xerogels were synthesized at different doping concentrations of Eu3+. In addition, nanoparticles synthesized by the sol–gel method and heat-treated up to 700 °C for one hour are described as novel and potential antioxidant materials, although, unfortunately, the luminescent properties of these materials were not studied.
There are some commonalities in the findings of these studies, for example, Lu2O3 was found to be an excellent candidate as a host matrix for luminescent lanthanide ions due to its favorable physical properties such as a high melting point, phase stability, and low thermal expansion, among others. Moreover, materials based on the Lu2O3:TR3+ system have extremely attractive optical characteristics stemming from 4f-5d electronic transitions, which leads to their recommendation for applications in areas such as advanced flat displays, high-security technologies, and biological labeling, due to their excellent dispersion and luminescence properties. Studying or proposing new synthesis methods and materials for the Lu2O3:TR3+ system, such as the epoxide-assisted sol–gel method of obtaining aerogels, can reduce the production costs of the system and yield more efficient materials that can compete with current commercial products.

2. Results and Discussion

In this work, through the epoxy-assisted sol–gel process, ceramic powders (CPs) were synthesized from xerogels and crystallized aerogels (CAs) using different proportions of Eu3+ ions (2, 5, 8, and 10 mol%) in Lu2O3 host matrices, as well as in a europium sesquioxide (Eu2O3) matrix, to enable their chemical, structural, and photoluminescent properties to be compared. The labels and the general characteristics of the synthesized samples are shown in Table 1.

2.1. Infrared (IR) Spectroscopy

Figure 1 presents the IR spectra of the CPs and CAs. The results show low signals corresponding to the -COO- interaction around 1500 and 1400 cm−1. This interaction is mainly attributable to atmospheric remains; however, the interaction was not constant and was more noticeable in the CPEu and CAEu samples. The main stretching vibration M-O band, which is associated with the interactions of the REs in the formation of REOs, was detected at a position of 585 cm−1 for the Lu2O3 host sample and at 535 cm−1 for the Eu2O3 matrix sample [49], while for all the doped samples, both the CPs and CAs, it was located at 575 cm−1. This change occurred due to the incorporation of Eu3+ in the doped samples.
This has not been reported in other investigations such as [4,51,52]; they mention that the band was located around 570 cm−1, but it was attributed only to the stretching vibrations of lutetium. Also, it can be supposed that if the concentration of Eu2O3 was increased in the Lu2O3 host matrix, this band would show a greater displacement toward the Eu2O3 position (535 cm−1).

2.2. X-Ray Diffraction (XRD)

The XRD patterns of both the CP and CA materials are shown in Figure 2a,b, where a series of peaks can be seen that correspond with the diffraction charts ICSD 98-004-0471 and ICSD 98-009-6953, for Lu2O3 and Eu2O3, respectively. The plane with the highest intensity (222) indicates high purity and well-defined crystalline structures of type C with an Ia-3 space group, like the one presented in Figure 2c. This is very important for REO luminescent materials because high crystallinity generally means fewer traps and stronger luminescence [48]. Furthermore, unlike what was seen in [8] where rare-earth aerogels were synthesized with a similar methodology, in our work, no secondary phases caused by Cl ions appeared. In particular, the planes suffered a slight displacement in relation to the Lu2O3 host matrix due to the incorporation of the Eu3+ ions that replaced Lu3+ [49].
As in the case of IR spectroscopy, if the concentration were to be increased further, the peaks would tend more toward the peaks in the Eu2O3 matrix. This behavior is due to the incorporation of the europium atoms, which causes deformations in the crystalline structure and results in different average sizes of the crystallites. These effects were calculated with the Debye–Scherrer equation and are presented in Table 2.
Although the sizes varied, there was a tendency to form new crystals on the part of both the CAs and the CPs. This occurred because, at a concentration of 2 mol%, the Lu2O3 crystallites tended to be smaller; at the next concentration (5 mol%) they reached an average size similar to that of the Lu2O3 host. When the Lu2O3 was ±17 nm, they formed with dimensions less than 10 nm (CPLu8 = 9.6 nm and CALu8 = 8.2 nm) to reach their final size in the samples doped at 10 mol%. It was not possible to calculate the average crystallite size for the Eu2O3 matrix. The results of the Rietveld analysis are presented in Table 2, which shows that the interplanar distance from the (222) plane was around 0.3 nm for all the samples.
The smaller crystallite size for the doped samples at 8 mol% could be related to this doping concentration via two possible effects: (1) reduced crystal growth rates and (2) enhanced nucleation sites. With reduced crystal growth rates, europium ions create lattice distortions in the Lu2O3 host, limiting crystal growth and forming larger grains. Similar behavior was reported by Nair et al. [53] and Shakirzyanov et al. [54], where the lattice distortion degree R allowed the quantification of the deviation of Lu-O bond lengths from their ideal values in the undoped structure. The doping-induced effect on REO bond length (L) and the lattice distortion degree (R) parameter were estimated using relations (1) and (2), respectively [55]:
L = a 3 4
R = | L doped L undoped | L doped   ×   100
where a is the lattice parameters obtained from XRD, L doped is the Lu-O bond length in the doped structure, and L undoped is the bond length in undoped Lu-O. Figure 3a,b shows the relationship between crystal size (nm) and lattice distortion degree (R) as a function of mol% Eu3+ for ceramic powders and crystallized aerogels.
When Lu3+ is substituted with Eu3+, a lattice distortion degree is observed that increases with the Eu3+ molar concentration, i.e., ranging from 0.67 to 0.57 for CPLu8 and CALu8. This is due to the different ion radii between Lu3+ and Eu3+. In both samples at 8 mol%, we can observe the atomic-scale distortions that lead to the formation of the smallest crystallite sizes observed for the two systems.
Lanthanide contraction refers to an unexpected decrease in the atomic and ionic radii of this family of elements [56]. It is caused by poor screening of the 6 s electrons from nuclear displacement by the 4f electrons, which results in smaller atoms (from lanthanum to lutetium) due to an increase in the effective nuclear charge to attract each electron. The unit cell parameters of the samples are presented in Table 3. These confirm the successful production of cubic crystal structures (Ia-3), and although the samples have different values, the behavior induced by the lanthanide contraction is interesting. The cell volume would be expected to increase as the amount of Eu3+ doping increases, which happens inversely because the size of the europium atom is larger than the lutetium atom.

2.3. Scanning Electron Microscopy (SEM)

Figure 4 shows micrographs of the CP samples CPLu (a and b) and CPLu10 (d and e). These exhibit a “collapsed” three-dimensional network made up of many agglomerated particles, with irregular shapes and sizes of less than 100 nm. The results of the elemental analysis by EDS for the host matrix sample CPLu (Figure 4c) only show the signals corresponding to the compound Lu2O3, while for the doped sample CPLu10 (Figure 4f), the signals associated with the presence of europium are also seen.
Naturally, the CAs (Figure 5) presented morphological differences. Although many particles with irregular shapes continued to appear, in this case, the microstructure is formed by an “open” interconnection of particles forming three-dimensional networks. In other work, these have been described as coral-shaped structures [18]. This type of particle dispersion is directly related to high porosity [38,57], specifically, the high porosity provided by the chains of pores that form between the particles due to the adequate extraction of CH3CH2OH during scCO2, but in the case of the Lu2O3:Eu3+ system, this type of morphology is reported here for the first time.
For the elemental analysis of the CALu and CALu10 aerogels (Figure 5c,f), the same congruence is maintained as in the CPs. Thus confirming the possibility of including Eu3+ in the composition of Lu2O3 matrices through the epoxide-assisted sol–gel process.

2.4. Transmission Electron Microscopy

The transmission micrographs show similarities to the scanning micrographs presented above. Figure 6 depicts agglomerated particles in the CPLu10 sample (Figure 6a,b), while the particles of CALu10 (Figure 6e,f) are more dispersed. The planes visible in these samples were indexed through Fourier transform SAED analysis to determine the interplanar distances of the areas selected (see the red dotted squares in Figure 6c,g). These presented an average of 2.9 nm (Figure 6d) and 2.99 nm (Figure 6h) for a section of 10 planes corresponding to the CPLu10 and CALu10 samples, respectively.
This result coincides with the results obtained in the Rietveld analysis (see Table 2) and confirms what was found through XRD about the existence of the (222) plane and the formation of high-purity C-type crystalline structures without secondary phases. This all agrees with what is described in [4,49,51].

2.5. Photoluminescent Properties

Figure 7 presents graphics of the CP (Figure 7a,b) and CA (Figure 7c,d) emission and excitation data. It is evident that the Lu2O3 matrix materials do not emit any type of signal on their own and that the photoluminescence is generated by the Eu3+ activator.
In the case of all the doped samples, a wavelength of 612 nm was able to excite the Eu3+ ions. While some slight signals corresponding to f transitions were present, it was the band around 250 nm corresponding to the charge transfer band (CTB) between Eu3+ and O2− that showed the greatest intensity. Therefore, the emission spectra were obtained using an incident wavelength of 250 nm. In this case, bands associated with the energy transitions 5D07F0, 5D07F1, 5D07F2, and 5D07F3 of the Eu3+ ion appeared around 580, 590, 612, and 650 nm, respectively, although it should be noted that in all the photoluminescent samples, the hypersensitive transition of 612 nm predominated due to the proper energy transfer at the Eu(C2) → Eu(S6) sites. This process is described in [58] for the interpretation of europium (III) spectra; it is the same as that seen in other studies [4,44,45,50,51,59] and is schematized in the energy transfer diagram in Figure 8. However, in this case, in both materials (the CPs and CAs), the samples doped at 10 mol% exhibited the most powerful photoluminescence compared to the samples at other doping concentrations, and mainly with the samples in the Eu2O3 matrix. In addition, quenching did not occur in any of the samples, primarily because only low concentrations of the dopant were used [46].
The International Commission on Illumination (CIE, its acronym in French) chromaticity coordinates were calculated. In Figure 9a, the CIE locations of the Eu2O3 matrix samples (CPEu and CAEu) are plotted; these are located in the reddish-orange emission zone. The locations of the samples with powerful photoluminescence (CPLu10 and CALu10) underwent a change in emission color, tending more toward orange-pink. In Table 4, all the values obtained for the CPs and CAs samples are listed, together with the data obtained for the chromaticity coordinates for other red phosphors. As can be seen in Table 4, only the chromaticity coordinates of the Eu2O3 matrices are close to the values of both the coordinates for the commercial phosphor Y2O2S:Eu3+ (0.658, 0.340) and the NTSC-recommended standard (0.67, 0.33). All the values for the doped samples are far from the red emission region, suggesting that the interactions between the Lu2O3 and Eu3+ ions are characterized by a powerful and pure reddish-orange color.
As mentioned, it can be inferred that doping with Eu3+ was the main cause of photoluminescence because the quantum yield from the Eu2O3 matrices (Figure 9b) was very low at only 2% for CPLu and 1% for CALu. However, the samples doped at 10 mol% greatly exceeded these values, with figures of 17% for CPLu10 and 37% for CALu10. Furthermore, the decay curves are well-fitted into a single exponential for the CP samples and into a double exponential for the CA samples.
Furthermore, from the intensity of the emission spectra, it was possible to calculate the symmetry ratio (R) of the 5D0-7F2 electric dipole transitions relative to the 5D0-7F1 magnetic dipole transitions for both the CPs and CAs:
R = I(5D07F2)/I(5D07F1)
The results are shown in Table 4 with the exception of the host matrices (CPLu and CALu), which did not present a luminescence response. The R values remained at around 4 for the doped samples, while for the CPEu and CAEu matrices, the values dropped to 2 and 1.45, respectively. This change in the asymmetric ratio suggests that the Eu3+ ions in the doped samples are localized in a non-symmetric environment in the crystal lattice, while in the Eu2O3 matrices, the Eu3+ is localized in a symmetric environment within the crystal lattice. It is important to note that R provides information about the symmetry environment of the crystallographic sites of the Eu3+ ions. Thus, if the value of R is high, Eu3+ tends to be located at asymmetric or non-centrosymmetric sites; if the value of R is low, Eu3+ tends to be located at high-symmetry sites or centrosymmetric sites. This could be why 10 mol% doping leads to better luminescent properties since, as seen in reports of other work [61,62], it is suggested that the greater the asymmetry of the Eu3+ ions, the stronger the Eu-O covalence, and hence, the greater and more powerful the emission intensity.

2.6. Analysis of Porosity by the Brunauer–Emmett–Teller (BET) Technique

The BET technique by nitrogen adsorption–desorption was used to calculate the pore volumes, the average pore diameters, and the specific surface areas of the CPLu10 and CALu10 samples. As established by Brunauer et al. [64], CPs exhibit a type-IV isotherm with an H4 hysteresis loop (Figure 10a). The surface area results were 5.36 and 5.17 m2·g−1 for CPLu10 and CALu10, respectively. The average pore diameter and volume values shown in Figure 10b for CPLu10 were 49 nm and 0.049 mL·g−1, respectively, while for CALu10, they were 44 nm and 0.015 mL·g−1.
The results suggest that crystallized aerogels have smaller diameters and pore volume sizes than ceramic powders. This could be due to the factors discussed in Section 2.3 on the results of the SEM analysis. In the case of the CPs, after heat treatment, a collapse occurred that caused particle agglomeration, which would mean a loss of porosity where the largest pores (macropores) persist. On the other hand, although the CAs also show a slight compaction from the heat treatment, they present a morphology with the shape of an interconnected three-dimensional structure as a result of the scCO2. The formation of this “open” structure must have resulted from the arrangement of the crystallites, and consequently, a greater number of micro- and mesopores were produced.
As explained by Rouquerol et al. [65], although the BET technique is widely used, it can produce errors when applied to microporous materials and should not really be used. Consequently, validly and reliably determining the porosity of aerogels is still considered a challenge, as they are synthetic materials with the highest porosity ever obtained. Because of this, it is likely that the adsorbate molecules underwent packing, which impeded gas permeability throughout the porous structure. Other factors that could affect the results are the type of pore and whether the material was adsorbent or not.
For everything described above, CALu10 was the most attractive and efficient sample, presenting some of the essential characteristics of a sensor, as mentioned in [1], such as high sensitivity, stability, good response times, and ease of synthesis. A good addition to this list would follow from testing whether these types of REO aerogels are non-toxic since, considering their nanometric properties, especially their porosity, the interaction of the surface area with the biological systems is still an area in which much exploration is needed. However, if the porosity of luminescent aerogels can be used to carry an active compound in a controlled manner, this class of materials could open the horizon for a new generation of products in nano biomedicine.

3. Conclusions

A novel series of crystalline materials consisting of a Lu2O3 system doped with Eu3+ and possessing photoluminescent properties were synthesized successfully by the epoxide-assisted sol–gel process. This method showed excellent results for the formation of stable gels, and subsequently, xerogels and aerogels, which can be crystallized so that materials of high purity and crystallinity are produced (in the form of ceramic powders and crystallized aerogels). Due to the similarity that exists between the CPs and CAs samples in IR spectroscopy, XRD, and EDS characterizations, it was determined that there are no chemical changes when modifying the molar% of Eu3+ and the difference is mainly reflected in the morphology due to the type of drying and the luminescent response of each system. Despite this, a correlation noted here that has not been reported in previous work concerns the size of the crystallites and how their formation was affected by increasing the concentration of Eu2O3 in the chemical composition of the materials. This also caused changes in the photoluminescence response, indicating that, in combination with the “open” and dispersed morphology presented by crystallized aerogels, it is possible to obtain a material with powerful and efficient photoluminescent properties, especially with doping at 10 mol%. Aerogels with photoluminescent properties have advantages over common biomarkers and drug carriers; therefore, some of the materials proposed in this project could have analytical applications in biomedical and related technologies; however, a barrier—or possibly a window of opportunity for researchers—can be noted in the lack of information relating to the toxicological damage that RE and aerogels can cause when they encounter biological systems.

4. Materials and Methods

To synthesize the gels, lutetium sesquioxide (Sigma-Aldrich, St. Louis, MA, USA, 99.9%) and europium sesquioxide (Sigma-Aldrich, St. Louis, MA, USA, 99.9%) were used. Hydrochloric acid (HCl, Fermont, México, 97%) was used to form the metallic salt. Ethanol (CH3CH2OH, J. T. Baker, BirthKansas City, MO, USA, 99.9%) served as the solvent, the gelation initiating agent was propylene oxide (CH3CHCH2O, Sigma-Aldrich, St. Louis, MA, USA, 99.9%), and the accelerator was citric acid monohydrate (C6H8O7•H2O, Sigma-Aldrich, St. Louis, MA, USA, 99.9%). Supercritical drying was carried out in a Quorum E3100 critical point dryer, and the CO2 used was acquired from the INFRA Group.
The following equipment and parameters were used to carry out the characterizations:
For IR spectroscopy, to determine the functional groups, a Perkin Elmer Spectrum 65 spectrometer was used, along with the potassium bromide (KBr) pellet technique. The transmittance was calculated for 2 min, covering wavelengths between 4000 and 400 cm−1.
X-ray diffraction was performed on a Bruker D8 ADVANCE diffractometer using the powder method. To establish the crystallographic-structural behavior, scanning was performed with Cu-Kα radiation covering the 2θ angle in a range of 20° to 80°. HighScore Plus software (v5.2) was used to conduct the Rietveld analysis and determine the cell parameters. The average size of the crystallites was obtained using the Debye–Scherrer equation.
Scanning electron microscopy was performed on selected samples of the crystallized aerogels, which were coated with gold using the sputter deposition technique for 5 s, with JEOL JSM-7800F equipment (Tokyo, Japan) at a voltage of 2 kV, and it was performed on selected ceramic powders with JEOL JSM-6701F equipment (Tokyo, Japan) at a voltage of 10 kV. To carry out the elemental chemical study by energy dispersive X-ray spectroscopy (EDS), these procedures took advantage of adapted detectors: an EBSD detector from EDAX for the crystallized aerogels and an INCAx-act detector from Oxford Instruments for the ceramic powders. Transmission electron microscopy (TEM) was performed with JEOL JEM-2100 equipment (Tokyo, Japan) under a bright field, and selected area electron diffraction (SAED) was performed using Gatan DigitalMicrograph software 3.01.
For photoluminescence analysis, a Hitachi F-7000 (Tokyo, Japan) fluorescence spectrophotometer with a 150-W xenon lamp was used. The excitation and emission spectra were recorded based on the interaction ranges of the activator ion (Eu3+). These were produced using wavelengths of EX: 250 nm and EM: 612 nm. The analysis time for each result was 60 nm/min, using a voltage of 400 V and maintaining the slit opening at 10 nm. Color Calculator software (v2.0) was employed to calculate the values (x, y) of the chromaticity coordinates and establish the colorimetry (CIE). The time decay (λem = 611 nm) was calculated with an SC-15 device with a front phase sample holder for the powders and films, and the quantum yield was calculated with an SC-30 Sphere Integrator Module. Both techniques were performed on an Edinburgh Instruments FS5 spectrofluorometer position of Lu2O3 matrices through the epoxide-assisted sol–gel process.
Finally, the nitrogen adsorption–desorption analysis was performed using the BET technique in a Quantachrome AutosorbIQ unit, performing desorption for 18 h at 423 K in a 9 mm cell.

4.1. Synthesis

4.1.1. The Sol–Gel Process

The precursors for the sol–gel process were in the form of chlorides (RECl3), which were obtained from a mixture of rare-earth sesquioxides (RE2O3) in 32.638 mmol of hydrochloric acid (HCl). Forming the host matrix of Lu2O3 and the matrix of Eu2O3 required only the mixture of 0.251 mmol and 0.284 mmol, respectively. Forming the different systems of doped samples, 2, 5, 8, and 10 mol% of Eu2O3 was mixed into the Lu2O3 host matrix. The reaction generated was as follows:
RE 2 O 3 + 6 HCl 2 RECl 3 + 3 H 2 O
Complete dispersion of RE2O3 (where for this project, RE can be Lu or Eu) in 6HCl was evidenced by the change in appearance from a whitish to a translucent solution, indicating the formation of the metal salt (2RECl3). Once the metal salt was formed, 34.782 mmol of CH3CH2OH was added as a solvent. At this point, the sol formed, and the hydrolysis and condensation reactions began. To promote condensation, propylene oxide (20 mmol) was added as an epoxy assistant [19,35]. This reaction was as follows:
Gels 10 00736 i001
Citric acid monohydrate dissolved in ethanol in a 1:10 ratio was added immediately. This acted as a catalyst that caused the formation of a whitish mixture with a thick and lumpy consistency, identifiable as an alcogel. The entire process was carried out under constant stirring at a temperature of 353 K. The alcogel was then aged at room temperature until gravity stopped the flow of the viscous mixture, which occurred within approximately 24 h [8].
After this, wet gels were obtained that could be manipulated in subsequent processes. Although, at this point, the gels were in their most stable form, great care had to be taken when handling them so as not to collapse the microstructure because a homogeneous three-dimensional network confers adequate structural properties, such as mechanical strength, porosity, transparency, and a sizable specific surface area [66].
Figure 11 shows the process, from the dispersion of the chloride salt (Figure 11a) and its change to a sol (Figure 11b), through its transformation into an alcogel (Figure 11c) and its transition to an aged monolithic wet gel (Figure 11d). At this stage, it is possible to obtain xerogels (Figure 11e), which, when thermally treated, are converted to ceramic powders (CPs) with photoluminescent properties (Figure 11f). Alternatively, by using a supercritical drying chamber (Figure 11g), it is possible to convert the aged gels into aerogels (Figure 11h), which when thermally treated become crystallized aerogels (CAs) that also have powerful photoluminescent properties (Figure 11i).

4.1.2. Supercritical Drying of the Aerogels (scCO2)

Supercritical conditions were established for the CO2 to act as a supercritical fluid [18]. This occurred at a pressure of 73.7 bars and a temperature of 305.15 K. Once the aged gels were placed inside the supercritical drying chamber, the interior was filled with CO2. As the gas gradually entered the gel cavities and became a supercritical fluid, the liquid phase (CH3CH2OH) was displaced. The time required for complete drying was 3 days, with the CO2 being replenished every 24 h to ensure complete elimination of the liquid phase. As mentioned in [36], CH3CH2OH can be extracted with relative ease due to the high solubility of ethanol in supercritical CO2.

4.1.3. Heat Treatment

Since materials obtained by the sol–gel process commonly occur in an amorphous form [67], both the xerogels and aerogels were subjected to thermal treatment to obtain the crystalline materials needed for the study: CPs and CAs. Heat treatment was performed at 1073 K for 2 h. This process facilitated the elimination of organic remains and solvent residues that remained anchored to the aged gels during the synthesis process and handling.
Figure 11e reveals that the xerogels did not undergo a significant change when converted into CPs (Figure 11f). While the aerogels (Figure 11h) suffered contraction when crystallized (Figure 11i), they were preserved in the form of monoliths, although they were very fragile to the touch.

Author Contributions

Conceptualization, F.d.J.C.R. and A.G.M.; methodology, A.D.A.M.; validation, J.G.M.; investigation and resources, A.G.M. and F.d.J.C.R.; data curation, J.G.M. and A.D.A.M.; writing—original draft preparation, A.D.A.M.; writing—review and editing, A.G.M. and F.d.J.C.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available upon request from the corresponding author. The data are not publicly available due to privacy or ethical restrictions.

Acknowledgments

The authors thank CONAHCYT for the financial support granted to project CB 2017-2018 A1-S-28234 and the Instituto Politécnico Nacional through SIP-IPN projects 20241486 and 20241489. A. D. Alcantar-Mendoza acknowledges the Conahcyt Ph D scholarship. The authors thank CNMN-IPN, Lab CREA-CIITEC, ESIQIE, CICATA, and the CIACyT-UASLP for their experimental support. The authors also would like to thank Henry Jankiewicz for the editing work that he did for this paper and M. García Murillo for her assistance.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Crawford, S.E.; Ohodnicki, P.R.; Baltrus, J.P. Materials for the photoluminescent sensing of rare earth elements: Challenges and opportunities. J. Mater. Chem. C 2020, 8, 7975–8006. [Google Scholar] [CrossRef]
  2. Maciel, G.S.; Rakov, N. Photon conversion in lanthanide-doped powder phosphors: Concepts and applications. RSC Adv. 2015, 5, 17283–17295. [Google Scholar] [CrossRef]
  3. Chen, Z.; Dong, G.; Barillaro, G.; Qiu, J.; Yang, Z. Emerging and perspectives in microlasers based on rare-earth ions activated micro-/nanomaterials. Prog. Mater. Sci. 2021, 121, 100814. [Google Scholar] [CrossRef]
  4. Zhao, Q.; Guo, N.; Jia, Y.; Lv, W.; Shao, B.; Jiao, M.; You, H. Facile surfactant-free synthesis and luminescent properties of hierarchical europium-doped lutetium oxide phosphors. J. Colloid Interface Sci. 2013, 394, 216–222. [Google Scholar] [CrossRef] [PubMed]
  5. Hossain, M.K.; Khan, M.I.; El-Denglawey, A. A review on biomedical applications, prospects, and challenges of rare earth oxides. Appl. Mater. Today 2021, 24, 101104. [Google Scholar] [CrossRef]
  6. Bünzli, J.-C.G. Lanthanide Photonics: Shaping the Nanoworld. Trends Chem. 2019, 1, 751–762. [Google Scholar] [CrossRef]
  7. Paderni, D.; Giorgi, L.; Fusi, V.; Formica, M.; Ambrosi, G.; Micheloni, M. Chemical sensors for rare earth metal ions. Coord. Chem. Rev. 2021, 429, 213639. [Google Scholar] [CrossRef]
  8. Clapsaddle, B.; Neumann, B.; Wittstock, A.; Sprehn, D.; Gash, A.; Satcher, J., Jr.; Simpson, R.; Bäumer, M. A sol-gel methodology for the preparation of lanthanide-oxide aerogels: Preparation and characterization. J. Sol-Gel Sci. Technol. 2012, 64, 381–389. [Google Scholar] [CrossRef]
  9. Leventis, N.; Vassilaras, P.; Fabrizio, E.F.; Dass, A. Polymer nanoencapsulated rare earth aerogels: Chemically complex but stoichiometrically similar core-shell superstructures with skeletal properties of pure compounds. J. Mater. Chem. 2007, 17, 1502–1508. [Google Scholar] [CrossRef]
  10. Rim, K.T.; Koo, K.H.; Park, J.S. Toxicological Evaluations of Rare Earths and Their Health Impacts to Workers: A Literature Review. Saf. Health Work 2013, 4, 12–26. [Google Scholar] [CrossRef]
  11. Yao, C.; Yao, C.; Tong, Y. Lanthanide ion-based luminescent nanomaterials for bioimaging. Trends Anal. Chem. 2012, 39, 60–71. [Google Scholar] [CrossRef]
  12. He, B.-X.; Liang, Y.; Xu, L.-W.; Shao, L.-B.; Liu, D.-G.; Yang, F.; Liang, G.-J. Extraction of REEs (Ce, Tb, Y, Eu) from Phosphors Waste by a Combined Alkali Roasting-Acid Leaching Process. Minerals 2021, 11, 437. [Google Scholar] [CrossRef]
  13. Zhou, J.; Leaño, J., Jr.; Liu, Z.; Jin, D.; Wong, K.-L.; Liu, R.-S.; Bünzli, J.-C. Impact of Lanthanide Nanomaterials on Photonic Devices and Smart Applications. Small 2012, 14, e1801882. [Google Scholar] [CrossRef] [PubMed]
  14. Tshikovhi, A.; Koao, L.F.; Malevu, T.D.; Linganiso, E.C.; Motaung, T.E. Dopants concentration on the properties of various host materials by sol-gel method: Critical review. Results Mater. 2023, 19, 100447. [Google Scholar] [CrossRef]
  15. Sakka, S. Birth of the sol–gel method: Early history. J. Sol-Gel Sci. Technol. 2022, 102, 478–481. [Google Scholar] [CrossRef]
  16. Olvera, A.; García, M.; López, P.; López, A.; Reyes, A.; Morales, Á.; Hernández, F.; Aguilera, L. Influence of Eu3+ doping content on antioxidant properties of Lu2O3 sol-gel derived nanoparticles. Mater. Sci. Eng. C 2016, 69, 850–855. [Google Scholar] [CrossRef]
  17. Kameneva, S.; Yorov, K.; Kamilov, R.; Kottsov, S.; Teplonogova, M.; Khamova, T.; Popkov, M.; Tronev, I.; Baranchikov, A.; Ivanov, V. Epoxide synthesis of binary rare earth oxide aerogels with high molar ratios (1:1) of Eu, Gd, and Yb. J. Sol-Gel Sci. Technol. 2023, 107, 586–597. [Google Scholar] [CrossRef]
  18. García, V.M.; García, A.; Carrillo, F.; Álvarez, R.I.; Madrigal, E. A New Ultrafine Luminescent La2O3:Eu3+ Aerogel. Gels 2023, 9, 615. [Google Scholar] [CrossRef]
  19. Chu, W.; Guo, S.; Zhang, J.; Liu, X.; Zhao, X.; Yi, X.; Liu, B. Rare earth lanthanum based aerogels with reduced chlorine ions by a modified epoxide gelation method. Chem. Phys. Lett. 2020, 761, 138072. [Google Scholar] [CrossRef]
  20. Worsley, M.; Ilsemann, J.; Gesing, T.; Zielasek, V.; Nelson, A.; Ferreira, R.; Carlos, L.; Gash, A.; Bäumer, M. Chlorine-free, monolithic lanthanide series rare earth oxide aerogels via epoxide-assisted sol-gel method. J. Sol-Gel Sci. Technol. 2018, 89, 176–188. [Google Scholar] [CrossRef]
  21. Kistler, S. Coherent Expanded Aerogels and Jellies. Nature 1931, 127, 741. [Google Scholar] [CrossRef]
  22. Stergar, J.; Maver, U. Review of aerogel-based materials in biomedical applications. J. Sol-Gel Sci. Technol. 2016, 77, 738–752. [Google Scholar] [CrossRef]
  23. Smirnova, I. Aerogels: Current status and challenges for the future. J. Supercrit. Fluids 2015, 106, 1. [Google Scholar] [CrossRef]
  24. Ferreira-Gonçalves, T.; Constantin, C.; Neagu, M.; Pinto, C.; Sabri, F.; Simón-Vázquez, R. Safety and efficacy assessment of aerogels for biomedical applications. Biomed. Pharmacother. 2021, 144, 112356. [Google Scholar] [CrossRef] [PubMed]
  25. Montes, S.; Maleki, H. Aerogels and their applications. In Colloidal Metal Oxide Nanoparticles Synthesis, Characterization and Applications, 1st ed.; Thomas, S., Tresa, A., Velayudhan, P., Eds.; Elsevier Inc.: Amsterdam, The Netherlands, 2020; Volume 1, pp. 337–399. [Google Scholar]
  26. Rusch, P.; Zámbó, D.; Bigall, N.C. Control over Structure and Properties in Nanocrystal Aerogels at the Nano-, Micro-, and Macroscale. Acc. Chem. Res. 2020, 53, 2414–2424. [Google Scholar] [CrossRef]
  27. Noman, M.T.; Amor, N.; Ali, A.; Petrik, S.; Coufal, R.; Adach, K.; Fijalkowski, M. Aerogels for Biomedical, Energy and Sensing Applications. Gels 2021, 7, 264. [Google Scholar] [CrossRef]
  28. Keller, J.; Wiemann, M.; Gröters, S.; Werle, K.; Vennemann, A.; Landsiedel, R.; Wohlleben, W. Aerogels are not regulated as nanomaterials, but can be assessed by tiered testing and grouping strategies for nanomaterials. Nanoscale Adv. 2021, 3, 3881–3893. [Google Scholar] [CrossRef]
  29. Bheekhun, N.; Rahim, A.; Talib, A.; Hassan, M.R. Aerogels in Aerospace: An Overview. Adv. Mater. Sci. Eng. 2013, 2013, 406065. [Google Scholar] [CrossRef]
  30. Fricke, J. Aerogels and their applications. J. Non-Cryst. Solids 1992, 147–148, 356–362. [Google Scholar] [CrossRef]
  31. Esquivel-Castro, T.A.; Ibarra-Alonso, M.C.; Oliva, J.; Martínez-Luévanos, A. Porous aerogel and core/shell nanoparticles for controlled drug delivery: A review. Mater. Sci. Eng. C 2019, 96, 915–940. [Google Scholar] [CrossRef]
  32. Innocenzi, P. Sol-gel processing for advanced ceramics, a perspective. Open Ceram. 2023, 16, 100477. [Google Scholar] [CrossRef]
  33. Pierre, A.C.; Pajonk, M. Chemistry of Aerogels and Their Applications. Chem. Rev. 2002, 102, 4243–4265. [Google Scholar] [CrossRef] [PubMed]
  34. Carrol, M.K.; Anderson, A.M. Twenty years of aerogel research at an undergraduate institution. J. Sol-Gel Sci. Technol. 2023, 1, 1–12. [Google Scholar] [CrossRef]
  35. Rechberger, F.; Niederberger, M. Synthesis of aerogels: From molecular routes to 3-dimensional nanoparticle assembly. Nanoscale Horiz. 2017, 2, 6–30. [Google Scholar] [CrossRef] [PubMed]
  36. García-González, C.A.; Camino-Rey, M.C.; Alnaief, M.; Zetzl, C.; Smirnova, I. Supercritical drying of aerogels using CO2: Effect of extraction time on the end material textural properties. J. Supercrit. Fluids 2012, 66, 297–306. [Google Scholar] [CrossRef]
  37. Lazrag, M.; Lemaitre, C.; Castel, C.; Hannachi, A.; Barth, D. Aerogel production by supercritical drying of organogels: Experimental study and modelling investigation of drying kinetics. J. Supercrit. Fluids 2018, 140, 394–405. [Google Scholar] [CrossRef]
  38. Scherer, G.W. Characterization of aerogels. Adv. Colloid Interface Sci. 1998, 76–77, 321–339. [Google Scholar] [CrossRef]
  39. Lebedev, A.E.; Lovskaya, D.D.; Menshutina, N.V. Modeling and scale-up of supercritical fluid processes. Part II: Supercritical drying of gel particles. J. Supercrit. Fluids 2021, 174, 105238. [Google Scholar] [CrossRef]
  40. Vareda, J.P.; Lamy-Mendes, A.; Durães, L. A reconsideration on the definition of the term aerogel based on current drying trends. Microporous Mesoporous Mater. 2018, 258, 211–216. [Google Scholar] [CrossRef]
  41. Zhang, H.; Chen, J.; Guo, H. Electrospinning synthesis and luminescent properties of Lu2O3:Eu3+ nanofibers. J. Rare Earths 2010, 28, 232–235. [Google Scholar] [CrossRef]
  42. Carrera, M.; García, A.; Carrillo, F.; García-Hernández, M.; Morales, A.; Velumani, S.; Rosa, E.; Kassiba, A. Lu2O3:Eu3+ glass ceramic films: Synthesis, structural and spectroscopic studies. Mater. Res. Bull. 2014, 51, 418–425. [Google Scholar] [CrossRef]
  43. Zhuang, L.; Zhang, Z.; Wang, Y.; Wan, H.; Xu, Z.; Mao, R.; Feng, H.; Zhao, J. Afterglow Behavior, Energy Transfer and Trap States in Lu2O3:Eu Single Crystal Scintillator. ECS J. Solid State Sci. Technol. 2018, 7, R17–R20. [Google Scholar] [CrossRef]
  44. Zych, E.; Trojan-Piegza, J. Anomalous activity of Eu3+ in S6 site of Lu2O3 in persistent luminescence. J. Lumin. 2007, 122–123, 335–338. [Google Scholar] [CrossRef]
  45. Feng, H.; Zhuang, L.; Huang, S.; Zhang, Z.; Xu, Z.; Mao, R.; Zhao, J. Correlation of afterglow, trap states and site preference in RE2O3:1%Eu (RE = Lu, Y, Sc) single crystal scintillators. J. Lumin. 2018, 209, 232–236. [Google Scholar] [CrossRef]
  46. Zhuang, L.; Feng, H.; Huang, S.; Zhang, Z.; Yong, W.; Mao, R.; Zhao, J. The luminescent properties comparison of RE2O3:Eu(RE=Lu,Y,Sc) with high and low Eu doping concentrations. J. Alloys Compd. 2019, 781, 302–307. [Google Scholar] [CrossRef]
  47. Won, H.; Kim, D.H. Luminescent Properties of Lu2O3:Eu3+ Nanophosphors Synthesized by Using Acid-Catalyzed Sol-Gel Method. New Phys. Sae Mulli 2013, 63, 319–323. [Google Scholar] [CrossRef]
  48. Jia, G.; You, H.; Zheng, Y.; Liu, K.; Guo, N.; Zhang, H. Synthesis and characterization of highly uniform Lu2O3:Ln3+ (Ln = Eu, Er, Yb) luminescent hollow microspheres. CrystEngComm 2010, 12, 2943–2948. [Google Scholar] [CrossRef]
  49. Gao, Y.; Gong, J.; Fan, M.; Fang, Q.; Wang, N.; Han, W.; Xu, Z. Large-scale synthesis of Lu2O3:Ln3+ (Ln3+ = Eu3+, Tb3+, Yb3+/Er3+, Yb3+/Tm3+, and Yb3+/Ho3+) microspheres and their photoluminescence properties. Mater. Res. Bull. 2012, 47, 4137–4145. [Google Scholar] [CrossRef]
  50. Popielarski, P.; Zeler, J.; Bolek, P.; Zorenko, T.; Paprocki, K.; Zych, E.; Zorenko, Y. Radio-, Thermo- and Photoluminescence Properties of Lu2O3:Eu and Lu2O3:Tb Nanopowder and Film Scintillators. Crystals 2019, 9, 148. [Google Scholar] [CrossRef]
  51. Tang, J.; Zhang, C.; Du, J.; Li, X.; Jia, X.; Jia, G. Controllable synthesis and luminescence properties of monodisperse lutetium oxide spheres with tunable particle sizes and multicolor emissions. J. Alloys Compd. 2021, 867, 159029. [Google Scholar] [CrossRef]
  52. Xu, Z.; Lin, J.; Sun, Y.; Ding, F.; Fan, H.; Shi, S.; Fang, Q.; Bi, Y.; Gao, Y. Facile Synthesis and Down-Conversion Emission of RE3+-Doped Lutetium Oxide Nanoparticles. J. Nanosci. Nanotechnol. 2017, 18, 2850–2855. [Google Scholar] [CrossRef]
  53. Nair, S.G.; Satapathy, J.; Kumar, N.P. Influence of synthesis, dopants, and structure on electrical properties of bismuth ferrite (BiFeO3). Appl. Phys. A 2020, 126, 836. [Google Scholar] [CrossRef]
  54. Shakirzyanov, R.I.; Volodina, N.O.; Kozlovskiy, A.L.; Zdorovets, M.V.; Shlimas, D.I.; Borgekov, D.B.; Garanin, Y.A. Study of the Structural, Electrical, and Mechanical Properties and Morphological Features of Y-Doped CeO2 Ceramics with Porous Structure. J. Compos. Sci. 2023, 7, 411. [Google Scholar] [CrossRef]
  55. Das, T.; Das, D.; Parashar, K.; Parashar, S.K.S.; Anupama, A.V.; Sahoo, B.; Das, B.K. Influence of Mg doping on structural, dielectric properties and Urbach energy in ZnO ceramics. J. Mater. Sci. Mater. Electron. 2023, 34, 2056. [Google Scholar] [CrossRef]
  56. Bart, C. What is the ‘Lanthanide Contraction’? Inorg. Chem. 2023, 63, 3713–3714. [Google Scholar] [CrossRef] [PubMed]
  57. Aegerter, M.; Leventis, N.; Koebel, M. Aerogels Handbook (Advances in Sol-Gel Derived Materials and Technologies); Springer: New York, NY, USA, 2011. [Google Scholar] [CrossRef]
  58. Binnemans, K. Interpretation of europium(III) spectra. Coord. Chem. Rev. 2015, 295, 1–45. [Google Scholar] [CrossRef]
  59. Thoř, T.; Rubešova, K.; Jakeš, V.; Cajzl, J.; Nádherný, L.; Mikolášova, D.; Beitlerová, A.; Nikl, M. Europium-doped Lu2O3 phosphors prepared by a sol-gel method. IOP Conf. Ser. Mater. Sci. Eng. 2019, 465, 012009. [Google Scholar] [CrossRef]
  60. García, V.M.; García, A.; Carrillo, F.; Cervantes, A.; Cayetano, N. Luminescent aerogels of Gd2O3:Eu3+ and Gd2O3:(Eu3+,Tb3+). Bull. Mater. Sci. 2023, 46, 75. [Google Scholar] [CrossRef]
  61. Iordanova, R.; Milanova, M.; Yordanova, A.; Aleksandrov, L.; Nedyalkov, N.; Kukeva, R.; Petrova, P. Structure and Luminescent Properties of Niobium-Modified ZnO-B2O3:Eu3+ Glass. Materials 2024, 17, 1415. [Google Scholar] [CrossRef]
  62. Aleksandrov, L.; Yordanova, A.; Milanova, M.; Iordanova, R.; Tzvetkov, P.; Markov, P.; Petrova, P. Glass-Ceramic Materials with Luminescent Properties in the System ZnO-B2O3-Nb2O5-Eu2O3. Molecules 2024, 29, 3452. [Google Scholar] [CrossRef]
  63. Trond, S.; Martin, J.; Stanavage, J.; Smith, A. Properties of Some Selected Europium-Activated Red Phosphors. J. Electrochem. Soc. 1969, 116, 1047–1050. [Google Scholar] [CrossRef]
  64. Brunauer, S.; Deming, L.S.; Deming, W.E.; Teller, E. On a Theory of the van der Waals Adsorption of Gases. J. Am. Chem. Soc. 1940, 62, 1723–1732. [Google Scholar] [CrossRef]
  65. Rouquerol, J.; Llewellyn, P.L.; Rouquerol, F. Is the BET Equation Applicable to Microporous Adsorbents? Stud. Surf. Sci. Catal. 2007, 160, 49–56. [Google Scholar] [CrossRef]
  66. Matter, F.; Luna, A.L.; Niederberger, M. From colloidal dispersions to aerogels: How to master nanoparticle gelation. Nano Today 2020, 30, 100827. [Google Scholar] [CrossRef]
  67. Cheng, W.; Rechberger, F.; Niederberger, M. Three-Dimensional Assembly of Yttrium Oxide Nanosheets into Luminescent Aerogel Monoliths with Outstanding Adsorption Properties. ACS Nano 2016, 10, 2467–2475. [Google Scholar] [CrossRef]
Figure 1. IR spectra of Lu2O3:Eu3+: (a) ceramic powders and (b) crystallized aerogels.
Figure 1. IR spectra of Lu2O3:Eu3+: (a) ceramic powders and (b) crystallized aerogels.
Gels 10 00736 g001
Figure 2. XRD patterns of the (a) ceramic powders, (b) crystallized aerogels of the Lu2O3:Eu3+ system and (c) schematic representation of the crystalline structure of C-type RE2O3.
Figure 2. XRD patterns of the (a) ceramic powders, (b) crystallized aerogels of the Lu2O3:Eu3+ system and (c) schematic representation of the crystalline structure of C-type RE2O3.
Gels 10 00736 g002
Figure 3. Crystallite size vs. lattice distortion degree for (a) ceramic powders and (b) crystallized aerogels.
Figure 3. Crystallite size vs. lattice distortion degree for (a) ceramic powders and (b) crystallized aerogels.
Gels 10 00736 g003
Figure 4. SEM images at different magnifications of (a) ×50,000 and (b) ×100,000, and (c) EDS analysis of the CPLu sample. SEM images at different magnifications of (d) ×50,000 and (e) ×100,000, and (f) EDS analysis of the CPLu10 sample.
Figure 4. SEM images at different magnifications of (a) ×50,000 and (b) ×100,000, and (c) EDS analysis of the CPLu sample. SEM images at different magnifications of (d) ×50,000 and (e) ×100,000, and (f) EDS analysis of the CPLu10 sample.
Gels 10 00736 g004
Figure 5. SEM images at different magnifications of (a) ×50,000 and (b) ×100,000, and (c) EDS analysis of the CALu sample. SEM images at different magnifications of (d) ×50,000 and (e) ×100,000, and (f) EDS analysis of the CALu10 sample.
Figure 5. SEM images at different magnifications of (a) ×50,000 and (b) ×100,000, and (c) EDS analysis of the CALu sample. SEM images at different magnifications of (d) ×50,000 and (e) ×100,000, and (f) EDS analysis of the CALu10 sample.
Gels 10 00736 g005
Figure 6. (a,b) TEM images of the CPLu10 sample at different magnifications, (c) HRTEM, and (d) SAED analysis; (e,f) TEM images of the CALu10 sample at different magnifications, (g) HRTEM, and (h) SAED analysis.
Figure 6. (a,b) TEM images of the CPLu10 sample at different magnifications, (c) HRTEM, and (d) SAED analysis; (e,f) TEM images of the CALu10 sample at different magnifications, (g) HRTEM, and (h) SAED analysis.
Gels 10 00736 g006
Figure 7. Photoluminescence study results: (a) excitation and (b) emissions of the ceramic powders, and (c) excitation and (d) emissions of the aerogels.
Figure 7. Photoluminescence study results: (a) excitation and (b) emissions of the ceramic powders, and (c) excitation and (d) emissions of the aerogels.
Gels 10 00736 g007
Figure 8. Energy-level diagram of the transfer energy of the Eu3+ in the CPs and CAs with the Lu2O3:Eu3+ system.
Figure 8. Energy-level diagram of the transfer energy of the Eu3+ in the CPs and CAs with the Lu2O3:Eu3+ system.
Gels 10 00736 g008
Figure 9. (a) CIE diagram and (b) results of the decay time and quantum yield of the Eu2O3 matrices and samples, showing strong photoluminescence in the CP and CA samples doped at 10 mol%.
Figure 9. (a) CIE diagram and (b) results of the decay time and quantum yield of the Eu2O3 matrices and samples, showing strong photoluminescence in the CP and CA samples doped at 10 mol%.
Gels 10 00736 g009
Figure 10. (a) N2 adsorption–desorption isotherms and (b) BJH pore-size distribution curves for the samples CPLu10 and CALu10.
Figure 10. (a) N2 adsorption–desorption isotherms and (b) BJH pore-size distribution curves for the samples CPLu10 and CALu10.
Gels 10 00736 g010
Figure 11. Stages in the sol–gel process used for the Lu2O3:Eu3+ system from metal salt to ceramic powders and crystallized aerogel: (a) metal salt, (b) sol, (c) wet gel, (d) aged monolithic gel, (e) general xerogel of the Lu2O3:Eu3+ system, (f) ceramic powders (the CPLu10 sample) under UV radiation, (g) supercritical drying chamber, (h) general aerogel of the Lu2O3:Eu3+ system, and (i) a crystallized aerogel (the CALu10 sample) under UV radiation.
Figure 11. Stages in the sol–gel process used for the Lu2O3:Eu3+ system from metal salt to ceramic powders and crystallized aerogel: (a) metal salt, (b) sol, (c) wet gel, (d) aged monolithic gel, (e) general xerogel of the Lu2O3:Eu3+ system, (f) ceramic powders (the CPLu10 sample) under UV radiation, (g) supercritical drying chamber, (h) general aerogel of the Lu2O3:Eu3+ system, and (i) a crystallized aerogel (the CALu10 sample) under UV radiation.
Gels 10 00736 g011
Table 1. Classification of the samples of ceramic powders and crystallized aerogels with the Lu2O3:Eu3+ system.
Table 1. Classification of the samples of ceramic powders and crystallized aerogels with the Lu2O3:Eu3+ system.
Ceramic Powder|Crystallized AerogelEu2O3 mol%SystemDescription
CPLu|CALu0Lu2O3Host matrix
CPLu2|CALu22Lu2O3:Eu3+Doping
CPLu5|CALu55
CPLu8|CALu88
CPLu10|CALu1010
CPEu|CAEu100Eu2O3Matrix
Table 2. Crystallographic data from the Debye–Scherrer equation and the Rietveld and SAED analyses.
Table 2. Crystallographic data from the Debye–Scherrer equation and the Rietveld and SAED analyses.
MaterialSampleCrystallite Size Average (nm)Rietveld Analysis
D-Spacing (222) Plane (nm)
SAED Analysis
D-Spacing (nm)
Ceramic
Powders
CPLu15.83-
CPLu213.113-
CPLu515.72.9-
CPLu89.63-
CPLu1010.332.9
CPEu-3.1-
Crystallized AerogelsCALu17.72.99-
CALu210.42.99-
CALu517.12.99-
CALu88.23-
CALu1013.52.992.99
CAEu-3.1-
Table 3. Lattice parameters and unit cell volumes obtained by Rietveld analysis for the CPs and CAs.
Table 3. Lattice parameters and unit cell volumes obtained by Rietveld analysis for the CPs and CAs.
Samplea/b/c [Å]Volume [Å3]α/β/γ [°]
Ceramic
Powders
CPLu10.38/10.38/10.381119.58490/90/90
CPLu210.40/10.40/10.401127.454
CPLu510.42/10.42/10.421133.297
CPLu810.45/10.45/10.451141.643
CPLu1010.47/10.47/10.471148.084
CPEu10.85/10.85/10.851280.276
Crystallized AerogelsCALu10.39/10.39/10.391124.719
CALu210.41/10.41/10.411128.958
CALu510.41/10.41/10.411130.718
CALu810.45/10.45/10.451141.607
CALu1010.45/10.45/10.451142.101
CAEu10.87/10.87/10.871287.275
Table 4. CIE chromaticity coordinates, relative luminescence intensity ratios (Rs) of the (5D0-7F2)/(5D0-7F1) transitions, and quantum yields for the CPs, CAs, and other Eu3+-doped oxide compounds. “n.a.” (not applicable).
Table 4. CIE chromaticity coordinates, relative luminescence intensity ratios (Rs) of the (5D0-7F2)/(5D0-7F1) transitions, and quantum yields for the CPs, CAs, and other Eu3+-doped oxide compounds. “n.a.” (not applicable).
SampleChromaticity
Coordinates (x, y)
RQuantum Yield (%)Ref.
Ceramic
Powders
CPLu---n.a.
CPLu2(0.4983, 0.3388)3.88-
CPLu5(0.5266, 0.3349)4.1-
CPLu8(0.5128, 0.3243)4.1-
CPLu10(0.4823, 0.3128)4.217.1
CPEu(0.6322, 0.3673)22.03
Crystallized AerogelsCALu---
CALu2(0.4855, 0.3284)4.1-
CALu5(0.4915, 0.3291)4.1-
CALu8(0.4635, 0.3184)4.09-
CALu10(0.5383, 0.3318)4.637.92
CAEu(0.6220, 0.3775)1.451.05
92% Gd-8% Eu-7.472.4[60]
50ZnO:47B2O3:
3Nb2O3:0.5Eu2O3
(0.656, 0.343)5.16-[61]
Glass-ceramics 25 h(0.652, 0.348)5.49-[62]
NTSC red phosphors standard(0.67, 0.33)--[61]
Y2O2S:Eu3+(0.658, 0.340)--[63]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alcantar Mendoza, A.D.; García Murillo, A.; Carrillo Romo, F.d.J.; Guzmán Mendoza, J. Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels. Gels 2024, 10, 736. https://doi.org/10.3390/gels10110736

AMA Style

Alcantar Mendoza AD, García Murillo A, Carrillo Romo FdJ, Guzmán Mendoza J. Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels. Gels. 2024; 10(11):736. https://doi.org/10.3390/gels10110736

Chicago/Turabian Style

Alcantar Mendoza, Alan D., Antonieta García Murillo, Felipe de J. Carrillo Romo, and José Guzmán Mendoza. 2024. "Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels" Gels 10, no. 11: 736. https://doi.org/10.3390/gels10110736

APA Style

Alcantar Mendoza, A. D., García Murillo, A., Carrillo Romo, F. d. J., & Guzmán Mendoza, J. (2024). Studies on the Powerful Photoluminescence of the Lu2O3:Eu3+ System in the Form of Ceramic Powders and Crystallized Aerogels. Gels, 10(11), 736. https://doi.org/10.3390/gels10110736

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