Next Article in Journal
Dynamics Analysis of Nonlinear Differential Equation Systems Applied to Low-Grade Gliomas and Their Treatment
Previous Article in Journal
Linking Error Estimation in Fixed Item Parameter Calibration: Theory and Application in Large-Scale Assessment Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Analysis of Erbium-Doped Silica-Based Glass-Ceramics Using Anomalous and Small-Angle X-Ray Scattering

by
Helena Cristina Vasconcelos
1,2,*,
Maria Meirelles
1,3,
Reşit Özmenteş
4 and
Luís Santos
5
1
Faculty of Science and Technology, University of the Azores, Ponta Delgada, S. Miguel, 9500-321 Azores, Portugal
2
Laboratory of Instrumentation, Biomedical Engineering and Radiation Physics (LIBPhys, UNL), Department of Physics, NOVA School of Science and Technology, 2829-516 Caparica, Portugal
3
Research Institute of Marine Sciences, University of the Azores (OKEANOS), Horta, Faial, 9901-862 Azores, Portugal
4
Vocational School of Health Services, Bitlis Eren University, 13100 Bitlis, Türkiye
5
Centro de Química Estrutural, Institute of Molecular Sciences, Departamento de Engenharia Química, Instituto Superior Técnico, University of Lisboa, Av. Rovisco Pais, 1049-001 Lisboa, Portugal
*
Author to whom correspondence should be addressed.
Foundations 2025, 5(1), 5; https://doi.org/10.3390/foundations5010005
Submission received: 14 November 2024 / Revised: 21 January 2025 / Accepted: 10 February 2025 / Published: 12 February 2025
(This article belongs to the Section Physical Sciences)

Abstract

This study employs advanced structural characterization techniques, including anomalous small-angle X-ray scattering (ASAXS), small-angle X-ray scattering (SAXS), and X-ray photoelectron spectroscopy (XPS), to investigate erbium (Er3+)-doped silica-based glass-ceramic thin films synthesized via the sol–gel method. This research examines the SiO2-TiO2 and SiO2-TiO2-PO2.5 systems, focusing on the formation, dispersion, and structural integration of Er3+-containing nanocrystals within the amorphous matrix under different thermal treatments. Synchrotron radiation tuned to the LIII absorption edge of erbium enabled ASAXS measurements, providing element-specific details about the localization of Er3+ ions. The findings confirm their migration into crystalline phases, such as erbium phosphate (EPO) and erbium titanate (ETO). SAXS and Guinier analysis quantified nanocrystal sizes, revealing trends influenced by their composition and heat treatment. Complementary XPS analysis of the Er 5p core-level states provided detailed information on the chemical and electronic environment of the Er3+ ions, confirming their stabilization within the crystalline structure. Transmission electron microscopy (TEM) highlighted the nanoscale morphology, verifying the aggregation of Er3+ ions into well-defined nanocrystals. The results offer a deeper understanding of their size, distribution, and interaction with the surrounding matrix.

1. Introduction

Recent advancements in photonic applications have increasingly focused on rare earth (RE) ion-doped materials due to their unique energy levels, which allow for a variety of radiative transitions across UV, visible, and NIR wavelengths. Common host materials include crystals, glasses, transparent ceramics, and nanocrystalline films. Among these, glass-ceramics (GCs)—amorphous matrices embedded with nanocrystals—are particularly promising, as they combine the properties of glass (like chemical durability and mechanical strength) with the optical properties of nanocrystals [1,2].
For free RE ions, intra-4f transitions are generally parity-forbidden, which leads to low transition probabilities and long luminescence lifetimes. These extended lifetimes are advantageous for applications such as upconversion lasers, where a sustained energy release is desirable [3]. Embedding RE ions in host materials can activate these transitions by introducing asymmetry in the surrounding crystal field, thereby increasing the luminescence intensity of lanthanide ions and enhancing their practical applicability [4]. However, a challenge with RE ions, particularly Er3+, is their low solubility in oxide materials, primarily due to mismatches between the ionic radii and the covalent bonds within the matrix. This mismatch often leads to ion clustering at higher concentrations, causing luminescence degradation via ion–ion coupling linked to erbium ion cross-relaxation processes in clustering Er–O–Er bonds [5] or the formation of optically inactive phases [1]. An alternative approach is to localize these ions within the nanocrystalline phase, which can significantly improve the material’s optical properties [1,2].
Among the GCs, silica–titania (ST) glass-ceramics have gained significant attention due to their compatibility with optical fibers and their tunable refractive index [6]. By adjusting the silica (SiO2)–to–titania (TiO2) ratio, the refractive index of ST glass-ceramics can be fine-tuned to enhance the light propagation and coupling efficiency in various optical circuits and photonic devices. This is critical for minimizing refractive index mismatches at interfaces, thereby reducing light scattering and improving the performance of components such as waveguides, amplifiers, and modulators used in integrated photonic systems.
The control of the nanocrystal size within the glass-ceramic matrix is critical for maintaining high transparency and minimizing optical losses caused by Rayleigh scattering. Ideally, nanocrystals should be kept below 100 nm in diameter [7]. Additionally, co-dopants like phosphorus (P) help form SiO2–TiO2–PO2.5 (STP) oxide matrices, which support higher concentrations of Er3+ ions without aggregation [8]. This improves solubility, extends the excited-state lifetimes of Er3+ ions, and reduces quenching effects. For instance, the crystallization of ErPO4 (EPO) in STP sol–gel glass matrices have shown enhanced fluorescence properties, attributed to the Er3+ ions’ environment within EPO crystallites [1]. Similarly, Er2Ti2O7 (ETO) crystallites can form in ST matrices, further increasing fluorescence emissions [1,9]. Co-dopants help distribute fluorescent ions uniformly within nanocrystals, minimizing clustering and enhancing optical performance in devices such as optical amplifiers [2].
Upon the heat treatment of sol–gel films, RE ions tend to migrate toward smaller nanocrystals, decreasing the average inter-ion distance (R) and intensifying energy transfer interactions due to the R−6 dependence [10,11,12]. This effect is particularly significant in erbium-doped glass compositions, which exhibit a strong fluorescence peak at 1540 nm, crucial for telecommunications applications. However, maintaining a uniform Er3+ distribution at concentrations above 0.5% remains challenging, as higher concentrations can lead to clustering and fluorescence quenching. Abdullah et al. [13] demonstrated that at a 6% Er concentration, silicate materials exhibited luminescence intensities indicative of the formation of Er2Si2O7 phases, contributing to enhanced luminescent properties. Meanwhile, Gao et al. [14] specifically highlighted that the γ-Er2Si2O7 phase exhibited the highest luminescent activity.
Small-angle X-ray scattering (SAXS) is a powerful technique to probe local structures, especially in nanoparticle systems. Anomalous small-angle X-ray scattering (ASAXS) further enhances sensitivity to specific elements or phases, providing a detailed view of crystallization processes and nanoscale phase separations in glassy systems [15]. This approach has been effectively utilized to study crystallization processes and nanoscale phase separations in glassy systems. For example, Hoell et al. [16] demonstrated the core–shell structure of BaF2 nanocrystals formed in phase-separated glasses, showcasing the utility of ASAXS for elucidating the structural evolution during crystallization. Gericke et al. [17] employed ASAXS in combination with absorption spectroscopy to observe nanostructural dynamics, providing detailed insights into compositional variations within the nanocrystals under heat treatment. Furthermore, Haas et al. [18] demonstrated the effectiveness of ASAXS in accessing the phase composition of oxyfluoride glass ceramics doped with Er3+/Yb3+, offering a nuanced understanding of the interplay between nanostructure and nanochemistry in these systems.
In this study, we employ both SAXS and ASAXS to investigate the dispersion of Er3+ ions in silica-based glass-ceramic thin films. The goal is to determine whether the Er3+ ions remain dispersed within the amorphous matrix, form clusters, or transition into ETO or EPO nanocrystals upon heat treatment. Specifically, we aim to perform the following:
  • Examine the dispersion of Er3+ ions: Do they remain uniformly distributed in the amorphous matrix, form clusters, or migrate into nanocrystalline phases?
  • Investigate how thermal treatments affect the distribution and structural organization of Er3+ ions, observed through SAXS and ASAXS.
  • Explore the local electronic environment of Er3+ ions using XPS, focusing on their binding energies and phase composition.
This study builds upon preliminary results from an initial investigation, where the use of SAXS and ASAXS to analyze Er3+ dispersion in silica-based glass-ceramics was presented solely at the national conference Física2000 (Portugal) [19] and was not formally published. We revisit these findings with a more comprehensive analysis, incorporating new data from XRD and XPS measurements. By investigating the distribution of nanocrystals and the localization of Er3+ ions within the glass-ceramic matrix, this work seeks to demonstrate how structural characterization techniques like SAXS and ASAXS can provide critical insights into material organization. These findings lay the groundwork for further studies aimed at leveraging nanostructural control in the development of photonic devices.

2. SAXS and ASAXS to Study Erbium-Doped Nanocrystals Embedded Within Glass

Small-angle X-ray scattering (SAXS) and anomalous SAXS (ASAXS) are powerful techniques for studying the nanoscale structure of doped glass-ceramic matrices. SAXS measures the scattering intensity I q as a function of the scattering vector q, which provides information about the structural features within the sample. By measuring the scattering of X-rays at small angles, SAXS provides detailed information about the size, shape, and spatial distribution of nanostructures, such as nanoparticles, aerosols, micelles, minerals, and particles synthesized through sol–gel reactions, among others [15]. This method is highly valuable for examining structural inhomogeneities in thin films, polymers, and other complex systems.
In SAXS, the scattering intensity is represented as I ( q ) , where q is the scattering vector, which provides information on the structural features within the sample. Here, the intensity depends solely on q, as SAXS is typically conducted at a fixed X-ray energy. The SAXS intensity I ( q ) is proportional to the electron density contrast Δ ρ between phases and follows:
I ( q ) Δ ρ 2 P ( q ) S ( q )
where P ( q ) is the form factor related to the particle shape and size, and S ( q ) is the structure factor accounting for inter-particle correlations. This dependence on electron density contrast allows SAXS to characterize the size and spatial distribution of nanocrystals and amorphous regions.
Building on SAXS, anomalous SAXS (ASAXS) extends this technique by tuning the X-ray energy near the absorption edge of a target element. Therefore, the scattering intensity is represented as I ( q , E ) , because it depends both on the scattering vector q and on the X-ray energy E. By tuning E near the absorption edge of a target element (Er, in this study), ASAXS selectively enhances the scattering contributions from that element. This method is particularly effective for studying composite samples with elements that exhibit spatial discontinuities, provided that the experimental setup can access the absorption edges of these elements [20]. This tuning provides element-specific contrast, isolating the Er3+ contribution to scattering, and thus allowing ASAXS to reveal details about Er3+ ’s localization, providing information about its local distribution within nanocrystalline and amorphous phases.
This energy tuning adjusts the anomalous scattering factor f E , selectively enhancing the scattering contribution from Er3+ ions. By plotting the square root of the summed scattering intensities ( Σ I ) 1 / 2 versus f E , one can separate element-specific contributions from the matrix background. This energy-dependent scattering approach allows us to isolate the contributions of erbium, providing quantitative information into Er3+ dispersion and nanocrystal formation. For instance, the thermal annealing of amorphous silicate glass resulted in the formation of cubic BaF2 nanocrystals embedded within the glass matrix. ASAXS studies conducted near the Ba LIII edge demonstrated that these nanocrystals fell within the nanometer size range and were enveloped by nanoshells of lower electron-density glass [21]. A similar thermally induced nanocrystalline phase also occurred in Cd(S,Se)-doped glasses, where ASAXS conducted near the selenium K edge, was employed to investigate the stoichiometry of nanocrystals smaller than 10 nm [22]. In this research, ASAXS is applied to investigate Er3+-doped thin ST and STP films synthesized using the sol–gel method.

2.1. General Principles

ASAXS extends SAXS by incorporating an energy-dependent scattering factor for selective element analysis. While SAXS measures scattering at a constant energy, ASAXS uses an energy (E) close to the absorption edge of the target element—in this case, Er. As E nears erbium’s LIII absorption edge (8357 eV), variations in the scattering factor f ( E ) selectively enhance scattering from regions containing Er3+. This yields the ASAXS intensity I ( q , E ) , with dual dependence on q and E. The scattering factor f ( E ) can be decomposed as:
f ( E ) = Z + f ( E ) + i f ( E )
where Z is the atomic number of Er, and f E and f ( E ) are the energy-dependent real and imaginary parts of the anomalous dispersion factor, respectively. Near the absorption edge, f E and f ( E ) undergo rapid changes, significantly altering the electron density contrast between Er3+ ions and the surrounding matrix, enhancing the sensitivity of ASAXS to Er3+, allowing it selective detection.
In ASAXS, the energy-dependent scattering intensity I ( q , E ) , which depends on both the scattering vector q and the X-ray energy E, can be represented as:
I q , E = f E r E + f m a t r i x q 2 . P ( q )
where f E r E is the scattering factor specific to Er3+, which varies significantly with energy E near its absorption edge; f m a t r i x q represents the constant scattering factor for the non-Er3+-containing matrix, which depends on q but does not vary with E; and P ( q ) represents the form factor, which is independent of E. Expanding this yields:
I q , E = f E r q , E 2 + f m a t r i x q 2 + 2 R e f E r q , E f m a t r i x * q . P ( q )
The expanded form of the scattering intensity includes contributions from the Er3+ nanocrystals f E r q , E 2 , matrix-only scattering f m a t r i x q 2 , and an interference term 2 R e f E r q , E f m a t r i x * q .
Thus, the scattering intensity of Er3+-doped thin films combines the individual scattering from Er3+ nanocrystals and the matrix, along with an interference term representing the interactions between the two components. Here, f E r q , E varies with both the scattering vector q and the X-ray energy E, especially near Er’s absorption edge, which induces changes in both the real and imaginary parts of the scattering factor. In contrast, f m a t r i x q remains constant with respect to energy.
The interference term 2 R e f E r q , E f m a t r i x * q accounts for the interaction between Er3+ and the matrix scattering contributions. The factor of 2 arises because the interference is symmetric and includes contributions from both f E r q , E f m a t r i x * q and its conjugate f m a t r i x q f E r * q , E , meaning it appears twice in the intensity calculation.
Only the real part of this product is physically measurable, representing the effective interference between Er3+ and matrix components.
The energy-dependent scattering intensity I ( q ,   E ) is essential for isolating the scattering contributions of Er3+ from other components in the material. Only the Er3+-specific terms f E and f ( E ) show significant changes near the LIII edge, while matrix scattering remains relatively stable. Standard references like [23] provide tabulated values of f E and f ( E ) for various elements, which are important for modeling scattering near absorption edges. Figure 1 illustrates the energy-dependent variations in f E and f ( E ) around Er’s L-edges (LIII, LII, LI) from 6 keV to 10 keV, showing abrupt shifts: f E becomes more negative, and f ( E ) more positive. These sharp transitions produce anomalous scattering effects that enhance Er’s contribution while reducing interference from other elements.

2.2. Energy-Dependent Scattering and the Linear Relationship ( Σ I ) 1 / 2 Versus f E

A fundamental aspect of ASAXS analysis is understanding how the scattering intensity varies as a function of X-ray energy, particularly near the absorption edges of the elements in question. This variation is driven by the real part of the anomalous scattering factor, f E , which has a pronounced impact near the absorption edge, whereas the imaginary part, f ( E ) , contributes minimally in this region. In the case of Er3+-doped materials, this energy dependence allows for the selective isolation of the scattering contributions from the Er3+ ions and an understanding of how they are distributed—whether they are dispersed uniformly in the amorphous matrix or incorporated into nanocrystals.
To quantify this, we express the total scattering intensity I ( q ,   E ) as a function of f E , which is directly related to the scattering contributions from Er3+. By integrating the scattering intensities over a range of q-values, we obtain the total integrated intensity Σ I , which can be written as:
Σ I = I E , q d q   ~   K ( A f E ) + B 2
where K is a constant, A represents the contribution from Er3+ ions, and B accounts for the scattering from the matrix. Since the scattering from Er3+ depends on where it is located (whether within nanocrystals or dispersed in the amorphous matrix), the terms A and B will vary accordingly.
To simplify this, we take the square root of Σ I , which gives:
( Σ I ) 1 / 2 = K ( A f E ) + B
This results in a linear relationship [19]:
( Σ I ) 1 / 2 = α f + β
where α = K A and β = K B .
The term α reflects the scattering contribution from Er3+, with higher values of α indicating that Er3+ is more concentrated in nanocrystals, and lower values suggesting that Er3+ is more uniformly dispersed within the amorphous matrix. The intercept β corresponds to the matrix’s baseline scattering, independent of Er3+’s presence.
The ratio β/α is particularly useful, as it indicates the relative amounts of Er3+ in nanocrystals versus the amorphous matrix. A higher β/α ratio suggests that Er3+ is more evenly distributed within the matrix, whereas a lower ratio indicates a stronger scattering contribution from Er3+, typically associated with its concentration in nanocrystals. Thus, the linear relationship described by Equation (7) enables the assessment of the Er3+ distribution.

2.3. Guinier Analysis and Radius of Gyration

In our Er3+-doped samples, we use the Guinier approximation in SAXS to estimate the size of Er3+-containing clusters. This approach yields information about the overall size of scattering entities, helping to distinguish between phases with different structural characteristics, such as nanocrystalline and amorphous regions. The analysis is focused on the low-q region, where qRg ≪ 1, which is particularly sensitive to the overall size of the scattering entities.
The Guinier plot (log I q   v s .   q2) enables us to determine the radius of gyration Rg of Er3+-containing regions, allowing us to estimate the size of these clusters. The Guinier equation, assuming an isotropic, approximately spherical distribution of scattering centers, is given by:
I q = I ( 0 ) e x p R g 2 q 2 3
where I q is the scattering intensity at a given scattering vector q; I ( 0 ) represents the extrapolated scattering intensity at zero angle (forward scattering); and q = 4 π   s i n ( θ ) λ with θ as the scattering angle and λ as the wavelength of the incident radiation.
In this equation, the radius of gyration Rg serves as an indicator of particle size, describing the distribution of the electron density within each scattering entity. This approximation holds for particles which are large compared to atomic scales but small enough to avoid inter-particle interference effects. Here, Rg effectively represents the average electron density distance from the particle’s center of mass.
From the linear region of the Guinier plot, we extract Rg by fitting the slope of log I(q) versus q2, providing a quantitative measure of the size of Er3+-containing clusters. For spherical particles, the linearity observed in the Guinier plot is typically narrow [25], indicating that the scattering entities exhibit a uniform size distribution, consistent with the idealized model of spherical inhomogeneities. When Er3+ ions aggregate into larger regions within the amorphous matrix, an increase in Rg reflects the formation of Er3+-rich regions, though these may not form fully organized nanocrystals. In contrast, if thermal treatment leads to the formation of Er3+ nanocrystals, they will exhibit a characteristic Rg value, indicating a defined size and more regular distribution. Conversely, if Er3+ ions remain well dispersed in the amorphous phase, a smaller Rg value will suggest minimal clustering.

2.4. Synergistic Role of ASAXS and Guinier Analysis for Erbium Distribution

When combined, SAXS and ASAXS provide a comprehensive view of Er3+ clustering in these thin films. ASAXS isolates the energy-dependent contribution of Er3+ to the scattering intensity, focusing on Er3+-rich regions. Meanwhile, Guinier analysis quantifies the size of these regions. Together, they provide a comprehensive understanding of the distribution of Er3+ in GCs, revealing how different thermal treatments affect Er3+ clustering at the nanoscale.

3. Materials and Methods

Thin films were deposited on Si substrates within the SiO2–TiO2–ErO1.5 and SiO2–TiO2–PO2.5–ErO1.5 systems using the sol–gel method, as described in [1]. The films were synthesized with target molar compositions, where SiO2 and TiO2 ratios were precisely controlled, and dopant levels were nominally set as follows:
A: 80SiO2–20TiO2 with 2 mol% ErO1.5;
B: 80SiO2–20TiO2 with 10 mol% PO2.5 and 2 mol% ErO1.5;
C: 98SiO2–2TiO2 with 10 mol% PO2.5 and 2 mol% ErO1.5;
D: 98SiO2–2TiO2 with 2 mol% ErO1.5.
The films, which were approximately 500 nm thick, underwent thermal treatments at 900 °C and 1000 °C for 30 and 60 min, respectively. For clarity, the samples are labeled based on treatment conditions (e.g., A1 for 900 °C/30 min, A2 for 1000 °C/60 min, etc.). These treatments produced transparent glass-ceramic films by precipitating nanocrystals without affecting their transparency.

3.1. SAXS and ASAXS Measurements

Measurements were conducted at the LURE facility (Orsay, Paris, France), now upgraded to the SWING beamline at the SOLEIL Synchrotron (Saint-Aubin, France), using synchrotron radiation at 1.85 GeV and a beam current of approximately 300 mA. Both SAXS and ASAXS measurements were performed on the same beamline. The primary distinction between the two techniques lies in the energy tuning used for ASAXS, which targets specific absorption edges for element-specific analysis, while SAXS uses a broader energy range. The D22 beamline spectrometer, equipped with a double-crystal monochromator (Ge(1,1,1) crystals), selected monochromatic radiation in the 5–15 keV range. For ASAXS, measurements were focused on energies from 8000 eV to 8347 eV, spanning 10 eV to 357 eV below the LIII absorption edge, enabling element-specific analysis. The experimental setup used a near-grazing incidence angle (0.3° ± 0.01°) and an energy resolution of 2 eV, ensuring precision in the measurements. The data acquisition setup was designed to ensure accurate measurements, supported by an adjustable aperture and guard slits to optimize the beam dimensions (H × V = 0.5 × 0.2 to 3 × 0.4 mm2) and minimize parasitic scattering [26].
The beam intensity was measured using Kapton foil-based monitors coupled with NaI scintillators, positioned both before and after the sample along the beam path. This setup ensured the accurate normalization of scattering intensities and compensated for variations in sample thickness and transmission. Calibration of the scattering vector (q) axis was achieved using crystalline samples with sharp Bragg peaks, while absolute intensity normalization utilized water as a reference, leveraging its known differential scattering cross-section of 1.65 × 10−2 cm−1 [26].
The experiments were conducted at room temperature, with the samples mounted in holders designed for solid materials. These holders ensured stable positioning and minimized mechanical vibrations during measurements. The vacuum chamber, equipped with a movable beamstop and a large output window, reduced background scattering and provided flexibility in detector configurations. Data collection utilized a linear position-sensitive detector capable of covering a q-range from 5 × 10−3−1 to 0.4 A˚−1, enabling efficient measurements without frequent adjustments [26]. More details about the standard protocols and advanced instrumentation that facilitated high-quality SAXS and ASAXS measurements at LURE-D22 can be found in [26]. Table 1 provides the specific energies, their shifts relative to the absorption edge, corresponding X-ray wavelengths, and the real part of the scattering factor for each energy.
This energy range enabled the comprehensive mapping of Er3+ scattering behavior, specifically targeting the anomalous scattering effects that reveal Er3+ distribution within the nanostructures of the films. The chosen energy steps (ΔE) allowed for the precise tracking of scattering intensity variations as the energy approached the LIII edge, revealing Er3+-rich nanostructures not visible at energies far from the absorption edge. The grazing incidence geometry enhanced the surface sensitivity and minimized the Si substrate contribution to less than 1% of the total signal.

3.2. Data Processing and Analysis

The real-time acquisition software applied automated corrections to the raw ASAXS data, accounting for detector response, absorption, and refraction effects [27]. The substrate signal, contributing less than 1% of the total, was excluded from these corrections. Scattered intensity I(q) was recorded as a function of the scattering vector q, and Guinier plots (log(I) vs. q2) were generated for each energy point. From these plots, the Guinier radius Rg was determined, providing an estimate of the average nanocrystal size, assuming spherical particles.
The integrated intensity (ΣI)1/2 was plotted against the anomalous scattering factor f E at each energy, following Equation (7). This linear relationship helped isolate Er3+’s contribution to the scattering intensity, revealing its distribution within the nanostructures—whether localized in nanocrystals or uniformly dispersed within the matrix.
To validate these results, complementary structural and morphological analyses were performed using grazing-incidence X-ray diffraction (GIXRD), X-ray photoelectron spectroscopy (XPS), and transmission electron microscopy (TEM). TEM micrographs, captured in bright-field mode using a Hitachi H-800 microscope (Hitachi, Tokyo, Japan) operating at 150 kV, allowed for a detailed examination of the sample morphology and nanocrystal distribution, revealing the structural characteristics of the samples. To analyze the diffraction pattern, the ODPIN online diffraction pattern indexing tool was utilized (https://www.odpin.com (accessed on 19 January 2025)). The crystal structure under investigation was modeled assuming a cubic system with the following unit cell parameters: a = 7.155 Å, b = 7.155 Å, c = 7.155 Å, and angles α = β = γ = 60° (https://legacy.materialsproject.org/materials/mp-5889 (accessed on 19 January 2025)). These parameters correspond to a cubic crystal system. The diffraction data were input into the ODPIN tool, which assisted in indexing the diffraction pattern and determining the corresponding crystallographic information based on the provided unit cell parameters.
GIXRD patterns of the films after heat treatment were recorded using a Siemens D-5000 diffractometer (Siemens, Munich, Germany) with CuKα radiation at a grazing incidence angle of 2°. The 2θ range was covered from 21° to 45° with a step size of 0.04°. The choice of the windows, at 21–37° and 25–45°, was deliberate. The first window captures the key diffraction peaks corresponding to a mix of the rutile phase (110) and ETO phase diffraction planes (311), (222), and (400). The second window isolates peaks corresponding only to the ETO phase diffraction planes (311), (222), (400), and the additional (331).
XPS measurements were conducted on an ESCALAB 250 spectrometer (Thermo Fisher Scientific K.K., Tokyo, Japan) with dual aluminum–magnesium anodes and monochromatic Al Kα radiation (hν = 1486.6 eV). Initially, a survey was performed exclusively on sample A1 to obtain an overview of its chemical composition. This survey was carried out with an energy resolution of 1 eV. Subsequently, detailed spectra were collected for the Er 5p peaks (5p3/2 and 5p1/2) in all selected samples using a higher energy resolution of 0.1 eV. All peaks were calibrated with respect to the C 1s peak, which was fixed at 284.4 eV. The procedure followed was as described in [28].

4. Results and Discussion

4.1. Crystallization Behavior

The samples analyzed in this study—labeled A, B, C, and D—consist of various compositions of SiO2, TiO2, PO2.5, and ErO1.5, selected for their established crystallization behavior and known phase formations during thermal treatments, as highlighted in prior research [1]. Table 2 provides a comparative overview of the crystallization phases observed in these samples following heat treatment. Sample A (80SiO2–20TiO2–2 mol% ErO1.5) remains amorphous at 900 °C (A1) but undergoes a phase transformation to a mixture of rutile (R) and erbium titanate (ETO) at 1000 °C (A2). Sample B (80SiO2–20TiO2–10 mol% PO2.5 and 2 mol% ErO1.5) retains an amorphous state at 900 °C (B1) and forms anatase (A) and erbium phosphate (EPO) at 1000 °C (B2). The presence of phosphorus likely facilitates the crystallization of the anatase and rutile phases, which may coexist.
In contrast, sample C (98SiO2–2TiO2–10 mol% PO2.5 and 2 mol% ErO1.5), with a lower TiO2 content, exhibits the crystallization of erbium phosphate (EPO), even at 900 °C (C1), and EPO becomes the dominant phase at 1000 °C (C2). Finaly, sample D (98SiO2–2TiO2 with 2 mol% ErO1.5) forms only erbium titanate (ETO) crystallites at both 900 °C and 1000 °C.
The X-ray diffraction (XRD) patterns of the samples are presented in Figure 2. Panel (a) shows the XRD data for sample A at 900 °C (A1) and 1000 °C (A2), highlighting the phase transitions observed in this composition. Panel (b) displays the XRD patterns for sample D at 900 °C (D1) and 1000 °C (D2), confirming the crystallization of erbium titanate (ETO) at both temperatures.
The XRD patterns in panel (a) of Figure 2 show a peak at approximately 27.5°, which corresponds to the (110) plane of rutile TiO2 (ICDD card no. 00-021-1276), confirming the presence of the rutile phase, a feature commonly observed in annealed sol–gel-derived ST glassy films [1]. Additional peaks were identified as belonging to Er2Ti2O7 (Erbium Titanate, ETO), rather than Er2O3 (Erbium Oxide) or metallic Ti (Titanium). Specifically, the peaks at 2θ values of 29.6°, 30.8°, 35.9°, and 39° correspond to the crystallographic planes (311), (222), (400), and (331) of Er2Ti2O7, as documented in ICDD card no. 01-073-1647, and are visible in both panels (a) and (b). As the annealing temperature increases, there is no significant shift in the position of the diffraction peaks, but the peak intensities increase notably, especially between 900 °C and 1000 °C. This increase in intensity suggests enhanced crystallinity and greater phase stability at higher temperatures. The sharp intensity of the (222) peak at 2θ = 29.6° specifically indicates a well-ordered matrix structure. This observation aligns with the findings from [9], which show that higher temperatures enhance atomic mobility, facilitating ion rearrangement into a more stable crystalline structure, as evidenced by sharper diffraction peaks. While the (222) peak shows a marked intensity increase, other weaker peaks exhibit a more modest rise with thermal treatment. This supports the conclusion that Er2Ti2O7 is the dominant crystalline phase, rather than isolated erbium oxide or titanium [29,30,31]. The absence of peaks corresponding to SiO2 and TiO2 indicates that these components remain amorphous, consistent with the lack of anatase or rutile phases at either temperature, as shown in Table 2. This further supports the conclusion that in the D1 and D2 samples, Er2Ti2O7 is the primary crystalline phase formed.
During the doping process, Er3+ ions likely bond with Ti in the presence of ambient oxygen, promoting the formation of erbium titanate (ETO). When Er and Ti are present in balanced proportions (e.g., sample D, 2% Er:2% Ti), only ETO crystallites form. This occurs because equal proportions of Er3+ and Ti4+ ions provide the stoichiometric balance and coordination environment required for ETO formation, in line with Pauling’s rules for ionic crystal stability [32]. Specifically, the balance between Er3+ and Ti4+ ions ensures proper coordination and charge neutrality, minimizing structural disorder. In this balanced system, each Er3+ ion is adequately coordinated by Ti4+ ions, stabilizing the crystalline lattice and following Pauling’s Coordination Principle (Rule 1), which describes how cations and anions arrange themselves for stable coordination.
In compositions where Ti > Er (e.g., sample A, 2% Er:20% Ti), a mixture of ETO and rutile (TiO2) phases is observed. The excess Ti4+ ions favor the formation of rutile, while some Er3+ ions remain dispersed within the amorphous matrix. This dispersion can be attributed to their role as network modifiers, disrupting the glass structure by breaking bridging oxygens (BOs) and forming non-bridging oxygens (NBOs) [33]. Although Er3+ ions form strong bonds with oxygen [34], these bonds alone are insufficient to counteract the structural dominance of TiO2, leading to the coexistence of rutile and dispersed Er3+ ions. The formation of rutile over ETO in Ti-rich compositions aligns with Pauling’s Electrostatic Valency Principle (Rule 2), where the excess Ti4+ ions dominate the lattice, creating a charge imbalance that destabilizes Er3+ coordination. These observations are consistent with previous studies [35], which emphasize the interplay between Ti4+ content and the stabilization of Er3+ crystalline structures.
The nanosized features of the investigated samples are clearly highlighted in the TEM micrographs presented in Figure 3, which provide bright-field images revealing the dispersion of nanoparticles within the samples. Panel (a) shows the microstructure of sample A annealed at 1000 °C (A2), while panel (b) displays sample D annealed at 1000 °C (D2).
The bright-field TEM images exhibit a strong contrast, with dark nanocrystals visible against the lighter amorphous matrix, confirming the presence of distinct nanocrystalline structures. Both panels reveal the predominance of spherical nanocrystals, which are relatively uniform in size, sub-100 nm, and exhibit some degree of clustering. In addition to the spherical particles, a smaller population of rod-shaped nanocrystals is evident, appearing elongated and less frequent. These rods are also sub-100 nm in size, suggesting a variation in growth mechanism or crystal orientation. These observations indicate that the annealing process at 1000 °C facilitated the nucleation and stabilization of spherical nanocrystals as the dominant phase, while the rod-shaped structures represent a minority phase, likely influenced by specific local conditions or compositional variations. The observed distribution and morphology are consistent with the processing conditions and annealing temperatures applied to the samples.

4.2. Nanostructure Characterization via SAXS

The SAXS measurements provided data on the size and distribution of nanocrystals in the sol–gel-derived thin films. Guinier plots were used to calculate the Guinier radius, which estimates the average size of the nanocrystals (Figure 4). The Guinier radii values for the various compositions, as summarized in Table 3, show clear trends across the different compositions and thermal treatments.
The Guinier radii for all samples, extracted from the SAXS measurements (Figure 4), ranged from 16.9 nm to 18.9 nm, indicating the formation of nanocrystals after thermal treatment. Samples A1 and A2 exhibited the smallest Rg values, with A1 showing the smallest overall Rg (16.90 nm). Samples B1 and B2 showed intermediate Rg values, which were slightly larger for the 1000 °C annealing condition (B2, 18.03 nm). Sample C (C1 and C2) displayed the largest Rg values, with C2 reaching the highest value (18.93 nm), suggesting that the addition of phosphorus facilitated more effective nanocrystal growth. Sample D (D1 and D2) also exhibited intermediate Rg values, with D1 being slightly larger (18.50 nm) than D2 (17.31 nm). These results highlight the impact of annealing temperature and composition on nanocrystal growth. The increasing Rg values observed for the samples annealed at 1000 °C, as compared to those treated at 900 °C, suggest enhanced crystallization at higher temperatures. The larger nanocrystal sizes observed in sample C may indicate that phosphorus addition promotes more effective nanocrystal growth. These trends are consistent with previous studies [36], which confirm the formation of nanocrystalline phases and further support the findings presented here.

4.3. Interpretation of the β/α Ratio

The ASAXS measurements were conducted near the LIII absorption edge of Er (8357 eV), which allowed for the isolation of Er3+’s contribution to nanocrystal formation. Since the K-edges of other elements in the matrix, such as phosphorus (P) at around 2153 eV, silicon (Si) at approximately 1840 eV, and titanium (Ti) near 4966 eV [24], were significantly far from the Er LIII edge, their contributions did not interfere with the analysis of Er, ensuring that the observed scattering data exclusively reflected the behavior of Er3+ within the nanostructures.
The parameters α, β, and the ratio β/α were extracted as outlined in Equation (7), providing information about the distribution of Er3+ within the nanocrystals and its interactions with other matrix components. The β/α ratio directly reflects how Er3+ is localized in the nanostructures, serving as a diagnostic for its concentration in the nanocrystals.
For sample A1, the higher β/α ratio suggests that Er3+ ions are more evenly dispersed throughout the amorphous matrix rather than concentrated within the nanocrystals, consistent with XRD results indicating an amorphous phase for this sample. This diffuse distribution of Er3+ lowers the electron density contrast between Er3+ and the surrounding matrix, resulting in a weaker and more gradual increase in scattering intensity, as shown in Figure 5. Consequently, the X-ray scattering effect is less pronounced, as the lack of concentrated Er3+ clusters reduces the sharpness of the scattering contrast. In contrast, sample A2 exhibits a lower β/α ratio, indicating a higher concentration of Er3+ within the nanocrystals. This concentrated presence of Er3+ within nanocrystals increases the electron density contrast, leading to a sharper scattering contrast and stronger intensity near the absorption edge. These observations align with the XRD results, which reveal a more substantial presence of Er3+ in the crystalline phases for sample A2.
The β/α ratios for sample pairs C (C1 and C2) and D (D1 and D2) remain nearly constant, as supported by the data in Figure 6 and Figure 7, with values comparable to those of sample A2. This stability, regardless of thermal variation, suggests a consistent concentration of Er3+ within the nanocrystals. The lower β/α ratios indicate the efficient incorporation of Er3+ into specific nanocrystalline phases, leading to a denser distribution of Er3+ within the crystalline structure rather than remaining in the amorphous matrix. This behavior likely points to a saturation effect, where the Er3+ concentration in the nanocrystals reaches a threshold and remains stable despite further thermal treatment.
The XRD results support this by confirming the presence of Er3+-containing crystalline phases in these samples. For the samples in series C, erbium phosphate (EPO) is observed, while erbium titanate (ETO) forms in samples from series D. The phase distinction suggests that Er3+ incorporation is composition-dependent, with phosphorus inhibiting the formation of ETO and favoring EPO in the samples with a 2% Er:2% Ti ratio. This selective phase formation is consistent with the stable β/α ratios observed in series C, where phosphorus promotes a more uniform and concentrated Er3+ distribution within the EPO phase.
In the B samples (Figure 8), the β/α ratio remains relatively stable between B1 (900 °C) and B2 (1000 °C), with a slight increase observed in B2. This suggests a mild tendency for Er3+ to disperse in the matrix at higher temperatures, although the effect is less pronounced than in sample A1. The intermediate β/α ratios in B1 and B2—lower than A1’s but higher than those of A2 and pairs C and D—indicate that Er3+ tends to remain in the matrix. However, the interaction between phosphorus and the high Ti content (20% Ti with 2% Er) in B2 seems to favor the crystallization of TiO2 phases, such as rutile, which competes with the incorporation of Er3+ into erbium phosphate (EPO) nanocrystals, thus influencing the distribution of Er3+. Interestingly, this competition is not fully reflected in the β/α ratio of 86 observed in B2, which is slightly higher than that of the other samples containing erbium crystallites.
Additionally, discrepancies arise between the XRD and ASAXS results, particularly for sample B1, which remains amorphous. Despite this, the β/α ratio of 73 for B1 is not comparable to that of sample A1, which underwent similar conditions. While XRD detects the crystallization of rutile in A2 and B1, ASAXS does not, as rutile lacks Er3+. This discrepancy can be attributed to the different detection principles of SAXS and ASAXS. SAXS depends primarily on the size and shape of nanostructures rather than their atomic structure, enabling it to detect features such as the formation of Er3+-free phases like rutile in B1, beyond just the ETO and EPO phases.
Notably, SAXS reveals a radius of gyration (Rg) of 16.9 nm for A1 and 18.02 nm for B1, even though both samples lack Er3+-rich precipitates or other crystalline phases. Given the sol–gel synthesis of these samples, these Rg values may correspond to residual porosity within the amorphous network rather than to crystalline precipitates. This suggests that SAXS is sensitive to early structural features, such as porosity, that may not be detectable by ASAXS, which is limited to detecting features containing Er3+.
This interpretation aligns with the findings in the literature; for example, reference [9] describes ETO films that retain sponge-like porosity even after high-temperature treatment, with pores around 10 nm persisting in thin film areas. This residual porosity, typical of sol–gel-derived films [8,37,38], could account for the SAXS-detected features in A1 and B1, underscoring SAXS’s ability to reveal subtle structural features such as pores. Further supporting this interpretation, a study in [25] examined activated carbon samples using SAXS. Similarly to our observations, SAXS detected residual porosity within the matrix of the carbon samples. The study reported spherical pores with sizes around 6 nm, suggesting that SAXS can reveal porosity in materials even in the absence of crystalline phases. This reinforces the idea that SAXS is sensitive to the structural features of the matrix, such as porosity, which are not necessarily captured by ASAXS, which focuses on crystalline phases.

4.4. X-Ray Photoelectron Spectroscopy (XPS) Analysis

X-ray photoelectron spectroscopy (XPS) is a crucial technique for probing the chemical state and electronic structure of Er in oxide materials, particularly for analyzing its 5p orbitals. These orbitals split into 5p3/2 and 5p1/2 due to spin–orbit coupling, resulting in two distinct binding energy peaks [39]. These peaks are sensitive to the Er’s chemical environment, which can vary in complex oxide matrices.
Erbium (Er) naturally prefers the +3 oxidation state due to its electron configuration. Like other rare earth elements, it has three electrons outside a stable, filled inner shell. By losing these three electrons, erbium attains a stable electronic configuration, making Er3+ the most chemically stable form. This is particularly true in oxygen-rich environments, such as oxides and silicates, which are commonly found in sol–gel glasses. The highly electronegative oxygen atoms in the glass matrix stabilize Er3+ because oxygen’s strong affinity for electrons makes further reductions unlikely. Additionally, the oxide network “locks” Er into the +3 oxidation state by forming stable Er–O bonds. To explore how changes in crystallinity and composition impact the electronic structure of Er3+, XPS was performed on four representative samples: A1, A2, D1, and D2. These samples were chosen to span a range of structural environments, from fully amorphous matrices to those containing the crystalline phases of TiO2 (rutile) and erbium titanate (Er2Ti2O7, ETO). Sample A1 was completely amorphous, A2 contained a mixture of rutile and ETO crystallites, D1 showed partial crystallization with remnants of ETO crystallites, and D2 contained well-developed nanocrystalline ETO within a glass matrix. This selection allowed for a comparative analysis of how varying crystallinity and phase composition influence the electronic environment and oxidation states of Er3+.
In general, the Er 5p3/2 and 5p1/2 peaks typically appear around 24.7 eV and 31.4 eV, respectively, as documented in the X-ray Data Booklet by Lawrence Berkeley National Laboratory [40]. Studies on Er-doped silicon have reported similar but slightly higher binding energies, with the 5p3/2 peak at approximately 26 eV and the 5p1/2 peak at 32 eV, due to variations in the local environment and matrix interactions [41]. These values provide a reference range for comparing Er-containing materials of varying crystallinity and coordination environments, as seen in our samples.
In panel (a) of Figure 9, a comprehensive survey scan of sample A1 was conducted, uncovering the key elements present, including the distinctive Er 5p peak. Building upon this initial finding, a more detailed XPS analysis was then performed, focusing specifically on the Er 5p peak (panel (b)) to delve deeper into the electronic properties of this element within the samples. The XPS spectra of the Er 5p3/2 and 5p1/2 peaks are displayed for samples A1, A2, D1, and D2, each showcasing the unique structural characteristics that significantly influence the local environment surrounding Er3+. Changes in the binding energies can reflect shifts in the chemical environment, crystallinity, oxidation state, and local bonding structure of Er in the different samples. For sample A1, the 5p3/2 and 5p1/2 peaks are broad, indicating the disordered environment of the amorphous matrix where Er atoms are randomly coordinated with oxygen. This broadening is indicative of the variability in local bonding, leading to a wider range of binding energies around the reference values. Additionally, the peaks are shifted to lower binding energies compared to other samples, reflecting the reduced electron density and weaker Er–O interactions in the disordered structure of A1. The broad peaks and lower binding energy suggest a less stable electronic environment around Er, and a weaker impact of spin–orbit coupling, as the disordered matrix likely diminishes the splitting of the 5p peaks. In contrast, sample A2, which contains both rutile and ETO crystallites, shows sharper 5p peaks, suggesting a more ordered local environment around the Er atoms. The binding energies observed in A2 are slightly higher than those in A1, indicating that the presence of crystalline phases, especially ETO, stabilizes the local electronic environment around Er3+, as the ordered lattice structure stabilizes the local electronic density around it, reducing the range of binding energy variations. The sharper peaks and higher binding energies in A2 suggest enhanced spin–orbit coupling, with better defined 5p3/2 and 5p1/2 peaks. Sample D1, which exhibits partial crystallization with remnants of ETO crystallites, shows 5p peaks that are partially resolved, lying between the broad profile of A1 and the sharper profile of A2. The binding energies in D1 are higher than in A1 but show slight variations due to the presence of both amorphous and crystalline regions. This partial crystallization introduces greater order into the structure, though some disordered regions still contribute to broader peaks. The higher binding energy in D1 compared to A1 and A2 may reflect a more stable oxidation state of Er3+, as the local environment around the ions becomes more crystalline.
Finally, sample D2, which contains well-developed nanocrystalline ETO within a glass matrix, exhibits the sharpest and most defined peaks of all samples, with the highest binding energies. The nanocrystalline environment around Er3+ ions provides a highly ordered structure, leading to consistent bonding with oxygen and, consequently, narrow XPS peaks due to the stable Er–O interactions.
This trend in binding energies (A1 < A2 < D1 < D2) reflects the progression from amorphous to increasingly crystalline matrices, accompanied by a rise in local structural order and stability in the oxidation state around Er3+ ions.

4.5. Corroboration with ASAXS and SAXS Analysis

The ASAXS and SAXS analyses corroborate the XPS findings by providing structural evidence for the varying degrees of order and crystallinity in each sample. The gradual increase in local structural order from A1 (fully amorphous) to D2 (nanocrystalline) is consistently reflected in both the scattering data and the XPS spectra, with each technique reinforcing the conclusion that crystallinity stabilizes the Er3+ electronic environment. This structural stability leads to narrower, higher binding energy peaks in the XPS spectra, underscoring the relationship between crystallinity, Er–O interaction strength, and electronic stability in sol–gel-derived glasses. This combined analysis provides valuable insights into the structural and chemical evolution of Er3+ within these materials as crystallinity increases.

5. Conclusions

This study provided a detailed analysis of erbium-doped silica-based glass-ceramic thin films synthesized via the sol–gel method, emphasizing the formation, dispersion, and stabilization of erbium-containing nanocrystals under varying thermal treatments. Advanced characterization techniques, including anomalous small-angle X-ray scattering (ASAXS), small-angle X-ray scattering (SAXS), X-ray photoelectron spectroscopy (XPS), and transmission electron microscopy (TEM), were employed to investigate the structural and chemical properties of the films.
The ASAXS results revealed distinct differences in the distribution of Er3+ ions across compositions and thermal conditions. The β/α ratios highlighted a higher concentration of Er3+ ions within nanocrystals for samples treated at higher temperatures and with specific dopant compositions, confirming their migration into crystalline phases such as erbium phosphate (EPO) and erbium titanate (ETO). The SAXS and Guinier analyses quantified the nanocrystal sizes, with the radii of gyration ranging from 16.9 nm to 18.9 nm, demonstrating the influence of phosphorus addition and annealing temperature on nanocrystal growth. TEM further validated these findings, showcasing the morphology and dispersion of spherical and rod-shaped nanocrystals within the amorphous matrix.
XPS provided detailed insights into the chemical environment and stabilization of Er3+ ions, confirming their incorporation into crystalline structures and their interaction with the surrounding matrix. The results demonstrated that thermal treatments and compositional adjustments play a crucial role in controlling the nanoscale organization and behavior of these systems.
This work advances the understanding of rare earth-doped glass-ceramics, offering a deeper understanding of the size, distribution, and interaction of erbium-containing nanocrystals within the amorphous matrix. The findings emphasize the importance of optimizing the composition and thermal treatment to achieve desirable nanostructural and chemical properties. These results lay the groundwork for the development of tailored glass-ceramic materials for advanced technological applications.

Author Contributions

Conceptualization, H.C.V., L.S. and R.Ö.; methodology, H.C.V. and L.S.; software, H.C.V. and L.S.; validation, M.M. and R.Ö.; investigation, M.M. and H.C.V.; writing—original draft preparation, R.Ö., M.M. and H.C.V.; writing—review and editing, H.C.V. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Strohhöfer, C.; Fick, J.; Vasconcelos, H.; Almeida, R. Active optical properties of Er-containing crystallites in sol–gel derived glass films. J. Non-Crystalline Solids 1998, 226, 182–191. [Google Scholar] [CrossRef]
  2. Lahoz, F.; Capuj, N.; Haro-González, P.; Martín, I.R.; Pérez-Rodríguez, C.; Cáceres, J.M. Stimulated emission in the red, green, and blue in a nanostructured glass ceramics. J. Appl. Phys. 2011, 109, 043102. [Google Scholar] [CrossRef]
  3. Jiang, B.; Hu, Y.; Ren, L.; Zhou, H.; Shi, L.; Zhang, X. Four- and five-photon upconversion lasing from rare earth elements under continuous-wave pump and room temperature. Nanophotonics 2022, 11, 4315–4322. [Google Scholar] [CrossRef] [PubMed]
  4. Hossain, M.K.; Ahmed, M.H.; Khan, I.; Miah, M.S.; Hossain, S. Recent Progress of Rare Earth Oxides for Sensor, Detector, and Electronic Device Applications: A Review. ACS Appl. Electron. Mater. 2021, 3, 4255–4283. [Google Scholar] [CrossRef]
  5. Vasconcelos, H.C.; Pinto, A.S. Fluorescence Properties of Rare-Earth-Doped Sol-Gel Glasses; IntechOpen: London, UK, 2017. [Google Scholar] [CrossRef]
  6. Zhao, W.; Jia, H.; Qu, J.; Yang, C.; Wang, Y.; Zhu, J.; Wu, H.; Liu, G. Sol-gel synthesis of TiO2-SiO2 hybrid films with tunable refractive index for broadband antireflective coatings covering the visible range. J. Sol-Gel Sci. Technol. 2023, 107, 105–121. [Google Scholar] [CrossRef]
  7. Almeida, R.M.; Morais, P.J.; Vasconcelos, H.C. Optical loss mechanisms in nanocomposite sol-gel planar waveguides. In Sol-Gel Optics IV; SPIE: Bellingham, WA, USA, 1997; Volume 3136. [Google Scholar] [CrossRef]
  8. Vasconcelos, H.; Meirelles, M.; Rivera-López, F. Erbium photoluminescence response related to nanoscale heterogeneities in sol-gel silicates. J. Rare Earths 2013, 31, 18–26. [Google Scholar] [CrossRef]
  9. Langlet, M.; Coutier, C.; Fick, J.; Audier, M.; Meffre, W.; Jacquier, B.; Rimet, R. Sol–gel thin film deposition and characterization of a new optically active compound: Er2Ti2O7. Opt. Mater. 2001, 16, 463–473. [Google Scholar] [CrossRef]
  10. Lahoz, F.; Pérez-Rodríguez, C.; Halder, A.; Das, S.; Paul, M.C.; Pal, M.; Bhadra, S.K.; Vasconcelos, H.C. Complete energy transfer due to rare-earth phase segregation in optical fiber preform glasses. J. Appl. Phys. 2011, 110, 083121. [Google Scholar] [CrossRef]
  11. Forster, T. Intermolecular Energy Migration and Fluorescence. Ann. Phys. 1948, 2, 55–75. [Google Scholar]
  12. Dexter, D.L. A Theory of Sensitized Luminescence in Solids. J. Chem. Phys. 1953, 21, 836–850. [Google Scholar] [CrossRef]
  13. Abdullah, A.; Benchafia, E.M.; Choi, D.; Abedrabbo, S. Synthesis and Characterization of Erbium-Doped Silica Films Obtained by an Acid–Base-Catalyzed Sol–Gel Process. Nanomaterials 2023, 13, 1508. [Google Scholar] [CrossRef] [PubMed]
  14. Gao, Y.; Shen, H.; Cao, J.; Li, D.; Yang, D. Control of the formation and luminescent properties of polymorphic erbium silicates on silicon. Opt. Mater. Express 2019, 9, 1716–1727. [Google Scholar] [CrossRef]
  15. Li, T.; Senesi, A.J.; Lee, B. Small Angle X-ray Scattering for Nanoparticle Research. Chem. Rev. 2016, 116, 11128–11180. [Google Scholar] [CrossRef] [PubMed]
  16. Hoell, A.; Raghuwanshi, V.S.; Bocker, C.; Herrmann, A.; Rüssel, C.; Höche, T. Crystallization of BaF2 from droplets of phase separated glass—Evidence of a core-shell structure by ASAXS. CrystEngComm 2020, 22, 5031–5039. [Google Scholar] [CrossRef]
  17. Gericke, E.; Wallacher, D.; Wendt, R.; Greco, G.; Krumrey, M.; Raoux, S.; Hoell, A.; Mascotto, S. Direct Observation of the Xenon Physisorption Process in Mesopores by Combining In Situ Anomalous Small-Angle X-ray Scattering and X-ray Absorption Spectroscopy. J. Phys. Chem. Lett. 2021, 12, 4018–4023. [Google Scholar] [CrossRef] [PubMed]
  18. Haas, S.; Hoell, A.; Wurth, R.; Rüssel, C.; Boesecke, P.; Vainio, U. Analysis of nanostructure and nanochemistry by ASAXS: Accessing phase composition of oxyfluoride glass ceramics doped with Er3+/Yb3+. Phys. Rev. B 2010, 81, 184207. [Google Scholar] [CrossRef]
  19. Santos, L.F.; Barros Marques, M.I.; Vasconcelos, H.C.; Almeida, R.M.; Lyon, O. Caracterização de Filmes Finos por SAXS e ASAXS. Física 2000 (Oral Presentation): 12ª Conferência nacional de Física: 10º Encontro Ibérico Para o Ensino da Física, Sociedade Portuguesa de Física, 27–30 September 2000, Figueira da Foz, Portugal. Abstract. 2000. Available online: https://www.yumpu.com/pt/document/read/5920105/programa-nautilus-universidade-de-coimbra (accessed on 12 November 2024).
  20. Sun, Y. Anomalous small-angle X-ray scattering for materials chemistry. Trends Chem. 2021, 3, 1045–1060. [Google Scholar] [CrossRef]
  21. Hoell, A.; Varga, Z.; Raghuwanshi, V.S.; Krumrey, M.; Bocker, C.; Rüssel, C. ASAXS study of CaF2 nanoparticles embedded in a silicate glass matrix. J. Appl. Crystallogr. 2014, 47, 60–66. [Google Scholar] [CrossRef]
  22. Ramos, A.; Lyon, O.; Levelut, C. Stoichiometry of Cd(S,Se) nanocrystals by anomalous small-angle x-ray scattering. J. Appl. Phys. 1995, 78, 6916–6922. [Google Scholar] [CrossRef]
  23. Sazaki, S. Anomalous Scattering Factors for Synchrotron Radiation Users. In Calculated Using Cromer and Liberman’s Method, KEK Report; National Lab. for High Energy Physics (Tsukuba, Japan): Tsukuba, Japan, 1984; pp. 82–83. [Google Scholar]
  24. Skuld. (n.d.). AS Periodic. University of Washington. Retrieved 10 November 2024. Available online: http://skuld.bmsc.washington.edu/scatter/AS_periodic.html (accessed on 10 November 2024).
  25. Blachnio, M.; Derylo-Marczewska, A.; Charmas, B.; Zienkiewicz-Strzalka, M.; Bogatyrov, V.; Galaburda, M. Activated Carbon from Agricultural Wastes for Adsorption of Organic Pollutants. Molecules 2020, 25, 5105. [Google Scholar] [CrossRef] [PubMed]
  26. Lesieur, P.; Lombardo, D.; Beauche, C.; Creoff, T.; Decamps, J.M.; Dubuisson, G.; Perilhous, S.; Lesieur, G.; Keller, M.; Ollivon, M. New Features at the LURE-D22 SAXS Beamline. Proceedings of the International School and Symposium on Small Angle Scattering (p. 62). Matrahaza, Hungary: Neutron Physics Department, Research Institute for Solid State Physics and Optics. Hosted by the Hungarian Academy of Sciences. 1998. Available online: https://inis.iaea.org/collection/NCLCollectionStore/_Public/30/036/30036868.pdf (accessed on 10 November 2024).
  27. Kutsch, B.; Lyon, O.; Schmitt, M.; Mennig, M.; Schmidt, H. Small-Angle X-ray Scattering Experiments in Grazing Incidence on Sol–Gel Coatings Containing Nano-Scaled Gold Colloids: A New Technique for Investigating Thin Coatings and Films. J. Appl. Cryst. 1997, 30, 948–956. [Google Scholar] [CrossRef]
  28. Gonçalves, M.C.; Pereira, J.C.; Matos, J.C.; Vasconcelos, H.C. Photonic Band Gap and Bactericide Performance of Amorphous Sol-Gel Titania: An Alternative to Crystalline TiO2. Molecules 2018, 23, 1677. [Google Scholar] [CrossRef] [PubMed]
  29. Bakhsh, A.; Maqsood, A. Sintering effects on structure, morphology, and electrical properties of sol-gel synthesized, nano-crystalline erbium oxide. Electron. Mater. Lett. 2012, 8, 605–608. [Google Scholar] [CrossRef]
  30. Bolzoni, L.; Yang, F. X-ray diffraction for phase identification in Ti-based alloys: Benefits and limitations. Phys. Scr. 2024, 99, 065024. [Google Scholar] [CrossRef]
  31. Gupta, S.K.; Sudarshan, K.; Jena, P.; Ghosh, P.S.; Yadav, A.K.; Jha, S.N.; Bhattacharyya, D. Bright aspects of defects and dark traits of dopants in the photoluminescence of Er2X2O7:Eu3+ (X = Ti and Zr) pyrochlore: An insight using EXAFS, positron annihilation and DFT. Mater. Adv. 2021, 2, 3075–3087. [Google Scholar] [CrossRef]
  32. Pauling, L. The principles determining the structure of complex ionic crystals. J. Am. Chem. Soc. 1929, 51, 1010–1026. [Google Scholar] [CrossRef]
  33. Bouabdalli, E.M.; El Jouad, M.; Touhtouh, S.; Chellakhi, A.; Hajjaji, A. Synthesis, structural, and optical behavior of erbium-doped silicophosphate glasses for photonics applications. Luminescence 2024, 39, e4802. [Google Scholar] [CrossRef] [PubMed]
  34. Kumar, B.K.; Babu, P.R.; Kavaz, E.; Kshetri, Y.K.; Kim, T.-H.; Anjos, V.d.C.d.; Raju, B.D.P. Spectroscopic characteristics and gain cross-section profiles of erbium-doped boro-tellurite glasses for optical amplifier applications. Opt. Mater. 2024, 148, 114950. [Google Scholar] [CrossRef]
  35. Coutier, C.; Audier, M.; Fick, J.; Rimet, R.; Langlet, M. Aerosol–gel preparation of optically active layers in the system Er/SiO2–TiO2. Thin Solid Films 2000, 372, 177–189. [Google Scholar] [CrossRef]
  36. Santos, L.; Vasconcelos, H.; Marques, M.B.; Almeida, R.M. Active Nanocrystals in Erbium-Doped Silica-Titania Sol-Gel Films. Mater. Sci. Forum 2004, 455-456, 545–549. [Google Scholar] [CrossRef]
  37. Vasconcelos, H.C. Optical Waveguides Based on Sol-Gel Coatings. In Waveguide Technologies in Photonics and Microwave Engineering; Steglich, P., Ed.; IntechOpen: London, UK, 2020; ISBN 978-1-83968-189-9. [Google Scholar]
  38. Orignac, X.; Vasconcelos, H.C.; Du, X.M.; Almeida, R.M. Influence of solvent concentration on the microstructure of SiO2−TiO2 sol-gel films. J. Sol-Gel Sci. Technol. 1997, 8, 243–248. [Google Scholar] [CrossRef]
  39. Bagus, P.S.; Nelin, C.J.; Al-Salik, Y.; Ilton, E.S.; Idriss, H. Multiplet splitting for the XPS of heavy elements: Dependence on oxidation state. Surf. Sci. 2016, 643, 142–149. [Google Scholar] [CrossRef]
  40. X-RAY DATA BOOKLET, Center for X-Ray Optics and Advanced Light Source. Lawrence Berkeley National Laboratory. Available online: https://xdb.lbl.gov/ (accessed on 4 November 2024).
  41. Tessler, L.R.; Piamonteze, C.; Iniguez, A.C.; de Siervo, A.; Landers, R.; Morais, J. UPS of a-Si:H: What is the energy of the Er 4f states? MRS Proc. 2000, 609, 111. [Google Scholar] [CrossRef]
Figure 1. Dependences of the real part (f′) and imaginary part (f″) of the X-ray scattering factor (f) on the X-ray energy for Er. The absorption edges displayed in the range of 6–10 keV correspond to the LIII, LII, and LI edges of the Er. The vertical transitions at these edges highlight the abrupt changes in both f′ and f′′ which contribute to the anomalous effect observed in ASAXS measurements. Figure 1 was obtained with the help of resources from the University of Washington [24].
Figure 1. Dependences of the real part (f′) and imaginary part (f″) of the X-ray scattering factor (f) on the X-ray energy for Er. The absorption edges displayed in the range of 6–10 keV correspond to the LIII, LII, and LI edges of the Er. The vertical transitions at these edges highlight the abrupt changes in both f′ and f′′ which contribute to the anomalous effect observed in ASAXS measurements. Figure 1 was obtained with the help of resources from the University of Washington [24].
Foundations 05 00005 g001
Figure 2. X-ray diffraction (XRD) patterns of the samples. (a) XRD of sample A annealed at 900 °C (A1) and 1000 °C (A2); and (b) XRD of sample D annealed at 900 °C (D1) and 1000 °C (D2).
Figure 2. X-ray diffraction (XRD) patterns of the samples. (a) XRD of sample A annealed at 900 °C (A1) and 1000 °C (A2); and (b) XRD of sample D annealed at 900 °C (D1) and 1000 °C (D2).
Foundations 05 00005 g002
Figure 3. TEM micrographs of samples annealed at 1000 °C, observed in bright-field mode. (a) Microstructure of sample A (A2) showing predominantly spherical nanocrystals with some clustering, along with a few rod-shaped structures. (b) Microstructure of sample D (D2), similarly exhibiting predominantly spherical nanocrystals and sparsely distributed rod-shaped particles. Both images are presented at a 100 nm scale. (c) Diffraction pattern of sample A (A2), displaying crystalline features with spots corresponding to the cubic symmetry of the ETO phase. (d) Diffraction pattern of sample D (D2), showing discrete diffraction spots consistent with a cubic symmetry (space group Fd 3 ¯ m), indicative of the Er2Ti2O7 (ETO) phase.
Figure 3. TEM micrographs of samples annealed at 1000 °C, observed in bright-field mode. (a) Microstructure of sample A (A2) showing predominantly spherical nanocrystals with some clustering, along with a few rod-shaped structures. (b) Microstructure of sample D (D2), similarly exhibiting predominantly spherical nanocrystals and sparsely distributed rod-shaped particles. Both images are presented at a 100 nm scale. (c) Diffraction pattern of sample A (A2), displaying crystalline features with spots corresponding to the cubic symmetry of the ETO phase. (d) Diffraction pattern of sample D (D2), showing discrete diffraction spots consistent with a cubic symmetry (space group Fd 3 ¯ m), indicative of the Er2Ti2O7 (ETO) phase.
Foundations 05 00005 g003
Figure 4. Guinier plots showing linear regions for the spherical nanocrystals, with the calculation of the gyration radius (Rg), based on the SAXS data for samples A1, A2, B1, B2, C1, C2, D1, and D2. The experimental SAXS data (circles) are fitted to the reference model (lines) based on Equation (8), enabling the evaluation of the Rg for the samples.
Figure 4. Guinier plots showing linear regions for the spherical nanocrystals, with the calculation of the gyration radius (Rg), based on the SAXS data for samples A1, A2, B1, B2, C1, C2, D1, and D2. The experimental SAXS data (circles) are fitted to the reference model (lines) based on Equation (8), enabling the evaluation of the Rg for the samples.
Foundations 05 00005 g004
Figure 5. ASAXS data for samples A1 (900 °C) and A2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E . Sample A2 has a steeper slope, indicating a higher Er3+ concentration within nanocrystals, while sample A1’s flatter slope suggests that Er3+ is more dispersed in the matrix. Adapted from [19].
Figure 5. ASAXS data for samples A1 (900 °C) and A2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E . Sample A2 has a steeper slope, indicating a higher Er3+ concentration within nanocrystals, while sample A1’s flatter slope suggests that Er3+ is more dispersed in the matrix. Adapted from [19].
Foundations 05 00005 g005
Figure 6. ASAXS data for samples C1 (900 °C) and C2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E , demonstrating a consistent β/α ratio. This indicates that the concentration of Er3+ in the nanocrystals is approximately the same across both temperatures, supporting the presence of stable Er3+ distributions in both samples. Adapted from [19].
Figure 6. ASAXS data for samples C1 (900 °C) and C2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E , demonstrating a consistent β/α ratio. This indicates that the concentration of Er3+ in the nanocrystals is approximately the same across both temperatures, supporting the presence of stable Er3+ distributions in both samples. Adapted from [19].
Foundations 05 00005 g006
Figure 7. ASAXS data for samples D1 (900 °C) and D2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E . The results demonstrate a similar β/α ratio to the sample pairs of C. Adapted from [19].
Figure 7. ASAXS data for samples D1 (900 °C) and D2 (1000 °C), showing scattering intensity ( Σ I ) 1 / 2 versus f E . The results demonstrate a similar β/α ratio to the sample pairs of C. Adapted from [19].
Foundations 05 00005 g007
Figure 8. ASAXS data for samples B1 (900 °C) and B2 (1000 °C) showing scattering intensity ( Σ I ) 1 / 2 versus f E , and demonstrating a high β/α ratio comparatively to the sample pairs of C, D, and sample A2. Adapted from [19].
Figure 8. ASAXS data for samples B1 (900 °C) and B2 (1000 °C) showing scattering intensity ( Σ I ) 1 / 2 versus f E , and demonstrating a high β/α ratio comparatively to the sample pairs of C, D, and sample A2. Adapted from [19].
Foundations 05 00005 g008
Figure 9. XPS spectra showing (a) a survey scan of sample A1, identifying key peaks including OKLL, O 1s, C 1s, Ti 2p, Si 2s, Si 2p, and Er 5p; and (b) a detailed comparison of the Er 5p1/2 and Er 5p3/2 peaks for samples A1, A2, D1, and D2, illustrating variations in binding energies and intensities.
Figure 9. XPS spectra showing (a) a survey scan of sample A1, identifying key peaks including OKLL, O 1s, C 1s, Ti 2p, Si 2s, Si 2p, and Er 5p; and (b) a detailed comparison of the Er 5p1/2 and Er 5p3/2 peaks for samples A1, A2, D1, and D2, illustrating variations in binding energies and intensities.
Foundations 05 00005 g009
Table 1. Diffusion factor f E for each energy of the incident X-ray beam in the ASAXS measurements. λ is the wavelength associated with this energy, and ΔE is the energy variation relative to the LIII absorption edge of erbium (8357 eV) [19].
Table 1. Diffusion factor f E for each energy of the incident X-ray beam in the ASAXS measurements. λ is the wavelength associated with this energy, and ΔE is the energy variation relative to the LIII absorption edge of erbium (8357 eV) [19].
Energy (eV)ΔE (eV)λ (Å) f E
80003571.5500−9.584
81572001.5200−10.822
82111461.5100−11.487
8265921.5001−12.474
8289681.4957−13.105
8313441.4914−14.015
8324331.4894−14.625
8336211.4874−15.539
8341161.4864−16.095
8347101.4853−17.059
Table 2. Crystallization phases observed in samples A, B, C, and D.
Table 2. Crystallization phases observed in samples A, B, C, and D.
Sample CompositionCrystallization BehaviorReferences
900 °C1000 °C
A: 80SiO2–20TiO2–2 mol% ErO1.5amR + ETO[1], This work
B: 80SiO2–20TiO2–10 mol% PO2.5 and 2 mol% ErO1.5amR + EPO[1]
C: 98SiO2–2TiO2–10 mol% PO2.5 and 2 mol% ErO1.5EPOEPO[1]
D: 98SiO2–2TiO2 with 2 mol% ErO1.5ETOETOThis work
am—amorphous; R—rutile (TiO2); ETO—erbium titanate (Er2Ti2O7); and EPO—erbium phosphate (ErPO4).
Table 3. Results of the linear fits and Guinier radii for various compositions. Adapted from [19].
Table 3. Results of the linear fits and Guinier radii for various compositions. Adapted from [19].
SampleA1A2B1B2C1C2D1D2
β/α230.956.973.686.157.556.551.565.9
R β / α 2 0.7230.9800.9800.9800.9800.9800.9800.980
Rg (nm)16.917.2318.0218.0318.9317.7918.5017.31
R2 represents the square of the correlation coefficient r.
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

Vasconcelos, H.C.; Meirelles, M.; Özmenteş, R.; Santos, L. Structural Analysis of Erbium-Doped Silica-Based Glass-Ceramics Using Anomalous and Small-Angle X-Ray Scattering. Foundations 2025, 5, 5. https://doi.org/10.3390/foundations5010005

AMA Style

Vasconcelos HC, Meirelles M, Özmenteş R, Santos L. Structural Analysis of Erbium-Doped Silica-Based Glass-Ceramics Using Anomalous and Small-Angle X-Ray Scattering. Foundations. 2025; 5(1):5. https://doi.org/10.3390/foundations5010005

Chicago/Turabian Style

Vasconcelos, Helena Cristina, Maria Meirelles, Reşit Özmenteş, and Luís Santos. 2025. "Structural Analysis of Erbium-Doped Silica-Based Glass-Ceramics Using Anomalous and Small-Angle X-Ray Scattering" Foundations 5, no. 1: 5. https://doi.org/10.3390/foundations5010005

APA Style

Vasconcelos, H. C., Meirelles, M., Özmenteş, R., & Santos, L. (2025). Structural Analysis of Erbium-Doped Silica-Based Glass-Ceramics Using Anomalous and Small-Angle X-Ray Scattering. Foundations, 5(1), 5. https://doi.org/10.3390/foundations5010005

Article Metrics

Back to TopTop