Next Article in Journal
Accelerating High-Entropy Alloy Design via Machine Learning: Predicting Yield Strength from Composition
Previous Article in Journal
Boosted NH3 Selective Catalytic Oxidation Activity over V-Pt-Ti Catalysts: Insight into Preparation Method Effects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fundamentals of Cubic Phase Synthesis in PbF2–EuF3 System

1
Department of Chemistry and Technology of Crystals Mendeleev, University of Chemical Technology of Russia, Miusskaya sq. 9, Moscow 125047, Russia
2
Prokhorov General Physics Institute of the Russian Academy of Sciences, Vavilova Str., 38, Moscow 119991, Russia
*
Authors to whom correspondence should be addressed.
Materials 2026, 19(1), 195; https://doi.org/10.3390/ma19010195
Submission received: 6 December 2025 / Revised: 25 December 2025 / Accepted: 30 December 2025 / Published: 5 January 2026
(This article belongs to the Section Optical and Photonic Materials)

Highlights

  • A high-purity cubic phase could be synthesized in the 0–37 mol% EuF3 composition range in a quasi-binary PbF2-EuF3 system.
  • The ordered rhombohedral R-phase exists in the concentration range from 37–39 to 43–44 mol% EuF3 in the quasi-binary PbF2-EuF3 system.
  • Isothermal cross-sections of T-X-Y projection of ternary Pb-Eu-F diagram in the 50–550 °C temperature range are reconstructed.

Abstract

Fluoride solid solutions exhibit exceptional optical and thermodynamic properties that make them valuable for advanced technological applications, and the PbF2-EuF3 system represents a particularly promising quasi-binary system for developing high-performance materials. However, the comprehensive understanding of the thermodynamic conditions governing phase equilibria and the precise boundaries of homogeneity regions in this system remains incomplete, limiting the rational design of single-phase materials with desired properties. Therefore, we conducted a comprehensive investigation of the thermodynamic conditions (temperature and composition) controlling the existence of cubic and rhombohedral phases within the homogeneity regions of the PbF2-EuF3 system. Solid solution samples were synthesized using both solid-phase synthesis and co-precipitation techniques from aqueous nitrate solutions. Phase equilibria were systematically investigated in two critical regions: the solvus line spanning 0–10 mol% EuF3 and the ordered rhombohedral R-phase region spanning 35–45 mol% EuF3. Structural characterization was performed at temperatures below the phase transition temperature in lead fluoride (365 °C) using X-ray phase analysis, optical probing, and Raman scattering. Our investigation successfully demonstrated the possibility of obtaining cubic preparations of high purity across the 0–37 mol% EuF3 composition range. Additionally, we precisely defined the region of existence of the ordered rhombohedral R-phase within the concentration range of 37–39 to 43–44 mol% EuF3. These findings provide essential thermodynamic data for the rational design of PbF2-EuF3 solid solutions and establish clear compositional boundaries for obtaining desired phase structures in this technologically important fluoride system.

Graphical Abstract

1. Introduction

Lead fluoride crystals have favorable properties such as high density, a non-hygroscopic and relatively chemically inert nature, and a short radiation length, while the scintillation light output is too small to obtain a photo-absorption peak under gamma ray excitation. Thus, it could not be used in medical imaging applications [1]. Lead fluoride crystals are widely used for the detection of beta rays due to their relatively small back scattering. Lead fluoride activated by europium (PbF2:Eu) is a light scintillator that is used for the detection of charged particles and soft gamma rays up to several hundred keV [2].
Timing performance of PbF2 crystals of various lengths and surface conditions coupled to SiPMs was evaluated against a reference detector with an optimized test setup using high-frequency readout and novel time walk correction, with special attention to the inherent limitations of single-photon Cherenkov detection only [3,4]. These materials also find application as up-conversion phosphors [5,6,7,8].
Ce-doped PbF2 crystals did not show intense photo- and radio-luminescence, and Eu- and Ho-doped ones showed several peaks excited under UV and 5.5-MeV alpha ray excitation, respectively [5].
In general, fluoride compounds and solid solutions in PbF2-REF3 systems (RE = rare earth element) are of interest for research as luminescent [9,10,11,12,13] and laser materials [14,15,16,17], especially effective in the mid-infrared range and in up-conversion imaging [13], as well as ionic conductors [18]. These systems have a wide crystallization region of the cubic fluorite phase Fm 3 ¯ m (over 20 mol% at 650–700 °C) [9,10], but the phase diagrams and the phase existence regions at low temperatures (below 650 °C) are insufficiently studied. The complexity of studying these phase diagrams lies in the fact that lead fluoride undergoes a phase transition at 335–360 °C. The high-temperature β-PbF2 polymorph crystallizes in a fluorite-like cubic structure (Fm 3 ¯ m), whereas the low-temperature α-PbF2 phase adopts an orthorhombic (Pnma) structure [9]. There is no reliable information about the homogeneity limits of PbF2, given the need to conduct research with highly pure materials. Typically, the declared chemical purity of the PbF2 under study is no more than 99.9 wt%.
PbF2 β ⟶ α phase transition is “frozen” because of the metastable state of β-PbF2 at room temperature. It allows using this material as a fluoride ion conductor and an optical material. When heated, the phase transition occurs easily, whereas when β-PbF2 is cooled from high temperatures, the phase transition is kinetically hindered.
In general, for the systems MF2–REF3 (where M = Ba2+, Sr2+, Ca2+, and Pb2+; RE = rare earth element), the presence of a rhombohedral ordered phase (R-phase) with the ideal composition M4RE3F17, existing in a relatively narrow concentration range of about 40 mol% REF3 [10,19,20,21,22,23], is assumed. According to the data [21], this phase is close to the fluorite phase, which complicates the identification of “rhombohedral fluorite”.
The phase diagram of the PbF2-EuF3 system is insufficiently explored—the high-temperature range of 650–1300 °C is only predicted [9] based on diagrams of other rare earth elements—both F- and R-phases were identified at 650 °C [10], and analysis conducted at 50 and 500 °C did not establish the presence of the R-phase, as this region was examined with a large interval of 10 mol%. Meanwhile, highly concentrated environments with reduced symmetry, activated by Eu, are promising materials for laser generation in the red spectral range [24].
Studies of solid solutions of PbF2-EuF3 and glass–ceramic materials have shown that the introduction of EuF3 in concentrations of 7–10 mol% stabilizes the high-temperature phase even when synthesized at temperatures below 300 °C [11,12]. Thus, on the T-X diagram of PbF2-EuF3 at temperatures below 365 °C and EuF3 concentrations of 0–10 mol%, there should be a line between the regions of crystallization of a single cubic phase solid solution Pb1−xRExF2+x and the region of crystallization of two phases (solvus line). In the two-phase region, the above-mentioned solid solution phase and a phase extremely close to α-PbF2 are in equilibrium, as the solubility of RE in the low-temperature phase of lead fluoride is negligible (Figure 1).
It is known that impurities even in concentrations of hundredths of a mol% can significantly shift the boundaries of the homogeneity region, including the temperatures of polymorphic transitions [25].
The overall aim of this study was to clarify the boundaries of the homogeneity region of the cubic phase for future development of synthesis technology of single-phase extra-pure cubic phase preparations in PbF2-EuF3 system. To complete this task, we investigated two regions of the phase diagram of the quasi-binary system PbF2-EuF3 within the composition range of 0–45 mol% EuF3 and identified the R-phase within the composition range of 35–45 mol% EuF3. To date, low-temperature phase equilibria and the homogeneity range of the cubic phase and R-phase in PbF2–EuF3 system are still poorly constrained, which limits the rational design of luminescent materials based on this system.

2. Materials and Methods

Preparations within the PbF2-EuF3 system with EuF3 content ranging from 0 to 12 mol% and from 35 to 45 mol% were synthesized at a low temperature using the co-precipitation technique [12,26]. The starting materials included Pb(NO3)2 (99.99 wt%, LANHIT, Moscow, Russia), Eu(NO3)3 (99.99 wt%, LANHIT, Moscow, Russia), and HF (99.9 wt%, TECH Systems, Moscow, Russia) that were additionally purified to 99.99999 wt%. Initially, lead and europium nitrate solutions were prepared in bidistilled water (0.08 mol/L). Subsequently, these solutions were meticulously mixed in predetermined proportions. The resulting solution was added dropwise at 10 mL/min into the hydrofluoric acid solution (5 vol%) with continuous stirring, and a 10-fold excess of hydrofluoric acid was added to maintain a constant pH and T = 25 °C during the process. The process involved continuous stirring of the mother liquor using a magnetic stirrer. At this stage, a chemical reaction has been conducting:
(1 − x)Pb(NO3)2 + xEu(NO3)3·6H2O + (2 + x)HF → Pb1−xEuxF2+x↓ + H2O,
The synthesis was carried out in a polypropylene reactor, and upon completion, the matrix solution was further stirred for one hour. The resulting precipitate was decanted, washed with bidistilled water until a negative reaction of diphenylamine to nitrate ions was observed, and then air-dried at 40–50 °C.
To determine the temperature of the phase transition in the concentration range of 0–9 mol%, EuF3 studies were performed on powder samples first obtained by co-precipitation at 50 °C and then subjected to heat treatment at certain temperatures of 200 °C, 300 °C, 330 °C, 350 °C, and 360 °C for 48 h in a resistance tube furnace in the first step. X-ray diffraction and luminescence spectroscopy were then analyzed. In the second step, the same samples were again heat-treated at the same temperature for an additional 24 h. A further set of studies was conducted. In all cases, the parameters remained unchanged. Therefore, we consider these data to be close to equilibrium. Because the transition from the high-temperature cubic phase to the low-temperature rhombic phase is “frozen,” quenching was not performed. The samples cooled from the treatment temperature to RT at an average rate of 50–100 °C/hour.
Preparations in the PbF2-EuF3 system with EuF3 concentrations ranging from 35 to 45 mol% were also obtained by solid-phase synthesis at 550 °C, which was higher than the reference α ⟶ β transition temperature. PbF2 (99.99% purity, LANHIT, Moscow, Russia) and EuF3 (99.99% purity, LANHIT, Moscow, Russia) were used as starting materials. Initially, the sampled powders were ground in a mortar to increase the contact area between the reacting substances. The solid-phase synthesis was conducted in a resistance handmade tube furnace at 550 °C for 2 h in corundum crucibles. Afterwards, the samples were cooled, and a series of studies were conducted. Then, the procedure of grinding was repeated, and the next heat treatment was performed under the same conditions for another 2 h. The same series of studies were repeated. Additionally, a fluorinating atmosphere was established [11,12,27,28] to prevent pyrohydrolysis. The results obtained after the first and second heat treatment were coincided.
The formation of a solid solution is described by the following chemical reaction.
(1 − x)PbF2 + xEuF3 → Pb1−xEuxF2+x
After completion of the first stage, the resulting sintered powders were re-ground and placed in the furnace under identical conditions.
X-ray diffraction analysis was conducted using an INERL Equinox 2000 X-ray diffractometer (Inel SAS, Artenay, France) (CuKα radiation with a wavelength of λ = 1.54 Å), with an accuracy of lattice parameter determination of +1%·and sensitivity up to 1% of the impurity phase. X-ray diffraction patterns were interpreted using Le Bail method (PCPDFWIN electronic catalog and JCPDS-ICDD database), and phase ratios were calculated using the Match! version 4.1 software. Volume fractions of the crystalline phase were determined with an error of ± 2%. All experiments were carried out at least 3 times (3–5 times). The results converged within the error limits. The errors fit within the symbol size in the presented data figures (see Figures 4, 5, 7, 10 and 12). During the analysis, the absence of partially oxidized phase (Pb2OF2) was additionally monitored based on the peak at 2Θ = 26.75° [11,29].
Energy-dispersive X-ray spectroscopy (EDS) microanalysis was performed to determine the actual composition of the samples using a scanning electron microscope VEGA3-LMU (TESCAN, Brno, Czech Republic) equipped with a lanthanum hexaboride thermionic cathode and an Oxford Instruments X-MAX-50 EDS detector, operating in the secondary electron mode. The AZTec software V.2.0 was utilized for data collection and processing. All measurements were conducted at room temperature, with no fewer than 8 points measured for each sample. Imaging was carried out at an accelerating voltage of 30 kV.
The impurity content of the starting materials and the resulting powders was determined by an inductively coupled plasma mass spectrometer NexION300D (PerkinElmer Inc., Waltham, MA, USA). Lead fluoride-based powders were dissolved in 20 mL of sulfuric acid (7N, 96%) purified by a surface distillation system Milestone DuoPUR (Milestone S.r.l., Sorisole, Italy) with microwave digestion of the sample in polytetrafluoroethylene autoclaves (DAP-100, PTFE, BERGHOFF GmbH&Co., Wenden, Germany) using a SPEEDWAVEFOUR microwave decomposition setup (BERGHOFF GmbH&Co., Wenden, Germany). The resulting solution was transferred to a polypropylene test tube and diluted with water. Deionized water was obtained using an Aquapuri 5–551 Series (Young Lin Instruments Co., Ltd., Hogye, Republic of Korea) and had an electrical resistance of 18.2 MΩ cm. The prepared solution was analyzed using inductively coupled plasma mass spectrometry (ICP-MS). Analytical measurements were carried out using an inductively coupled plasma mass spectrometer NexION300D (PerkinElmer Inc., Waltham, MA, USA) [29]. The resulting powders had a purity of greater than 99.99 wt% (see Table S1 for impurity analysis).
Europium (Eu3+) photoluminescence (PL) spectra were recorded using a Fluorolog FL3-22 spectrofluorometer (Horiba Jobin Yvon, Longjumeau, France) in the wavelength range of 400–700 nm with a 0.1 nm step, excited by a diode laser (λ = 377 nm).
Raman scattering spectra (RS) were investigated using a HORIBA LabRam Soleil instrument (Horiba Jobin Yvon GmbH, Bensheim, Germany) with the ability to excite RS with lasers of three wavelengths: 377 nm, 532 nm, and 785 nm. In our case, only the 785 nm excitation wavelength was used, since others caused the excitation of Eu3+ fluorescence [30]. Measurements were carried out in the range of 149–3500 cm−1 with a 2 cm−1 step.
Infrared (IR) transmittance measurements were performed using a Tensor 27 Fourier-transform infrared spectrometer (FTIR) (Bruker Optics Inc., Billerica, MA, USA) in the range of 400–8000 cm−1 with a 1.9 cm−1 step.

3. Results and Discussion

3.1. Position of the Solvus Line on the PbF2-EuF3 Diagram

The samples synthesized by the co-precipitation method have appeared as fine white powders. The appearance did not change depending on the nominal europium content, indirectly indicating the absence of oxidized phases. Analysis of the phase composition of the samples after synthesis at nominal concentrations from 0.5% to 7 mol% EuF3 revealed the presence of two phases simultaneously: α-PbF2 and cubic Pb1−xEuxF2+x. From 8 to 12 mol%, there was only one cubic phase of the solid solution with parameters corresponding to the nominal composition (Figure 2). In 0–0.5 mol% EuF3 composition range, the X-ray reflections belonging to the rhombic phase corresponded to pure α-PbF2.
It is worth noting that the reflections of the cubic solid solution Pb1−xEuxF2+x are shifted relative to the reflections of nominally pure cubic β-PbF2 at 2Θ large angles, which indicates the compression of cubic lattice parameters because of the incorporation of smaller Eu3+ ions into a solid-solution crystal lattice.
The lattice parameter for these solid solutions obeys a linear equation (Vegard’s law):
a = 5.940 + kEu·x,
where 5.940 Å—the β-PbF2 lattice parameter; x—the mol% of RE; and the coefficient kEu = −0.00237 as determined in [10,11].
The lattice parameter of the obtained solid solutions in the range of nominal EuF3 content from 1 to 7 mol% was found to be 5.919 ± 0.003 Å, corresponding to the EuF3 content of 7.0–9.5 mol% according to the Vegard law.
The examination of the PXRD patterns after heat treatments revealed (Figure 3, Figure S1, Table S2 for the volume fraction of the cubic phase) that the proportion of the cubic phase began to increase starting from 300 °C. For the sample with a nominal composition of Pb0.97Eu0.03F2.03, at 330 °C, the volume fraction of α-PbF2 is small compared to a solid-solution cubic phase (Figure 3), which is consistent with the findings of a previous study [10].
A sharp increase in the proportion of the cubic phase is observed at temperatures above 330 °C; samples with a nominal EuF3 content exceeding 5 mol% are already single-phased in this temperature range. In nominally pure PbF2, the rhombic phase is detected even after heat treatment at 360 °C (Figure 4).
Remarkably, the lattice parameters of the cubic phase after heat treatment show a tendency to increase (Figure 5), indicating an approach of the lattice parameters of solid solutions obtained by solid-phase synthesis at 500 °C [11].
Such lattice parameter behavior and the unchanged properties of the powders after an additional 24 h heat treatment indicate that the obtained values are close to equilibrium state.
Eu3+ is a luminescent-active impurity, and its spectrum is sensitive to changes in the symmetry of the surrounding environment. Thanks to this, the Eu3+ ion is often used as a spectroscopic probe, making samples with Eu3+ concentration of 0.1 mol% [31]. It is considered that, at such concentrations, the additional introduction of Eu does not significantly alter the structure. In our case, the Eu concentration is knowingly higher and varies from sample to sample, so the use of all the fine capabilities of spectroscopic probing analysis may be incorrect; hence, we used only an estimation of local symmetry by the asymmetry coefficient [11,12].
The electric dipole transition 5D07F2 (~ 612 nm) in the Eu3+ ion is highly sensitive. The magnetic dipole transition 5D07F1 (~ 590 nm), permitted in terms of parity, has an intensity that is practically independent of the point symmetry of the luminescent center and its environment. To characterize the local environment of Eu3+ ions, the asymmetry coefficient R21 is used. By definition, it is the ratio of the intensities of the highly sensitive electric dipole transition 5D07F2 and the magnetic dipole transition 5D07F1 [12]:
R 21 =   I E D D 5 0 F 7 2 I M D D 5 0 F 7 1 ,
The greater the value of this ratio, the less symmetrical the environment of the europium ion in this matrix. Dominance in the intensity of the band corresponding to the magnetic dipole transition 5D07F1 over the electric dipole transition 5D07F2 indicates an environment of the Eu3+ ion that is close to centrosymmetric, as observed in solid solutions (Figure 6).
The transition between the energy levels 5D0 and 7F0 is also forbidden by the selection rules for electronic dipole transitions; moreover, these levels are degenerate and therefore have zero Stark splitting. This transition is practically absent in the fluorescence spectra of solid solutions, confirming a symmetry close to cubic for the Eu3+ optical centers.
As the annealing temperature increases, the asymmetry coefficient behaves differently depending on the nominal composition (Figure 7).
In samples with a nominal content of EuF3 ranging from 3 to 9 mol%, there is a gradual decrease in R21, indicating an increase in symmetry. In samples with 0.5 and 1 mol% EuF3, initially up to 200 °C, R21 increases due to structure relaxation, followed by a sharp decrease in R21, indicating increased centrosymmetry. In the phase transition region, R21 in all samples ranges from 0.30 to 0.35 (see Table S3 for values of the asymmetry coefficient R21). Thus, the structural data obtained from XRD and luminescence studies coincide.
Figure 8 schematically represents a fragment of the T-X diagram of the quasi-binary system PbF2-EuF3 in the investigated range of temperatures and compositions. It is evident that, with the increase in the nominal content of EuF3, the temperature of the phase transition sharply decreases. It is worth noting that the phase transition in all investigated compositions is “frozen”, as in nominally pure PbF2 [9], meaning that no phase separation occurs upon cooling for a long time (at least, no changes in the structure and luminescence of annealed powders have occurred over one year of exposition. See Figure S3).

3.2. Determining the Existence Region of the Rhombohedral R-Phase in the PbF2-EuF3 System

To determine the existence of the R-phase, samples were synthesized in the nominal Eu concentration range of 35–45 at% using the co-precipitation and solid-phase methods. In both cases, the synthesized samples appeared as white powders. It was particularly important in the case solid-phase synthesis, indicating the absence of oxidation processes and the formation of lead oxyfluorides and oxides.
X-ray spectral microanalysis allowed determining that the deviations from the theoretically calculated compositions of the samples during synthesis do not exceed 0.05 at% (see Table S4 for SEM-EDX analysis). The SEM analysis of powder images shows that the particle sizes obtained by the co-precipitation technique vary widely from several to tens of microns, with irregular shapes (Table S4). Finer powders (ranging from 0.5 to 1 micron) were obtained through solid-phase synthesis, attributed to several grinding steps during synthesis.
When studying the diffraction patterns of samples obtained by solid-phase synthesis (Figure 9a) and co-precipitation from aqueous nitrate solutions (Figure 9b), the presence of solid solutions with parameters different from the high-temperature cubic β-PbF2 was confirmed, which was especially noticeable at 2Θ large values. It results from lattice parameters.
To confirm the assumption of the presence of the R-phase, the calculation of the unit cell parameter was carried out based on the obtained XRD data. Initially, the peaks were indexed within the framework of a cubic cell (Fm 3 ¯ m) with parameters close to the high-temperature modification of lead fluoride. The obtained values of the unit cell parameter for samples obtained by solid-phase synthesis and the co-precipitation technique are presented in Table S5.
Graphically analyzing the obtained results (Figure 10a), a significant difference is noticeable between the theoretical values (green line in Figure 10) obtained according to Vegard’s law for a hypothetical solid solution Pb1−xEuxF2+x and the values of the unit cell parameter determined for samples synthesized by the co-precipitation and solid-phase methods. According to [32], the deviation is better observed in terms of the volume of the crystalline cell (Figure 10b), and correspondingly in linear equations (Röttgers’s rule):
V = V0 + kV·x,
where V0—the unit cell volume of the fluorite matrix MF2 (for β-PbF2 209,58 Å3); x—the mol% of RE; kV—the coefficient.
kV = 3a0k,
Thus, from Equations (3) and (6), kVEu = −0.2509.
In the range of 40–45 mol% EuF3, deviations in cell parameter and volume from the calculated values are observed. Scanning with a larger step [11] indicated that the violation of Vegard’s law begins at 36 ± 2 mol% EuF3, but a significantly different phase, based on the rhombic (Pnma) modification of europium trifluoride Er1-yPbyF3-y, can only be observed starting from 60 mol% EuF3. Thus, deviations from linear laws in the 40–45 mol% EuF3 range may be associated with the existence of a phase in this range, with reflections close to the Pb1−xEuxF2+x cubic solid solution, i.e., the R-phase.
We discovered that deviations from the linear law for samples obtained by solid-phase synthesis and co-precipitation were in opposite directions. A slight increase in the cell parameter for solid-phase samples corresponding to a solid solution with lower Eu concentration may indicate the presence of a narrow two-phase equilibrium region of the cubic solid solution Pb1−xEuxF2+x and its distorted modification, the R-phase. A decrease in the cell parameter for samples obtained by the co-precipitation technique may be related to the non-equilibrium nature of the obtained phases.
The fluorescence spectra of powder samples in this concentration range were also investigated (Figure 11).
In 37–45 mol% EuF3 concentration range, the asymmetry coefficient was significantly higher than that in the range up to 10 mol% EuF3. For samples with composition (100 − x)PbF2−xEuF3 (x = 41–45 mol% EuF3), the coefficient R21 gradually increases (see Figure 12), indicating a slight increase in asymmetry due to the heterovalent substitution in the solid solution. However, with further increase in Eu concentration, R21 undergoes a sharp increase. Thus, the boundary at which the local structure begins to distort significantly is approximately 39 mol% Eu.
Additionally, vibrational spectra were investigated in this system. The complete vibrational representation for crystals with a fluorite structure can be expressed as follows [33]:
Γ = F1u (IR) + F2g (R),
where F1u represents a triply degenerate, antisymmetric vibration with respect to the center of symmetry, active in the infrared spectrum, and F2g represents a triply degenerate, symmetric vibration with respect to the center of symmetry, active in the Raman spectrum.
The FTIR spectroscopy (Figure S2) in our case does not allow determining the vibrational modes of cubic PbF2, as the F1u vibrations lie around 347 cm−1 (less than 400 cm−1) [34]. However, it did reveal the presence of nitro groups [NO3] [35] and hydroxyl groups [OH] [35] in samples obtained by co-precipitation (see Figure S2, and Table S6 for absorption spectra details), indicating the high sensitivity of the method, as the qualitative reaction of diphenylamine was negative, and the samples were dried for an extended period.
In the Raman spectra of crystals of undistorted fluorite-type, a single line will be observed, and the frequency of this line for a crystal of nominally pure cubic β-PbF2 falls within the range of 256–259 cm−1 [36], corresponding to the F2g vibration mode. This is clearly seen in the spectra of single-phase samples with concentrations of 5 and 11 mol% EuF3. When comparing the Raman spectra of samples with Eu concentrations ranging from 37 to 45 mol% EuF3, a noticeable shift in the peak maximum and broadening of the peaks can be observed (Figure 13, see Table S7 for Raman spectra analysis). The Raman spectra present a superposition of one cubic phase maximum and several R-phase maxima. Since the R-phase is a rhombic deformed cubic phase with similar lattice parameters, its Raman spectrum represents the broadening and splitting of one cubic phase band, as shown for the Ba4Y3F17 phase in a similar system [37]. Thus, a sharp broadening of the Raman band with a shift to the region of high energies indicates the formation of the R-phase.
One can observe a significant shift in the peak maximum for samples containing Eu 37–43 at%, indicating the presence of another phase structurally similar (Figure 13). Since the phase of the cubic solid solution Pb1−xEuxF2+x and the R-phase are very close in structure and parameters, the splitting of the band into close components leads to its broadening. The narrowing of the band in the region of 43 at% Eu may be associated with the crystallization region of only the R-phase.
Summarizing the results of phase equilibria analysis, we reconstructed the T-X-Y diagram of Pb-Eu-F ternary system within the PbF2-EuF3 section (Figure 14). In this study, we proved the general view of T-X-Y diagram [11] but expanded the temperature range from 50 to 550 °C and clarified the concentration ranges of homogeneity limits for cubic and rhombic PbF2:Eu and R-phase. For better understanding, we presented the fragment of T-X-Y diagram nearby the PbF2-EuF3 quasi-binary cross-section in a variable scale with the large logarithmic scale near PbF2 and the linear scale starting from 10 mol% EuF3. The homogeneity limits were determined along the PbF2-EuF3 quasi-binary cross-section in mol% EuF3. Their widths towards the Metal-F direction were less than 1 mol%, which is typical for the homogeneity ranges of fluoride binary compounds [9]. Also, cross-sections of T-X-Y projection of ternary Pb-Eu-F diagram are presented schematically for certain temperatures. For better understanding, the areas of trivariant equilibria, i.e., phase homogeneity areas, are presented on a large scale, contrary to the common view when they presented as dots on the actual scale.
The main features of the analyzed system were determined, and it was established that the homogeneity region of the R-phase is about 1 mol.% and spans 44–45 mol% EuF3, slightly changing with temperature over the entire temperature range under consideration.
The cubic solid solution Pb1−xEuxF2+x spans 7.5–38 mol% EuF3 at 50–200 °C with a possible retrograde solvus behavior at low temperatures (dash line in Figure 14). It then expands from 7.5 to 37 mol% EuF3 at 300 °C. Starting at 330 °C, it reaches the T-Pb-F plane and no longer deviates from it with increasing temperature. The Eu-rich boundary of the cubic solid solution Pb1−xEuxF2+x shifts from 37 to 32 mol% EuF3 at 550 °C.
Simultaneously, PbF2:Eu rhombic phase expanded from 0.1 to 1 mol% EuF3 at a temperature rise from 50 to 280 °C. After then, it is quickly compressed to zero at 360 °C. In the entire temperature range under consideration, the homogeneity region of the orthorhombic PbF2 phase does not detach from the T-Pb-F plane.
So, we established that the stable cubic phase in PbF2-EuF3 quasi-binary system could be obtained at all temperatures starting from RT to 550 °C, and it exists in a wide concentration range up to 38 mol% of EuF3.
More detail bivariant and monovariant equilibria with Pb1−xEuxF2+x cubic solid-solution phase and R-phase are plotted in Figure 15 for the T-X-Y cross-section at 200 °C. We can see Scubic solid solutionSR-phaseV bivariant equilibrium in 38–44 mol% EuF3 range, which slightly broadens with an increase in temperature (see Figure 14).
The only currently unclear issue is the existence of Sα-PbF2Sβ-PbF2SR-phaseV monovariant equilibrium at low temperatures (Figure 15—red triangle). We believe this issue is important for the future operation of devices based on a cubic solid solution Pb1−xEuxF2+x phase, but for successful design of materials based on the cubic phase, detailed investigations of 3D homogeneity range must be carried out. And the first challenge in these investigations is having a method of measurement of nonstoichiometry for fluorine similar to those used for oxygen for the high-temperature superconducting YBa2Cu3O7-δ phase [38].

4. Conclusions

Fluoride solid solutions in lead–europium systems are of significant interest for optical and electronic applications due to their unique luminescent properties and structural characteristics. Moreover, understanding phase relationships and structural transitions in these systems is crucial for optimizing their functional properties. However, the complete phase diagram and structural behavior across the full composition range of the (100 − x)PbF2−xEuF3 quasi-binary system remains incompletely characterized, particularly regarding phase transition mechanisms and local structural environments. Therefore, we conducted a comprehensive investigation of phase relationships and structural transitions across two distinct composition ranges of the (100 − x)PbF2−xEuF3 system using multiple synthesis and characterization approaches.
For the low-europium range (x = 0–12 mol% EuF3), samples were synthesized by co-precipitation from aqueous solutions. Phase composition analysis revealed pure α-PbF2 and cubic solid solution Pb0.93Eu0.07F2.07 for x = 0.5–7 mol% and cubic solid solution Pb0.93Eu0.07F2.07 for x = 8–12 mol%. Thermal treatments at 200–400 °C enabled the determination of phase transition temperatures for samples with x = 0–7 mol%. Investigation of Eu3+ local structure showed significant symmetry increases during thermal treatments at 300–350 °C, allowing solvus line position determination.
For the high-europium range (x = 35 to −45 mol% EuF3), samples were synthesized using both solid-phase and liquid-phase methods. X-ray diffraction analysis revealed deviations from Vegard’s law and Röttgers’s rule in the 39–44 mol% Eu range. Spectral-luminescent analysis co-precipitated samples (37–44 mol% EuF3) showed significant changes in asymmetry coefficient R21, indicating altered europium ion environment symmetry. Raman spectroscopy confirmed a phase slightly different from the cubic structure in the 37–43 at% Eu range. The results establish that two close phases coexist in the x = 36–42 mol% EuF3 range: cubic Pb0.64Eu0.36F2.36 and rhombically distorted R-phase. At 43 mol% EuF3, only the R-phase appears to be present, evidenced by Raman band narrowing, while higher EuF3 concentrations establish phase equilibrium among R-phase, Eu1−γPbγF3−γ, and EuF3.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma19010195/s1, Figure S1: X-ray powder diffraction patterns of sample synthesized in the PbF2-EuF3 quasi-binary system by the co-precipitation technique and annealed at different temperatures; Figure S2: Fragment of the IR transmittance spectra of samples in the system (100 − x)PbF2−xEuF3 (x = 35–45 mol.% EuF3), obtained by the co-precipitation technique; Figure S3: PL spectra of as-synthesized samples and after 1-year exposition under normal RT conditions; Table S1: Impurity element concentrations in the samples of PbF2 and Pb0.65Eu0.35F2.35 solid solution determined by ICM-MS; Table S2: Volume fraction of the cubic phase during heat treatment; Table S3: Values of the asymmetry coefficient R21 for samples with and without heat treatment; Table S4: Results of scanning electron microscopy supported by EDS analysis; Table S5: Cell parameter of Pb1−xEuxF2+x samples (x = 35–45 mol%); Table S6: Absorption band values for (100 − x)PbF2−xEuF3 samples (x = 35–45 mol.% EuF3); Table S7: Raman spectra of solid solutions with different Eu content.

Author Contributions

Conceptualization, O.P. and I.A.; methodology, S.Z., K.R. and O.P.; software, M.B. and O.P.; validation, M.M. and I.A.; formal analysis, K.R. and I.A.; investigation, S.Z., K.R., M.M. and M.B.; resources, R.A.; data curation, K.R.; writing—original draft preparation, M.B. and I.A.; writing—review and editing, O.P. and R.A.; visualization, M.M. and M.B.; supervision, I.A.; project administration, I.A.; funding acquisition, R.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Science and Higher Education of Russia through the project FSSM-2025-0006.

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Materials. Further inquiries can be directed to the corresponding authors.

Acknowledgments

The authors are grateful to the National Analytic Certification Laboratory of Mendeleev Center of D. Mendeleev University of Chemical Technology of Russia for ICP-MS, XRD, FTIR, and luminescence measurements.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ntoupis, V.; Linardatos, D.; Saatsakis, G.; Kalyvas, N.; Bakas, A.; Fountos, G.; Kandarakis, I.; Michail, C.; Valais, I. Response of Lead Fluoride (PbF2) Crystal under X-Ray and Gamma Ray Radiation. Photonics 2023, 10, 57. [Google Scholar] [CrossRef]
  2. PbF2—Lead Fluoride Scintillator Crystal|Advatech UK. Available online: https://www.advatech-uk.co.uk/pbf2.html (accessed on 13 November 2025).
  3. A Roadmap for Sole Cherenkov Radiators with SiPMs in TOF-PET—IOPscience. Available online: https://iopscience.iop.org/article/10.1088/1361-6560/ac212a (accessed on 20 November 2025).
  4. Cantone, C.; Carsi, S.; Ceravolo, S.; Di Meco, E.; Diociaiuti, E.; Frank, I.; Kholodenko, S.; Martellotti, S.; Mirra, M.; Monti-Guarnieri, P.; et al. Beam Test, Simulation, and Performance Evaluation of PbF2 and PWO-UF Crystals with SiPM Readout for a Semi-Homogeneous Calorimeter Prototype with Longitudinal Segmentation. Front. Phys. 2023, 11, 1223183. [Google Scholar] [CrossRef]
  5. Li, D.; Mo, J.; Wang, C.; Liu, W.; Ge, H.; Han, D.; Hao, A.; Chai, B.; She, J. Screen Printing of Upconversion NaYF4:Yb3+/Eu3+ with Li+ Doped for Anti-Counterfeiting Application. Chin. Opt. Lett. 2020, 18, 110501. [Google Scholar] [CrossRef]
  6. Xiong, J.; Li, G.; Zhang, J.; Li, D.; Pun, E.Y.B.; Lin, H. Fluorescence Regulation Derived from Eu3+ in Miscible-Order Fluoride-Phosphate Blocky Phosphor. Nanotechnology 2021, 32, 435705. [Google Scholar] [CrossRef]
  7. Manasa, P.; Ran, F.; Basavapoornima, C.; Depuru, S.R.; Jayasankar, C.K. Optical Characteristics of (Eu3+,Nd3+) Co-Doped Leadfluorosilicate Glasses for Enhanced Photonic Device Applications. J. Lumin. 2020, 223, 117210. [Google Scholar] [CrossRef]
  8. Kurosawa, S.; Yokota, Y.; Yanagida, T.; Yoshikawa, A. Optical and Scintillation Property of Ce, Ho and Eu-Doped PbF2. Radiat. Meas. 2013, 55, 120–123. [Google Scholar] [CrossRef]
  9. Buchinskaya, I.I.; Fedorov, P.P. Lead Difluoride and Related Systems. Russ. Chem. Rev. 2004, 73, 371–400. [Google Scholar] [CrossRef]
  10. Tyagi, A.K.; Patwe, S.J.; Achary, S.N.; Mallia, M.B. Phase Relation Studies in Pb1−xM′xF2+x Systems (0.0 ≤ x ≤ 1.0; M′ = Nd3+, Eu3+ and Er3+). J. Solid State Chem. 2004, 177, 1746–1757. [Google Scholar] [CrossRef]
  11. Petrova, O.B.; Mayakova, M.N.; Smirnov, V.A.; Runina, K.I.; Avetisov, R.I.; Avetissov, I.C. Luminescent Properties of Solid Solutions in the PbF2-EuF3 and PbF2–ErF3 Systems. J. Lumin. 2021, 238, 118262. [Google Scholar] [CrossRef]
  12. Sevostjanova, T.S.; Khomyakov, A.V.; Mayakova, M.N.; Voronov, V.V.; Petrova, O.B. Luminescent Properties of Solid Solutions in the PbF2–Euf3 System and Lead Fluoroborate Glass Ceramics Doped with Eu3+ Ions. Opt. Spectrosc. 2017, 123, 733–742. [Google Scholar] [CrossRef]
  13. Savikin, A.P.; Egorov, A.S.; Budruev, A.V.; Perunin, I.Y.; Grishin, I.A. Visualization of 1.908-Μm Radiation of a Tm:YLF Laser Using PbF2-Based Ceramics Doped with Ho3+ Ions. Tech. Phys. Lett. 2016, 42, 1083–1086. [Google Scholar] [CrossRef]
  14. Liao, J.; Chen, Q.; Niu, X.; Zhang, P.; Tan, H.; Ma, F.; Li, Z.; Zhu, S.; Hang, Y.; Yang, Q.; et al. Energy Transfer and Cross-Relaxation Induced Efficient 2.78 Μm Emission in Er3+/Tm3+: PbF2 Mid-Infrared Laser Crystal. Crystals 2021, 11, 1024. [Google Scholar] [CrossRef]
  15. Huang, X.; Wang, Y.; Zhang, P.; Su, Z.; Xu, J.; Xin, K.; Hang, Y.; Zhu, S.; Yin, H.; Li, Z.; et al. Efficiently Strengthen and Broaden 3 Μm Fluorescence in PbF2 Crystal by Er3+/Ho3+ as Co-Luminescence Centers and Pr3+ Deactivation. J. Alloys Compd. 2019, 811, 152027. [Google Scholar] [CrossRef]
  16. Li, X.; Zhang, P.; Yin, H.; Zhu, S.; Li, Z.; Hang, Y.; Chen, Z. Sensitization and Deactivation Effects of Nd3+ on the Er3+: 27 Μm Emission in PbF2 Crystal. Opt. Mater. Express 2019, 9, 1698. [Google Scholar] [CrossRef]
  17. Zhou, M.; Zhang, P.; Niu, X.; Liao, J.; Chen, Q.; Zhu, S.; Hang, Y.; Yang, Q.; Yin, H.; Li, Z.; et al. Ultra-Broadband and Enhanced near-Infrared Emission in Bi/Er Co-Doped PbF2 Laser Crystal. J. Alloys Compd. 2022, 895, 162704. [Google Scholar] [CrossRef]
  18. Sorokin, N.I.; Ivanovskaya, N.A.; Buchinskaya, I.I. Ionic Conductivity of Cold Pressed Nanoceramics Pr0.9Pb0.1F2.9 Obtained by Mechanosynthesis of Components. Phys. Solid State 2023, 65, 101. [Google Scholar] [CrossRef]
  19. Sobolev, B.P. The Rare Earth Trifluorides: The High Temperature Chemistry of the Rare Earth Trifluorides. Part 2. Introduction to Materials Science of Multicomponent Metal Fluoride Crystals; Arxius de les seccions de ciències; Institut d’Estudis Catalans: Barcelona, Spain, 2001; ISBN 84-7283-610-X. [Google Scholar]
  20. Achary, S.N.; Patwe, S.J.; Tyagi, A.K. Powder XRD Study of Ba4Eu3F17: A New Anion Rich Fluorite Related Mixed Fluoride. Powder Diffr. 2002, 17, 225–229. [Google Scholar] [CrossRef]
  21. Dombrovski, E.N.; Serov, T.V.; Abakumov, A.M.; Ardashnikova, E.I.; Dolgikh, V.A.; Van Tendeloo, G. The Structural Investigation of Ba4Bi3F17. J. Solid State Chem. 2004, 177, 312–318. [Google Scholar] [CrossRef]
  22. Patwe, S.; Achary, S.; Tyagi, A. Synthesis and Characterization of Ba1−xErxF2+x (0.00 ≤ x ≤ 1.00). Mater. Res. Bull. 2002, 37, 2243–2253. [Google Scholar] [CrossRef]
  23. Patwe, S.J.; Achary, S.N.; Tyagi, A.K. Synthesis and Characterization of Y1-XPb2+1.5xF7 (−1.33 ≤ x ≤ 1.0). Mater. Res. Bull. 2001, 36, 597–605. [Google Scholar] [CrossRef]
  24. Loiko, P.; Rytz, D.; Schwung, S.; Pues, P.; Jüstel, T.; Doualan, J.-L.; Camy, P. Watt-Level Europium Laser at 703 Nm. Opt. Lett. 2021, 46, 2702. [Google Scholar] [CrossRef]
  25. Scott, S.; Barnes, H. Sphalerite-Wurtzite Equilibria and Stoichiometry. Geochim. Cosmochim. Acta 1972, 36, 1275–1295. [Google Scholar] [CrossRef]
  26. Mayakova, M.N.; Voronov, V.V.; Iskhakova, L.D.; Kuznetsov, S.V.; Fedorov, P.P. Low-Temperature Phase Formation in the BaF2-CeF3 System. J. Fluor. Chem. 2016, 187, 33–39. [Google Scholar] [CrossRef]
  27. Karimov, D.N.; Sorokin, N.I.; Chernov, S.P.; Sobolev, B.P. Growth of MgF2 Optical Crystals and Their Ionic Conductivity in the As-Grown State and after Partial Pyrohydrolysis. Crystallogr. Rep. 2014, 59, 928–932. [Google Scholar] [CrossRef]
  28. Inaguma, Y.; Ueda, K.; Katsumata, T.; Noda, Y. Low-Temperature Formation of Pb2OF2 with O/F Anion Ordering by Solid State Reaction. J. Solid State Chem. 2019, 277, 363–367. [Google Scholar] [CrossRef]
  29. Strekalov, P.V.; Mayakova, M.N.; Runina, K.I.; Petrova, O.B. Organic Phosphor and Lead Fluoride Based Luminescent Hybrids. Tsvetnye Met. 2021, 10, 25–31. [Google Scholar] [CrossRef]
  30. Borik, M.A.; Gerasimov, M.V.; Kulebyakin, A.V.; Larina, N.A.; Lomonova, E.E.; Milovich, F.O.; Myzina, V.A.; Ryabochkina, P.A.; Sidorova, N.V.; Tabachkova, N.Y. Structure and Phase Transformations in Scandia, Yttria, Ytterbia and Ceria-Doped Zirconia-Based Solid Solutions during Directional Melt Crystallization. J. Alloys Compd. 2020, 844, 156040. [Google Scholar] [CrossRef]
  31. Crystallography Open Database. Available online: https://www.crystallography.net/cod/1530196.html (accessed on 5 December 2025).
  32. Sorokin, N.I. Molar Volume Correlation between Nonstoichiometric M1 − xRxF2 + x (0 ≤ x ≤ 0.5) and Ordered MmRnF2m + 3n (m/n = 8/6, 9/5) Phases in Systems MF2–RF3 (M = Ca, Sr, Ba, Pb; and R = Rare-Earth Elements). Russ. J. Inorg. Chem. 2019, 64, 351–356. [Google Scholar] [CrossRef]
  33. Anderson, A. The Raman Effect: Applications; M. Dekker: New York, NY, USA, 1971. [Google Scholar]
  34. Bazhenov, A.V.; Smirnova, I.S.; Fursova, T.N.; Maksimuk, M.Y.; Kulakov, A.B.; Bdikin, I.K. Optical Phonon Spectra of PbF2 Single Crystals. Phys. Solid State 2000, 42, 41–50. [Google Scholar] [CrossRef]
  35. Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination Compounds; Wiley: Hoboken, NJ, USA, 2008. [Google Scholar] [CrossRef]
  36. Thangadurai, P.; Ramasamy, S.; Kesavamoorthy, R. Raman Studies in Nanocrystalline Lead (II) Fluoride. J. Phys. Condens. Matter 2005, 17, 863–874. [Google Scholar] [CrossRef]
  37. Kuznetsov, S.V.; Nizamutdinov, A.S.; Madirov, E.I.; Voronov, V.V.; Tsoy, K.S.; Khadiev, A.R.; Yapryntsev, A.D.; Ivanov, V.K.; Kharintsev, S.S.; Semashko, V.V. Near Infrared Down-Conversion Luminescence of Ba4Y3F17:Yb3+:Eu3+ Nanoparticles under Ultraviolet Excitation. Nanosyst. Phys. Chem. Math. 2020, 11, 316–323. [Google Scholar] [CrossRef]
  38. Kishio, K.; Shimoyama, J.; Hasegawa, T.; Kitazawa, K.; Fueki, K. Determination of Oxygen Nonstoichiometry in a High-T c Superconductor Ba2YCu3O7-δ. Jpn. J. Appl. Phys. 1987, 26, L1228. [Google Scholar] [CrossRef]
Figure 1. T-x diagram of PbF2-EuF3 quasi-binary system with the corresponding literature references for different temperature ranges: F—cubic phase; R—orthorhombic phase; L—liquid phase.
Figure 1. T-x diagram of PbF2-EuF3 quasi-binary system with the corresponding literature references for different temperature ranges: F—cubic phase; R—orthorhombic phase; L—liquid phase.
Materials 19 00195 g001
Figure 2. X-ray diffraction patterns of the (1 − x)(PbF2)x(EuF3) samples with 1–9 mol% EuF3 nominal content. Dot lines and Miller indices indicate the reflections of the cubic solid-solution phase, and black bars are reference data for pure cubic β-PbF2 according to COD-1530196 [31].
Figure 2. X-ray diffraction patterns of the (1 − x)(PbF2)x(EuF3) samples with 1–9 mol% EuF3 nominal content. Dot lines and Miller indices indicate the reflections of the cubic solid-solution phase, and black bars are reference data for pure cubic β-PbF2 according to COD-1530196 [31].
Materials 19 00195 g002
Figure 3. X-ray diffraction patterns of Pb0.97Eu0.03F2.03 samples after annealing at different temperatures (dashed lines and Miller indices indicate reflections of the cubic solid-solution phase).
Figure 3. X-ray diffraction patterns of Pb0.97Eu0.03F2.03 samples after annealing at different temperatures (dashed lines and Miller indices indicate reflections of the cubic solid-solution phase).
Materials 19 00195 g003
Figure 4. Dependence of cubic phase volume fraction on temperature for (1 − x)(PbF2)x(EuF3) samples.
Figure 4. Dependence of cubic phase volume fraction on temperature for (1 − x)(PbF2)x(EuF3) samples.
Materials 19 00195 g004
Figure 5. Variation in lattice parameters of the cubic phase after heat treatment of samples with nominal composition Pb1−xEuxF2+x.
Figure 5. Variation in lattice parameters of the cubic phase after heat treatment of samples with nominal composition Pb1−xEuxF2+x.
Materials 19 00195 g005
Figure 6. Normalized PL spectra of Pb0.97Eu0.03F2.03 solid solutions after heat treatment at different temperatures (λexc = 377 nm). The inset shows a diagram of the Eu3+ levels.
Figure 6. Normalized PL spectra of Pb0.97Eu0.03F2.03 solid solutions after heat treatment at different temperatures (λexc = 377 nm). The inset shows a diagram of the Eu3+ levels.
Materials 19 00195 g006
Figure 7. Changes in the asymmetry coefficient after the annealing at different temperatures of samples with a nominal composition of Pb1−xEuxF2+x.
Figure 7. Changes in the asymmetry coefficient after the annealing at different temperatures of samples with a nominal composition of Pb1−xEuxF2+x.
Materials 19 00195 g007
Figure 8. Fragment of the T-X diagram of the quasi-binary system PbF2-EuF3, according to the results of XRD analysis: 1—cubic phase (Fm 3 ¯ m); 2—rhombic phase (Pnma).
Figure 8. Fragment of the T-X diagram of the quasi-binary system PbF2-EuF3, according to the results of XRD analysis: 1—cubic phase (Fm 3 ¯ m); 2—rhombic phase (Pnma).
Materials 19 00195 g008
Figure 9. X-ray diffraction patterns of samples obtained by the solid-phase synthesis (a) and co-precipitation technique (b) in the quasi-binary system PbF2-EuF3 (dashed lines and Miller indices indicate nominally pure β-PbF2 according to COD-1530196 [31]).
Figure 9. X-ray diffraction patterns of samples obtained by the solid-phase synthesis (a) and co-precipitation technique (b) in the quasi-binary system PbF2-EuF3 (dashed lines and Miller indices indicate nominally pure β-PbF2 according to COD-1530196 [31]).
Materials 19 00195 g009
Figure 10. Lattice parameter (a) and lattice volume (b) for solid-solution (100 − x)PbF2−xEuF3 (x = 35–45 mol% EuF3) samples obtained by the co-precipitation (3) and solid-phase methods (2) and calculated by Vegard’s law (1–green line).
Figure 10. Lattice parameter (a) and lattice volume (b) for solid-solution (100 − x)PbF2−xEuF3 (x = 35–45 mol% EuF3) samples obtained by the co-precipitation (3) and solid-phase methods (2) and calculated by Vegard’s law (1–green line).
Materials 19 00195 g010
Figure 11. Normalized PL spectra of samples in the (100 − x)PbF2−xEuF3 system (x = 37–45 mol% EuF3), obtained by the co-precipitation technique (λexc = 377 nm).
Figure 11. Normalized PL spectra of samples in the (100 − x)PbF2−xEuF3 system (x = 37–45 mol% EuF3), obtained by the co-precipitation technique (λexc = 377 nm).
Materials 19 00195 g011
Figure 12. Dependence of the asymmetry coefficient R21 on the EuF3 content in the system (100 − x)PbF2−xEuF3 (x = 37–45 mol% EuF3), obtained by the co-precipitation technique.
Figure 12. Dependence of the asymmetry coefficient R21 on the EuF3 content in the system (100 − x)PbF2−xEuF3 (x = 37–45 mol% EuF3), obtained by the co-precipitation technique.
Materials 19 00195 g012
Figure 13. Overall view of the normalized and Gaussian-fitted Raman spectra of (100 − x)PbF2−xEuF3 samples (x = 37–45 mol% EuF3), obtained by the co-precipitation technique. The insertions show the dependences of Δνmax and FWHM vs. EuF3 concentration. The green line is attributed with the nominal pure cubic β-PbF2 [33].
Figure 13. Overall view of the normalized and Gaussian-fitted Raman spectra of (100 − x)PbF2−xEuF3 samples (x = 37–45 mol% EuF3), obtained by the co-precipitation technique. The insertions show the dependences of Δνmax and FWHM vs. EuF3 concentration. The green line is attributed with the nominal pure cubic β-PbF2 [33].
Materials 19 00195 g013
Figure 14. T-X diagram near quasi-binary PbF2-EuF3 section of ternary Pb-Eu-F system in variable scale. The sketches of isothermal cross-sections of T-X-Y projection of ternary Pb-Eu-F diagram are presented schematically.
Figure 14. T-X diagram near quasi-binary PbF2-EuF3 section of ternary Pb-Eu-F system in variable scale. The sketches of isothermal cross-sections of T-X-Y projection of ternary Pb-Eu-F diagram are presented schematically.
Materials 19 00195 g014
Figure 15. Fragments of the possible isothermal cross-sections of the T-X-Y projection of the ternary Pb-Eu-F diagram in the 50–200 °C temperature range.
Figure 15. Fragments of the possible isothermal cross-sections of the T-X-Y projection of the ternary Pb-Eu-F diagram in the 50–200 °C temperature range.
Materials 19 00195 g015
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

Zykova, S.; Runina, K.; Mayakova, M.; Berezina, M.; Petrova, O.; Avetisov, R.; Avetissov, I. Fundamentals of Cubic Phase Synthesis in PbF2–EuF3 System. Materials 2026, 19, 195. https://doi.org/10.3390/ma19010195

AMA Style

Zykova S, Runina K, Mayakova M, Berezina M, Petrova O, Avetisov R, Avetissov I. Fundamentals of Cubic Phase Synthesis in PbF2–EuF3 System. Materials. 2026; 19(1):195. https://doi.org/10.3390/ma19010195

Chicago/Turabian Style

Zykova, Sofia, Kristina Runina, Mariya Mayakova, Maria Berezina, Olga Petrova, Roman Avetisov, and Igor Avetissov. 2026. "Fundamentals of Cubic Phase Synthesis in PbF2–EuF3 System" Materials 19, no. 1: 195. https://doi.org/10.3390/ma19010195

APA Style

Zykova, S., Runina, K., Mayakova, M., Berezina, M., Petrova, O., Avetisov, R., & Avetissov, I. (2026). Fundamentals of Cubic Phase Synthesis in PbF2–EuF3 System. Materials, 19(1), 195. https://doi.org/10.3390/ma19010195

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