Next Article in Journal
A Green Electromagnetic Energy Harvester with Up-Frequency and Unidirectional Rotation for Smart Pavement
Next Article in Special Issue
Preparation and Properties of Boron Modified Phenolic Resin for Automotive Friction Materials
Previous Article in Journal
Effect of Macro Fibers on Chloride Permeability and Damage of Concrete Under Uniaxial Compression
Previous Article in Special Issue
Microstructure and Bioactivity of Ca- and Mg-Modified Silicon Oxycarbide-Based Amorphous Ceramics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Glass-Ceramics of the Lithium Aluminosilicate System Nucleated by TiO2: The Role of Redox Conditions of Glass Melting in Phase Transformations and Properties

1
Institute of Machinery, Materials, and Transport, Peter the Great St. Petersburg Polytechnic University, St. Petersburg 195251, Russia
2
S.I. Vavilov State Optical Institute, St. Petersburg 192171, Russia
3
Ioffe Institute, St. Petersburg 194021, Russia
4
International Research and Education Center for Physics of Nanostructures, ITMO University, St. Petersburg 199034, Russia
5
Department of Glass and Glass-Ceramics, Mendeleev University of Chemical Technology, Moscow 125047, Russia
6
D.V. Efremov Institute of Electrophysical Apparatus, St. Petersburg 189631, Russia
*
Author to whom correspondence should be addressed.
Materials 2025, 18(4), 785; https://doi.org/10.3390/ma18040785
Submission received: 16 January 2025 / Revised: 4 February 2025 / Accepted: 7 February 2025 / Published: 11 February 2025

Abstract

:
Lithium aluminosilicate glasses nucleated by TiO2 are usually melted in oxidizing conditions. The reducing conditions of glass melting, which allow to obtain ions of variable valence in lower oxidation states, can influence the ability of titania to provide proper phase assemblage, structure and properties of lithium aluminosilicate glass-ceramics. The aim of this study is to reveal this influence. The model glass containing TiO2 was melted with and without the addition of As2O3. Using heat treatments between 680 °C and 1300 °C, XRD, SEM and DSC data, Raman and absorption spectroscopy, transparent glass-ceramics based on nanocrystals of β-quartz solid solutions (ss) and/or γ-Al2O3 with spinel structure and opaque glass-ceramics based on nanocrystals of β-spodumene ss were obtained and characterized. Three-phase immiscibility develops during secondary heat treatments. Al2TiO5 crystallizes from aluminotitanate amorphous regions simultaneously with the appearance of β-quartz ss, while traces of anatase and then rutile appear at elevated temperatures. Phase assemblage and the sequence of phase transformations do not depend on the redox conditions of glass melting, while the rate of these transformations is significantly higher in glass melted without the addition of As2O3. Absorption in glass melted without the addition of As2O3 and the corresponding glass-ceramics originate from octahedrally coordinated Ti3+ ions and Ti3+-Ti4+ pairs in glass and nanocrystals of γ-Al2O3, Al2TiO5 and β-quartz ss. Transparent glass-ceramics with a thermal expansion coefficient of ~0.3 × 10−6 K−1 were obtained from both glasses.

1. Introduction

Transparent lithium aluminosilicate glass-ceramics combine the unique properties of transparency and a close-to-zero thermal expansion coefficient [1,2,3]. Developed by Stookey more than 60 years ago [4], glass-ceramics of the lithium aluminosilicate system still remain a subject of intensive studies [5,6,7,8,9,10,11,12,13,14,15]. Stookey found out [4] that titania, TiO2, is an effective nucleating agent promoting crystallization of lithium aluminosilicate glasses of special compositions upon heat treatments. The role of titania as a nucleating agent of the lithium aluminosilicate system was studied by different structure-sensitive methods [16,17,18,19,20,21,22] and mechanisms controlling the formation of titania-doped glass-ceramics were suggested. It was demonstrated that they depend on the base glass composition, the nature and concentration of doping ions, and the amount of the nucleating agent. Being a polyvalent ion, titanium can be found in glasses in the form of Ti4+ and Ti3+ ions. Oxidation states of titanium ions depend on the glass melting temperature, oxygen partial pressure in the glass-melting furnace, the concentration of titanium ions and the glass composition [18,23,24,25,26,27,28,29]. Using titania as a nucleator, oxidizing agents are usually added to the glass batch to avoid unwanted coloration caused by traces of Ti3+ ions formed during the glass melting at high temperatures [1,2]. However, the reducing atmosphere of glass melting is required to obtain polyvalent ions in lower oxidation states. Therefore, it is worth knowing how the reducing atmosphere of glass melting influences phase transformations in glasses of the lithium aluminosilicate system nucleated by TiO2 and properties of corresponding glass-ceramics.
There are studies of phase transformations in titania-containing glasses of magnesium aluminosilicate [18,29,30], fresnoite [31,32,33], and zinc aluminosilicate [34] systems melted under different redox conditions. In these studies, it was demonstrated that variation of the glass melting conditions influences kinetics of liquid–liquid phase separation and of crystallization of metastable phases, compositions and structures of titania-containing phases and generally does not influence crystallization of the stable equilibrium phases [18,29,34]. The aim of the present work is to study the effect of redox conditions of glass melting on the phase composition and properties of lithium aluminosilicate glass-ceramics nucleated by titania, which will help to develop transparent thermal shock resistant glass-ceramics of the lithium aluminosilicate system containing polyvalent ions in lower oxidation states important for advanced photonic applications [35].

2. Materials and Methods

2.1. Materials Preparation

Model glass with the composition 12 Li2O∙24 Al2O3∙64 SiO2 (mol%) [17,19] nucleated by 6 mol% TiO2 added on top of the base composition was melted in oxidizing (with addition of 0.5 wt% As2O3) and neutral (without addition of arsenic trioxide) conditions in a laboratory electric furnace in crucibles made of quartz ceramics at 1580 °C for 4 h (h) with stirring and casted onto a metal plate. The raw materials were reagent grade oxides and lithium carbonate manufactured by Neva-Reactive Co, Saint-Petersburg, Russian Federation. The weight of the glass batch was 400 g. The glasses were annealed at 640 °C for 1 h and cooled with an annealing furnace to room temperature. Small portions of the glasses were quenched by pressing with a metal plate and used for the differential scanning calorimetry (DSC) study. Annealed glasses were transparent and different in color. The glass melted under neutral conditions, referred to as LAS, exhibited a brownish-grey color, while the LASox glass, prepared under oxidizing conditions, displayed a light-yellow tint. The color difference is clearly seen in photographs of the as-cast glasses with a thickness of ca. 6 cm (see Figure 1).
The glasses were cut into pieces with a diamond saw and isothermally heat-treated at temperatures between 680 °C and 1200 °C by single and two-stage heat treatments for 6 h at each stage.

2.2. Materials Characterization

2.2.1. The Differential Scanning Calorimetry

The differential scanning calorimetry study was performed using a simultaneous thermal analyzer NETZSCH STA 449 F3 Jupiter, Netzsch Geratebau GmbH, Selb, Germany, with an Ar dynamic flow atmosphere. The temperature was increased from room temperature to 1300 °C at a heating rate of 10 °C⋅min−1. The quenched glasses and those subjected to single stage holdings at 680 °C, 700 °C, and 720 °C for 6 h, with a weight of about 15–20 mg were employed. To determine the character of crystalline phases responsible for exothermic effects on the DSC curves, quenched and heat-treated glasses with a weight of ca. 100 mg were heated in the furnace of the thermal analyzer up to corresponding temperatures. The glass-ceramics obtained were taken out from the cooled furnace and studied by X-ray diffraction (XRD) analysis.

2.2.2. XRD Analysis

The XRD study of the powdered samples was performed using a Shimadzu XRD-6000 diffractometer, Shimadzu, Kyoto, Japan, with Cu Kα radiation and a Ni filter (λ = 1.5406 Å). The mean crystal sizes were calculated from the broadening of X-ray peaks according to Scherrer’s equation:
D = Kλ/Δ(2θ)cosθ,
where λ is the wavelength of X-ray radiation, θ is the diffraction angle, Δ(2θ) is the width of peak at half of its maximum, and K is the constant assumed to be 1 [36]. The mean size of γ-Al2O3 crystals with a spinel structure was estimated using the peak with Miller’s indices hkl (440). The mean size of tieilite, Al2TiO5, crystals was determined from the peak with Miller’s indices (020). The mean size of crystals of β-quartz ss was determined from the peak with Miller’s indices (220). The mean size of crystals of β-spodumene ss was calculated using the peak with indices (102). An error of the calculation of the mean size of crystallites is about 5–10%.
The lattice parameter a of spinel nanocrystals was estimated from the position of the peak with hkl indices (440). The lattice parameters of Al2TiO5 were estimated using the peaks with indices (002), (020), (110), (023) and (200). The lattice parameters of β-quartz ss were calculated from the peaks with indices (110) and (211). The lattice parameters of β-spodumene ss were calculated from the peaks with hkl indices (111) and (102). An error in the estimation of the lattice parameters is ±0.003 Å.

2.2.3. Raman Spectroscopy

Unpolarized Raman spectra were obtained using a backscattering configuration with a confocal InVia Renishaw Raman microscope, Renishaw, Wotton-under-Edge, England. The setup included a Leica objective with a magnification of ×50 (N.A. = 0.75), a thermoelectrically cooled CCD camera, and an edge filter. For excitation, a laser with a wavelength of 488 nm (from an Ar+ ion source) was employed. The spectral resolution achieved during the measurements was 2 cm−1. Each spectrum was obtained by averaging data from 10 acquisitions, each lasting 10 s. The samples were either transparent polished flat-parallel plates having a thickness of ~1 mm or opaque pieces of materials.

2.2.4. Scanning Electron Microscopy

The morphology of the initial and heat-treated glasses was obtained by scanning electron microscopy (SEM) using a Tescan Vega 3 SBH microscope, Tescan, Brno, Czech Republic. For the study, the surfaces of the same transparent polished plates and opaque bulk samples, which were used for measuring Raman spectra, were preliminarily cleaned in isopropyl alcohol and benzene, etched in a solution of HF for about 2 s and washed in distilled water. ImageJ software version 1.51j8 [37] was used to calculate the particle sizes.

2.2.5. Absorption Spectroscopy

A Shimadzu UV-3600 spectrophotometer, Shimadzu, Japan, was employed for recording absorption spectra within the wavelength range from 250 to 3300 nm using the same polished ~1 mm thick flat-parallel plates.

2.2.6. The Linear Coefficient of Thermal Expansion

The thermal expansion coefficients of the glasses and glass-ceramics were estimated using a Linseis L 75 vs. 1000 dilatometer, Linseis Messgeräte GmbH, Selb, Germany. The samples were rods with a length ranging from 30 to 50 mm and flat-parallel planes with sections measuring ~5 × 5 mm. The measurements were performed at room temperature up to 320 °C with a heating rate of 5 °C⋅min−1.

3. Results

The polished plates of transparent glasses and glass-ceramics and the pieces of opaque glass-ceramics are shown in Figure 2. Note that the thin polished plate of the glass made from the brownish-grey as-cast LAS glass has a light grey coloration because of a six times smaller thickness. The initial glasses, heat-treated at 680 °C for 6 h and by two-stage heat treatments with the second hold at temperatures between 720 °C and 900 °C, are transparent. The LAS glass-ceramics obtained at 950 °C and at 1000 °C are also transparent, while LASox glass-ceramics obtained at these temperatures are translucent. Both glass-ceramics obtained at 1100 °C are translucent, and those prepared at 1200 °C are opaque. Therefore, glasses melted under neutral conditions lose transparency at higher secondary heat treatment temperatures than glasses melted under oxidizing conditions. The opaque LAS glass-ceramic prepared at 1200 °C is grey colored and has a white surface layer, see Figure 2a.

3.1. Study by Differential Scanning Calorimetry

Figure 3 shows the DSC scans of the quenched LAS and LASox glasses, as well as these glasses heat-treated at 680 °C, 700 °C and 720 °C for 6 h. The shapes of the corresponding DCS curves of LAS and LASox glasses are nearly similar, being different in the values of characteristic crystallization temperatures, see Figure 3 and Table 1. The DSC curve of the quenched LASox glass shows three peaks with markedly different intensities. The first one is a broad peak of a low intensity with Tmax = 802 °C, the second one is a narrow and intense peak with crystallization onset temperature Ton = 851 °C, and crystallization maximum temperature Tmax = 874 °C. The third one is a weak exothermic peak with Tmax = 1186 °C. The first peak is usually assigned to nucleation of the main crystalline phase, the second one to the crystallization of the lithium aluminosilicate solid solution (ss) with β-quarts structure (β-quarts ss), and the third one to the transformation of β-quartz ss into β-spodumene ss [38,39]. The DSC curve of the LAS glass exhibits a plateau instead of the first peak, see Figure 3 and Table 1. All characteristic crystallization temperatures of the LAS glass are lower than those for the LASox glass, while the glass transition temperatures of both glasses are similar, see Table 1. Therefore, the redox conditions of glass melting affect crystallization temperatures and do not affect the transition temperature of glasses.
Preliminary heat treatments lead to a drastic change in the appearance of DSC curves, see Figure 3. The character of this change for LAS and LASox samples is the same, differing in the temperatures of thermal effects. Instead of a single narrow intense peak, two broad exothermic peaks of lower intensities appear. The characteristic temperatures for both glasses preliminarily heat-treated at 680 °C, 700 °C, and 720 °C are shown in Figure 3 and Table 1. Tmax of the second broad exothermic peak is 35–40 °C higher than this temperature for the initial quenched glass. With the increase in the temperature of preliminary heat treatment, there is a redistribution of intensities of these two peaks in favor of the second one, see Figure 3. Tmax of the second peak and Tg gradually increase, see Table 1.
The reason for this behavior will be explained below while discussing the phase composition of the samples. The characteristic Tg, Ton and Tmax temperatures of glasses preliminary heat-treated at 680 °C are similar, being somewhat higher for the LAS glass than for the corresponding LASox samples.
The XRD pattern of the glass-ceramic obtained by heating the quenched LAS glass in the furnace of the DSC instrument up to the temperature of the sharp exothermic peak at 880 °C, see Figure 4a, testifies to the crystallization of β-quarts ss, traces of tieilite, Al2TiO5, and spinel, see Figure 4b. On the XRD pattern of the sample heated up to the temperature of the third exothermic effect at 1175 °C, additional diffraction peaks appear as compared with the pattern of the sample heated up to 880 °C, manifesting in a transformation of the β-quarts ss into the β-spodumene ss while spinel traces disappeared. Thus, the small high-temperature peak evolving at ~1175 °C is due to crystallization of β-spodumene ss, see Figure 4b.
Figure 5a shows the DSC signals of the LAS glass preliminarily heat-treated at 680 °C for 6 h with stops at temperatures of 787 °C, 838 °C, 858 °C, and 935 °C. The corresponding XRD patterns are presented in Figure 5b. The material obtained by the heat treatment at 680 °C, 6 h + 787 °C, 0 h is X-ray amorphous. The sample obtained by the heat treatment at 680 °C, 6 h + 838 °C, 0 h contains nanocrystals of γ-Al2O3 with a spinel structure and traces of β-quarts ss. It is speculated that this exothermic peak is caused by the crystallization of spinel, while traces of β-quarts ss appear during the cooling of the sample in the furnace, because the samples heated to the specified temperature could not be immediately removed from the hot furnace of the DSC instrument. The XRD pattern of the sample obtained by the heat treatment at 680 °C, 6 h + 858 °C, 0 h, corresponding to the end of the first exothermic peak and the beginning of the second one, contains peaks of nanocrystals of γ-Al2O3, Al2TiO5 and β-quarts ss. The XRD pattern of glass-ceramic obtained by the heat treatment at 680 °C, 6 h + 935 °C, 0 h proves the crystallization of β-quarts ss and Al2TiO5. The presence of γ-Al2O3 cannot be ruled out on this XRD pattern because the most intense peak of γ-Al2O3 overlaps the peaks of β-quarts ss and is therefore difficult to detect.
The XRD pattern of the glass preliminarily heat-treated at 680 °C for 6 h and heated in the DSC furnace up to 858 °C is significantly different from the XRD pattern of the quenched glass heated in the DSC furnace up to 880 °C. The XRD pattern of the glass preliminarily heat-treated at 680 °C for 6 h and heated in the DSC furnace up to 935 °C is similar to the XRD pattern of the quenched glass heated in the DSC furnace up to 880 °C, see Figure 6. Thus, the first peak on the DSC curve of the preliminary heat-treated sample is associated with the crystallization of the phase with the spinel structure, and the second peak is caused by the crystallization of β-quartz ss and Al2TiO5. On the DSC curve of the quenched sample there is no exothermic peak associated with the crystallization of the phase with spinel structure. Therefore, preliminary heat treatments provoke spinel crystallization and increase the crystallization temperature of β-quartz ss and Al2TiO5.

3.2. XRD Study

Figure 7a,b shows XRD patterns of the LAS and LASox glasses and glasses of the same compositions, which underwent heat treatment at 680 °C and two-stage heat treatments with a temperature at the second stage between 720 °C and 1200 °C. The holding time at each stage was 6 h. The figures allow the following of the formation of the main crystalline phase, β-quartz ss, at temperatures between 780 °C and 1000 °C, and crystallization of β-spodumene ss at temperatures between 1100 °C and 1200 °C. Figure 7c,d gives a closer look at the XRD patterns of the glasses and glass-ceramics fabricated by heat treatments at the second stage up to 800 °C. Initial LAS and LASox glasses are X-ray amorphous with a maximum amorphous halo located at 2θ = 23.1° and 23.3°, respectively. Glasses heat-treated at the nucleation stage of 680 °C for 6 h are X-ray amorphous, and their XRD patterns are similar to those of glass precursors. The two-stage heat treatment with the second stage at 720 °C for 6 h leads to crystallization of a small fraction of crystallites with spinel structure, manifested by an appearance of a broad peak with Miller indices hkl (440) at 2θ ≅ 66.8°.
Its low intensity prevents estimation of the lattice parameter and the size of the spinel crystallites. The position of the amorphous halo shifts to a smaller angle 2θ = 22.8° for both samples. After the second stage heat treatment at 750 °C for 6 h, spinel peaks become more pronounced, which indicates an increase in the volume fraction of this phase. The lattice parameters of spinel nanocrystals are anorm = 7.918 Å and aox = 7.916 Å, and sizes are Dnorm = 4.5 nm and Dox = 7.3 nm, respectively. Traces of β-quartz ss are also seen on the patterns. The appearance of spinel and traces of β-quartz ss cause a further change in the composition of the residual glass phase, which is manifested by a shift in the maximum of the amorphous halo to smaller angles, indicating that the residual glass becomes enriched in silica. After heat treatments at the second stage at 780 °C and 800 °C, the β-quartz ss becomes the predominant crystalline phase, while spinel and tieilite, Al2TiO5, nanocrystals are also found in the XRD patterns. The position of the amorphous halo of residual glass is shifted to 2θ = 22.6°and then to 2θ = 22.5° with increasing heat-treatment temperature, see Figure 7c,d.
The lattice parameters and mean crystallite sizes of spinel obtained at different heat-treatment temperatures are presented in Table 2. It is believed that spinel has the composition and structure of γ-Al2O3. The lattice parameter a of the unit cell of the cubic modification of Al2O3 usually takes the value a = 7.900–7.908 Å [40]. With a slight oxygen deficiency, the parameter a of γ-Al2O3 becomes equal to 7.911 Å (ICDD PDF card #79-1558) and 7.914 Å (ICDD PDF card 79-1557) [41]. The lattice parameter a of γ-Al2O3 crystallites in glass-ceramics changes from 7.915 Å to 7.925 Å and increases with increasing the heat-treatment temperature. The difference in the parameter a of γ-Al2O3 nanocrystals in LAS and LASox glass-ceramics is very small. Nevertheless, the lattice parameter a of γ-Al2O3 in LAS glass-ceramics is slightly higher than in LASox ones. The sizes of spinel crystallites increase with increasing the heat-treatment temperature ranging from 4.5 to 14.0 nm, see Table 2.
Figure 8a shows that in spite of the same phase assemblage of glass-ceramics obtained by two-stage heat treatment at 680 °C, 6 h + 780 °C, 6 h, the crystallinity fractions in the LAS glass-ceramic is much higher than in the LASox one. This difference is levelled out by the heat treatment at 680 °C, 6 h + 800 °C, 6 h, see Figure 8b.
Figure 7e,f show the XRD patterns of glass-ceramics obtained by two-stage heat treatments with a second-stage temperature between 850 °C and 1200 °C. After the heat treatment at 850 °C the crystallinity fraction increases to such extent that the amorphous halo disappears. Volume fractions of β-quartz ss and Al2TiO5 gradually increase. β-spodumene ss crystallizes at the expense of β-quartz ss at 1100 °C. Mullite (ICDD PDF card # 79-1458) appears at the same heat-treatment temperature. The lattice parameters and mean crystallite sizes of β-quartz ss and β-spodumene ss are presented in Table 2. Their variation with heat-treatment temperature is similar for the LAS and the LASox glasses.
As was mentioned above, the appearance of the surface and the volume of the LAS glass-ceramics obtained by the heat treatment at 1200 °C are different. The surface is a dense, white-colored layer while the volume is grey colored, see Figure 2. Traces of rutile (ICDD PDF card #78-1509) are found on the XRD patterns taken from the grey part of the LAS glass-ceramic. The appearance of rutile is accompanied by some decrease in the tieilite fraction. We failed to find a difference in the phase assemblages of the surface and volume of LASox glass-ceramic obtained by the same heat treatment. The XRD pattern shows a tiny fraction of rutile in this sample.
Let us discuss the features of crystallization of tieilite, Al2TiO5. As mentioned above, the onset of the crystallization of β-quartz solid solution in glasses at a temperature of 780 °C is accompanied by the appearance of a small fraction of Al2TiO5 crystallites with mean sizes of ~6 and 8 nm for LAS and LASox glass-ceramics, respectively, see Table 3. Increasing the heat-treatment temperature up to 1200 °C leads to an increase in the size of Al2TiO5 crystallites up to ~40 nm, a slow increase in their fraction and a change in their lattice parameters, see Table 3. The values of the lattice parameter do not change with the heat-treatment temperature within the measurement error and are independent of the redox conditions of glass melting. Lattice parameters b, c and the volume increase with the heat-treatment temperature. An analysis of the evolution of lattice parameters shows that the lattice of a Al2TiO5 unit cell, which has the shape of a rectangular parallelepiped elongated in [010] and [001] directions, becomes even more elongated in the same directions under the influence of high temperature. The lattice parameters of tieilite in LAS and LASox glass-ceramics are similar. The character of their variation with temperature is also similar. The redox conditions of glass melting do not significantly affect the structural transformations in tieilite crystallites. A pronounced change in lattice parameters after the heat treatment at 1200 °C found in the present study can be a prerequisite to decomposition of tieilite with exsolution of rutile [42]. Indeed, as mentioned above, traces of rutile are found in LASox and in the volume of the LAS glass-ceramics obtained by the heat treatment at 1200 °C. The comparison of XRD patterns of the LAS and the LASox glass-ceramics obtained by the heat treatment at 1200 °C demonstrates that LASox glass-ceramic contains a smaller fraction of rutile and a higher fraction of mulite than the LAS glass-ceramic, compare Figure 7e,f.
The sequence of phase transformations revealed by XRD study is similar for both glasses: precursor glasses and glasses subjected to the nucleation treatment are X-ray amorphous; nanocrystals of γ-Al2O3 with spinel structure and sizes ranging from 4.5 nm to 14 nm evolve during heat treatments in the second stage in the temperature range from 720 °C to 800 °C; β-quartz ss and the crystalline phase of the nucleating agent, tieilite, appear additionally to spinel during the heat treatment at 780 °C; β-quartz ss are obtained by heating in the interval between 780 °C and 1000 °C; the glass-ceramic prepared at 1100 °C contains β-spodumene ss instead of β-quartz ss and traces of mullite; β-spodumene ss is the main crystalline phase of glass-ceramics produced in the temperature range between 1100 °C and 1200 °C; tieilite nanocrystals with a mean size ranging from ~6 nm to ~45 nm are found in glass-ceramics prepared in the temperature interval from 800 °C to 1200 °C; during the heat treatment at 1200 °C, the rutile crystallites appear in the phase assemblage of glass-ceramics.
In spite of the same phase assemblage of glass-ceramics prepared from glasses melted in different redox conditions, the kinetics of phase transformations and spinel lattice parameters are slightly different. The spinel lattice parameters in the LAS glass-ceramics are larger than in the LASox glass-ceramics. The crystallinity fractions in the LAS glass-ceramic obtained by two-stage heat treatment at 680 °C, 6 h + 780 °C, 6 h is much higher than in the LASox one. However, this difference is levelled out by the heat treatment at 680 °C, 6 h + 800° C, 6 h. The rutile crystallinity fraction in the LAS glass-ceramics is larger than in the LASox one, while the mullite fraction is smaller.

3.3. Study by Raman Spectroscopy

Figure 9a,b show Raman spectra of the initial and heat-treated LAS glass. Raman spectra of initial LAS and LASox glasses are similar and contain broad bands with maxima at 482 cm−1, 800 cm−1, 910 cm−1 and ~1015 cm−1. The wing of the latter band extends to 1200 cm−1. A similar spectrum was obtained in ref. [19] for the glass of the same composition nucleated by 7 mol% TiO2 and melted in oxidizing conditions. The bands at 482 cm−1, 800 cm−1, and ~1000–1200 cm−1 are due to vibrations of bonds in the tetrahedrons of the aluminosilicate network in the glass structure [43], and the band at ~910 cm−1 is due to vibrations of [TiO4] tetrahedra embedded in this network [44]. After the heat treatment at the nucleation stage at a temperature of 680 °C, minor changes are noticed in the Raman spectrum, see Figure 9a,b. The high-frequency band slightly broadens and the band with a maximum at 800 cm−1 somewhat increases in intensity compared to the band with a maximum at 910 cm−1, which indicates the development of liquid–liquid phase separation of the precursor glass [19]. The position of the band at 482 cm−1 is not influenced by this heat treatment.
After the two-stage heat treatment with a temperature of 720 °C at the second stage, significant changes are observed in the Raman spectrum, see Figure 9a,b. The bands with maxima at 910 cm−1, 800 cm−1, and 482 cm−1 move their positions to 890 cm−1, 810 cm−1, and 472 cm−1, respectively. Instead of the band with a maximum at ca. 1015 cm−1, a band at 1055 cm−1 appears. The intensity of the band at 472 cm−1 increases as compared with the similar band in spectra of precursor glass and glass heat treated at the nucleation stage. The intensity of the band at ~810 cm−1 increases relatively to the band at ca. 890 cm−1. The increase in the intensity of this band at the expense of the band with a maximum at ca. 890 cm−1 is due to the superposition of vibrations of the [TiO5] and [TiO6] groups in amorphous aluminotitanate regions on a weak band in the 800 cm−1 region, corresponding to vibrations of the tetrahedrons of the aluminosilicate network [17,19,20]. These changes are associated with a further development of the liquid–liquid phase separation with the formation of aluminotitanate amorphous regions [19,20] and a change in the composition of the aluminosilicate glass network, which is in accordance with a change in the position of the amorphous halo in the corresponding XRD pattern, see Figure 7c. Crystallization of γ-Al2O3 with spinel structure revealed by XRD analysis does not show itself in the Raman spectrum of this glass-ceramic. Running ahead, one will say that spectral features of γ-Al2O3, were not detected in Raman spectra of samples obtained by heat treatments between 750 °C and 800 °C as well, in spite of the fact that XRD analysis reliably identified them. The reason is that bond vibrations in aluminate spinel crystals are very weak compared to bond vibrations in titanium-containing compounds. According to refs [45,46], γ-Al2O3 exhibits narrow peaks with maxima at 315, 410, 520, 713 and 835 cm−1 in the Raman spectrum of γ-Al2O3 corresponding to vibrations of the Al–O bond in tetrahedral structural units of AlO4 [45]. Based on calculations presented in ref. [47], the strongest Raman peak for γ-Al2O3 with a spinel structure is located at ~401 cm−1. There are also several bands of medium intensities ranging from 100 cm−1 to 900 cm−1 [47]. In Raman spectra presented in Figure 9b there is a band with a maximum at 403 cm−1, which could be the spectroscopic sign of γ-Al2O3. However, this band can be also caused by vibrations in tieilite crystals, because this band appears simultaneously with other bands at 250 cm−1, 310 cm−1, and 890 cm−1, characteristic of tieilite, which crystallizes at a larger temperature interval of heat treatments than γ-Al2O3 (see below). It is worth mentioning that the spectroscopic signs of crystals with a spinel structure were found only in Raman spectra of spinel-based glass-ceramics of magnesium [48] and zinc aluminosilicate systems [34], where the spinel crystallinity fraction was significantly higher than in the present case.
In the Raman spectrum of the glass-ceramic obtained by the heat treatment at the second stage at 750 °C, the bands with maxima at 472 cm−1 and 815 cm−1 are enhanced, positions of the band maxima change from 810 cm−1 to 815 cm−1 and from 1055 cm−1 to 1068 cm−1, and a very weak band appears at ~144 cm−1. The latter peak can be attributed to the most intense vibration in the Raman spectrum of the metastable modification of TiO2, anatase [17].
In the Raman spectrum of the sample obtained by the heat treatment at the second stage at a temperature of 780 °C, the band at ~144 cm−1, caused by vibrations in anatase nanocrystals, narrows, the band at 480 cm−1 narrows and intensifies, and the band at ~1080 cm−1 appears. The last two bands are related to vibrations in nanocrystals of β-quartz ss [17,19,49,50,51]. As mentioned above, in the spectrum of these glass-ceramics there is a number of bands with maxima at 250 cm−1, 310 cm−1, 403 cm−1, 540 cm−1, and 890 cm−1. They appear simultaneously, and their intensities increase continuously with the increasing temperature of heat treatment. These bands belong to vibrations in tieilite nanocrystals [17,19], which is in accordance with the XRD data. The Raman spectrum of glass-ceramic obtained by heat treatment with a temperature of 900 °C at the second stage shows two anatase bands at ~144 cm−1 and ~638 cm−1 manifesting the maximum anatase crystallinity fraction achieved by this heat treatment, see Figure 9a,c. After increasing the temperature of the second hold to 1000 °C, the bands related to anatase disappear. Note that the XRD peaks of anatase coincide in position with the main peaks of β-quartz ss and its small amount cannot be detected by the XRD data [17,19].
As the temperature at the second stage of heat treatment increases, the intensities of the bands at ~460 cm−1 and ~1080 cm−1 increase, and their peaks become narrower, see Figure 9c, which indicates the development of β-quartz ss crystallization. Their positions constantly move to longer wave numbers manifesting enrichment of these crystals with silica [50]. After the second stage heat treatment at 1100 °C, these bands shift to 490 cm−1 and 1120 cm−1, which can be interpreted as recrystallization of β-quartz ss into β-spodumene ss [17,19]. Starting from the heat treatment at 800 °C, a weak band appears at 110 cm−1 simultaneously with increasing intensity and narrowing of the bands attributed to β-quartz ss. This band can correspond to external vibrations in nanocrystals of β-quartz ss [51]. In Raman spectra of glass-ceramics obtained by heat treatments at 1100 °C and at 1200 °C, weak bands at ~118 cm−1 and ~190 cm−1 were observed that can be attributed to external vibrations in nanocrystals of β-spodumene ss [49], see Figure 9d.
Figure 9d shows the bands at ~260 cm−1, ~445 cm−1 and ~600 cm−1, which are related to crystallization of rutile [17] at the expense of tieilite in the bulk of glass-ceramic obtained by the heat treatment at a second stage at 1200 °C. The signs of rutile crystals were not found in the spectrum from the white surface of this sample, which is in accordance with XRD data. XRD patterns of glass-ceramics obtained by heat treatments at 1100 °C and 1200 °C contain small mullite fractions, see Figure 7e,f. Raman spectra of synthetic mullites of different compositions have similar spectra with the strongest bands located at 415 cm−1, 600 cm−1, and 980 cm−1 [52]. They can be hidden in the contours of the corresponding broad bands in Raman spectra, see Figure 10f.
Raman spectra of glass-ceramics obtained by the same heat treatments that differ most strongly from each other according to Raman spectroscopy data are presented in Figure 10a–f. Raman spectra of glass precursors, as well as the spectra of glasses heat-treated at 680 °C for 6 h are similar to each other. The spectra of glass-ceramics obtained by the two-stage heat treatment at 680 °C and at 720 °C for 6 h differ from each other by the ratio of band intensities in the region of high wave numbers, see Figure 10a. The increase in the relative intensity of the band with a maximum at 800 cm−1 compared to the intensity of the band at 890 cm−1 in the spectrum of the LAS glass-ceramic suggests that the rate of liquid–liquid phase separation in this glass during this heat treatment is higher than that in the LASox glass, i.e., neutral conditions of glass melting speed up the development of liquid–liquid phase separation in this glass. Crystallization of the titanate phases of anatase and tieilite are sped up in the LAS glass ceramized by the heat treatment at the second stage at 780 °C as compared with the LASox glass, see Figure 10b. The same tendency remains true for glass-ceramics obtained by heat treatment with the second stage at 900 °C, see Figure 10c. The higher intensity of the anatase peaks in the Raman spectrum of the LASox glass-ceramic obtained by heat treatment at 950 °C, see Figure 10d, means that anatase crystallization in the LASox glass reaches its maximum and will decrease at higher temperatures while the maximum anatase content in the LAS glass-ceramics was reached at a previous holding temperature. The traces of anatase crystals remain in the LASox glass-ceramic obtained by heat treatment at 1000 °C while the LAS glass-ceramic does not show spectral features of this phase, and intensities of peaks assigned to tieilite crystals are higher for the LAS glass-ceramic than for the LASox one, see Figure 10e. Figure 10f shows that the LAS glass-ceramic prepared by heat treatment at 1200 °C demonstrates rutile bands of higher intensities as compared with the LASox glass-ceramic.
Though Raman spectra of precursor glasses and glasses heat-treated at 680 °C for 6 h are similar to each other, a comparison of the Raman spectra of the materials obtained by two-stage heat treatments unambiguously indicates that, despite the fact that the sequence of transformations in the titania-containing phase is independent of the redox conditions of glass melting, the rate of these transformations is significantly higher at ceramming of the LAS glass.

3.4. Morphology Characterization by SEM

SEM analysis (Figure 11a) reveals the presence of inhomogeneous regions within the bulk of the initial amorphous LAS glass. The calculated size distribution shows a broad profile, with the mean size of the inhomogeneous regions being approximately 27 nm (Figure 12). These inhomogeneities may indicate liquid–liquid phase separation during glass formation. The broad size distribution may arise from overlapping size distributions of chemically distinct regions. Such regions are likely precursors to the crystallization of various phases upon subsequent heat treatments.
Following the single-step heat treatment at 680 °C for 6 h, the LAS glass remains X-ray amorphous. However, its SEM image (Figure 11b) shows increased inhomogeneity, clearly distinguishing two types of regions. The first type consists of small, spherical, bright regions with a narrow size distribution averaging 16 nm (Figure 12). The second predomonant type features larger, irregular, dark regions with an average size of 74 nm (Figure 12).
After two-stage heat treatment at 680 °C for 6 h and 750 °C for 6 h, the large inhomogeneous regions are not seen on the SEM image anymore. Numerous spherical particles appear in the SEM image of the LAS glass-ceramic, averaging 23 nm in size (Figure 11c). The broad shape of the size distribution suggests overlap from multiple chemically distinct particle populations, see Figure 12. XRD analysis confirms the crystallization of the spinel phase with a mean crystallite size of 4.5 nm.
The heat treatment at 680 °C for 6 h and 780 °C for 6 h promotes significant crystallization of the LAS glass. The SEM image of this glass-ceramic is shown in Figure 11d. XRD analysis confirms the formation of three crystalline phases, β-quartz ss, spinel, and Al2TiO5, see Figure 7c, with mean crystallite sizes of 8 nm, 26 nm and ~6 nm, respectively, see Table 2 and Table 3. The particle size distribution suggests the possible overlap of multiple size distributions corresponding to different phases, see Figure 12. The mean particle size is approximately 21 nm, which is close to the mean size of the predominant crystalline phase of β-quartz ss.
The two-step heat treatment at 680 °C and 800 °C for 6 h results in more extensive crystallization of the LAS glass, see Figure 7e. XRD analysis confirms the presence of crystalline phases of β-quartz ss, spinel, and Al₂TiO₅. The SEM image reveals significant etching of the material, characterized by numerous large caverns, which may indicate silica depletion, see Figure 11f. The mean particle size is 31 nm, see Figure 12, which is also close to the mean size of the predominant crystalline phase of β-quartz ss, which is 25 nm, see Table 2.
Following the two-step heat treatment at 680 °C and 1000 °C for 6 h, XRD analysis shows the formation of β-quartz ss, mullite, and Al2TiO5. Signs of the residual glass phase become less prominent. The particle size distribution, similar to previous samples, exhibits a broad profile, likely due to overlapping distributions from different phases, with an average particle size of 30 nm. The size of the predominant crystalline phase of β-quartz ss is 28 nm, see Table 2.
Finally, after two-step heat treatment at 680 °C and 1200 °C for 6 h, the SEM image of the opaque LAS sample shows the presence of micron-sized, needle-like crystals and agglomerates of spherical crystals, see Figure 11h. The needle-like crystals are very similar to crystals of β-spodumene ss that were crystallized in the lithium aluminosilicate glass of a different composition during its heat treatment at 1350 °C [12]. XRD analysis confirms the crystallization of β-spodumene (ss), mullite, and Al2TiO5 phases, see Figure 7e.
The influence of redox conditions of glass melting on the phase assamblage of glass-ceramics is revealed by the comparison of SEM images of LAS and LASox glass-ceramics obtained by two-stage heat treatment at 680 °C and 780 °C for 6 h. In accordance with XRD data, see Figure 8a, the SEM image of the LASox glass-ceramic shows a significantly reduced crystallinity fraction, see Figure 11e. The particle size distribution for this sample also indicates overlapping distributions, with a mean particle size of 16 nm, which is smaller than the mean particle size of the LAS glass-ceramic. Considering that the main contribution to the increase in mean crystallite size comes from larger crystals of β-quartz ss, the smaller mean crystallite size in the LASox glass-ceramic indicates a smaller fraction of these crystals.

3.5. Optical Spectroscopy

Light losses in phase separated glasses and glass-ceramics are determined by absorption due to coloring ions and light scattering on the interfaces of regions of inhomogeneity inherent to such materials. It is generally accepted that the coloration of glass-ceramics containing titania as a nucleating agent is mainly caused by intervalent charge transfer transitions between titanium and iron impurity ions in different oxidation states [8,53]. Though the influence of absorption due to the Fe3+-Ti3+ and Fe2+-Ti4+ intervalence charge transitions on absorption spectra of our materials cannot be excluded, only absorption due to titanium ions was considered in this study. A similar approach was suggested in [53]. Titanium ions in glasses exist in two oxidation states. The presence of Ti3+ and Ti4+ ions, and Ti3+-Ti4+ pairs defines the absorption of glasses under study. Ti4+ ion has an electronic configuration of 3d0 and does not demonstrate absorption bands due to d–d transitions in the ligand field. Ti4+ ions participate in O-Ti4+ charge transfer bands positioning in the UV range of spectrum at ~300 nm [54] and in intervalence charge transfer Ti4+-Ti3+ transitions responsible for coloration in the visible spectral range [55]. Ti3+ ions have an electronic configuration of 3d1 and demonstrate one broad band in the visible range of spectrum due to the 2T2g → Eg transition of the Ti3+ ions in octahedral symmetry. The band often has a shoulder at the long-wavelength side of the absorption band caused by the Jahn–Teller effect [23,25,26,27,55]. The absorption band of Ti3+ ion in tetrahedral (Td) symmetry caused by the 2T2g → Eg transition is located in the near-infrared region of the spectra [55]. The O-Ti3+ charge transfer band is expected in UV range of spectrum at ~240 nm [54].
Absorption spectra of initial and heat-treated glasses are shown in Figure 13a–h. For the convenience of comparison, the spectra of the LAS and LASox glasses and glass-ceramics obtained by the same heat-treatment schedule are presented in Figure 14a–h. The spectra are formed by the absorption edge in the UV spectral range, see Figure 13e,f, intense absorption of OH groups spanning from 2700 nm to 3300 nm, see Figure 13g,h, and light losses in the visible and near-IR spectral range, see Figure 13a–c, which have different origin in the LAS and the LASox glass-ceramics. Note that the LASox glass-ceramics obtained by heat treatments at 950 °C and 1000 °C at the second stage cracked during heat treatment, and their spectra were not recorded.
For the initial LAS glass, the UV absorption edge is observed at ~330 nm. In the spectrum of the glass heat-treated at the nucleation stage at 680 °C for 6 h it moves to longer wavelengths by 5 nm. In the spectrum of glass-ceramics obtained by the two-stage heat treatment with a temperature of 720 °C at the second stage, the position of the absorption edge is found at ~343 nm and it is nearly the same after heat treatments at 750 °C and 780 °C. After heat treatment with a temperature of 800 °C at the second stage, the absorption edge shifts to shorter wavelengths, to 334 nm. After increasing the temperature of the second stage to 850 °C and 900 °C, the absorption edge shifts again to longer wavelengths. Its position is equal to 347 nm and 351 nm, respectively. After heat treatments at 950 °C and 1000 °C, the position of the absorption edge shifts again to shorter wavelengths, see Figure 13e.
The position of the absorption edge for the LASox glass and its variation with the heat-treatment temperature are surprisingly similar to those of the LAS glass, see Figure 13e,f and Figure 14a–e,g,h. The position of the absorption edge differs only for glass-ceramics obtained by the heat treatment at the second stage at 800 °C, see Figure 14f. The absorption edge in the spectrum of this LASox glass-ceramic is shifted to longer wavelengths compared to its position in the spectrum of the LAS sample. The nonmonotonic transmittance variation during the crystallization of glasses inclined to liquid–liquid phase separation was described in ref. [56] and assigned to incoherent scattering that takes place in a material containing amorphous and crystallized regions of inhomogeneity. The authors of ref. [56] demonstrated that the extinction coefficient can reach a maximum when the crystallinity fraction is 0.5–1.0 and then decrease due to the presence of elements of ordering in the relative position of the crystallites.
The same raw materials were used for the preparation of both glasses, which means that the iron content in both glasses was similar. The absorption edge in precursor glass and in glass-ceramics originates from O2−→Ti4+ and O2−→Ti3+ charge transfer bands [54]. Taking into account that the O2−→Ti4+ charge transfer band is located at longer wavelengths than the O2−→Ti3+ one, it can be suggested that the content of Ti4+ ions in the LAS and the LASox glasses is very similar, i.e., the content of Ti3+ in the LAS glass is rather low. In glass-ceramics containing β-quarts ss, scattering losses are superimposed with the absorption edge.
A broad absorption band is found in the spectrum of the LAS precursor glass. Its intensity increases after heat treatment at the nucleation stage and with the increasing of the heat-treatment temperature at the second stage up to 780 °C, see Figure 13a,c. After further heat treatments the absorption band becomes broader, and its intensity somewhat decreases. The appearance of this absorption band can be explained by Ti3+ ions distributed between different amorphous and crystalline phases. Since the absorption edge and light scattering in the obtained multiphase materials are superimposed on the short-wave part of the absorption spectrum, absorption due to these losses was subtracted from the experimental absorption spectra of the LAS samples using the corresponding absorption spectra of LASox samples for substraction. The difference spectra are shown in Figure 13d. The spectrum of the precursor glass is a typical absorption spectrum of Ti3+ ions in silicate glasses with a broad peak with a maximum at ~480 nm attributed to d–d transition 2T2g2Eg of 3d1 electron of Ti3+ ions in octahedral coordination in the ligand field affected by the Jahn–Teller effect [25,26,27]. The tail with a maximum at ~800 nm is assigned to intervalence charge transfer in Ti3+-Ti4+ pairs [57,58]. The absorption intensity increases after the preliminary heat treatment, the band maximum shifts to ~500 nm, which is connected with the participation of Ti3+ ions in the liquid–liquid phase separation and entering into the aluminotitanate amorphous regions. With the increase in the heat-treatment temperature at the second stage, the broadband absorption in the visible and near-IR region intensifies, and its maximum shifts to ~730 nm in the spectra of glass-ceramics obtained by heat treatments at 750 °C and 780 °C at the second stage. Similar to the absorption of Ti3+ ions in corrund crystals [59], this absorption is attributed to Ti3+ ions in the ligand field of octahedral symmetry (the shoulder at ~480 nm) and Ti3+-Ti4+ pairs in crystals of γ-Al2O3. In the spectrum of corrundum, the Ti4+ ions in Al3+ sites are considered to be charge-compensated by Al vacancies with one vacancy for every three Ti4+ ions [60]. An increase of the lattice parameters of γ-Al2O3 with heat-treatment temperature found by the XRD analysis can be due to the entering of the crystals of γ-Al2O3 by Ti3+ ions (ionic radius in octahedral coordination is 0.81 Å) and Ti4+ ions (ionic radius in octahedral coordination is 0.745 Å). Note that the ionic radii of Al3+ in octahedral coordination is 0.675 Å.
The absorption band shape changes with the raising of the heat-treatment temperature. The spectrum is still mainly formed by the absorption of Ti3+-Ti4+ pairs. Taking into account the crystallization of β-quartz ss and tieilite, Al2TiO5, this change in the shape of the absorption band can be connected with the distribution of titanium ions between the crystalline phases. Spinel does not crystallize in glass-ceramics obtained by heat treatments at 850–1000 °C at the second stage. So the broadband spectra of glass-ceramics obtained by these heat treatments are connected with titanium ions in β-quartz ss and tieilite.
The light losses in the LASox glass-ceramics are mostly caused by light scattering, which has a non-monotoneous dependence on the heat-treatment temperature, as was mentioned above.
Figure 13g,h show wide asymmetric absorption bands in the spectral region from ~2700 nm to 3300 nm caused by the presence of OH-groups in glasses and glass-ceramics. Heat treatment at the nucleation stage at 680 °C for 6 h has no effect on the shape and intensity of this band. Two-stage heat treatments with a temperature from 720 °C to 780 °C at the second stage cause a successive change in the spectrum of OH-groups. The intensity in the range from 2750 nm to 3050 nm increases and the intensity in the range from 3050 nm to 3300 nm decreases. Further narrowing of the OH-groups absorption band and growth of its intensity is observed in spectra of glass-ceramics obtained during heat treatments in the range of crystallization temperatures from 800 °C to 1000 °C. Comparison of the absorption bands of OH-groups in glasses melted in different redox conditions and in corresponding glass-ceramics showed that the intensities of the absorption bands of OH-groups are higher in inital and heat-treated LAS glass than in LASox samples. The absorption band shape is the same for the materials obtained by the same heat-treatment shedule, see Figure 14a–h, which is explained by the similarity of their phase compositions.

3.6. Estimation of the Coefficient of Thermal Expansion

Table 4 lists the coefficients of thermal expansion (CTE) for the initial and heat-treated glasses of the LAS and the LASox compositions in the temperature range between 20 °C and 320 °C. The CTE values of precursor glasses are similar and equal to ~4.3 × 10−6 K−1. After the nucleation heat treatment at 680 °C for 6 h the CTE values slightly decrease to become ~4.1 × 10−6 K−1. There is a gradual increase of CTE values with the increase in the ceramming temperature to 720 °C and then to 750 °C. Glass-ceramics based on spinel nanocrystals and prepared by the consequent three-stage heat treatment schedule 680 °C, 6 h + 720 °C, 6 h + 750 °C, 6 h have the highest CTE value of ~5.2 × 10−6 K−1. The CTE value starts to decrease as the first portions of β-quartz ss crystallize (heat treatment with the last hold at 780 °C for 6 h) and reaches the lowest value of ~0.3 × 10−6 K−1 for the glass-ceramics obtained by ceramming at the last stage of heat treatment at 950 °C for 6 h. After heat treatment of the precursor glass at 1200 °C, the CTE increases again to a value of 1.46 × 10−6 K−1 due to the crystallization of β-spodumene ss. The dependences of the change in the CTE value on the temperature of heat treatment for samples obtained by the crystallization of glasses melted under neutral and oxidizing conditions are similar, which can be explained by the similarity of their phase assemblage.

4. Discussion

Glasses melted in neutral and oxidizing conditions are X-ray amorphous and inhomogeneous according to SEM microscopy findings. Inhomogeneity regions have a broad size distribution with a mean size of 27 nm. In accordance with previous studies [19,20], it is suggested that liquid–liquid phase separation occurs during glass melt casting and annealing. Absorption spectra of glasses melted in different redox conditions are different by the appearance of a broadband absorption in the LAS glass mainly assigned to Ti3+ ions in distorted octahedral coordination and Ti3+-Ti4+ pairs in the glass structure. One may suggest that these ions are distributed between the lithium aluminosilicate glass matrix and amorphous regions of inhomogeneity composed of octahedrally coordinated titanium and aluminum ions even with the formation of Ti3+-Ti4+ clusters [18]. Based on the similarity of XRD patterns, Raman spectra and glass transition temperatures, it was speculated that glasses melted in neutral and oxidizing conditions have a similar structure because of the low content of Ti3+ ions. Nevertheless, the role of Ti3+ ions in liquid–liquid phase separation and crystallization of LAS glass is clearly seen.
The glasses obtained by heat treatment at 680 °C for 6 h do not contain any crystalline phase, their XRD patterns are similar, and their Raman spectra are near similar to each other and to those of precursor glasses. However, there is a tremendous difference in the behavior of their DSC curves, their morphology revealed in the SEM study and their absorption spectra as compared with those of precursor glasses. It allows us to conclude that liquid–liquid phase separation continues to develop in both glasses during the nucleation heat treatment.
According to data of SEM analysis, two types of amorphous regions with mean sizes of 16 and 74 nm are formed in the glass during heat treatment at 680 °C for 6 h. A similar structure formation was not found in previously studied glasses of the lithium aluminosilicate system. The formation of large amorphous aluminotitanate regions of inhomogeneity in magnesium ([20] and refs therein) and zinc ([20] and refs therein) aluminosilicate glasses nucleated by titania was previously revealed by a combination of small-angle X-ray scattering and Raman spectroscopy data. Those materials also contained smaller size phase-separated regions enriched in aluminates of corresponding cations ([20] and refs therein). One can suggest a similar character of phase separation in glasses under study and suppose that Ti3+ ions are distributed between large aluminotitanate amorphous regions, smaller size aluminate regions and residual glass.
Judging by the data of Raman spectroscopy, glasses cerammed using a two-stage heat treatment with a temperature at the second hold of 720 °C have a well-developed liquid–liquid phase separated structure. During this phase separation, spinel nanocrystals appear in both glasses, and the composition of the residual lithium aluminosilicate glass slightly changes, which is manifested by a change in the position of the amorphous halo in the XRD pattern and the position of the band at ~480 cm−1 in the Raman spectrum. Spinel is the sole crystalline phase in transparent glass-ceramics obtained by heat treatments at the second stage at 720 °C and 750 °C. Upon crystallization of spinel, titanium ions Ti3+ enter its structure, which is clearly seen in the absorption spectra of the LAS glass-ceramics. The change of the position of the maximum of absorption band to longer wavelengths suggests that absorption spectrum in glass-ceramics is mainly formed by the Ti3+-Ti4+ intervalent charge transfer band. This is a spectroscopic confirmation of the presence of Ti4+ ions in the spinel structure.
Therefore, based on the DSC, SEM and XRD data, it was concluded that preliminary heat treatment at 680 °C, resulting in the formation of a reach phase-separated structure, provokes spinel crystallization during further heat treatments between 720 °C and 800 °C. According to the similarity of absorption spectra of these glass-ceramics and spectra of Ti3+-doped Al2O3, it was speculated that Ti3+ ions enter into spinel nanocrystals. Taking into consideration that, according to the findings of the Raman spectroscopy, the rate of phase transformations in the titania-containing phase in glass melted in neutral conditions is significantly higher than in glass melted in oxidizing conditions, it was suggested that Ti3+ ions become a component of amorphous aluminotitanate phase-separated regions.
Previously, the decrease in intensity of the first broad peak in the DSC curves of glasses nucleated at 700 °C and its near disappearance from DCS curves of glasses preliminarily heat-treated at 720 °C for 6 h was mentioned. The sample preliminarily heat-treated at 720 °C already contained spinel nanocrystals. Therefore, in this sample only high temperature phases of β-quartz ss, tieilite and β-spodumene ss crystallized, which resulted in the appearance of the second and the third peaks and the absence of the first peak on the DSC curve.
A specific feature of the morphology of the materials under study is the similarity of the mean size of inhomogeneous regions in the precursor glass and in glass-ceramics obtained by two-stage heat treatments at temperatures up to 1000 °C. Sizes of β-quartz ss crystallized in both glasses in the temperature range from 780 °C to 1000 °C are similar to each other. Our findings are in line with the results presented in ref. [38] and devoted to the study of the lithium aluminosilicate glass nucleated by a mixture of TiO2 and ZrO2, where the constancy of the size of β-quartz ss crystallized in the temperature range from 860 °C to 960 °C was revealed and discussed.
The comparison of XRD patterns of the LAS and the LASox glass-ceramics obtained at 1200 °C demonstrates that LAS glass-ceramic contains a larger fraction of rutile and a smaller fraction of mullite than the LASox glass-ceramic. Previously it was revealed the difference in lattice parameters of rutile crystallized in zinc aluminosilicate glasses nucleated by TiO2 and melted in different redox conditions [34]. In ref. [34] this difference was explained by the presence of Ti3+ ions in the rutile structure. Here, it is suggested that in lithium aluminosilicate glass-ceramics Ti3+ ions also participate in the formation of rutile nanocrystals, which results in the facilitation of their crystallization. Note that the role of small additives of transition metal ions in kinetics of crystallization in glass-ceramics was discussed in ref. [18].
The rate of phase transformations resulted in the crystallization of β-quartz ss is significantly higher during the heat treatment of the glass melted in neutral conditions. However, the sequence of phase transformations and phase assemblage of glass-ceramics melted in different redox conditions are independent of the redox conditions of glass melting. This is very promising for the development of transparent lithium aluminosilicate glass-ceramics with close-to-zero thermal expansion coefficient and doped with various functional ions in the lower oxidation states. Crystallization of β-quartz ss leads to a decreasing of the value of the thermal expansion coefficient. The lowest coefficient of thermal expansion was obtained by the heat treatment at 950 °C. It was ~0.3 × 10−6 K−1.

5. Conclusions

The glass of the lithium aluminosilicate system containing titanium oxide as a nucleating agent was melted with and without the addition of As2O3, i.e., under oxidizing and neutral conditions. The glasses were heat-treated in the temperature range from 680 °C to 1300 °C to obtain glass-ceramics. Glass-ceramics developed from glasses melted without the addition of As2O3 are black-colored, and those fabricated from glasses melted with addition of As2O3 are colorless.
Transparent glass-ceramics based on β-quartz ss and/or γ-Al2O3 nanocrystals and opaque glass-ceramics based on β-spodumene ss nanocrystals were obtained. Crystallization of β-quartz ss was accompanied by nanocrystals of tieilite, Al2TiO5. Crystallization of β-spodumene ss was accompanied by tieilite, mullite (at 1100 °C) and rutile (at 1200 °C).
Heat treatment at the nucleation stage causes crystallization of γ-Al2O3 with a spinel structure and an increase in the crystallization temperature of β-quartz ss and Al2TiO5 and decrease in the temperature of β-spodumene ss crystallization. Glasses melted under neutral conditions lose transparency at higher secondary heat-treatment temperatures than glasses melted under oxidizing conditions.
X-ray amorphous precursor glasses have an inhomogeneous structure with a mean size of regions of inhomogeneity of ca. 27 nm. Preliminary heat treatment at 680 °C results in the formation of a reach bimodal liquid–liquid phase separated structure with mean sizes of inhomogeneous regions of ca. 16 nm and 74 nm. Titanium ions enter both types of inhomogeneous regions. An appearance of a bimodal structure of inhomogeneous regions in the result of heat treatment at the nucleation stage suggests the development of a three-phase immiscibility in the lithium aluminosilicate glass with a high excess of alumina over lithium oxide, i.e., the formation of two amorphous phases, the aluminotitanate phase and the aluminate phase, dispersed in a high-silica matrix. It is the first time that three-phase immiscibility is revealed in glasses of the lithium aluminosilicate system.
It is the first time that the formation of nanocrystals of γ-Al2O3 with a spinel structure appeared during heat treatments at the second stage in the temperature interval between 720 °C and 800 °C was documented. Titanium ions are located in the spinel structure as Ti3+ ions in the ligand field of octahedral symmetry and in Ti3+-Ti4+ pairs. Ti3+ ions also enter amorphous aluminotitanate phase-separated regions and facilitate phase transformations in this phase with formation of tieilite, anatase and rutile.
The phase assemblage and sequence of phase transformations is independent of the redox conditions of glass melting; however, the rate of these transformations is significantly higher when ceramming the glass melted in neutral conditions. This is very promising for the development of transparent glass-ceramics with close-to-zero thermal expansion coefficient and doped with various functional ions in lower oxidation states.
The variation of the coefficient of thermal expansion with heat-treatment temperature reflects the variation of the phase composition of the developed materials, does not depend on the redox conditions of glass melting and reaches as low value as ~0.3 × 10−6 K−1 for glass-ceramics based on nanocrystals of β-quartz ss.

Author Contributions

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

Funding

This research received no funding.

Data Availability Statement

The original contributions presented in this study are included in the article. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Höland, W.; Beall, G.H. Glass-Ceramic Technology; Wiley: Hoboken, NJ, USA, 2012; pp. 92–97, 269–279. [Google Scholar]
  2. Bach, H.; Krause, D. Low Thermal Expansion Glass Ceramics; Springer-Verlag: Berlin/Heidelberg, Germany, 2005. [Google Scholar]
  3. Petzoldt, J.; Pannhorst, W. Chemistry and structure of glass-ceramic materials for high precision optical applications. J. Non-Cryst. Solids 1991, 129, 191–198. [Google Scholar] [CrossRef]
  4. Stookey, S.D. Method of Making Ceramics and Product Thereof. U.S. Patent 2,920,971, 12 January 1960. [Google Scholar]
  5. Beall, G.H.; Karstetter, B.R.; Rittler, H.L. Crystallization and chemical strengthening of stuffed β-quartz glass-ceramics. J. Am. Ceram. Soc. 1967, 5, 181–190. [Google Scholar] [CrossRef]
  6. Zanotto, E.D. A bright future for glass-ceramics. Am. Ceram. Soc. Bull. 2010, 89, 19–27. [Google Scholar]
  7. Fernandez-Martin, C.; Bruno, G.; Crochet, A.; Ovono, D.O.; Comte, M.; Hennet, L. Nucleation and growth of nanocrystals in glass-ceramics: An in situ SANS perspective. J. Am. Ceram. Soc. 2012, 95, 1304–1312. [Google Scholar] [CrossRef]
  8. Nakane, S.; Kawamoto, K. Coloration mechanism of Fe ions in β-quartz s.s. glass-ceramics with TiO2 and ZrO2 as nucleation agents. Front. Mater. 2017, 4, 14–19. [Google Scholar] [CrossRef]
  9. Fu, Q.; Wheaton, B.R.; Geisinger, K.L.; Credle, A.J.; Wang, J. Crystallization, microstructure, and viscosity evolutions in lithium aluminosilicate glass-ceramics. Front. Mater. 2016, 3, 49. [Google Scholar] [CrossRef]
  10. Kleebusch, E.; Patzig, C.; Krause, M.; Hu, Y.; Höche, T.; Rüssel, C. The formation of nanocrystalline ZrO2 nuclei in a Li2O-Al2O3-SiO2 glass—A combined XANES and TEM study. Sci. Rep. 2017, 7, 10869. [Google Scholar] [CrossRef]
  11. Zandona, A.; Patzig, C.; Rüdinger, B.; Hochrein, O.; Deubener, J. TiO2(B) nanocrystals in Ti-doped lithium aluminosilicate glasses. J. Non-Cryst. Solids 2019, 2, 100025. [Google Scholar] [CrossRef]
  12. Dymshits, O.; Bachina, A.; Alekseeva, I.; Golubkov, V.; Tsenter, M.; Zapalova, S.; Bogdanov, K.; Danilovich, D.; Zhilin, A. Phase transformations upon formation of transparent lithium alumosilicate glass-ceramics nucleated by yttrium niobates. Ceramics 2023, 6, 1490–1507. [Google Scholar] [CrossRef]
  13. Vigier, M.; Deniard, P.; Gautron, E.; Gautier, N.; Genevois, C.; Ory, S.; Allix, M.; Kacem, I.B.; Jobic, S. Microstructural insights on lithium aluminum silicate (LAS) glass ceramics. Ceram. Int. 2024, 50, 29011–29015. [Google Scholar] [CrossRef]
  14. Guo, Y.; Wang, J.; Ruan, J.; Han, J.; Xie, J.; Liu, C. Microstructure and ion-exchange properties of glass-ceramics containing ZnAl2O4 and β-quartz solid solution nanocrystals. J. Eur. Ceram. Soc. 2021, 41, 5331–5340. [Google Scholar] [CrossRef]
  15. Beall, G.H.; Pinckney, L.R. Nanophase glass-ceramics. J. Am. Ceram. Soc. 1999, 82, 5–16. [Google Scholar] [CrossRef]
  16. Doherty, P.D. Direct observation of the crystallization of Li2O-Al2O3-SiO2 glasses containing TiO2. J. Am. Ceram. Soc. 1967, 50, 77–80. [Google Scholar] [CrossRef]
  17. Alekseeva, I.P.; Belyaevskaya, N.M.; Bobovich, Y.S.; Tsenter, M.Y.; Chuvaeva, T.I. Recording, interpretation, and some examples of application of Raman spectra for glass ceramics activate with titanium (IV) oxide. Opt. Spectrosc. 1978, 45, 927–936. [Google Scholar]
  18. Khodakovskaya, R.Y. Chemistry of Titanium-Containing Glasses and Glass Ceramics; Chemistry: Moscow, Russia, 1978; pp. 190–198. (In Russian) [Google Scholar]
  19. Dymshits, O.S.; Zhilin, A.A.; Petrov, V.I.; Tsenter, M.Y.; Chuvaeva, T.I.; Golubkov, V.V. A Raman spectroscopic study of phase transformations in titanium-containing lithium aluminosilicate glasses. Glass Phys. Chem. 1998, 24, 79–96. [Google Scholar]
  20. Alekseeva, I.; Dymshits, O.; Tsenter, M.; Zhilin, A.; Golubkov, V.; Denisov, I.; Skoptsov, N.; Malyarevich, A.; Yumashev, K. Optical applications of glass-ceramics. J. Non-Cryst. Solids 2010, 356, 3042–3058. [Google Scholar] [CrossRef]
  21. Kleebusch, E.; Patzig, C.; Krause, M.; Hu, Y.; Höche, T.; Rüssel, C. The effect of TiO2 on nucleation and crystallization of a Li2O-Al2O3-SiO2 glass investigated by XANES and STEM. Sci. Rep. 2018, 8, 2929. [Google Scholar] [CrossRef]
  22. Kleebusch, E.; Patzig, C.; Höche, T.; Rüssel, C. The evidence of phase separation droplets in the crystallization process of a Li2O-Al2O3-SiO2 glass with TiO2 as nucleating agent—An X-ray diffraction and (S)TEM-study supported by EDX-analysis. Ceram. Int. 2018, 44, 2919–2926. [Google Scholar] [CrossRef]
  23. Johnston, W.D. Oxidation-reduction equilibria in molten Na2O·2SiO2 glass. J. Am. Ceram. Soc. 1965, 48, 184–190. [Google Scholar] [CrossRef]
  24. Rawal, B.S.; MacCrone, R.K. Optical absorption in an equimolar barium borosilicate glass containing titanium lons. J. Non-Cryst. Solids 1978, 28, 337–345. [Google Scholar] [CrossRef]
  25. Kurkjian, C.R.; Peterson, G.E. An EPR study of Ti3++-Ti4+ in TiO2-SiO2 glasses. Phys. Chem. Glas. 1974, 15, 12–17. [Google Scholar]
  26. Morinaga, K.; Yoshida, H.; Takebe, H. Compositional dependence of absorption spectra of Ti3+ in silicate, borate, and phosphate glasses. J. Am. Ceram. Soc. 1994, 77, 3113–3118. [Google Scholar] [CrossRef]
  27. El-Shafi, N.A.; Morsi, M.M. Optical absorption and infrared studies of some silicate glasses containing titanium. J. Mater. Sci. 1997, 32, 5185–5189. [Google Scholar] [CrossRef]
  28. Andrade, L.H.C.; Lima, S.M.; Novatski, A.; Neto, A.M.; Bento, A.C.; Baesso, M.L.; Gandra, F.C.G.; Guyot, Y.; Boulon, G. Spectroscopic assignments of Ti3+ and Ti4+ in titanium-doped OH_ free low-silica calcium aluminosilicate glass and role of structural defects on the observed long lifetime and high fluorescence of Ti3+ ions. Phys. Rev. B 2008, 78, 224202. [Google Scholar] [CrossRef]
  29. Vogel, W. Glass Chemistry; Springer: Berlin/Heidelberg, Germany; New York, NY, USA, 1994; pp. 313–314. [Google Scholar] [CrossRef]
  30. Höland, W.; Wange, P.; Naumann, K.; Vogel, J.; Carl, G.; Jana, C.; Götz, W. Control of phase formation processes in glass-ceramics for medicine and technology. J. Non-Cryst. Solids 1991, 129, 152–162. [Google Scholar] [CrossRef]
  31. Wisniewski, W.; Thieme, K.; Rüssel, C. Fresnoite glass-ceramics—A review. Prog. Mater. Sci. 2018, 98, 68–107. [Google Scholar] [CrossRef]
  32. Keding, R.; Rüssel, C. Electrochemical nucleation for the preparation of oriented glass ceramics. J. Non-Cryst. Solids 1997, 219, 136–141. [Google Scholar] [CrossRef]
  33. Höche, T.; Kleebe, H.-J.; Brydson, R. Can fresnoite (Ba2TiSi2O8) incorporate Ti3+ when crystallizing from highly reduced melts? Philos. Mag. A 2001, 81, 825–839. [Google Scholar] [CrossRef]
  34. Eremeev, K.; Dymshits, O.; Alekseeva, I.; Khubetsov, A.; Zapalova, S.; Tsenter, M.; Basyrova, L.; Serres, J.M.; Mateos, X.; Loiko, P.; et al. Effect of redox conditions of glass melting on the structure and the properties of titanium-containing gahnite glass-ceramics. J. Eur. Ceram. Soc. 2024, 44, 3362–3380. [Google Scholar] [CrossRef]
  35. Lipatiev, A.; Fedotov, S.; Lotarev, S.; Naumov, A.; Lipateva, T.; Savinkov, V.; Shakhgildyan, G.; Sigaev, V. Direct laser writing of depressed-cladding waveguides in extremely low expansion lithium aluminosilicate glass-ceramics. Opt. Laser Technol. 2021, 138, 106846. [Google Scholar] [CrossRef]
  36. Lipson, H.; Steeple, H. Interpretation of X-Ray Powder Patterns; McMillan: London, UK; New York, NY, USA, 1970; p. 344. [Google Scholar]
  37. Schneider, C.A.; Rasband, W.S.; Eliceiri, K.W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 2012, 9, 671–675. [Google Scholar] [CrossRef] [PubMed]
  38. Ovono Ovono, D.; Berre, S.; Pradeau, P.; Comte, M.; Bruno, G. Study of the crystallization kinetics of LAS glass by differential scanning calorimetry, X-ray diffraction, and beam bending viscometry. Thermochim. Acta 2012, 527, 158–164. [Google Scholar] [CrossRef]
  39. Ovono Ovono, D.; Bruno, G.; Pradeau, P.; Berre, S. Conditions for crystallization of LAS glass-ceramics as a function of nucleating agent amount and heat treatment. Int. J. Appl. Glass Sci. 2013, 4, 20–30. [Google Scholar] [CrossRef]
  40. Shirasuka, K.; Yanagida, H.; Yamaguchi, G. The preparation of η-alumina and its structure. Mater. Sci. 1976, 84, 610–613. [Google Scholar] [CrossRef]
  41. Zhou, R.-S.; Snyder, R.L. Structures and transformation mechanisms of the η, γ and θ transition aluminas. Acta Cryst. 1991, B47, 617–630. [Google Scholar] [CrossRef]
  42. Skala, R.D.; Li, D.; Low, I.M. Diffraction, structure and phase stability studies on aluminium titanate. J. Eur. Ceram. Soc. 2009, 29, 67–75. [Google Scholar] [CrossRef]
  43. Galeeener, F.L.; Leadbetter, A.J.; Stringfellow, M.W. Comparison of the neutron, Raman and infrared vibrational spectra of vitreous SiO2, GeO2 and BeF2. Phys. Rev. B 1983, 27, 1052–1078. [Google Scholar] [CrossRef]
  44. Bobovich, Y.S. Spectroscopic invistigation of titanium coordination in some vitreous materials. Opt. Spectrosc. 1963, 14, 647–654. [Google Scholar]
  45. Thomas, P.V.; Ramakrishnan, V.; Vaidyan, V.K. Oxidation studies of aluminum thin films by Raman spectroscopy. Thin Solid Film. 1989, 170, 35–40. [Google Scholar] [CrossRef]
  46. Baronskiy, M.; Rastorguev, A.; Zhuzhgov, A.; Kostyukov, A.; Krivoruchko, O.; Snytnikov, V. Photoluminescence and Raman spectroscopy studies of low-temperature γ-Al2O3 phases synthesized from different precursors. Opt. Mater. 2016, 53, 87–93. [Google Scholar] [CrossRef]
  47. Liu, Y.; Cheng, B.; Wang, K.; Ling, G.; Cai, J.; Song, C.; Han, G. Study of Raman spectra for γ-Al2O3 models by using first-principles method. Solid State Commun. 2014, 178, 16–22. [Google Scholar] [CrossRef]
  48. Basyrova, L.; Bukina, V.; Balabanov, S.; Belyaev, A.; Drobotenko, V.; Dymshits, O.; Alekseeva, I.; Tsenter, M.; Zapalova, S.; Khubetsov, A.; et al. Synthesis, structure and spectroscopy of Fe2+: MgAl2O4 transparent ceramics and glass-ceramics. J. Lumin. 2021, 236, 118090. [Google Scholar] [CrossRef]
  49. Sharma, K.; Simons, B. Raman study of crystalline polymorphs and glasses of spodumene composition quenched from various pressures. Am. Mineral. 1981, 66, 118–126. [Google Scholar]
  50. Alekseeva, I.; Dymshits, O.; Ermakov, V.; Zhilin, A.; Petrov, V.; Tsenter, M. Raman spectroscopy quantifying the composition of stuffed β-quartz derivative phases in lithium aluminosilicate glass-ceramics. J. Non-Cryst. Solids 2008, 354, 4932–4939. [Google Scholar] [CrossRef]
  51. Sprengard, R.; Binder, K.; Brandle, M.; Fotheringham, U.; Sauer, J.; Pannhorst, W. On the interpretation of the experimental Raman spectrum of β-eucryptite LiAlSiO4 from atomistic computer modeling. J. Non-Cryst. Solids 2000, 274, 264–270. [Google Scholar] [CrossRef]
  52. Bost, N.; Duraipandian, S.; Guimbretière, G.; Poirier, J. Raman spectra of synthetic and natural mullite. Vib. Spectrosc. 2016, 82, 50–52. [Google Scholar] [CrossRef]
  53. Chavoutier, M.; Caurant, D.; Majérus, O.; Boulesteix, R.; Loiseau, P.; Jousseaume, C.; Brunet, E.; Lecomte, E. Effect of TiO2 content on the crystallization and the color of (ZrO2, TiO2)-doped Li2O-Al2O3-SiO2 glasses. J. Non-Cryst. Solids 2014, 384, 15–24. [Google Scholar] [CrossRef]
  54. Loeffler, B.M.; Burns, R.G.; Tossell, J.A.; Vaughan, D.J.; Johnson, K.H. Charge transfer in lunar materials: Interpretation of ultraviolet-visible spectral properties of the moon. In Proceedings of the Lunar Science Conference, 5th, Houston, TX, USA, 18–22 March 1974; Volume 5, pp. 3007–3016. [Google Scholar]
  55. Marfunin, A.S. Physics of Minerals and Inorganic Materials: An Introduction; Springer-Verlag: Berlin/Heidelberg, Germany; New York, NY, USA, 1979; p. 340. [Google Scholar]
  56. Petrovski, G.T.; Zhilin, A.A.; Shepilov, M.P.; B’en, V.B.; Li, K.K. Nonmonotonic transmittance variation of a material during the crystallization of liquating glasses. J. Opt. Technol. 2003, 70, 857–863. [Google Scholar] [CrossRef]
  57. Sanchez, A.; Strauss, A.J.; Aggarwal, R.L.; Fahey, R.E. Crystal Growth, Spectroscopy, and laser characteristics of Ti: Al2O3. IEEE J. Quantum Electron. 1988, 24, 995–1002. [Google Scholar] [CrossRef]
  58. Ripenko, V.V.; Khomenko, V.M.; Vyshnevskyi, O.A.; Kosorukov, O.O. Synthesis and physical properties of Li-Al-Si glass-ceramic materials from petalite of Polohivka Deposit (the Ukrainian Shield). Mineral. J. 2017, 39, 53–62. [Google Scholar] [CrossRef]
  59. Moulton, P.F.; Cederberg, J.G.; Stevens, K.T.; Foundos, G.; Koselja, M.; Preclikova, J. Characterization of absorption bands in Ti:sapphire crystals. Opt. Mater. Express. 2019, 9, 2216–2251. [Google Scholar] [CrossRef]
  60. Mohapatra, S.K.; Kröger, F.A. Defect structure of α-Al2O3 doped with titanium. J. Am. Ceram. Soc. 1977, 60, 381–387. [Google Scholar] [CrossRef]
Figure 1. Photographs of as-casted glasses with thickness of ca. 6 mm: (a) LAS; (b) LASox. Glass notations are shown in figure.
Figure 1. Photographs of as-casted glasses with thickness of ca. 6 mm: (a) LAS; (b) LASox. Glass notations are shown in figure.
Materials 18 00785 g001
Figure 2. Photographs of samples of the precursor glasses, glasses heat-treated at 680 °C and glass-ceramics prepared by heat treatments with the first stage at 680 °C and the second stage between 720 °C and 1200 °C: (a) LAS; (b) LASox. The regimes of heat treatments are shown in figure. Heat-treatment duration at each stage is 6 h. The thickness of transparent polished plates is 1 mm.
Figure 2. Photographs of samples of the precursor glasses, glasses heat-treated at 680 °C and glass-ceramics prepared by heat treatments with the first stage at 680 °C and the second stage between 720 °C and 1200 °C: (a) LAS; (b) LASox. The regimes of heat treatments are shown in figure. Heat-treatment duration at each stage is 6 h. The thickness of transparent polished plates is 1 mm.
Materials 18 00785 g002
Figure 3. DSC signals as time functions for quenched and heat-treated LAS and LASox glasses. Tg denotes the glass transition temperature, Ton is the onset crystallization temperature, Tmax is the maximum crystallization temperature. The curves are shifted for the convenience of observation.
Figure 3. DSC signals as time functions for quenched and heat-treated LAS and LASox glasses. Tg denotes the glass transition temperature, Ton is the onset crystallization temperature, Tmax is the maximum crystallization temperature. The curves are shifted for the convenience of observation.
Materials 18 00785 g003
Figure 4. (a) DSC signals as time functions for the quenched LAS glass with stops at temperatures of 880 °C and 1175 °C; (b) XRD patterns of thus obtained LAS glass-ceramics. The curves are shifted for the convenience of observation.
Figure 4. (a) DSC signals as time functions for the quenched LAS glass with stops at temperatures of 880 °C and 1175 °C; (b) XRD patterns of thus obtained LAS glass-ceramics. The curves are shifted for the convenience of observation.
Materials 18 00785 g004
Figure 5. (a) DSC signals as time functions for the LAS glass preliminary heated at 680 °C for 6 h with stops at temperatures of 787 °C, 838 °C, 858 °C, and 935 °C; (b) XRD patterns of thus obtained LAS glass-ceramics. The curves are shifted for the convenience of observation.
Figure 5. (a) DSC signals as time functions for the LAS glass preliminary heated at 680 °C for 6 h with stops at temperatures of 787 °C, 838 °C, 858 °C, and 935 °C; (b) XRD patterns of thus obtained LAS glass-ceramics. The curves are shifted for the convenience of observation.
Materials 18 00785 g005
Figure 6. XRD patterns of the LAS quenched glass, and the glass heat-treated at 680 °C for 6 h and then heated in the furnace of the DSC instrument. The heat-treatment schedules are shown in the figure. The patterns are shifted for the convenience of observation.
Figure 6. XRD patterns of the LAS quenched glass, and the glass heat-treated at 680 °C for 6 h and then heated in the furnace of the DSC instrument. The heat-treatment schedules are shown in the figure. The patterns are shifted for the convenience of observation.
Materials 18 00785 g006
Figure 7. (a,b) XRD patterns of LAS and LASox glasses and glass-ceramics: (a) the LAS; (b) the LASox; (c,d) a closer look on XRD patterns of glasses and glass-ceramics obtained by heat treatments at the second stage up to 800 °C: (c) the LAS; (d) the LASox; (e,f) a closer look on XRD patterns of glass-ceramics obtained by heat treatments at the second stage between 850 °C and 1200 °C: (e) the LAS; (f) the LASox. The heat-treatment temperature at the first stage is 680 °C, duration of each stage is 6 h. The patterns are shifted for the convenience of observation.
Figure 7. (a,b) XRD patterns of LAS and LASox glasses and glass-ceramics: (a) the LAS; (b) the LASox; (c,d) a closer look on XRD patterns of glasses and glass-ceramics obtained by heat treatments at the second stage up to 800 °C: (c) the LAS; (d) the LASox; (e,f) a closer look on XRD patterns of glass-ceramics obtained by heat treatments at the second stage between 850 °C and 1200 °C: (e) the LAS; (f) the LASox. The heat-treatment temperature at the first stage is 680 °C, duration of each stage is 6 h. The patterns are shifted for the convenience of observation.
Materials 18 00785 g007
Figure 8. XRD patterns of the LAS and the LASox glass-ceramics obtained by two-stage heat treatments at: (a) 680 °C, 6 h + 780° C, 6 h; (b) 680 °C, 6 h + 800 °C, 6 h. The patterns are shifted for the convenience of observation.
Figure 8. XRD patterns of the LAS and the LASox glass-ceramics obtained by two-stage heat treatments at: (a) 680 °C, 6 h + 780° C, 6 h; (b) 680 °C, 6 h + 800 °C, 6 h. The patterns are shifted for the convenience of observation.
Materials 18 00785 g008
Figure 9. Raman spectra of the LAS glass and glass-ceramics obtained by single and two-stage heat treatments: (a) from the glass to 1200 °C; (b) from the glass up to 780 °C; (c) from 800 °C up to 950 °C; (d) from 1000 °C up to 1200 °C. The time of heat treatment at each hold is 6 h. The spectra are shifted for the convenience of observation.
Figure 9. Raman spectra of the LAS glass and glass-ceramics obtained by single and two-stage heat treatments: (a) from the glass to 1200 °C; (b) from the glass up to 780 °C; (c) from 800 °C up to 950 °C; (d) from 1000 °C up to 1200 °C. The time of heat treatment at each hold is 6 h. The spectra are shifted for the convenience of observation.
Materials 18 00785 g009
Figure 10. Raman spectra the LAS and the LASox glass-ceramics obtained by two-stage heat treatments at: (a) 680 °C + 720 °C; (b) 680 °C + 780 °C; (c) 680 °C + 900 °C; (d) 680 °C + 950 °C; (e) 680 °C + 1000 °C; (f) 680 °C + 1200 °C. The timeof heat treatment at each stage is 6 h. The spectra are shifted for the convenience of observation.
Figure 10. Raman spectra the LAS and the LASox glass-ceramics obtained by two-stage heat treatments at: (a) 680 °C + 720 °C; (b) 680 °C + 780 °C; (c) 680 °C + 900 °C; (d) 680 °C + 950 °C; (e) 680 °C + 1000 °C; (f) 680 °C + 1200 °C. The timeof heat treatment at each stage is 6 h. The spectra are shifted for the convenience of observation.
Materials 18 00785 g010
Figure 11. SEM images of the samples of: (a) the initial LAS glass; (b) the LAS glass after heat treatment at 680 °C; (c) the LAS glass ceramized at 680 °C and at 750 °C; (d) the LAS glass ceramized at 680 °C and at 780 °C; (e) the LASox glass ceramized at 680 °C and at 780 °C; (f) the LAS glass ceramized at 680 °C and at 800 °C; (g) the LAS glass ceramized at 680 °C and at 1000 °C; (h) the LAS glass ceramized at 680 °C and at 1200 °C. The holding time is 6 h.
Figure 11. SEM images of the samples of: (a) the initial LAS glass; (b) the LAS glass after heat treatment at 680 °C; (c) the LAS glass ceramized at 680 °C and at 750 °C; (d) the LAS glass ceramized at 680 °C and at 780 °C; (e) the LASox glass ceramized at 680 °C and at 780 °C; (f) the LAS glass ceramized at 680 °C and at 800 °C; (g) the LAS glass ceramized at 680 °C and at 1000 °C; (h) the LAS glass ceramized at 680 °C and at 1200 °C. The holding time is 6 h.
Materials 18 00785 g011
Figure 12. Particle size distributions in the initial LAS glass and LAS and LASox glass-ceramics calculated from SEM data presented in Figure 11. Heat-treatment schedules and mean crystallite sizes are listed in figure.
Figure 12. Particle size distributions in the initial LAS glass and LAS and LASox glass-ceramics calculated from SEM data presented in Figure 11. Heat-treatment schedules and mean crystallite sizes are listed in figure.
Materials 18 00785 g012
Figure 13. Absorption spectra in different spectral ranges of the samples of: (a) the initial LAS glass and glass heat-treated up to 1000 °C in the spectral range from 250 nm to 3300 nm; (b) the initial LASox glass and glass heat-treated up to 900 °C in the spectral range from 250 nm to 3300 nm; (c) the LAS glass and glass heat-treated up to 780 °C in the spectral range from 300 nm to 1750 nm; (d) the subtracted spectra; (e) the LAS glass and glass heat-treated up to 1000 °C in the spectral range from 300 nm to 460 nm; (f) the LASox glass and glass heat-treated up to 900 °C in the spectral range from 300 nm to 475 nm; (g) the LAS glass and glass heat-treated up to 1000 °C in the spectral range from 2600 nm to 3300 nm; (h) the LASox glass and glass heat-treated up to 900 °C in the spectral range from 2600 nm to 3300 nm. The duration of each hold is 6 h.
Figure 13. Absorption spectra in different spectral ranges of the samples of: (a) the initial LAS glass and glass heat-treated up to 1000 °C in the spectral range from 250 nm to 3300 nm; (b) the initial LASox glass and glass heat-treated up to 900 °C in the spectral range from 250 nm to 3300 nm; (c) the LAS glass and glass heat-treated up to 780 °C in the spectral range from 300 nm to 1750 nm; (d) the subtracted spectra; (e) the LAS glass and glass heat-treated up to 1000 °C in the spectral range from 300 nm to 460 nm; (f) the LASox glass and glass heat-treated up to 900 °C in the spectral range from 300 nm to 475 nm; (g) the LAS glass and glass heat-treated up to 1000 °C in the spectral range from 2600 nm to 3300 nm; (h) the LASox glass and glass heat-treated up to 900 °C in the spectral range from 2600 nm to 3300 nm. The duration of each hold is 6 h.
Materials 18 00785 g013aMaterials 18 00785 g013b
Figure 14. Absorption spectra of the samples of: (a) the initial LAS and LASox glasses; (b) the LAS and LASox glasses heat-treated at 680 °C; (c) the LAS and LASox glasses heat-treated at 680 °C + 720 °C; (d) the LAS and LASox glasses heat-treated at 680 °C + 750 °C; (e) the LAS and LASox glasses heat-treated at 680 °C + 780 °C; (f) the LAS and LASox glasses heat-treated at 680 °C + 800 °C; (g) the LAS and LASox glasses heat-treated at 680 °C + 850 °C; (h) the LAS and LASox glasses heat-treated at 680 °C + 900 °C. The duration of each hold is 6 h.
Figure 14. Absorption spectra of the samples of: (a) the initial LAS and LASox glasses; (b) the LAS and LASox glasses heat-treated at 680 °C; (c) the LAS and LASox glasses heat-treated at 680 °C + 720 °C; (d) the LAS and LASox glasses heat-treated at 680 °C + 750 °C; (e) the LAS and LASox glasses heat-treated at 680 °C + 780 °C; (f) the LAS and LASox glasses heat-treated at 680 °C + 800 °C; (g) the LAS and LASox glasses heat-treated at 680 °C + 850 °C; (h) the LAS and LASox glasses heat-treated at 680 °C + 900 °C. The duration of each hold is 6 h.
Materials 18 00785 g014aMaterials 18 00785 g014b
Table 1. Characteristic temperatures of initial quenched and heat-treated glasses shown in DSC curves.
Table 1. Characteristic temperatures of initial quenched and heat-treated glasses shown in DSC curves.
Heat-Treatment RegimeTg, °CTon1, °CTmax1, °CTon2, °CTmax2, °CTmax3, °C
The LAS glass
quenched698--8478671175
680 °C, 6 h6998108398829091142
700 °C, 6 h7057648078789001135
720 °C, 6 h7087658158718901139
The LASox glass
quenched699-8028518741186
680 °C, 6 h6957998318769011136
700 °C, 6 h7047798098798991141
720 °C, 6 h7087558288769001136
Table 2. Lattice parameters and mean crystallite sizes of γ-Al2O3, β-quartz ss and β-spodumene ss in LAS and LASox glass-ceramics.
Table 2. Lattice parameters and mean crystallite sizes of γ-Al2O3, β-quartz ss and β-spodumene ss in LAS and LASox glass-ceramics.
Heat-Treatment Regimeγ-Al2O3β-Quartz ssβ-Spodumene ss
a, ÅD, nma, Åc, ÅD, nma, Åc, ÅD, nm
The LAS glass
680 °C, 6 h + 720° C, 6 h
680 °C, 6 h + 750 °C, 6 h7.9184.5
680 °C, 6 h + 780° C, 6 h7.9218.05.2075.33626
680 °C, 6 h + 800 °C, 6 h7.92514.05.1965.35025
680 °C, 6 h + 850 °C, 6 h 5.1895.37022
680 °C, 6 h + 1000 °C, 6 h 5.1865.36228
680 °C, 6 h + 1100 °C, 6 h 7.5529.14545
680 °C, 6 h + 1200 °C, 6 h 7.5419.14345
The LASox glass
680 °C, 6 h + 720° C, 6 h
680 °C, 6 h + 750 °C, 6 h7.9167.3 28
680 °C, 6 h + 780° C, 6 h7.9159.45.2075.31926
680 °C, 6 h + 800 °C, 6 h7.92110.35.1975.33726
680 °C, 6 h + 850 °C, 6 h 5.1925.36521
680 °C, 6 h + 1000 °C, 6 h 5.1875.36827
680 °C, 6 h + 1100 °C, 6 h 7.5509.15336
680 °C, 6 h + 1200 °C, 6 h 7.5519.15036
Table 3. Lattice parameters and mean crystallite sizes of Al2TiO5 in LAS and LASox glass-ceramics.
Table 3. Lattice parameters and mean crystallite sizes of Al2TiO5 in LAS and LASox glass-ceramics.
Heat-Treatment RegimeAl2TiO5
The LAS GlassThe LASox Glass
a, Åb, Åc, ÅV, Å3D, nma, Åb, Åc, ÅV, Å3D, nm
680 °C, 6 h + 780° C, 6 h3.5949.3569.557321.36.23.5939.5419.577321.48.0
680 °C, 6 h + 800 °C, 6 h3.5959.3659.564322.07.13.5949.3589.577322.18.7
680 °C, 6 h + 850 °C, 6 h3.5929.3789.560322.08.33.5959.3859.589323.510.2
680 °C, 6 h + 900 °C, 6 h3.5969.3739.570322.69.03.5949.4189.619325.612.0
680 °C, 6 h + 950 °C, 6 h3.5979.4549.590326.19.43.5949.4219.628326.012.0
680 °C, 6 h + 1000 °C, 6 h3.5989.4569.624327.412.23.5949.4579.633327.413.7
680 °C, 6 h + 1100 °C, 6 h3.5969.4569.663328.618.83.5959.4609.677329.119.4
680 °C, 6 h + 1200 °C, 6 h3.5919.4889.716331.044.13.5929.4899.708330.942.0
Table 4. Coefficients of thermal expansion (CTE) for precursor and heat-treated glasses of the LAS and the LASox compositions in the temperature range from room temperature to 320 °C.
Table 4. Coefficients of thermal expansion (CTE) for precursor and heat-treated glasses of the LAS and the LASox compositions in the temperature range from room temperature to 320 °C.
Heat-Treatment RegimeCTE20-320 (×10−6 K−1)
LASLASox
Precursor glasses4.35 ± 0.084.25 ± 0.08
680 °C, 6 h4.19 ± 0.084.21 ± 0.08
680 °C, 6 h + 720 °C, 6 h4.82 ± 0.094.57 ± 0.09
680 °C, 6 h + 720 °C, 6 h + 750 °C, 6 h5.36 ± 0.105.18 ± 0.10
680 °C, 6 h + 720 °C, 6 h + 750 °C, 6 h + 780 °C, 6 h4.30 ± 0.084.85 ± 0.09
680 °C, 6 h + 780 °C, 6 h-4.85 ± 0.10
720 °C, 6 h + 800 °C, 6 h1.15 ± 0.021.24 ± 0.01
720 °C, 6 h + 850 °C, 6 h0.61 ± 0.010.45 ± 0.08
680 °C, 6 h +720 °C, 6 h +750 °C, 6 h +950 °C, 6 h0.30 ± 0.010.19 ± 0.01
680 °C, 6 h +1200 °C, 6 h1.49 ± 0.031.42 ± 0.03
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

Maltsev, S.; Dymshits, O.; Alekseeva, I.; Volokitina, A.; Tenevich, M.; Bachina, A.; Bogdanov, K.; Zapalova, S.; Shakhgildyan, G.; Zhilin, A. Glass-Ceramics of the Lithium Aluminosilicate System Nucleated by TiO2: The Role of Redox Conditions of Glass Melting in Phase Transformations and Properties. Materials 2025, 18, 785. https://doi.org/10.3390/ma18040785

AMA Style

Maltsev S, Dymshits O, Alekseeva I, Volokitina A, Tenevich M, Bachina A, Bogdanov K, Zapalova S, Shakhgildyan G, Zhilin A. Glass-Ceramics of the Lithium Aluminosilicate System Nucleated by TiO2: The Role of Redox Conditions of Glass Melting in Phase Transformations and Properties. Materials. 2025; 18(4):785. https://doi.org/10.3390/ma18040785

Chicago/Turabian Style

Maltsev, Stanislav, Olga Dymshits, Irina Alekseeva, Anna Volokitina, Maksim Tenevich, Anastasia Bachina, Kirill Bogdanov, Svetlana Zapalova, Georgiy Shakhgildyan, and Aleksandr Zhilin. 2025. "Glass-Ceramics of the Lithium Aluminosilicate System Nucleated by TiO2: The Role of Redox Conditions of Glass Melting in Phase Transformations and Properties" Materials 18, no. 4: 785. https://doi.org/10.3390/ma18040785

APA Style

Maltsev, S., Dymshits, O., Alekseeva, I., Volokitina, A., Tenevich, M., Bachina, A., Bogdanov, K., Zapalova, S., Shakhgildyan, G., & Zhilin, A. (2025). Glass-Ceramics of the Lithium Aluminosilicate System Nucleated by TiO2: The Role of Redox Conditions of Glass Melting in Phase Transformations and Properties. Materials, 18(4), 785. https://doi.org/10.3390/ma18040785

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