Next Article in Journal
Self-Lubricating Materials for Extreme Condition Applications
Previous Article in Journal
Hydrothermal Co-Processing of Coal Fly Ash Cenospheres and Soluble Sr(II) as Environmentally Sustainable Approach to Sr-90 Immobilization in a Mineral-like Form
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

First Principles Calculations of Atomic and Electronic Structure of TiAl3+- and TiAl2+-Doped YAlO3

1
Institute of Solid State Physics, University of Latvia, Kengaraga Street 8, LV-1063 Riga, Latvia
2
College of Sciences & CQUPT-BUL Innovation Institute, Chongqing University of Posts and Telecommunications, Chongqing 400065, China
3
Institute of Physics, University of Tartu, W. Ostwald Str. 1, 50411 Tartu, Estonia
4
Academy of Romanian Scientists, Ilfov Str. No. 3, 050044 Bucharest, Romania
5
Faculty of Science and Technology, Jan Długosz University, Armii Krajowej 13/15, PL-42200 Częstochowa, Poland
6
Department of Inorganic Chemistry, Uzhhorod National University, 46 Pidhirna Str., 88000 Uzhhorod, Ukraine
*
Author to whom correspondence should be addressed.
Materials 2021, 14(19), 5589; https://doi.org/10.3390/ma14195589
Submission received: 24 August 2021 / Revised: 21 September 2021 / Accepted: 22 September 2021 / Published: 26 September 2021

Abstract

:
In this paper, the density functional theory accompanied with linear combination of atomic orbitals (LCAO) method is applied to study the atomic and electronic structure of the Ti3+ and Ti2+ ions substituted for the host Al atom in orthorhombic Pbnm bulk YAlO3 crystals. The disordered crystalline structure of YAlO3 was modelled in a large supercell containing 160 atoms, allowing simulation of a substitutional dopant with a concentration of about 3%. In the case of the Ti2+-doped YAlO3, compensated F-center (oxygen vacancy with two trapped electrons) is inserted close to the Ti to make the unit cell neutral. Changes of the interatomic distances and angles between the chemical bonds in the defect-containing lattices were analyzed and quantified. The positions of various defect levels in the host band gap were determined.

1. Introduction

It has long been well known that compounds of Y, Al and O form the following three crystal structures: Y3Al5O12 (YAG), Y4Al2O9 (YAM) and YAlO3 (YAP). There are numerous studies of their spectroscopic properties, in pure form and with different impurities. For example, Ce3+ spectra in YAG were studied recently in Refs. [1,2,3,4]. The same Ce3+ impurity in YAM was a subject of spectroscopic investigations in Refs. [5,6], whereas triply (Yb3+/Ho3+/Tm3+) doped YAM nanoparticles were investigated in Ref. [7].
Yttrium aluminum perovskite (YAP) plays an important role in materials science research as an excellent model system for in-depth studies of the defect formation and their influence on its optical properties, on the one hand, and development and improvement of the already existing applications, on the other. Such an interest is explained by two main factors. First of all, there exists the possibility of doping with many transition metal (TM) and rare earth (RE) ions, and not all crystals can offer such an opportunity. Various TM ions can occupy the Al site, whereas RE ions can be easily incorporated at the Y site. The second important circumstance is that YAP has a wide band gap, which in different publications was reported to range from 8.5 eV to 9.0 eV [8,9]. Many defects and/or impurity ion energy levels can be located within the forbidden gap, which will manifest themselves in the appearance of additional absorption/emission peaks in the defect-containing YAP spectra [9,10,11,12,13,14].
Historically, interest in YAP as an attractive material for doping with transition metal and rare earth ions has been driven due to its potential application [13,14] in solid-state lasers capable of operating at shorter wavelengths than classical Al2O3:Ti lasers [15]. It is expected that tunable laser crystal material, such as Ti-doped YAlO3, can demonstrate absorption range from 350 to 550 nm, as well as fluorescence range from 540 to 800 nm [16]. Because of these expectations, research on optical, luminescence and laser properties has been under the study of several research centers. Yamaga et al. [15] have performed detailed low temperature measurements of the zero-phonon lines of Ti3+ ions in YAP, YAG and Al2O3 at 10 K. The emission spectrum of Ti3+ in YAP consists of an intense zero-phonon line at 539.7 nm and with full width at half maximum (FWHM) of 30 cm−1 accompanied by a rather weak line at shorter wavelength 537.1 nm with FWHM of 55 cm−1. It was reported that the zero-phonon lines of Ti3+:YAP are strongly polarized perpendicular to the tetragonal axis and those of Ti3+:Al2O3 parallel to the trigonal axis. Detailed study of the donor and acceptor states and corresponding charge-transfer (CT) transitions in YAP:Ti3+/Ti4+ was done by Basun at al [9]. In particular, through the conductivity, optical absorption and luminescence measurements, it was found that the Ti3+ ground state is situated at about 4.60 ± 0.1 eV below the conduction band. Furthermore, a titanium-bound exciton band at about 3.60 eV above the Ti3+ ground state was identified. Another important result is that CT transition from the Ti4+ to Ti3+ ground state was found to be at 4.20 ± 0.1 eV. The results obtained suggest that the band gap energy Eg of YAP is 8.8 ± 0.2 eV. This value is in good agreement with data of Lushchik et al. [8]; from there we can conclude that Eg is in the range 8.5–9.0 eV. Two different processes, namely, Ti3+/Ti4+ pair formation and corresponding hole trapping at oxygen ions adjacent to a lattice defect are responsible for the broad parasitic photo-induced absorption [17].
A few examples that emphasize the diversity of applications of the doped YAP host are as follows: (i) scintillating applications (especially when doped with the La3+, Ce3+, Pr3+ ions) [18,19,20,21,22,23,24,25,26,27,28]; (ii) thermoluminescent material (YAP with Mn ions) [29,30,31,32,33,34]; (iii) solid-state laser applications [35,36] and (iv) electroluminescent devices [37]. Various aspects of the doped YAP crystal were studied theoretically, such as crystal field splittings of the Mn3+ and Mn4+ energy levels [38], splittings of the Ce3+ 4f and 5d states [39] and Ce3+ 4f ground state position in the YAP band gap [40,41]. The titanium impurities in YAP were studied experimentally in a number of references, for example, Ti3+ and Ti4+ excitation spectra [42,43,44,45], EPR spectra of Ti3+ [17,46] and charge transfer processes [9]. The trap defect levels in the YAP band gap were analyzed experimentally in Refs. [18,47]. The existing measurements of the structural optical, thermal and mechanical properties of pure and Ti-doped YAlO3 important for practical applications are summarized in Table 1.
As seen in the above given references, a vast majority of publications were focused on the Ti3+ or Ti4+ ions. However, if the Ti3+ ions capture an electron, another type of impurity center, Ti2+, will be formed. Such a center in YAP is less studied, and the present paper offers a theoretical insight into its properties.
One of the key factors that plays a crucial role in assessing perspectives of a given material for its optical applications is the location of the impurity ion ground state in the host band gap. In view of this, in the present paper we continue our earlier studies of the defects electronic structure in the YAP crystal [34,38,39,48,49] by performing the first principles calculations for the pristine YAP, Ti3+ and Ti2+ doped YAP as well as for the F color center (the oxygen vacancy with two electrons) in the YAP lattice. All calculations have been performed within the formalism of hybrid density functional theory (DFT). Comparison of the calculated results for the defect-containing YAP with those for the pristine material allowed us to establish the influence of defects on the structural and electronic properties of the studied host material.
Table 1. Structural and optical properties of orthorhombic YAP and YAP:Ti.
Table 1. Structural and optical properties of orthorhombic YAP and YAP:Ti.
a0, b0, c0 (Å)5.179, 5.329, 7.370[50]
Y sites/cm31.97 × 1022
Bulk modulus, B (GPa)192[50]
Hardness (Knoop value)1310[13]
Melting point (°C)1875[13]
Thermal conductivity (W/cm-°C at 25°C)0.11[13]
Thermal expansion coefficient (10−6/°C, at 25°C)2.2[13]
Optical transmission (absorption coefficient < 1.0 cm−1)0.29–5.9 μ[13]
Refractive index (6328 Å)
(a = z, b = y, c = x)
na = 1.97
nb = 1.96
nc = 1.94
[13]
Band gap energy, eV8.5–9.0
8.8 ± 0.2
[8]
[9]
Ti-ion related optical absorption bands (eV)7.08 (Ti3+-related)
5.76 (Ti4+-related)
5.27 (Ti4+-related)
5.39 (Ti3+ → Ti4+)
4.42 (Ti3+-related)
4.42 (2T2g → 4S, Ti3+)
4.20 (Ti4+ → Ti3+)
2.88 (2T2g2E, Ti3+)
2.87 (Ti3+)
2.81 (Ti3+)
2.53 (Ti3+)
2.52 (Ti3+)
2.50 (2T2g2E, Ti3+)
[9]
[9]
[9]
[16]
[9]
[43]
[9]
[43]
[16]
[9]
[9]
[16]
[43]
Ti-ion luminescence band3.02 (Ti4+)
2.10 (Ti3+, d-d transition)
2.06 (Ti3+, d-d transition)
1.99 eV (Ti3+, d-d transition)
[9]
[43]
[9]
[16]
F center:
optical absorption (eV)
luminescence (eV)

5.84 and 5.15
2.95

[51]
[51]
F+ center:
optical absorption (eV)luminescence (eV)

6.5, 5.63, and 4.3
3.49

[51]
[51]
The paper is organized as follows: Section 2 contains all relevant details of the YAP structure and calculating settings, Section 3 describes all calculated results in relation to the experimental data and (when available) calculated values, and, finally, Section 4 concludes the paper with a short summary.

2. Computational Details

The DFT approach is based on the Hartree-Fock method using the HSE06 Hamiltonian [52,53] as implemented in total energy CRYSTAL17 computer code [54], which was used in all calculations. The computer simulation began from complete geometry optimization to be sure our modelling was able to reproduce the basic crystal parameters, e.g., lattice constant and optical band gap of ideal YAP. Then, the influence of the presence of the Ti3+ dopant and the Ti2+ dopant and its compensating F-center on YAP atomic and electronic structures were studied using a supercell (SC) approach. The localized Gaussian-type functions (GTFs) in the form of basis set (BS) centered on atomic nuclei for expansion of crystalline orbitals as linear combinations of atomic orbitals (CO LCAO) were employed. The hybrid exchange–correlation functional HSE06 has been used according to direction given in Refs. [52,53,54]. The Triple-Zeta Valence with Polarization quality BSs for titanium, alumina and oxygen were taken from Ref. [55], while for yttrium the Hay-Wadt effective core pseudopotential BS was taken directly from Ref. [54]. Use of LCAO GTF approach greatly facilitates the calculation of effective charges and bond population from the wave functions using the Mulliken population scheme [56,57].
YAP possesses a perovskite structure of the GdFeO3 type [58]. Its orthorhombic unit cell (Pbnm space group) contains four formula units with two non-equivalent oxygen positions. In order to simulate the isolated substitution defect, the orthorhombic unit cell was extended to a 2 × 2 × 2 SC. The SC retains the orthorhombic symmetry and contains 160 atoms with the distance between periodically repeated defects of 10.38 Å [48]. To model the F-center, we employed “ghost” BS [53], meaning the atomic orbitals of removed oxygen are left behind. This technique facilitates the accurate description of the electron distribution within the vacancy and allows us to mimic the F+-center, which is a compensating defect to simulate a Ti2+ substitutional dopant. The equilibrium geometry is obtained using an analytical optimization method as implemented in the CRYSTAL code [54]. To provide a balanced summation in both direct and reciprocal lattices, the reciprocal space integration has been performed by sampling the SCs Brillouin zone with a 3 × 3 × 3 Pack-Monkhorst mesh [59] that gives 14 k-points in total. The calculations have been considered as converged only when the total energy differs by less than 10−8 a.u. in two successive cycles of the SCF (Self-Consistent Field) procedure. Within the self-consistency, the accuracies (tolerances) of 10−8 were chosen for calculations of Coulomb and exchange integrals [54]. The visualizing software VESTA (https://jp-minerals.org/vesta/en/, accessed on 25 September 2021) was used to present equilibrium YAP structures in Figure 1.

3. Results and Discussion

3.1. Structural Properties

Figure 1 shows the considered supercells for each of the above-described structural models. Influence of the structural defects (substitutional atoms and vacancies) on the YAP lattice geometry is considered by comparing the optimized interatomic distances and angles between chemical bonds for all above-described structural models.
Table 2 lists the lattice constants and atomic coordinates obtained in this study, as well as their experimental values [50], calculated using the HSE06 exchange-correlation function within DFT [60]. It can be easily seen that the equilibrium lattice constants (a0 = 5.189 Å, b0 = 5.317 Å, and c0 = 7.388 Å) calculated in this study are in very good agreement with both those experimentally measured (a0 = 5.179 Å, b0 = 5.329 Å, and c0 = 7.370 Å [50]) and those previously calculated (a0 = 5.179 Å, b0 = 5.342 Å, and c0 = 7.367 Å [60]). In addition, the experimental and calculated fractional coordinates of all ions in a unit cell expressed in the crystal lattice constants’ units are also very close to one another. Thus, we conclude that the chosen computational approach allows us to obtain reliable data for the Ti-doped YAP crystal.
The nearest environment of the Al3+ ions in YAlO3 is a deformed octahedron made of six oxygens. The second coordination sphere around Al3+ ions contains six Y3+ ions. The oxygen ions laying on the opposite sides of the central Al3+ ion are located on the same straight line with the central ion, making the angles of 180 degrees—this is an important observation, which will be used to illustrate the difference with the defect-containing YAlO3 lattice.
Isovalent substitution of the Al3+ ion for the Ti3+ ion is accompanied by the expansion of the TiO6 octahedron, which is in line with a greater Ti3+ ionic radius in six-fold coordination (0.67 Å, if compared with 0.535 Å for Al3+ [61]. All O–Ti–O angles inside the TiO6 octahedron are not changed, which is anticipated since the electric charges of the substituting and substituted ions are the same. Therefore, such an isovalent doping leads only to an expansion of the crystal lattice while keeping all angles between the chemical bonds the same, as in the neat lattice.
Presence of the oxygen vacancy or substitution of the Al3+ by the Ti2+ ion considerably lowers the symmetry of the AlO6 or TiO6 octahedron because the oxygen ions located on opposite sides of the central ion are no longer on the same straight line, deviating from the 180° angles (Table 3). In the long run, such symmetry lowering may lead to the enhancement of the emission lines of impurity ions located at those sites. It should be noted that the presence of the oxygen vacancy considerably shortens the distance to the nearest Al (or Ti) ion (Figure 2 and Table 3).

3.2. Electronic Properties

In the ternary compounds, such as YAP, two pairs of bonding atoms can be selected, like Y–O and Al–O in this particular case. Their chemical properties are not identical, as they depend on the atomic numbers, charges, electron configurations and electronegativities of the ions involved into these bond formations. The peculiar bonding features can be analyzed by considering the effective Mulliken charges and bond orders as follows. The calculated Mulliken charges (Table 4) deviate from those prescribed by the formal valences derived from the chemical formula (+3 for Y and Al, −2 for O), which is an indication of the covalent bonds between the ions in YAlO3 crystal lattice. The charge of the yttrium ions is closer to the “+3” value than that of the aluminium ions; therefore, the Y–O ions are more ionic, whereas the Al–O ions are more covalent. The oxygen ions effective charges for a pristine YAlO3 also follow the geometrical structure: the charges of the ions located at the same distance from the central Al3+ ion are the same. Moreover, the absolute values of the ions located further from the central ion are slightly greater than those ones of the closer ions. This can be explained by weaker overlap of the oxygen and aluminium wave functions and increased degree of ionicity of the corresponding Al–O bonds. F-centers in pristine and Ti2+-doped YAP attract 0.53 e and 0.66 e, correspondingly.
The atomic orbitals of the constituting atoms in the YAP lattice are hybridized because of chemical bond formation. Due to these effects, the oxygen 2p ions produce a minor contribution to the conduction band, whereas the cation’s unfilled states can be traced down by their small percentage in the valence band. These effects are different in the Y–O and Al–O pairs and are represented in detail in Table 4, which lists the results of the Mulliken effective charge and bond order calculations.
Figure 3 and Figure 4 show band structures and density of states (DOS) calculated for pristine, F-center containing, Ti3+- and Ti2+-YAP. The direct Γ-Γ band gap calculated in this study for pristine YAP has a width of 7.26 eV [48] and is in good agreement with the experimentally detected band gap of 8.5 eV [8]. The band structures calculated for all four perovskite structures under study look similar and agree with band structures published previously in the literature using different ab initio methods [60]. Position of the bottom of the conduction band (CB) practically does not change with respect to the top of the valence band (VB), while the levels induced by defects are present within the YAP’s optical gap. The VB top and CB bottom of pristine YAP consist of the O 2p and Y 4d states with an admixture of the Al 3p orbitals, respectively.
In the case of F-center (Figure 3b and Figure 4b), an occupied level is located at 3.36 eV above of the VB top. This level has a very small dispersion of 25 meV, meaning the F-center can be treated as an isolated defect within the framework of the chosen model. The defect level induced by F-center consist mainly of admixture of F-center orbitals with the O 2p and Al 3p orbitals (Figure 4b) that is in line with relatively short Al–F-center bond of 1.22 Å (Table 3). An empty level also induced by F-center in pristine YAP is located very close to its bottom of the CB (Figure 4b) and contains also an admixture of the Y 4d states. In fact, this empty level may form the bottom of the CB, making YAP to act as an indirect Γ-Z semiconductor.
The Ti3+ ion substituting the host Al ion in the YAP lattice induces three defect levels inside its optical gap (Figure 4c). Due to distortion of oxygen octahedra around the Ti dopant, the eg-t2g splitting of the Ti 3d orbitals takes place. The occupied t2g level is located 5.32 eV above the VB top and has the dispersion (energy interval, in which this level is localized) of 20 meV. This level consists mainly of the Ti 3d states with a small admixture of the O 2p states. The other two empty levels are located close to each other at 6.31 and 6.40 eV above the top of the VB, respectively. Their dispersions are 5 and 10 meV, respectively. Empty levels induced by the Ti3+ dopant consist of the Ti 3d states with a small admixture of the O 2p and Al 3p orbitals (Figure 3c).
YAP doped with the Ti2+ ions contains a compensating F-center to force a unit cell to be neutral. Presence of the F-center induces an occupied level in the YAP crystal at 3.20 eV above its VB top (Figure 3d). This level has a dispersion of 25 meV and consists of an admixture of F-center states, O 2p and Ti 3d orbitals (Figure 4d). The other three levels induced by the Ti2+ dopant are located at 4.98, 5.93, and 6.30 eV above the VB top, respectively (Figure 3d). The level at 4.98 eV is occupied, while two others are vacant. These three levels’ dispersions are 27, 7, and 38 meV, respectively.
Presence of defects, which affects the effective charges of the surrounding ions, also modifies slightly the VB profile, as is shown by Figure 4.
We note that the YAP host cannot be directly doped with the Ti2+. However, formation of Ti2+F-center complexes is expected in YAP during, e.g., hard irradiation. Formation of titanium ions in this oxidation state is physically unavoidable for at least two reasons: (i) the perovskite crystals always contain oxygen vacancies. Removal of the negative charge would imply removal of the positive charge as well, hence Ti2+ can appear instead of Ti3+; (ii) the charge transfer transitions from the oxygen ions to the Ti3+ ions would lead to the formation of the Ti2+. According to the Mulliken population analysis performed in this study (Table 4), we predict the formation of Ti2+ ion with Mulliken effective charge of 1.07 e if the F-center is created in the Ti-doped YAP crystal. The Mulliken effective charge calculated for Ti3+-doped YAP if no F-center is present is 1.56 e (Table 4). Fermi level of Ti3+-doped YAP is positioned 5.31 eV above the top of the VB of the undoped crystal, while in case of possible formation of Ti2+F-center complex the Fermi level is located 4.94 eV above of the VB top of pristine YAP or 0.37 eV below the Fermi level of Ti3+-doped crystal, assuming the coexistence of the Ti3+/Ti2+ dopants in YAP under, e.g., ionizing irradiation conditions when formation of oxygen vacancies (F-centers) is highly expected. In addition, we note that the formation of the Ti2+-dopant in YAP without the presence of the F-center is hardly possible and thus is beyond the scope of this study.
Finally, we must emphasize that understanding where the impurity levels are located in the band gap, how these levels change their position with a change in the charge state of the impurity ion and how they are located in relation to each other, as well as in relation to point defects (F-centers)—all this is very important for an accurate and detailed description of the processes of photo- and thermally stimulated conversion of point defects in wide-gap halides, oxides and perovskites [62,63,64,65,66,67,68,69,70,71,72,73,74,75,76].

4. Conclusions

Detailed calculations of the structural and electronic properties of the pristine defect-containing YAlO3 were performed in the present paper. Among the considered defects were F-center (oxygen vacancy), Ti3+ and Ti2+ ions, introduced instead of the Al3+ ion. Optimization of the crystal structures for each considered case has led to the following conclusions:
(i)
Isovalent replacement of the Al3+ ion by the Ti3+ ion keeps the symmetry of the substitutional site; the only effect is a slight increase of the Ti–O distances as compared to the Al–O ones because of the difference in ionic radii of the Al3+ and Ti3+ ions. Our calculated position of the Ti3+ ground state in the YAP band gap (5.32 eV above the valence band top) agrees nicely with the value of about 5.21 eV derived by Rogers and Dorenbos in Ref. [77].
(ii)
Appearance of the oxygen vacancy or replacement of the Al3+ ion by the Ti2+ ion with simultaneous formation of the charge compensating defects lowers the symmetry of the substitutional site by modifying the angles between the chemical bonds in the AlO6 or TiO6 octahedra.
(iii)
Creation of the defects in the pristine YAlO3 structure leads to the formation of localized dispersionless defects’ energy levels in the host band gap. These levels are located in the central region of the band gap and closer to the CB bottom; they can significantly modify the host’s optical properties by causing additional defect-related absorption peaks in the optical spectra.
We note that appearance of the substitutional defects and F-centers in the ideal YAP crystal structure will be accompanied by (i) emergence of local vibrational modes and (ii) splitting of the host’s degenerated vibrational modes due to the symmetry lowering around these point defects, which is currently out of scope of the present paper.

Author Contributions

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

Funding

M.G.B. appreciates support from the Chongqing Recruitment Program for 100 Overseas Innovative Talents (grant no. 2015013), the Program for the Foreign Experts (grant no. W2017011), Wenfeng High-end Talents Project (grant no. W2016-01) offered by the Chongqing University of Posts and Telecommunications (CQUPT), Estonian Research Council grant PUT PRG111, European Regional Development Fund (TK141), and NCN project 2018/31/B/ST4/00924. This study was supported by a grant from Latvian Research Council No. LZP-2018/1-0214 (for AIP). Institute of Solid State Physics, University of Latvia as the Center of Excellence has received funding from the European Union’s Horizon 2020 Framework Program H2020-WIDESPREAD-01-2016-2017-TeamingPhase2 under Grant Agreement No. 739508, project CAMART2.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw/processed data required to reproduce these findings cannot be shared at this time as the data also form a part of an ongoing study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Przybylińska, H.; Ma, C.G.; Brik, M.G.; Kamińska, A.; Sybilski, P.; Wittlin, A.; Berkowski, M.; Zorenko, Y.; Gorbenko, V.; Wrzesinski, H.; et al. Electronic structure of Ce3+ multicenters in yttrium aluminum garnets. Appl. Phys. Lett. 2013, 102, 241112. [Google Scholar] [CrossRef] [Green Version]
  2. Polisadova, E.; Valiev, D.; Vaganov, V.; Oleshko, V.; Han, T.; Zhang, C.; Burachenko, A.; Popov, A.I. Time-resolved cathodoluminescence spectroscopy of YAG and YAG:Ce3+ phosphors. Opt. Mater. 2019, 96, 109289. [Google Scholar] [CrossRef]
  3. Karipbayev, Z.T.; Lisitsyn, V.M.; Mussakhanov, D.A.; Alpyssova, G.K.; Popov, A.I.; Polisadova, E.F.; Elsts, E.; Akilbekov, A.T.; Kukenova, A.B.; Kemere, M.; et al. Time-resolved luminescence of YAG:Ce and YAGG:Ce ceramics prepared by electron beam assisted synthesis. Nucl. Instrum. Methods B 2020, 479, 222–228. [Google Scholar] [CrossRef]
  4. Lisitsyn, V.M.; Tulegenova, A.T.; Lisitsyna, L.A.; Vaganov, V.A.; Soshchin, N.P.; Polisadova, E.F.; Abdullin, K.A.; Yangyang, J. Photo and cathodoluminescence of commercial YAG:Ce based phosphors in UV region. Nucl. Instrum. Methods B 2020, 478, 120–124. [Google Scholar] [CrossRef]
  5. Wang, Y.; Hrubiak, R.; Turczyński, S.; Pawlak, D.A.; Malinowski, M.; Włodarczyk, D.; Kosyl, K.M.; Paszkowicz, W.; Przyybylińska, H.; Wittlin, A.; et al. Spectroscopic properties and martensitic phase transition of Y4Al2O9: Ce single crystals under high pressure. Acta Mater. 2019, 165, 346–361. [Google Scholar] [CrossRef]
  6. Fetliński, B.; Turczyński, S.; Malinowski, M.; Szczepański, P. Down-Shifting in the YAM:Ce3++ Yb3+ System for Solar Cells. Materials 2021, 14, 2753. [Google Scholar] [CrossRef]
  7. Gomes, P.F.; Silva, T.C.; Maia, L.J.; Carvalho, J.F. Synthesis and visible down-and up-conversion emissions from Yb3+/Ho3+/Tm3+ Co-Doped Y4Al2O9 (YAM) nanocrystalline particles. J. Lumin. 2020, 227, 117554. [Google Scholar] [CrossRef]
  8. Lushchik, C.; Feldbach, E.; Frorip, A.; Kirm, M.; Lushchik, A.; Maaroos, A.; Martinson, I. Multiplication of electronic excitations in CaO and YAlO3 crystals with free and self-trapped excitons. J. Phys. Condens. Matter 1994, 6, 11177. [Google Scholar] [CrossRef]
  9. Basun, S.A.; Danger, T.; Kaplyanskii, A.A.; McClure, D.S.; Petermann, K.; Wong, W.C. Optical and photoelectrical studies of charge-transfer processes in YAlO3: Ti crystals. Phys. Rev. B 1996, 54, 6141. [Google Scholar] [CrossRef]
  10. Shimizu, Y.; Ueda, K.; Inaguma, Y. Photoluminescence excitation spectra of lanthanide doped YAlO3 in vacuum ultraviolet region. Opt. Mater. 2017, 66, 327–331. [Google Scholar] [CrossRef]
  11. Eglitis, R.I.; Popov, A.I. Comparative Ab initio calculations for ABO3 perovskite (001), (011) and (111) as well as YAlO3 (001) surfaces and F centers. J. Nano-Electron. Phys. 2019, 11, 01001. [Google Scholar] [CrossRef] [Green Version]
  12. Zorenko, Y.V.; Voloshinovskiĭ, A.S.; Konstankevych, I.V. Luminescence of F+ and F centers in YAlO3. Opt. Spectrosc. 2004, 96, 532–537. [Google Scholar] [CrossRef]
  13. Weber, M.J.; Bass, M.; Andringa, K.; Monchamp, R.R.; Comperchio, E. Czochralski growth and properties of YAlO3 laser crystals. Appl. Phys. Lett. 1969, 15, 342–345. [Google Scholar] [CrossRef]
  14. Massey, G.A. Criterion for selection of cw laser host materials to increase available power in the fundamental mode. Appl. Phys. Lett. 1970, 17, 213–215. [Google Scholar] [CrossRef]
  15. Yamaga, M.; Henderson, B.; O’Donnell, K.P.; Rasheed, F.; Gao, Y.; Cockayne, B. Polarization of emission spectra from Ti3+-doped oxide crystals-II. Zero-phonon emission lines. Appl. Phys. B 1991, 52, 225–229. [Google Scholar] [CrossRef]
  16. Wegner, T.; Petermann, K. Excited state absorption of Ti3+: YAlO3. Appl. Phys. B 1989, 49, 275–278. [Google Scholar] [CrossRef]
  17. Chosrovian, H.; Rentsch, S.; Stösser, R. ESR investigations and mechanisms of light-induced optical absorption in Ti3+: YAlO3 crystals. Phys. Status Solidi A 1992, 130, 463–473. [Google Scholar] [CrossRef]
  18. Moore, M.E.; Delzer, C.; Watts, J.; Musicó, B.L.; Xu, C.; Collette, R.M.; Rack, P.D.; Zhang, Y.; Melcher, C.L.; McConchie, S.; et al. Studying the effects of thermally diffusing Ce into the surface of YAlO3 for associated particle imaging. Nucl. Instrum. Methods B 2020, 473, 55–61. [Google Scholar] [CrossRef]
  19. Nakanishi, K.; Yamamoto, S.; Kamada, K.; Yoshikawa, A. Performance evaluation of YAlO3 scintillator plates with different Ce concentrations. Appl. Radiat. Isot. 2021, 168, 109483. [Google Scholar] [CrossRef] [PubMed]
  20. Dorenbos, P. Fundamental Limitations in the Performance of Ce3+–, Pr3+–, and Eu2+–Activated Scintillators. IEEE Trans. Nucl. Sci. 2010, 57, 1162–1167. [Google Scholar] [CrossRef]
  21. Baryshevsky, V.G.; Korzhik, M.V.; Moroz, V.I.; Pavlenko, V.B.; Fyodorov, A.A.; Smirnova, S.A.; Egorycheva, O.A.; Kachanov, V.A. YAlO3:Ce-fast-acting scintillators for detection of ionizing radiation. Nucl. Instrum. Methods B 1991, 58, 291–293. [Google Scholar] [CrossRef]
  22. Vedda, A.; Martini, M.; Meinardi, F.; Chval, J.; Dusek, M.; Mares, J.A.; Mihokova, E.; Nikl, M. Tunneling process in thermally stimulated luminescence of mixed LuxY1−xAlO3:Ce crystals. Phys. Rev. B 2000, 61, 8081. [Google Scholar] [CrossRef]
  23. Zorenko, Y.; Gorbenko, V.; Konstankevych, I.; Voznjak, T.; Savchyn, V.; Nikl, M.; Mares, J.A.; Nejezchleb, K.; Mikhailin, V.; Kolobanov, V.; et al. Peculiarities of luminescence and scintillation properties of YAP:Ce and LuAP:Ce single crystals and single crystalline films. Radiat. Meas. 2007, 42, 528–532. [Google Scholar] [CrossRef]
  24. De Queiroz, T.B.; Ferrari, C.R.; Ulbrich, D.; Doyle, R.; De Camargo, A.S.S. Luminescence characteristics of YAP:Ce scintillator powders and composites. Opt. Mater. 2010, 32, 1480–1484. [Google Scholar] [CrossRef]
  25. Zhuravleva, M.; Novoselov, A.; Yoshikawa, A.; Pejchal, J.; Nikl, M.; Fukuda, T. Crystal growth and scintillation properties of Pr-doped YAlO3. Opt. Mater. 2007, 30, 171–173. [Google Scholar] [CrossRef]
  26. Gieszczyk, W.; Mrozik, A.; Bilski, P.; Vistovskyy, V.; Voloshinovskii, A.; Paprocki, K.; Zorenko, T.; Zorenko, Y. Scintillation and Energy-Storage Properties of Micro-Pulling-Down Grown Crystals of Sc3+-and La3+-Doped YAlO3 Perovskite. Crystals 2020, 10, 385. [Google Scholar] [CrossRef]
  27. Rathaiah, M.; Kucera, M.; Pejchal, J.; Beitlerova, A.; Kucerkova, R.; Nikl, M. Epitaxial growth, photoluminescence and scintillation properties of Gd3+ co-doped YAlO3:Ce3+ films. Radiat. Meas. 2019, 121, 86–90. [Google Scholar] [CrossRef]
  28. Sreebunpeng, K.; Janthon, P.; Chewpraditkul, W.; Szczesniak, T.; Nikl, M.; Yoshikawa, A. Scintillation characteristics of YAlO3:Pr perovskite single crystals. Opt. Mater. 2020, 108, 110161. [Google Scholar] [CrossRef]
  29. Zhydachevskii, Y.; Durygin, A.; Suchocki, A.; Matkovskii, A.; Sugak, D.; Bilski, P.; Warchol, S. Mn-doped YAlO3 crystal: A new potential TLD phosphor. Nucl. Instrum. Methods B 2005, 227, 545–550. [Google Scholar] [CrossRef]
  30. Zhydachevskii, Y.; Durygin, A.; Suchocki, A.; Matkovskii, A.; Sugak, D.; Loutts, G.B.; Noginov, M.A. Radiation and thermally induced effects in YAlO3:Mn crystals. J. Lumin. 2004, 109, 39–49. [Google Scholar] [CrossRef]
  31. Marciniak, L.; Trejgis, K. Luminescence lifetime thermometry with Mn3+–Mn4+ co-doped nanocrystals. J. Mater. Chem. C 2018, 6, 7092–7100. [Google Scholar] [CrossRef]
  32. Zhydachevskyy, Y.; Martynyuk, N.; Popov, A.I.; Sugak, D.; Bilski, P.; Ubizskii, S.; Berkowski, M.; Suchocki, A. Thermally induced fading of Mn-doped YAP nanoceramics. J. Phys. Conf. Ser. 2018, 987, 012009. [Google Scholar] [CrossRef]
  33. Afanassyev, D.; Ubizskii, S.; Zhydachevskyy, Y.; Luchechko, A.; Popov, A.I.; Suchocki, A. Time-resolved pulsed OSL of ceramic YAP:Mn phosphors. Integr. Ferroelectr. 2019, 196, 24–31. [Google Scholar] [CrossRef]
  34. Piskunov, S.; Isakoviča, I.; Popov, A.I. Atomic structure of manganese-doped yttrium orthoaluminate. Nucl. Instrum. Methods B 2018, 434, 6–8. [Google Scholar] [CrossRef]
  35. Kück, S. Laser-related spectroscopy of ion-doped crystals for tunable solid-state lasers. Appl. Phys. B 2001, 72, 515–562. [Google Scholar] [CrossRef]
  36. Liu, B.; Wang, Q.; Xu, X.; Xu, J.; Zhu, F.; Zhou, H. Polarized spectral properties of Sm: YAlO3 single crystal for reddish-orange laser. Opt. Mater. 2020, 99, 109510. [Google Scholar] [CrossRef]
  37. Shimizu, Y.; Ueda, K. Ultraviolet emission from YAlO3:Gd3+ thin film electroluminescent devices fabricated on perovskite-type oxide substrates. Opt. Mater. 2019, 91, 371–375. [Google Scholar] [CrossRef]
  38. Brik, M.G.; Sildos, I.; Berkowski, M.; Suchocki, A. Spectroscopic and crystal field studies of YAlO3 single crystals doped with Mn ions. J. Phys. Condens. Matter 2008, 21, 025404. [Google Scholar] [CrossRef]
  39. Wang, Y.; Włodarczyk, D.; Li, L.; Wittlin, A.; Przybylińska, H.; Sybilski, P.; Zhydachevskii, Y.; Ma, C.-G.; Brik, M.G.; Malinowski, M.; et al. Electronic structure of Ce3+ in yttrium and lutetium orthoaluminate crystals and single crystal layers. J. Alloy. Compd. 2017, 723, 157–163. [Google Scholar] [CrossRef]
  40. Brik, M.G.; Sildos, I.; Kiisk, V. Calculations of physical properties of pure and doped crystals: Ab initio and semi-empirical methods in application to YAlO3:Ce3+ and TiO2. J. Lumin. 2011, 131, 396–403. [Google Scholar] [CrossRef]
  41. Brik, M.G.; Ma, C.G.; Piasecki, M.; Suchocki, A. Locating impurity and defect levels in the host band gap by first-principles calculations: Pure and Ce3+-doped YAlO3. Opt. Mater. 2021, 113, 110843. [Google Scholar] [CrossRef]
  42. Danger, T.; Petermann, K.; Schwentner, N.; Sliwinski, G.; Wong, W.C. UV-spectroscopy and band structure of Ti:YAlO3. J. Lumin. 1997, 72, 171–173. [Google Scholar] [CrossRef]
  43. Baryshnikov, V.I.; Kvapil, J.; Sarukura, N.; Segawa, Y. Especialities of Ti3+ excitation in YAlO3 crystals. J. Lumin. 1997, 72, 157–158. [Google Scholar] [CrossRef]
  44. Grunwaldt, J.D.; Gunβer, W.; Danger, T.; Petermann, K.; Reller, A. Determination of a new titanium site in Ti:YAlO3. J. Phys. Chem. Solids 1994, 55, 699–705. [Google Scholar] [CrossRef]
  45. Ando, M.; Sakaguchi, T.; Yamanaka, A.; Kawabe, Y.; Hanamura, E. Electroluminescence of Oxygen-Deficient YAlO3 Crystals with Dopants. Jpn. J. Appl. Phys. 2006, 45, 3659. [Google Scholar] [CrossRef]
  46. Lin, J.Z.; Wu, S.Y.; Fu, Q.; Zhang, H.M. Studies of the local structure and the g factors for the orthorhombic Ti3+ center in YAP. Mod. Phys. Lett. B 2007, 21, 737–743. [Google Scholar] [CrossRef]
  47. van Vuuren, A.J.; Saifulin, M.M.; Skuratov, V.A.; O’Connell, J.H.; Aralbayeva, G.; Dauletbekova, A.; Zdorovets, M. The influence of stopping power and temperature on latent track formation in YAP and YAG. Nucl. Instrum. Methods B 2019, 460, 67–73. [Google Scholar] [CrossRef]
  48. Piskunov, S.; Isakoviča, I.; Popov, A.I. Electronic structure of MnAl3+-and MnAl2+-doped YAlO3: Prediction from the first principles. Opt. Mater. 2018, 85, 162–166. [Google Scholar] [CrossRef]
  49. Piskunov, S.; Isakoviča, I.; Putnina, M.; Popov, A.I. Ab initio calculations of the electronic structure for Mn2+-doped YAlO3 crystals. Low Temp. Phys. 2020, 46, 1160–1164. [Google Scholar] [CrossRef]
  50. Ross, N.L.; Zhao, J.; Angel, R.J. High-pressure single-crystal X-ray diffraction study of YAlO3 perovskite. J. Solid State Chem. 2004, 177, 1276–1284. [Google Scholar] [CrossRef]
  51. Popov, A.I.; Kotomin, E.A.; Maier, J. Basic properties of the F-type centers in halides, oxides and perovskites. Nucl. Instrum. Methods B 2010, 268, 3084–3089. [Google Scholar] [CrossRef]
  52. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Krukau, A.V.; Vydrov, O.A.; Izmaylov, A.F.; Scuseria, G.E. Influence of the exchange screening parameter on the performance of screened hybrid functionals. J. Chem. Phys. 2006, 125, 224106. [Google Scholar] [CrossRef] [PubMed]
  54. Dovesi, R.; Saunders, V.R.; Roetti, C.; Orlando, R.; Zicovich-Wilson, C.M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N.M.; Bush, I.J.; et al. Crystal’17 User’s Manual. 2018. Available online: https://www.crystal.unito.it/manuals/crystal17.pdf (accessed on 24 August 2021).
  55. Peintinger, M.F.; Oliveira, D.V.; Bredow, T. Consistent Gaussian basis sets of triple-zeta valence with polarization quality for solid-state calculations. J. Comput. Chem. 2013, 34, 451–459. [Google Scholar] [CrossRef] [PubMed]
  56. Mulliken, R.S. Electronic population analysis on LCAO–MO molecular wave functions. I. J. Chem. Phys. 1955, 23, 1833–1840. [Google Scholar] [CrossRef] [Green Version]
  57. Mulliken, R.S. Electronic population analysis on LCAO–MO molecular wave functions. II. Overlap populations, bond orders, and covalent bond energies. J. Chem. Phys. 1955, 23, 1841–1846. [Google Scholar] [CrossRef]
  58. Ross, N.L. Distortion of GdFeO3-type perovskites with pressure: A study of YAlO3 to 5 GPa. Phase Transit. 1996, 58, 27–41. [Google Scholar] [CrossRef]
  59. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integrations. Phys. Rev. B 1976, 13, 5188. [Google Scholar] [CrossRef]
  60. Ning, L.; Cheng, W.; Zhou, C.; Duan, C.; Zhang, Y. Energetic, Optical, and Electronic Properties of Intrinsic Electron-Trapping Defects in YAlO3: A Hybrid DFT Study. J. Physiclal Chem. C 2014, 118, 19940–19947. [Google Scholar] [CrossRef]
  61. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. A 1976, 32, 751–767. [Google Scholar] [CrossRef]
  62. Nikiforov, S.V.; Kortov, V.S.; Petrov, M.O. Luminescent and dosimetric properties of ultrafine magnesium oxide ceramics after high dose irradiation. Radiat. Meas. 2016, 90, 252–256. [Google Scholar] [CrossRef]
  63. Kortov, V.S.; Milman, I.I.; Moiseykin, E.V.; Nikiforov, S.V.; Ovchinnikov, M.M. Deep-trap competition model for TL in α-Al2O3:C heating stage. Radiat. Prot. Dosim. 2006, 119, 41–44. [Google Scholar] [CrossRef]
  64. Gonzalez, R.; Monge, M.A.; Santiuste, J.E.M.; Pareja, R.; Chen, Y.; Kotomin, E.; Kukla, M.M.; Popov, A.I. Photoconversion of F-type centers in thermochemically reduced MgO single crystals. Phys. Rev. B 1999, 59, 4786–4790. [Google Scholar] [CrossRef] [Green Version]
  65. Kotomin, E.A.; Popov, A.I.; Stashans, A.A. Novel model for F+ to F photoconversion in corundum crystals. J. Phys. Condens. Matter 1994, 6, L569–L573. [Google Scholar] [CrossRef]
  66. Usseinov, A.B.; Zhukovskii, Y.F.; Kotomin, E.A.; Akilbekov, A.T.; Zdorovets, M.V.; Baubekova, G.M.; Karipbayev, Z.T. Transition levels of acceptor impurities in ZnO crystals by DFT-LCAO calculations. J. Phys. Conf. Ser. 2018, 1115, 042064. [Google Scholar] [CrossRef]
  67. Uklein, A.V.; Multian, V.V.; Kuz’micheva, G.M.; Linnik, R.P.; Lisnyak, V.V.; Popov, A.I.; Gayvoronsky, V.Y. Nonlinear optical response of bulk ZnO crystals with different content of intrinsic defects. Opt. Mater. 2018, 84, 738–747. [Google Scholar] [CrossRef]
  68. Osinkin, D.A.; Khodimchuk, A.V.; Porotnikova, N.M.; Bogdanovich, N.M.; Fetisov, A.V.; Ananyev, M.V. Rate-Determining Steps of Oxygen Surface Exchange Kinetics on Sr2Fe1.5Mo0.5O6−δ. Energies 2020, 13, 250. [Google Scholar] [CrossRef] [Green Version]
  69. Porotnikova, N.M.; Eremin, V.A.; Farlenkov, A.S.; Kurumchin, E.K.; Sherstobitova, E.A.; Kochubey, D.I.; Ananyev, M.V. Effect of AO Segregation on Catalytical Activity of La0.7A0.3MnO3±δ (A = Ca, Sr, Ba) Regarding Oxygen Reduction Reaction. Catal. Lett. 2018, 148, 2839–2847. [Google Scholar] [CrossRef]
  70. Kotomin, E.A.; Piskunov, S.; Zhukovskii, Y.F.; Eglitis, R.I.; Gopejenko, A.; Ellis, D.E. The electronic properties of an oxygen vacancy at ZrO2-terminated (001) surfaces of a cubic PbZrO3: Computer simulations from the first principles. Phys. Chem. Chem. Phys. 2008, 10, 4258–4263. [Google Scholar] [CrossRef] [PubMed]
  71. Carrasco, J.; Illas, F.; Lopez, N.; Kotomin, E.A.; Zhukovskii, Y.F.; Piskunov, S.; Maier, J.; Hermansson, K. First principles simulations of F centers in cubic SrTiO3. Phys. Status Solidi C 2005, 2, 153–158. [Google Scholar] [CrossRef]
  72. Zhumatayeva, I.Z.; Kenzhina, I.E.; Kozlovskiy, A.L.; Zdorovets, M.V. The study of the prospects for the use of Li0. 15Sr0. 85TiO3 ceramics. J. Mater. Sci. Mater. Electron. 2020, 31, 6764–6772. [Google Scholar] [CrossRef]
  73. Kozlovskiy, A.L.; Kenzhina, I.E.; Zdorovets, M.V.; Saiymova, M.; Tishkevich, D.I.; Trukhanov, S.V.; Trukhanov, A.V. Synthesis, phase composition and structural and conductive properties of ferroelectric microparticles based on ATiOx (A = Ba, Ca, Sr). Ceram. Int. 2019, 45, 17236–17242. [Google Scholar] [CrossRef]
  74. Kato, S.; Nakajima, N.; Yasui, S.; Yasuhara, S.; Fu, D.; Adachi, J.; Nitani, H.; Takeichi, Y.; Anspoks, A. Dielectric response of BaTiO3 electronic states under AC fields via microsecond time-resolved X-ray absorption spectroscopy. Acta Materialia 2021, 207, 116681. [Google Scholar] [CrossRef]
  75. Alekseev, V.; Liber, V.; Starikov, A.; Anspoks, A.; Auzins, E.; Klotins, E.; Kotleris, J. High efficiency angular deflection of the laser beam: PLZT/intracavity array. Ferroelectrics 1992, 131, 301–306. [Google Scholar] [CrossRef]
  76. Anspoks, A.; Bocharov, D.; Purans, J.; Rocca, F.; Sarakovskis, A.; Trepakov, V.; Dejneka, A.; Itoh, M. Local structure studies of SrTi16O3 and SrTi18O3. Phys. Scr. 2014, 89, 044002. [Google Scholar] [CrossRef]
  77. Rogers, E.G.; Dorenbos, P. Vacuum energy referred Ti3+/4+ donor/acceptor states in insulating and semiconducting inorganic compounds. J. Lumin. 2014, 153, 40–45. [Google Scholar] [CrossRef]
Figure 1. Equilibrium structure of 2 × 2 × 2 supercell of orthorhombic Pbnm YAlO3 with (a) F-center, (b) Ti3+ ion and (c) Ti2+ ion substituted for host Al3+ atom. The supercell contains 160 atoms in a periodically repeated unit. The Ti dopant is shown as a dark-grey ball, the compensating F-center is a small pink ball, and the Y, Al, and O sublattices are shown in light blue, brown, and red, respectively. The “b” crystallographic axis is perpendicular to the figure plane and is directed outward.
Figure 1. Equilibrium structure of 2 × 2 × 2 supercell of orthorhombic Pbnm YAlO3 with (a) F-center, (b) Ti3+ ion and (c) Ti2+ ion substituted for host Al3+ atom. The supercell contains 160 atoms in a periodically repeated unit. The Ti dopant is shown as a dark-grey ball, the compensating F-center is a small pink ball, and the Y, Al, and O sublattices are shown in light blue, brown, and red, respectively. The “b” crystallographic axis is perpendicular to the figure plane and is directed outward.
Materials 14 05589 g001
Figure 2. Equilibrium positions of oxygen octahedra in the 2 × 2 × 2 supercell structure of orthorhombic Pbnm YAlO3. (a) pristine YAlO3, (b) YAlO3 with F-center, (c) YAlO3 containing Ti3+ ion and (d) YAlO3 containing Ti2+ ion substituted for host Al3+ ion. See Table 3 for bond lengths and angles.
Figure 2. Equilibrium positions of oxygen octahedra in the 2 × 2 × 2 supercell structure of orthorhombic Pbnm YAlO3. (a) pristine YAlO3, (b) YAlO3 with F-center, (c) YAlO3 containing Ti3+ ion and (d) YAlO3 containing Ti2+ ion substituted for host Al3+ ion. See Table 3 for bond lengths and angles.
Materials 14 05589 g002
Figure 3. Electronic band structure as calculated for (a) perfect orthorhombic YAlO3 bulk crystal, (b) F-center and YAlO3, (c) Al-site Ti3+-doped YAlO3 and (d) Al-site Ti2+-doped YAlO3. Zero at the energy scale corresponds to the top of the valence band. Inset shows schematically the Brillouin zone and high symmetry points of Pbnm space group.
Figure 3. Electronic band structure as calculated for (a) perfect orthorhombic YAlO3 bulk crystal, (b) F-center and YAlO3, (c) Al-site Ti3+-doped YAlO3 and (d) Al-site Ti2+-doped YAlO3. Zero at the energy scale corresponds to the top of the valence band. Inset shows schematically the Brillouin zone and high symmetry points of Pbnm space group.
Materials 14 05589 g003
Figure 4. Equilibrium positions of oxygen octahedra in the 2 × 2 × 2 supercell structure of orthorhombic Pbnm YAlO3. (a) pristine YAlO3, (b) YAlO3 with F-center, (c) YAlO3 containing Ti3+ ion and (d) YAlO3 containing Ti2+ ion substituted for host Al3+ ion. See Table 3 for bond lengths and angles.
Figure 4. Equilibrium positions of oxygen octahedra in the 2 × 2 × 2 supercell structure of orthorhombic Pbnm YAlO3. (a) pristine YAlO3, (b) YAlO3 with F-center, (c) YAlO3 containing Ti3+ ion and (d) YAlO3 containing Ti2+ ion substituted for host Al3+ ion. See Table 3 for bond lengths and angles.
Materials 14 05589 g004
Table 2. Structural characteristics of YAlO3 crystal in orthorhombically distorted Pbnm perovskite structure.
Table 2. Structural characteristics of YAlO3 crystal in orthorhombically distorted Pbnm perovskite structure.
Lattice Constants and VolumeExperiment, Ref. [50]Calc. (HSE06), Ref [60]This Study
a0, Å5.1795.1795.189
b0, Å5.3295.3425.317
c0, Å7.3707.3677.388
V, Å3203.49203.82203.86
Fractional Coordinates (In Units of the Lattice Constants)
Atoms:xyzxyzxyz
Y−0.0120.0530.25−0.0120.0550.25−0.0110.0500.25
Al00.5000.5000.50
O10.0840.4770.250.0840.4780.250.0820.4790.25
O20.7050.2950.0440.7060.2940.0420.7080.2920.044
Table 3. Atomic structure of pristine and Ti-doped YAlO3 crystals in orthorhombically distorted perovskite structure. Bond lengths l are given in Å and bond angles α are in degrees. For atomic notations please see Figure 2.
Table 3. Atomic structure of pristine and Ti-doped YAlO3 crystals in orthorhombically distorted perovskite structure. Bond lengths l are given in Å and bond angles α are in degrees. For atomic notations please see Figure 2.
YAlO3YAlO3:FYAlO3:Ti3+YAlO3:Ti2+
lAl,Ti-O_I1.921.922.011.94
lAl,Ti-O_II1.921.922.011.96
lAl,Ti-O_III1.901.901.992.00
lAl,Ti-O_IV1.901.901.992.00
lAl,Ti-O_V,F1.901.221.971.21
lAl,Ti-O_VI1.901.921.972.08
lAl,Ti-Y_I3.022.983.022.98
lAl,Ti-Y_II3.023.043.023.10
lAl,Ti-Y_III3.153.123.173.15
lAl,Ti-Y_IV3.153.163.173.25
lAl,Ti-Y_V3.243.173.303.14
lAl,Ti-Y_VI3.243.263.303.33
α0_I-Al,Ti-O_II180.0178.6180.0175.4
α0_III-Al,Ti-O_IV180.0178.5180.0172.8
α0_V,F-Al,Ti-O_VI180.0174.4180.0174.7
αY_I-Al,Ti-Y_II180.0178.9180.0175.4
αY_III-Al,Ti-Y_IV180.0179.0180.0175.8
αY_V-Al,Ti-Y_VI180.0179.4180.0176.8
Table 4. Mulliken population analysis of pristine and Ti-doped YAlO3 crystals in orthorhombically distorted perovskite structure. Q is Mulliken effective charge in e, bond populations P are in e. Negative bond population means electronic repulsion. For atomic notations please see Figure 2.
Table 4. Mulliken population analysis of pristine and Ti-doped YAlO3 crystals in orthorhombically distorted perovskite structure. Q is Mulliken effective charge in e, bond populations P are in e. Negative bond population means electronic repulsion. For atomic notations please see Figure 2.
YAlO3YAlO3:FYAlO3:Ti3+YAlO3:Ti2+
QY_I2.462.422.452.40
QY_II2.462.462.452.46
QY_III2.462.452.452.45
QY_IV2.462.462.452.46
QY_V2.462.452.462.43
QY_VI2.462.462.462.46
QAl,Ti1.771.261.561.07
QO_I−1.42−1.44−1.39−1.35
QO_II−1.42−1.43−1.39−1.37
QO_III−1.40−1.43−1.37−1.38
QO_IV−1.40−1.44−1.37−1.39
QO_V,F−1.40−0.53−1.35−0.66
QO_VI−1.40−1.41−1.35−1.38
PAl,Ti-O_I0.280.380.160.11
PAl,Ti-O_II0.280.370.160.15
PAl,Ti-O_III0.270.340.150.05
PAl,Ti-O_IV0.270.350.150.09
PAl,Ti-O_V,F0.27−0.620.16−1.44
PAl,Ti-O_VI0.270.180.16−0.24
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Piskunov, S.; Gopejenko, A.; Pankratov, V.; Isakoviča, I.; Ma, C.-G.; Brik, M.G.; Piasecki, M.; Popov, A.I. First Principles Calculations of Atomic and Electronic Structure of TiAl3+- and TiAl2+-Doped YAlO3. Materials 2021, 14, 5589. https://doi.org/10.3390/ma14195589

AMA Style

Piskunov S, Gopejenko A, Pankratov V, Isakoviča I, Ma C-G, Brik MG, Piasecki M, Popov AI. First Principles Calculations of Atomic and Electronic Structure of TiAl3+- and TiAl2+-Doped YAlO3. Materials. 2021; 14(19):5589. https://doi.org/10.3390/ma14195589

Chicago/Turabian Style

Piskunov, Sergei, Aleksejs Gopejenko, Vladimir Pankratov, Inta Isakoviča, Chong-Geng Ma, Mikhail G. Brik, Michal Piasecki, and Anatoli I. Popov. 2021. "First Principles Calculations of Atomic and Electronic Structure of TiAl3+- and TiAl2+-Doped YAlO3" Materials 14, no. 19: 5589. https://doi.org/10.3390/ma14195589

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