Next Article in Journal
Modeling the Effect of Magnesium Content on the Electrical Conductivity and Hardness of Technical Aluminum (1050) Alloy
Previous Article in Journal
Optical Temperature Sensor Capabilities of the Green Upconverted Luminescence of Er3+ in La3NbO7 Ceramic Powders
Previous Article in Special Issue
Studies of Optical, Dielectric, Ferroelectric, and Structural Phase Transitions in 0.9[KNbO3]-0.1 [BaNi1/2Nb1/2O3−δ]
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Second-Order Raman Scattering in Ferroelectric Ceramic Solid Solutions LiNbxTa1−xO3

1
Tananaev Institute of Chemistry—Subdivision of the Federal Research Centre, Kola Science Centre of the Russian Academy of Sciences, 184206 Apatity City, Russia
2
P. N. Lebedev Physical Institute of the Russian Academy of Sciences, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(4), 456; https://doi.org/10.3390/cryst12040456
Submission received: 25 February 2022 / Revised: 21 March 2022 / Accepted: 22 March 2022 / Published: 24 March 2022

Abstract

:
In the second-order Raman spectra of ceramic solid solutions, LiNbxTa1−xO3 weak overtone bands of fully symmetric fundamental polar excitations were observed for the first time. The frequencies of the two bands exceeded the value of the overtone frequency corresponding to the fully symmetrical vibration 4A1(z). The possibility of the existence of phonon bound states of the antipolar type in the vibrational spectrum of LiNbxTa1−xO3 ceramics is predicted.

1. Introduction

Lithium niobate (LiNbO3) and tantalate (LiTaO3) are among the most studied oxide compounds in modern materials science. This interest is due to the wide applications of these materials in functional electronics. The forms in which these materials are used in practice are different—single- or polycrystals, thin films, micro- and nanopowders, and ceramics. Moreover, they can have different chemical compositions—congruent or stoichiometric, and contain various dopants of metal ions. It is obvious that such a variety of forms and applications requires additional research to modify and optimize material properties.
LiNb1−xTaxO3 (LNT) solid solutions have recently been studied (see, for example, [1,2,3]), while they open up prospects for combining the advantages of both materials. In particular, the LNT can be expected to exhibit high piezoelectric coefficients (close to LiNbO3) and the temperature stability of the LiTaO3 properties. However, the growth of LNT single crystals is a difficult task due to deviation from stoichiometry inherent in LiNbO3 and LiTaO3 crystals, nonhomogeneous distribution of cations caused by fluctuations in the temperature fields in the crystal growth zone and differences in the melting temperatures of both components of the system.
In addition, a LNT complex compound is attractive for both practical applications and fundamental researches, for example, in determining dependence of the LiNb1−xTaxO3 structural characteristics on the composition.
The system of ferroelectric solid solutions LiNbxTa1−xO3 is a subject of intensive research by different methods because it may be the base of ceramic materials with ferroelectric, superionic, semiconductor properties and cross-effects. The components of the LiNbxTa1−xO3 system are phases of variable composition and are characterized by a strongly defective crystal structure of both the cationic and anionic sublattices. The structure of defects largely forms the physical (especially ferroelectric and nonlinear optical) characteristics of materials.
Raman spectroscopy plays a significant role in the study of crystalline systems with a disordered structure and the processes of disordering of their structural units. The Raman spectra are highly sensitive to changes in the interactions between the structural units of a crystal and, consequently, to various rearrangements of the crystal structure, which occur, in particular, upon changes in its composition and temperature, as well as to anharmonicity effects. In this case, it is important to study simultaneously the Raman spectra of the first and second order, which are especially sensitive to various fine processes of disordering of the crystal structure.
The first-order Raman spectra of single crystals and ceramic solid solutions with the general formula LiNbxTa1−xO3 have been studied in a number of works [4,5,6,7,8,9,10,11,12,13,14,15,16,17]. In the article [4], the change in the local composition of the LiNb1−xTaxO3 single crystal was analyzed by the Raman spectroscopy. In articles [5,6,7,8,9,10,11,12], the Raman spectra of such crystals were analyzed at various values of the niobium concentration x = 0–1. The concentration dependences of the frequencies of the fundamental modes of A1(z)- and E(x,y)-symmetries were studied. In [13,14], the temperature dependence of the Raman spectra of a LiTa0.9Nb0.1O3 single crystal was studied. The presence of a second-order ferroelectric–paraelectric phase transition of the order–disorder type was established at temperatures above 900 K. In this case, a "relay" transfer of softness from the fundamental optical phonon of A1(z) symmetry to a non-fundamental vibration with a frequency of 120 cm−1 was observed. The article [15] is devoted to the numerical calculation from first principles (ab initio) of the frequencies of the fundamental modes of the A1(z)-symmetry of a LiNbxTa1−xO3 mixed crystal. The presence of a linear increase in the calculated values of frequencies during the transition from lithium tantalate to lithium niobate is shown. At the same time, the experimental data obtained in [12] reveal a weak nonlinear dependence of the frequencies in the Raman spectra of LiNbxTa1−xO3 as a function of x. The Raman spectra of LiNb0.5Ta0.5O3:xPr nanocrystals embedded in a boron-based glass matrix are reported in [16]. An increase in the concentration x of praseodymium from 0 to 0.01 leads to an insignificant frequency shift of the two lines and the appearance of one new Raman line. A study of the concentration dependence of the Raman spectra of LiNbxTa1−xO3 nanocrystals 19–45 nm in size introduced into a boron-based glass matrix was performed in [17]. An increase in the niobium concentration led to a shift of some Raman lines to the high-frequency region. Thus, in papers [4,5,6,7,8,9,10,11,12,13,14,15,16,17], the Raman spectra of the first order (0–1200 cm−1) of ceramic solid solutions of the LiNbxTa1−xO3 system were analyzed upon excitation (or relaxation) of single-particle states of polar modes.
In [18,19], the manifestation of overtone bands in the Raman spectra of diamond and LiNbO3:Tb crystals was reported for the first time in the spectral range 1300–2700 cm−1. In the second-order Raman spectra studied earlier in a number of crystals [20,21,22,23], pairs of phonons with oppositely directed quasi-momenta, corresponding to the boundary or inner region of the Brillouin zone, appeared. Previously, in the second-order Raman spectra of a number of crystals, manifestations of phonon pairs were detected, which were interpreted as anomalies in the density of bound and free two-phonon states.
In this paper, we set the task of studying the full Raman spectrum of polar modes in LiNbxTa1−xO3 ceramic solid solutions excited in the backscattering geometry in a wide spectral region, including both the range of fundamental vibrations (150–1000 cm−1) and the region of the second-order spectrum (100–150 and 1000–2000 cm−1).

2. Materials and Methods

Samples of LiNbxTa1−xO3 ceramic solid solutions in the form of pellets up to 10 mm in diameter and 1 mm thick were prepared by standard ceramic technology by high-temperature solid-phase synthesis [12], Figure 1. Table 1 shows the structural characteristics of the crystal lattices of the samples under study.
The setup for recording of Raman spectra at room temperature using a backscattering geometry with BWS465-785H spectrometer was presented in [18,19]. Raman spectra were excited by 785 nm cw semiconductor laser with a 200 mW power. Laser radiation was delivered to the sample surface using one of the channels of a two-wire optical fiber, and the Raman signal was delivered through the second channel to a selective light filter and then to the input of the spectrometer. We used a BWS465-785H spectrometer with a detection range of 50–2850 cm−1.

3. Results and Discussion

Figure 2 shows the Raman spectrum of the LiNb0.99Ta0.01O3 ceramic solid solution in the frequency range 50–2000 cm−1 recorded in the backscattering geometry.
As can be seen from Figure 2, the presented Raman spectrum of a ceramic solid solution with a small addition of Ta contains mainly fundamental modes corresponding to first-order Raman processes [24,25,26,27] in the range of 50–1000 cm−1. It should be noted that in the presented Raman spectrum, as well as in the spectra of a nominally pure stoichiometric or congruent LiNbO3 crystal [12,24,25,26,27,28,29], there are no second-order bands in the region above 1000 cm−1.
Figure 3a–c shows the recorded Raman spectra of the three studied samples LiNb0.4Ta0.6O3, LiNb0.5Ta0.5O3 and LiNb0.6Ta0.4O3 obtained at two different points of each sample in the backscattering geometry. In these figures, curves (1) and (2) correspond to different points of the studied samples.
As can be seen from Figure 3, in the region of the Raman spectra (100–1000 cm−1), where fundamental vibrations of the crystal lattice are manifested, a number of lines of different intensity are observed. It can be seen from this figure that in the Raman spectra of ceramic solid solutions LiNbxTa1−xO3 at approximately the same concentration of Nb and Ta (x ≈ 0.5) in the overtone region of the Raman spectra (1000–2000 cm−1) several second-order bands are present. The frequencies of the two bands are significantly higher than the value of the overtone (1736 cm−1) corresponding to the fully symmetrical vibration of 868 cm−1. In this case, the relative intensity of the two highest frequency bands (in the region of 1700–2000 cm−1) is the lowest for the LiNb0.5Ta0.5O3 sample (Figure 3b). It should be noted that exactly in this concentration range (0.2 < y < 0.8), according to the data of [12], a strong disordering of the solid solution structure (maximum at y ≈ 0.5) and a nonlinear dependence of the main characteristics (frequency, width, intensity and shape parameter) of the spectral lines takes place. The observed nonlinearity in the behavior of the main parameters of the spectral lines allows us to conclude that the transformation of the structure of LiNbO3 ceramics into the structure of LiTaO3 ceramics in the system of solid solutions LiNbxTa1−xO3 with decreasing x occurs continuously through the region of increased structural disorder (0.2 < y < 0.8) with the maximum value of the unit cell deformation at y ≈ 0.5. No structural phase transition is observed in this case.
Table 2 shows the assignment of the Raman lines recorded in this work by symmetry types based on polarization measurements [7,28,29,30,31,32] performed for single-crystal samples.
As can be seen from this table, the assignment was not performed for all registered first-order Raman lines. In particular, in one of the first works on Raman scattering in a single-domain lithium niobate crystal, the 175 cm−1 line was assigned to the mixed mode [27]. Moreover, in [28,29], this line refers to the fundamental 2E(x,y) mode of stoichiometric lithium niobate. In [30], this line is also assigned to the fundamental 2E(x,y) mode based on theoretical calculations and experimental results [31] for stoichiometric 7LiNbO3 and 6LiNbO3 single crystals. The lines near 820 cm−1, which appear in the region of oxygen carcass vibrations, apparently correspond to the Li3NbO4 phase. Such a phase can manifest itself due to the decomposition of a phase with a variable composition of lithium metaniobate [32].
Figure 4 shows the concentration dependences of the observed second-order Raman lines of LNT ceramic solid solutions. As can be seen from this figure, the change in the frequencies of the second-order Raman lines has a multidirectional character.
The scheme for the formation of bound states of phonons in crystals was considered in theoretical works [18,19,33,34,35]. In accordance with these works, the density of two-phonon states ρ2(ω) is calculated using one-phonon Green’s functions D1(k,ω) according to Formulas (1)–(5):
D 1 k , ω = ω k 2 1 ω ω k + 1 2 i Γ 1 ω + ω k 1 2 i Γ
F ω = i 2 π 4 d 3 k D 1 k , ω ω 1 D 1 k , ω 1 d ω 1 .  
Here, k -phonon wave vector, ω-frequency in rad/s, ω( k )-dispersion dependence of the optical phonon frequency on the wave vector, i-imaginary unit, Γ-phonon damping constant and ω1 is the integration variable.
The density of single-phonon states ρ1(ω1) in the parabolic approximation of the optical phonon dispersion law near the center of the Brillouin zone ωω0 has the form:
ρ 1 ω 1 = b ω 0 ω 1
In this expression, the frequencies ω1 of phonons of the optical branch of the 4A1(z) -phonon can take values from ω1 = ω0 − ∆ to ω1 = ω0; b is normalizations coefficient and ∆ is the integration range of ω1 taken into account in Equation (2). For the function F(ω), we get:
F ω = 1 4 ω 0 2 b 0 Δ x ω 2 ω 0 x + i Γ d x
The second-order Raman intensity is proportional to the density of two-phonon states ρ2(ω):
ρ 2 ω = 1 π ω 0 2 I m 2 F ω 1 Λ 4 F ω
here Λ4 is the fourth-order anharmonicity parameter of the crystal lattice.
Let us introduce a dimensionless variable u:
u = (ω0ω1)/ω0
For definiteness, we require the fulfillment of the normalization condition:
0 d ρ 1 u d u = 1
From here, taking into account (3), with the relative width of the integration range d = ∆/ω0, we obtain
b = 3 2 d 3
Let us introduce dimensionless function f(ω) taking into account (4, 6–8):
f ω = b 0 d u 2 u + a d u = 3 2 d 3 d a 2 a r c t g 2 d a
here a = (ω −2ω0 + )/ω0, γ = Γ/ω0–phonon damping coefficient. Accordingly, relation (5) up to a constant factor takes the form:
ρ 2 ω = I m 2 f ω 1 λ 4 f ω
where λ4 is the dimensionless anharmonicity parameter.
Let us consider, for definiteness, the overtone state of the fully symmetric mode 4A1(z) with the frequency ν0 = 868 cm−1 (ω0 = 2πcν0 = 1.64 × 1014 rad/s). The results of a comparison between the calculated Raman spectrum in the overtone region 2ν0 = 1736 cm−1 and the experimentally observed one for the LiNb0.6Ta0.4O3 ceramic solid solution are shown in Figure 5. The choice of this sample is due to the highest observed relative intensity of the second-order Raman scattering. The dashed arrow in this figure indicates the exact position of the 2ν0 overtone.
We can see from this figure, that the band at 1772 cm−1 (curve 2 in Figure 5), is well approximated by the bound state of the polar mode 4A1(z) at ν0 = 868 cm−1 with the following parameters: d = 0.05, γ = 0.03 and λ4 = 0.07. For the 1924 cm−1 band (curve 3 in Figure 5), satisfactory agreement is achieved at d = 0.05, γ = 0.03 and λ4 = 0.25. The integration region d in Formula (9) is chosen in accordance with the frequency range occupied by the 4A1(z) branch on the phonon dispersion curve ω(k).

4. Conclusions

It was found for the first time that in the Raman spectra of ferroelectric ceramic solid solutions of lithium niobate-tantalate with different [Nb]/[Ta] concentrations, there are second-order bands in the overtone region (1000–2000 cm−1) of fundamental polar modes, whereas the Raman spectrum of the LiNb0.99Ta0.01O3 ceramic solid solution in the range of 1000–2000 cm−1 contains only a weak continuous background for all recorded modes. We have found two bands with frequencies that exceed the exact value of the overtone (868 × 2 = 1736 cm−1) of the 4A1(z) mode. Theoretical modeling performed testifies that these bands are due to biphonons. Further studies of biphonons are of interest for establishing the conditions for the generation of entangled and bound states of two phonons in spontaneous and stimulated Raman scattering in ferroelectric and piezoelectric crystals and ceramics.

Author Contributions

Conceptualization, N.S. and P.S.; methodology, N.S., M.P. and P.S.; software, P.S.; formal analysis, P.S.; investigation, A.P.; resources, N.S., M.P.; data curation, A.P.; writing—original draft preparation, P.S., A.P., N.S.; writing—review and editing, M.P.; visualization, A.P.; supervision, N.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Ministry of Science and Higher Education Russian Federation scientific topic No FMEZ-2022-0016 and RFBR grant No. 20-52-04001 Bel_mol_a.

Data Availability Statement

The raw data required to reproduce these findings are available from corresponding author N.S. on a reasonable request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Xue, D.; Betzler, K.; Hesse, H. Dielectric properties of lithium niobate-tantalate crystals. Solid State Commun. 2000, 115, 581–585. [Google Scholar] [CrossRef]
  2. Wood, I.G.; Daniels, P.; Brown, R.H.; Glazer, A.M. Optical birefringence study of the ferroelectric phase transition in lithium niobate tantalate mixed crystals: LiNb1-xTaxO3. J. Phys. Condens. Matter 2008, 20, 235237. [Google Scholar] [CrossRef]
  3. Roshchupkin, D.; Emelin, E.; Plotitcyna, O.; Rashid, F.; Irzhak, D.; Karandashev, V.; Orlova, T.; Targonskaya, N.; Sakharov, S.; Mololkin, A.; et al. Single crystals of ferroelectric lithium niobate–tantalate LiNb1–xTaxO3 solid solutions for high-temperature sensor and actuator applications. Acta Crystallogr. B 2020, 76, 1071–1076. [Google Scholar] [CrossRef]
  4. Irzhak, A.; Irzhak, D.; Khvostikov, V.; Pundikov, K.; Roshchupkin, D.; Fahrtdinov, R. Effect of local changes in the composition of the LiNb1-xTaxO3 single crystal on the Raman spectra. J. Raman Spectrosc. 2022, in press. [CrossRef]
  5. Vasylechko, L.; Sydorchuk, V.; Lakhnik, A.; Suhak, Y.; Wlodarczyk, D.; Hurskyy, S.; Yakhnevych, U.; Zhydachevskyy, Y.; Sugak, D.; Syvorotka, I.; et al. Investigations of LiNb1−xTaxO3 nanopowders obtained with mechanochemical method. Crystals 2021, 11, 755. [Google Scholar] [CrossRef]
  6. Buryy, O.; Vasylechko, L.; Sydorchuk, V.; Lakhnik, A.; Suhak, Y.; Wlodarczyk, D.; Hurskyj, S.; Yakhnevych, U.; Zhydachevskyy, Y.; Sugak, D.; et al. Crystal structure, Raman spectra and electrical conductivity of LiNb1–xTaxO3 nanopowders obtained with high-energy ball milling. J. Nano-Electron. Phys. 2021, 13, 02038. [Google Scholar] [CrossRef]
  7. Bartasyte, A.; Margueron, S.; Glazer, A.M.; Simon, E.; Gregora, I.; Huband, S.; Thomas, P.A. Vibrational modes and overlap matrix of LiNb1−xTaxO3 mixed crystals. Phys. Rev. B 2019, 99, 094306. [Google Scholar] [CrossRef]
  8. Rüsing, M.; Sanna, S.; Neufeld, S.; Berth, G.; Schmidt, W.G.; Zrenner, A.; Yu, H.; Wang, Y.; Zhang, H. Vibrational properties of LiNb1-xTaxO3 mixed crystals. Phys. Rev. B 2016, 93, 184305. [Google Scholar] [CrossRef]
  9. Bartasyte, A.; Glazer, A.; Wondre, F.; Prabhakaran, D.; Thomas, P.; Huband, S.; Keeble, D.; Margueron, S. Growth of LiNb1−xTaxO3 solid solution crystals. Mater. Chem. Phys. 2012, 134, 728–735. [Google Scholar] [CrossRef]
  10. Sidorov, N.; Palatnikov, M.; Serebryakov, Y.; Rogovoi, V.; Melnik, N. Structural ordering and ferroelectric properties of LiTaxNb1-xO3 solid solutions. Ferroelectrics 1996, 188, 31–40. [Google Scholar] [CrossRef]
  11. Serebryakov, Y.; Sidorov, N.; Palatnikov, M.; Lebold, V.; Savchenko, Y.; Kalinnikov, V. The influence of Mg2+, Gd3+ and Ta5+ admixtures on cation structural ordering in lithium niobate single crystals. Ferroelecrrics 1995, 167, 181–189. [Google Scholar] [CrossRef]
  12. Sidorov, N.V.; Serebryakov, Y.A.; Palatnikov, M.N.; Mel’nik, N.N. Characteristic features of the structural ordering and ferroelectric properties of the solid solutions LiTaxNb1-xO3. Phys. Solid State 1995, 37, 1910–1915. [Google Scholar]
  13. Ge, Y.-C.; Zhao, C.-Z. Raman study of the phase transition in LiTa0.9Nb0.1O3 single crystal. J. Raman Spectrosc. 1995, 26, 975–979. [Google Scholar] [CrossRef]
  14. Ge, Y.-C.; Zhao, C.-Z. Raman spectra study on LiTa0.9Nb0.1O3 single crystal. Spectrosc. Lett. 1995, 28, 451–458. [Google Scholar] [CrossRef]
  15. Sanna, S.; Riefer, A.; Neufeld, S.; Schmidt, W.G.; Berth, G.; Rüsing, M.; Widhalm, A.; Zrenner, A. Vibrational fingerprints of LiNbO3-LiTaO3 mixed crystals. Ferroelectrics 2013, 447, 63–68. [Google Scholar] [CrossRef]
  16. Jaschin, P.W.; Varma, K.B.R. Enhanced second harmonic generation and photoluminescence in Pr-doped LiNb0.5Ta0.5O3 nanocrystals embedded in a borate based glass. J. Appl. Phys. 2017, 122, 083107. [Google Scholar] [CrossRef]
  17. Jaschin, P.W.; Varma, K.B.R. Structural evolution and second harmonic properties of lithium niobate–tantalate nanocrystals embedded in a borate glass. J. Non-Cryst. Solids 2016, 434, 41–52. [Google Scholar] [CrossRef]
  18. Gorelik, V.S.; Pyatyshev, A.Y. Raman scattering in diamond nano- and microcrystals, synthesized at high temperatures and high pressures. Diam. Relat. Mater. 2020, 110, 108104. [Google Scholar] [CrossRef]
  19. Sidorov, N.V.; Palatnikov, M.N.; Gorelik, V.S.; Sverbil, P.P. Second-order Raman spectra of a LiNbO3:Tb crystal. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 266, 120445. [Google Scholar] [CrossRef]
  20. Anik’ev, A.A.; Umarov, B.S.; Scott, J.F. Analysis of two-phonon resonances in and Raman spectroscopy of ammonium chloride. J. Raman Spectrosc. 1985, 16, 315–318. [Google Scholar] [CrossRef]
  21. Anik’ev, A.A.; Gorelik, V.S.; Umarov, B.S. Effects of resonance interaction between a soft mode and two-phonon excitations in Raman spectra of quartz. Sov. Phys. Solid State 1984, 26, 1679–1682. [Google Scholar]
  22. Sushchinsky, M.M.; Gorelik, V.S.; Maximov, O.P. Higher-order Raman spectra of GaP. J. Raman Spectrosc. 1978, 7, 26–30. [Google Scholar] [CrossRef]
  23. Gorelik, V.; Maximov, O.; Mitin, G.; Sushchinsky, M.; Lebedev, P. Bound and many-particle states in polariton Raman spectra of NH4Cl crystals. Solid State Commun. 1977, 21, 615–619. [Google Scholar] [CrossRef]
  24. Nogueira, B.A.; Milani, A.; Castiglioni, C.; Fausto, R. The correlation between experimental polarized Raman spectra and their density functional theory prediction in the LCAO framework: The R3c LiNbO3 crystal as a test case. J. Raman Spectrosc. 2021, 52, 995–1010. [Google Scholar] [CrossRef]
  25. Margueron, S.; Bartasyte, A.; Glazer, A.M.; Simon, E.; Hlinka, J.; Gregora, I.; Gleize, J. Resolved E-symmetry zone-centre phonons in LiTaO3 and LiNbO3. J. Appl. Phys. 2012, 111, 104105. [Google Scholar] [CrossRef]
  26. Sidorov, N.; Palatnikov, M.; Kadetova, A. Raman scattering in non-stoichiometric lithium niobate crystals with a low photorefractive effect. Crystals 2019, 9, 535. [Google Scholar] [CrossRef] [Green Version]
  27. Barker, A.S., Jr.; Loudon, R. Dielectric properties and optical phonons in LiNbO3. Phys. Rev. 1967, 158, 433–445. [Google Scholar] [CrossRef]
  28. Ridah, A.; Bourson, P.; Fontana, M.D.; Malovichko, G. The composition dependence of the Raman spectrum and new assignment of the phonons in LiNbO3. J. Phys. Condens. Matter 1997, 9, 9687–9693. [Google Scholar] [CrossRef]
  29. Sidorov, N.; Yanichev, A.; Palatnikov, M.; Manukovskaya, D. Comparative investigation of LiNbO3 crystals Raman spectra in the temperature range 100–400 K. Vib. Spectrosc. 2018, 96, 19–25. [Google Scholar] [CrossRef]
  30. Caciuc, V.; Postnikov, A.V.; Borstel, G. Ab initio structure and zone-center phonons in LiNbO3. Phys. Rev. B 2000, 61, 8806–8813. [Google Scholar] [CrossRef]
  31. Repelin, Y.; Husson, E.; Bennani, F.; Proust, C. Raman spectroscopy of lithium niobate and lithium tantalate. Force field calculations. J. Phys. Chem. Solids 1999, 60, 819–825. [Google Scholar] [CrossRef]
  32. Svaasand, L.O.; Eriksrud, M.; Nakken, G.; Grande, A.P. Solid-solution range of LiNbO3. J. Cryst. Growth 1974, 22, 230–232. [Google Scholar] [CrossRef]
  33. Ruvalds, J.; Zawadowski, A. Two-phonon resonances and hybridization of the resonance with single-phonon states. Phys. Rev. B 1970, 2, 1172–1175. [Google Scholar] [CrossRef]
  34. Ruvalds, J.; Zawadowski, A. Indirect coupling and antiresonance of two optic phonons. Phys. Rev. Lett. 1970, 24, 1111–1114. [Google Scholar] [CrossRef]
  35. Ruvalds, J.; Zawadowski, A. Resonances of two phonons from different dispersion branches. Solid State Commun. 1971, 9, 129–132. [Google Scholar] [CrossRef]
Figure 1. Picture of studied samples.
Figure 1. Picture of studied samples.
Crystals 12 00456 g001
Figure 2. Raman spectrum of the LiNb0.99Ta0.01O3 ceramic solid solution for the backscattering geometry.
Figure 2. Raman spectrum of the LiNb0.99Ta0.01O3 ceramic solid solution for the backscattering geometry.
Crystals 12 00456 g002
Figure 3. Raman spectra at the backscattering geometry in the region of fundamental and overtone vibrations of LiNb0.4Ta0.6O3 (a), LiNb0.5Ta0.5O3 (b) and LiNb0.6Ta0.4O3 (c) ceramic samples. Curves (1) and (2) correspond to two different points of the corresponding sample.
Figure 3. Raman spectra at the backscattering geometry in the region of fundamental and overtone vibrations of LiNb0.4Ta0.6O3 (a), LiNb0.5Ta0.5O3 (b) and LiNb0.6Ta0.4O3 (c) ceramic samples. Curves (1) and (2) correspond to two different points of the corresponding sample.
Crystals 12 00456 g003aCrystals 12 00456 g003b
Figure 4. Dependence of second-order Raman lines frequency on the composition of LNT ceramic solid solutions. Lines (1–9) correspond to the frequencies of Raman lines marked in Figure 3a–c in the region of 1000–2000 cm−1: (1)-1034, (2)-1111, (3)-1178, (4)-1245, (5)-1334, (6)-1456, (7)-1591, (8)-1771 and (9)-1920 cm−1.
Figure 4. Dependence of second-order Raman lines frequency on the composition of LNT ceramic solid solutions. Lines (1–9) correspond to the frequencies of Raman lines marked in Figure 3a–c in the region of 1000–2000 cm−1: (1)-1034, (2)-1111, (3)-1178, (4)-1245, (5)-1334, (6)-1456, (7)-1591, (8)-1771 and (9)-1920 cm−1.
Crystals 12 00456 g004
Figure 5. Comparison of the spectral Raman intensity of the LiNb0.6Ta0.4O3 ceramic solid solution (curve 1) in the overtone region with the calculated ρ2(ν) dependences (2) and (3).
Figure 5. Comparison of the spectral Raman intensity of the LiNb0.6Ta0.4O3 ceramic solid solution (curve 1) in the overtone region with the calculated ρ2(ν) dependences (2) and (3).
Crystals 12 00456 g005
Table 1. Structural characteristics of the crystal lattices of the studied ceramic solid solutions [8].
Table 1. Structural characteristics of the crystal lattices of the studied ceramic solid solutions [8].
Samplea, Åc, Å
LiNb0.4Ta0.6O35.2113.98
LiNb0.5Ta0.5O35.1513.8
LiNb0.6Ta0.4O35.2214.04
Table 2. Frequencies of polar modes of LNT ceramic solid solutions registered in this work and their assignment taking into account the results of [7,28,29,30,31,32].
Table 2. Frequencies of polar modes of LNT ceramic solid solutions registered in this work and their assignment taking into account the results of [7,28,29,30,31,32].
ν, cm−1Assignment
LiNb0.4Ta0.6O3LiNb0.5Ta0.5O3LiNb0.6Ta0.4O3
142146 (148)144 (146)1E(x,y)
175 (177)169 (171)169 (175)2E(x,y) [28,29,30,31]
2192262221A1(z)
292 (294)294 (296)292 (294)3E(x,y)
3173183184E(x,y)
3483433442A1(z); 3A1(z)
378374375 (376)6E(x,y)
453 (455)445 (447)4477E(x,y)
603610 (612)6084A1(z)
662662657 (662)8E(x,y); 9E(x,y)
811 (815)815 (820)818 (820)Li3NbO4 phase [32]
865 (866)8688684A1(z)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sidorov, N.; Palatnikov, M.; Pyatyshev, A.; Sverbil, P. Second-Order Raman Scattering in Ferroelectric Ceramic Solid Solutions LiNbxTa1−xO3. Crystals 2022, 12, 456. https://doi.org/10.3390/cryst12040456

AMA Style

Sidorov N, Palatnikov M, Pyatyshev A, Sverbil P. Second-Order Raman Scattering in Ferroelectric Ceramic Solid Solutions LiNbxTa1−xO3. Crystals. 2022; 12(4):456. https://doi.org/10.3390/cryst12040456

Chicago/Turabian Style

Sidorov, Nikolay, Mikhail Palatnikov, Alexander Pyatyshev, and Pavel Sverbil. 2022. "Second-Order Raman Scattering in Ferroelectric Ceramic Solid Solutions LiNbxTa1−xO3" Crystals 12, no. 4: 456. https://doi.org/10.3390/cryst12040456

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