Next Article in Journal
Green Synthesis of Ni-Cu-Zn Based Nanosized Metal Oxides for Photocatalytic and Sensor Applications
Next Article in Special Issue
Electrical Conductivity and Dielectric Relaxation in Ag1−xLixNbO3
Previous Article in Journal
Study on Micro-Structure and Tensile Mechanical Properties of Dissimilar Metal Weld Joint Connecting Steam Generator Nozzle and Safe-End
Previous Article in Special Issue
Structural Properties and Dielectric Hysteresis of Molecular Organic Ferroelectric Grown from Different Solvents
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Raman Response of Quantum Critical Ferroelectric Pb-Doped SrTiO3

by
Ekaterina D. Linnik
1,2,
Alexey S. Mikheykin
1,
Diego Rubi
3,
Vladimir B. Shirokov
4,
Daoud Mezzane
2,5,
Svitlana V. Kondovych
2,
Igor A. Lukyanchuk
1,2 and
Anna G. Razumnaya
1,*
1
Faculty of Physics, Southern Federal University, 344090 Rostov-on-Don, Russia
2
Laboratoire de Physique de la Matière Condensée, Université de Picardie Jules Verne, 80080 Amiens, France
3
Instituto de Nanociencia y Nanotecnología, CONICET-CNEA, San Martín 1650, Argentina
4
Southern Scientific Center of Russian Academy of Science, 344006 Rostov-on-Don, Russia
5
iMED-Lab, Cadi Ayyad University, Marrakesh 40000, Morocco
*
Author to whom correspondence should be addressed.
Crystals 2021, 11(12), 1469; https://doi.org/10.3390/cryst11121469
Submission received: 31 October 2021 / Revised: 21 November 2021 / Accepted: 24 November 2021 / Published: 26 November 2021
(This article belongs to the Special Issue Research and Development of Ferroelectric Material)

Abstract

:
A quantum paraelectric SrTiO3 is a material situated in close proximity to a quantum critical point (QCP) of ferroelectric transition in which the critical temperature to the ferroelectric state is suppressed down to 0 K. However, the understanding of the behavior of the phase transition in the vicinity of this point remains challenging. Using the concentration x of Pb in solid solution Sr1−xPbxTiO3 (PSTx) as a tuning parameter and applying the combination of Raman and dielectric spectroscopy methods, we approach the QCP in PSTx and study the interplay of classical and quantum phenomena in the region of criticality. We obtain the critical temperature of PSTx and the evolution of the temperature-dependent dynamical properties of the system as a function of x to reveal the mechanism of the transition. We show that the ferroelectric transition occurs gradually through the emergence of the polar nanoregions inside the non-polar tetragonal phase with their further expansion on cooling. We also study the ferroelastic cubic-to-tetragonal structural transition, occurring at higher temperatures, and show that its properties are almost concentration-independent and not affected by the quantum criticality.

1. Introduction

There has been significant interest in quantum phenomena in strontium titanate oxide, SrTiO3, over the past decades. It has been observed that, upon cooling up to 30 K, the dielectric permittivity of this material follows the Curie–Weiss law with the characteristic Curie–Weiss temperature of ~37 K but then deviates from this behavior, saturating to a giant value of more than 20 000 at ~3 K [1]. Critical lattice dynamics reveals similar behavior. It has been observed that, unlike in conventional ferroelectrics, the frequency of the soft mode does not become unstable but also saturates at low temperatures [2,3]. It is widely accepted [1,4,5] that ferroelectric ordering in SrTiO3 at low temperatures is suppressed by the quantum fluctuations of the polar Ti ions, preventing their stabilization in the off-center potential energy minima positions, hence the formation of the ferroelectric state. In other words, SrTiO3 belongs to the noble systems in which the behavior is defined by quantum properties, widely known as quantum materials [6]. It is believed that the regime of proximity to the quantum critical point (QCP) [7] is realized in SrTiO3, which makes this material not only a natural laboratory to study the fundamental properties of quantum criticality but also a suitable system for application in the emerging non-von Neumann quantum and neuromorphic computing circuits.
The pristine SrTiO3 is located slightly outside the QCP. To achieve the genuine quantum ferroelectric transition at T = 0 and to study the crossover from quantum to classical criticality, one should apply the continuous driving force that suppresses the quantum fluctuation, restoring the ferroelectric order. Various means of stabilizing ferroelectricity in SrTiO3, including substrate-induced [8] and pressure-induced [9] strains, substitution of 18O isotope [4], and defect generation [10] were suggested. However, the most commonly used way is chemical doping with the substation of Sr by cation A (A = Ca, Ba, Pb, etc.) with the formation of Sr1−xAxTiO3 solid solution [11,12,13]. The advantage of this method is the well-targeted attainability of the QCP and the ferroelectric phase at a certain x, although the question of whether the polar state at a small x arises because of the polar ion displacement or because of the freezing of impurity-induced dipoles has been debated for many decades [14,15,16]. It is believed, however, that in Sr-based compounds, the quantum effects are strongly developed and prevent the frozen-glass behavior [12] that we also assume in our article.
In the present study, we explore the Sr1−xPbxTiO3 (PSTx) solid solutions with x = 0.005, 0.02, 0.04 by means of Raman and dielectric spectroscopies to investigate the effect of quantum fluctuations and the interaction of vibrational modes on the phase states in the vicinity of the QCP. We established that ferroelectric transition already takes place for the lowest Pb concentration x = 0.005 at the critical temperature of ~16 K. We demonstrate that this solid solution is very close to the QCP where the thermal and quantum fluctuations compete. Solid solutions with x = 0.02 and 0.04 possess higher critical temperatures, 56 K and 72 K, respectively, and are characterized by the more classical ferroelectric behavior. For all three Pb concentrations, we show that the ferroelectric transition occurs gradually through the emergence of the polar nanoregions inside the non-polar paraelectric phase. We give a detailed analysis of the temperature-dependent Raman spectra for these compounds in the temperature interval of 12–300 K and demonstrate that they also reveal a structural transition from the cubic P m 3 ¯ m to the non-polar tetragonal I4/mcm phase at a temperature of ~90 K.

2. Materials and Methods

The ceramic Sr1−xPbxTiO3 solid solutions with x = 0.005, 0.02, 0.04 were prepared from stoichiometric mixtures of SrCo3, TiO2, and PbO oxides according to the standard technique of solid-phase synthesis. The samples were calcined in the platinum crucible in the lead oxide ambient at T = 1400 K for 3 h and sintered at T = 1650 K for 1 h. Samples were pressed at 50 MPa in the form of 2 mm thick pellets with a diameter of 10 mm.
Synchrotron powder diffraction data were collected at the P24(EH2) PETRA III beamline, DESY (Hamburg, Germany), using a 2D marCCD 165 detector at temperature of 250 K. The monochromatic beam at wavelength λ = 0.5607 Å was slitted down to 300 × 300 μm2. The 200 μm diameter glass capillaries were used for mounting the finely ground powder samples. The sample-to-detector distance (D ≈ 75 mm) and the detector parameters were calibrated using a LaB6 NIST reference powder sample; calibration and subsequent integration were conducted using the Dioptas software [17]. The Rietveld refinement of the collected data was conducted using the FullProf software [18]. The pseudo-Voigt profile was used to model the Bragg reflection profile. At 250 K, the cubic unit cell parameters of the PSTx samples with x = 0.005, 0.02, 0.04 were 0.3910 nm, 0.3908 nm, and 0.3903, respectively. No parasitic phases were detected (see Supplementary materials).
The Raman spectra were excited in the range of 12–800 cm−1 using argon laser radiation with wavelength λ = 514.5 nm. They were recorded in a backscattering geometry using a Jobin Yvon T64000 spectrometer equipped with CCD and photon counting detector in the temperature range 15–150 K. An optical microscope with a 50× objective was used to focus the incident light as a 2 μm diameter spot on the sample. Low-temperature dependencies of the Raman spectra were received using a closed-cycle helium cryostat with temperature stability of ± 0.1 K. The measured Raman spectra were corrected for the Bose–Einstein temperature factor ( e ω k T 1 ) 1 .
Dielectric measurements at different temperatures were performed with an Agilent E4980 LCR-meter hooked to a Quantum Design Physical Properties Measurement System. Both sides of the pellets were covered with silver paste to create a capacitor-like structure. Capacitance C was measured at 10 kHz by assuming an equivalent parallel RC circuit, as is the standard for low leakage samples, and the corresponding relative permittivity εr was calculated by means of the equation εr = (C x d)/(ε0 x Ac), where d and Ac are the dielectric thickness and the capacitor area, respectively. The measured temperature range was between 2 K and 290 K.

3. Results

At room temperature, the SrTiO3 crystallizes in a cubic O h 1 symmetry with the P m 3 ¯ m space group. The optical representation of the Oh cubic phase includes three triply degenerated ferroelectric polar F1u modes and one triply degenerated “silent” F2u-symmetry mode: Γcub = 3F1u + F2u. According to the symmetry selection rules, neither of them are Raman active. The ferroelectric F1u modes are conventionally called a Slater mode [19], an Axe mode [20], and a Last mode [21]. Moreover, each F1u mode splits into transverse (TO) and longitudinal (LO) components due to the long-range dipole–dipole interaction. Although they are not Raman active, they can be observed by various techniques, such as Hyper-Raman and neutron scattering, and IR spectroscopy [22,23]. A Slater mode describes mutually opposite vibrational displacements of the Ti and O ions along the z-axis. According to infrared, hyper-Raman, and neutron data, the TO1 and LO1 components of the Slater mode are observed at the frequencies of ~88 cm−1 and ~175 cm−1, respectively [24]. An Axe mode corresponds to the oxygen vibrations leading to the bending of the oxygen frame. The TO2 and LO2 components of the Axe mode are found at the frequencies of ~175 and ~266 cm−1, respectively. A Last mode represents the mutual vibrations of the TiO6 octahedra framework and the Sr ions along the z-axis. Its TO4 and LO4 components can be detected at frequencies of ~545 and ~795 cm−1, respectively [24]. The so-called “silent” F2u mode is a structural mode describing oxygen vibrations in the xy-plane. This mode does not exhibit a dipole moment or the polarizability change of the molecule [25].
Upon cooling down to 105 K, the SrTiO3 undergoes the non-polar structural phase transition from the cubic O h 1 to the tetragonal D 4 h 18 phase with the I4/mcm space group caused by the antiphase rotations of the oxygen octahedra with multiple increases in the unit cell [26,27]. Such a tetragonal distortion leads to the splitting of the Slater, Axe, and Last modes with F1u-symmetry into the A2u and Eu components. A “silent” F2u mode transforms to the B2u and Eu doublets. Since oxygen rotation does not lead to the ferroactive ion displacement, the A2u and Eu modes are not Raman active. Note that the hydrostatic and chemical pressure or the epitaxial strains may induce the ferroelectric phase transition in the SrTiO3 with symmetry lowering to the orthorhombic C2v phase with the mm2 point group [28]. In this ferroelectric phase, each of the A2u and B2u modes transforms into the modes with B2 symmetry, and the Eu modes split into the A1 and B1 modes. All of these modes are Raman active.
Moreover, at the structural phase transition to the tetragonal phase, we observe additional lines in the Raman spectra at 100 K. Their appearance is explained by the multiplication of the SrTiO3 unit cell leading to the folding of the Brillouin zone; therefore, the modes from the R-point reflect the zone center. These modes can be interpreted as the A2g, [A2g + B1g], A1g, and Eg, [A1u + Eu] which are observed at frequencies of ~795, 229, 52, 40, and 436, respectively, and the two groups of [B1g + Eg] modes are observed at the two frequencies of ~144 and 440 cm−1 [29].
Figure 1 shows the temperature dependences of the Raman spectra of the PSTx solid solutions in a frequency range of 15–800 cm−1. At 100 K, the Raman spectra contain a set of peaks corresponding to the two-phonon scattering lines observed in pure SrTiO3 with the cubic phase P m 3 ¯ m [30]. The lowest frequency line at ~75 cm−1 corresponds to the difference combination band that is caused by the phonon’s creation and annihilation processes [30]. Aside from a different combination band, the frequency region of 250–400 cm−1 consists of the lines corresponding to the creation of phonons from the different bands (the so-called additional combination band) and to the creation of the phonons from the same phonon branch (the so-called overtones). The frequency range of 600–800 cm−1 also contains additional combination bands and overtones [30]. Since at 100 K the first-order Raman lines are not observed, we conclude that the PSTx has a cubic structure in which a non-polar P m 3 ¯ m phase has no vibrations related to the molecule polarizability changes.
Upon cooling, the Raman spectra of the PSTx reveal the wing of the low-frequency line centered at the frequency below 15 cm−1 in the temperature range of 90–12 K (see Figure 1). This line can be attributed to the Eg mode, which also indicates a structural phase transition at 90 K. The same mode was observed by Taniguchi et al. [31] in the oxygen-isotope-exchanged SrTiO3 at a temperature of ~100 K at the frequency of ~8 cm−1. Note that the frequency and intensity of the Eg mode increased upon cooling.
Moreover, at ~90 K, we observed the weak peak of A1g mode at frequency ~25 cm−1 that increases with the temperature decreasing (as shown in Figure 2). The emergence of this peak evidences the structural transition to the non-polar tetragonal I4/mcm phase. Thus, the temperature of the structural transition is coherent with the T-x phase diagram for PSTx solid solutions [32], where the transition temperature in the PSTx with x = 0.005, 0.02, 0.04 is observed at ~90 K.
The results obtained at low temperatures in [33] showed that the Raman spectra of the SrTiO3 contain the B1g and Eg modes at ~144 and ~440 cm−1 corresponding to the Ti-O vibrations that are also responsible for the transition to the tetragonal I4/mcm phase. However, in our experiment, these modes were not observed, presumably because of suppression of the Ti ions’ motions at a small amount of the Pb dopant.
The comparison of our Raman spectra of the PSTx with the small x, the spectra of the pure SrTiO3, and the spectra of the PSTx with x = 0.2 is shown in Figure 3. We observed that even a small amount of the Pb dopant leads to drastic spectral changes in comparison with the pure SrTiO3 ceramic. Note that the Raman spectrum of the PSTx with x = 0.2 contains a clear dip at ~67 cm−1 between two intensive peaks at ~24 cm−1 and ~86 cm−1. These peaks can be attributed to ferroelectric TO1 modes of the E and A1 symmetry that correspond to the tetragonal P4mm phase as shown in [34]. As can be seen in Figure 3, Raman spectra of the PSTx with x = 0.005, 0.02, 0.04 also reveal two intensive peaks, centered at ~22–29 cm−1 and at ~58–65 cm−1 separated by the dip at ~45 cm−1. According to the phase T-x diagram [32], low-x PSTx have the non-polar I4/mcm phase at low temperatures.
According to the symmetry selection rules for the I4/mcm phase, these lines can be assigned to the TO1 mode, which splits into the two ferroelectric Eu and A2u components. Note that, in pure oxygen-isotope-exchanged SrTiO3, the same modes appear at lower temperatures and frequencies, although these modes are forbidden in the Raman spectra of the paraelectric SrTiO3.
Analyzing the higher frequency region 150–800 cm−1 of the Raman spectra of the PSTx at low temperatures, we observe the two narrow lines at ~170 and ~540 cm−1 appearing on the background of the two-phonon scattering. Their intensity rises strongly with the temperature decreasing. These lines refer to the transverse components of the Axe and Last modes, TO2 and TO4, respectively, which are prohibited by the symmetry selection rules in paraelectric SrTiO3. The appearance of these modes in the Raman spectra of paraelectric perovskite materials is usually attributed to the pre-transitional polarization fluctuations. For instance, in the Sr1-xCaxTiO3 solid solutions [11] the appearance of these modes is associated with the ferroelectric phase transition [33]. Additionally, it was shown that, in the pure SrTiO3 ceramics or the BSTx solid solutions, the TO2 and TO4 lines indicate the presence of the polar nanoregions originated from the microdeformations between the stochastically oriented microcrystallites, impurities, or oxygen vacancies [3].
Moreover, we observe the emergence of the sharp peak at ~263 cm−1 upon cooling. This line is attributed to the TO3 mode, originating from the F2u “silent” structural mode. In the Raman spectra of the PSTx with x = 0.005, 0.02, and 0.04, the “silent” mode appears at 60, 50, and 80 K, respectively. Since this mode is not related to the polarizability change, its emergence may be associated with the structural transformation into the orthorhombic phase. The longitudinal LO3 component of a “silent” mode is also observed in the Raman spectra at 474 cm−1, but it is too blurred for the correct estimation of its parameters.

4. Discussion

For more detailed analysis of the Raman spectra, we deconvoluted them on the two-phonon and first-order Raman scattering using the damping oscillator functions (DHO model) and the Fano profile:
I ( ω ) = I 0 + i A i Γ 0 i ω 0 i 2 ω ( ω 0 i 2 ω 2 ) 2 + Γ 0 i 2 ω 2 + A j ( q + E ( ω ) ) 2 1 + E ( ω ) 2 ,
where I0 is the background intensity, ω0i is the frequency of the oscillator, Γ0i is the damping (half-width) of the oscillator, Ai and Aj are the strengths (intensity) of the oscillator and Fano profile, q is the asymmetry parameter of the Fano resonance, and E ( ω ) = 2 ( ω ω 0 ) Γ 0 i .
Note that the TO2 polar phonon mode observed at ~170 cm−1 demonstrates an asymmetric profile of the Fano resonance with the parameter q > 0. Such asymmetry was also previously observed in the SrTiO3-based materials synthesized in various designs [35]. Usually, the Fano resonance with positive q is a feature of the interaction between a discrete phonon state and a continuous excitation background, such as multiphonon states [36].
According to the deconvolution of the Raman spectra, the temperature dependence of the frequency of the polar Slater modes with Eu and A2u components, denoted as TO1, is shown in Figure 4. Note that the Eu component splits into A1 and B1 doublets in the orthorhombic phase as the temperature decreases. As a result, we observe three TO1 modes at low frequencies, demonstrating a soft mode behavior when the temperature changes. The temperature behavior of the TO1 mode frequency (about 70 cm−1) reveals the minimum critical temperatures ~18 K and ~50 K for the x = 0.005, ~50 K for the x = 0.02, and ~70 K for the x = 0.04. It is known that such soft mode behavior may indicate the macroscopic phase transition in the ferroelectrics [37,38]. Our results specify the symmetry of the low-T ferroelectric phase of PSTx as an orthorhombic one that is in line with the symmetry of the ferroelectric phases in other near-quantum ferroelectrics [11,28,29,31] but is in contrast to [32] where this phase was suggested as tetragonal. Note, however, that the suggestion in [32] is based on the simple extrapolation of the tetragonal symmetry P4mm of PbTiO3 to PSTx compounds. The TO2 and TO4 vibrations observed in the vicinity of 170 and 540 cm−1 are hard modes that slightly depend on the temperature changes.
The temperature changes for the intensity of the polar TO1, TO2, and the TO4 modes of PSTx with x = 0.005, 0.02, and 0.04 are shown in Figure 5. It is worth noting that the three soft TO1 modes reveal a strong interaction of phonons; therefore, we sum up the values of their intensities and present the total intensity of the collective TO1 modes on the top row in Figure 5. Upon cooling, the intensities of all polar modes reveal a sharp increase, demonstrating a kink-like behavior. As can be seen in Figure 5, the bending points at the x = 0.005, 0.02, and 0.04 are observed at temperatures of ~90 K, ~75 K, and ~100 K, respectively. As was shown for the Sr1−xBaxTiO3 solid solutions with small Ba content [3], the presence of prohibited lines in the paraelectric phase points out the existence of the polar nanoregions in the system. The abrupt growth in their intensity corresponds to the expansion of their dimensions. Note that studies of the SrTiO3 by resonant piezoelectric spectroscopy [39] also revealed the polar resonances at temperatures below 80 K, which become stronger on further cooling. This was interpreted as the appearance of polarity in the ferroelastic domain walls. Such thermally induced effects are consistent with the temperatures of the emergence of the polar nanoregions in our Raman data. Thus, in the case of the PSTx, we can assume that the Pb-concentration increase also leads to the rise in critical temperature of the expansion of the polar nanoregions. For the concentrations x = 0.02 and 0.04, we do observe such behavior. However, the temperature behavior for the sample with x = 0.005 is different, presumably due to the manifestation of the quantum effects.
Figure 6a shows how the low-temperature region of existence of the ferroelectric phase and the region of the emergence of the polar nanoregions (shaded area) evolve as a function of Pb concentration. The filled red and green circles present the characteristic temperatures obtained from Raman data, delimiting these regions. The filled blue circles present the independent measurements of the ferroelectric transition temperatures, obtained as maxima in the temperature dependencies of the dielectric constants (Figure 6b). For comparison, we also present the transition temperatures obtained in [32] from dielectric measurements (empty blue circles). All the data are in good agreement with each other; we found that the ferroelectric transition occurs at temperatures of ~16 K, 56 K, and 72 K for the samples with x = 0.005, 0.02, and 0.04, respectively.
The dependence of the ferroelectric transition temperatures on concentration was fitted by the theoretical formula for the system with quantum criticality [40], T c = A ( x x c ) 1 2 , with A 399 and xc 0.0025 (red dashed line in Figure 6a). The extrapolated critical concentration, xc 0.0025, corresponds to the quantum critical point in which a concentration-driving quantum phase transition from the paraelectric to ferroelectric phase occurs at zero temperature as a function of increasing concentration [4,41,42].
For PSTx with x = 0.005, the ferroelectric transition occurs at a finite temperature, Tc = 16–18 K, that is, however, substantially reduced with respect to the temperature at which the polar nanoregions emerge (~90 K) and with respect of the temperature of the critical soft mode slope change (~50 K). We assume therefore that the quantum effects are also relevant at x = 0.005 and that the transition is conditioned by competing interaction between thermal and quantum fluctuations. With increasing x, the ferroelectric transition temperature grows and the interval of pre-transitional polarization fluctuations shrinks. The system recovers classical thermal behavior.

5. Conclusions

Raman spectra of the PSTx solid solutions with x = 0.005, 0.02, and 0.04 were studied at T~12–300 K to obtain insight on the nature of the phase transitions emerging in this temperature interval. It was established that all samples undergo the structural transition from the cubic P m 3 ¯ m to the non-polar tetragonal I4/mcm phase at a temperature of ~90 K. Another transition corresponding to the transformation to the ferroelectric orthorhombic phase takes place at temperatures of ~16 K, 56 K, and 72 K for x = 0.005, 0.02, and 0.04, respectively. This transition was shown to occur gradually through the emergence of the polar nanoregions inside the non-polar tetragonal phase with their further expansion on cooling. The results were confirmed by dielectric measurements. It was revealed that the PSTx with the lowest Pb concentration x = 0.005 is close to the quantum critical point (QCP), where the thermal and quantum fluctuations compete.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/cryst11121469/s1, Figure S1: Diffraction pattern and refined model of profile for Sr1−xPbxTiO3 with x = 0.005, Figure S2: Diffraction pattern and refined model of profile for Sr1−xPbxTiO3 with x = 0.02, Figure S3: Diffraction pattern and refined model of profile for Sr1−xPbxTiO3 with x = 0.04, Table S1: Refined lattice parameters from Le Bail fit, maximal strains with anisotropy and reliability factors with all non-excluded points.

Author Contributions

Conceptualization, E.D.L., I.A.L. and A.G.R.; methodology, A.S.M., D.M., A.G.R. and D.R.; formal analysis, E.D.L., S.V.K. and A.G.R.; investigation, E.D.L., D.M., A.S.M. and D.R.; data curation, S.V.K., V.B.S. and A.G.R.; writing—original draft preparation, E.D.L., I.A.L. and A.G.R.; writing—review and editing, I.A.L., V.B.S. and A.G.R.; project administration, A.G.R. and D.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was financially supported by the Ministry of Science and Higher Education of the Russian Federation, grant agreement number 075-15-2021-953 from 5/10/2021, and by the European Union H2020-MSCA-RISE-MELON action, project number 872631.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Acknowledgments

D.R. thanks Beatriz Noheda, Groningen University, for the access to the equipment to perform the capacitance vs. temperature measurements and the MELON action for visiting financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Muller, K.A.; Burkard, H. SrTiO3: An intrinsic quantum paraelectric below 4 K. Phys. Rev. B 1979, 19, 3593–3602. [Google Scholar] [CrossRef]
  2. Yamanaka, A.; Kataoka, M.; Ibana, Y.; Inoue, K.; Hehlen, B.; Courtens, E. Evidence for competing orderings in strontium titanate from hyper-Raman scattering spectroscopy. Europhys. Lett. 2000, 50, 688–694. [Google Scholar] [CrossRef]
  3. Linnik, E.D.; Lukyanchuk, I.A.; Mikheykin, A.S.; Ragulya, A.V.; Gorshunov, B.P.; Alyabyeva, L.N.; Zhukov, S.S.; Mezzane, D.; Razumnaya, A.G. Crystal structure and the spectral response of the Ba-doped SrTiO3 incipient ferroelectrics. Phys. Status Solidi B 2021, 258, 2100010. [Google Scholar] [CrossRef]
  4. Rowley, S.E.; Spalek, L.J.; Smith, R.P.; Dean, M.P.M.; Itoh, M.; Scott, J.F.; Lonzarich, G.G.; Saxena, S.S. Ferroelectric quantum criticality. Nat. Phys. 2014, 10, 367–372. [Google Scholar] [CrossRef] [Green Version]
  5. Nova, T.F.; Disa, A.S.; Fechner, M.; Cavalleri, A. Metastable ferroelectricity in optically strained SrTiO3. Science 2019, 364, 1075–1079. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Cava, R.; de Leon, N.; Xie, W. Introduction: Quantum Materials. Chem. Rev. 2021, 121, 2777–2779. [Google Scholar] [CrossRef] [PubMed]
  7. Carr, L.D. Understanding Quantum Phase Transitions; CRC Press: Boca Raton, FL, USA, 2010; p. 756. [Google Scholar]
  8. Haeni, J.H.; Irvin, P.; Chang, W.; Uecker, R.; Reiche, P.; Li, Y.L.; Choudhury, S.; Tian, W.; Hawley, M.E.; Craigo, B.; et al. Room-temperature ferroelectricity in strained SrTiO3. Nature 2004, 430, 758–761. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Coak, M.J.; Hainesa, C.R.S.; Liua, C.; Rowley, S.E.; Lonzaricha, G.G.; Saxena, S.S. Quantum critical phenomena in a compressible displacive ferroelectric. Proc. Natl. Acad. Sci. USA 2020, 117, 12707–12712. [Google Scholar] [CrossRef]
  10. Lee, D.; Lu, H.; Gu, Y.; Choi, S.-Y.; Li, S.-D.; Ryu, S.; Paudel, T.R.; Song, K.; Mikheev, E.; Lee, S.; et al. Emergence of room-temperature ferroelectricity at reduced dimensions. Science 2015, 349, 1314–1317. [Google Scholar] [CrossRef] [Green Version]
  11. Bianchi, U.; Kleemann, W.; Bednorz, J.G. Raman scattering of ferroelectric Sr1-xCaxTiO3, x = 0.007. J. Phys. Condens. Matter. 1994, 6, 1229–1238. [Google Scholar] [CrossRef]
  12. Lemanov, V.V. Phase transitions and dielectric relaxation in incipient ferroelectrics with the perovskite structure. Ferroelectrics 2007, 346, 97–109. [Google Scholar] [CrossRef]
  13. Wang, J.X.; Zhang, C.; Liu, J.-M. Revisiting the phase transitions in BaxSr1-xTiO3 at low doping range (x ≤ 0.1). J. Alloys Compd. 2018, 749, 276–282. [Google Scholar] [CrossRef]
  14. Höchli, U.T. Dynamics of Freezing Electric Dipoles. Phys. Rev. Lett. 1982, 48, 1494. [Google Scholar] [CrossRef]
  15. Van der Klink, J.J.; Borsa, F. NMR study of the quasi-reorientational dynamics of Li ions in KTaO3:Li. Phys. Rev. B 1984, 30, 52. [Google Scholar] [CrossRef]
  16. Wang, R.; Inaguma, Y.; Itoh, M. Dielectric properties and phase transition mechanisms in Sr1-xBaxTiO3 solid solution at low doping concentration. Mater. Res. Bull. 2001, 36, 1693–1701. [Google Scholar] [CrossRef]
  17. Prescher, C.; Prakapenka, V.B. DIOPTAS: A program for reduction of two-dimensional X-ray diffraction data and data exploration. High Press. Res. 2015, 35, 223–230. [Google Scholar] [CrossRef]
  18. Rodriguez-Carvajal, J. Recent developments of the program FULLPROF. Comm. Powder Diffr. (IUCr) Newsl. 2001, 26, 12–19. [Google Scholar]
  19. Slater, J.C. The Lorentz correction in Barium titanate. Phys. Rev. 1950, 78, 748–761. [Google Scholar] [CrossRef]
  20. Axe, J.D. Apparent ionic charges and vibrational eigenmodes of BaTiO3 and other perovskites. Phys. Rev. 1957, 157, 429–435. [Google Scholar] [CrossRef]
  21. Last, J.T. Infrared-absorption studies on barium titanate and related materials. Phys. Rev. 1957, 105, 1740–1750. [Google Scholar] [CrossRef]
  22. Vogt, H.; Rossbroich, G. Accurate determination of the far-infrared dispersion in SrTiO3 by hyper-Raman spectroscopy. Phys. Rev. B 1981, 24, 3086–3091. [Google Scholar] [CrossRef]
  23. Hlinka, J.; Petzelt, J.; Kamba, S.; Noujni, D.; Ostapchuk, T. Infrared dielectric response of relaxor ferroelectrics. Phase Transit. 2006, 79, 41–78. [Google Scholar] [CrossRef]
  24. Barker, A.S. Temperature dependence of the transverse and longitudinal optic mode frequencies and charges in SrTiO3 and BaTiO3. Phys. Rev. 1966, 145, 391–399. [Google Scholar] [CrossRef]
  25. Hellwig, H.; Sehirlioglu, A.; Payne, D.A.; Han, P. Hyper-Raman-active soft mode in Pb(Mg1/3Nb2/3)0.73Ti0.27O3. Phys. Rev. B 2006, 73, 094126. [Google Scholar] [CrossRef]
  26. Cowley, R.A.; Buyers, W.J.L.; Dolling, G. Relationship of normal modes of vibration of strontium titanate and its antiferroelectric phase transition at 110 K. Solid State Commun. 1969, 7, 181–184. [Google Scholar] [CrossRef]
  27. Shirane, G.; Yamada, Y. Lattice-Dynamical Study of the 110 °K Phase Transition in SrTiO3. Phys. Rev. 1969, 177, 858–863. [Google Scholar] [CrossRef]
  28. Uwe, H.; Sakudo, T. Stress-induced ferroelectricity and soft phonon modes in SrTiO3. Phys. Rev. B 1976, 13, 271–286. [Google Scholar] [CrossRef]
  29. Shigenari, T.; Abe, K.; Takemoto, T.; Sanaka, O.; Akaike, T.; Sakai, Y.; Wang, R.; Itoh, M. Raman spectra of the ferroelectric phase of SrTi18O3: Symmetry and domains below Tc and the origin of the phase transition. Phys. Rev. B 2006, 74, 174121. [Google Scholar] [CrossRef]
  30. Nilsen, W.G.; Skinner, J.G. Raman Spectrum of Strontium Titanate. J. Chem. Phys. 1968, 48, 2240–2248. [Google Scholar] [CrossRef]
  31. Taniguchi, H.; Itoh, M.; Yagi, T. Ideal Soft Mode-Type Quantum Phase Transition and Phase Coexistence at Quantum Critical Point in 18O-Exchanged SrTiO3. Phys. Rev. Lett. 2007, 99, 017602. [Google Scholar] [CrossRef]
  32. Lemanov, V.V. Phase transitions in SrTiO3 quantum paraelectric with impurities. Ferroelectrics 1999, 226, 133–146. [Google Scholar] [CrossRef]
  33. Petzelt, J.; Ostapchuk, T.; Gregora, I.; Rychetsky, I.; Hoffmann-Eifert, S.; Pronin, A.V.; Yuzyuk, Y.; Gorshunov, B.P.; Kamba, S.; Bovtun, V.; et al. Dielectric, infrared, and Raman response of undoped SrTiO3 ceramics: Evidence of polar grain boundaries. Phys. Rev. B 2001, 64, 184111. [Google Scholar] [CrossRef]
  34. Jain, M.; Yuzyuk, Y.I.; Katiyar, R.S.; Somiya, Y.; Bhalla, A.S. Local symmetry breaking in PbxSr1−xTiO3 ceramics and composites studied by Raman spectroscopy. J. Appl. Phys. 2005, 98, 024116. [Google Scholar] [CrossRef]
  35. Sirenko, A.A.; Akimov, I.A.; Fox, J.R.; Clark, A.M.; Li, H.-C.; Si, W.; Xi, X.X. Observation of the first-order Raman scattering in SrTiO3 thin films. Phys. Rev. Lett. 1999, 82, 4500–4503. [Google Scholar] [CrossRef] [Green Version]
  36. Gupta, S.; Katiyar, R.S. Temperature-dependent structural characterization of sol–gel deposited strontium titanate (SrTiO3) thin films using Raman spectroscopy. J. Raman Spectrosc. 2001, 32, 885–891. [Google Scholar] [CrossRef]
  37. Lyddane, R.H.; Sachs, R.G.; Teller, E. On the Polar Vibrations of Alkali Halides. Phys. Rev. 1941, 59, 673–679. [Google Scholar] [CrossRef]
  38. Cochran, W.; Cowley, R.A. Dielectric constants and lattice vibrations. J. Phys. Chem. Solids 1962, 23, 447–459. [Google Scholar] [CrossRef]
  39. Salje, E.K.H.; Aktas, O.; Carpenter, M.A. Domains within domains and walls within walls: Evidence for polar domains in cryogenic SrTiO3. Phys. Rev. Lett. 2013, 111, 247603. [Google Scholar] [CrossRef]
  40. Morf, R.; Schneider, T.; Stoll, E. Nonuniversal critical behavior and its suppression by quantum fluctuations. Phys. Rev. B 1977, 16, 184111. [Google Scholar] [CrossRef]
  41. Rischau, C.W.; Lin, X.; Grams, C.P.; Finck, D.; Harms, S.; Engelmayer, J.; Lorenz, T.; Gallais, Y.; Fauqué, B.; Hemberger, J.; et al. A ferroelectric quantum phase transition inside the superconducting dome of Sr1-xCaxTiO3. Nat. Phys. 2017, 13, 643–648. [Google Scholar] [CrossRef]
  42. Rowley, S.; Smith, R.; Dean, M.; Spalek, L.; Sutherland, M.; Saxena, M.; Alireza, P.; Ko, C.; Liu, C.; Pugh, E.; et al. Ferromagnetic and ferroelectric quantum phase transitions. Phys. Status Solidi B 2010, 247, 469–475. [Google Scholar] [CrossRef]
Figure 1. Temperature-dependent Raman spectra of the Sr1−xPbxTiO3 solid solutions at concentrations x = 0.005, 0.02, 0.04 in the temperature range 12–100 K.
Figure 1. Temperature-dependent Raman spectra of the Sr1−xPbxTiO3 solid solutions at concentrations x = 0.005, 0.02, 0.04 in the temperature range 12–100 K.
Crystals 11 01469 g001
Figure 2. The temperature evolution of the Raman spectra of the Sr1−xPbxTiO3 solid solutions at concentrations x = 0.005, 0.02, 0.04 in the temperature range 50–100 K. The experimental data are shown by the black circles. The red lines show the results of spectra deconvolution. The stars indicate the A1g mode.
Figure 2. The temperature evolution of the Raman spectra of the Sr1−xPbxTiO3 solid solutions at concentrations x = 0.005, 0.02, 0.04 in the temperature range 50–100 K. The experimental data are shown by the black circles. The red lines show the results of spectra deconvolution. The stars indicate the A1g mode.
Crystals 11 01469 g002
Figure 3. The Raman spectra of the PSTx solid solutions with small amount of Pb (x = 0.005, 0.02, 0.04) and their comparison with the Raman spectra of the pure SrTiO3 and the PSTx with x = 0.2 at 50 K.
Figure 3. The Raman spectra of the PSTx solid solutions with small amount of Pb (x = 0.005, 0.02, 0.04) and their comparison with the Raman spectra of the pure SrTiO3 and the PSTx with x = 0.2 at 50 K.
Crystals 11 01469 g003
Figure 4. The temperature evolution of the TO1 mode frequency of the Sr1−xPbxTiO3 solid solutions at selected concentrations x = 0.005, 0.02, 0.04. The dashed lines indicate the changes in slope for the temperature dependencies of the soft mode frequency.
Figure 4. The temperature evolution of the TO1 mode frequency of the Sr1−xPbxTiO3 solid solutions at selected concentrations x = 0.005, 0.02, 0.04. The dashed lines indicate the changes in slope for the temperature dependencies of the soft mode frequency.
Crystals 11 01469 g004
Figure 5. The temperature evolution of intensities of the TO1 modes (upper row) and of the TO2 and TO4 modes (bottom row) for the Sr1−xPbxTiO3 solid solutions at x = 0.005, 0.02, 0.04. The dashed lines show the interpolation of the modes intensity.
Figure 5. The temperature evolution of intensities of the TO1 modes (upper row) and of the TO2 and TO4 modes (bottom row) for the Sr1−xPbxTiO3 solid solutions at x = 0.005, 0.02, 0.04. The dashed lines show the interpolation of the modes intensity.
Crystals 11 01469 g005
Figure 6. (a) Characteristic temperatures of the emergence of the polar nanoregions (red circles) and critical temperatures of the ferroelectric transition (green and blue circles) for different Pb concentrations x. The filled red and green circles correspond to the Raman measurements, and the filled blue circles to the dielectric measurements. The empty blue circles correspond to the dielectric data obtained in Ref. [32]. The high-temperature area of the plot corresponds to the paraelectric phase of PSTx, the intermediate shaded area corresponds to the emergence of the polar nanoregions, and the low-temperature area corresponds to the ferroelectric phase. The dashed red line gives the fit of the results by theoretical formula, T c = A ( x x c ) 1 2 . (b) Temperature dependence of the dielectric permittivity for different Pb concentrations x.
Figure 6. (a) Characteristic temperatures of the emergence of the polar nanoregions (red circles) and critical temperatures of the ferroelectric transition (green and blue circles) for different Pb concentrations x. The filled red and green circles correspond to the Raman measurements, and the filled blue circles to the dielectric measurements. The empty blue circles correspond to the dielectric data obtained in Ref. [32]. The high-temperature area of the plot corresponds to the paraelectric phase of PSTx, the intermediate shaded area corresponds to the emergence of the polar nanoregions, and the low-temperature area corresponds to the ferroelectric phase. The dashed red line gives the fit of the results by theoretical formula, T c = A ( x x c ) 1 2 . (b) Temperature dependence of the dielectric permittivity for different Pb concentrations x.
Crystals 11 01469 g006
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Linnik, E.D.; Mikheykin, A.S.; Rubi, D.; Shirokov, V.B.; Mezzane, D.; Kondovych, S.V.; Lukyanchuk, I.A.; Razumnaya, A.G. Raman Response of Quantum Critical Ferroelectric Pb-Doped SrTiO3. Crystals 2021, 11, 1469. https://doi.org/10.3390/cryst11121469

AMA Style

Linnik ED, Mikheykin AS, Rubi D, Shirokov VB, Mezzane D, Kondovych SV, Lukyanchuk IA, Razumnaya AG. Raman Response of Quantum Critical Ferroelectric Pb-Doped SrTiO3. Crystals. 2021; 11(12):1469. https://doi.org/10.3390/cryst11121469

Chicago/Turabian Style

Linnik, Ekaterina D., Alexey S. Mikheykin, Diego Rubi, Vladimir B. Shirokov, Daoud Mezzane, Svitlana V. Kondovych, Igor A. Lukyanchuk, and Anna G. Razumnaya. 2021. "Raman Response of Quantum Critical Ferroelectric Pb-Doped SrTiO3" Crystals 11, no. 12: 1469. https://doi.org/10.3390/cryst11121469

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