Soft Mode Spec of FE and MF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

REVIEW ARTICLE | FEBRUARY 17 2021

Soft-mode spectroscopy of ferroelectrics and multiferroics:


A review
Special Collection: 100 Years of Ferroelectricity — a Celebration

S. Kamba 

APL Mater. 9, 020704 (2021)


https://doi.org/10.1063/5.0036066


View Export
Online Citation

30 October 2024 15:21:18


APL Materials RESEARCH UPDATE scitation.org/journal/apm

Soft-mode spectroscopy of ferroelectrics


and multiferroics: A review
Cite as: APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066
Submitted: 3 November 2020 • Accepted: 30 December 2020 •
Published Online: 17 February 2021

S. Kambaa)

AFFILIATIONS
Institute of Physics of the Czech Academy of Sciences, Na Slovance 2, 182 21 Prague 8, Czech Republic

Note: This paper is part of the Special Topic on 100 Years of Ferroelectricity—A Celebration.
a)
Author to whom correspondence should be addressed: kamba@fzu.cz

ABSTRACT
This article summarizes the results of the investigations of the dynamics of ferroelectric (FE) phase transitions (PTs) obtained in Prague
during the last 25 years. After a short introduction, explaining differences between displacive and order-disorder types of FE PTs, the results
of the broadband dielectric, THz, and mainly IR spectroscopic investigations of hydrogen-bonded FEs, BaTiO3 , relaxor FEs, strained incipient
FEs, and various multiferroics are reviewed. The high sensitivity of the IR spectroscopy to polar phonons was demonstrated in ultrathin films,

30 October 2024 15:21:18


which allowed us to reveal strain-induced FE PTs. Electrically active magnons (i.e., electromagnons) can be observed in the IR and Raman
spectra of multiferroics. Their frequencies soften on heating toward temperatures of magnetic PTs similarly as phonons in displacive FEs. As
expected, the electromagnons can be dependent on the external magnetic field. This behavior has been demonstrated in BiFeO3 , SrMn7 O12 ,
and multiferroics with Y- and Z-type hexaferrite crystal structures.
© 2021 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/5.0036066

I. INTRODUCTION quantum paraelectric SrTiO3 above 90 K. By the way, most of the


experimental works performed in the 1960s refer to soft mode
Let us start first with a short historical remark. The concept of studies of quantum paraelectrics SrTiO3 and KTaO3 probably due
FE soft mode, i.e., one or several polar phonons, whose frequency to the relatively high soft mode frequency and large temperature
tends to zero or greatly decrease near the temperature of structural changes observed in these materials. Regarding the experimental
PT T c , is very old, and it is frequently used for the explanation of the techniques, mostly inelastic neutron scattering and Raman scat-
dynamics of PTs. Classical Cochran’s papers1–3 from 1959 to 1961 tering were used. Far infrared (IR) spectroscopy was less used,
are usually considered as the first papers about this topic, but Ander- but it became more popular in the 1970s with the development
son4 came with the idea of soft mode already in 1958. He published of fast Fourier spectroscopy. Nevertheless, this technique is more
his paper only in Russian, and although the paper appeared only in appropriate than other techniques because the soft mode is IR
conference proceeding two years later,4 it is rather frequently cited in active in both paraelectric and FE phases, in contrast to Raman
Eastern Europe. Nevertheless, one must accept that the first original scattering, where the soft mode is mostly active only in the FE
ideas about the FE soft mode were published by Ginzburg already in phase or it can be activated in the paraelectric phase by an
1949,5–7 but at that period, the Russian papers were not translated to external electric field. Inelastic neutron scattering has the advan-
English; therefore, these articles were forgotten. tage that it allows us to measure the dispersion of the soft opti-
The first observation of the soft mode is even older. Landsberg cal phonon branch in the whole Brillouin zone, but this mea-
and Mandelstam8 saw some soft modes in Raman spectra near the surement is much more time consuming and matchlessly more
α ↔ β transition in quartz already in 1929, but they explained its expensive, and moreover, large single crystals are needed for such
temperature behavior inaccurately. Therefore, the papers of Barker experiments.
and Tinkham,9 as well as that of Cowley10 published in 1962, are In the beginning of the 1970s, the soft mode spectroscopy was
considered as the first papers, which experimentally confirmed the used frequently for the study of dynamics of phase transitions—see
soft mode theory. Let us note that both papers concern the study of the review papers11–13 and a monography.14 It was shown that the

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-1


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

optical soft mode is a dynamical origin of only displacive struc- expressed as a sum of damped harmonic oscillators and Debye
tural PTs. In the case of order–disorder PTs (see below), the soft relaxations,14
mode is not an optical phonon, but it has a relaxational character,
and slowing down of the relaxation frequency is seen in microwave ε ∗(ω) = ε∞ + εph + εdr
(MW), radio frequency, or low-frequency ranges on cooling toward Δεj ω2TOj
n m ΔεRj ωRj
T c .11,13 = ε∞ + ∑j + ∑j . (2)
The demand of miniaturization of the new capacitors in inte- ω2TOj − ω2 + iωγj ωRj + iω
grated circuits requires new dielectric materials with high permittiv-
ity. In the case of non-volatile memories, the capacitors or transistor ωTOj , γj , and Δεj denote the transverse frequency, damping, and
gates should be moreover FE. However, the permittivity in ceramics dielectric contribution to the static permittivity of the j-th polar
and thin films is markedly reduced compared to the single crystals phonon, respectively, and ΔεRj and ωRj mark the dielectric strength
due to the low-permittivity grain boundaries, dead layers between and relaxation frequency of the j-th relaxation, respectively. In the
the thin films and substrates, as well as the strain in the films caused case of displacive PTs, the relaxation contribution εdr should be zero
by a misfit between the substrates and films.15 Understanding the in the paraelectric phase and the increase in permittivity near T c [see
permittivity reduction is necessary for processing good quality thin Eq. (1)] is caused by softening (i.e., decrease in the frequency) of one
films and ceramics, and broadband dielectric spectroscopy helps a of the polar phonons. If the phonons are not coupled, the oscillator
lot in this case. strengths f j = Δεj ω2TOj of all polar modes are expected to be roughly
In the 1990s, a giant piezoelectric effect was revealed in FE temperature independent. It means, for example, that if the soft
relaxor-based single crystals,16 which allowed for the construction of mode frequency ωSM reduces twice, ΔεSM increases four times, etc.
new and much more efficient piezoelectric devices. Since that time, The temperature dependence of the ωSM can be easily obtained from
the interest in relaxors rapidly increased, and mainly, the broad- the Curie–Weiss law [Eq. (1)] and Lyddane–Sachs–Teller relation,18
band dielectric spectroscopy (including our works) contributed a lot which expresses the relation between static permittivity ε0 and high
to the understanding of the mechanisms of huge permittivity, giant frequency permittivity ε∞ with longitudinal ωLO and transverse ωTO
piezoelectric effect, and structural PTs in these materials. phonon frequencies. For materials with two atoms per unit cell, it
Within the last 25 years, our Prague’s group investigated the has the form18
dynamics of structural PTs in many FE materials, which allowed us ε0 ω2
to generalize the soft mode behavior: In reality, the SM never softens = 2LO . (3)

30 October 2024 15:21:18


ε∞ ωTO
completely, and some finite frequency can be observed at T c . Several
reasons can be given for the softening saturation:17 ε∞ and ωLO are usually only slightly temperature dependent, and
● First-order nature of the structural PT. because ε0 diverges at T c [see Eq. (1)], the transverse phonon fre-
● Quantum effects at low temperatures in incipient quency ωTO should decrease to zero at T c . If we assume ε∞ and ωLO
ferroelectrics. to be constant and if we put ωTO = ωSM , we obtain the Cochran law
● The polar soft mode is not the critical soft mode of the for the temperature dependence of the soft mode frequency,1
PT. The proper soft mode comes from outside of the Bril-
louin zone center. This is the case of antiferroelectric PTs, ω2SM = A(T − Tc ), (4)
incommensurate FE transitions, or improper FE PTs.
● The most frequent case is that the central mode type disper- where A is the Cochran constant. For materials with more atoms
sion (in other words, a dielectric relaxation) appears near T c per unit cell, the generalized Lyddane–Sachs–Teller relation has the
below the optical SM frequency. Both modes are coupled, form14
and the central mode is mainly responsible for the dielectric ε0
2
n ωLOj
anomaly near T c . This case occurs due to defects or more = ∏j=1 2 . (5)
ε∞ ωTOj
often at the crossover from the displacive to order–disorder
type PTs. In the case of non-coupled phonons, only one transverse
In this work, we will review results mostly from the last 25 years. phonon frequency is temperature dependent and one can obtain
again the Cochran law [Eq. (4)]. Although the concept of the optical
soft mode as the dynamical origin of the PT is accepted since 1960,
II. SCIENTIFIC BACKGROUND there are only several pure displacive ferroelectrics known (e.g.,
In usual proper ferroelectrics, the low-frequency permittivity PbTiO3 19 ), where the low-frequency dielectric anomaly is caused
above the second order PT temperature T c follows the Curie–Weiss only by softening of the optical soft phonon. Also for pure incip-
law,14 ient ferroelectrics such as SrTiO3 9,20,21 and KTaO3 ,20,22 where the
C soft mode does not soften completely, because the FE PT is not
ε′ = , (1) reached in these materials, the soft mode is completely responsi-
T − Tc
ble for the anomalous temperature dependence of the low-frequency
where C is the Curie–Weiss constant. The complex permittivity ε∗ permittivity.
= ε′ + iε′′ is determined by a sum of contributions of the high- If the high-temperature paraelectric structure is disordered, i.e.,
frequency electronic polarization (ε∞ ), polar phonons (εph ), and if some kind of atoms occupy two or more equivalent positions with
possibly of some dielectric relaxations (εdr ), which can be usually the probability less than 1, the atoms usually order below T c only

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-2


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

in some of the mentioned sites and we say that the system under-
goes an order–disorder phase transition at T c . If the ordering of ions
is connected with the appearance of spontaneous polarization, the
PT is FE. In this case, the phonons may exhibit only small temper-
ature anomalies and the dynamical origin of the PT is the critical
slowing-down of some relaxation frequency ωRj [see the last term in
Eq. (2)] near T c , which expresses the hopping motion of the dis-
ordered atoms among the equivalent positions in the paraelectric
phase. The soft relaxation frequency ωRs in the disordered materi-
als plays the role of the optical soft mode in displacive ferroelectrics
and follows the simple linear temperature dependence above T c ,13

ωRs = A′ (T − Tc ). (6)

Below T c , the soft dielectric relaxation usually disappears from


the spectra, but another relaxation may appear due to the dynamics
of FE domain walls. If we take into account that the relaxation fre-
quency cannot increase up to infinity on heating, we should add to
the linear dependence [Eq. (6)] some saturation frequency ωsat . The
modified law for the soft relaxation frequency has then the form23

Rs = [A (T − Tc )] + ω−1
−1
ω−1 ′
sat . (7)

The relaxation strength f Rs = ΔεRs ωRs of the soft relaxation in


the paraelectric phase can be assumed temperature independent,
and due to the slowing down of ωRs , one can obtain the Curie–Weiss
increase in the static permittivity. Into a group of materials with FIG. 1. Temperature dependence of the real part of permittivity in the vicinity
order–disorder PTs belong triglycine sulfate, sodium nitrite,11,14 or of order–disorder FE PT in (C3 N2 H5 )5 Bi2 Cl11 . Reproduced with permission from

30 October 2024 15:21:18


(C3 N2 H5 )5 Bi2 Cl11 .24 The typical feature of the order–disorder PT is A. Piecha, G. Bator, and R. Jakubas, J. Phys.: Condens. Matter 17, L411–L417
(2005). Copyright 2005 IOP Publishing.
the shape of dielectric anomaly at T c . The Curie–Weiss behavior is
fulfilled only for ε′ measured at frequencies below ωRs . If the mea-
sured frequency is above ωRs , the Debye relaxator [Eq. (2)] stops to
contribute into static ε′ and a decrease in permittivity is observed in
the vicinity of PT and a minimum in ε′ (T) develops at T c . For exam- which will be discussed here, other examples can be found in the
ple, in (C3 N2 H5 )5 Bi2 Cl11 , the relaxation frequency softens down to reviews.11,13,29
10 kHz, and therefore, the minimum in ε′ (T) is seen at T c in Fig. 1 It seems nowadays that the PTs with a pure soft optical mode
at frequencies above 10 kHz. In many materials with order–disorder or a pure soft relaxation mode are rather rare. Most of the materials
PTs, the relaxation frequency softens only down to the microwave exhibit a crossover between displacive and order–disorder PT, i.e.,
region, and therefore, the minimum in ε′ (T) is not seen in dielec- the phonons exhibit anomalies near T c ; however, their contributions
tric data taken below 1 MHz, i.e., in the most frequently studied to permittivity do not explain the complete dielectric anomaly near
frequency range. T c . Additional soft relaxation (i.e., central mode) frequently appears
Many materials were reported as classical displacive or near T c , and only the sum of both contributions is responsible for
order–disorder ferroelectrics; however, their detailed investigations the dielectric anomaly near T c .
have shown that they usually exhibit some crossover between both
kinds of mechanisms of the PTs. For example, an underdamped
III. INVESTIGATION OF HYDROGEN BONDED
optical soft mode was observed in the THz spectra of tris-sarcosine FERROELECTRICS—TARTRATES AND KDP
calcium chloride (TSCC);25 therefore, this material was assumed to
be displacive FE. However, the careful reinvestigation of the THz Sodium potassium tartrate tetrahydrate NaKC4 H4 O6 ⋅ 4H2 O,
spectra of TSCC and comparison with the low-frequency dielectric called Rochelle salt (RS), is the first known ferroelectrics. Valashek
data revealed a soft relaxation (central mode) close to T c , which is revealed a FE hysteresis loop in this compound already in 192030
responsible for critical deviations from the Curie–Weiss law and but published in detail in 1921.31 Although RS was studied by many
the Cochran law near T c .26 The central mode was assigned to crit- authors within the last 100 years, the understanding of the mecha-
ical polar cluster dynamics. On the other hand, also many ferro- nism leading to the formation of a state with a spontaneous polariza-
electrics with order–disorder PTs show critical slowing down of the tion was a long time puzzle. In contrast to all other ferroelectrics, its
relaxation, which is coupled with optical phonons or other higher- FE phase (monoclinic space group P21 -C22 , Z = 4) exists only between
frequency relaxations. It means that the driving force of the PT T c1 = 255 K and T c2 = 297 K. Above and below these tempera-
is the latter excitation, although the dielectric anomaly near T c is tures, RS is paraelectric with the same orthorhombic space group
mainly due to the soft relaxation. As the examples KDP-type ferro- P21 21 2-D32 , Z = 4. Older neutron diffraction data32 showed that the
electrics,12 BaTiO3 ,27 Rochelle salt,28 or LiNaGe4 O9 23 can be named, whole tartrate and crystal water molecules are disordered in the

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-3


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

paraelectric phase with a small disordering amplitude. Newer rein-


vestigation using synchrotron x-ray diffraction revealed a disorder
of only K atoms and three water O atoms in the paraelectric phase,
while the bonded H atoms in water are not disordered.33 A Debye
relaxation mode was observed in the MW region, and its relax-
ation frequency linearly decreases with temperature near T c1 and
T c2 ;34–36 therefore, the PTs in RS were regarded as classical examples
of the order–disorder type. However, later, it was shown that at low
temperatures, the relaxation hardens to the THz range and remark-
ably underdamps so that it has phonon (i.e., resonant) character.37
Because its oscillator strength f SM = ΔεSM ω2SM is reduced on cooling
below T c1 by one order of magnitude, Volkov et al.38 suggested that
this mode is probably linearly coupled with another optical phonon
in the FIR range. It means that the transition is more likely displacive
than order–disorder transition.
The lattice dynamics of RS was also investigated by means of IR,
Raman, and inelastic neutron scattering spectroscopies.28,39 The soft
mode near 20 cm−1 was observed at low temperatures, and this mode
is linearly coupled with the phonon near 75 cm−1 [see Fig. 2(a)].
The higher frequency mode partially softens on heating from low
temperatures to T c1 , and it causes the softening and rising damping
of the mode near 20 cm−1 . Moreover, the soft mode near 20 cm−1
is coupled to an overdamped mode with a temperature indepen-
dent bare frequency of 13 cm−1 . Temperature dependent coupling of
these modes causes critical slowing down of the dielectric relaxation,
which is responsible for the strong dielectric anomalies seen near T c1
and T c2 . It means that the dynamical origin of the FE PTs in RS is not

30 October 2024 15:21:18


only a dielectric relaxation but also unstable optical phonons near
20 cm−1 and 75 cm−1 . Therefore, the PTs in RS can be treated as
a mixture of displacive and order–disorder type.28 This conclusion
was also confirmed by IR measurements at high hydrostatic pres-
sures, where the largest influence of the pressure on the phonon
frequency near 75 cm−1 was observed.40 The frequency of this
mode increases with pressure, which corresponds with the published
increase in critical temperatures with pressure.41
Chemically relative to RS is lithium thallium tartrate monohy- FIG. 2. (a) Temperature dependence of the mode frequencies obtained from the
drate (LTT)—LiTlC4 H4 O6 ⋅ H2 O. It exhibits only one FE PT, and fit of E∣∣a polarized spectra with the coupled-oscillator model. Modes with the bare
its critical temperature T c = 11 K is one of the lowest known.42 The frequencies ω1 and ω2 are coupled below 140 K, while the modes with frequencies
free permittivity ε′a obeys the Curie–Weiss law well above T c , but in ω1 and ω0 are coupled at higher temperatures, and this coupling induces a soft
contrast to other ferroelectrics, ε′a decreases only slightly below T c .43 critical dielectric relaxation ωR near both T c1 and T c2 . (b) Dielectric loss spectra
ε′′ 206 and THz37 spectra. Repro-
The permittivity is very sensitive to the mechanical boundary condi- a obtained from the fits of experimental microwave
duced with permission from S. Kamba, G. Schaack, and J. Petzelt, Phys. Rev. B
tion, so whereas ε′a reaches ∼5000 at T c , the clamped ε′a (T c ) is only 51, 14998–15007 (1995). Copyright 1995 American Physical Society.
∼30.44 Both the elastic compliance at constant electric field sE44 and
the piezoelectric coefficient d14 become exceedingly large close to
T c (note that LTT is piezoelectric already in the paraelectric phase
above T c ) and appear to have the highest magnitude measured in the PT is purely displacive, although the optical soft mode does
any materials.44 not soften completely due to its strong coupling with a transverse
The optical soft mode in LTT was first observed in far IR spec- acoustic mode.
tra by Gerbaux et al.45 and investigated in more detail by Volkov The piezoelectric paraelectric phase has an orthorhombic struc-
et al.46 The latter authors have shown that its frequency softens from ture (P21 21 2–D32 , Z = 4),48,49 but the structure of the FE phase was not
21 cm−1 at 300 K to 9 cm−1 at T c , but its dielectric contribution Δε known. Therefore, new IR, Raman, dielectric, and structural mea-
was two orders of magnitude smaller that the observed free per- surements of LTT were performed. The high-pressure dielectric data
mittivity near T c . Therefore, Volkov et al.46 claimed that another show that T c strongly increases with hydrostatic pressure (up to 54 K
dispersion of the relaxation type should exist in the MW region and at 370 MPa).50 Permittivity ε′a (T) exhibits an unusual plateau below
the PT is of mixed displacive and order–disorder type. Nevertheless, T c and then drops down in dependence on pressure by 5 K–15 K
Hayashi et al.47 have shown that the free ε′a is enhanced due to piezo- below T c (Fig. 3). FE hysteresis loops are only slim in the range
electric resonances in the MW region and that the high frequency of plateau, giving evidence about easy movement of domain walls,
permittivity corresponds very well to the THz ε′a (T).46 Therefore, but below the temperature of the drop down (T c2 ), the domains are

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-4


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

increase in T c with hydrostatic pressure was observed, in the case


of the LAT crystal, T c decreases only slightly down to 91 K at
400 MPa.53
One of the most studied hydrogen-bonded ferroelectrics is
potassium dihydrogen phosphate KH2 PO4 (KDP). At 123 K, it
undergoes the FE PT from the tetragonal (space group I 4̄¯ 2d) to
orthorhombic FE phase with the Fdd2 space group.54 Structural
data show that the hydrogen atoms in O–H⋅ ⋅ ⋅O hydrogen bonds
move in a double minimum potential. Its shape has been determined
by Lawrence and Robertson55 based on earlier published mid-IR
transmission spectra.
Many far IR reflectivity studies have been published.56–63
Since the lowest frequency band is overdamped and it can be
well described by the Debye relaxation model whose frequency
exhibits slowing down toward T c ,64 the PT was first assumed to
be of an order–disorder type. This statement was also supported
by microwave and submillimeter dielectric studies.64,65 The E∣∣c
polarized IR spectra can be fitted by two strongly coupled oscilla-
tors.56,61,62 The internal ν4 polar mode polarized along the ferro-
electric c axis slightly softens toward T c and couples with the inter-
site tunneling proton motions perpendicular to the c axis but also
with other polar modes of the same symmetry. The coupled exci-
tation evolves to yield the overdamped FE mode that triggers the
FE PT.62 According to Kobayashi’s theory,66 the lower frequency
oscillator represents predominantly the H tunneling mode, while the
higher frequency mode is predominantly the K+ −PO4 lattice vibra-
tion along the c-axis. Afterward, Blinc and Žekš, based on NMR and

30 October 2024 15:21:18


neutron scattering data, interpreted the FE soft mode by coupling of
K vibration with the whole H2 PO4 group.67
The temperature dependence of the phonon frequencies
obtained from the fits of IR reflectivity spectra using the factorized
form of the dielectric function is shown in Fig. 4(a). The frequency
of the overdamped SM mode is much higher than the relaxation fre-
FIG. 3. (a) Temperature dependence of the permittivity ε′a of LTT at various pres- quency ωR obtained from the Debye fit of submillimeter dielectric
sures. (b) Pressure dependence of the soft-mode frequencies at different temper- spectra64 [see Fig. 4(b)]. The fit of IR spectra gives ωTO frequency,
atures obtained from the Raman b(cb)c spectra. The broken curves in the high- but due to a large phonon damping γTO , ωTO is an ambiguous
pressure FE phase are the results of the square-root fits (equivalent to the Cochran
parameter (strongly influenced by γTO ) and the frequency of the
law as a function of pressure). Reproduced with permission from Kamba et al., J.
Phys.: Condens. Matter 8, 4631–4642 (1996). Copyright 1996 IOP Publishing. dielectric loss maximum corresponding to ωR = ω2TO /γTO has a bet-
ter physical meaning.13 One can also see from Fig. 4(b) that T c rises
and ωR decreases with deuteration of KDP. Extremely low ωR is also
a reason why the SM reflection band is hardly seen in far-IR spectra
frozen and the sample cannot be switched.50 The unusual temper- of fully deuterated KDP.56
ature independence of ε′a seen between T c and T c2 was theoreti- The dynamics of the PT in KDP has also been frequently inves-
cally explained by Tagantsev with a bilinear coupling between the tigated using Raman spectroscopy because the paraelectric phase is
order parameter (soft mode) and strain.51 The optical soft mode is also piezoelectric and the selection rules for IR and Raman activi-
strongly temperature and pressure dependent. It exhibits minimum ties of phonons are the same. Similarly, as in the IR spectra, strongly
frequency at T c but no anomaly near T c2 ,50 i.e., no structural change coupled soft phonons have been observed in xy (i.e., B2 symme-
occurs at T c2 , only the coupling between the soft mode and strains try) Raman spectra.12,68–74 Since it has been found that the over-
changes. The symmetry of the FE phase at a hydrostatic pressure of damped soft mode becomes underdamped at room temperature by
120 MPa was investigated using an elastic neutron scattering.52 The the application of hydrostatic pressure, the soft mode is not dif-
monoclinic FE structure with a symmetry P21 11 was observed below fuse but a collective, propagation excitation.69 Due to this fact and
both T c and T c2 , with a larger monoclinic distortion below T c2 than also due to coupling of the soft overdamped mode with the sec-
below T c . ond underdamped mode, the FE PT in KDP can be considered as
Chemically related to LTT is lithium ammonium tartrate a crossover between displacive and order–disorder type, which was
monohydrate (LAT). It exhibits the FE and ferroelastic PTs at T c theoretically proposed by Kobayashi in 1968.66 A similar mecha-
= 98 K. IR and Raman spectra of the LAT were investigated down nism of the FE PT has been observed also in Raman and IR spec-
to 15 K, and no soft mode was observed53 because the PT is proper tra of other KDP-type crystals as RbH2 PO4 and KH2 AsO4 and
ferroelastic and improper FE. In contrast to LTT where a substantial their deuterated analogs.12,61,62,70,72 On the other hand, NH4 H2 PO4

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-5


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

article and can be found, e.g., in the reviews of Höchli et al.75 and
Petzelt et al.76

IV. LEAD-BASED RELAXOR FERROELECTRICS


Unlike classic ferroelectrics that exhibit a peak in permittiv-
ity ε′ (T) at the temperature of PT, which is frequency independent
in displacive ferroelectrics, its height depends on frequency, but
its temperature T c does not change (in the case of order–disorder
PTs—see Fig. 1), there is another class of materials, which exhibits
a high and frequency dependent peak in ε′ (T)—see, e.g., Fig. 5.
These materials frequently do not exhibit any FE PT without an
external electric field, but since they show many similarities with
ferroelectrics, they are called relaxor ferroelectrics.
Since the discovery of the first relaxor FE PbMg1/3 Nb2/3 O3
(PMN) by Smolenskii and Agranovskaya77 in the end of 1950s, many
materials with similar dielectric properties were found. Isupov78
tried to explain the peculiar dielectric anomaly in relaxors by a
diffuse PT, i.e., by a distribution of the FE PT temperatures due
to microscopic inhomogeneities in the samples. Nevertheless, later,
it was shown that the relaxor dielectric anomaly is not necessary
accompanied by a PT. For example, model relaxor FE PMN exhibits
a non-polar cubic structure down to liquid helium temperatures79
and it becomes FE only under cooling in a bias electric field.80 In
1983, Burns and Dacol measured temperature dependences of the
optical refractive index in various relaxors and observed a devia-
tion from the linear temperature dependence some 300 K above the

30 October 2024 15:21:18


temperature T m , where ε′ (T) shows maximum.81 They explained
it by a creation of polar clusters, and since that time, the temper-
ature T d is called the Burns temperature. Nowadays, it is believed
that most of the peculiar physical properties of relaxors are some-
how connected with the polar nanoclusters, which appear probably

FIG. 4. (a) Temperature dependence of the longitudinal (full circles) and trans-
verse (open circles) optical lowest-frequency phonon frequencies of KDP polarized
along the FE c axis. The inset shows the temperature dependence of the oscillator
strength of the overdamped soft mode. (b) Temperature dependence of the relax-
ation frequencies ωR = 1/τ C in partially deuterated KDP obtained from the Debye
fit [Eq. (2)] of submillimeter spectra. Deuterium concentration x in KH2(1−x) D2x PO4
is listed in the figure. Reproduced with permission from P. Simon, F. Gervais, and
E. Courtens, Phys. Rev. B 37, 1969–1979 (1988). Copyright 1988 American Phys-
ical Society and Volkov et al., Ferroelectrics 25, 531–534 (1980). Copyright 1980
Taylor & Francis.
FIG. 5. Temperature dependence of permittivity in relaxor FE PbMg1/3 Nb2/3 O3
and NH4 H2 AsO4 are antiferroelectrics and their solid solutions depicted at various frequencies. The ε0 -EMA curve is the result of calculation
with KDP, RbH2 PO4 , and KH2 AsO4 exhibit very interesting dipolar of the phonon static ε0 using the effective medium approximation described in
Ref. 103. Reproduced with permission from Bovtun et al., J. Eur. Ceram. Soc. 26,
glass behavior. Nevertheless, the discussion of their detail structural,
2867–2875 (2006). Copyright 2006 European Ceramic Society.
dielectric, and lattice dynamics properties is beyond the scope of this

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-6


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

due to some quenched random electric fields82 from cations of dif- indicates increasing correlation among the polar clusters. The
ferent valences, which are sitting in the equivalent perovskite sites mean (τ 0 ) and upper (τ 1 ) relaxation times diverge according to the
(e.g., Mg2+ and Nb5+ in PMN); however, details are still not fully Vogel–Fulcher law,96
understood.
In 1997, Park and Shrout16 showed that relaxor-based UVF
τ0,1 (T) = τL exp( ), (8)
PbMg1/3 Nb2/3 O3 –PbTiO3 (PMN-PT) and PbZn1/3 Nb2/3 O3 –PbTiO3 T − TVF
(PZN-PT) single crystals exhibit ultrahigh strain and piezoelectric
behavior in electric field, and since that time, the interest for relax- with the same freezing temperature T VF = 230 ± 5 K for both ceram-
ors rapidly increased because these materials show many technical ics but different activation energies U VF of 1370 K and 1040 K for
applications in piezoelectric devices and microelectronics. Besides the PLZT of 8/65/35 and 9.5/65/35, respectively. τ L in Eq. (8) marks
many research papers published on relaxors, several reviews also the limiting high-temperature relaxation time, which is of the order
appeared. We can refer the reader to the reviews of Cross,83,84 Ye,85 of the reciprocal lowest optical phonon frequency. The shortest
Samara,86,87 and Bokov,88,89 where the present knowledge of relax- relaxation time τ 2 is about 10−12 s and remains almost temperature
ors is overviewed. There are also several reviews about the lattice independent [Fig. 6(a)]. Below room temperature, the loss spec-
dynamics of relaxors. The results of Raman scattering were reviewed tra ε′′ (ω) become essentially frequency independent and the per-
by Siny et al.;90 inelastic neutron scattering results were reviewed by mittivity increases linearly with decreasing logarithm of frequency
Shirane and Gehring91 and Cowley et al.;92 and IR and MW dielec- [Figs. 6(b) and 6(c)].97 This behavior was successfully explained
tric dispersion was reviewed by Kamba and Petzelt,93 Hlinka et al.,94 by a constant distribution g(τ) of equally strong Debye relaxations
Petzelt et al.,95 and Buixaderas et al.29 between upper (τ 1 ) and lower (τ 2 ) limits of relaxation times. It was
It was demonstrated that for understanding the peculiar dielec- shown that if the measured frequency range is much narrower than
tric dispersion in relaxors, a broadband dielectric spectroscopy from the distribution g(τ), the complex permittivity can be expressed as96
sub-Hz up to THz and IR frequencies is necessary. First of all, let
us describe the properties of lanthanum modified lead zirconate ε′ (ω) = ε∞ − B(T) ln(ωτ2 ),
titanate (Pb1−x Lax )(Zry Ti1−y )1−x/4 O3 (PLZT 100(x/y/1−y)) with (9)
π
x = 8 and 9.5 and y = 65.96 Both ceramics exhibit broad and strongly ε′′ (ω) = B(T),
frequency dependent peaks in ε′ (T) and ε′′ (T) seen between 325 K 2
(at 100 Hz) and 500 K (at 36 GHz). A single broad and symmetric where B(T) is the frequency-independent but temperature-

30 October 2024 15:21:18


dispersion that occurs below the polar phonon frequencies was dependent parameter describing the magnitude of g(τ). This model
fitted with the Cole–Cole formula and with the uniform distri- was successfully used not only for the description of the low-
bution of Debye relaxations. On decreasing the temperature, the temperature dielectric properties of relaxor PLZT96,97 [Fig. 6(b)],
distribution of relaxation times becomes extremely broad (10 orders PMN,100 and PbMg1/3 Ta2/3 O3 (PMT)98 but also for fitting of the
of magnitude at 300 K in PLZT 9.5/65/35—see Fig. 4, left), which anomalous broad dielectric relaxation in Bi1.5 Zn1.0 Nb1.5 O7 with a

FIG. 6. (a) Temperature dependence of the mean (τ 0 ), upper (τ 1 ), and lower (τ 2 ) relaxation times in relaxor PLZT determined from the uniform distribution model of Debye
relaxation frequencies. The frequency dependence of the (b) real and (c) imaginary parts of the complex permittivity in PLZT 8/65/35 at various temperatures close and
below freezing temperature T VF . Reproduced with permission from Kamba et al., J. Phys.: Condens. Matter 12, 497–519 (2000). Copyright 2000 IOP Publishing and
Rychetský et al., J. Phys.: Condens. Matter 15, 6017–6030 (2003). Copyright 2003 IOP Publishing.

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-7


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

cubic pyrochlore structure,99 which is not a relaxor FE. In this case, polar nanoregions is highly improbable.107 Hlinka et al. explained
the upper relaxation time follows the Arrhenius law and the relax- the phonon waterfall phenomena by a coupling of the acoustic and
ation was assigned to the local hopping of atoms in the A and O′ optical phonon branches. Observation of different qwf in various
positions of the pyrochlore structure among several local potential Brillouin zones was explained by different structure factors.107
minima. Nevertheless, below T VF and above T d , a soft optical phonon
The archetypical and best-studied relaxor FE is PMN because was observed in inelastic neutron scattering spectra of PMN with
its solid solution with PbTiO3 is material with the highest piezo- extrapolated critical temperature around T ∗ = 400 K.104 The same
electric coefficients.16 The typical relaxor peak in ε′ (T) measured optical soft mode was also observed in the IR spectra, but in con-
between 3 mHz and 420 GHz is shown in Fig. 5.100 The frequency trast to the inelastic neutron scattering spectra, the underdamped
plots of the dielectric function (Fig. 7) reveal similar behavior as in soft phonon was observed at all temperatures (including tempera-
PLZT: a slowing down and broadening of the dielectric relaxation tures between T VF and T d ) not only in PMN (Fig. 8)108 but also in
on cooling.101 Again, in the logarithmic scale, linearly dependent PLZT,96 PMT,98 and PbSc1/2 Ta1/2 O3 .109 Earlier spectroscopic data108
ε′ (ω) and frequency independent loss spectra ε′′ (ω) are seen below have shown that below 400 K, the soft phonon in PMN follows the
the freezing temperature T VF ≅ 230 K. Qualitatively similar dielec- Cochran law with the extrapolated critical temperature correspond-
tric behavior was observed in all relaxors, and it was interpreted ing to the Burns temperature T d [Fig. 8(a)]. Nevertheless, at higher
as follows:102 The dielectric relaxation appears in the spectra below temperatures, the soft mode hardens slower than the Curie–Weiss
the Burns temperature (T d = 600 K–700 K in various relaxors) and law and the Lyddane–Sachs–Teller relation describing the experi-
expresses a dynamical flipping of the cluster polarization. Below the mental ε′ (T) required. Moreover, an additional overdamped cen-
freezing temperature T VF , the clusters freeze out and the flipping of tral mode was observed below T d [Fig. 8(a)], whose activation was
polarization disappears. Nevertheless, the polar cluster boundaries explained by the splitting of the soft mode due to a local breaking
can still vibrate (breeze) below T VF , but since it is a temperature of cubic symmetry in polar clusters.108 Based on inelastic neutron
activated process over energy barriers with some distribution, the scattering experiments, Vakhrushev and Shappiro proposed the soft
relaxation times (or frequencies) exhibit also a distribution g(τ), mode splitting also above T d , and they claimed that both compo-
which broadens on cooling. This low-temperature g(τ) is responsible nents exhibit minimum near T d , but their phonon parameters had
for the dielectric dispersion observed below T VF (e.g., below 200 K high error bars due to the high phonon damping above T d .
in Fig. 7). The recent detailed analysis of the dielectric dispersion For that reason, new detailed THz and IR studies of relaxor
between T VF and T d revealed two relaxations, which were well fit- PMN and PMT have been performed up to 900 K and the spectra

30 October 2024 15:21:18


ted using two Cole–Cole formulas that express the flipping and were fitted using the Bruggeman effective medium formula.103,110
breezing of polar nanoregions.103 The former one disappears below Six polar phonons observed at 900 K gave evidence about the non-
T VF (τ 1 diverges), and the latter one appears below T ∗ , where the cubic local structure above T d since the cubic perovskite structure
static polar nanoclusters start to appear (depending on the material, allows only three polar phonons in the IR spectra. The E-component
T ∗ = 350 K–450 K) and its g(τ) anomalously broadens below the of the soft mode polarized perpendicularly to local polarization
freezing temperature, as explained above. (P∣∣[111]cub ) in polar nanodomains was clearly observed. This mode
Phonons in relaxor ferroelectrics were also frequently inves- is overdamped at all the temperatures, while the A1 component is
tigated. Inelastic neutron scattering spectra revealed a drop down underdamped at all temperatures. Both soft phonons exhibit clear
of the soft optical phonon branch and its merging with an acoustic shifts with temperature. The E mode component exhibits softening
phonon branch at the critical wave vector qwf and its disappearance toward T ∗ = 350 K and 400 K in PMT and PMN, respectively, and
below qwf at temperatures between T VF and T d .104 This effect was hardening below this temperature [Fig. 8(b)].
called a phonon waterfall, and it was at first explained by a phonon The static permittivity ε′0– calculated from the E mode contri-
scattering on the polar nanocluster boundaries.105,106 Later, inelas- butions follows the Curie–Weiss behavior above 400 K [Fig. 8(b)],
tic neutron scattering studies found that the critical wave vector qwf , which corresponds to previous observations above T d .111 The
below which the phonon waterfall appears, depends on the choice of deviation from the Curie–Weiss dependence, previously observed
the Brillouin zone, and therefore, the relation of qwf to the size of the below T d in the low- and radio frequency dielectric data,111 can be

FIG. 7. Frequency dependence of real and imaginary


parts of the dielectric function in PbMg1/3 Nb2/3 O3 taken at
various temperatures. Reproduced with permission from
Bovtun et al., Ferroelectrics 298, 23–30 (2004). Copyright
2004 Taylor & Francis.

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-8


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

a cubic structure (Pm3m). On decreasing the temperature, the PT


to FE tetragonal phase (P4mm) occurs at ∼400 K. The transition
is of the first order, predominantly displacive with many features
involving also an order–disorder mechanism.112–115 In the paraelec-
tric phase, signatures of polar nanoregions typical for relaxors were
revealed.116,117 On further decreasing the temperature, another first-
order FE PT to orthorhombic phase (Amm2) occurs at ∼280 K, and
finally, the rhombohedral structure (R3m) appears below ∼180 K.118
The corresponding spontaneous polarization Ps changes stepwise
from (001) to (110) and (111) directions, respectively. Temperature
and broadband-frequency dependencies of permittivity are shown
in Fig. 9.
First, far IR studies of the FE soft mode were published by Bal-
lantyne119 and Barker120 already at the beginning of the 1960s. Both
authors observed an overdamped optical soft mode below 200 ○ C,
which followed Eq. (6) typical for disordered systems. Luspin
et al.27,121 measured the IR reflectivity up to 1350 K and found that
the soft mode frequency follows the Cochran law [Eq. (4)] above
∼520 K, but below this temperature, the softening ceases with ωSM
∼ 63 cm−1 and the soft-mode strength saturates at Δε ≈ 600. Since
the static permittivity is one order of magnitude higher near T c
(see Fig. 9), some additional soft excitation must exist below the
saturated soft mode.122 This is the typical behavior at a crossover
from displacive to order–disorder phase transition.27 The over-
damped soft excitation was really revealed in hyper-Raman123 and
submillimeter12 spectra, and in combination with newer THz and
IR spectra,124,125 the softening toward T c of the overdamped lower-

30 October 2024 15:21:18


frequency excitation was revealed [marked in Fig. 10(b) as a CM]
and also supported by first-principles-based calculations.124 The
split of the soft mode in the paraelectric phase proofs a non-cubic
local symmetry (polar nanoregions) above T c1 . This observation
was also supported by a nonlinear temperature dependence of the
index of refraction116 and by acoustic emission measurements.117
The local polarization in the paraelectric phase is usually modeled

FIG. 8. (a) Temperature dependences of the polar phonon frequencies in PMN.


Solid points mark data obtained on the PMN thin film, and open circles and squares
are soft mode frequencies from inelastic neutron scattering and IR reflectivity of the
PMN single crystal. (b) Temperature dependence of the reciprocal static permittiv-
ity 1/ε0– (solid squares) obtained from the EMA fit of E-component IR and THz
spectra of PMT. Solid lines are the results of the Curie–Weiss fit. Green solid cir-
cles mark the overdamped E component of the soft mode frequency characterized
by a relaxation frequency ωR = ω2E /γE . Inset: 1/ε0– of PMT compared with that of
PMN. Reproduced with permission from Kamba et al., J. Phys.: Condens. Matter
17, 3965–3974 (2005). Copyright 2005 IOP Publishing and Nuzhnyy et al., Appl.
Phys. Lett. 114, 182901 (2019). Copyright 2019 AIP Publishing LLC.

explained by the appearance of dielectric relaxations at lower fre-


quencies,103 which mask the high-frequency dielectric peak at T ∗
(see the EMA peak calculated using an effective media approxi-
mation in Fig. 5). The E mode softening toward T ∗ is a hint of a
local symmetry-lowering of the crystal structure within the polar
nanodomains.
FIG. 9. Temperature dependencies of the dielectric permittivity (a) and loss (b)
in BaTiO3 ceramics plotted at selected frequencies. Broadband spectra of dielec-
V. DYNAMICS OF FERROELECTRIC PHASE tric permittivity (c) and loss (d) in BaTiO3 ceramics. Soft and central modes are
TRANSITIONS IN BaTiO3 marked together with dielectric relaxations R1 and R2. Reproduced with permis-
sion from Bovtun et al., Phys. Rev. Mater. 5, 014404 (2021). Copyright 2021
BaTiO3 (BT) is the most extensively studied FE material with
American Physical Society.
many actual and potential applications. Its paraelectric phase has

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-9


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

FIG. 10. Schematic diagrams of the model of Ti off-centering creating spontaneous polarization. Models of (a) the rhombohedral, (b) the orthorhombic, (c) the tetragonal,
and (d) the cubic phase. (e) Temperature dependence and assignment of the TO polar mode frequencies in BaTiO3 coarse-grain ceramics. CM marks the soft central
mode. Reproduced with permission from K. Tsuda, R. Sano, and M. Tanaka, Phys. Rev. B 86, 214106 (2012). Copyright 2012 American Physical Society and Petzelt et al.,
Ferroelectrics 469, 14–25 (2014). Copyright 2014 Taylor & Francis.

30 October 2024 15:21:18


by a dynamic disorder of Ti cations among eight equivalent positions incipient ferroelectrics are SrTiO3 and KTaO3 , and because their
[Figs. 10(a)–10(d)].126 Four and two equivalent positions are statisti- quantum paraelectric phases are very sensitive on any perturbations
cally occupied by Ti in the FE tetragonal and orthorhombic phases, such as doping, strain, or electric field, the FE phase can be easily
respectively, while only one position causes FE polarization along induced in these systems.86
the [111] direction in the rhombohedral phase [see Fig. 10(a)]. The Choice of the substrate is the key factor for the strain as the lat-
dynamic disorder but simultaneously displacive character of the PT tice mismatch leads to tensile or compressive strain in thin films. In
from the paraelectric to FE phase in BaTiO3 was confirmed by syn- 2004, it was proved that 1% tensile strain caused by the DyScO3 sub-
chrotron x-ray diffuse scattering and its modeling using molecular strate can induce a FE PT near 300 K in SrTiO3 .128 It was shown that
dynamics simulations.127 the PT is driven by the in-plane polarized FE soft mode (Fig. 11),
As mentioned above, the soft phonon anomalies cannot com- i.e., it is of the displacive type.129,130 The soft mode frequency
pletely explain huge dielectric anomalies near the PT temperatures strongly hardens in the external electric field, and therefore, the per-
shown in Fig. 9. Broadband dielectric spectra reveal two additional mittivity is highly tunable using the field.131 On the other hand,
dielectric relaxations R1 and R2 in the MW and radio frequency the ∼1% compressive strain caused by NdGaO3 132 or (LaAlO3 )0.29 -
regions.122 The R1 relaxation undergoes a critical slowing down to (SrAl1/2 Ta1/2 O3 )0.71 (LSAT)133 substrates caused dramatic stiffening
T c1 , which correlates with the increase in its dielectric contribution of the in-plane polarized soft-mode frequency (Fig. 11), so the FE
ΔεR1 and is, in fact, mainly responsible for the Curie–Weiss law close PT cannot be revealed from the IR spectra. Nevertheless, the soft
to T c1 . The R1 frequency ωR1 above T c1 sharply increases on heating phonon polarized normal to the film plane (unfortunately, not active
and merges with the central mode. Its temperature dependence can in the near-normal IR reflectance geometry) should induce the FE
be well fitted with a combination of the Arrhenius law and critical PT in compressively strained SrTiO3 grown on a LSAT substrate
slowing-down of the pre-exponential frequency.122 because the FE polarization P∣∣c was revealed in second harmonic
generation studies.134 The influence of various substrates, film thick-
nesses, grain size, and their boundaries in films and ceramics of
VI. STRAIN-INDUCED FERROELECTRIC PHASE SrTiO3 on phonons was reviewed in Ref. 135.
TRANSITIONS IN INCIPIENT FERROELECTRICS SrTiO3 is the n = ∞ member of the Srn+1 Tin O3n+1 homologous
Incipient ferroelectrics are paraelectric materials exhibiting a series. Bulk Srn+1 Tin O3n+1 crystallizing in the Ruddlesden–Popper
large increase in the permittivity on cooling due to a phonon soft- structure [Fig. 12(a)] exhibits incipient FE behavior, and its per-
ening as follows from the Lyddane–Sachs–Teller relations [Eq. (5)]. mittivity increases with the number of perovskite layers n.136 In
The permittivity usually saturates at low temperatures, but the 1% tensile strained Srn+1 Tin O3n+1 films grown on DyScO3 , FE PTs
FE PT does not occur due to quantum fluctuations; hence, these were observed in samples with n ≥ 3 with T c increasing with the
materials are also called quantum paraelectrics. The most famous rising number of perovskite layers n.137 THz transmission and IR

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-10


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

FIG. 11. Temperature dependences of the soft mode frequencies in the SrTiO3
crystal and strained films grown on various substrates. Compressively strained
films grown on NdGaO3 or LSAT substrates exhibit stiffening of the soft mode with
respect to the crystal, while tensile strained SrTiO3 /DyScO3 exhibits a soft mode
anomaly and appearance of the central mode near 300 K due to a high lattice
anharmonicity near strain-induced FE PT. Below ∼180 K, two new modes activate
above 120 cm−1 in this film due to the antiferrodistortive PT within the ferroelectric
phase. Reproduced with permission from J. Petzelt and S. Kamba, Ferroelectrics
503, 19–44 (2016). Copyright 2016 Taylor & Francis.

reflectivity spectra reveal the FE soft mode, which drives the FE PT,

30 October 2024 15:21:18


but near T c , the dielectric strength of a central mode dominates.138
The n = 6 film (Sr7 Ti6 O19 ) shows T c = 180 K, high electric-field tun-
ability of permittivity, and exceptionally low dielectric loss at 300 K,
which overcomes all the known tunable MW dielectrics.137 A
detailed analysis of the MW, IR, and THz spectra shows that the low
room-temperature MW loss in this film has the origin in the absence
of the central mode at 300 K [Fig. 12(b)], while the high tunability
of permittivity is due to the sensitivity of the soft mode to the elec-
tric field.138 The strain quickly relaxes with increasing Srn+1 Tin O3n+1
film thickness, which limits its application. Recently, it was proved
that if one SrTiO3 layer in Srn+1 Tin O3n+1 is replaced by BaTiO3 , the
chemical pressure reduces the strain, thicker films can be grown, and
simultaneously, excellent MW properties are kept.139 Today’s best FIG. 12. (a) Schematics of the crystal structure of a unit cell of the n = 1–6
room-temperature MW tunable dielectric parameters were achieved and n = ∞ members of the Srn+1 Tin O3n+1 homologous series. (b) Complex
dielectric spectra of a strained Sr7 Ti6 O19 thin film measured at different temper-
in films of Sr6 BaTi6 O19 , where unprecedented low loss at frequencies atures. The points are experimental data between 3 MHz and 40 GHz. The curves
up to 125 GHz was observed.139 This makes this film very attractive show the results of the fits to the experimental MW and far IR spectra. At room
for high-frequency applications in 5G telecommunications. temperature, the MW loss is low because it is determined only by the phonon
The second most popular incipient FE is KTaO3 , where only contributions [Eq. (2)], while at lower temperatures, a dielectric relaxation (cen-
two publications were devoted to the IR and THz SM studies,140,141 tral mode) enhances the MW loss. Reproduced with permission from Goian et al.,
but the soft mode was also studied in detail using the hyper-Raman Phys. Rev. B 90, 174105 (2014). Copyright 2014 American Physical Society.
scattering.20,142 In Ref. 141, the chemical solution deposited poly-
crystalline (grain size ∼ 160 nm) film (d ≈ 200 nm) on the (0001)
sapphire substrate, partially (001) oriented, was shown to be FE with VII. SOFT-MODE SPECTROSCOPY IN MULTIFERROICS
a diffuse T c ≈ 60 K. The reason for it was shown to be the com- WITH DISPLACIVE FERROELECTRIC PHASE
pressive strain due to the larger thermal expansion of the sapphire TRANSITIONS
compared to KTaO3 (the film was annealed at 900 ○ C), which could Multiferroics are materials exhibiting magnetic and FE order
induce a FE phase with preferential out-of-plane orientation of Ps . simultaneously. Since the magnetization can be controlled by the
Despite it, a minimum in the frequency of the in-plane polarized electric field (i.e., magnetoelectric effect) in these materials, they
soft mode was observed near T c ,141 while in crystal and ceramics, have high promising potential for applications in spintronics and
the soft mode gradually softens on cooling and levels off at ∼20 cm−1 nonvolatile memories, and therefore, they are intensively studied.
below 20 K.20,143 In this chapter, we will focus on type-I multiferroics, where FE

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-11


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

FIG. 13. (a) Complex dielectric spectra


of BiFeO3 ceramics at selected temper-
atures. The dots are experimental THz
data; lines are the results of the IR reflec-
tivity fits. Note the shifts of ε′′ peaks
to lower frequencies due to the phonon
softening on heating. (b) Temperature
dependence of the electromagnon fre-
quencies in BiFeO3 determined from
the fits of THz spectra compared with
published frequencies obtained from
Raman scattering and far IR spec-
tra. Reproduced with permission from
Kamba et al., Phys. Rev. B 75, 024403
(2007). Copyright 2007 American Phys-
ical Society and Skiadopoulou et al.,
Phys. Rev. B 91, 174108 (2015). Copy-
right 2015 American Physical Society.

and magnetic phases appear independently, and therefore, their can be called electromagnons. They are activated in the spectra from
magnetoelectric coupling is relatively small. outside of the Brillouin zone center due to the incommensurate
The most studied multiferroic is BiFeO3 because it is one of the modulation of spins in the antiferromagnetic phase.152 The electro-
few single-phase room-temperature multiferroics with high FE T c magnons are activated due to a dynamic magnetoelectric coupling,

30 October 2024 15:21:18


= 1120 K and antiferromagnetic Néel temperature T N = 643 K.144 and they give rise to unidirectional light propagation in the THz
Unfortunately, its magnetoelectric coupling is weak and the 1.2% region.151
change in its permittivity with the magnetic field (magnetodielec- Since there are few multiferroics in nature and they have mostly
tric effect) observed above 200 K is mainly caused by a combi- only small magnetoelectric coupling, Fennie and Rabe153 suggested
nation of magnetoresistance influencing related Maxwell–Wagner theoretically a new route for preparation of multiferroics with a
dielectric relaxation in electrically leaky samples.145 In the insu- strong magnetoelectric coupling. They proposed to use a biaxial
lating phase (i.e., below 200 K), the magnetodielectric effect drops strain in epitaxial thin films for inducing the FE and ferromagnetic
down below 0.3% in the magnetic field of 9 T. The effect is small states in materials, which are in the bulk form paraelectric and anti-
because the linear magnetoelectric effect is forbidden in incommen- ferromagnetic. The basic condition for such materials is the strong
surately modulated cycloidal magnetic structures. Only above 19 T, spin–phonon coupling. Based on their first-principles calculations,
the antiferromagnetic structure becomes unmodulated and the lin- Fennie and Rabe153 proposed to use EuTiO3 because this mate-
ear magnetoelectric effect is allowed.146 rial is in its bulk form an incipient FE and its permittivity strongly
IR reflectivity and THz transmission spectra revealed 13 polar changes at the antiferromagnetic PT due to the strong spin–phonon
phonons, which exactly correspond to the rhombohedral R3c sym- coupling.154 The temperature dependence of the permittivity in the
metry.145 On heating, the phonons broaden and gradually merge EuTiO3 crystal was successfully explained by the strongly anhar-
in three bands allowed in the cubic perovskite Pm3m structure monic behavior of the soft mode.155,156 The bulk EuTiO3 ceram-
expected in the paraelectric phase. The spectra taken up to 900 K ics undergoes an antiferrodistortive PT from the high-temperature
show an optical soft mode, which is responsible for an increase in cubic Pm3m (Z = 1) structure to the tetragonal I4/mcm (Z = 1)
the low-frequency permittivity with rising temperature [Fig. 13(a)]. phase near 280 K.157,158 For that reason, the soft mode splits in the
Phonon softening is only partial, suggesting that the phase transition tetragonal phase and both components of the soft mode soften on
into the paraelectric phase will be of the first order.145 It is con- cooling. Lee et al.159 grew the epitaxial EuTiO3 thin films on DyScO3
sistent with the theoretical prediction147 of the first-order PT into substrates, and due to the 1% tensile strain, the soft mode (seen
intermediate antiferrodistortive tetragonal phase with the I4/mcm in the IR reflectance) is much softer than in the bulk, reaching a
space group. Very recent perturbed angular correlation experimen- minimum frequency at T c = 250 K [Fig. 14(a)]. Due to the soft
tal studies148 revealed the first-order PT but into the orthorhombic mode anomaly, the permittivity exhibits a maximum [Fig. 14(b)],
Pbnm paraelectric phase with a twice smaller unit cell, i.e., the PT typical for the displacive FE PTs. FE polarization was calculated
is improper FE. The cubic Pm3m phase appears near 1200 K, but from the first principles159 and also estimated based on the value of
already higher than 10 K, the crystal melts.149 T c 160 with the resulting value of 20 μC/cm2 –30 μC/cm2 . The non-
THz and Raman spectra revealed the number of spin excita- centrosymmetric structure of the thin film has also been proven
tions, whose frequencies soften toward T N [see Fig. 13(b)].150 These by the observation of the second harmonic generation signal below
magnons are electrically and magnetically active151 and therefore T c . Magneto-optic Kerr effect studies revealed a strain-induced

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-12


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

on LSAT. Very recently, the same soft-mode frequency shift with


the magnetic field was revealed also in the THz spectra of EuTiO3
ceramics.162 Moreover, a shift of ferromagnetic resonance from the
microwave into THz region was discovered if the magnetic compo-
nent of THz radiation was polarized perpendicularly to the external
magnetic field.162
A FE PT can be induced also in (Eu1−x Bax )TiO3 solid solu-
tion.163 (Eu0.5 Ba0.5 )TiO3 and (Eu0.5 Ba0.25 Sr0.25 )TiO3 exhibit dis-
placive FE PTs at 215 K and 130 K, respectively, because the FE soft
modes were observed in THz and IR spectra.164,165 In both mate-
rials, the Néel temperature T N remains close below 2 K. Since T N
is very low and the internal electric field is of order of 10 MV/cm
in the FE phase and this field (and the FE polarization) can
be switched at low temperatures by an external electric field of
3–10 kV/cm−1 , the system was suggested for measurement of an
electron dipole moment, which could be revealed in observation of
the linear magnetoelectric effect in the paramagnetic state.163 This
effect is from symmetry reasons forbidden in the paramagnetic state,
but if the linear magnetoelectric effect would be observed, it could
be explained by a non-zero electron dipole moment that violates
both space-inversion and time-reversal symmetry. These are basic
conditions for the linear magnetoelectric effect. A change in magne-
tization with the electric field was really observed in (Eu0.5 Ba0.5 )TiO3
at 4 K,166 but since the sample heats in the field above the coercive
field of 10 kV/cm, the effect could be explained by a small tempera-
ture change. Nevertheless, the measurement allowed us to determine
that the electron dipole moment should be smaller than 6 × 10−25 e

30 October 2024 15:21:18


cm (e marks the charge of electrons).166 (Eu0.5 Ba0.25 Sr0.25 )TiO3 could
be more suitable for such studies because its coercive field is three
times lower so that the sample heating could be smaller.165
Ten years ago, Bousquet et al. used density functional the-
ory and predicted a strain-induced displacive FE PT in the ferro-
FIG. 14. (a) Temperature dependence of the lowest-frequency phonons and (b) magnetic EuO, but since its optical phonon is highly stable, the
static phonon permittivity in the EuTiO3 film grown on DyScO3 . Soft mode anomaly critical strain was predicted to be around 4%.167 Researches from
induces dielectric anomaly at T c , and two new modes activate in the IR spectra
below 150 K due to an antiferrodistortive PT at T AFD . Reproduced with permission
Cornell University grew strained EuO thin films using molecular-
from Lee et al., Nature 466, 954–959 (2010). Copyright 2010 Springer Nature. beam epitaxy on various substrates to impose different epitaxial
strains, and the Prague group investigated the phonon properties
of the strained EuO films using IR reflectance spectroscopy. It was
found that although the phonon frequency decreased with ten-
ferromagnetic state below 4.3 K. All these studies159 confirmed the sile strain as predicted, the previously predicted critical 4% tensile
theoretical predictions that the strain can induce FE and ferromag- strain was insufficient to induce a FE phonon instability.168 Hear-
netic order in paraelectric antiferromagnets. Nevertheless, the mag- ing this news, Bousquet performed new calculations using a more
netic Curie temperature T c = 4.3 K is too low for practical applica- advanced hybrid functional and found that more than 5.8% ten-
tions, despite the FE critical temperature close to room temperature, sile strain is required for the phonon instability and resulting FE
which can even be enhanced by a higher strain. state.168 Biaxially straining EuO more than 5% is challenging as such
Permittivity in EuTiO3 dramatically changes with the magnetic films tend to relax after the epitaxial growth of only a few mono-
field.154 Since the high value of permittivity is caused by the soft layers. This challenge was overcome by growing superlattices made
mode contribution, it is natural to expect that the soft mode fre- of EuO and BaO. The latter material has a larger lattice parame-
quency will change with the magnetic field. The first attempt to see ter than EuO and therefore helps maintain the tensile strain in the
the soft mode tuning in the magnetic field in ceramics was not suc- EuO thin layers. In (EuO)2 /(BaO)2 superlattices grown epitaxially
cessful due to a very broad reflectivity band.155 Nevertheless, in com- on LSAT substrates, 6.4% biaxial tensile strain in the EuO layers
pressively strained EuTiO3 films grown on LSAT substrates, the soft was achieved and a FE soft mode anomaly was observed at 100 K
phonon is significantly stiffened above 100 cm−1 and shows a rather [Fig. 15(a)].168 It induces a peak in the dielectric permittivity typical
narrow reflectance peak, whose change with the magnetic field and of a FE PT [see Fig. 15(b)]. Magnetic measurements simultaneously
temperature can be studied with much higher accuracy than in the proved that the superlattices remain ferromagnetic, i.e., they are
bulk samples.161 Actually, reliable tuning of the soft mode frequency multiferroic.168
by 2 cm−1 was detected in the IR reflectance spectra,161 explaining High ferromagnetic and FE critical temperatures were theoret-
in this way the magnetodielectric effect in the EuTiO3 film grown ically predicted in the strained SrMnO3 thin films.169 The prediction

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-13


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

FIG. 15. (a) Temperature dependence of the EuO optical phonon frequency in epitaxial (001)-oriented EuO films with various nominal strain levels imposed by underlying
(001) YSZ, (110) YAlO3 , (001) Si, and (001) LSAT substrates. +5.6% and +6.4% nominal tensile strain was reached in [(EuO)2 /(BaO)6 ]28 /Si and [(EuO)2 /(BaO)2 ]35 /LSAT
superlattices, respectively. The EuO phonon in the superlattice with the highest strain (+6.4%) exhibits softening typical for a displacive FE PT. Upon undergoing the PT,
the single optical phonon splits into three phonons due to the reduction of the symmetry of the (EuO)2 unit cell that occurs below T c . (b) Schematic illustration of the
crystal structure of the (EuO)2 /(BaO)2 superlattice on an (001) LSAT perovskite substrate in which the (EuO)2 layer is under biaxial tension and the (BaO)2 layer is under
biaxial compression. The directions of both strains are marked by arrows. Eu, Ba, and O atoms are shown in blue, green, and red, respectively. (c) Temperature and strain
dependence of the static relative permittivity of the EuO films and EuO layers in the (EuO)x /(BaO)y superlattices. The permittivity values were calculated from fits of the IR
reflectance. The FE critical temperature observed for the [(EuO)2 /(BaO)2 ]35 superlattice strained to LSAT is marked by the arrow. Reproduced with permission from Goian
et al., Commun. Mater. 1, 74 (2020). Copyright 2020 Springer Nature.

30 October 2024 15:21:18


was supported by extremely strong spin–phonon coupling in this of several strained SrMnO3 thin films but did not obtain reliable
material, which was observed near T N in the IR spectra of SrMnO3 phonon parameters, presumably due to the high phonon damp-
ceramics.170 The film with 1.7% tensile strain grown on the LSAT ing in this material. Ba substitution of Sr induces negative chemi-
substrate exhibited FE T c ≈ 400 K and Néel temperature T N cal pressure in the lattice, and therefore, the FE PT was observed
< 200 K.171 Nevertheless, these films were much conducting, which in Sr1−x Bax MnO3 crystals172 and ceramics173 near 350 K. The soft
did not allow us to study the magnetoelectric coupling. Since the mode exhibits minimum at this temperature173,174 and splits in the
FE PT should be of displacive type, we investigated the IR spectra FE phase due to the change of symmetry from the cubic to tetragonal

FIG. 16. (a) Temperature depen-


dence of the phonon frequencies in
Sr0.55 Ba0.45 MnO3 ceramics obtained
from the fits of IR and THz spectra. The
area corresponding to T c is hatched.
Note the soft mode minimum near T c
and anomalies of the two soft mode
components near T N . (b) Complex
dielectric spectra obtained from the fits
of IR reflectivity and THz permittivity
spectra at various temperatures. The
experimental THz data are marked by
dots. Reproduced with permission from
Goian et al., J. Phys.: Condens. Matter
28, 175901 (2016). Copyright 2016 IOP
Publishing.

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-14


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

30 October 2024 15:21:18


FIG. 17. Low-temperature crystallographic structure of 2H–BaMnO3 featuring face-sharing MnO6 octahedrons (space group P63 cm) from (a) side view and (b) top view.
(c) Temperature dependence of dielectric permittivity of 2H–BaMnO3 ceramics measured at various frequencies and in or without the bias electric field. (d) Real and (e)
imaginary parts of complex THz dielectric permittivity showing activation and hardening of a new soft phonon below T c = 130 K. (f) Temperature dependence of the soft
mode frequency obtained from the THz and Raman spectra. The dashed line is the result of the fit using the Cochran law. Reproduced with permission from J. Varignon and
P. Ghosez, Phys. Rev. B 87, 140403(R) (2013). Copyright 2013 American Physical Society and Kamba et al., Phys. Rev. B 95, 174103 (2017). Copyright 2017 American
Physical Society.

phase.173 Both modes exhibit also other anomalies at T N due to the down and the distribution of relaxation frequencies of domain wall
strong spin–phonon coupling (Fig. 16). vibrations broadens. This explains the relaxor-like dielectric disper-
Many type-I multiferroics exhibit improper FE PT connected sion seen below 60 K in Fig. 17(c). The contribution of domain wall
with a multiplication of unit cell below T c . YMnO3 175 and also the motion in ε′ reduces under the bias electric field [Fig. 17(c)] because
famous BiFeO3 147 belong to this family. We investigated a two- the FE domain size increases and domain wall concentration reduces
layered BaMnO3 , which at high temperatures crystallizes in the non- under the field. An antiferromagnetic order appears at 53 K, but
polar hexagonal P63 /mmc structure and its unit cell triples below short-range magnetic correlations were observed in the EPR spectra
T c = 130 K and the space group changes to P63 cm.176,177 Unlike in up to 230 K.176
all proper ferroelectrics, where a peak in ε′ (T) is seen at T c , only
a change in slope in ε′ (T) was observed at T c [Fig. 17(c)] because
the PT is driven by a soft Brillouin zone-boundary phonon of K 3 VIII. SPIN-INDUCED MULTIFERROICS
symmetry with a wave vector q = (1/3, 1/3, 0). This IR inactive In 2003, a new class of multiferroics was discovered in
phonon is coupled with the hard zone-center Γ 2 − mode. Owing to TbMnO3 . Kimura et al.178 observed a narrow and small peak in
the Brillouin-zone folding below T c , the soft mode becomes a zone- ε′ (T) at the temperature, where the cycloidal magnetic structure
center mode in the FE phase, which activates in the IR and Raman of TbMnO3 changes its modulation. Its FE polarization was 3 or 4
spectra and hardens on cooling. This phonon has a small dielec- orders of magnitudes lower than in classical ferroelectrics, but since
tric strength, so it only partially explains the increase in ε′ below T c the polarization direction was possible to control by the direction
(Fig. 17).176 An additional dielectric relaxation, observed in the kHz of the external magnetic field, the authors explained it as caused by
and microwave region, is probably due to the vibration of the FE the spin interactions. Due to this fact, the polarization is more sensi-
domain walls. On cooling, this temperature-activated process slows tive to the magnetic field and the magnetoelectric coupling is huge.

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-15


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

Shortly after this observation, many new spin-induced ferroelectrics


were discovered, called now type-II multiferroics.179 Since the order
parameter of the FE PT transition is not polarization, these multi-
ferroics belong to improper or pseudoproper ferroelectrics.180 For
that reason, one cannot expect FE soft phonons or critical dielectric
relaxation in these materials, and therefore, the dielectric anoma-
lies are small at T c . Nevertheless, in one paper, a critical relaxation
in multiferroic MnWO4 was reported, but it was observed in the
MW region only less than 1 K above the T c .181 This soft dielectric
excitation was explained by an electromagnon181 coupled with the
polar phonon. Just this excitation, activated in the dielectric spec-
tra in a close proximity to the FE PT, is responsible for the tiny
dielectric anomaly at T c in MnWO4 and most probably also in other
spin-induced ferroelectrics.
The electromagnons (i.e., electrically active magnons) were first
time discovered by Pimenov et al. in THz spectra of TbMnO3 and
DyMnO3 ,182 and since that time, they were observed in many mul-
tiferroics.179,183–185 The characteristic feature of the electromagnon
is its activity in the dielectric THz or MW spectra and its cou-
pling with polar phonons. It means that it receives its dielec-
tric strength from the IR active phonons. Due to off-diagonal
dynamic magnetoelectric coupling, THz or MW absorption of
electromagnons exhibits a directional dichroism, i.e., the absorp-
tion depends on the propagation direction of the electromag-
netic radiation.179 The electromagnons were observed in multifer-
roic perovskite manganites,182–184 akermanites,186,187 BiFeO3 ,150–152
ε phase of Fe2 O3 ,188 Ni3 TeO6 ,189 Ni3−x Cox TeO6 ,190 quadruple

30 October 2024 15:21:18


perovskites CaMn7 O12 191 and SrMn7 O12 ,192 Y-type hexaferrites
FIG. 18. (a) Temperature dependence of the Raman scattering spectra showing
BaSrCoZnFe11 AlO22 ,193,194 and Ba2 Mg2 Fe12 O22 ,195,196 Z-type hexa- the electromagnon in the Y-type hexaferrite BaSrZnCoFe11 AlO22 . (b) Magnetic
ferrite (Bax Sr1−x )3 Co2 Fe24 O41 ,197,198 and others.179 field dependence of the extinction coefficient of the Y-type hexaferrite in the THz
Interestingly, the Y-type hexaferrite exhibits an asymmetric region measured at 150 K. In the field above 4 T, the electromagnon doublet disap-
change in static FE polarization P with magnetic field H (P(H) pears and a ferromagnetic resonance (FMR) appears in the low-frequency part of
= −P(−H)), whereas the Z-type hexaferrite shows a symmetric the spectra. Reproduced with permission from Vít et al., Phys. Rev. B 97, 134406
change (P(H) = P(−H)). These facts were clarified recently: the static (2018). Copyright 2018 American Physical Society.
magnetoelectric coupling in Y-type hexaferrites can be explained
only by the inverse Dzyaloshinskii–Moriya interaction, while addi-
tional p-d hybridization becomes dominant in the Z-type hexafer- resonance (FMR), which linearly shifts from the MW to THz region
rite.199 The magnetoelectric coupling is giant200,201 because the mag- with rising magnetic field.193
netic structures of hexaferrites are extremely sensitive to the external Recent MW studies of the magnetic permeability revealed the
magnetic field. THz and Raman spectroscopic studies of Y- and FMR in the Z-type hexaferrite (Bax Sr1−x )3 Co2 Fe24 O41 at frequen-
Z-type hexaferrites revealed the electromagnons in both kinds of cies below 2 GHz (measured without the magnetic field).202 Its fre-
the spectra.193,194,197,198 It is plausible because all polar excitations quency showed a linear decrease on heating toward the magnetic PT
should be both IR and Raman active in the non-centrosymmetric temperature of 500 K, where the magnetic structure changes from
FE phase. Analytical calculations confirm that the electromagnons conical to collinear. FMR studies performed near 9 GHz with a mag-
are activated by the exchange striction.193,196 These calculations also netic field up to 10 kOe also confirmed the PT near 500 K. Moreover,
allowed us to determine selection rules for electromagnons in var- FMR spectra were tunable by an external dc and ac electric field
ious magnetic structures of hexaferrites. Figure 18 shows that the that allowed us to determine the value of the linear magnetoelec-
electromagnon in Y-type hexaferrites softens on heating toward the tric coefficient αE = 390 ps/m at 170 K.202 The value of αE was one
magnetic (and simultaneously FE) PT near room temperature so order of magnitude lower than that reported at 10 K201 but still
that it behaves similarly as the FE soft phonon in displacive ferro- one or even four orders of magnitude higher than in other “high-
electrics. Since the electromagnon is activated in the THz and Raman temperature” multiferroics. The sensitivity of the FMR on electric
spectra only in an alternating longitudinal conical magnetic struc- field gives evidence that the FMR has an electromagnon character.
ture and this structure disappears in the magnetic fields above 4 T, In 2012, an unusually strong spin-order induced FE polariza-
where a collinear magnetic structure develops, the intensity of the tion was revealed in CaMn7 O12 below 90 K. Subsequently, it was
electromagnon weakens and its frequency softens with the magnetic shown that the quadruple perovskites (AMn3 )Mn4 O12 (A2+ = Cd, Sr,
field, and finally, this excitation disappears in the field above 4 T Pb) exhibit a similar sequence of structural and magnetic PTs.203–205
[Fig. 18(b)]. On the other hand, a new excitation appears in the THz Near 400 K, they undergo a charge-ordering metal–insulator PT
spectra below 0.4 THz in the field above 4 T. This is a ferromagnetic from a cubic to a rhombohedral R3 structure. The next PT gives

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-16


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm


FIG. 19. Temperature dependence of (a) the low-frequency excitations and (b) their plasma frequencies Ωj = Δεj ωj observed in the THz and IR spectra of SrMn7 O12 .
The open symbols correspond to spin excitations or phonons activated by breaking of the inversion center, whereas the remaining phonons are marked by solid symbols.
(c) Real and (d) imaginary parts of the complex refractive index of SrMn7 O12 ceramics obtained from the THz time-domain spectroscopy. Reproduced with permission from

30 October 2024 15:21:18


Kamba et al., Phys. Rev. B 99, 184108 (2019). Copyright 2019 American Physical Society.

rise to the Mn orbital ordering and an incommensurate structural with the MW dielectric relaxation so that this system belongs to the
modulation along the c axis below ∼250 K. On further cooling, the crossover between order–disorder and displacive type of the PT.28
samples show two magnetic PTs near 90 K and 40 K with PT temper- On the other hand, another related tartrate—lithium thallium tar-
atures depending on the chemical composition. The antiferromag- trate monohydrate—shows a typical behavior of the displacive fer-
netic phases are helical and incommensurately modulated. IR, THz, roelectrics.50 The crossover between displacive and order–disorder
and Raman response of AMn7 O12 (A = Ca, Sr) ceramics revealed PT exhibits also BaTiO3 .122,127 Lead-based relaxor ferroelectrics with
qualitatively similar dramatic changes in phonon spectra with tem- perovskite structures show a strong and soft dielectric relaxation
perature reflecting changes in phonon selection rules in various whose distribution of relaxation frequencies anomalously broadens
crystal structures (Fig. 17) including new phonons appearing in the on cooling toward the freezing temperatures and covers the frequen-
spin-order-induced FE phases.191,192 The strongest variations occur cies from the sub-Hz up to THz region even below the freezing
in THz spectra near the two magnetic PTs [see Figs. 19(c) and 19(d)]. temperatures.96,100 Although the relaxors are cubic in the whole tem-
They activate new modes in the spectra, with resonance frequen- perature interval, the phonons are split at all temperatures due to
cies and intensities changing with temperature and magnetic field. the locally broken center of symmetry in polar nanodomains and
Below the lowest-temperature magnetic PT, a transfer of plasma fre- the lowest frequency phonon undergoes a critical softening toward
quencies Ω1 and Ω2 (defined as square roots of oscillator strength) T ∗ = 350 K–400 K, giving evidence about a local PT in polar nan-
from the low-frequency phonons to these excitations was observed odomains.103,110 In incipient FE SrTiO3 , Srn+1 Tin O3n+1 , and KTaO3 ,
[see Fig. 19(b)]; therefore, they were assigned to electromagnons frequencies of the soft phonons strongly reduce on cooling and sat-
because they must be polar. urate near liquid helium temperatures due to quantum fluctuations,
but it was shown that the epitaxial strain in thin films can induce
displacive FE PTs, which can be revealed in the phonon behav-
IX. CONCLUSIONS
ior.129,130,137,138,141 In antiferromagnetic and incipient ferroelectric
In this article, the basic concept of the FE soft modes driving EuTiO3 , 1% tensile strain induces not only a displacive FE PT at 250
the FE PTs is explained and demonstrated in materials investigated K, but the antiferromagnetic structure changes also to a ferromag-
mainly in Prague within the last 25 years. In Rochelle salt—the first netic one due to a strong spin–phonon coupling.159 In ferromagnetic
known FE crystal—it has been shown that although this material EuO, a huge tensile strain (more than 6%) was needed to induce a
was for a long time considered a typical example of the system with displacive FE PT,168 even if the phonon is highly temperature sta-
an order–disorder PT, two coupled polar phonons show anomalous ble in the bulk EuO. Incomplete phonon softening was observed on
softening with rising temperature and these phonons are coupled heating in the multiferroic BiFeO3 connected with the first-order

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-17


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

and improper FE PT at 1100 K.145 The displacive character of the 8


G. S. Landsberg and L. I. Mandelstam, “Lichtzerstreuung in krystallen bei hoher
FE PT was revealed in the multiferroic Sr0.55 Ba0.45 MnO3 .173 The soft temperatur,” Z. Phys. 60, 364–375 (1930).
9
mode exhibits also a strong anomaly at the antiferromagnetic PT A. S. Barker and M. Tinkham, “Far-infrared ferroelectric vibration mode in
due to a large spin–phonon coupling in this material.173 In multifer- SrTiO3 ,” Phys. Rev. 125, 1527–1530 (1962).
10
R. A. Cowley, “Temperature dependence of a transverse optic mode in stron-
roic hexagonal BaMnO3 , the FE soft mode was observed in the THz
tium titanate,” Phys. Rev. Lett. 9, 159–161 (1962).
and Raman spectra but only below T c = 130 K because the improper 11
J. F. Scott, “Soft-mode spectroscopy: Experimental studies of structural phase
FE PT related with tripling of the unit cell occurs below T c so that transitions,” Rev. Mod. Phys. 46, 83–128 (1974).
the soft mode has a wave vector from the Brillouin zone boundary 12
G. A. Samara and P. S. Peercy, “The study of soft-mode transitions at high
in the paraelectric phase and therefore cannot be IR/Raman active pressure,” Solid State Phys. 36, 1–118 (1982).
above T c .176 Electromagnons were detected in the THz and Raman 13
J. Petzelt, G. V. Kozlov, and A. A. Volkov, “Dielectric-spectroscopy of paraelec-
spectra of the multiferroic BiFeO3 150,151,152 but also in spin-induced tric soft modes,” Ferroelectrics 73, 101–123 (1987).
14
multiferroics with the Y- and Z-hexaferrite crystal structures,193,197 M. E. Lines and A. M. Glass, Principles and Applications of Ferroelectrics and
as well as in quadruple perovskite SrMn7 O12 .192 Their frequencies Related Materials (Clarendon Press, Oxford, 1977).
15
soften on heating toward the temperatures of the magnetic PTs sim- J. Petzelt, S. Kamba, and J. Hlinka, “Ferroelectric soft modes in ceramics and
films, review,” in New Development in Advanced Functional Ceramics, edited by
ilar to the soft phonons in ferroelectrics. In hexaferrites, the mag- L. Mitoseriu (Transworld Research Network, Trivandrum, 2007), pp. 387–421.
netic structures are highly sensitive to the magnetic field; therefore, 16
S.-E. Park and T. R. Shrout, “Ultrahigh strain and piezoelectric behavior in
the electromagnons disappear from the spectra at higher magnetic relaxor based ferroelectric single crystals,” J. Appl. Phys. 82, 1804–1811 (1997).
fields.193,197 17
S. Kamba, E. Buixaderas, T. Ostapchuk, and J. Petzelt, “Ferroelectric soft modes
All these results demonstrate the high ability and sensitivity of and dynamic central modes near some phase transitions,” Ferroelectrics 268,
IR and THz spectroscopy for determining the type of FE PTs in var- 163–168 (2002).
18
ious materials, not only in the bulk form but also in the ultrathin R. H. Lyddane, R. G. Sachs, and E. Teller, “On the polar vibrations of alkali
films, which can be highly strained and exhibit a behavior completely halides,” Phys. Rev. 59, 673–676 (1941).
19
J. Hlinka, B. Hehlen, A. Kania, and I. Gregora, “Soft mode in cubic PbTiO3 by
different from the bulk samples.
hyper-Raman scattering,” Phys. Rev. B 87, 064101 (2013).
20
H. Vogt, “Refined treatment of the model of linearly coupled anharmonic
DEDICATION oscillators and its application to the temperature dependence of the zone-center
soft-mode frequencies of KTaO3 and SrTiO3 ,” Phys. Rev. B 51, 8046–8059 (1995).
This article is dedicated to my teacher and great friend Jan 21
J. Petzelt, T. Ostapchuk, I. Gregora, I. Rychetsky, S. Hoffmann-Eifert, A. Pronin,

30 October 2024 15:21:18


Petzelt on the occasion of his 80th birthday. Y. Yuzyuk, B. Gorshunov, S. Kamba, V. Bovtun, J. Pokorny, M. Savinov, V.
Porokhonskyy, D. Rafaja, P. Vanek, A. Almeida, M. Chaves, A. Volkov, M.
ACKNOWLEDGMENTS Dressel, and R. Waser, “Dielectric, infrared, and Raman response of undoped
SrTiO3 ceramics: Evidence of polar grain boundaries,” Phys. Rev. B 64, 184111
I acknowledge the support of the Czech Science Founda- (2001).
tion (Project No. 18-09265S) and the Operational Programme 22
Y. Ichikawa, M. Nagai, and K. Tanaka, “Direct observation of the soft-mode
Research, Development and Education financed by the Euro- dispersion in the incipient ferroelectric KTaO3 ,” Phys. Rev. B 71, 092106 (2005).
23
pean Structural and Investment Funds and the Czech Min- E. Buixaderas, S. Kamba, I. Gregora, P. Vanek, J. Petzelt, T. Yamaguchi,
istry of Education, Youth and Sports (Project No. SOLID 21- and M. Wada, “Far-infrared and Raman studies of the ferroelectric phase
CZ.02.1.01/0.0/0.0/16_019/0000760). transition in LiNaGe4 O9 ,” Phys. Status Solidi B 214, 441–452 (1999); avail-
able at https://onlinelibrary.wiley.com/doi/10.1002/%28SICI%291521-3951%
28199908%29214%3A2%3C441%3A%3AAID-PSSB441%3E3.0.CO%3B2-X
24
A. Piecha, G. Bator, and R. Jakubas, “Critical slowing down of low-frequency
DATA AVAILABILITY dielectric relaxation in ferroelectric (C3 N2 H5 )5 Bi2 Cl11 ,” J. Phys.: Condens. Matter
The data that support the findings of this study are available 17, L411–L417 (2005).
25
from the corresponding author upon reasonable request. G. V. Kozlov, A. A. Volkov, J. F. Scott, G. E. Feldkamp, and J. Petzelt,
“Millimeter-wavelength spectroscopy of the ferroelectric phase transition in tris-
sarcosine calcium chloride,” Phys. Rev. B 28, 255–261 (1983).
26
REFERENCES J. Petzelt, A. A. Volkov, Y. G. Goncharov, J. Albers, and A. Klöpperpieper,
“Dynamic critical behaviour in TSCC,” Solid State Commun. 73, 5–9 (1990).
1 27
W. Cochran, “Crystal stability and the theory of ferroelectricity,” Phys. Rev. Lett. K. A. Müller, Y. Luspin, J. L. Servoin, and F. Gervais, “Displacive-order-
3, 412–414 (1959). disorder crossover at the ferroelectric-paraelectric phase transitions of BaTiO3 and
2
W. Cochran, “Crystal stability and the theory of ferroelectricity,” Adv. Phys. 9, LiTaO3 ,” J. Phys. Lett. 43, 537–542 (1982).
387–423 (1960). 28
S. Kamba, G. Schaack, and J. Petzelt, “Vibrational spectroscopy and soft-mode
3
W. Cochran, “Crystal stability and the theory of ferroelectricity part II. Piezoelec- behavior in Rochelle salt,” Phys. Rev. B 51, 14998–15007 (1995).
tric crystals,” Adv. Phys. 10, 401–420 (1961). 29
E. Buixaderas, S. Kamba, and J. Petzelt, “Lattice dynamics and central-mode
4
P. W. Anderson, “Qualitative description of the phase transition in BaTiO3 -type phenomena in the dielectric response of ferroelectrics and related materials,”
ferroelectrics,” in Fizika Dielectrikov, edited by G. I. Skanavi (Akademii Nauk Ferroelectrics 308, 131–192 (2004).
USSR, Moscow, 1960). 30
J. Valasek, “Piezoelectric and allied phenomena in Rochelle salt,” Phys. Rev. 15,
5
V. L. Ginzburg, “O polyarizatsii i piezoeffekte titanata bariya vblizi tochki 537–538 (1920).
segnetoelectricheskogo perekhoda,” JETP 19, 36–41 (1949). 31
J. Valasek, “Piezo-electric and allied phenomena in Rochelle salt,” Phys. Rev. 17,
6
V. L. Ginzburg, “Teoria segnetoelektriceskich javlenij,” Usp. Fiz. Nauk 38, 475–481 (1921).
32
490–525 (1949). Y. Iwata, S. Mitani, and I. Shibuya, “Neutron diffraction analysis on a possi-
7
V. L. Ginzburg, “Phase transitions in ferroelectrics (some historical remarks),” ble disordered structure of paraelectric Rochelle salt,” Ferroelectrics 107, 287–292
Ferroelectrics 267, 23–32 (2010). (1990).

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-18


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

33 56
F. Mo, R. H. Mathiesen, J. A. Beukes, and K. M. Vu, “Rochelle salt—A structural T. Kawamura and A. Mitsuishi, “Far-infrared and Raman spectra of KD2 PO4 ,”
reinvestigation with improved tools. I. The high-temperature para-electric phase Opt. Commun. 10, 337–340 (1974).
57
at 308 K,” IUCrJ 2, 19–28 (2015). D. A. Ledsham, W. G. Chambers, and T. J. Parker, “Far infrared measurements
34
H. Akao and T. Sasaki, “Dielectric dispersion of Rochelle salt in the microwave on KH2 PO4 using dispersive reflection spectroscopy,” Infrared Phys. 17, 165–172
region,” J. Chem. Phys. 23, 2210–2214 (1955). (1976).
35 58
H. E. Müser and J. Potthaest, “Zum dielektrischen Verhalten von Seignette- M. El Sherif, F. Brehat, B. Wyncke, and A. Hadni, “Far infrared reflectivity
salz im Bereich der Dezimeter- und Zentimeterwellen,” Phys. Status Solidi B 24, spectra of a KH2 PO4 single crystal,” Int. J. Infrared Millimeter Waves 5, 815–827
109–113 (1967). (1984).
59
36
F. Sandy and R. V. Jones, “Dielectric relaxation of Rochelle salt,” Phys. Rev. 168, B. Wyncke and F. Brehat, “Temperature dependence of the lattice modes
481–493 (1968). polarised in a direction perpendicular to the c axis of potassium dihydrogen
37
A. A. Volkov, G. V. Kozlov, E. B. Kryukova, and J. Petzelt, “Low-temperature phosphate,” J. Phys. C: Solid State Phys. 19, 2649–2661 (1986).
60
transformations of the relaxational soft modes in crystals of the Rochelle salt F. Brehat and B. Wyncke, “Low-frequency modes in KH2 PO4 -type crystals,” Int.
family,” Sov. Phys. JETP 63, 110–114 (1986); available at http://jetp.ac.ru/cgi- J. Infrared Millimeter Waves 8, 155 (1987).
61
bin/dn/e_063_01_0110.pdf F. Gervais and A. Simon, “Infrared spectroscopy of KH2 PO4 -type
38
A. A. Volkov, G. V. Kozlov, E. B. Kryukova, and A. A. Sobyanin, ferroelectrics,” Ferroelectrics 72, 77–93 (1987).
62
“New results on the dynamics of Rochelle salt crystals (a system with a P. Simon, F. Gervais, and E. Courtens, “Paraelectric-ferroelectric phase transi-
“double” critical point),” Sov. Phys. Usp. 29, 574–575 (1986); available at tions of KH2 PO4 , RbH2 PO4 , and KH2 AsO4 studied by infrared reflectivity,” Phys.
https://iopscience.iop.org/article/10.1070/PU1986v029n06ABEH003423/pdf Rev. B 37, 1969–1979 (1988).
63
39
J. Hlinka, J. Kulda, S. Kamba, and J. Petzelt, “Resonant soft mode in Rochelle F. Brehat and B. Wyncke, “Far-infrared reflectivity spectrum of potassium dihy-
salt by inelastic neutron scattering,” Phys. Rev. B 63, 052102 (2001). drogen phosphate, polarized along the c axis, reinvestigated in the ferroelectric
40 phase,” Infrared Phys. 33, 307–312 (1992).
S. Kamba, G. Schaack, J. Petzelt, and B. Brezina, “Soft-mode behaviour in 64
Rochelle salt and lithium thallium tartrate monohydrate at high pressure,” Fer- A. A. Volkov, G. V. Kozlov, S. P. Lebedev, and A. M. Prokhorov, “Proton modes
roelectrics 186, 181 (1996). in the crystals of KH2 PO4 family,” Ferroelectrics 25, 531–534 (1980).
65
41
K. Gesi and K. Ozawa, “Effect of hydrostatic pressure on the phase transitions in M. Horioka and R. Abe, “Dielectric dispersion in microwave region of
(Rochelle salt)1-x (ammonium Rochelle salt)x mixed crystal system,” J. Phys. Soc. KH2 PO4 ,” Ferroelectrics 108, 267–270 (1990).
66
Jpn. 48, 2003–2006 (1980). K. K. Kobayashi, “Dynamical theory of the phase transition in KH2 PO4 -type
42 ferroelectric crystals,” J. Phys. Soc. Jpn. 24, 497–508 (1968).
B. T. Matthias and J. K. Hulm, “New ferroelectric tartrates,” Phys. Rev. 82, 67
108–109 (1951). R. Blinc and B. Žekš, “Proton-lattice interactions and the soft mode in
43 KH2 PO4 ,” J. Phys. C: Solid State Phys. 15, 4661–4670 (1982).
J. Fousek, L. E. Cross, and K. Seely, “Some properties of the ferroelectric lithium 68
I. P. Kaminow and T. C. Damen, “Temperature dependence of the ferroelectric

30 October 2024 15:21:18


thallium tartrate,” Ferroelectrics 1, 63–70 (1970).
44 mode in KH2 PO4 ,” Phys. Rev. Lett. 20, 1105–1108 (1968).
E. Sawaguchi and L. E. Cross, “Electromechanical coupling effects on the dielec- 69
tric properties and ferroelectric phase transition in lithium thallium tartrate,” P. S. Peercy, “Observation of an underdamped ‘soft’ mode in potassium dihy-
Ferroelectrics 2, 37–46 (1971). drogen phosphate,” Phys. Rev. Lett. 31, 379–382 (1973).
70
45 P. S. Peercy, “Evaluation of the coupled proton-optic-mode model for KH2 PO4
X. Gerbaux, A. Hadni, J. Pierron, and S. Messaadi, “Soft mode and far infrared
and RbH2 PO4 ,” Phys. Rev. B 9, 4868–4871 (1974).
spectra of ferroelectric lithium thallium tartrate at low temperatures,” Int. J. 71
Infrared Millimeter Waves 6, 131–140 (1985). P. S. Peercy, “Soft mode and coupled modes in the ferroelectric phase of KDP,”
46 Solid State Commun. 16, 439–442 (1975).
A. A. Volkov, Y. G. Goncharov, G. V. Kozlov, J. Petzelt, J. Fousek, and 72
B. Brezina, “Temperature unstable mode in the submillimeter wavelength spec- R. S. Katiyar, M. A. F. Scarparo, R. Srivastava, and S. P. S. Porto, “Search of soft
trum of lithium-thallium tartrate,” Sov. Phys. Solid State 28, 1794 (1986); available modes in crystals having KDP structure,” Ferroelectrics 16, 213–218 (1977).
73
at http://www.mathnet.ru/links/376550a4a9947bd06e14211fc405fbbc/ftt689.pdf Y. Tominaga, “Study on ferroelectric phase transition of KH2 PO4 by Raman
47 scattering—Objection to the proton tunneling mode,” Ferroelectrics 52, 91–100
K. Hayashi, K. Deguchi, and E. Nakamura, “Elastic and dielectric studies on
(1983).
the phase transition of ferroelectric LiTlC4 H4 O6 ⋅H2 O,” J. Phys. Soc. Jpn. 61, 74
1357–1361 (1992). T. Shigenari, “Comment on the interpretation of Raman spectrum of KDP,”
48 Ferroelectrics 532, 6–12 (2019).
G. J. Mccarthy, L. H. Schlegel, and E. Sawaguchi, “Crystal data for lithium 75
U. T. Höchli, K. Knorr, and A. Loidl, “Orientational glasses,” Adv. Phys. 39,
thallium tartrate monohydrate,” J. Appl. Crystallogr. 4, 180–181 (1971).
49 405–615 (1990).
M. I. Kay, “The structure of the paraelectric phase of lithium thallium tar- 76
J. Petzelt, S. Kamba, A. V. Sinitksi, A. G. Pimenov, A. A. Volkov, G. V. Kozlov,
trate monohydrate by neutron diffraction data,” Ferroelectrics 19, 159–164
and R. Kind, “Far-infrared and near-millimetre dielectric response of DRADP-
(1978).
50
50 dipolar glass compared with that of RADP,” J. Phys.: Condens. Matter 5,
S. Kamba, G. Schaack, J. Petzelt, and B. Brezina, “High-pressure dielectric 3573–3586 (1993).
and lattice vibration studies of the phase transition in lithium thallium tartrate 77
G. A. Smolenskii and A. I. Agranovskaya, Sov. Phys.–Tech. Phys. 3, 1380 (1958).
monohydrate (LTT),” J. Phys.: Condens. Matter 8, 4631–4642 (1996). 78
51
V. A. Isupov, “Causes of phase-transition broadening and the nature of dielec-
A. K. Tagantsev, “Susceptibility anomaly in films with bilinear coupling between tric polarization in some ferroelectrics,” Sov. Phys.-Sol. State 5, 136–140 (1963).
order parameter and strain,” Phys. Rev. Lett. 94, 247603 (2005). 79
N. De Mathan, E. Husson, G. Calvarn, J. R. Gavarri, A. W. Hewat, and A. Morell,
52
S. Kamba, J. Kulda, V. Petříček, G. McIntyre, and J. P. Kiat, “High-pressure “A structural model for the relaxor PbMg1/3 Nb2/3 O3 at 5 K,” J. Phys.: Condens.
structural and dielectric studies of the phase transitions in lithium thallium Matter 3, 8159–8171 (1991).
tartrate monohydrate,” J. Phys.: Condens. Matter 14, 4045–4055 (2002). 80
G. Calvarin, E. Husson, and Z. G. Ye, “X-ray study of the electric field-induced
53
S. Kamba, B. Brezina, J. Petzelt, and G. Schaack, “Study of the phase transition in phase transition in single crystal Pb(Mg1/3 Nb2/3 )O3 ,” Ferroelectrics 165, 349–358
lithium ammonium tartrate monohydrate (LAT) by means of infrared and Raman (1995).
spectroscopy,” J. Phys.: Condens. Matter 8, 8669–8679 (1996). 81
G. Burns and F. H. Dacol, “Glassy polarization behavior in ferroelectric com-
54
R. J. Nelmes, G. M. Meyer, and J. E. Tibballs, “The crystal structure of tetragonal pounds Pb(Mg1/3 Nb2/3 )O3 and Pb(Zn1/3 Nb2/3 )O3 ,” Solid State Commun. 48,
KH2 PO4 and KD2 PO4 as a function of temperature,” J. Phys. C: Solid State Phys. 853–856 (1983).
15, 59–75 (1982). 82
V. Westphal, W. Kleemann, and M. D. Glinchuk, “Diffuse phase-
55
M. C. Lawrence and G. N. Robertson, “Estimating the proton potential in KDP transitions and random-field-induced domain states of the relaxor ferroelectric
from infrared and crystallographic data,” Ferroelectrics 34, 179–186 (1981). PbMg1/3 Nb2/3 O3 ,” Phys. Rev. Lett. 68, 847–850 (1992).

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-19


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

83 106
L. E. Cross, “Relaxor ferroelectrics,” Ferroelectrics 76, 241–267 (1987). P. M. Gehring, S. Wakimoto, Z. G. Ye, and G. Shirane, “Soft mode dynamics
84 above and below the Burns temperature in the relaxor Pb(Mg1/3 Nb2/3 )O3 ,” Phys.
L. E. Cross, “Relaxor ferroelectrics: An overview,” Ferroelectrics 151, 305–320
(1994). Rev. Lett. 87, 277601 (2001).
85 107
Z.-G. Ye, Key Engineering Materials, edited by C. Boulesteix (Trans Tech J. Hlinka, S. Kamba, J. Petzelt, J. Kulda, C. Randall, and S. Zhang, “Origin of
Publications, Switzerland, 1998), Vol. 155-156, pp. 81–122. the ‘Waterfall’ effect in phonon dispersion of relaxor perovskites,” Phys. Rev. Lett.
86
G. A. Samara, “Ferroelectricity revisited—Advances in materials and physics,” 91, 107602 (2003).
108
in Solid State Physics, Advances in Research and Applications, edited by H. Ehren- S. Kamba, M. Kempa, V. Bovtun, J. Petzelt, K. Brinkman, and N. Setter, “Soft
reich and F. Spaepen (Academic Press, San Diego, 2001), Vol. 56, pp. 240–458. and central mode behaviour in PbMg1/3 Nb2/3 O3 relaxor ferroelectric,” J. Phys.:
87 Condens. Matter 17, 3965–3974 (2005).
G. A. Samara, “The relaxational properties of compositionally disordered ABO3
109
perovskites,” J. Phys.: Condens. Matter 15, R367–R411 (2003). S. Kamba, M. Berta, M. Kempa, J. Hlinka, J. Petzelt, K. Brinkman, and N. Setter,
88 “Far-infrared soft-mode behavior in PbSc1/2 Ta1/2 O3 thin films,” J. Appl. Phys. 98,
A. A. Bokov and Z.-G. Ye, “Recent progress in relaxor ferroelectrics with
perovskite structure,” J. Mater. Sci. 41, 31–52 (2006). 074103 (2005).
89 110
A. A. Bokov and Z.-G. Ye, “Dielectric relaxation in relaxor ferroelectrics,” D. Nuzhnyy, J. Petzelt, V. Bovtun, S. Kamba, and J. Hlinka, “Soft mode driven
J. Adv. Dielectr. 02, 1241010 (2012). local ferroelectric transition in lead-based relaxors,” Appl. Phys. Lett. 114, 182901
90
I. G. Siny, R. S. Katiyar, and A. S. Bhalla, Ferroelectr. Rev. 2, 51 (2000). (2019).
111
91
G. Shirane and P. M. Gehring, Morphotropic Phase Boundary Perovskites, High D. Viehland, S. J. Jang, L. E. Cross, and M. Wuttig, “Deviation from Curie-
Strain Piezoelectrics, and Dielectric Ceramics, Ceramic Transactions, edited by R. Weiss behavior in relaxor ferroelectrics,” Phys. Rev. B 46, 8003–8006 (1992).
112
Guo, K. M. Nair, W. Wong-Ng, A. Bhalla, D. Viehland, D. Suvorov, C. Wu, and M. E. Lines and A. M. Glass, Principles and Applications of Ferroelectrics and
S.-I. Hirano (The American Ceramic Society, 2003), Vol. 136, pp. 17–35. Related Materials (Oxford University Press, Oxford, 2001).
92 113
R. A. Cowley, S. N. Gvasaliya, S. G. Lushnikov, B. Roessli, and G. M. Rotaru, R. Comes, M. Lambert, and A. Guinier, “Chain structure of BaTiO3 and
“Relaxing with relaxors: A review of relaxor ferroelectrics,” Adv. Phys. 60, 229–327 KNbO3 ,” Solid State Commun. 6, 715–719 (1968).
114
(2011). A. Zi˛ebińska, D. Rytz, K. Szot, M. Gorny, and K. Roleder, “Birefringence above
93 T c in single crystals of barium titanate,” J. Phys.: Condens. Matter 20, 142202
S. Kamba and J. Petzelt, Piezoelectric Single Crystals and Their Application,
edited by S. Trolier-McKinstry, L. E. Cross, and Y. Yamashita (Penn State (2008).
115
University, 2004), pp. 257–275. J. Hlinka, T. Ostapchuk, D. Nuzhnyy, J. Petzelt, P. Kuzel, C. Kadlec, P. Vanek,
94 I. Ponomareva, and L. Bellaiche, “Coexistence of the phonon and relaxation soft
J. Hlinka, J. Petzelt, S. Kamba, D. Noujni, and T. Ostapchuk, “Infrared dielectric
response of relaxor ferroelectrics,” Phase Transitions 79, 41–78 (2006). modes in the terahertz dielectric response of tetragonal BaTiO3 ,” Phys. Rev. Lett.
95
J. Petzelt, D. Nuzhnyy, V. Bovtun, M. Kempa, M. Savinov, S. Kamba, and J. 101, 167402 (2008).
116
Hlinka, “Lattice dynamics and dielectric spectroscopy of BZT and NBT lead-free G. Burns and F. H. Dacol, “Polarization in the cubic phase of BaTiO3 ,” Solid
perovskite relaxors - comparison with lead-based relaxors,” Phase Transitions 88, State Commun. 42, 9–12 (1982).

30 October 2024 15:21:18


117
320–332 (2015). E. Dul’kin, J. Petzelt, S. Kamba, E. Mojaev, and M. Roth, “Relaxor-like behavior
96 of BaTiO3 crystals from acoustic emission study,” Appl. Phys. Lett. 97, 032903
S. Kamba, V. Bovtun, J. Petzelt, I. Rychetsky, R. Mizaras, A. Brilingas, J. Banys,
J. Grigas, and M. Kosec, “Dielectric dispersion of the relaxor PLZT ceramics in (2010).
118
the frequency range 20 Hz-100 THz,” J. Phys.: Condens. Matter 12, 497–519 F. Jona and G. Shirane, Ferroelectric Crystals (Pergamon Press, Oxford, 1962).
119
(2000). J. M. Ballantyne, “Frequency and temperature response of the polarization of
97
I. Rychetský, S. Kamba, V. Porokhonskyy, A. Pashkin, M. Savinov, V. Bovtun, barium titanate,” Phys. Rev. 136, A429–A436 (1964).
120
J. Petzelt, M. Kosec, and M. Dressel, “Frequency-independent dielectric losses A. S. Barker, “Temperature dependence of the transverse and longitudinal
(1/f noise) in PLZT relaxors at low temperatures,” J. Phys.: Condens. Matter 15, optic mode frequencies and charges in SrTiO3 and BaTiO3 ,” Phys. Rev. 145,
6017–6030 (2003). 391–399 (1966).
98 121
S. Kamba, D. Nuzhnyy, S. Veljko, V. Bovtun, J. Petzelt, Y. L. Wang, N. Setter, Y. Luspin, J. L. Servoin, and F. Gervais, “Soft mode spectroscopy in barium
J. Levoska, M. Tyunina, J. Macutkevic, and J. Banys, “Dielectric relaxation and titanate,” J. Phys. C: Solid State Phys. 13, 3761–3773 (1980).
polar phonon softening in relaxor ferroelectric PbMg1/3 Ta2/3 O3 ,” J. Appl. Phys. 122
V. Bovtun, D. Nuzhnyy, M. Kempa, T. Ostapchuk, V. Skoromets, J. Suchan-
102, 074106 (2007). icz, P. Czaja, J. Petzelt, and S. Kamba, “Ferroelectric soft mode and microwave
99
S. Kamba, V. Porokhonskyy, A. Pashkin, V. Bovtun, J. Petzelt, J. Nino, S. Trolier- dielectric relaxation in in BaTiO3 -PbMg1/3 Nb2/3 O3 ceramics,” Phys. Rev. Mater.
Mckinstry, M. Lanagan, and C. Randall, “Anomalous broad dielectric relaxation 5, 014404 (2021).
in Bi1.5 Zn1.0 Nb1.5 O7 pyrochlore,” Phys. Rev. B 66, 054106 (2002). 123
H. Vogt, J. A. Sanjurjo, and G. Rossbroich, “Soft-mode spectroscopy in cubic
100
V. Bovtun, S. Veljko, S. Kamba, J. Petzelt, S. Vakhrushev, Y. Yakymenko, BaTiO3 by hyper-Raman scattering,” Phys. Rev. B 26, 5904–5910 (1982).
K. Brinkman, and N. Setter, “Broad-band dielectric response of PbMg1/3 Nb2/3 O3 124
I. Ponomareva, L. Bellaiche, T. Ostapchuk, J. Hlinka, and J. Petzelt, “Terahertz
relaxor ferroelectrics: Single crystals, ceramics and thin films,” J. Eur. Ceram. Soc. dielectric response of cubic BaTiO3 ,” Phys. Rev. B 77, 012102 (2008).
26, 2867–2875 (2006). 125
J. Petzelt, D. Nuzhnyy, M. Savinov, V. Bovtun, M. Kempa, T. Ostapchuk,
101
V. Bovtun, S. Kamba, A. Pashkin, M. Savinov, P. Samoukhina, J. Petzelt, I. P. J. Hlinka, G. Canu, and V. Buscaglia, “Broadband dielectric spectroscopy of
Bykov, and M. D. Glinchuk, “Central-peak components and polar soft mode in Ba(Zr,Ti)O3 : Dynamics of relaxors and diffuse ferroelectrics,” Ferroelectrics 469,
relaxor PbMg1/3 Nb2/3 O3 crystals,” Ferroelectrics 298, 23–30 (2004). 14–25 (2014).
102 126
V. Bovtun, J. Petzelt, V. Porokhonskyy, S. Kamba, and Y. Yakimenko, K. Tsuda, R. Sano, and M. Tanaka, “Nanoscale local structures of rhombohe-
“Structure of the dielectric spectrum of relaxor ferroelectrics,” J. Eur. Ceram. Soc. dral symmetry in the orthorhombic and tetragonal phases of BaTiO3 studied by
21, 1307–1311 (2001). convergent-beam electron diffraction,” Phys. Rev. B 86, 214106 (2012).
103 127
D. Nuzhnyy, J. Petzelt, V. Bovtun, M. Kempa, S. Kamba, and J. Hlinka, M. Pasciak, T. R. Welberry, J. Kulda, S. Leoni, and J. Hlinka, “Dynamic
“Infrared, terahertz, and microwave spectroscopy of the soft and central modes displacement disorder of cubic BaTiO3 ,” Phys. Rev. Lett. 120, 167601 (2018).
in Pb(Mg1/3 Nb2/3 )O3 ,” Phys. Rev. B 96, 174113 (2017). 128
J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, and A. K. Tagant-
104
S. Wakimoto, C. Stock, R. J. Birgeneau, Z. G. Ye, W. Chen, W. J. L. Buyers, P. M. sev, “Room-temperature ferroelectricity in strained SrTiO3 ,” Nature 430, 758–761
Gehring, and G. Shirane, “Ferroelectric ordering in the relaxor Pb(Mg1/3 Nb2/3 )O3 (2004).
as evidenced by low-temperature phonon anomalies,” Phys. Rev. B 65, 172105 129
D. Nuzhnyy, J. Petzelt, S. Kamba, P. Kužel, C. Kadlec, V. Bovtun,
(2002). M. Kempa, J. Schubert, C. M. Brooks, and D. G. Schlom, “Soft mode behavior in
105
P. M. Gehring, S.-E. Park, and G. Shirane, “Soft phonon anomalies in the SrTiO3 /DyScO3 thin films: Evidence of ferroelectric and antiferrodistortive phase
relaxor ferroelectric Pb(Zn1/3 Nb2/3 )0.92 Ti0.08 O3 ,” Phys. Rev. Lett. 84, 5216 (2000). transitions,” Appl. Phys. Lett. 95, 232902 (2009).

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-20


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

130
C.-H. Lee, V. Skoromets, M. D. Biegalski, S. Lei, R. Haislmaier, M. Bernhagen, Laulainen, and D. C. Lupascu, “Multiferroic bismuth ferrite: Perturbed angular
R. Uecker, X. Xi, V. Gopalan, X. Martí, S. Kamba, P. Kužel, and D. G. Schlom, correlation studies on its ferroic α−β phase transition,” Phys. Rev. B 102, 224110
“Effect of stoichiometry on the dielectric properties and soft mode behavior of (2020).
149
strained epitaxial SrTiO3 thin films on DyScO3 substrates,” Appl. Phys. Lett. 102, R. Palai, R. S. Katiyar, H. Schmid, P. Tissot, S. J. Clark, J. Robertson, S. A. T.
082905 (2013). Redfern, G. Catalan, and J. F. Scott, “β phase and γ−β metal-insulator transition
131
C. Kadlec, V. Skoromets, F. Kadlec, H. Němec, J. Hlinka, J. Schubert, in multiferroic BiFeO3 ,” Phys. Rev. B 77, 014110 (2008).
150
G. Panaitov, and P. Kužel, “Temperature and electric field tuning of the ferro- S. Skiadopoulou, V. Goian, C. Kadlec, F. Kadlec, X. F. Bai, I. C. Infante,
electric soft mode in a strained SrTiO3 /DyScO3 heterostructure,” Phys. Rev. B 80, B. Dkhil, C. Adamo, D. G. Schlom, and S. Kamba, “Spin and lattice excitations
174116 (2009). of a BiFeO3 thin film and ceramics,” Phys. Rev. B 91, 174108 (2015).
132 151
D. Nuzhnyy, J. Petzelt, S. Kamba, X. Martí, T. Čechal, C. M. Brooks, and D. G. I. Kezsmarki, U. Nagel, S. Bordacs, R. S. Fishman, J. H. Lee, H. T. Yi, S. W.
Schlom, “Infrared phonon spectroscopy of a compressively strained (001) SrTiO3 Cheong, and T. Room, “Optical diode effect at spin-wave excitations of the room-
film grown on a (110) NdGaO3 substrate,” J. Phys.: Condens. Matter 23, 045901 temperature multiferroic BiFeO3 ,” Phys. Rev. Lett. 115, 127203 (2015).
152
(2011). M. Cazayous, Y. Gallais, A. Sacuto, R. De Sousa, D. Lebeugle, and D. Colson,
133 “Possible observation of cycloidal electromagnons in BiFeO3 ,” Phys. Rev. Lett.
D. Nuzhnyy, J. Petzelt, S. Kamba, T. Yamada, M. Tyunina, A. K. Tagantsev, J.
Levoska, and N. Setter, “Polar phonons in some compressively stressed epitaxial 101, 037601–037604 (2008).
153
and polycrystalline SrTiO3 thin films,” J. Electroceram. 22, 297–301 (2009). C. J. Fennie and K. M. Rabe, “Magnetic and electric phase control in epitaxial
134 EuTiO3 from first principles,” Phys. Rev. Lett. 97, 267602 (2006).
R. C. Haislmaier, R. Engel-Herbert, and V. Gopalan, “Stoichiometry as key
154
to ferroelectricity in compressively strained SrTiO3 films,” Appl. Phys. Lett. 109, T. Katsufuji and H. Takagi, “Coupling between magnetism and dielectric
032901 (2016). properties in quantum paraelectric EuTiO3 ,” Phys. Rev. B 64, 054415 (2001).
135 155
J. Petzelt and S. Kamba, “Far infrared and terahertz spectroscopy of ferroelec- S. Kamba, D. Nuzhnyy, P. Vaněk, M. Savinov, K. Knížek, Z. Shen, E. Šantavá,
tric soft modes in thin films: A review,” Ferroelectrics 503, 19–44 (2016). K. Maca, M. Sadowski, and J. Petzelt, “Magnetodielectric effect and optic soft
136
D. Noujni, S. Kamba, A. Pashkin, V. Bovtun, J. Petzelt, A.-K. Axelsson, N. M. mode behaviour in quantum paraelectric EuTiO3 ceramics,” Europhys. Lett. 80,
Alford, P. L. Wise, and I. M. Reaney, “Temperature dependence of microwave 27002 (2007).
156
and THz dielectric response in Srn+1 Tin O3n+1 (n = 1–4),” Integr. Ferroelectr. 62, V. Goian, S. Kamba, J. Hlinka, P. Vaněk, A. A. Belik, T. Kolodiazhnyi, and
199–203 (2004). J. Petzelt, “Polar phonon mixing in magnetoelectric EuTiO3 ,” Eur. Phys. J. B 71,
137
C.-H. Lee, N. D. Orloff, T. Birol, Y. Zhu, V. Goian, E. Rocas, R. Haislmaier, 429–433 (2009).
157
E. Vlahos, J. A. Mundy, L. F. Kourkoutis, Y. Nie, M. D. Biegalski, J. Zhang, V. Goian, S. Kamba, O. Pacherová, J. Drahokoupil, L. Palatinus, M. Dušek,
M. Bernhagen, N. A. Benedek, Y. Kim, J. D. Brock, R. Uecker, X. X. Xi, V. Gopalan, J. Rohlíček, M. Savinov, F. Laufek, W. Schranz, A. Fuith, M. Kachlík, K. Maca,
D. Nuzhnyy, S. Kamba, D. A. Muller, I. Takeuchi, J. C. Booth, C. J. Fennie, and A. Shkabko, L. Sagarna, A. Weidenkaff, and A. A. Belik, “Antiferrodistortive phase
D. G. Schlom, “Exploiting dimensionality and defect mitigation to create tunable transition in EuTiO3 ,” Phys. Rev. B 86, 054112 (2012).
158
J. Köhler, R. Dinnebier, and A. Bussmann-Holder, “Structural instability of

30 October 2024 15:21:18


microwave dielectrics,” Nature 502, 532–536 (2013).
138
V. Goian, S. Kamba, N. Orloff, T. Birol, C. H. Lee, D. Nuzhnyy, J. C. Booth, EuTiO3 from X-ray powder diffraction,” Phase Transitions 85, 949–955 (2012).
159
M. Bernhagen, R. Uecker, and D. G. Schlom, “Influence of the central mode and J. H. Lee, L. Fang, E. Vlahos, X. Ke, Y. W. Jung, L. F. Kourkoutis, J.-W. Kim, P. J.
soft phonon on the microwave dielectric loss near the strain-induced ferroelectric Ryan, T. Heeg, M. Roeckerath, V. Goian, M. Bernhagen, R. Uecker, P. C. Hammel,
phase transitions in Srn+1 Tin O3n+1 ,” Phys. Rev. B 90, 174105 (2014). K. M. Rabe, S. Kamba, J. Schubert, J. W. Freeland, D. A. Muller, C. J. Fennie,
139 P. Schiffer, V. Gopalan, E. Johnston-Halperin, and D. G. Schlom, “A strong fer-
N. M. Dawley, E. J. Marksz, A. M. Hagerstrom, G. H. Olsen, M. E. Holtz,
roelectric ferromagnet created by means of spin-lattice coupling,” Nature 466,
V. Goian, C. Kadlec, J. Zhang, X. Lu, J. A. Drisko, R. Uecker, S. Ganschow, C. J.
954–959 (2010).
Long, J. C. Booth, S. Kamba, C. J. Fennie, D. A. Muller, N. D. Orloff, and D. G. 160
Schlom, “Targeted chemical pressure yields tuneable millimetre-wave dielectric,” J. H. Lee, L. Fang, E. Vlahos, X. Ke, Y. W. Jung, L. F. Kourkoutis, J.-W. Kim, P. J.
Nat. Mater. 19, 176–181 (2020). Ryan, T. Heeg, M. Roeckerath, V. Goian, M. Bernhagen, R. Uecker, P. C. Hammel,
140 K. M. Rabe, S. Kamba, J. Schubert, J. W. Freeland, D. A. Muller, C. J. Fennie, P.
V. Železný, J. Buršík, and P. Vaněk, “Preparation and infrared characterization
Schiffer, V. Gopalan, E. Johnston-Halperin, and D. G. Schlom, “Addendum to
of potassium tantalate thin films,” J. Eur. Ceram. Soc. 25, 2155–2159 (2005).
141 ‘A strong ferroelectric ferromagnet created by means of spin-lattice coupling,’”
V. Skoromets, S. Glinšek, V. Bovtun, M. Kempa, J. Petzelt, S. Kamba, B. Malič, Nature 476, 114 (2011).
M. Kosec, and P. Kužel, “Ferroelectric phase transition in polycrystalline KTaO3 161
S. Kamba, V. Goian, M. Orlita, D. Nuzhnyy, J. H. Lee, D. G. Schlom, K. Z.
thin film revealed by terahertz spectroscopy,” Appl. Phys. Lett. 99, 052908 (2011).
142 Rushchanskii, M. Lezaic, T. Birol, C. J. Fennie, P. Gemeiner, B. Dkhil, V. Bovtun,
H. Vogt, “Evidence of defect-induced polarization clusters in nominally M. Kempa, J. Hlinka, and J. Petzelt, “Magnetodielectric effect and phonon proper-
pure KTaO3 from low-temperature Raman and hyper-Raman spectra,” J. Phys.: ties of compressively strained EuTiO3 thin films deposited on (001) (LaAlO3 )0.29 -
Condens. Matter 3, 3697–3709 (1991). (SrAl1/2 Ta1/2 O3 )0.71 ,” Phys. Rev. B 85, 094435 (2012).
143
S. Glinšek, D. Nuzhnyy, J. Petzelt, B. Malič, S. Kamba, V. Bovtun, M. Kempa, 162
D. Repček, C. Kadlec, F. Kadlec, M. Savinov, M. Kachlík, J. Drahokoupil,
V. Skoromets, P. Kužel, I. Gregora, and M. Kosec, “Lattice dynamics and broad- P. Proschek, J. Prokleška, K. Maca, and S. Kamba, “Seemingly anisotropic mag-
band dielectric properties of the KTaO3 ceramics,” J. Appl. Phys. 111, 104101 netodielectric effect in isotropic EuTiO3 ceramics,” Phys. Rev. B 102, 144402
(2012). (2020).
144
G. Catalan and J. F. Scott, “Physics and applications of bismuth ferrite,” Adv. 163
K. Z. Rushchanskii, S. Kamba, V. Goian, P. Vaněk, M. Savinov, J. Prokleška,
Mater. 21, 2463–2485 (2009). D. Nuzhnyy, K. Knížek, F. Laufek, S. Eckel, S. K. Lamoreaux, A. O. Sushkov, M.
145
S. Kamba, D. Nuzhnyy, M. Savinov, J. Šebek, J. Petzelt, J. Prokleška, R. Hau- Ležaić, and N. A. Spaldin, “A multiferroic material to search for the permanent
mont, and J. Kreisel, “Infrared and terahertz studies of polar phonons and mag- electric dipole moment of the electron,” Nat. Mater. 9, 649–654 (2010).
netodielectric effect in multiferroic BiFeO3 ceramics,” Phys. Rev. B 75, 024403 164
V. Goian, S. Kamba, D. Nuzhnyy, P. Vaněk, M. Kempa, V. Bovtun, K. Knížek,
(2007). J. Prokleška, F. Borodavka, M. Ledinský, and I. Gregora, “Dielectric, magnetic
146
Y. F. Popov, A. M. Kadomtseva, G. P. Vorob’ev, and A. K. Zvezdin, “Discovery and structural properties of novel multiferroic Eu0.5 Ba0.5 TiO3 ceramics,” J. Phys.:
of the linear magnetoelectric effect in magnetic ferroelectric BiFeO3 in a strong Condens. Matter 23, 025904 (2011).
magnetic field,” Ferroelectrics 162, 135–140 (1994). 165
V. Goian, S. Kamba, P. Vaněk, M. Savinov, C. Kadlec, and J. Prokleška,
147
I. A. Kornev, S. Lisenkov, R. Haumont, B. Dkhil, and L. Bellaiche, “Finite- “Magnetic and dielectric properties of multiferroic Eu0.5 Ba0.25 Sr0.25 TiO3
temperature properties of multiferroic BiFeO3 ,” Phys. Rev. Lett. 99, 227602 ceramics,” Phase Transitions 86, 191–199 (2013).
(2007). 166
S. Eckel, A. O. Sushkov, and S. K. Lamoreaux, “Limit on the electron electric
148
G. Marschick, J. Schell, B. Stöger, J. N. Gonçalves, M. O. Karabasov, dipole moment using paramagnetic ferroelectric Eu0.5 Ba0.5 TiO3 ,” Phys. Rev. Lett.
D. Zyabkin, A. Welker, M. Escobar C, D. Gärtner, I. Efe, R. A. Santos, J. E. M. 109, 193003 (2012).

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-21


© Author(s) 2021
APL Materials RESEARCH UPDATE scitation.org/journal/apm

167 189
E. Bousquet, N. A. Spaldin, and P. Ghosez, “Strain-induced ferroelectricity in S. Skiadopoulou, F. Borodavka, C. Kadlec, F. Kadlec, M. Retuerto, Z. Deng,
simple rocksalt binary oxides,” Phys. Rev. Lett. 104, 037601 (2010). M. Greenblatt, and S. Kamba, “Magnetoelectric excitations in multiferroic
168 Ni3 TeO6 ,” Phys. Rev. B 95, 184435 (2017).
V. Goian, R. Held, E. Bousquet, Y. Yuan, A. Melville, H. Zhou, V. Gopalan,
190
P. Ghosez, N. A. Spaldin, D. G. Schlom, and S. Kamba, “Making EuO multiferroic S. Skiadopoulou, M. Retuerto, F. Borodavka, C. Kadlec, F. Kadlec, M. Míšek,
by epitaxial strain engineering,” Commun. Mater. 1, 74 (2020). J. Prokleška, Z. Deng, X. Tan, C. Frank, J. A. Alonso, M. T. Fernandez-Diaz,
169
J. H. Lee and K. M. Rabe, “Epitaxial-strain-induced multiferroicity in SrMnO3 M. Croft, F. Orlandi, P. Manuel, E. Mccabe, D. Legut, M. Greenblatt, and
from first principles,” Phys. Rev. Lett. 104, 207204 (2010). S. Kamba, “Structural, magnetic, and spin dynamical properties of the polar
170 antiferromagnets Ni3−x Cox TeO6 (x = 1, 2),” Phys. Rev. B 101, 014429 (2020).
S. Kamba, V. Goian, V. Skoromets, J. Hejtmánek, V. Bovtun, M. Kempa,
191
F. Borodavka, P. Vaněk, A. A. Belik, J. H. Lee, O. Pacherová, and K. M. Rabe, F. Kadlec, V. Goian, C. Kadlec, M. Kempa, P. Vaněk, J. Taylor, S. Rols,
“Strong spin-phonon coupling in infrared and Raman spectra of SrMnO3 ,” Phys. J. Prokleška, M. Orlita, and S. Kamba, “Possible coupling between magnons and
Rev. B 89, 064308 (2014). phonons in multiferroic CaMn7 O12 ,” Phys. Rev. B 90, 054307 (2014).
171 192
C. Becher et al., “Strain-induced coupling of electrical polarization and struc- S. Kamba, V. Goian, F. Kadlec, D. Nuzhnyy, C. Kadlec, J. Vít, F. Borodavka,
tural defects in SrMnO3 films,” Nat. Nanotechnol. 10, 661–665 (2015). I. S. Glazkova, and A. A. Belik, “Changes in spin and lattice dynamics induced by
172
H. Sakai, J. Fujioka, T. Fukuda, D. Okuyama, D. Hashizume, F. Kagawa, magnetic and structural phase transitions in multiferroic SrMn7 O12 ,” Phys. Rev. B
H. Nakao, Y. Murakami, T. Arima, A. Q. R. Baron, Y. Taguchi, and Y. Tokura, 99, 184108 (2019).
“Displacement-type ferroelectricity with off-center magnetic ions in perovskite 193
J. Vít, F. Kadlec, C. Kadlec, F. Borodavka, Y. S. Chai, K. Zhai, Y. Sun, and
Sr1−x Bax MnO3 ,” Phys. Rev. Lett. 107, 137601 (2011). S. Kamba, “Electromagnon in the Y-type hexaferrite BaSrCoZnFe11 AlO22 ,” Phys.
173
V. Goian, F. Kadlec, C. Kadlec, B. Dabrowski, S. Kolesnik, O. Chmaissem, Rev. B 97, 134406 (2018).
D. Nuzhnyy, M. Kempa, V. Bovtun, M. Savinov, J. Hejtmánek, J. Prokleška, and 194
H. Shishikura, Y. Tokunaga, Y. Takahashi, R. Masuda, Y. Taguchi, Y. Kaneko,
S. Kamba, “Spectroscopic studies of the ferroelectric and magnetic phase tran- and Y. Tokura, “Electromagnon resonance at room temperature with gigantic
sitions in multiferroic Sr1−x Bax MnO3 ,” J. Phys.: Condens. Matter 28, 175901 magnetochromism,” Phys. Rev. Appl. 9, 044033 (2018).
(2016). 195
174 N. Kida, S. Kumakura, S. Ishiwata, Y. Taguchi, and Y. Tokura, “Gigantic
H. Sakai, J. Fujioka, T. Fukuda, M. S. Bahramy, D. Okuyama, R. Arita, terahertz magnetochromism via electromagnons in the hexaferrite magnet
T. Arima, A. Q. R. Baron, Y. Taguchi, and Y. Tokura, “Soft phonon mode Ba2 Mg2 Fe12 O22 ,” Phys. Rev. B 83, 064422 (2011).
coupled with antiferromagnetic order in incipient-ferroelectric Mott insulators 196
T. Nakajima, Y. Takahashi, S. Kibayashi, M. Matsuda, K. Kakurai, S. Ishiwata,
Sr1−x Bax MnO3 ,” Phys. Rev. B 86, 104407 (2012).
175 Y. Taguchi, Y. Tokura, and T.-h. Arima, “Electromagnon excitation in the field-
C. J. Fennie and K. M. Rabe, “Ferroelectric transition in YMnO3 from first
induced noncollinear ferrimagnetic phase of Ba2 Mg2 Fe12 O22 studied by polar-
principles,” Phys. Rev. B 72, 100103(R) (2005).
176 ized inelastic neutron scattering and terahertz time-domain optical spectroscopy,”
S. Kamba, D. Nuzhnyy, M. Savinov, P. Toledano, V. Laguta, P. Brazda, Phys. Rev. B 93, 035119 (2016).
L. Palatinus, F. Kadlec, F. Borodavka, C. Kadlec, P. Bednyakov, V. Bovtun, 197
F. Kadlec, C. Kadlec, J. Vít, F. Borodavka, M. Kempa, J. Prokleška, J. Buršík,

30 October 2024 15:21:18


M. Kempa, D. Kriegner, J. Drahokoupil, J. Kroupa, J. Prokleska, K. Chapagain,
R. Uhrecký, S. Rols, Y. S. Chai, K. Zhai, Y. Sun, J. Drahokoupil, V. Goian, and S.
B. Dabrowski, and V. Goian, “Unusual ferroelectric and magnetic phases in
Kamba, “Electromagnon in the Z-type hexaferrite (Bax Sr1−x )3 Co2 Fe24 O41 ,” Phys.
multiferroic 2H-BaMnO3 ceramics,” Phys. Rev. B 95, 174103 (2017).
177 Rev. B 94, 024419 (2016).
J. Varignon and P. Ghosez, “Improper ferroelectricity and multiferroism in 198
2H-BaMnO3 ,” Phys. Rev. B 87, 140403(R) (2013). S. H. Chun, K. W. Shin, H. J. Kim, S. Jung, J. Park, Y.-M. Bahk, H.-R.
178 Park, J. Kyoung, D.-H. Choi, D.-S. Kim, G.-S. Park, J. F. Mitchell, and K. H.
T. Kimura, T. Goto, H. Shintani, K. Ishizaka, T. Arima, and Y. Tokura,
Kim, “Electromagnon with sensitive terahertz magnetochromism in a room-
“Magnetic control of FE polarization,” Nature 426, 55–58 (2003).
179 temperature magnetoelectric hexaferrite,” Phys. Rev. Lett. 120, 027202 (2018).
Y. Tokura, S. Seki, and N. Nagaosa, “Multiferroics of spin origin,” Rep. Prog. 199
Phys. 77, 076501 (2014). Y. S. Chai, S. H. Chun, J. Z. Cong, and K. H. Kim, “Magnetoelectricity in multi-
180
P. Tolédano, “Pseudo-proper ferroelectricity and magnetoelectric effects in ferroic hexaferrites as understood by crystal symmetry analyses,” Phys. Rev. B 98,
TbMnO3 ,” Phys. Rev. B 79, 094416 (2009). 104416 (2018).
200
181
D. Niermann, C. P. Grams, P. Becker, L. Bohaty, H. Schenck, and J. Hemberger, K. Zhai, Y. Wu, S. Shen, W. Tian, H. Cao, Y. Chai, B. C. Chakoumakos,
“Critical slowing down near the multiferroic phase transition in MnWO4 ,” Phys. D. Shang, L. Yan, F. Wang, and Y. Sun, “Giant magnetoelectric effects achieved
Rev. Lett. 114, 037204 (2015). by tuning spin cone symmetry in Y-type hexaferrites,” Nat. Commun. 8, 519
182 (2017).
A. Pimenov, A. A. Mukhin, V. Y. Ivanov, V. D. Travkin, A. M. Balbashov, and 201
A. Loidl, “Possible evidence for electromagnons in multiferroic manganites,” Nat. S. H. Chun, Y. S. Chai, B.-G. Jeon, H. J. Kim, Y. S. Oh, I. Kim, H. Kim, B. J. Jeon,
Phys. 2, 97–100 (2006). S. Y. Haam, J.-Y. Park, S. H. Lee, J.-H. Chung, J.-H. Park, and K. H. Kim, “Electric
183 field control of nonvolatile four-state magnetization at room temperature,” Phys.
N. Kida, Y. Takahashi, J. S. Lee, R. Shimano, Y. Yamasaki, Y. Kaneko,
Rev. Lett. 108, 177201 (2012).
S. Miyahara, N. Furukawa, T. Arima, and Y. Tokura, “Terahertz time-domain 202
spectroscopy of electromagnons in multiferroic perovskite manganites,” J. Opt. V. Laguta, M. Kempa, V. Bovtun, J. Buršík, K. Zhai, Y. Sun, and S. Kamba,
Soc. Am. B 26, A35–A51 (2009). “Magnetoelectric coupling in multiferroic Z-type hexaferrite revealed by electric-
184
A. M. Shuvaev, A. A. Mukhin, and A. Pimenov, “Magnetic and magnetoelec- field-modulated magnetic resonance studies,” J. Mater. Sci. 55, 7624–7633
tric excitations in multiferroic manganites,” J. Phys.: Condens. Matter 23, 113201 (2020).
203
(2011). R. D. Johnson, L. C. Chapon, D. D. Khalyavin, P. Manuel, P. G. Radaelli, and
185
V. N. Krivoruchko, “Electrically active magnetic excitations in C. Martin, “Giant improper ferroelectricity in the ferroaxial magnet CaMn7 O12 ,”
antiferromagnets,” Low Temp. Phys. 38, 807 (2012). Phys. Rev. Lett. 108, 067201 (2012).
186 204
S. Bordács, I. Kézsmárki, D. Szaller, L. Demkó, N. Kida, H. Murakawa, Y. Y. S. Glazkova, N. Terada, Y. Matsushita, Y. Katsuya, M. Tanaka, A. V. Sobolev,
Onose, R. Shimano, T. Rõõm, U. Nagel, S. Miyahara, N. Furukawa, and Y. Tokura, I. A. Presniakov, and A. A. Belik, “High-pressure synthesis, crystal structures, and
“Chirality of matter shows up via spin excitations,” Nat. Phys. 8, 734–738 (2012). properties of CdMn7 O12 and SrMn7 O12 perovskites,” Inorg. Chem. 54, 9081–9091
187
I. Kezsmarki, D. Szaller, S. Bordacs, V. Kocsis, Y. Tokunaga, Y. Taguchi, H. (2015).
205
Murakawa, Y. Tokura, H. Engelkamp, T. Room, and U. Nagel, “One-way trans- A. A. Belik, Y. S. Glazkova, N. Terada, Y. Matsushita, A. V. Sobolev, I. A.
parency of four-coloured spin-wave excitations in multiferroic materials,” Nat. Presniakov, N. Tsujii, S. Nimori, K. Takehana, and Y. Imanaka, “Spin-driven
Commun. 5, 3203 (2014). multiferroic properties of PbMn7 O12 perovskite,” Inorg. Chem. 55, 6169–6177
188 (2016).
C. Kadlec, F. Kadlec, V. Goian, M. Gich, M. Kempa, S. Rols, M. Savinov,
206
J. Prokleska, M. Orlita, and S. Kamba, “Electromagnon in ferrimagnetic ε-Fe2 O3 M. Horioka, Y. Satuma, and H. Yanagihara, “Dielectric dispersion and soft
nanograin ceramics,” Phys. Rev. B 88, 104301 (2013). mode in Rochelle salt,” J. Phys. Soc. Jpn. 62, 2233–2236 (1993).

APL Mater. 9, 020704 (2021); doi: 10.1063/5.0036066 9, 020704-22


© Author(s) 2021

You might also like