You are on page 1of 37

MR43CH16-Tagantsev

ARI

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

ANNUAL
REVIEWS

24 May 2013

15:27

Further

Click here for quick links to


Annual Reviews content online,
including:
Other articles in this volume
Top cited articles
Top downloaded articles
Our comprehensive search

Flexoelectric Effect in Solids


Pavlo Zubko,1 Gustau Catalan,2,3
and Alexander K. Tagantsev4
1
Department of Condensed Matter Physics, University of Geneva, CH-1211 Geneva 4,
Switzerland; email: pavlo.zubko@unige.ch
2

Institut Catal`a de Recerca i Estudis Avancats (ICREA), Barcelona 08010, Catalunya, Spain

Institut Catal`a de Nanoci`encia i Nanotecnologia (ICN2), CSIC-ICN, Campus Universitat


Autonoma de Barcelona, Bellaterra 08193, Spain; email: gustau.catalan@cin2.es
4
Ceramics Laboratory, Swiss Federal Institute of Technology (EPFL), Lausanne 1015,
Switzerland; email: alexander.tagantsev@ep.ch

Annu. Rev. Mater. Res. 2013. 43:387421

Keywords

The Annual Review of Materials Research is online at


matsci.annualreviews.org

strain gradients, electromechanical coupling, piezoelectricity,


ferroelectricity, domain walls

This articles doi:


10.1146/annurev-matsci-071312-121634
c 2013 by Annual Reviews.
Copyright 
All rights reserved

Abstract
Flexoelectricitythe coupling between polarization and strain gradientsis
a universal effect allowed by symmetry in all materials. Following its discovery several decades ago, studies of exoelectricity in solids have been scarce
due to the seemingly small magnitude of this effect in bulk samples. The development of nanoscale technologies, however, has renewed the interest in
exoelectricity, as the large strain gradients often present at the nanoscale can
lead to strong exoelectric effects. Here we review the fundamentals of the
exoelectric effect in solids, discuss its presence in many nanoscale systems,
and look at potential applications of this electromechanical phenomenon.
The review also emphasizes the many open questions and unresolved issues
in this developing eld.

387

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

1. INTRODUCTION

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Piezoelectricity: the
linear response of
polarization to
mechanical strain and
vice versa; possible
only in
noncentrosymmetric
materials

Flexoelectricity is a property of all insulators whereby they polarize when subject to an inhomogeneous deformation. The exoelectric coupling is between polarization and strain gradient, rather
than between polarization and homogeneous strain. This difference is crucial to understanding
both the advantages and the limitations of exoelectricity compared with its close relative,
piezoelectricity.
Strain, like stress, does not by itself break centrosymmetry: If a material is centrosymmetric
to start with, it will continue to be centrosymmetric under a homogeneous deformation. This
is intuitively clear: When a plate of a centrosymmetric material is subjected to a homogeneous
deformation (Figure 1a), we cannot rationalize the appearance of polarization, because there is no
preferred sense of direction for the polarization vector. By contrast, a strain gradient does break
centrosymmetry. Under a strain gradient (e.g., one resulting from bending, as in Figure 1b),
the top and bottom surfaces of the plate are no longer equivalent and therefore dene a sense
of direction for the induced polarization vector. Mathematically, the exoelectric effect is controlled by a fourth-rank tensor and is therefore allowed in materials of any symmetry, whereas
piezoelectricity is controlled by a third-rank tensor, which is allowed only in materials that are
noncentrosymmetric. In piezoelectrics, the polarization thus occurs due to the low symmetry of
the material rather than due to the symmetry-breaking effect of the perturbation.
Flexoelectricity as a strain gradientdriven breaking of the local centrosymmetry can also be
visualized at the microscopic level. For a simple ionic lattice, such as that sketched in Figure 1b, a
vertical gradient of in-plane strain (induced, for instance, by bending) may cause the central cation
to be squeezed up like a pea inside a peapod, breaking the local centrosymmetry and inducing
polarity. This analogy, based purely on steric considerations, is crude but close to the idea of
Bursian & Zaikovskii (1) that the central Ti ion in the ferroelectric perovskite BaTiO3 must shift
up on bending or, vice versa, that bending must appear due to the presence of a Ti ion in the
upper side of the unit cell, which causes it to expand while the more empty lower half contracts.

P=0

P0

Figure 1
Cartoon illustrating the microscopic mechanism of exoelectricity. (a) Uniform strain (e.g., that due to
uniaxial compression) does not break inversion symmetry and hence cannot generate polarization in a
centrosymmetric material. (b) A strain gradient (e.g., that due to bending), however, does break inversion
symmetry and can therefore lead to relative displacements of the centers of negative and positive charges and
to the appearance of polarization in any material.

388

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

An alternative picture is provided by Harris (2), who noticed that the strain gradient in a shock
wave makes unequal the distances between the atomic planes, resulting in a local breaking of
centrosymmetry.
The above are all simple ionic pictures, with rigid ions shifting and causing polarization. The
exoelectric effect is nevertheless a subtle physical phenomenon, and its intuitive vision is often
deceptive. For example, the conclusion about the sign of the exoelectric response that one might
draw from the cartoon shown in Figure 1b may readily be wrong (3, 4).
Of course, strain gradients do not affect only ionic positionsan asymmetric redistribution of
the electron density will take place as well, contributing to the total polarization. The exoelectricity in graphene (5) is controlled by this mechanism. The ionic and electronic components of
exoelectricity are complementary, but in this article we keep an ionic picture for simplicity. We
also exclude from this article the exoelectricity of liquid crystals (6) [which is actually the eld in
which the term exoelectricity was rst coined and was later adopted by Indenbom et al. (7) for
solid dielectrics] and biological materials (810), as they originate from different physics and have
very different applications (11).
So why should we care about exoelectricity? Electromechanical properties play an essential
role in the physics of solids and their practical application. Until recently, when referring to
electromechanical properties, one generally meant piezoelectricity and electrostriction, whereas
exoelectricity was hardly mentioned on account of its relative weakness. However, it is becoming
clear that this neglect is not justied, for several reasons. First, exoelectricity, in contrast to piezoelectricity, is a universal property allowed by symmetry in any structure and therefore broadens the
choice of materials that can be used for electromechanical sensors and actuators. Second, reduced
dimensions imply larger gradients: A strain difference over a small distance gives a large strain
gradient. The small length scales involved in nanotechnology thus lead to the growing impact of
exoelectricity, which at the nanoscale may even be competitive with that of piezoelectricity. In
addition, a number of experiments have reported giant exoelectric coupling constants exceeding
theoretical estimates by several orders of magnitude. Finally, the polar nature of the exoelectric effect means that strain gradients can effectively play the role of an equivalent electric eld
and can be used, for example, to switch the spontaneous polarization of a ferroelectric material.
By contrast, switching of polarization by homogeneous strain is allowed by symmetry only in a
limited class of ferroelectrics (those that are piezoelectrics in the paraelectric phase). Therefore,
exoelectricity is not just a substitute for piezoelectricity at the nanoscale; it also enables additional
electromechanical functionalities not available otherwise.
Several reviews on exoelectricity have been published, and the reader is referred to References 3, 4, and 1216. The subject is, however, evolving very rapidly, with many new developments
happening in the past few years, particularly in the area of nanoscale materials such as thin lms.
Here we therefore focus more attention on recent developments, both in experiment and in theory, as well as on controversial or unresolved issues. We begin with a brief introduction to the
phenomenological theory of bulk exoelectricity in Section 2, where many of the characteristic
features of exoelectricity are derived. The concepts discussed here are referred to throughout the
rest of the review. In Section 3 we focus on the importance of nite size effects that are inherent
in all real exoelectric systems. In Section 4, we then turn our attention to the magnitude of the
exoelectric response and discuss the progress made and challenges faced in the development of
a microscopic theory of exoelectricity. Section 5 is devoted to experimental studies of exoelectricity in both bulk and nanoscale materials, and Section 6 addresses possible applications based
on the exoelectric effect. Finally, we conclude this review with a brief summary of open questions
and proposed future research in Section 7.

www.annualreviews.org Flexoelectric Effect in Solids

Electrostriction: the
contribution to the
strain proportional to
the square of the
polarization (or the
square of an external
electric eld); possible
in any material

389

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

2. PHENOMENOLOGICAL THEORY OF THE BULK EFFECT


Theoretical work on exoelectricity dates back to the seminal papers by Mashkevich & Tolpygo
(17, 18), who rst proposed the effect, and by Kogan (19), who formulated the rst phenomenological theory. For a historical overview of the early developments in the theory of exoelectricity,
the reader is referred to the comprehensive review by Tagantsev (4) as well as to the more recent,
concise summary by Maranganti et al. (13). Here we briey summarize the basic phenomenological description of the direct and converse exoelectric effects, emphasizing the differences between
exoelectricity and piezoelectricity, and the analogies between strain gradients and electric elds.

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

2.1. Static Response


In contrast to the piezoelectric response, the exoelectric effect in the static (e.g., in a bent plate)
and dynamic (e.g., in a sound wave) situations generally requires separate treatments (4, 20). Let
us start with the static case.
2.1.1. Constitutive equations. We introduce the static exoelectric effect as a linear response
of polarization Pi to a strain gradient u kl / x j in the absence of electric eld. It is governed by a
fourth-rank exoelectric tensor, ijkl ,


Pi
,
1.
klij =
(u kl / x j ) E=0
where E denotes electric eld. This effect can readily be incorporated into the Landau phenomenological framework. Despite the need for tensors for the description of exoelectricity, an insight
into this phenomenon can be gained by considering a 1D model involving one component of
polarization and one of strain, where the tensor sufxes can be omitted.
In such a model, the macroscopic description of the static bulk exoelectric response can be
obtained by generalizing the thermodynamic potential used for the description of the piezoelectric
response by introducing a linear coupling between the polarization and strain gradient and vice
versa into the system. In the most general form, a thermodynamic potential density suitable for
such a description reads
G =

1 2 c 2
u
P
P + u Pu f1 P
f2 u
PE u,
2
2
x
x

2.

where is the relevant stress component. This expansion does not contain anharmonic terms
(i.e., the electrostriction term is omitted). When nonlinear effects are of interest, these terms can
readily be incorporated into the framework, as in, for example, References 2123. In particular,
electrostrictive coupling terms are always allowed by symmetry and are essential for the description
of perovskite ferroelectrics; for these materials, the linear piezoelectric term Pu in Equation 2
has to be substituted by the electrostrictive coupling term quP2 (24).
If we set to zero the coefcients for the gradient-containing terms, the bulk equations of state
of the material can be found by a simple minimization of the potential density in Equation 2 with
respect to the polarization and strain. Such minimization leads to the standard linear electromechanical equations of a piezoelectric:
P = E + eu,

3.

= eE + c u,

4.

where e = is the so-called strain-charge piezoelectric coefcient and c E = c 2 is the elastic


constant at xed electric eld. From this we see that the term Pu in Equation 2 controls the bulk
390

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

piezoelectric response. We drop this term in further discussion so as to see the electromechanical
consequences of pure exoelectricity without piezoelectricity. However, this discussion can be
readily generalized to piezoelectrics by taking this term into account.
Thus, we address exoelectricity by using the thermodynamic potential density from
Equation 2 with = 0. Here it is convenient to present G as the sum of two contributions:
f1 + f2 (u P)
,
2
x


P
f
1 2 c 2
u
P + u
u
=
P
PE u,
2
2
2
x
x
Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org
Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

G = 

5.
6.

where f f1 f2 is the exocoupling coefcient (a tensor in the general case).


Now that the potential density contains gradient terms, to get the equation of state, one should

minimize the thermodynamic potential of the sample as a whole, G d V (integrating over the
volume of the sample); i.e, one should apply the Euler equations G / A ddx (G /( A/ x)) =
0, where A stands for the variables of the problem (in this case P and u). Such minimization yields
the bulk constitutive electromechanical equations (7)
u
,
x

7.

P
,
x

8.

P = E +
= cu +

f.

9.

In Equation 7, the rst right-hand-side term describes the dielectric response with the clamped
dielectric susceptibility , and the second right-hand-side term describes the exoelectric response
with the exoelectric coefcient . In Equation 8, the rst right-hand-side term describes Hooks
law, with the elastic constant at xed polarization c, and the second right-hand-side term describes
the converse exoelectric response. The latter term has the physical meaning of a linear response
of stress (or strain) to a polarization gradient in a mechanically free sample and was studied by
Mindlin (25) and others within the mechanics-of-materials community, in which the theory of
(converse) exoelectricity appears to have developed independently of that of the condensed matter
physics community outlined here (for details, see Reference 13). Converse exoelectricity may be
particularly important in piezoresponse force microscopy experiments in which the voltage applied
to the conducting tip of an atomic force microscope gives rise to large and highly nonuniform
electric elds that, through converse exoelectricity, may give rise to an apparent piezoelectric
response, even in nonpiezoelectric materials (26).
u
As is clear from Equation 7, the vector f ux (or fijkl xijl in full tensor notation) has the same effect
on polarization as the external electric eld E. It is sometimes termed the exoelectric eld and
is a convenient concept when one is considering the strength of exoelectric poling effects. For
example, when polarization switching or poling caused by strain gradients is discussed (2731), it
is this quantity that should be compared with the coercive or built-in electric elds.
The last term in Equation 5 does not contribute to the bulk constitutive electromechanical
equations. This can be concluded directly from the fact that its contribution to the thermodynamic potential of the sample can be transformed to an integral over the surface of the sample:

12 ( f1 + f2 ) uPdS. Thus, the thermodynamic potential density in Equation 6 provides a full
phenomenological description of the static bulk exoelectric effect. Consequently, both the direct and converse exoelectric effects in Equations 7 and 8 are described by the same coefcient
f f1 f2 .
www.annualreviews.org Flexoelectric Effect in Solids

391

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

The bulk exoelectric effect describes a strain gradientinduced polarization analogous to the
strain-induced polarization of the piezoelectric effect. The constitutive equations (Equations 79)
also highlight an important feature of the bulk exoelectric effect: Equation 9 explicitly shows that
the bulk exoelectric coefcient is proportional to the dielectric permittivity of the material (7,
19, 32, 33), suggesting that the exoelectric response should be enhanced in materials with high
dielectric constants (high-K materials) such as ferroelectrics.

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

2.1.2. Direct and converse flexoelectricity. It is instructive to rewrite the constitutive equations
by taking into account that the exoelectric effect is relatively weak so that P in Equation 8 can be
replaced with E. This gives a set of constitutive equations,
P = E +

u
,
x

10a.

E
+ c u,
10b.
x
suitable for comparison with those for the piezoelectric response, which for our one-component
1D model are given by Equations 3 and 4. Although for both effects the direct and converse
responses are controlled by exactly the same coefcient, there exists a strong asymmetry between
the converse and direct exoelectric effects. Specically, for the direct effect, in the absence of an
electric eld, a strain gradient induces a homogeneous polarization. However, for the converse
effect, in a mechanically free sample, a homogeneous electric eld does not cause a strain gradient
or any other linear mechanical response (of course, nonlinear effects such as electrostriction still
exist). This asymmetry is in strong contrast to the symmetric piezoelectric effect, in which strain
induces polarization and electric eld induces stress, as is clear from Equations 3 and 4.
As pointed out in Section 1, however, exoelectricity is a subtle phenomenon, and the aforementioned asymmetry of the bulk exoelectric effect does not lead to an asymmetry of the linear
electromechanical response of a sample as a whole. Actually, a homogeneous electric eld does
cause an inhomogeneous deformation for nite-size samples, as is discussed in Section 3.2.
It is worth making a note on terminology: Whereas in the liquid crystals community the term
converse exoelectricity generally refers to bending of the material induced by an electric eld, in
the solid-state community this term is generally reserved for the appearance of stress or strain due
to a polarization gradient, as discussed above [although Indenbom et al. (7) used the term inverse
exoelectric effect to refer to polarization-induced bending in their article].
=

2.2. Dynamic Response


The dynamic bulk exoelectric response can be described by minimizing the action (T  +
u )d Vd t (the integral is taken over the volume of the sample and time). Here  comes from
Equation 6, and the kinetic energy density is dened as (20)
T =

U + M U P,
2

11.

where and U stand for the density and the acoustic displacement, respectively (in the 1D model,
u = U / x); the dot over P and U refers to the time derivative. Such minimization with respect
to polarization and strain leads to the following equations of motion for these variables:
u

M U,
x
u
2P
U = c
M P +
.
x
x2
P = E +

392

Zubko

Catalan

Tagantsev

12.
13.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

= s + d

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Figure 2
Illustration of the dynamic exoelectric effect. An acoustic wave passing through a medium generates a
time-dependent strain gradient that generates a polar displacement of the ions. For an acoustic wave, the
acceleration of the atoms is proportional to the strain gradient. Thus, on top of the static exoelectric
response, s ux , the acceleration of ions of different masses gives rise to an additional polar displacement,
2

d (m1 m2 ) tU2 (m1 m2 ) ux .

In Equation 13, the last two terms play an essential role in the lattice dynamics, controlling, for
example, the specic shape of the dispersion curve for acoustic phonons in perovskite ferroelectrics
(34). However, for macroscopic situations, in which the typical spatial scales are much larger
than the lattice constant, these terms can be neglected. Then, eliminating U from the set of
Equations 12 and 13, we get the following equation for the polarization accompanying the strain
gradient in a moving medium (e.g., in the case of an acoustic wave):
P = E + ( + d )

u
,
x

14.

where
d = c M/

15.

is the coefcient describing the dynamic exoelectric response. The physical meaning of the
dynamic exoelectric effect is clear from Equation 12: It is the polarization induced by the acceleration of the medium. Melnikovsky (35) also recently discussed polarization due to acceleration.
On the microscopic level, the dynamic exoelectric effect can be related to the mass difference of
the ions constituting the material (2, 20).
The M coefcient can be calculated in terms of the dynamic lattice theory (20). Figure 2
illustrates the static and dynamic contributions to the polarization in an acoustic wave. As is clear
from Equation 15, the dynamic contribution scales as the bulk dielectric constant of the material,
just like the static one. Of key importance is that these contributions are expected to be of the same
order of magnitude, according to estimates (20). The dynamic contribution to the exoelectric
effect makes the exoelectric effect qualitatively different from the piezoelectric effect, for which
the polarization and strain in a moving medium are linked by the same relation as in the static
case, i.e., P = E + eu.
The above treatment gives a qualitatively correct picture of the exoelectric effect in a sound
wave in which the amplitude of the polarization is controlled by the sum of the static and dynamic
contributions. Remarkably, both exoelectric coefcients are independent of the frequency of
the wave. Unlike the exoelectric coefcients, however, the magnitude of the polarization does
depend on frequency, as the strain gradients in two acoustic waves of the same amplitude but
different frequency will be different.
For the quasi-static case, however, our 1D treatment is oversimplied. Here a full tensorial
treatment is needed to evaluate the contribution of dynamic exoelectricity to the total polarization
response (P. V. Yudin & A. K. Tagantsev, submitted manuscript). It can then be shown that this
www.annualreviews.org Flexoelectric Effect in Solids

393

ARI

24 May 2013

15:27

contribution can be neglected in the quasi-static regime, i.e., when the modulation frequency is
lower than that of the relevant acoustic resonance of the sample. The treatment of this effect is
formally similar to that of the piezoelectric resonance in a nite sample.
The framework formulated above for the 1D model can readily be generalized to the real 3D
situation. Now the coefcients f, , and d become fourth-rank tensors symmetric with respect
to the permutation of the rst pair of sufxes. As these tensors are of even rank, they can be
nonzero in materials of any symmetry, including amorphous substances (36). The number of their
independent components is controlled by the symmetry of the material (see, e.g., References 37
39). For example, this number is two for isotropic materials and three for nonpiezoelectric cubic
materials. Because these tensors do not exhibit permutation symmetry with respect to the second
pair of sufxes, the two-sufx Voigt notation cannot be consistently introduced. Nevertheless, this
article uses such two-sufx representation (e.g., 11 1111 ), as is customarily done in papers on
exoelectricity.

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

3. FLEXOELECTRICITY IN FINITE SAMPLES


Real samples are always nite, and thus a complete phenomenological description must include
the effects of surfaces or interfaces. In this section we briey discuss two effects associated with
the nite size of a real sample: (a) the contribution of surfaces to bending-induced polarization
and (b) the polarization-induced bending that appears in thin samples. Although our discussion
explicitly focuses on the model system of a cylindrically bent capacitor, the results derived here have
important implications for all nite exoelectric systems, including thin lms and nanodevices,
which are discussed in the later sections.

3.1. Contribution of Surface Piezoelectricity


In a nite sample there will always be surface contributions to any effect. Typically, these are small,
as they are usually controlled by the surface/volume ratio, but in some cases they can compete
with the bulk contribution of another, weaker effect. For example, in centrosymmetric materials,
due to the symmetry-breaking impact of the surface, a thin surface layer of thickness generally
becomes piezoelectric and can mimic the bulk exoelectric response. Let us discuss this effect
by considering the cylindrical bending of the thin parallel-plate capacitor sketched in Figure 3.
The piezoelectric coefcients, e, of the surface layers on the opposite sides of the plate should be
of opposite signs (as controlled by the opposite orientation of the surface normal); the same is
valid for the bending-induced strains in these layers. Therefore, the induced polarizations in these
layers are of the same sign.
The normal component of the electric displacement D = 0 E + P across any dielectric interface
must be preserved, i.e., Dz = P + 0 E = Ph + 0 Eh , where 0 is the vacuum permittivity.
Therefore, the presence of a polarization P within the piezoelectric surface layers must give rise
to internal electric elds in the sample, thus polarizing it. For a short-circuited capacitor, the
potential difference = 2E + h Eh across the capacitor must vanish; here h is the thickness of
the nonpiezoelectric bulk (see Figure 3). The electric displacement induced by the strain gradient
can then be calculated as (40)
u 11
hh
,
16.
D = e
2h + h x3
where = 0 + is the dielectric constant of the surface layer and h is that of the bulk. For
thin-enough surface layers (/ h  /h ), Equation 16 yields the effective exoelectric coefcient
394

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Ph

Eh

Ez

Figure 3
Surface piezoelectricity. Upon bending, as shown on the left, the tensile/compressive strains in the
top/bottom surface layers give rise to a polarization P in the piezoelectric surface layers of thickness .
Because the normal component of the electric displacement must remain constant, this surface polarization
gives rise to electric elds Eh and thus to a polarization Ph within the nonpiezoelectric bulk. The measured
average polarization of the whole structure therefore depends not only on the dielectric properties of the
piezoelectric surface layers but also on those of the bulk. The potential and eld E proles along the
sample thickness are shown on the right.

associated with surface piezoelectricity:


eff
1133 = e

h
.

17.

Thus, the polarization of the system arising from surface piezoelectricity is sensitive to the bulk
value of the dielectric constant and, for thin-enough surfaces layers, is independent of the surface/
volume ratio.
Finally, to evaluate the size of this surface effect, let us get an estimate for the effective exocoupling coefcient f eff eff /h e/ . For a conservative lower-bound estimate, we consider
the surface layer to be atomically thin ( = a few angstroms). Then, using e 1 C m2 and
/0 10, we nd f eff of a few volts. This value is approximately the typical value of the components of the exocoupling tensor fijkl 110 V (see Section 4). Thus, surface piezoelectricity can
readily compete with bulk exoelectricity.

3.2. Polarization-Induced Bending


As discussed in Section 2, the electromechanical constitutive equations (Equations 10a and 10b)
describing the bulk exoelectric effect suggest an asymmetry between the direct and converse
exoelectric responses. Specically, in the absence of an electric eld, a strain gradient induces
a homogeneous polarization, whereas a homogeneous electric eld does not induce a strain
gradient. Meanwhile, a thermodynamic analysis by Bursian & Trunov (33) shows that this is
not the case and that a plate of centrosymmetric material should bend under the action of the
eld due to exoelectric coupling, as experimentally observed in BaTiO3 platelets in both the
paraelectric and ferroelectric phases (1).
The reason for this discrepancy is that the polarization-induced bending predicted by Bursian
& Trunov (33) is a nonlocal effect that can be obtained only by considering the thermodynamics
of the nite-size sample as a whole. Thus, this effect is not captured by the local theory developed
in Section 2.1.
www.annualreviews.org Flexoelectric Effect in Solids

395

ARI

24 May 2013

15:27

The origin of the polarization-induced bending moment can be understood by considering


what happens at the sample surface. If the polarization at the surface of the plate changes
continuously from its bulk value P to zero outside the plate, then according to Equation 8, the
polarization gradients at the plate surfaces must create some stress via the converse exoelectric
effect, resulting in forces applied to the opposite surfaces of the plate (40). The mechanical
moment of these forces causes the bending predicted by the Bursian-Trunov approach. The
Bursian-Trunov theory, although written as a bulk theory, actually treats the total free energy
of the plate, implicitly taking into account the surface effects (i.e., the aforementioned forces
applied to the surfaces). A similar bending effect was considered by Eliseev et al. (41). It can thus
be shown that a bending-mode exoelectric sensor working as an actuator will be characterized
by the same effective piezoelectric constant (40).
For more general boundary conditions, the situation is more complex and requires the use of
modied mechanical boundary conditions (40, 42). Modied boundary conditions for the polarization were derived by Indenbom et al. (7) and by Eliseev et al. (41), whereas modied mechanical
boundary conditions were obtained by Yurkov (42). The paper by Indenbom et al. (7) also gave
alternative modied mechanical boundary conditions, although their explicit derivation was not
provided. Failure to take into account these exoelectricity-induced modications of mechanical
boundary conditions can lead to explicit conicts with thermodynamics. This applies, for instance,
to the results in Reference 41, in which the validity of classical mechanical boundary conditions
was postulated. In view of the importance of the modied boundary conditions, it would now
be useful to reexamine the ideas of Cross and coworkers [who proposed that, in piezoelectric
composites based on the exoelectric effect, a separation of the direct and converse piezoelectric
effects can be achieved (12, 118)] by explicitly taking the boundary conditions into account. The
experimental observation of polarization-induced bending is discussed in Section 5.2.4.

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

4. MICROSCOPIC CALCULATIONS OF BULK FLEXOELECTRICITY


Above, we do not address the magnitude of the exoelectric effect. A rough, order-of-magnitude,
theoretical estimate of the exocoupling coefcient was rst provided by Kogan (19). Subsequently, other approaches led to essentially the same result (2, 45). The most straightforward way
to arrive at Kogans estimate is to consider a simple lattice of point charges q with interatomic
spacing a. Let this lattice be distorted by an atomic-scale strain gradient of order 1/a and with
an atomic-scale polarization of the order of (ea)/a3 . Such a strong perturbation is expected to
2
modify the energy density in the material, which is of the order of 4q0 a a13 , by a comparable
amount. Assigning this energy change to the exoelectric term f P ux yields Kogans estimate of
the exocoupling coefcient,
f q /(4 0 a) 110 V,

18.

using the electronic charge for q and a of the order of a few angstroms.
Since Kogans (19) seminal work in 1964, there have been a number of attempts to properly
quantify the exoelectric response in solids. On the microscopic level, static bulk exoelectricity
can be viewed as strain gradientdriven charge redistribution in the material. Roughly speaking,
one can distinguish two contributions to this redistribution: ionic and electronic. The rst
attempts to quantify the ionic contribution were undertaken by Askar et al. (46), who used
a shell model for the lattice dynamics of ionic crystals to calculate a full set of exocoupling
coefcients for a number of biatomic cubic crystals. In the late 1980s, Tagantsev (4, 20) then
developed a microscopic theory of ionic exoelectricity based on a rigid-ion model. The rst steps
toward a microscopic description of the electronic contribution to exoelectricity were recently
396

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

made by Resta (47) and Hong & Vanderbilt (48), who generalized Martins (49) treatment of
piezoelectricity to the exoelectric effect.
To gain some microscopic understanding of the ionic contribution to exoelectricity, let us
begin by briey discussing the theory developed by Tagantsev (4, 20). When a crystalline lattice
is deformed, the ith component of the displacement of the nth atom, wn,i , can be expressed as a
sum of two terms:
 Rn, j
U i
int
d y j + wn,i
,
19.
wn,i =
xj
x 0j
where x 0j and Rn, j are the coordinates of an immobile reference point and that of the nth atom
before the deformation. Here, summation over the three Cartesian coordinates is assumed for
all repeating sufxes. The rst right-hand-side term in Equation 19 represents the contribution
of the unsymmetrized strain U i / x j taken in the elastic medium approximation, also known as
external strain. In the case of homogeneous strain, the external strain has a simple meaning: The
displacement of each atom is proportional to the distance to it from a certain immobile point. For
a material in which all atoms are centers of inversion, the external strain is sufcient to describe all
atomic displacements caused by a homogeneous deformation. Consider, for instance, the simple
centrosymmetric diatomic lattice sketched in Figure 4a with c = 2a. Under the homogeneous
deformation shown in Figure 4b, centrosymmetry is preserved (c = 2a), and the displacements
of both types of atoms follow the external strain eld, which is represented by the deformation of
the square lattice in Figure 4a.
However, if the material is noncentrosymmetric or if the deformation is inhomogeneous,
the two types of atoms are no longer constrained by symmetry and may undergo additional
internal displacements within the deformed unit cell. These internal displacementsthe difference between the real displacement of an atom and that calculated from the local external
strainare known as internal strains (50) and are described by the second right-hand-side term in
Equation 19. Figure 4c shows the same centrosymmetric atomic lattice stretched inhomogeneously along x1 . The longitudinal strain gradient breaks the centrosymmetry, and thus the displacements of the atoms no longer have to follow the external strain eld.
For a piezoelectric (noncentrosymmetric) material, the internal strains are responsible for the
change in polarization under homogeneous strain and are, to lowest order, proportional to the
macroscopic strain u ij = (U i / x j + U j / xi )/2. For a centrosymmetric material, the internal
strain of the nth atom is, to lowest order, proportional to the elastic gradient:
int
ex
ikl
wn,
j = wn, j N n, j

u ik
.
xl

20.

If the body is considered to be made up of point charges Qn (this includes all charges, not only
ions) with the coordinates Rn,i , the response of the polarization to an inhomogeneous deformation
can be found by calculating the variation of the average dipole-moment density caused by this
perturbation:


1
Q n (Rn,i + wn,i ) V 1
Q n Rn,i ,
21.
Pi = V n
n

where V and Vn are the sample volume before and after the deformation, respectively. Substituting Equation 19 into 21, one nds the polarization response to be composed of two contributions:
One is conditioned by internal strain, and the other by external strain. The rst describes piezoelectricity (in noncentrosymmetric materials). It also describes the bulk exoelectric effect with a
exoelectric coefcient given by

Q n N n,iklj ,
22.
ikjl = v 1
n

www.annualreviews.org Flexoelectric Effect in Solids

397

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

a
a
x3

x1
c = uc
a = ua

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

w int

Figure 4
When uniformly strained along x1 , the centrosymmetric lattice in panel a will deform as shown in panel b.
All atomic displacements are restricted by symmetry and will follow the elastic medium approximation (blue
lattice). For example, if the homogeneous strain is u = u 11 , then c = uc and a = ua. However, an
inhomogeneous deformation lifts this symmetry restriction, and the atomic displacements will not, in
general, follow the elastic medium approximation (as shown for the black atoms in panel c). The difference
w int between the actual displacement (a) and that expected from the elastic medium approximation (ua)
gives rise to internal strain.

where the summation is taken over the ions in a unit cell of volume v. The interpretation of
the second contribution is not straightforward. It is termination dependent, and according to
Tagantsev (4, 20), it can be interpreted as the surface exoelectric effect. Resta (47) recently questioned the existence of this effect [and claimed that the dynamic contribution to the exoelectric
effect (see Section 2.2) does not exist either]. Its contribution, however, is not expected to be enhanced in high-K materials and is therefore of minor importance for applications; we thus exclude
it from further consideration.
The tensor N n,iklj above, which links the internal strains and strain gradients, can be calculated
from the dynamic matrix of the material (20), which in turn can be obtained from ab initio lattice
dynamics simulations. Using this tensor and the transverse Born ionic charges (obtained, e.g.,
from Berry-phase calculations), one can obtain the tensor by using Equation 22. This approach
was implemented by Maranganti & Sharma (51) to calculate the exoelectric coefcients for a
number of materials.
An alternative way to obtain the tensor consists of direct calculations of the polarization
response in an inhomogeneously deformed crystalline lattice. In view of the periodic boundary
398

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

conditions typically required for rst-principles calculations, consideration of a periodic distribution of the strain gradient (as the source of a static wave of external strains) is a reasonable
option. Then once the transverse Born ionic charges are available, the amplitude of the polarization wave can be found. Hong et al. (52) directly implemented this approach in their calculations
for some ferroelectric ABO3 perovskites. These authors introduced the static strain wave via xing
the positions of the A-site atoms as a sinusoidal function of the distance. Both the direction of the
atomic displacements and the modulation direction were parallel to a cubic crystallographic axis
so that the simulated exoelectric response was related to the 11 component of the exoelectric
tensor. One drawback of this approach is that such a simulation does not provide all the information needed to evaluate this component in view of the depolarizing-eld problem. Specically,
such conditions of the simulation imply conservation of the longitudinal component of electrical
displacement, and the effective coefcient D extracted from such a simulation is related to the
true coefcient via D = /(1 + /0 ). Thus, in the case of high-K materials (like ferroelectric
perovskites) with  0 , the calculated D does not yield , whereas D /0 appears to be close
to the exocoupling coefcient f = / . To calculate the coefcient within this framework,
additional simulations of are required.
Ponomareva et al. (53) circumvented the depolarizing-eld problem in nite-temperature simulations of and f tensors. Here the effect was addressed by employing Monte Carlo simulations
with an ab initiocalculated effective Hamiltonian; the contribution of the depolarizing energy
was deliberately eliminated. A drawback of this work is that, in the ab initio calculations of the
f tensor (also used in the Monte Carlo simulations of ), only the interaction between the local
dipole and strain inside one unit cell was taken into account, and such approximation can readily
entail some 50% inaccuracy.
A disadvantage of the three aforementioned methods is that the purely electronic contribution
to bulk exoelectricity is missing. This is probably a minor drawback when it comes to high-K
materials, in which the ionic contribution is enhanced and dominates the total tensor, but
a complete theory should also incorporate the electronic contributions. Recently, Resta (47)
demonstrated how the classical work by Martin (49) can be extended to the description of
the purely electronic contribution to bulk exoelectricity. Also inspired by Martins work,
Hong & Vanderbilt (48) calculated the electronic contribution from rst principles for a
number of materials. A remarkable feature of this contribution is that, in contrast to the
ionic one, it appears even in the absence of internal strains. These calculations were also
performed under conditions of xed electrical displacement so that the evaluated electronic
exoelectric coefcient is underestimated by a factor of (1 + el /0 ), where el is the electronic
contribution to the permittivity. First-principles calculations of the purely electronic contribution to the total exoelectric response (including possible surface conditioned contributions)
were also performed by Dumitrica et al. (54) and by Kalinin & Meunier (5) for carbon
systems.

5. EXPERIMENTAL STUDIES OF FLEXOELECTRICITY


5.1. Quantifying Flexoelectricity in Bulk
The static exoelectric response is most commonly measured by using some variation of the two
methods sketched in Figure 5. The rst method consists of dynamically bending the material in
a cantilever-beam geometry to generate a transverse strain gradient, as shown in Figure 5a. The
exoelectric polarization can then be measured by recording the displacement current owing
www.annualreviews.org Flexoelectric Effect in Solids

399

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

b
F
A

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

x3

u11

u
P = ~12 11
x3

u
P = ~11 33
x3

Figure 5
Methods most commonly used to quantify the exoelectric response. (a) The cantilever bending method
used to measure the effective transverse exoelectric coefcient 12 . (b) The pyramid-compression method
used to obtain the effective longitudinal exoelectric coefcient 11 . Metallic electrodes are shaded in gray.

between the metallic plates with a lock-in amplier. In this way the coefcient 12 , where
P3 = 12

u 11
,
x3

can be calculated; here the exoelectric tensor is expressed by using the two-sufx notation. The
cantilever bending method has been used extensively to investigate exoelectricity (also referred
to as bending piezoelectricity) in polymers (5557). Ma & Cross (5862) recently used this method
to systematically quantify the exoelectric response in a number of perovskite ceramics. Variations
of the method involving three- or four-point bending geometries have also been successfully
employed (12, 63, 64). Importantly, the cross section of a bent beam is modied due to anticlastic
bending (as sketched in Figure 5a), giving rise to nonzero u22 and u33 components and their
corresponding gradients, related to u11 through the Poisson ratio . Thus, 12 is an effective exoelectric coefcient involving a combination of exoelectric tensor components that depends on
the precise geometry of the system (33, 64, 65). In the case of an isotropic bent beam, for example,
12 = 11 + (1 )12 .
A second method for measuring direct exoelectricity, employed by the Cross group (12),
involves uniaxial compression of a truncated pyramidshaped sample, as illustrated in Figure 5b.
The stress 33 = F/A, generated by the pair of forces F, is different at the top and bottom surfaces
of the truncated pyramid due to their different areas A, setting up a longitudinal strain gradient
33
. Again, 11 is an effective coefcient.
and thus generating a exoelectric polarization P3 = 11 u
x3
If we assume that 33 is the only nonzero stress component (14), for an isotropic material this
coefcient can be related to the exoelectric tensor components through
11 = 11 212 .
In practice, the pyramid-compression approach is complicated by the fact that the strain gradient
is strongly inhomogeneous (66) (concentrated mainly at the sample edges), making it difcult to

extract reliable values of .


400

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Conversely, application of a bias across such a structure gives rise to a nonuniform electric eld
distribution and hence polarization gradients that, in turn, generate strain in the sample through
the converse exoelectric effect, which can be measured by using interferometric techniques (12,
67). Such measurements always include contributions from electrostriction, which usually dominate the signal. However, the eld dependence is different for electrostriction (which is quadratic
in eld) and exoelectricity (which is linear in eld), and therefore the two effects can, in principle,
be separated. Fu and coworkers (67) used this method to measure the converse exoelectric effect
and thus to estimate the exoelectric coefcient 11 for (Ba,Sr)TiO3 . The result was in excellent
agreement with measurements of the direct exoelectric effect. Hana and colleagues (68, 69) used
a similar method to study converse exoelectricity in Pb(Mg1/3 Nb2/3 )O3 -PbTiO3 (PMN-PT).
Extracting the full exoelectric tensor is nevertheless challenging, even for the simplest cubic
dielectrics with only three independent coefcients. Zubko et al. (64, 65) employed a dynamical
mechanical analyzer in the three-point bending conguration to generate exoelectric polarization in nonpiezoelectric SrTiO3 single crystals of different crystallographic orientations. However,
pure bending experiments yield only two independent equations for the three exoelectric tensor
components and thus must be combined with a different method to obtain all three tensor components (65). One such method involves accurate measurements of transverse acoustic phonon
frequencies, which are renormalized by the exoelectric terms and can thus be used to quantify
the exoelectric coefcients (21, 87). This method, however, will generally give the sum of the
static and dynamic responses.
The exoelectric coefcients for a number of perovskite ceramics are listed in Table 1.
Particularly high exoelectric coefcients (10100 C m1 ) have been measured close to the
ferroelectric-to-paraelectric phase transitions of (Ba,Sr)TiO3 , relaxor PMN-PT, and (Pb,Sr)TiO3
ceramics, for which the dielectric constants reach values exceeding 10,00020,000. Measurements
of the exoelectric response as a function of temperature conrm the expected scaling of with
, as illustrated in Figure 6 for several perovskite compounds in their paraelectric phases. The
exact proportionality between and predicted by Equation 9, however, does not always hold,
as can be seen most clearly in Figure 6b. Due to the large differences in dielectric permittivities
of different compounds, it is more instructive to compare the normalized (or exocoupling)

Table 1 Measured experimental coefficients for perovskite ceramics in the paraelectric phasea
Compound
BaTiO3
Ba0.67 Sr0.33 TiO3

Coefficient

Method

Value (C m1 )

/0

f = / (V)

12 (T c + 3.4 K)

CB

50 (62)

10,000

560 (62)

11

PC (0.5 Hz)

150 (12)

20,000

850

11

CFE (400 Hz)

120 (67)
700

12

CB (1 Hz)

100 (60)

16,000

Ba0.65 Sr0.35 TiO3

12

CB

8.5 (144)

4,100

234

PbMg0.33 Nb0.67 O3

12

CB

34 (58)

13,000

2645

PMN-PT

11

PC (410 Hz)

612 (69)

21,000

PC (0 Hz)

2050 (69)

3265
110270

Pb0.3 Sr0.7 TiO3

11

PC (0.5 Hz)

20 (12)

13,500

Pb(Zr,Ti)O3

12

4PB

0.5 (63)

2,200

CB (1 Hz)

1.4 (61)

170
25
72

The values in the last column were calculated from the measured exoelectric coefcient and the experimental dielectric susceptibilities. References in the
fourth and six columns are in parentheses. All values are room temperature values unless stated otherwise. Measurement techniques: CB, cantilever
bending; PC, pyramid compression; CFE, converse exoelectric effect; 4PB, four-point bending.

www.annualreviews.org Flexoelectric Effect in Solids

401

MR43CH16-Tagantsev

24 May 2013

15:27

(Ba,Sr)TiO3 ceramics
0.16

140

11 (C m1)

0.04
0.0004

0.0008

15,000

0.0010

Strain gradient (m1)

80

14,000

20,000

28C
30C
35C
40C
50C
60C
70C

0.08

10,000
60
40

5,000

20
12,000

8,000
10

6,000
4,000

20
0

20

30

40

50

60

70

2,000
20

Temperature (C)

30

40

50

60

70

Temperature (C)

SrTiO3 single crystals


4.0

60

|P3| 1010 (C m2)

50

40

I/z0 (pA m1)

10,000

15

30

3.5

100
101

3.0
4

111

2.5

2
0
0.00 0.02

2.0
0.04

0.06

0.08

0.10

Strain gradient (m1)

0.12

1.5

20

Capacitance (nF)

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

0.12

0.00
0.0000

100

(Pb,Sr)TiO3 ceramics

Relative permittivity

120

23C

Relative permittivity

Polarization
(C m2)

160

11 (C m1)

ARI

1.0
10
0.5
0
250

200

150

100

50

0.0
50

Temperature (C)
Figure 6
Temperature evolution of the effective longitudinal exoelectric coefcient and the dielectric permittivity of (a) (Ba,Sr)TiO3 and
(b) (Pb,Sr)TiO3 ceramics above the Curie temperature. The inset in panel a shows the linear relationship between polarization and
strain gradient. Adapted from Reference 12 with kind permission from Springer Science+Business Media. (c) Temperature evolution of
the ratio of exoelectric current I and bending displacement z0 (with I/z0 12 ) and capacitance of a (100)-oriented crystal of SrTiO3 .
The anomaly at 105 K is related to ferroelastic domain wall motion in SrTiO3 . The inset shows the linear relationship between
polarization and strain gradients for three crystals of different orientation. Reprinted with permission from Reference 64. Copyright
2007, the American Physical Society.

coefcients f = / measured in volts; these coefcients were predicted by Kogan (19) to be of


the order of f q/4 0 a = 110 V (Equation 18) for simple ionic solids.
For all the materials listed in Table 1, the absolute magnitude of the response greatly exceeds
this simple theoretical estimate. BaTiO3 -based ceramics, in particular, show enormous exoelectric and exocoupling coefcients. Such large exoelectric coefcients are particularly puzzling,
as theoretical considerations suggest that exocoupling coefcients in excess of 1015 V should
402

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

make the perovskite structure unstable to the formation of incommensurate phases; this instability
is discussed in more detail in Section 5.2.5. For PMN-PT, the measured coefcients vary by an
order of magnitude, depending on the measurement method used (68, 69).
Experimental data for single crystals are shown in Table 2 together with theoretical coefcients
calculated by using density functional theory (DFT). Although direct comparison with ceramics
is difcult due to the scarcity of single-crystal data and the different measurement techniques
used in most cases, the exocoupling coefcients for single crystals are generally considerably
lower. For SrTiO3 the only monocrystalline material whose full static exoelectric tensor has
been quantiedthe magnitude of the exoelectric response is in good agreement with Kogans
estimate (64, 65) (although there is still some uncertainty about the signs of some tensor components). Neutron and Brillouin scattering measurements on BaTiO3 and KTaO3 single crystals
give values of the same order of magnitude (34, 7072). However, all the values given in Tables 1
and 2 should be treated as order-of-magnitude estimates only. Challenges to properly quantifying
exoelectricity arise from difculties in separating the bulk exoelectric response from the contributions of surface piezoelectricity and dynamic exoelectricity. It has been generally assumed that
in high-K dielectrics, the contributions from surface piezoelectricity should not be signicant, as
they have not been expected to scale with the bulk dielectric constant of the material. This, however, appears not to be the case, as is discussed in Section 3.1, and thus all the experimental bulk
static exoelectric coefcients may well be affected by this contribution. Meanwhile, neutron and
Brillouin scattering measurements of phonon dispersion curves include the dynamic exoelectric
effect. Moreover, they cannot be used to determine the signs of the coefcients (they can be used
to determine only the relative signs of some of the coefcients) (P.V. Yudin & A.K. Tagantsev,
submitted manuscript).
We deliberately exclude polymers and elastomers from Tables 1 and 2, as the nature of the
effect is very different in these materials and is beyond the scope of this review. Nevertheless,
it is worth mentioning that the exibility of these materials makes them in principle attractive
candidates for future exoelectric devices and has motivated many studies on exoelectricity in
polymers. Recently, Baskaran et al. (66, 73, 74) reported extremely large exoelectric coefcients
for polyvinylidene uoride (PVDF) thin lms, but there is still no consensus on the best methods
for quantifying exoelectricity in these materials, leading to very large variations in the values
reported for different polymers by different groups (5557, 75, 76). For PVDF alone, the reported
coefcients range from 13 nC m1 (76) to 80 C m1 and higher (66, 73, 74).
Theoretically quantifying the exoelectric tensor has been equally challenging. Nevertheless,
a number of different approaches have been developed to calculate the exoelectric coefcients,
as reviewed in Section 4. The results of such calculations for a variety of materials are summarized
in Table 2.
For SrTiO3 , the theoretical and experimental values are of the same order of magnitude,
although there is still signicant disagreement (including in the signs) between values obtained
by using different theoretical methods. For other materials, notably BaTiO3 , the disagreement
between theory and experiment is much larger. In all cases, however, the calculations yield f-values
comparable to or smaller than Kogans estimate (Equation 18) and thus cannot account for the
large coefcients measured for ferroelectric ceramics.
In summary, despite signicant experimental and theoretical effort, reliable quantication of
the exoelectric response remains challenging. Although there is some convergence on the order
of magnitude of the response for some monocrystalline oxides, there are still large, unexplained
discrepancies between experimental data on ceramics and experimental data on single crystals
and equally large differences between theoretical values obtained by using different techniques.
The importance of building an accurate database of exoelectric tensor components for a wide
www.annualreviews.org Flexoelectric Effect in Solids

403

404

Zubko

Catalan

Tagantsev

12.5

GaAs

47

300

2.53

11 (r.t.)
5.3 (146)

7(e) (64, 65)


5.8 (21, 65)

12 (r.t.)
|44 | (r.t.)
11 12
|44 |
11 12
44

0.2(e) (64, 65)

11 (r.t.)

/0

Experiment
Value (nC m1 )

Bi12 TiO20

PbTiO3

KTaO3

SrTiO3

Ba0.5 Sr0.5 TiO3

|11 12 |
|44 |

Coefficient

12.5

0: 1.8 (34, 70, 72)


2.9: 2.5 (34, 70, 72)

2.6
2.2
1.2 to 1.4 (70, 145)
1.22; 2.4 (70, 87, 145)

0.08

<7.8 (70, 71)


<0.15 (70, 71)

f = / (V)

1020 (53)
610 (53)

11
12
44

11
12
44

11

11

12
44

11

+0.51
+0.85
0.84
+0.50
+0.26
+0.17

(51)
102(c) (51)
102(d) (51)
102(c) (51)
102(d) (51)
102(c) (51)

102(d)

3.7(d) (51)
3.6(d) (51)

0.26(d) (51)

+0.15(d) (51)
5.5(d) (51)
1.9(d) (51)

12
44

Theory
Value (nC m1 )

11

f (V)

(Continued )

0.046
0.077
0.076
0.045
0.024
0.015

20(b) (48)

17(b) (48)

150(a) (52)
16(b) (48)

5.1 (53)
3.3 (53)
0.045 (53)

40(a) (52)
16(b) (48)

24 May 2013

BaTiO3

Compound

ARI

Single crystals

Table 2 Experimental and theoretical flexoelectric coefficients for single crystalsa

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev
15:27

8.2

4.6

9.7
5.9

11.9

ZnS

KCl

MgO

NaCl

Si

11

11
12
44

+0.41
(51)
0.12 102(c) (51)
0.21 102(c) (51)

12(b) (48)

5(b) (48)
0.079
0.023
0.040

11(b) (48)

0.098
0.030
0.057

+0.40 102(c) (51)


0.12 102(c) (51)
0.23 102(c) (51)

11
12
44
102(c)

0.043
0.021
0.084

0.31 102(c) (51)


1.5 102(c) (51)
0.61 102(c) (51)

11
12
44

24 May 2013

11

0.048
0.022
0.034

+0.47 102(c) (51)


+0.31 102(c) (51)
0.34 102(c) (51)

11
12
44

ARI

a
Underlined values were calculated by using experimental values of . Values marked (a) and (b) were computed under conditions of xed electrical displacement (rather than xed electric eld)
and correspond to D / f. Values marked (b) include only the electronic contribution. The theoretical coefcients from Reference 51 labeled (c) were calculated using a shell model, whereas
those labeled (d) were obtained from ab initio calculations. Coefcients 11 and 12 marked by (e) were calculated assuming that 44 is positive.

11.1

GaP

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev
15:27

www.annualreviews.org Flexoelectric Effect in Solids

405

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

range of materials should not be overlooked, because, as we see next, exoelectric effects are
ubiquitous at the nanoscale and must be properly included for accurate modeling of nanoscale
structures and devices.

5.2. Manifestations of the Flexoelectric Effect in Solids

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Although precise quantication of exoelectricity remains a challenge, its importance in modifying the functional properties of materials is becoming more and more apparent, particularly
at the nanoscale. In this section we examine the effects of exoelectricity on the electrical and
mechanical behavior of materials and discuss the different experimental manifestations of the
exoelectric effect.
5.2.1. Strain gradients in thin films. As we see from Tables 1 and 2, the largest exoelectric coefcients have been measured in ferroelectric materials, as they normally have large dielectric constants. If we also bear in mind that gradients are inversely proportional to size, ferroelectric thin lms are the obvious place to look for large exoelectric effects. Indeed, some
of the most dramatic manifestations of exoelectricity have been reported in ferroelectric thin
lms.
All thin lms are grown on rigid substrates, and stresses thus appear due to the mismatches
between lattice parameters and thermal expansion coefcients of the lm and the substrate. The
substrate-induced strain state in the lms may be homogeneous if they are sufciently thin; however, whenever the thickness exceeds some critical value, the lms may relieve the stress by formation of mist dislocations (77) or by twinning if the lm is ferroelastic. Both these relaxation
mechanisms are highly inhomogeneous and generate exoelectric effects.
The effect of strain gradients on the dielectric constant of thin lms was studied by Catalan
et al. (22, 78). Their assumption, based on earlier work by Kim et al. (79) and consistent with X-ray
diffraction analysis (78), was that the mismatch strain does not relax suddenly at the surface but
rather relaxes exponentially through the lm. When the exoelectricity caused by the exponential
strain relaxation is introduced into the free energy of the system, it causes an orders-of-magnitude
decrease in the dielectric constant and complete smearing of the dielectric peak at the ferroelectric
Curie temperature. These results can be understood by recalling that the effect of the exoelectricity is analogous to that of applying an external electric eld (see Section 2.1), and external elds
decrease the permittivity by saturating the polarization.
Sharma and colleagues (80) also analyzed the adverse impact of exoelectricity on the dielectric
constant of thin lms by using atomistic calculations. These authors showed that the gradient
arising from intrinsic surface tension leads to a exoelectrically induced lowering of the permittivity
that mimics the well-known dead-layer effect (81). Thus, even for perfectly coherent thin lms
with well-matched electrodes, surface exoelectricity still acts to prevent the permittivity of the
lms from being as large as that of bulk samples. Flexoelectricity has also been predicted to increase
the critical thickness (or to decrease the critical temperature) for ferroelectricity in thin lms (82).
Experimentally, the connection between surface gradients and polarization was also pointed out
by Scott (83), who showed that surface stresses induced by polishing can cause the appearance
of polar modes in otherwise centrosymmetric single crystals. Kholkin et al. (84) also showed that
the grain boundaries of polycrystalline SrTiO3 (a centrosymmetric material) have a piezoelectric
response consistent with surface gradientinduced exoelectricity.
5.2.2. Flexoelectric poling. By analogy with electric elds, strain gradients skew the thermodynamic potential, as illustrated in Figure 7, and may lead to preferential poling of the material. For
406

Zubko

Catalan

Tagantsev

ARI

24 May 2013

15:27

Energy

MR43CH16-Tagantsev

u = 0
x

u 0
x

10 mTorr

100 mTorr

350 mTorr

10

P (C cm2 )

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Polarization

10
3 2 1

3 2 1

3 2 1

E (MV cm1)
Figure 7
The effects of strain gradients in ferroelectric materials can mimic those of an electric eld. Just like an
electric eld, a strain gradient leads to the skewing of the ferroelectric double-well potential, favoring one of
the two polarization states. This can lead to a shift in the ferroelectric hysteresis loop along the eld axis. In
Reference 30, progressively larger strain gradients were induced in HoMnO3 thin lms by increasing the
oxygen pressure during growth from 10 to 350 mTorr. Adapted with permission from Reference 30.
Copyright 2011, American Institute of Physics.

example, the Lubomirsky group (36) reported that the poling of quasi-amorphous BaTiO3 upon
cooling was assisted by exoelectricity. In ferroelectrics, strain gradients can lead to asymmetric
polarization-eld hysteresis loops that are typically observed due to the presence of an internal eld
(28, 29). This effect is expected to be particularly appreciable in ferroelectric thin lms in which
the mist strain due to the underlying substrate is relaxed though the formation of dislocations,
giving rise to large strain gradients (85). Experimentally, several authors have discussed the effect
of strain gradients on polarity (27, 30, 31, 86), and Lee et al. (30) in particular directly showed the
correlation between asymmetric hysteresis loops (and domain population) and strain gradients,
measured by using grazing incidence X-ray diffraction, in HoMnO3 thin lms. Different strain
gradients were induced by changing the oxygen pressure during growth. The effective electric
eld due to these strain gradients leads to a reorientation of defect dipoles, which in turn result in
the observed shifts of the hysteresis loops once the samples are cooled to room temperature (see
Figure 7). An extreme manifestation of the effect of exoelectricity on polarity is the discovery
that strain gradients can be used to actively switch the polarization of a ferroelectric (27, 31, 86).
This phenomenon is further discussed in Section 6.3.
www.annualreviews.org Flexoelectric Effect in Solids

407

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Flexoelectricity appears to play an important role in a number of poling phenomena. However,


in many cases (27, 36, 86) the magnitude of the exoelectric coefcients required to explain the
experimental observations purely by exoelectricity exceeds physically reasonable values by several
orders of magnitude. Thus, other contributions may also play a role.

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

5.2.3. Elasticity and lattice dynamics. As we discuss above, both the static and dynamic exoelectric effects lead to a modication of the phonon dispersion curves (34). This effect was experimentally studied in single crystals of SrTiO3 (87) and KTaO3 (72) and was used to estimate
the total (static + dynamic) bulk exocoupling coefcients for these materials (see Table 2). The
inuence of exoelectricity on the propagation of nonlinear strain waves was also studied theoretically by Mirzade (88), although in this case the expansion of the thermodynamic potential was
formulated in terms of the electric eld rather than in terms of the polarization, and only the static
exoelectric contribution was considered.
Flexoelectricity may also affect the elastic properties of materials and thus the measurements of
the elastic constants. In particular, the electrostatic cost of generating exoelectric charge causes
a depth-dependent stiffening in nanoindentation experiments, owing to the very inhomogeneous
nature of the stress introduced by a nanoindenter. The Sharma group (8991) studied this effect
both experimentally and theoretically for several materials. For BaTiO3 the authors used their
nanoindentation results to estimate the transverse exoelectric coefcient and found it to be
approximately 4 C m1 , in good agreement with other experimental results for this material.
5.2.4. Polarization-induced bending. During the late 1960s and early 1970s, Bursian and
coworkers studied the phenomenon of polarization-induced bending, the theoretical aspects of
which are discussed in Section 3.2. Arguing that the development of a polarization should lead to
an inhomogeneous deformation of the perovskite unit cell, Bursian & Zaikovskii (1) demonstrated
the effect by observing the bending of a 2.5-m-thick ferroelectric BaTiO3 plate as its polarization
was switched by an applied electric eld (Figure 8a).
By measuring the curvature induced by the electric eld E, one can also get an estimate of
the exoelectric coefcient, which is given by (33)
12 =

Gd 2

,
E 12(1 2 )

23.

where G is the Youngs modulus, d is the crystal thickness, and is the Poisson ratio. From the data
of Bursian & Zaikovskii (1), reproduced in Figure 8b, we estimate 12 to be of the order of 0.1
1 C m1 in the paraelectric phase of BaTiO3 , close to the phase transition. Taking the dielectric
constant of BaTiO3 to be approximately 10,000 at these temperatures, we obtain a exocoupling
coefcient f 110 V, which is of the order of Kogans (19) estimate.
Later, Bursian (93) proposed another manifestation of the converse exoelectric effect. For
regular 180 ferroelectric domain structures, domains of opposite polarization would distort in
the opposite sense, leading to no net macroscopic bending of the sample, as sketched in Figure 8c.
Again, this effect is expected to be most pronounced in thin ferroelectric layers. Superlattices
composed of alternating ultrathin ferroelectric layers and nonferroelectric layers may thus be
an ideal system to observe this phenomenon. Indeed, such inhomogeneous strain distributions
have been predicted via DFT calculations of PbTiO3 /SrTiO3 superlattices by Aguado-Puente
& Junquera (92). These considerations imply that, in thin ferroelectrics, there may be a strong
elastic component to the formation of 180 domains (93) that is usually assumed to be dictated
only by the electrostatic boundary conditions. This inhomogeneous strain may help explain the
large coherence lengths for the domains observed in PbTiO3 /SrTiO3 superlattices (92, 94).
408

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

(m1)

150

40

80

40
40

80

E (kV cm1)

Curvature, (m1)

80

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

40

80

20
20

80

100

180

Temperature (C)

Figure 8
(a) Polarization-induced bending of a 2.5-m-thick BaTiO3 crystal in the ferroelectric phase at room
temperature (1). The sample curvature exhibits hysteresis due to the hysteretic response of the ferroelectric
polarization. Reprinted from Reference 1. (b) Field-induced curvature as a function of temperature across
the ferroelectric-to-paraelectric phase transition. Reprinted from Reference 1. (c) Local bending of a
polydomain ferroelectric envisaged by Bursian (93). Adapted from Reference 93 with permission from
Taylor & Francis, Ltd. (http://www.tandf.co.uk/journals).

5.2.5. Domain boundary flexoelectricity. Interfaces, including surfaces and domain boundaries, can give rise to a multitude of fascinating physical phenomena that are absent from the
bulk of the host materials (9597). Recent discoveries of novel functionalities within ferroelectric
and ferroelastic domain walls have generated tremendous excitement within the nanoelectronics
community (97, 98). Such domain walls are intrinsically very narrow, and their internal gradients
are therefore intrinsically very large, which, in turn, should lead to large exoelectric and
exomagnetic effects (see sidebar on exomagnetism). For example, even for SrTiO3 , with its
very modest tetragonality (c /a 1.00056 at less than 105 K) (99), the strain gradients within the
ferroelastic domain walls can reach values of the order of 105 106 m1 (100), which translates
into polarization values of 0.11 C cm2 larger than the spontaneous polarization of many
multiferroic ferroelectrics. Detailed phenomenological calculations by Morozovska et al. (23),
which take into account the various couplings between polarization, rotations of the TiO6
octahedra, strain, strain gradients, and depolarizing elds in a fully self-consistent manner, yield
somewhat lower but still sizable estimates. As the temperature is reduced to less than 105 K, the
calculated exoelectric polarization increases to 0.01 C cm2 within 90 twin boundaries and
to 0.1 C cm2 in the antiphase boundaries of SrTiO3 . Moreover, at still lower temperatures
(50 K for antiphase boundaries and 40 K for twin walls), a switchable ferroelectric polarization
of several microcoulombs per square centimeter develops due to the biquadratic coupling
term ijkl Pi P j k l , where  is the order parameter describing oxygen rotations (21, 23). The
www.annualreviews.org Flexoelectric Effect in Solids

409

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

FLEXOMAGNETISM
A linear coupling between strain gradient and magnetization is also possible and is termed exomagnetism. The
exomagnetic effect in solids was recently examined theoretically by Lukashev & Sabirianov (105) and by Eliseev
and coworkers (41, 107). Microscopically, the effect arises from the rearrangements of atomic spins following
atomic displacements due to inhomogeneous deformation (105). In contrast to exoelectricity, which is a general
phenomenon allowed in solids of any symmetry, the exomagnetic effect is restricted to a limited number of magnetic
symmetry classes (107). An indirect exomagnetic effect must also be present in all magnetoelectrics: Because in
these materials polarization and magnetization are coupled, any gradient-induced polarization must induce some
magnetization.

inhomogeneous exoelectric polarization at such domain walls may help explain the apparent
suppression of the exoelectric response observed in SrTiO3 single crystals at less than 105 K and
attributed to polar or charged domain walls (64). Polar domain walls were also recently invoked
(101) to explain the low-temperature elastic anomalies of SrTiO3 (102). V. Janovec (private
communication) recently remarked that all ferroelastic domain walls must, by symmetry, be polar.
The recent discovery of phase coexistence in highly strained BiFeO3 thin lms offers an exciting
opportunity for engineering large strain gradients on the nanometer scale. At the phase boundary
between the tetragonal-like T-phase and the rhombohedral-like R-phase, the out-of-plane lattice

parameter changes from 4.1 A to 4.6 A over a distance of 10 perovskite unit cells (40 A)
7
1
(103), giving rise to a local strain gradient in excess of 10 m (104). Interestingly, in addition
to large exoelectric effects, one may also expect signicant exomagnetic effects (see sidebar on
exomagnetism) in this system, as bulk BiFeO3 is both ferroelectric and antiferromagnetic at room
temperature. Zhang et al. (104) proposed that such exomagnetic effects at phase boundaries may
be responsible for the enhanced magnetism observed in their mixed-phase BiFeO3 lms.
Flexoelectricity can also have dramatic consequences for the conductivity of domain walls
a topic that is currently generating an enormous amount of interest within the ferroelectrics
community (97, 106). Even in the nominally uncharged 180 ferroelectric domain walls, the
exoelectric effect can lead to the development of inhomogeneous polarization perpendicular
to the domain wall. This polarization gives rise to an internal depolarizing eld that leads to a
redistribution of the free carriers within the semiconducting material and either accumulation or
depletion of carriers around domain walls (108, 109).
Within the Landau-Ginzburg-Devonshire formalism, the description of domain walls involves
the inclusion of the gradient term 2g ( Px )2 in the free-energy expansion (Equation 6), where g
determines the domain wall energy. The effect of the exoelectric coupling is to renormalize this
gradient coefcient g g f 2 s , where s is the elastic compliance (21, 22, 41). Hence, if f is
large enough, the domain wall energy can become negative, making the material unstable with
respect to the development of modulated phases. Borisevich et al. (110) argued that this is precisely
what happens at the ferroelectric-antiferroelectric phase boundary in Sm-substituted BiFeO3 thin
lms. In terms of lattice dynamics theory, the instability arises from the interaction (or mode
coupling) between the optical and acoustic phonons, as originally discussed by Axe et al. (34).
For large-enough exoelectric couplings, the optical branch pushes the acoustic branch to zero
frequency at a certain point with q
= 0 in the Brillouin zone, leading to an incommensurate phase.
Just like within ferroelastic domain walls, strain gradients are also very large around dislocations
and are thus expected to give rise to large local exoelectric polarizations (7). Nanoscale gradients
were also proposed to play a role in the appearance of polarization in relaxors (111).
410

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

5.2.6. Electromechanics of ionic conductors. Although most of our discussion focuses on


ferroelectric materials, which tend to display the largest exoelectric coefcients, exoelectricity
is also highly relevant to other systems such as ionic conductors. For example, Morozovska
et al. (112) showed theoretically that exoelectric effects should play an important role in the
electrotechnical properties of lithium batterytype materials. For the simple case of a parallel
plate capacitor lled with such a material, the exoelectric tensor directly links the displacement
of the top electrode to the total electronic charge injected into the system. The exoelectric
contribution to the overall displacement is comparable to that of the so-called deformation
potential and may even exceed it (112).

6. TOWARD APPLICATIONS OF FLEXOELECTRICITY


6.1. Piezoelectric Metamaterials and Nanodevices
The revival of exoelectricity can in part be attributed to the realization of Fousek et al. (113) that
a nanocomposite with built-in shape gradients can produce an effective piezoelectric response
irrespective of whether its constituents are piezoelectric. Flexoelectricity can thus be transformed
from a nuisance into a useful functional property, opening the applications of piezoelectricity to
a whole new class of materials. In 2007, Cross and coworkers led a patent for the simple device
illustrated in Figure 9a (12, 113, 114). This device consists of an array of truncated dielectric
pyramids with a high exoelectric coefcient, like those discussed in Section 5.1, embedded
in another medium (which could simply be air) and sandwiched between two metallic plates.
When the plates are compressed, a stress gradient is generated in each pyramid, inducing a
exoelectric polarization and thus an effective piezoelectric response. Devices with dimensions in
the 10100-m range have been fabricated by using (Ba,Sr)TiO3 as the dielectric, and effective
piezoelectric coefcients up to 40 pC N1 , along with the expected scaling with size, have
been demonstrated (114, 115). A similar type of metamaterial was studied theoretically by
the Sharma group (116). This group considered a nonpiezoelectric elastic matrix with nanoscale
inhomogeneities (nanoinclusions) that lead to local strain gradients. To achieve an effective
macroscopic piezoelectric effect, the shape and distribution of such inhomogeneities must be
noncentrosymmetric to avoid the cancellation of the polarization locally induced by the strain

b
Tungsten
Electrode strip
(Ba,Sr)TiO3
(Ba,Sr)TiO3
Electrode strip
Tungsten

Figure 9
(a) A piezoelectric composite based on the longitudinal exoelectric effect. Adapted from Reference 12 with kind permission from
Springer Science+Business Media. (b) A exure-mode exoelectric piezocomposite. Adapted with permission from Reference 118.
Copyright 2009, American Institute of Physics.
www.annualreviews.org Flexoelectric Effect in Solids

411

ARI

24 May 2013

15:27

gradients. Recently, Chandratre & Sharma (117) proposed that even graphene could in principle
be made piezoelectric by breaking the centrosymmetry of the hexagonal lattice with nanoscale
holes.
A piezoelectric composite based on the transverse, rather than the longitudinal, exoelectric
effect is sketched in Figure 9b (118). The charge generated by bending the exoelectric material
is collected by metallic strips located at positions where the strain gradient is maximum. Effective
piezoelectric coefcients of several thousand picocoulombs per newton were reported. Although
care must be taken when comparing the effective piezoelectric coefcients of such bending-mode
composites with those of standard bulk piezoelectrics measured under compression, these results
are very promising for future applications of exoelectric composites.
The expected enhancement of exoelectricity at the nanoscale has also stimulated theoretical
work on piezoelectric nanodevices. The Sharma group has led the eld with atomistic calculations
on various types of nanostructures, including nanocantilevers, nanocomposites, and superlattices.
The group has shown, for example, that exoelectric effects could enhance the effective piezoelectric coefcient of BaTiO3 nanobeams (119) and as much as double the energy-harvesting potential
of piezoelectric PZT nanocantilevers (120).
The past decade has seen some tremendous advances in the fabrication and characterization
of oxide nanocomposites. The regular arrays of dielectric nanopyramids envisaged by Cross (12)
have been realized by a number of groups (121, 122), as have nanowires (123), tricolor superlattices (124), buckled nanoribbons (125), and other nanoshapes and nanocomposites in which large
exoelectric effects could be observed (126, 127). The effects of exoelectricity, however, have
rarely been considered when such structures have been characterized (127, 128).

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

6.2. Strain Gradient Engineering


The use of epitaxial strain to modify the properties of thin lms is a mature area of research
(129), and strain engineering has some impressive achievements to show, from doubling the
critical temperature of some superconductors (130) to turning SrTiO3 into a room temperature
ferroelectric (131). There has been much less work on strain gradient engineering, but the growing
awareness of exoelectricity is starting to change this situation.
Sharma et al. (132) proposed a layered model for a piezoelectric metamaterial based on nonpiezoelectric components. To achieve macroscopic piezoelectricity, the macroscopic inversion
symmetry of the system must be broken, for example, by adding a third component layer to
form so-called tricolor superlattices (134) or by having a net compositional gradient between the
top and bottom layers of the lm. The piezoelectric performance of the system is conditioned
by the common action of exoelectricity and the inhomogeneity of the polarization and strain
(strain gradient) at the interfaces. This work can be viewed as a fresh look at an older concept of
compositionally graded materials (133).
A different type of gradient can be engineered by growing lms with ferroelastic twins (domains
with alternate directions of spontaneous deformation). Twinning is a well-known strain relaxation
mechanism in thin lms, whereby the orientation and size of the twins can be controlled with the
substrate mismatch and lm thickness (135, 136). Catalan et al. (137) noticed that the nanotwins
that appear in a clamped lm generate strong lateral strain gradients with associated exoelectric
polarizations of the order of C cm2 . Such polarizations are comparable to the size of the spontaneous ferroelectric polarization, and thus the total polarization (exoelectric plus ferroelectric)
can be substantially different, in magnitude and direction, from the pure ferroelectric one (see
Figure 10). The authors (137) propose that the titled polarization of such twinned lms may
give rise to enhanced piezoelectricity, in analogy to what has been shown for morphotropic phase
412

Zubko

Catalan

Tagantsev

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

a
a
c

Substrate

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

y
PbTiO3
x

DyScO3

2 nm

Figure 10
Epitaxial clamping of a ferroelastic lm to a at substrate inhibits the natural distortions (a) of the
ferroelastic domain structure and gives rise to a highly inhomogeneous strain distribution and therefore to
exoelectric polarization. High-resolution transmission electron microscopy allows the positions of atoms to
be determined with subangstrom resolution (b) and thus allows one to map out the strain (color map in panel
c) and polarization distributions (arrows in panel c) within the ferroelastic domains. Adapted with permission
from Reference 137. Copyright 2011, Macmillan.

boundary piezoelectrics (138141). This hypothesis, however, remains to be experimentally


substantiated.

6.3. Mechanical Polarization Switching


Properly harnessed, exoelectricity enables the control of polarity, as its effect is equivalent to
that of an external electric eld, as explained above. In 1969, Bursian and coworkers (86) discussed
the possibility of ferroelectric switching driven by a strain gradient and showed that the bending
of a few-micrometers-thick plate of BaTiO3 can result in the reversal of the sign of its pyroelectric coefcient. The active control of polarity by strain gradients in thin lms was reported by
Gruverman et al. (27), who demonstrated that the polarization state of a ferroelectric Pb(Zr,Ti)O3
www.annualreviews.org Flexoelectric Effect in Solids

413

ARI

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

24 May 2013

15:27

200 nm

Figure 11
Highly concentrated stress elds under an atomic force microscope tip pressing on the sample surface
produce strain gradients that are equivalent to an electric eld sufcient to reverse the polarization of a
ferroelectric thin lm. Nanometric domains can thus be written by simple mechanical pressure, without any
charge injection. From Reference 31. Adapted with permission from AAAS.

capacitor can be reversed by strain gradients generated by bending of the underlying Si substrate.
Recently, Lu et al. (31) used the inhomogeneous deformation caused by pushing with the tip of an
atomic force microscope to switch the polarization of an ultrathin BaTiO3 lm (see Figure 11).
As the authors argue, the conversion of mechanical pressure into readable information is conceptually analogous to the functioning of a nanoscopic typewriter. The exoelectric switching of
polarization has a useful advantage in that it removes the need for applying large electric elds
and thus eliminates associated problems such as leakage and breakdown.

7. UNRESOLVED ISSUES AND FUTURE TRENDS


As Niels Bohr famously said, Prediction is very difcult, especially about the future. Anticipating
the future research in any eld is a tricky and somewhat futile exercise, but there are obvious gaps
that need to be lled, and these give a good indicator of where research may be or should be
concentrating.
At the time of writing, there is not yet a universal consensus regarding the size, or even the
sign, of the exoelectric tensor components for any material. For the best-studied one, SrTiO3 ,
there is some convergence regarding the order of magnitude of the coefcients, with experimental
414

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

values being consistent with Kogans (19) estimate and some theoretical models. The signs of
the coefcients, however, have not been settled, and there is still a signicant variation in the
values predicted by different models. For other systems, such as BaTiO3 , the situation is even
worse: Experimentally, not all the individual tensor components are available, and the calculations and the measurements disagree by several orders of magnitude (52). Solving the reason for
these discrepancies and nding reliable methods for measuring exoelectricity should be priorities
both for experimentalists and for theoreticians. We need to build a robust catalog of exoelectric
coefcients for all materials of technological interest so that exoelectricity can be reliably incorporated into the calculations of performance at the nanoscale. The best exoelectrics may not have
been discovered, and given the universality of the effect, their search should not be restricted to
oxides.
The many experimental observations discussed in this review clearly highlight the important
role of exoelectricity in the electromechanical properties of materials. At the same time, their
interpretation in terms of exoelectricity is often not straightforward. A number of ndings, ranging from measurements of exoelectricity in ceramics to some of the poling effects in ferroelectric
and quasi-amorphous lms, suggest values of bulk exoelectric coefcients that greatly exceed the
criteria for stability of these materials. A clarication of this issue is crucial to both our fundamental
understanding and our effective harvesting of exoelectricity.
There have been predictions that exoelectricity should lead to signicantly enhanced piezoelectric performance at the nanoscale (12, 115, 119, 120, 142). Yet direct evidence for such performance is still scarce; the truncated pyramid composites of Penn State (114) are probably not
small enough, whereas the twinning-induced exoelectric rotation of polarization (137) has not
yet been complemented by a direct measurement of the expected concomitant piezoelectric enhancement due to the difculties posed by the very reduced thickness of the epitaxially clamped
samples. Meanwhile, the theoretical nanocantilevers with enhanced piezoelectric performance
proposed by the Sharma group (119, 120) have not been experimentally realized, although recent
developments in the fabrication of nanotubes, nanorods, and other nanoscale objects could make
such nanocantilevers a reality in the near future. At present, the only devices that appear to be
delivering impressive performances are the exural composites made at Penn State and shown in
Figure 9b (118). We foresee that, within the applied physics and materials engineering community, there will continue to be a strong drive to demonstrate the role of exoelectricity in improved
electromechanical energy harvesting. The recent discovery of exoelectric switching of polarization by nanoindentation (31) will also likely inspire further research on the mechanical control of
polarity.
There are also other, perhaps more peripheral but equally inspiring, subjects for the more
adventurous scientists. Flexomagnetism (see sidebar) has been theoretically proposed, but there
is no conclusive proof of its existence, measurement of its magnitude, or demonstration of its
applications. Large strain gradients inside domain walls or near dislocations should also lead to
enormous exoelectric effects that have not received much attention.
The above are but a few suggestions; however, this review does not aim to be prescriptive.
Rather, we would like to nish by reiterating the reasons why we think exoelectricity is important and likely to grow. It is a universal property of all materials irrespective of symmetry or
composition, including magnetic and biological materials. It is, moreover, an effect that grows as
device size [and dimensionality (143)] diminishes, being large at the nanoscale, and is therefore
important for understanding and manipulating the properties of nanodevices. It affords functionalities, such as control over polarity, that are not generally available with homogeneous strain
alone. And last, but not least, it is, as we hope to have shown here, still a work in progress, and
there are few things more stimulating to a scientist than an unresolved problem.
www.annualreviews.org Flexoelectric Effect in Solids

415

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

DISCLOSURE STATEMENT
The authors are not aware of any afliations, memberships, funding, or nancial holdings that
might be perceived as affecting the objectivity of this review.

ACKNOWLEDGMENTS

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

The authors are indebted to Dragan Damjanovic, John F. Fu, Anna Morozovska, Jim Scott,
Max Stengel, and Jean-Marc Triscone for illuminating discussions and helpful comments on the
manuscript. The authors also gratefully acknowledge funding from the Swiss National Science
Foundation (A.K.T. and P.Z.), from MaNEP (P.Z.), from a European Research Council Starting
Grant (G.C.), and from the Leverhulme Trust (G.C. and P.Z.).

LITERATURE CITED

1. Bursian EV,
Zaikovskii OI. 1968. Changes in the curvature of a ferroelectric lm due to polarization.
Sov. Phys. Solid State 10(5):112124
2. Harris P. 1965. Mechanism for the shock polarization of dielectrics. J. Appl. Phys. 36(3):73941
3. Tagantsev AK. 1987. Pyroelectric, piezoelectric, exoelectric, and thermal polarization effects in ionic
crystals. Sov. Phys. Usp. 30:588603
4. Tagantsev AK. 1991. Electric polarization in crystals and its response to thermal and elastic perturbations.
Phase Transit. 35(34):119203
5. Kalinin SV, Meunier V. 2008. Electronic exoelectricity in low-dimensional systems. Phys. Rev. B
77:033403
6. Meyer RB. 1969. Piezoelectric effects in liquid crystals. Phys. Rev. Lett. 22:91821
7. Indenbom VL, Loginov EB, Osipov MA. 1981. Flexoelectric effect and crystal structure. Kristalograja
26:115762
8. Petrov AG. 2002. Flexoelectricity of model and living membranes. Biochim. Biophys. Acta 1561(1):125
9. Petrov AG. 2006. Electricity and mechanics of biomembrane systems: exoelectricity in living membranes. Anal. Chim. Acta 568(12):7083
10. Breneman KD, Brownell WE, Rabbitt RD. 2009. Hair cell bundles: exoelectric motors of the inner
ear. PLoS ONE 4:e5201
11. Rudquist P, Lagerwall ST. 2012. Applications of exoelectricity. In Flexoelectricity in Liquid Crystals.
Theory, Experiments and Applications, ed. A Buka, N Eber, pp. 21148. London: Imperial College Press
12. Cross L. 2006. Flexoelectric effects: charge separation in insulating solids subjected to elastic strain
gradients. J. Mater. Sci. 41:5363
13. Maranganti R, Sharma ND, Sharma P. 2006. Electromechanical coupling in nonpiezoelectric materials
due to nanoscale nonlocal size effects: Greens function solutions and embedded inclusions. Phys. Rev. B
74:014110
14. Ma W. 2010. Flexoelectric charge separation and size dependent piezoelectricity in dielectric solids.
Phys. Status Solid. B 247(1):21318
15. Nguyen TH, Mao S, Yeh Y-W, Purohit PK, McAlpine MC. 2013. Nanoscale exoelectricity. Adv.
Mater. 25:94674
16. Lee D, Noh TW. 2012. Giant exoelectric effect through interfacial strain relaxation. Philos. Trans. R.
Soc. Lond. Ser. A 370:494457
17. Mashkevich VS, Tolpygo KB. 1957. Electrical, optical and elastic properties of diamond type crystals.
I. Sov. Phys. JETP 5(3):43539
18. Tolpygo KB. 1963. Long wavelength oscillations of diamond-type crystals including long range forces.
Sov. Phys. Solid State 4:1297305
19. Kogan SM. 1964. Piezoelectric effect during inhomogeneous deformation and acoustic scattering of
carriers in crystals. Sov. Phys. Solid State 5(10):206970
416

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

20. Tagantsev AK. 1986. Piezoelectricity and exoelectricity in crystalline dielectrics. Phys. Rev. B 34:5883
89
21. Tagantsev AK, Courtens E, Arzel L. 2001. Prediction of a low-temperature ferroelectric instability in
antiphase domain boundaries of strontium titanate. Phys. Rev. B 64:224107
22. Catalan G, Sinnamon LJ, Gregg JM. 2004. The effect of exoelectricity on the dielectric properties of
inhomogeneously strained ferroelectric thin lms. J. Phys. Condens. Matter 16(13):2253
23. Morozovska AN, Eliseev EA, Glinchuk MD, Chen L-Q, Gopalan V. 2012. Interfacial polarization and
pyroelectricity in antiferrodistortive structures induced by a exoelectric effect and rotostriction. Phys.
Rev. B 85:094107
24. Devonshire AF. 1949. Theory of barium titanate. Philos. Mag. 40(309):104063
25. Mindlin RD. 1968. Polarization gradient in elastic dielectrics. Int. J. Solids Struct. 4(6):63742
26. Ma W. 2008. A study of exoelectric coupling associated internal electric eld and stress in thin lm
ferroelectrics. Phys. Status Solid. B 245(4):76168
27. Gruverman A, Rodriguez BJ, Kingon AI, Nemanich RJ, Tagantsev AK, et al. 2003. Mechanical stress
effect on imprint behavior of integrated ferroelectric capacitors. Appl. Phys. Lett. 83(4):72830
28. Tagantsev AK, Gerra G. 2006. Interface-induced phenomena in polarization response of ferroelectric
thin lms. J. Appl. Phys. 100:051607
29. Cao H-X, Lo VC, Li Z-Y. 2006. Simulation of exoelectricity effect on imprint behaviour of ferroelectric
thin lms. Solid State Commun. 138:4048
30. Lee D, Yoon A, Jang SY, Yoon J-G, Chung J-S, et al. 2011. Giant exoelectric effect in ferroelectric
epitaxial thin lms. Phys. Rev. Lett. 107:057602
31. Lu H, Bark C-W, Esque de los Ojos D, Alcala J, Eom CB, et al. 2012. Mechanical writing of ferroelectric
polarization. Science 336(6077):5961
32. Tagantsev AK. 1985. Theory of exoelectric effect in crystals. Sov. Phys. JETP 61:(6):124654
Trunov NN. 1974. Nonlocal piezoelectric effect. Sov. Phys. Solid State 10:76062
33. Bursian EV,
34. Axe JD, Harada J, Shirane G. 1970. Anomalous acoustic dispersion in centrosymmetric crystals with soft
optic phonons. Phys. Rev. B 1:122734
35. Melnikovsky LA. 2007. Polarization of dielectrics by acceleration. J. Low Temp. Phys. 148:55964
36. Lyahovitskaya V, Feldman Y, Zon I, Wachtel E, Gartsman K, et al. 2005. Formation and thermal stability
of quasi-amorphous thin lms. Phys. Rev. B 71:094205
37. Sirotin YI, Shaskolskaya MP. 1982. Fundamentals of Crystal Physics. Moscow: Mir
38. Le Quang H, He Q-C. 2011. The number and types of all possible rotational symmetries for exoelectric
tensors. Proc. R. Soc. Lond. Ser. A 467:236986
39. Shu L, Wei X, Pang T, Yao X, Wang C. 2011. Symmetry of exoelectric coefcients in crystalline
medium. J. Appl. Phys. 110(10):104106
40. Tagantsev AK, Yurkov AS. 2012. Flexoelectric effect in nite samples. J. Appl. Phys. 112(4):044103
41. Eliseev EA, Morozovska AN, Glinchuk MD, Blinc R. 2009. Spontaneous exoelectric/exomagnetic
effect in nanoferroics. Phys. Rev. B 79:165433
42. Yurkov A. 2011. Elastic boundary conditions in the presence of the exoelectric effect. JETP Lett.
94:45558
43. Fu JY, Cross LE. 2007. On the exoelectric effects in solid dielectrics: theories and applications. Ferroelectrics 354:23845
44. Fu JY, Cross LE. 2007. Separate control of direct and converse piezoelectric effects in exoelectric
piezoelectric composites. Appl. Phys. Lett. 91:162903
45. Klc A, Marvan M. 2004. Theoretical study of the exoelectric effect based on a simple model of ferroelectric material. Integr. Ferroelectr. 63(1):15559
46. Askar A, Lee PCY, Cakmak AS. 1970. Lattice-dynamics approach to the theory of elastic dielectrics with
polarization gradient. Phys. Rev. B 1(8):3525
47. Resta R. 2010. Towards a bulk theory of exoelectricity. Phys. Rev. Lett. 105:127601
48. Hong J, Vanderbilt D. 2011. First-principles theory of frozen-ion exoelectricity. Phys. Rev. B 84:180101
49. Martin RM. 1972. Piezoelectricity. Phys. Rev. B 5(4):1607
50. Born M, Huang K. 1962. Dynamical Theory of Crystal Lattices. Oxford, UK: Oxford Univ. Press
www.annualreviews.org Flexoelectric Effect in Solids

417

ARI

24 May 2013

15:27

51. Maranganti R, Sharma P. 2009. Atomistic determination of exoelectric properties of crystalline dielectrics. Phys. Rev. B 80:054109
52. Hong J, Catalan G, Scott JF, Artacho E. 2010. The exoelectricity of barium and strontium titanates
from rst principles. J. Phys. Condens. Matter 22(11):112201
53. Ponomareva I, Tagantsev AK, Bellaiche L. 2012. Finite-temperature exoelectricity in ferroelectric thin
lms from rst principles. Phys. Rev. B 85:104101
54. Dumitrica T, Landis CM, Yakobson BI. 2002. Curvature-induced polarization in carbon nanoshells.
Chem. Phys. Lett. 360(12):18288
55. Ibe T. 1974. Bending piezoelectricity in polytetrauoroethylene. Jpn. J. Appl. Phys. 13:19798
56. Breger L, Furukawa T, Fukada E. 1976. Bending piezoelectricity in polyvinylidene uoride. Jpn. J. Appl.
Phys. 15:223940
57. Fukada E, Sessler GM, West JE, Berraissoul A, Gunther
P. 1987. Bending piezoelectricity in monomorph

polymer lms. J. Appl. Phys. 62:364346


58. Ma W, Cross LE. 2001. Observation of the exoelectric effect in relaxor Pb(Mg1/3 Nb2/3 )O3 ceramics.
Appl. Phys. Lett. 78(19):292021
59. Ma W, Cross LE. 2001. Large exoelectric polarization in ceramic lead magnesium niobate. Appl. Phys.
Lett. 79(26):442022
60. Ma W, Cross LE. 2002. Flexoelectric polarization of barium strontium titanate in the paraelectric state.
Appl. Phys. Lett. 81(18):344042
61. Ma W, Cross LE. 2005. Flexoelectric effect in ceramic lead zirconate titanate. Appl. Phys. Lett.
86(7):072905
62. Ma W, Cross LE. 2006. Flexoelectricity of barium titanate. Appl. Phys. Lett. 88(23):232902
63. Ma W, Cross LE. 2003. Strain-gradient-induced electric polarization in lead zirconate titanate ceramics.
Appl. Phys. Lett. 82(19):329395
64. Zubko P, Catalan G, Buckley A, Welche PRL, Scott JF. 2007. Strain-gradient-induced polarization in
SrTiO3 single crystals. Phys. Rev. Lett. 99:167601
65. Zubko P, Catalan G, Buckley A, Welche PRL, Scott JF. 2007. Strain-gradient-induced polarization in
SrTiO3 single crystals. Phys. Rev. Lett. 99:167601. Erratum. 2008. Phys. Rev. Lett. 100:199906
66. Baskaran S, He X, Chen Q, Fu JY. 2011. Experimental studies on the direct exoelectric effect in
alpha-phase polyvinylidene uoride lms. Appl. Phys. Lett. 98(24):242901
67. Fu JY, Zhu W, Li N, Cross LE. 2006. Experimental studies of the converse exoelectric effect induced
by inhomogeneous electric eld in a barium strontium titanate composition. J. Appl. Phys. 100(2):024112
68. Hana P, Marvan M, Burianova L, Zhang SJ, Furman E, Shrout TR. 2006. Study of the inverse exoelectric
phenomena in ceramic lead magnesium niobate-lead titanate. Ferroelectrics 336:13744
69. Hana P. 2007. Study of exoelectric phenomenon from direct and from inverse exoelectric behavior of
PMNT ceramic. Ferroelectrics 351:196203
70. Vaks VG. 1973. Introduction to the Microscopic Theory of Ferroelectrics. Moscow: Nauka
71. Harada J, Axe J, Shirane G. 1971. Neutron-scattering study of soft modes in cubic BaTiO3 . Phys. Rev. B
4:15562
72. Farhi E, Tagantsev AK, Currat R, Hehlen B, Courtens E, Boatner LA. 2000. Low energy phonon
spectrum and its parameterization in pure KTaO3 below 80 K. Eur. Phys. J. B 15(4):61523
73. Baskaran S, Ramachandran N, He X, Thiruvannamalai S, Lee HJ, et al. 2011. Giant exoelectricity in
polyvinylidene uoride lms. Phys. Lett. A 375(20):208284
74. Baskaran S, He X, Wang Y, Fu JY. 2012. Strain gradient induced electric polarization in alpha-phase
polyvinylidene uoride lms under bending conditions. J. Appl. Phys. 111(1):014109
75. Marvan M, Havranek A. 1988. Flexoelectric effect in elastomers. Progr. Colloid Polym. Sci. 78:3336
76. Chu B, Salem DR. 2012. Flexoelectricity in several thermoplastic and thermosetting polymers. Appl.
Phys. Lett. 101:103905
77. Matthews JW, Blakeslee AE. 1974. Defects in epitaxial multilayers. I. Mist dislocations. J. Cryst. Growth
27:11825
78. Catalan G, Noheda B, McAneney J, Sinnamon LJ, Gregg JM. 2005. Strain gradients in epitaxial ferroelectrics. Phys. Rev. B 72:020102

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

418

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

79. Kim HJ, Hoon Oh S, Jang HM. 1999. Thermodynamic theory of stress distribution in epitaxial
Pb(Zr, Ti)O3 thin lms. Appl. Phys. Lett. 75(20):319597
80. Majdoub MS, Maranganti R, Sharma P. 2009. Understanding the origins of the intrinsic dead layer effect
in nanocapacitors. Phys. Rev. B 79:115412
81. Sinnamon LJ, Bowman RM, Gregg JM. 2001. Investigation of dead-layer thickness in
SrRuO3 /Ba0.5 Sr0.5 TiO3 /Au thin-lm capacitors. Appl. Phys. Lett. 78(12):172426
82. Zhou H, Hong J, Zhang Y, Li F, Pei Y, Fang D. 2012. Flexoelectricity induced increase of critical
thickness in epitaxial ferroelectric thin lms. J. Phys. Condens. Matter 407(17):337781
83. Scott JF. 1968. Lattice perturbations in CaWO4 and CaMoO4 . J. Chem. Phys. 48:874
84. Kholkin A, Bdikin I, Ostapchuk T, Petzelt J. 2008. Room temperature surface piezoelectricity in SrTiO3
ceramics via piezoresponse force microscopy. Appl. Phys. Lett. 93(22):222905
85. Tagantsev AK, Cross LE, Fousek J. 2010. Domains in Ferroic Crystals and Thin Films. New York: Springer
Oi Z, Makarov KV. 1969. Ferroelectric plate polarization by bending. Izv. Akad. Nauk SSSR
86. Bursian EV,
Ser. Fiz. 33(7):1098102
87. Hehlen B, Arzel L, Tagantsev AK, Courtens E, Inaba Y, et al. 1998. Brillouin-scattering observation of
the TA-TO coupling in SrTiO3 . Phys. Rev. B 57(22):R1398992
88. Mirzade FK. 2007. Inuence of exoelectricity on the propagation of nonlinear strain waves in solids.
Phys. Status Solid. 244(2):52944
89. Gharbi M, Sun ZH, Sharma P, White K. 2009. The origins of electromechanical indentation size effect
in ferroelectrics. Appl. Phys. Lett. 95(14):142901
90. Gharbi M, Sun ZH, Sharma P, White K, El-Borgi S. 2011. Flexoelectric properties of ferroelectrics and
the nanoindentation size-effect. Int. J. Solids Struct. 48(2):24956
91. Robinson CR, White KW, Sharma P. 2012. Elucidating the mechanism for indentation size-effect in
dielectrics. Appl. Phys. Lett. 101(12):122901
92. Aguado-Puente P, Junquera J. 2012. Structural and energetic properties of domains in PbTiO3 /SrTiO3
superlattices from rst principles. Phys. Rev. B 85:184105

93. Bursian EV.


2004. The importance of the unlocal piezoeffect in domain structure formation in ferroelectrics. Ferroelectrics 307(1):17779
94. Zubko P, Jecklin N, Torres-Pardo A, Aguado-Puente P, Gloter A, et al. 2012. Electrostatic coupling and
local structural distortions at interfaces in ferroelectric/paraelectric superlattices. Nano Lett. 12:284651
95. Zubko P, Gariglio S, Gabay M, Ghosez P, Triscone J-M. 2011. Interface physics in complex oxide
heterostructures. Annu. Rev. Condens. Matter Phys. 2:14165
96. Bibes M, Villegas J, Barthelemy A. 2011. Ultrathin oxide lms and interfaces for electronics and spintronics. Adv. Phys. 60:584
97. Catalan G, Seidel J, Ramesh R, Scott JF. 2012. Domain wall nanoelectronics. Rev. Mod. Phys. 84:11956
98. Salje E, Zhang H. 2009. Domain boundary engineering. Phase Transit. 82:45269
99. Lytle FW. 1964. X-ray diffractometry of low-temperature phase transformations in strontium titanate.
J. Appl. Phys. 35(7):221215
100. Cao W, Barsch GR. 1990. Landau-Ginzburg model of interphase boundaries in improper ferroelastic
18 symmetry. Phys. Rev. B 41:433448
perovskites of d 4h
101. Scott JF, Salje EKH, Carpenter MA. 2012. Domain wall damping and elastic softening in SrTiO3 :
evidence for polar twin walls. Phys. Rev. Lett. 109:187601
102. Kityk AV, Schranz W, Sondergeld P, Havlik D, Salje EKH, Scott JF. 2000. Low-frequency superelasticity
and nonlinear elastic behavior of SrTiO3 crystals. Phys. Rev. B 61:94656
103. Zeches RJ, Rossell MD, Zhang JX, Hatt AJ, He Q, et al. 2009. A strain-driven morphotropic phase
boundary in BiFeO3 . Science 326:97780
104. Zhang JX, Zeches RJ, He Q, Chu Y-H, Ramesh R. 2012. Nanoscale phase boundary: a new twist to
novel functionalities. Nanoscale 4:6196204
105. Lukashev P, Sabirianov RF. 2010. Flexomagnetic effect in frustrated triangular magnetic structures. Phys.
Rev. B 82:094417
106. Seidel J, Martin LW, He Q, Zhan Q, Chu YH, et al. 2009. Conduction at domain walls in oxide
multiferroics. Nat. Mater. 8:22934
www.annualreviews.org Flexoelectric Effect in Solids

419

ARI

24 May 2013

15:27

107. Eliseev EA, Glinchuk MD, Khist V, Skorokhod VV, Blinc R, Morozovska AN. 2011. Linear magnetoelectric coupling and ferroelectricity induced by the exomagnetic effect in ferroics. Phys. Rev. B
84:174112
108. Eliseev EA, Morozovska AN, Svechnikov GS, Maksymovych P, Kalinin SV. 2012. Domain wall conduction in multiaxial ferroelectrics: impact of the wall tilt, curvature, exoelectric coupling, electrostriction,
proximity and nite size effects. Phys. Rev. B 85:045312
109. Maksymovych P, Morozovska AN, Yu P, Eliseev EA, Chu Y-H, et al. 2012. Kalinin. Tunable metallic
conductance in ferroelectric nanodomains. Nano Lett. 12(1):20913
110. Borisevich AY, Eliseev EA, Morozovska AN, Cheng C-J, Lin J-Y, et al. 2012. Atomic-scale evolution
of modulated phases at the ferroelectricantiferroelectric morphotropic phase boundary controlled by
exoelectric interaction. Nat. Commun. 3:775
111. Ahn SJ, Kim J-J, Kim J-H, Choo W-K. 2003. Origin of polar domains in ferroelectric relaxors. J. Korean
Phys. Soc. 42:S100911
112. Morozovska AN, Eliseev EA, Tagantsev AK, Bravina SL, Chen L-Q, Kalinin SV. 2011. Thermodynamics
of electromechanically coupled mixed ionic-electronic conductors: deformation potential, Vegard strains,
and exoelectric effect. Phys. Rev. B 83:195313
113. Fousek J, Cross LE, Litvin DB. 1999. Possible piezoelectric composites based on the exoelectric effect.
Mater. Lett. 39(5):28791
114. Fu JY, Zhu W, Li N, Smith NB, Cross LE. 2007. Gradient scaling phenomenon in microsize exoelectric
piezoelectric composites. Appl. Phys. Lett. 91(18):182910
115. Zhu W, Fu JY, Li N, Cross L. 2006. Piezoelectric composite based on the enhanced exoelectric effects.
Appl. Phys. Lett. 89(19):192904
116. Sharma ND, Maranganti R, Sharma P. 2007. On the possibility of piezoelectric nanocomposites without
using piezoelectric materials. J. Mech. Phys. Solids 55(11):232850
117. Chandratre S, Sharma P. 2012. Coaxing graphene to be piezoelectric. Appl. Phys. Lett. 100(2):023114
118. Chu B, Zhu W, Li N, Cross LE. 2009. Flexure mode exoelectric piezoelectric composites. J. Appl. Phys.
106(10):104109
T. 2008. Enhanced size-dependent piezoelectricity and elasticity in
119. Majdoub MS, Sharma P, C
agin
nanostructures due to the exoelectric effect. Phys. Rev. B 77:125424. Erratum. 2008. Phys. Rev. B
79:119904
T. 2008. Dramatic enhancement in energy harvesting for a narrow
120. Majdoub MS, Sharma P, C
agin
range of dimensions in piezoelectric nanostructures. Phys. Rev. B 78:121407. Erratum. 2009. Phys. Rev.
B 79:159901(E)
121. Szafraniak I, Harnagea C, Scholz R, Bhattacharyya S, Hesse D, Alexe M. 2003. Ferroelectric epitaxial
nanocrystals obtained by a self-patterning method. Appl. Phys. Lett. 83(11):221113
122. Nonomura H, Nagata M, Fujisawa H, Shimizu M, Niu H, Honda K. 2005. Structural control of selfassembled PbTiO3 nanoislands fabricated by metalorganic chemical vapor deposition. Appl. Phys. Lett.
86(16):163106
123. Wang ZL, Song J. 2006. Piezoelectric nanogenerators based on zinc oxide nanowire arrays. Science
312:24246
124. Lee HN, Christen HM, Chisholm MF, Rouleau CM, Lowndes DH. 2005. Strong polarization enhancement in asymmetric three-component ferroelectric superlattices. Nature 433:39599
125. Yi Qi, Kim J, Nguyen TD, Lisko B, Purohit PK, McAlpine MC. 2011. Enhanced piezoelectricity and
stretchability in energy harvesting devices fabricated from buckled PZT ribbons. Nano Lett. 11(3):1331
36
126. Wang ZL. 2004. Zinc oxide nanostructures: growth, properties and applications. J. Phys. Condens. Matter
16:R82958
127. Liu C, Hu S, Shen S. 2012. Effect of exoelectricity on electrostatic potential in a bent piezoelectric
nanowire. Smart Mater. Struct. 21:115024
128. Nonnenmann SS, Leaffer OD, Gallo EM, Coster MT, Spanier JE. 2010. Finite curvature-mediated
ferroelectricity. Nano Lett. 10(2):54246
129. Schlom DG, Chen L-Q, Eom C-B, Rabe KM, Streiffer SK, Triscone J-M. 2007. Strain tuning of
ferroelectric thin lms. Annu. Rev. Mater. Res. 37:589626

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

420

Zubko

Catalan

Tagantsev

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

MR43CH16-Tagantsev

ARI

24 May 2013

15:27

130. Locquet J-P, Perret J, Fompeyrine J, Machler E, Seo JW, Van Tendeloo G. 1998. Doubling the critical
temperature of La1.9 Sr0.1 CuO4 using epitaxial strain. Nature 394:45356
131. Haeni JH, Irvin P, Chang W, Uecker R, Reiche P, et al. 2004. Room-temperature ferroelectricity in
strained SrTiO3 . Nature 430:75861
132. Sharma ND, Landis CM, Sharma P. 2010. Piezoelectric thin-lm superlattices without using piezoelectric materials. J. Appl. Phys. 108(2):024304
133. Ban Z-G, Alpay SP, Mantese JV. 2003. Fundamentals of graded ferroic materials and devices. Phys. Rev.
B 67:184104
134. Sai N, Meyer B, Vanderbilt D. 2000. Compositional inversion symmetry breaking in ferroelectric perovskites. Phys. Rev. Lett. 84:563639
135. Speck JS, Pompe W. 1994. Domain congurations due to multiple mist relaxation mechanisms in
epitaxial ferroelectric thin lms. I. Theory. J. Appl. Phys. 76(1):46676
136. Speck JS, Seifert A, Pompe W, Ramesh R. 1994. Domain congurations due to multiple mist relaxation
mechanisms in epitaxial ferroelectric thin lms. II. Experimental verication and implications. J. Appl.
Phys. 76(1):47783
137. Catalan G, Lubk A, Vlooswijk AHG, Snoeck E, Magen C, et al. 2011. Flexoelectric rotation of polarization in ferroelectric thin lms. Nat. Mater. 10:96367
138. Bellaiche L, Garcia A, Vanderbilt D. 2000. Finite-temperature properties of Pb(Zr1x Tix )O3 alloys
from rst principles. Phys. Rev. Lett. 84:542730
139. Fu H, Cohen RE. 2000. Polarization rotation mechanism for ultrahigh electromechanical response in
single-crystal piezoelectrics. Nature 403:28183
140. Guo R, Cross LE, Park S-E, Noheda B, Cox DE, Shirane G. 2000. Origin of the high piezoelectric
response in PbZr1x Tix O3 . Phys. Rev. Lett. 84:542326
141. Speck JS, Pompe W. 1994. Domain congurations due to multiple mist relaxation mechanisms in
epitaxial ferroelectric thin lms. I. Theory. J. Appl. Phys.76:46676
142. Chu B, Zhu W, Li N, Cross LE. 2010. Flexoelectric compositea new prospect for lead-free piezoelectrics. Funct. Mater. Lett.3(1):7981
143. Gregg JM. 2012. Stressing ferroelectrics. Science 335:4142
144. Huang W, Kim K, Zhang S, Yuan F-G, Jiang X. 2011. Scaling effect of exoelectric (Ba,Sr)TiO3
microcantilevers. Phys. Status Solid. Rapid Res. Lett. 5(9):35052
145. Cowley RA. 1964. Lattice dynamics and phase transitions of strontium titanate. Phys. Rev. A 134(4):981
97
146. Shandarov SM, Shmakov SS, Burimov NI, Syuvaeva OS, Kargin YF, Petrov VM. 2012. Detection of
the contribution of the inverse exoelectric effect to the photorefractive response in a bismuth titanium
oxide single crystal. JETP Lett. 95(12):61821

www.annualreviews.org Flexoelectric Effect in Solids

421

MR43-FrontMatter

ARI

1 June 2013

17:7

Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org


Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Contents

Annual Review of
Materials Research
Volume 43, 2013

Computational Materials (Richard LeSar & Simon Phillpot,


Keynote Topic Editors)
Computational Approaches for the Dynamics of Structure Formation
in Self-Assembling Polymeric Materials
Marcus Muller

and Juan J. de Pablo p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1


Density Functional Theory Models for Radiation Damage
S.L. Dudarev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p35
Electronic-Structure Theory of Organic Semiconductors:
Charge-Transport Parameters and Metal/Organic Interfaces
Veaceslav Coropceanu, Hong Li, Paul Winget, Lingyun Zhu, and Jean-Luc Bredas p p p p p63
Phase-Field Model for Microstructure Evolution at the
Mesoscopic Scale
Ingo Steinbach p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p89
Reactive Potentials for Advanced Atomistic Simulations
Tao Liang, Yun Kyung Shin, Yu-Ting Cheng, Dundar E. Yilmaz,
Karthik Guda Vishnu, Osvalds Verners, Chenyu Zou, Simon R. Phillpot,
Susan B. Sinnott, and Adri C.T. van Duin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 109
Simulating Mechanical Behavior of Ceramics Under
Extreme Conditions
I. Szlufarska, K.T. Ramesh, and D.H. Warner p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 131
Uncertainty Quantication in Multiscale Simulation of Materials:
A Prospective
Aleksandr Chernatynskiy, Simon R. Phillpot, and Richard LeSar p p p p p p p p p p p p p p p p p p p p p p p 157
Modern Optical Microscopy Techniques in Materials Research
(Venkatraman Gopalan, Keynote Topic Editor)
Nanoscale Hard X-Ray Microscopy Methods for Materials Studies
Martin Holt, Ross Harder, Robert Winarski, and Volker Rose p p p p p p p p p p p p p p p p p p p p p p p p p p p 183
Nonlinear Optical Microscopy of Single Nanostructures
Libai Huang and Ji-Xin Cheng p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 213
v

MR43-FrontMatter

ARI

1 June 2013

17:7

Real-Time, Subwavelength Terahertz Imaging


F. Blanchard, A. Doi, T. Tanaka, and K. Tanaka p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 237
Superresolution Multidimensional Imaging with Structured
Illumination Microscopy
Aurelie Jost and Rainer Heintzmann p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 261
Vesicle Photonics
A.E. Vasdekis, E.A. Scott, S. Roke, J.A. Hubbell, and D. Psaltis p p p p p p p p p p p p p p p p p p p p p p p p 283
Annu. Rev. Mater. Res. 2013.43:387-421. Downloaded from www.annualreviews.org
Access provided by WIB6283 - Universitaetsbibliothek Dortmund on 05/11/15. For personal use only.

Current Interest
Bionanomaterials and Bioinspired Nanostructures for
Selective Vapor Sensing
Radislav Potyrailo and Rajesh R. Naik p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 307
Electroplating Using Ionic Liquids
Andrew P. Abbott, Gero Frisch, and Karl S. Ryder p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 335
Engineering Crystal Morphology
Preshit Dandekar, Zubin B. Kuvadia, and Michael F. Doherty p p p p p p p p p p p p p p p p p p p p p p p p p p 359
Flexoelectric Effect in Solids
Pavlo Zubko, Gustau Catalan, and Alexander K. Tagantsev p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 387
Mesoscale Domains and Nature of the Relaxor State by Piezoresponse
Force Microscopy
V.V. Shvartsman, B. Dkhil, and A.L. Kholkin p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 423
Nanowire Heterostructures
Jerome K. Hyun, Shixiong Zhang, and Lincoln J. Lauhon p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 451
Phosphors for Solid-State White Lighting
Nathan C. George, Kristin A. Denault, and Ram Seshadri p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 481
Polymer Electrolytes
Daniel T. Hallinan Jr. and Nitash P. Balsara p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 503
Templated Chemically Deposited Semiconductor Optical
Fiber Materials
Justin R. Sparks, Pier J.A. Sazio, Venkatraman Gopalan,
and John V. Badding p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 527
Water VaporMediated Volatilization of High-Temperature Materials
Peter J. Meschter, Elizabeth J. Opila, and Nathan S. Jacobson p p p p p p p p p p p p p p p p p p p p p p p p p p 559
The Yin-Yang of Rigidity Sensing: How Forces and Mechanical
Properties Regulate the Cellular Response to Materials
Ingmar Schoen, Beth L. Pruitt, and Viola Vogel p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 589

vi

Contents

You might also like