Domain Walls From Fundamental Properties To Nanotechnology Concepts Dennis Meier Jan Seidel Marty Gregg Ramamoorthy Ramesh Instant Download
Domain Walls From Fundamental Properties To Nanotechnology Concepts Dennis Meier Jan Seidel Marty Gregg Ramamoorthy Ramesh Instant Download
https://ebookbell.com/product/domain-walls-from-fundamental-
properties-to-nanotechnology-concepts-dennis-meier-jan-seidel-
marty-gregg-ramamoorthy-ramesh-22122314
https://ebookbell.com/product/ferroelectric-domain-walls-statics-
dynamics-and-functionalities-revealed-by-atomic-force-microscopy-1st-
edition-jill-guyonnet-auth-4662864
https://ebookbell.com/product/kinks-and-domain-walls-an-introduction-
to-classical-and-quantum-solitons-1st-edition-tanmay-vachaspati-905548
https://ebookbell.com/product/topological-structures-in-ferroic-
materials-domain-walls-vortices-and-skyrmions-1st-edition-jan-seidel-
eds-5354744
https://ebookbell.com/product/domainspecific-development-with-visual-
studio-dsl-tools-steve-cook-974580
The Ear In The Wall Arthur Benjamin Reeve
https://ebookbell.com/product/the-ear-in-the-wall-arthur-benjamin-
reeve-1650288
https://ebookbell.com/product/bioenergy-1st-edition-demain-arnold-l-
wall-judy-d-harwood-5307800
https://ebookbell.com/product/domainspecific-conceptual-modeling-
concepts-methods-and-adoxx-tools-dimitris-karagiannis-46346498
https://ebookbell.com/product/domaindriven-design-with-golang-use-
golang-to-create-simple-maintainable-systems-to-solve-complex-
business-problems-1st-edition-matthew-boyle-47508642
https://ebookbell.com/product/domainspecific-languages-effective-
modeling-automation-and-reuse-andrzej-wsowski-47661738
D O M A I N WA L L S
Series on Semiconductor Science and Technology
Series Editors
1
3
Great Clarendon Street, Oxford, OX2 6DP,
United Kingdom
Oxford University Press is a department of the University of Oxford.
It furthers the University’s objective of excellence in research, scholarship,
and education by publishing worldwide. Oxford is a registered trade mark of
Oxford University Press in the UK and in certain other countries
© Oxford University Press 2020
The moral rights of the authors have been asserted
First Edition published in 2020
Impression: 1
All rights reserved. No part of this publication may be reproduced, stored in
a retrieval system, or transmitted, in any form or by any means, without the
prior permission in writing of Oxford University Press, or as expressly permitted
by law, by licence or under terms agreed with the appropriate reprographics
rights organization. Enquiries concerning reproduction outside the scope of the
above should be sent to the Rights Department, Oxford University Press, at the
address above
You must not circulate this work in any other form
and you must impose this same condition on any acquirer
Published in the United States of America by Oxford University Press
198 Madison Avenue, New York, NY 10016, United States of America
British Library Cataloguing in Publication Data
Data available
Library of Congress Control Number: 2019957865
ISBN 978–0–19–886249–9
DOI: 10.1093/oso/9780198862499.001.0001
Printed and bound by
CPI Group (UK) Ltd, Croydon, CR0 4YY
Links to third party websites are provided by Oxford in good faith and
for information only. Oxford disclaims any responsibility for the materials
contained in any third party website referenced in this work.
Preface
Technological evolution and revolution are both driven by the discovery of new func-
tionalities, new materials and the design of yet smaller, faster, and more energy-efficient
components. Progress is being made at a breathtaking pace, stimulated by the rapidly
growing demand for more powerful and readily available information technology: high-
speed internet and data-streaming, home automation, tablets and smartphones are now
“necessities” for our everyday lives. Consumer expectations for progressively more data
storage and exchange appear to be insatiable.
Oxide electronics is a promising and relatively new field, that has the potential to
trigger major advances in information technology. Oxide materials offer a multitude of
applications including spintronics, thermoelectrics and power harvesting, which arise
from the broad spectrum of tunable phenomena they exhibit, including magnetism,
multiferroicity, and superconductivity.
Oxide interfaces are particularly intriguing. Here, low local symmetry combined
with an increased susceptibility to external fields leads to unusual physical properties
distinct from those of the homogeneous bulk. In this context, but not limited to oxides,
ferroelectric domain walls have attracted recent attention as a completely new type of
functional interface. In addition to their functional properties, such walls are spatially
mobile and can be created, moved, and erased on demand. This unique degree of
flexibility enables domain walls to take an active role in future devices and hold great
potential as multifunctional 2D systems for nanoelectronics. With domain walls as
reconfigurable electronic 2D components, a new generation of adaptive nanotechnology
and flexible circuitry becomes possible, that can be altered and upgraded throughout
the lifetime of the device. Thus, what started out as fundamental research, at the limit of
accessibility, is finally maturing into a promising concept for next-generation technology.
This book provides a state-of-the-art overview about the significant progress that
has been made in ferroelectric domain wall research over the last decade and evaluates
emerging application possibilities in information technology. Bringing together world-
leading scientists from complementary disciplines, the book gives a broad overview
of how domain walls can be used as functional nano-objects with distinct physical
properties; it also illustrates how domain walls have shifted from being muses for
scientific curiosity into becoming key objects of interest to technology developers.
Different chapters also highlight the close relationship between the progress and the
development of cutting-edge experimental and theoretical analysis tools.
Going beyond the currently available literature, the book identifies major questions
and challenges that will influence research on domain walls, refining the reader’s picture
of the state of the art.
vi Preface
Sharing our excitement about ferroelectric domain walls with you, we and our co-
authors are hoping that you will enjoy reading this comprehensive work and become
curious to find out how far we can go, in the years to come, to establish a new technology
paradigm.
Dennis Meier,
Jan Seidel,
J. Marty Gregg, and
Ramamoorthy Ramesh
1
Physical Properties inside Domain
Walls
Basic Principles and Scanning Probe Measurements
1.1 Introduction
Although domain wall properties are material specific, two features are common to all
of them.
First: by symmetry, a domain wall cannot just stop in the middle of a crystal. Any domain
wall ends at an interface (the surface of the crystal, grain boundary), in another domain
wall (forming a needle domain), or on itself (forming a bubble domain) (Figure 1.1).
Thus, despite being nanoscopically thin, they can be macroscopically long, providing
a continuous path between different interfaces of a crystal irrespective of how big the
crystal is. This “topologically protected” percolation path is most useful for transport
applications (Lee and Salje 2005; Seidel et al. 2009).
Second: domain walls are mobile; they shift their position as domains grow or shrink
in response to external fields. This mobility sets domain walls apart from other types of
interfaces, and is a useful feature that can be exploited in devices where the wall is moved
into and out of a reading unit, as in the “racetrack memory” concept proposed by Stuart
Parkin and co-workers (Parkin et al. 2008), or the domain wall logic devices explored
by the group of Russell Cowburn in Cambridge (Allwood et al. 2005). This mobility
property means that domain walls need not be regarded as just a transport medium
(a connector) for, say, electrical currents, but also as a “container” of information that can
itself be moved into and out of the reading head, carrying with it whatever wall-specific
physical property is of interest, such as, e.g. internal magnetization or polarization.
G. Catalan and N. Domingo, Physical Properties inside Domain: Basic Principles and Scanning Probe Measurements In: Domain Walls: From
Fundamental Properties to Nanotechnology Concepts. Edited by: Dennis Meier, Jan Seidel, Marty Gregg, and Ramamoorthy Ramesh, Oxford
University Press (2020). © G. Catalan and N. Domingo.
DOI: 10.1093/oso/9780198862499.003.0001
2 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
Figure 1.1 Domain wall configurations. (a) Domain walls ending in the middle of a crystal lead to
unresolved configurations. Domains walls ending at (b) surfaces or interfaces, and (c) in another domain
wall forming needle domains or (d) on itself, forming bubble domains.
where k plays the role of the “exchange constant.” This energy contribution is quadratic
as it cannot depend on whether the gradient is positive or negative: the wall energy must
be invariant under space inversion, as it obviously does not depend on whether you cross
the wall from left to right or from right to left.
Minimization of the gradient of the order parameter to minimize the energy favors
broadening of domain walls (broader walls means smaller gradients), but this broadening
comes at the expense of having more material misaligned with respect to the ideal ordered
state, and this also costs energy. Since the Landau free energy, F, of the homogeneous
ferroic state is
a 2 b 4
F= + + O 6 , (1.2)
2 2
where a and b are constants and is the order parameter, the equilibrium thickness
of the domain wall δ can be found by variational minimization of the total energy G,
including gradient and homogeneous terms, integrated across the domain wall thickness:
∞
a 2 b 4 k ∂ 2
G = + + (1.3)
2 2 2 ∂x
−∞
with the center of the wall at x = 0. Here, k is a constant and the boundary conditions
are (∞) = − (−∞) = 0 , with 0 being the homogeneous monodomain state. The
solution for this equation is (Mitsui and Furuichi 1953; Zhirnov 1959)
x
= 0 tanh (1.4)
λ
with
a
0 = − (1.5)
b
From the polarization profile defined by Equation (1.4), the domain wall thickness δ
can in good approximation be defined as twice the correlation length, δ = 2λ. Moreover,
taking into account that the second derivative of the free energy, in this case with respect
to the polarization P, yields the permittivity χ:
∂ 2F 1
χ= =− , (1.7)
∂P 2 2a
4 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
x
(x) = 0 (−δ/2 < x < δ/2) . (1.9)
δ/2
In this approximation, the gradient of the order parameter is equal to the total change of
the order parameter, 20 , divided by the wall thickness, δ, so the energy density U (per
unit volume) associated with the gradient is
2
Ugradient = 1 2 k 20 δ . (1.10)
The other term of the energy density is the standard electrostatic (or magnetostatic, or
elastic) energy, which is proportional to the square of the order parameter:
The total energy density per unit area of the wall (σ ) is obtained by integrating the volume
energy density across the domain wall thickness:
2
δ/2
1 20 1 −1 0 2 1 −1 2
σ= 2 k δ + 2 χ (x) 2
dx = 2k + χ 0 δ. (1.12)
−δ/2 δ 6
∂σ P0 2 1
= 0 = −2k 2 + χ −1 P0 2 , (1.13)
∂δ δ 6
we obtain an expression that is surprisingly close to Equation (1.8) despite the simplifi-
cations:
√
δ = 2 3 kχ. (1.14)
i) The domain wall thickness is inversely proportional to the anisotropy of the order
parameter (the quadratic term that leads the free energy term in the Landau
expansion). This term is generally smaller for magnets than for ferroelectrics and
ferroelastics, for which structural coupling is intrinsically strong. In consequence,
ferroelectric and ferroelastic domain walls are thinner than magnetic walls. The
orders of magnitude for the most common ferroic domain walls from thinnest
to thickest would be ferroelectric domain walls, with δ ∼ 1 nm, followed by
ferroelastic domain walls with δ ∈ (1 − 10) nm, and finally magnetic domain walls
with typical thicknesses in the range δ ∈ (10 − 100) nm.
ii) The domain wall thickness is directly proportional to the susceptibility of the
order parameter (Equation (1.8)). As the phase transition is approached, the
susceptibility will tend to diverge, and hence so will the domain wall thickness.
This is intuitive: near the phase transition, the anisotropy energy (conceived as
the depth of the double well responsible for the ferroic state) is reduced, and
so is the penalty for departing from the polarized state, resulting in broadened
domain walls. Another way to look at this is to think of the domain wall as a
layer of paraphase sandwiched between ferroic domains; as the actual paraphase
is approached (as the temperature increases), the domain walls increase their
thickness (Chrosh and Salje 1999) and eventually, at the transition temperature
TC , occupy the entire material.
0.4 0.4
(a) (b)
Pt(C m–2)
0.0 0.0
–0.2 –0.2
–0.4 –0.4
–0.4
–0.2
0.0
0.2
0.4
–0.4
–0.2
0.0
0.2
0.4
Pr(C m–2) Pr(C m–2)
Figure 1.2 Domain wall trajectories (in phase space) for orthorhombic 120◦ walls (left) and
rhombohedral 180◦ walls in BaTiO3 , superimposed with the corresponding free energy surfaces. The
vertical axis is the polarization along the tetragonal <001> direction, and the horizontal axis is the
polarization along the rhombohedral <111> direction. Note that the domain wall trajectory is not
straight, and as the polarization switches it takes a detour through a local minimum in which the polar
symmetry of the wall is different from that of the domains.
Source: Figure from Marton et al. 2011, copyright by the American Physics Society.
where P0 is the spontaneous polarization, and Q12 and Q11 are, respectively, the trans-
verse and longitudinal electrostrictive coefficients. Notice that, in general, the transverse
coefficient is negative Q12 < 0, and the longitudinal one is positive Q11 > 0. Since the
wall is coherently clamped between rigid domains, the strains sz and sy must remain
unchanged, while the component perpendicular to the domain wall plane, sx , is allowed
to relax. Given that the electrostrictive contraction is suppressed at the wall (that is,
Q12 → 0 at the domain wall), sx will relax to a bigger lattice parameter, and thus the
lattice will expand perpendicular to the wall, leading to a cubic-like like structure such as
the one of the corresponding paraphrase for this case. This can have consequences for
Physical Properties inside Domain Walls 7
Domain Wall
Figure 1.3 In perovskite oxides, octahedral tilts appear in order to fit the oxygen octahedral inside the
perovskite pseudocubic unit cell. If the unit cell is expanded—as must happen inside a ferroelectric
domain wall—then the tilts will be reduced.
Source: Figure from Catalan 2012b, copyright by Taylor and Francis, 2012.
the mechanical and thermal properties; longer bonds generally mean softer bonds and
slower phonons, so one might expect the domain walls to be mechanically softer and
to have different thermal conductivity compared to the bulk. The internal strain of the
wall can also affect other important structural parameters, such as perovskite octahedral
rotations, as illustrated in Figure 1.3.
α 2 β 4 a b
G = P + P + K(P)2 + M 2 + M 4 + A(M)2 + γ P 2 M 2 (1.16)
2 4 2 4
Here it is important to notice two things. First, the last term is always positive: the order
parameters are both squared and thus positive, and the coupling term γ must be positive
if the material has a paraphase. Second, because this term is positive, it increases the total
energy of the system. Accordingly, in order to minimize the energy, if one of the two order
parameters is nonzero, the other one should be zero; if both of them were nonzero at the
same time, then the energy would be higher.
Let us imagine that the system is already in the ferroelectric state (P = 0). We can now
group together the terms multiplying M 2 , so that the coefficient of the magnetic order
parameter is rewritten as
8 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
a
G ∼ + γ P2 M2. (1.17)
2
Typically, a depends on temperature as a0 (T -T C ) (Curie Weiss law), where a0 is
a constant. Here, T C is the Curie temperature, where the system would become
magnetically ordered. Notice that there can be a range of temperatures T < TC where
a/2 is negative (which would favor magnetic order), but γ P 2 is positive and its absolute
value is greater than a/2. This condition is fulfilled for T ∗ < T < TC , where
2γ P 2
T ∗ ≡ Tc − (1.18)
a0
is the renormalized (ferroelectrically modulated) magnetic ordering temperature. From
this treatment, we can make two important observations:
1. When T ∗ < T < TC , the material will be magnetically disordered in the ferroelec-
tric domains, but magnetically ordered inside the domain walls, because the γ P 2
term is suppressed. The resulting polarization and magnetization profiles will look
as in Figure 1.4. Needless to say, the energy contribution from order parameter
coupling will affect the thickness of the domain walls (Goltsev et al. 2003) and, by
extension, the domain size scaling properties (Catalan et al. 2008).
2. It follows from the above treatment that there can be phase transitions inside
domain walls at a different critical temperature (T C ) from that of the bulk (T ∗ ).
The concept of phase transitions inside domain walls was first put forward for
(a) (b)
1.0
0.8
LMP(x)
[a.u.] 0.6
0.4
1
0.2
0.8
0.6 0.0
0.4 –5 –4 –3 –2 –1 0 1 2 3 4 5
–0.2
x
0.2
–0.4
0.0 0.4
0.2 –0.6
P
–1 0.0
–0.5 u.]
0.0 –0.2 ) [a. –0.8 M
P(x 0.5 –0.4 (x
) [a.u 1 M
.] –1.0
Figure 1.4 Calculated polarization and magnetization profiles across a ferroelectric domain wall in a
magnetoelectric with biquadratic coupling between P and M. Adapted from Daraktchiev et al. 2010.
Copyright by the American Physics Society.
Physical Properties inside Domain Walls 9
The key physical insight is that when a component of the primary order parameter
goes through zero (as happens inside a domain wall), it permits the emergence of
the secondary order parameter that was otherwise frustrated by the primary one
(Houchmandzadeh et al. 1991; Tagantsev et al. 2001; Daraktchiev et al. 2010).
3 μm
10 nm
Figure 1.5 Correlation of DW magnetism over a vortex network in multiferroic domain walls of
hexagonal ErMnO3.. (a) AFM topography image, (b) is the room temperature PFM image showing up
and down domains, and (c) shows the MFM image taken at a low temperature and under a magnetic
field of 0.1 T, scanning at a constant height of 40 nm above the surface. The sketch in (d) corresponds to
the domain walls’ net magnetic moments over the entire vortex network.
Source: Adapted from Geng et al. 2012. Copyright by the American Chemical Society.
force microscopy (MFM), for the study of local magnetization in the domain walls;
electric force microscopy (EFM) and Kelvin probe force microscopy (KPFM) for
the study of electrostatic fields and surface potentials; conductive–AFM (C-AFM)
for charge transport measurements; and contact resonance frequency (CRF) or force
spectroscopy techniques for mechanical characterization. In addition, the analysis of any
of the exposed physical magnitudes by AFM can be done as a function of in situ changes
in temperature, external electric magnetic fields, light, or environmental control (see, e.g.
Chapters 11 and 13 for examples).
1.4.2 Polarization
The internal symmetry of domain walls is different from that of the domains they
separate and therefore, by von Neumann’s principle, the symmetry-allowed functional
properties inside domain walls can also be different from those in the domains. The
most obvious examples pertain to polarity. Domain walls separating opposite polarities
Physical Properties inside Domain Walls 11
(i.e. 180◦ walls in ferroelectrics or ferromagnets) must go to zero polarization (along the
easy axis) in their center –although the manner in which the polar inversion is achieved
can be very different, ranging from rigid rotations (Néel and Bloch walls) to change in
magnitude without rotation (see Figure 1.8).1
Conversely, in antipolar materials (antiferroelectrics, antiferromagnetics), any an-
tiphase boundary must, by definition, contain a parallel pair of dipoles, and thus the
domain walls are polar (Li 1956). These two extreme opposites of no polarization inside
a ferroelectric wall and polarization inside an antiferroelectric wall are illustrated in
Figure 1.6.
The case of ferroelastics is more subtle but still amenable to symmetry-based analyses.
In particular, any ferroelastic domain wall must break the mirror symmetry along the
spontaneous strain directions it separates. Consequently, all ferroelastic domain walls
must be piezoelectric, even if the ferroelastic material is macroscopically centrosymmet-
ric ( Janovec et al. 1999). The strong strain gradients associated with ferroelastic domain
walls can also lead to the emergence of ferroelectric polarization at the wall, as observed
for the nonpolar SrTiO3 (Zubko et al. 2007) and CaTiO3 (Gonçalves-Ferreira et al.
2008; Van Aert et al. 2012). One consequence of domain wall piezoelectricity is that a
high concentration of domain walls in a sample can result in a macroscopically enhanced
piezoelectric coefficient (Wada et al. 2006; Hlinka et al. 2009).
Piezoelectricity at the domain walls can be detected by PFM, exploiting the converse
piezoelectric effect: the AFM tip used as a top mobile electrode in contact with the
samples applies an AC electric field and simultaneously senses the sample mechanical
expansion and contraction. The disappearance of polarization inside the 180◦ domain
walls of ferroelectrics has other physical consequences. The first and most obvious one
is that, since there is no polarity inside the wall, there cannot be piezoelectricity either.
Consequently, the amplitude of the signal in PFM images will be zero at the domain walls
where the polarity switches sign, i.e. at the 180◦ domain walls (see Figure 1.7).
(a) DW DW (b) DW DW
Figure 1.6 Sketches of domain walls in (a) polar and (b) antipolar materials. In the first case, the
polarization disappears inside the walls of polar materials; conversely, polarization appears inside the
walls of antipolar materials. In both cases, the polar properties of the walls are exactly the opposite to
those of the domains.
1 Even though the polarization projection along the easy axis must be zero at the center of the wall, it can have
components along the other axes; that would be the case for Néel walls and Bloch walls as shown in Figure 1.8.
12 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
nm
60
200 200 100
40
50
100 20 100
0
0
0 0
0 100 200 300 400 500 0 100 200 300 400 500
nm nm
Figure 1.7 Amplitude and phase PFM images of 180◦ domain walls. The phase of the induced
oscillation is correlated with the sign of the ferroelectric polarization. The lack of amplitude signal at the
domain wall separating two opposite domains indicates the lack of piezoelectricity as the center of the
wall is nonpolar, as illustrated in Figure 1.6a.
Figure 1.8 Different types of domain walls: (a) Ising type, (b) Bloch type, and (c) Néel type.
(a) 10 pm (b)
180 deg
5 pm
0 deg
0 pm
(c) (d) (e) (f) 1 pA
0 pA
Time
conductive than the bulk, whereas tail-to-tail walls will be less conductive, the converse
being true for p-type semiconductors. This phenomenon was theoretically predicted by
Eliseev et al. (2011) and experimentally confirmed by Meier et al. (2012).
Domain wall conductivity can also be modified by chemical doping (when the doping
goes preferentially into the domain wall, this is referred to as “decoration”). This was
famously demonstrated by Aird and Salje’s detection of superconductivity along domain
walls in sodium-doped WO3 (Aird and Salje 1998). As well as being dopants, ions can
be charge carriers themselves. A signature of ionic conductivity is its slow dynamics, as
illustrated in Figure 1.9. It is also worth mentioning that domain wall decoration does not
14 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
1.4.4. Magnetotransport
Any magnetic semiconductor (and, by extension, most magnetoelectric multiferroics,
given that they tend to be wide-bandgap semiconductors) is likely to display some
form of magnetoresistance. This is the case with the domain walls of multiferroic
BiFeO3 (He et al. 2012; Domingo et al. 2017). It is important to note, however, that
while all magnetic materials can be magnetoresistive, magnetism is not a prerequisite
Physical Properties inside Domain Walls 15
500 nm
(f)
(b) (d)
500
400
300
nm
200
100
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
μm
140
ΔX (nm)
70
0 (g)
50
0.75
twin domain
PhaseMFM (deg)
Current (nA)
0
0.50
0.2
–50 0
twin wall 0.25
–0.2
–0.3 0 0.3
–100 0.00
–1 –0.5 0 0.5 1
0.00 0.25 0.50 0.75 1.00 1.25
Voltage (V)
μm
Figure 1.10 Correlation between strain and enhanced magnetic and conductive properties at domain
walls of LaSrMnO3 domain walls. (a) Topography AFM image of the film and (b) C-AFM image
showing an enhanced conductivity at the domain walls, observed as thin black lines, together with the
corresponding I(V) curves at the domains and domain walls. The ferroelastic twin walls can be observed
as patterns in OC-SEM images in (c). Finally, the magnetic properties of the strained twin walls are
shown in images d–g; (d) shows an MFM image together with the associated profile section
demonstrating the correlation of the magnetic pattern with that of the twin walls. The right column
shows the MFM images at different distances from the surface: (e) d = 35 nm, (f) d = 20 nm, and
(g) d = 10 nm. As the magnetic tip approaches the surface, it interacts more strongly with the magnetic
moment of the domain walls until it is finally able to switch it, enhancing the magnetic contrast.
16 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
for magnetoresistance, because the Hall Effect can modify the current paths in any
thin disordered semiconductor (Parish and Littlewood 2003). Since “thin disordered
semiconductor” is a good description of any ferroelectric domain wall, it follows that
it may in principle be possible to measure magnetoresistance in the domain walls of
any ferroelectric even if it is not multiferroic, a hypothesis that remains to be tested.
A distinguishing feature of this “geometric magnetoresistance” is that it is positive
(resistance increases with increasing magnetic field). This is the opposite of what is found
in magnetic conductors, for which field-induced spin alignment reduces the magnetic
scattering of the charge carriers. Another feature is that the effect should be maximum
when the magnetic field is perpendicular to the transport plane, i.e. perpendicular to
the domain wall. Both of these features have been detected in in-situ magnetotransport
measurements of single domain walls of BiFeO3 (Domingo et al. 2017), suggesting a
geometric origin for their room-temperature magnetoresistance (see Figure 1.11).
HDW1
I V
HDW2 A
125° 0 pA
(d) (e)
17,5 17,5
DW1 Parallel DW DW2 Perpendicular DW
15,0 15,0
H = 0T H = 0T
12,5 H = 1T 12,5 H = 1T
I (pA)
V
V
7,5 A 7,5
A
5,0 5,0
2,5 2,5
0,0 0,0
1,0 1,2 1,4 1,6 1,8 2,0 1,0 1,2 1,4 1,6 1,8 2,0
V (V) V (V)
Figure 1.11 Magnetotransport measurements on BiFeO3 domain walls. (a) Scheme of the combined
C-AFM setup with an external magentic field. (b) The amplitude PFM image shows two families of
a1/a2 domains with perpendicular relative orientation with respect to the external applied magnetic
field. (c) All the domain walls are observed to be conductive, but their level of conductivity changes when
a magnetic field is applied either parallel as in (d) or perpendicular to the longitudinal domain walls as
shown in (e).
Source: Adapted from Domingo et al. 2017.
Physical Properties inside Domain Walls 17
200 Hz
0 Hz
2.5 μm
–200 Hz
Figure 1.12 Contact resonance frequency image of a pattern of stripe domains on a periodically poled
LiNbO3 single crystal. The domain walls stand out as dark lines as depicted by a decrease in the contact
resonance frequency of the cantilever when the tip is scanning over the wall. This feature can be
correlated with a decrease in the domain wall stiffness with respect to that of the adjacent domains.
18 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
The reason for the mechanical contrast of the walls is, once again, the coupling
between order parameters, in this case, strain and polarization, which are necessarily
coupled in any piezoelectric and thus in any ferroelectric. Inducing a deformation in a
ferroelectric changes its polarization (ferroelectrics are piezoelectric), and this change
in polarization has an electrostatic energy cost. The domain wall, itself, however, is not
polar, and therefore it is not piezoelectric, meaning that there is no electrostatic cost for
its deformation. Stating that the deformation of the domain wall has a reduced energy
cost is just a physicist’s way of saying that deforming a ferroelectric wall is easy. In other
words, as observed, the wall will respond more softly to the atomic force microscope
indentation than the domains.
..........................................................................................
REFERENCES
Aird A., Salje E. K. H., “Sheet superconductivity in twin walls: experimental evidence of WO3-x ,”
J. of Phys. Cond. Mat. 10, L377 (1998).
Allwood D. A., Xiong G., Faulkner C. C., Atkinson D., Petit D., Cowburn R. P., “Magnetic domain-
wall logic,” Science 309, 1688 (2005).
Balcells L. L., Paradinas M., Bagués N., Domingo N., Moreno R., Galceran R., Walls M., Santiso
J., Konstantinovic Z., Pomar A., Casanove M. -J., Ocal C., Martínez B., Sandiumenge F.,
“Enhanced conduction and ferromagnetic order at (100)-type twin walls in La0.7 Sr0.3 MnO3
thin films,” Phys. Rev. B 92, 075111 (2015).
Bartels M., Hagen V., Burianek M., Getzlaff M., Bismayer U., Wiesendanger R., “Impurity-
induced resistivity of ferroelastic domain walls in doped lead phosphate,” J. Phys.: Condens.
Mat. 15, 957–962 (2003).
Bursill L. A., Peng J. L., “Electron microscopic studies of ferroelectric crystals,” Ferroelectrics 70,
191 (1986).
Bursill L. A., Peng J. L., Feng D., “HREM study of (100) ferroelectric domain-walls in potassium
niobate,” Philos. Mag. A 48, 953 (1983).
Catalan G., “On the link between octahedral rotations and conductivity in the domain walls of
BiFeO3 ,” Ferroelectrics 433, 65–73 (2012b).
Catalan G., Béa H., Fusil S., Bibes M., Paruch P., Barthélémy A., Scott J. F., “Fractal dimension
and size scaling of domains in thin films of multiferroic BiFeO3 ,” Phys. Rev. Lett. 100, 027602
(2008).
Catalan G, Seidel J., Ramesh R., Scott J. F., “Domain wall nanoelectronics,” Rev. Mod. Phys. 84,
119–156 (2012a).
Choi T., Horibe Y., Yi H. T., Choi Y. J., Wu W., Cheong S.-W., “Insulating interlocked ferroelectric
and structural antiphase domain walls in multiferroic YMnO3 ,” Nat. Mat. 9, 253 (2010).
Chrosh J., Salje E. K. H., “Temperature dependence of the domain wall width in LaAlO3 ,” J. Appl.
Phys. 85, 722 (1999).
Cordero-Edwards K., Domingo N., Abdollahi A., Sort J., Catalan G., “Ferroelectrics as Smart
Mechanical Materials,” Adv. Mater. 29, 1702210 (2017).
Daraktchiev M., Catalan G., Scott J. F., “Landau theory of ferroelectric domain walls in magneto-
electrics,” Ferroelectrics 375, 122 (2008).
20 References
Daraktchiev M., Catalan G., Scott J. F., “Landau theory of domain wall magnetoelectricity,” Phys.
Rev. B 81, 224118 (2010).
Diéguez O., Aguado-Puente P., Junquera J., Íñiguez J., “Domain walls in a perovskite oxide with
two primary structural order parameters: first-principles study of BiFeO3 ,” Phys. Rev. B 87,
024102 (2013).
Domingo N., Farokhipoor S., Santiso J., Noheda B., Catalan G., “Domain wall magnetoresistance
in BiFeO3 thin films measured by scanning probe microscopy,” J. Phys.: Condens. Mat. 29,
334003 (2017).
Eliseev E. A., Morozovska A. N., Svechnikov G. S., Venkatraman Gopalan, Shur V. Ya., “Static
conductivity of charged domain walls in uniaxial ferroelectric semiconductors,” Phys. Rev. B
83, 235313 (2011).
Farokhipoor S., Noheda B., “Conduction through 71◦ domain walls in BiFeO3 thin films,” Phys.
Rev. Lett. 107, 127601 (2011).
Floquet N., Valot C. M., Mesnier M. T., Niepce J. C., Normand L., Thorel A., Kilaas R.,
“Ferroelectric domain walls in BaTiO3 : fingerprints in XRPD diagrams and quantitative
HRTEM image analysis,” J. Physique III 7, 1105 (1997).
Foeth M., Sfera A., Stadelmann P., Buffat P. -A., “A comparison of HREM and weak beam trans-
mission electron microscopy for the quantitative measurement of the thickness of ferroelectric
domain walls,” J. Electron Microsc. 48(6), 717–723 (1999).
Fousek J., “Permissible domain walls in ferroelectric species,” Czech. J. Phys. 9, 955 (1971).
Fousek J., Janovec V., “The orientation of domain walls in twinned ferroelectric crystals,” J. Appl.
Phys. 40, 135 (1969).
Geng Y., Lee N., Choi Y. J., Cheong S.-W., Weida Wu, “Collective magnetism at multiferroic vortex
domain walls,” Nano Lett. 12, 6055 (2012).
Goltsev A. V., Pisarev R. V., Lottermoser Th., Fiebig M., “Structure and interaction of antiferro-
magnetic domain walls in hexagonal YMnO3 ,” Phys. Rev. Lett. 90, 177204 (2003).
Gonçalves-Ferreira L., Redfern S. A. T., Artacho E., Salje E. K. H., “Ferrielectric twin walls in
CaTiO3 ,” Phys. Rev. Lett. 101, 097602 (2008).
Gruverman A., Wu D., Scott J. F., “Piezoresponse force microscopy studies of switching behavior
of ferroelectric capacitors on a 100-ns time scale,” Phys. Rev. Lett. 100, 097601 (2008).
Gureev M. Y., Tagantsev A. K., Setter N., “Head-to-head and tail-to-tail 180◦ domain walls in an
isolated ferroelectric,” Phys. Rev. B 83, 184104 (2011).
He L., Vanderbilt D., “First-principles study of oxygen-vacancy pinning of domain walls in
PbTiO3 ,” Phys. Rev. B 68, 134103 (2003).
He Q., Yeh C.-H., Yang J.-C., Singh-Bhalla G., Liang C.-W., Chiu P.-W., Catalan G.,
Martin L. W., Chu Y.-H., Scott J. F., Ramesh R., “Magnetotransport at domain walls in BiFeO3 ,”
Phys. Rev. Lett. 108, 067203 (2012).
Hlinka J., Ondrejkovic P., Marton P., “The piezoelectric response of nanotwinned BaTiO3 ,”
Nanotechnology 20, 105709 (2009).
Houchmandzadeh B., Lajzerowicz J., Salje E. K. H., “Order parameter coupling and chirality of
domain walls,” J. Phys.: Condens. Mat. 3, 5163 (1991).
Janovec V., Richterová L., Privratska J., “Polar properties of compatible ferroelastic domain walls,”
Ferroelectrics 222, 331 (1999).
Jia C. L., Mi S. B., Urban K., Vrejoiu I., Alexe M., Hesse D., “Atomic-scale study of electric dipoles
near charged and uncharged domain walls in ferroelectric films,” Nat. Mater. 7, 57 (2008).
Lajzerowicz J., Niez J. J., “Phase transition in a domain wall,” J. Phys. Lett. Paris 40, L165 (1979).
Physical Properties inside Domain Walls 21
Lee D., Behera R. K., Wu P., Xu H., Li Y. L., Sinnott S. B., Phillpot S. R., Chen L. Q.,
Gopalan V., “Mixed Bloch-Néel-Ising character of 180◦ ferroelectric domain walls,” Phys. Rev.
B 80, 060102(R) (2009).
Lee W. T., Salje E. K. H., “Chemical turnstile,” Appl. Phys. Lett. 87, 143110 (2005).
Logginov A. S., Meshkov G. A., Nikolaev A. V., Nikolaeva E. P., Pyatakov A. P., Zvezdin A. K.,
“Room temperature magnetoelectric control of micromagnetic structure in iron garnet films,”
Appl. Phys. Lett. 93, 182510 (2008).
Lu H., Bark C.-W., Esque de los Ojos D., Alcala J., Eom C. B., Catalan G., Gruverman A.,
“Mechanical writing of ferroelectric polarization,” Science 336, 59–61 (2012).
Lubk A., Gemming S., Spaldin N. A., “First-principles study of ferroelectric domain walls in
multiferroic bismuth ferrite,” Phys. Rev. B 80, 104110 (2009).
Marton P., Rychetsky I., Hlinka J., “Domain walls of ferroelectric BaTiO3 within the Ginzburg-
Landau-Devonshire phenomenological model,” Phys. Rev. B 81, 144125 (2010).
Marton P., Rychetsky I., Hlinka J., “Erratum: domain walls of ferroelectric BaTiO3 within the
Ginzburg-Landau-Devonshire phenomenological model,” Phys. Rev. B 84, 139906(E) (2011).
Meier D., Seidel J., Cano A., Delaney K., Kumagai Y., Mostovoy M., Spaldin N. A., Ramesh R.,
Fiebig M., “Anisotropic conductance at improper ferroelectric domain walls.” Nat. Mater. 11,
284 (2012).
Mitsui T., Furuichi J., “Domain structure of rochelle salt and KH2 PO4 ,” Phys. Rev. 90, 193–202
(1953).
Nelson C. T., Gao P., Jokisaari J. R., Heikes C., Adamo C., Melville A., Baek S.-H.,
Folkman C. M., Winchester B., Gu Y., Liu Y., et al., “Domain dynamics during ferroelectric
switching,” Science 334, 968–971 (2011).
Padilla J., Zhong W., Vanderbilt D., “First-principles investigation of 180◦ domain walls in
BaTiO3 ,” Phys. Rev. B 53, R5969 (1996).
Parish M. M., Littlewood P. B., “Non-saturating magnetoresistance in heavily disordered semicon-
ductors,” Nature 426, 162–165 (2003).
Parkin S. S. P., Hayashi M., Thomas L., “Magnetic domain-wall racetrack memory,” Science 320,
190 (2008).
Privratska J., “Possible appearance of spontaneous polarization and/or magnetization in domain
walls associated with non-magnetic and non-ferroelectric domain pairs,” Ferroelectrics 353,
116 (2007).
Privratska J., Janovec V., “Pyromagnetic domain walls connecting antiferromagnetic non-
ferroelastic magnetoelectric domains,” Ferroelectrics 204, 321 (1997).
Privratska J., Janovec V., “Spontaneous polarization and/or magnetization in non-ferroelastic
domain walls: symmetry predictions,” Ferroelectrics 222, 23 (1999).
Salje E. K. H., Phase Transitions in Ferroelastic and Co-Elastic Materials (Cambridge University
Press, Cambridge, 1993).
Seidel J., Martin L. W., He Q., Zhan Q., Chu Y.-H., Rother A., Hawkridge M. E., Maksymovych
P., Yu P., Gajek M., Balke N., Kalinin S. V., Gemming S., Wang F., Catalan G., Scott J. F.,
Spaldin N. A., Orenstein J., Ramesh R., “Conduction at domain walls in oxide multiferroics,”
Nat. Mater. 8, 229 (2009).
Shilo D., Ravichandran G., Bhattacharya K., “Investigation of twin-wall structure at the nanometre
scale using atomic force microscopy,” Nat. Mater. 3, 453–457 (2004).
Sluka T., Tagantsev A. K., Bednyakov P., Setter N., “Free electron gas at charged domain walls in
insulating BaTiO3 ,” Nat. Commun. 4, 1808 (2013).
22 References
Stefani Ch., Ponet L., Shapovalov K., Chen P., Langenberg E., Schlom D. G., Artyurhin S.,
Stengel M., Domingo N., Catalan G., “Ferroelectric 180 degree walls are mechanically softer
than the domains they separate,” arXiv:2005.04249 (2020).
Stepkova V., Marton P., Hlinka J., “Stress-induced phase transition in ferroelectric domain walls
of BaTiO3 ,” J. Phys.: Condens. Mat. 24, 212201 (2012).
Tagantsev A. K., Courtens E., Arzel L., “Prediction of a low-temperature ferroelectric instability
in antiphase domain boundaries of strontium titanate,” Phys. Rev. B 64, 224107 (2001).
Van Aert S., Turner S., Delville R., Schryvers D., Van Tendeloo G., Salje E. K. H., “Direct obser-
vation of ferrielectricity at ferroelastic domain boundaries in CaTiO3 by electron microscopy,”
Adv. Mater. 24, 523 (2012).
Wada S., Yako K., Yokoo K., Kakemoto H., Tsurumi T., “Domain wall engineering in barium
titanate single crystals for enhanced piezoelectric properties,” Ferroelectrics 334, 17 (2006).
Wojdeł J. C., Íñiguez J., “Ferroelectric transitions at ferroelectric domain walls found from first
principles,” Phys. Rev. Lett. 112, 247603 (2014).
Xiao Y., Shenoy V. B., Bhattacharya K., “Depletion layers and domain walls in semiconducting
ferroelectric thin films,” Phys. Rev. Lett. 95, 247603 (2005).
Zhirnov V. A., “Contribution to the theory of domain walls in ferroelectrics,” Sov. Phys. JETP 35,
822 (1959).
Zubko P., Catalan G., Buckley A., Welche P. R. L., Scott J. F., “Strain-gradient-induced polarization
in SrTiO3 single crystals,” Phys. Rev. Lett. 99, 167601 (2007).
Zubko P., Catalan G., Tagantsev A. K., “Flexoelectric effect in solids,” Annu. Rev. Mater. Res. 43,
387–421 (2013).
2
Novel Phases at Domain Walls
S. Farokhipoor1 , C. Magen2 , D. Rubi3 , and
B. Noheda
1 Zernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4,
9747AG Groningen, The Netherlands
2 Instituto de Ciencia de Materiales de Aragón, CSIC—Universidad de Zaragoza,
Consejo Nacional de Investigaciones Científicas y Técnicas, Gral. Paz 1499, San Martín,
Argentina
S. Farokhipoor, C. Magen, D. Rubi, and B. Noheda, Novel Phases at Domain Walls In: Domain Walls: From Fundamental Properties to
Nanotechnology Concepts. Edited by: Dennis Meier, Jan Seidel, Marty Gregg, and Ramamoorthy Ramesh, Oxford University Press (2020).
© S. Farokhipoor, C. Magen, D. Rubi, and B. Noheda.
DOI: 10.1093/oso/9780198862499.003.0002
24 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
a b
Tb
O
Mn
Figure 2.1 Prototype orthorhombic crystal structure of TbMnO3 . The oxygen octahedra contained in
one unit cell are shaded. Mn3+ cations are placed inside the octahedra, and the Tb3+ cations are located
in the voids between the octahedra.
electrical properties (Goodenough 1955; Coey et al. 1999; Salamon and Jaime 2001).
Even though we focus on a perovskite manganite, TbMnO3 , the effects presented here
may be generalized to different types of materials with similar constraints.
TbMnO3 has a distorted perovskite structure with orthorhombic symmetry, as shown
in Figure 2.1. In this particular atomic arrangement, the magnetic interactions between
the 3d electrons of Mn3+ ions take place by superexchange mediated by the oxygen 2p
orbitals, and TbMnO3 shows magnetic frustration due to the competition between ferro-
magnetic and antiferromagnetic exchange interactions involving the magnetic moments
at near neighbors’ and next near neighbors’ Mn3+ sites. This causes antiferromagnetic
ordering below TN ≈ 41K, first in the form of a sinusoidal spin density wave and,
upon further cooling, as a cycloidal spin modulation, below TN ≈ 25K. The cycloid
arises together with a relative displacement of the oxygens, due to the presence of
the inverse Dzyaloshinskii–Moriya interaction, breaking the inversion symmetry and
inducing ferroelectricity (Kimura et al. 2003; Kenzelmann et al. 2005).
Perovskite-based crystals are rarely homogeneous. Symmetry lowering from the high-
temperature cubic (parent) phase gives rise to the formation of crystallographic domains
or twins, that is, regions of identical crystal structure but different crystal orientation.
Because different crystal orientations lead to different strain states, these domains are
ferroelastic (stress can be used to tune the deformation between two or three different
states) (Salje 1993). In bulk materials and in the absence of external forces, transitioning
Novel Phases at Domain Walls 25
to the low-symmetry phase will take place by random nucleation, and subsequent growth,
of domains at different parts of the crystal. The transition is completed when all domains
meet, forming domain walls. In order to minimize the elastic energy, domain walls are
formed along those atomic planes that are closely oriented in contiguous domains and
can be brought to coincide by small adaptations (see also Chapters 1 and 4). If the crystal
is free, this can take place by tilting of the twin domains, but if, on the other hand, the
crystal is confined to a bulkier substrate, as in the case of epitaxial growth of thin films,
twinning can be accompanied by very large strain gradients (Catalan et al. 2011). In this
case of epitaxial growth, for an appropriate lattice mismatch between the film and the
substrate, domain patterns can be formed by the alternation of two or more domains in a
periodic fashion, rather than by random nucleation, in order to achieve lattice matching
of the film with the substrate (Roitburd 1976). In this way, periodic arrays of domain
walls are introduced during the transition to the distorted phase (Tagantsev et al. 2010).
Due to the crucial role that structural details play in determining the ferroic order
parameters, it is naturally expected that the properties of multiferroics, and in particular
manganites, are greatly modified (and possibly improved) via strain-engineered domain
formation, as addressed in Chapter 5. Next to the structural and elastic degrees of
freedom, in multiferroic materials, the magnetic and electric boundary conditions also
play a crucial role in determining the formation of magnetic and ferroelectric domains,
respectively. Depolarization or demagnetization fields that develop in the material as
a result of the presence of a homogeneous polarization or magnetization, respectively,
become stronger as the dimensions perpendicular to the net moment decrease, and
thus there is a critical size below which a homogeneous magnetization or polarization
cannot be stabilized. The material, then, will break into domains, the size of which will
be determined by the balance between the dipolar/magnetic energy of the domains and
the anisotropy energy, which determines the domain wall formation energy.
The simplest model that proposes this balance was reported for a slab of magnetic
material by Landau and Lifshitz (Landau and Lifshitz 1935) and Kittel (Kittel 1946) by
considering that the domain energy density is determined by the demagnetization field
of the spins perpendicular to the surface. If the magnetic anisotropy is large enough,
stripe domains, separated by 180◦ domain walls, form; while for lower anisotropy, closure
domains and vortices are preferred, all of them causing the net magnetization to vanish.
Added terms to the energy, such as the Dzyaloshinskii–Moriya (or asymmetric exchange)
contribution, lead to other topological defects like skyrmions (Nagaosa and Tokura
2013). These considerations can be also applied to the case of ferroelectric domains
(Mitsui and Furuichi 1953) and can be extended to magnetoelectric multiferroics
(Daraktchiev et al. 2008). Ferroelectrics are largely anisotropic, compared to their
magnetic counterparts, resulting in much narrower domain walls (often just one atom
thick) and making it considerably more challenging to observe closure domains, vortices
or skyrmions, which have only recently been reported in ferroelectrics (Nahas et al. 2015;
Das et al. 2019).
As described in Schilling et al. (Schilling at al. 2006), the free energy of a ferroic
material should contain a term that describes the energy density of the domains and an-
other one that represents the domain wall formation energy. The former is proportional
26 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
108
107
106
105
w2(nm2)
104
103
102
101
100
BaTiO3 (ferroelastic)
104
103
102
101
100
100 101 102 103 104 105 106
film thickness (nm)
Figure 2.2 (a) Square of the domain width as a function of the thickness, demonstrating that the
quadratic dependence spans six orders of magnitude in thickness and that it is valid both for
ferromagnets and for ferroelectrics. (b) If the square of the domain width is divided by the domain wall
width, the quadratic law is independent of the nature of the dipolar interactions, and universality is
revealed.
Source: Taken from Catalan et al. 2012. Reproduced with permission from the American Physical Society.
to the domain size w, while the latter is proportional to both the number of domain
walls (and thus proportional to 1/w) and the domain wall area (and thus proportional to
the film thickness, d). If other contributions, such as the wall–wall interactions, charge
screening, etc. can be neglected, then the derivative of the free energy with these two
terms leads to an equilibrium domain size of the form w = Ad 1/2 , where A depends on
the ratio between the anisotropy of the order parameter and the dipolar interaction or, in
other words, on the domain wall width. This quadratic behavior can be shown in Figure
2.2(a) for different ferroelectric and ferromagnetic materials. It can also be observed in
this figure that ferromagnetic systems give rise to a larger A pre-factor (≈10–100 nm1/2 )
than that found for ferroelectric/ferroelastic systems (~1–5 nm1/2 ), for which a strong
anisotropy exists, reflecting the very different sizes of the domain walls in both cases. If
the domain wall width is eliminated from the thickness scaling (see Figure 2.2(b)), then
Novel Phases at Domain Walls 27
Tb3+
Mn3+
O2–
Sr2+
Ti4+
O2–
Figure 2.3 A sketch of the atomic arrangement around a TbMnO3 (110)o domain wall between two
orthorhombic domains (with the ao and bo directions interchanged with respect to each other). In the
domains, the pseudocubic unit cells of orthorhombic TbMnO3 form in zigzag fashion along the [001]
direction. At the twin domain, the zigzag is mirrored, creating a short Tb-Tb bond distance at every
other Tb atom of the domain wall. The experiments show that, at those sites with short bond distances,
the Tb3+ cation is replaced with Mn3+ . The cubic unit cells in the bottom row represent the SrTiO3
underlying substrate.
Source: Farokhipoor et al. (2014). Reproduced with permission from Springer Nature.
6
Remanent Min
Min(H = 0) (μ B f.u.–1)
0.4
4
0.2
2
0.0 0
0.0 0.2
1/d (nm–1)
Figure 2.4 Density of the domain walls and in-plane magnetic moment at H = 0 versus the inverse of
the thickness.
Source: Farokhipoor et al. (2014). Reproduced with permission from Springer Nature.
30 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
2 nm
(Tb1–xMnx)MnO3
TbMnO3
SrTiO3 (001)
Figure 2.5 High-angle annular dark field (HAADF) STEM image of region of the novel (encircled)
phase, equivalent to ~12 consecutive domain walls.
EELS) spectrum imaging shows that, at the domain walls, every other Tb3+ atomic
column is substituted by Mn3+ ions along the growth direction. Indeed, as shown in the
figure, the accommodation of opposite zigzag patterns at the domain walls gives rise to
too short Tb3+ -Tb3+ bond distances for every other Tb atom, along the [001] direction.
This unfavorable configuration induces the structure to rearrange itself by substituting
Tb3+ by the smaller Mn3+ cation in order to locally release the strain at the boundaries
of the twins (see also Chapter 3).
This novel phase is not only restricted to a single isolated 2D domain wall. In some
regions, the strain conditions are such that this structure spreads over extended regions of
the film, as though a succession of domain walls occurs. This is illustrated in Figure 2.5, in
which the region separating two domains is not an atomic flat domain wall, but a volume
equivalent to approximately 12 domain walls in width. The approximate stoichiometry of
this new structure, assuming that the chemical substitution of alternate Tb columns along
[001] direction is complete and there are no oxygen vacancies, would be (Tbx Mn1-x )
MnO3 , with x ∼ 0.5.
(a) (b)
Spectrum Image
2 nm
HAADF Mn O
(c)
80
Relat. Composition (at.%)
70 O
60
Mn
30
20
0 1 2 3 4 5
x(nm)
Figure 2.6 STEM-EELS spectrum image of a region of 25-nm-thick TbMnO3 thin film grown on
SrTiO3 . (a) High-angle annular dark field (HAADF) reference image; (b) HAADF signal (in grey scale),
Mn L2,3 signal, and O K content; (c) profile of relative Mn/O ratio quantified from (b), determined
along the white arrow in (a) by integrating 20 pixels horizontally. Dashed lines correspond to the
nominal stoichiometry (Mn:O = 1:3).
calculated inelastic scattering cross section of the different species can provide the
compositional ratio between two or more chemical elements in a material (Egerton
2011). Figure 2.6 shows the STEM-EELS spectrum images collected in a region
containing two domain walls in a 25-nm-thick TbMnO3 film grown on SrTiO3 . While
the estimated Mn:O ratio of the domains approaches the nominal 1:3 ratio (25% Mn at.,
75% O at.) the oxygen content at the domain walls drops to approximately 70% O at.
This non-stoichiometry of the wall is caused by Mn replacement of Tb columns, which
is not compensated with higher O content and, therefore, increases the local Mn content
at the domain wall.
The off-stoichiometry of the TbMnO3 domain walls would impact the Mn valence,
too. The Mn oxidation state has been frequently studied by analyzing the EELS fine
structure of the O-K edge or the Mn L2,3 edge, which are extremely dependent on
the hybridization of the Mn-3d and O-2p orbitals, and thus on the oxygen content
32 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
Figure 2.7 Estimation of Mn nominal valence obtained from the EELS fine structure analysis of
TbMnO3 region containing a domain wall. (a) HAADF image of the region; (b) Mn valence determined
from the shift of the O-K pre-peak; (c) relative energy shift of the L3 and L2 Mn edges; (d, e) examples of
O-K and Mn L2,3 edges in the different regions of the domains and the domain wall.
Source: Adapted from Farokhipoor et al. 2014. Reproduced with permission from Springer Nature.
and chemical environment of Mn. The combined change in stoichiometry and novel
crystal environment for Mn can be translated into a variation of the nominal Mn valence.
This can be tracked by following the previous calibration performed in isoelectronic
structures (lanthanum-calcium manganites as a function of the La/Ca content) in which
the position of the O-K pre-peak relative to the main O peak is analyzed (Varela et al.
2006). By Gaussian fitting of both peaks, the energy difference between the maxima can
be measured and correlated to the Mn oxidation state (Varela et al. 2006). Similarly, the
energy difference between the L2 and L3 white lines of the EELS Mn edge also depends
on the Mn valence. Figure 2.7 also shows a decrease in the Mn valence at the domain
wall; while the domains average a nominal Mn oxidation state of approximately +3.2,
the wall exhibits an average Mn valence of +3.0.
Even though a direct correlation between the stoichiometry and the nominal valence
values obtained here cannot be made, both point to the same direction; i.e. it is plausible
that Mn in the X columns would be less positively charged (that is, a +2 character),
which would reduce the oxidation state of the defect to +3, assuming that the Mn in the
B positions have the same valence as that measured in the domains. The evolution of the
relative positions of the L3 and L2 white lines of Mn is consistent with this scenario; a
shift of L3 of about 1 eV to lower energy losses is observed, which has been reported in
the literature and was associated with lower Mn valence (Schmid and Mader 2006).
All this points toward a quite unique driving mechanism behind the periodic structure
that is observed in TbMnO3 /SrTiO3 , notably different from any other type of ordering
reported in thin film heterostructures. Indeed, other studies present lattice-modulated
patterns involving oxygen vacancies ordering, for instance, in cobaltites (Choi et al. 2012;
Novel Phases at Domain Walls 33
Gazquez et al. 2013; Biškup et al. 2014; Kwon et al. 2014). For LaCoO3 , the origin
of magnetism is controversial, with one report ascribing it to the presence of oxygen
vacancies (Gazquez et al. 2013; Biškup et al. 2014), whereas other reports correlate
it with the high spin state of Co3+ (Choi et al. 2012; Kwon et al. 2014). In these
works, the stripes look more like a periodically structured region with different chemical
composition due to oxygen vacancies, and not necessarily located at the domain walls
(Choi et al. 2012; Biškup et al. 2014; Kwon et al. 2014). Nevertheless, we postulate
that this phenomenon could be more generally present in (001)-oriented A3+ B3+ O3
orthorhombic perovskites, since they all show an A-cation zigzag configuration along the
[001] direction and the same valence state for the small and large cations. In particular,
multivalent B-cations, such as Mn, Ni, or Cr, should allow for more flexible coordination
and formation of new phases at the domain walls. A similar magnetic trend with thickness
has been observed in orthorhombic ScMnO3 grown on LaAlO3 ; a more detailed TEM
and neutron diffraction studies is still pending (Wang et al. 2015).
2.4 Summary
In this chapter, we describe a route for the stabilization of novel two-dimensional
phases at structural domain walls that self-assemble during growth, driven by the large
stresses present at the ferroelastic twin boundaries. In the case of TbMnO3 /SrTiO3
heterostructures, the change in chemistry takes place by selective substitution of half
the Tb cations at the domain wall by Mn cations, with an accompanying decrease in
Mn valence at the domain wall sites. This two-dimensional phase orders magnetically
at low temperatures. We believe that this mechanism should also be active in other
orthorhombic A3+ B3+ O3 perovskites.
..........................................................................................
REFERENCES
Biškup N., Salafranca J., Mehta V., Oxley M. P., Suzuki Y., Pennycook S. J., Pantelides S. T.,
Varela M., “Insulating ferromagnetic LaCoO3−δ films: a phase induced by ordering of oxygen
vacancies,” Phys. Rev. Lett. 112, 087202 (2014).
Bode M., Vedmedenko E. Y., Von Bergmann K., Kubetzka A., Ferriani P., Heinze S., Wiesendanger
R., “Atomic spin structure of antiferromagnetic domain walls,” Nat. Mater. 5, 477 (2006).
Brinkman A., Huijben M., van Zalk M., Huijben J., Zeitler U., Maan J. C., van de Wiel W. G.,
Rijnders G., Blank D. H. A., Hilgenkamp H., “Magnetic effects at the interface between non-
magnetic oxides,” Nat. Mater. 6, 493 (2007).
Catalan G., Lubk A., Vlooswijk A. H. G., Snoeck E., Magen C., Janssens A., Rispens G., Blank
D. H. A., Noheda B., “Flexoelectric rotation of polarization in ferroelectric thin films,” Nat.
Mater. 10, 963 (2011).
Catalan G., Lukyanchuk I., Schilling A., Gregg J. M., Scott J. F., J. Mater. Sci. 44, 5307 (2009).
Catalan G., Seidel J., Ramesh R., Scott J. F., “Domain wall nanoelectronics,” Rev. Modern Phys. 84,
119 (2012).
34 References
Choi W. S., Kwon J.-H., Jeen H, Hamann-Borrero J. E., Radi A., Macke S., Sutarto R., He F.,
Sawatzky G. A., Hinkov V., Kim M., Lee H. N., “Strain-induced spin states in atomically
ordered cobaltites,” Nano. Lett. 12, 4966 (2012).
Coey J. M. D., Viret M., von Molnar S., “Mixed-valence manganites,” Adv. Phys. 48, 167 (1999).
Daraktchiev M., Catalan G., Scott J. F., “Landau theory of ferroelectric domain walls in magneto-
electrics,” Ferroelectrics 375, 122 (2008).
Das S., Tang Y. L., Hong Z., Gonçalves M. A. P., McCarter M. R., Klewe C., Nguyen K. X.,
Gómez-Ortiz F., Shafer P., Arenholz E., Stoica V. A., Hsu S.-L., Wang B., Ophus C., Liu J. F.,
Nelson C. T., Saremi S., Prasad B., Mei A. B., Schlom D. G., Íñiguez J., García-Fernández
P., Muller D. A., Chen L. Q., Junquera J., Martin L. W., Ramesh R., “Observation of room-
temperature polar skyrmions,” Nature 568, 368 (2019).
Daumont C. J. M., Farokhipoor S., Ferri A., Wojdel J. C., Iniguez J., Kooi B. J., Noheda B., “Tuning
the atomic and domain structure of epitaxial films of multiferroic BiFeO3 ,” Phys. Rev. B 81,
144115 (2010).
Daumont C. J. M., Mannix D., Venkatesan S., Rubi D., Catalan G., Kooi B. J., De Hosson
J. Th. M., Noheda B., “Epitaxial TbMnO3 thin films on SrTiO3 substrates: a structural study,”
J. Phys.: Condens. Mat. 18, 182001 (2009).
Dong S., Yu R., Yunoki S., Liu J.-M., Dagotto E., “Double-exchange model study of multiferroic
RMnO3 perovskites,” Eur. Phys. J. B 71, 339 (2009).
Egerton R. F., Electron Energy-Loss Spectroscopy in the Electron Microscope, 3rd ed. (Springer,
Boston, 2011).
Farokhipoor S., Magén C., Venkatesan S., Iñiguez J., Daumont C. J. M., Rubi D., Snoeck E.,
Mostovoy M., de Graaf C., Müller A., Döblinger M., Scheu C., Noheda B., “Artificial chemical
and magnetic structure at the domain walls of an epitaxial oxide,” Nature 515, 379 (2014).
Fontcuberta J., “Multiferroic RMnO3 thin films,” C. R. Phys. 16, 204 (2015).
Gazquez J., Bose S., Sharma M., Torija M. A., Pennycock S. J., Leighton C., Varela M., “Lattice
mismatch accommodation via oxygen vacancy ordering in epitaxial La0.5 Sr0.5 CoO3-δ thin
films,” APL Mater. 1, 012105 (2013).
Goodenough J. B., “Theory of the role of covalence in the perovskite-type manganites [La, M(II)]
MnO3 ,” Phys. Rev. 100, 564 (1955).
Kenzelmann M., Harris A. B., Jonas S., Broholm C., Schefer J., Kim S. B., Zhang C. L., Cheong
S.-W., Vajk O. P., Lynn J. W., “Magnetic Inversion Symmetry Breaking and Ferroelectricity in
TbMnO3 ,” Phys. Rev. Lett. 95, 087206 (2005).
Kimura Y., Goto T., Shintani H., Ishizaka K., Arima T., Tokura Y., “Magnetic control of
ferroelectric polarization,” Nature 426, 55 (2003).
Kittel C., “Theory of the structure of ferromagnetic domains in films and small particles,” Phys.
Rev 70, 965 (1946).
Kwon J.-H., Choi W. S., Kwon Y.-K., Jung R., Zuo J.-M., Lee H. N., Kim M., “Nanoscale spin-
state ordering in LaCoO3 epitaxial thin films,” Chem. Mater. 26, 2496 (2014).
Landau L., Lifshitz E., “On the theory of the dispersion of magnetic permeability in ferromagnetic
bodies,” Phys. Z. Sowjet. 8, 153 (1935).
Ma Y., Dai Y., Guo M., Niu Ch., Zhu Y., Huang B., “Evidence of the existence of magnetism in
pristine VX2 monolayers (X = S, Se) and their strain-induced tunable magnetic properties,”
ACS Nano. 6, 1695 (2012).
Marti X., Sanchez F., Skumryev V., Laukhin V., Ferrater C., Garcia-Cuence M. V., Varela
M., Fontcuberta J., “Crystal texture selection in epitaxies of orthorhombic antiferromagnetic
YMnO3 films,” Thin Solid Films 516, 4899 (2008).
Marti X., Skumryev V., Cattoni A., Bertacco R., Laukhin V., Ferrater C., Garcia-Cuenca M. V.,
Varela M., Sanchez F., Fontcuberta J., “Ferromagnetism in epitaxial orthorhombic YMnO3 thin
films,” J. Magn. Magn.Mater 321, 1719 (2009).
Novel Phases at Domain Walls 35
Marti X., Skumryev V., Ferrater C., Garcia-Cuence M. V., Varela M., Sanchez F., Fontcuberta J.,
“Emergence of ferromagnetism in antiferromagnetic TbMnO3 by epitaxial strain,” Appl. Phys.
Lett. 96, 222505 (2010).
Mitsui T., Furuichi J., “Domain structure of rochelle salt and KH2 PO4 ,” Phys. Rev. 90, 193 (1953).
Nagaosa N., Tokura Y., “Topological properties and dynamics of magnetic skyrmions,” Nat.
Nanotechnol. 8, 899 (2013).
Nahas Y., Prokhorenko S., Louis L., Gui Z., Kornev I., Bellaiche L., “Discovery of stable
skyrmionic state in ferroelectric nanocomposites,” Nat. Comm. 6, 8542 (2015).
Pennycook S. J., Nellist P. D., Scanning Transmission Electron Microscopy (Springer, New York,
2011).
Pertsev N. A., Zembilgotov A. G., “Energetics and geometry of 90◦ domain structures in epitaxial
ferroelectric and ferroelastic films,” J. Appl. Phys. 78, 6170 (1995).
Pompe W., Gong X., Suo Z., Speck J. S., “Elastic energy release due to domain formation in the
strained epitaxy of ferroelectric and ferroelastic films,” J. Appl. Phys. 74, 6012 (1993).
Roitburd A. L., “Equilibrium structure of epitaxial layers,” Phys. Status Solidi A 37, 329 (1976).
Rubi D., de Graaf C., Daumont C. J. M., Mannix D., Broer R., Noheda B., “Ferromagnetism and
increased ionicity in epitaxially grown TbMnO3 films,” Phys. Rev. B 79, 014416 (2009).
Rubi D., Venkatesan S., Kooi B. J., De Hosson J. T. M., Palstra T. T. M., Noheda B., “Magnetic
and dielectric properties of YbMnO3 perovskite thin films,” Phys. Rev. B 78, 020408 (2008).
Salamon M. B., Jaime M., “The physics of manganites: structure and transport,” Rev. Mod. Phys.
73, 583 (2001).
Salje E. K., Phase Transitions in Ferroelastic and Co-elastic Crystals (Cambridge University Press,
Cambridge, UK, p. 296, 1993).
Schilling A., Adams T. B., Bowman R. M., Gregg J. M., Catalan G., Scott J. F., “Scaling of domain
periodicity with thickness measured in BaTiO3 single crystal lamellae and comparison with
other ferroics,” Phys. Rev. B 74, 024115 (2006).
Schmid H. K., Mader W., “Oxidation states of Mn and Fe in various compound oxide systems,”
Micron. 37, 426 (2006).
Tagantsev A. K., Cross L. E., Fousek J., Domains in Ferroic Crystals and Thin Films (Springer, New
York, 2010).
Takahashi K. S., Kawasaki M., Tokura Y., “Interface ferromagnetism in oxide superlattices of
CaMnO3 /CaRuO3 ,” Appl. Phys. Lett. 79, 1324 (2001).
Varela M., Pennycook T. J., Tian W., Mandrus D., Pennycook S. J., Peña V., Sefrioui Z., Santamaria
J., “Atomic scale characterization of complex oxide interfaces,” J. Mater. Sci. 41, 4389 (2006).
Venkatesan S., Daumont C. J. M., Kooi B. J., Noheda B., de Hosson J. Th. M., “Nanoscale domain
evolution in thin films of multiferroic TbMnO3 ,” Phys. Rev. B 80, 214111 (2009).
Wang F., Zhang Y. Q., Liu W., Ning X. K., Bai Y., Dai Z. M., Ma S., Zhao X. G., Li K., Zhang
Z. D., “Abnormal magnetic ordering and ferromagnetism in perovskite ScMnO3 film,” Appl.
Phys. Lett. 106, 232906 (2015).
Yang H. X., Hallal A., Terrade D., Waintal X., Roche S., Chshiev M., “Proximity effects induced
in graphene by magnetic insulators: first-principles calculations on spin filtering and exchange-
splitting gaps,” Phys. Rev. Lett. 110, 046603 (2013).
3
First-Principles Studies of Structural
Domain Walls
J. Íñiguez1,2
1 MaterialsResearch and Technology Department, Luxembourg Institute of Science and
Technology (LIST), 5 avenue des Hauts-Fourneaux, L-4362 Esch/Alzette, Luxemburg
2 Department of Physics and Materials Science, University of Luxembourg, 41 rue du Brill,
3.1 Introduction
Structural domain walls are critical to the properties of multi-domain states in ferroelec-
tric and ferroelastic materials. For one thing, the mobility of the walls, or the lack thereof,
largely determines the response of a compound to external stimuli. For another, walls
may have interesting functional properties of their own, either intrinsic or resulting from
the various types of defects they usually attract (see also Chapter 1.4). While not new
(Salje 2010), the latter concept has become particularly popular in recent times, together
with the notion that suitably engineered walls may provide a path toward the production
of novel densely packed nanodevices. This “the wall is the device” philosophy (Catalan
et al. 2012) motivates most of the recent first-principles work on the topic, and is a focus
of this chapter.
Why should walls present peculiar properties that set them apart from the domains?
Why should they be the generally preferred location of point defects? The wall atomic
structure is not a low-energy one; rather, it is one forced by the structural boundary
conditions imposed by the surrounding domains that meet (clash) at the walls. Hence, the
walls are at a relatively high-energy state in which the dominant interatomic couplings in
the material are somewhat frustrated. Two expectations follow. First, the properties of the
wall may depart from those of the domains, because their structures are different. Second,
when it comes to accommodating defects, the walls seem to have a starting advantage
as compared to the domains. In a domain, a defect creates a structural disruption that
usually has a strong penalty associated with it, as it prevents the material from adopting
its lowest-energy configuration. In contrast, the energy of the walls is not expected to
increase as much in the presence of a defect1 ; in fact, it is even conceivable that certain
defects might allow a wall to optimize its structure and energy (Zhao and Íñiguez 2019).
1 This hand-waving argument is supported by a recent simulation study of PbTiO ’s walls (Paillard et al.
3
2017); yet, exceptions to the rule are known too (Skjærvø et al. 2018).
J. Íñiguez, First-Principles Studies of Structural Domain Walls In: Domain Walls: From Fundamental Properties to Nanotechnology Concepts.
Edited by: Dennis Meier, Jan Seidel, Marty Gregg, and Ramamoorthy Ramesh, Oxford University Press (2020). © J. Íñiguez.
DOI: 10.1093/oso/9780198862499.003.0003
First-Principles Studies of Structural Domain Walls 37
A B O
Figure 3.1 Most common structural instabilities in perovskites. Structure of an ABO3 cubic
perovskite. A cations (shaded balls) occupy the cell corners, B cations (white balls) the cell center, and
oxygen anions (black balls) the face centers. In panel (a), the arrows represent the local atomic
displacements that are typical of a ferroelectric instability. Panels (b) and (c) show typical
antiferrodistortive tiltings of the O6 octahedral groups, in antiphase and in phase, respectively.
What do we know about the structure of structural walls? Fortunately, more and more,
thanks to the ongoing revolution in local-probe techniques. Yet, a detailed experimental
characterization of the materials (devices) in the relevant conditions (of operation)
remains a great challenge, and our detailed knowledge is still limited. Despite that, domain
boundaries have long been discussed at a theoretical level, traditionally in the framework
of Ginzburg–Landau models (Lajzerowicz and Niez 1979; Houchmandzadeh et al.
1991; Tagantsev et al. 2010). Within such approaches, the picture of the wall structure
is straightforward: This is a region in which the ferroic order parameter (e.g. the electric
polarization of Figure 3.1(a)) changes sign as we move from one domain to another;
hence, being associated with a locally null order parameter, the wall looks like the high-
symmetry phase of the compound (see also Chapter 4). A refinement of this picture is
available for materials presenting two competing orders (e.g. the electric polarization of
Figure 3.1(a) and the tilts of the O6 octahedra of Figures 3.1(b) and 3.1(c)); in such
cases, as already mentioned in Chapter 1.3, the weaker order parameter may appear
at the wall region, where the strongest dominant order vanishes (see Figure 3.2(a))
(Houchmandzadeh et al. 1991; Tagantsev et al. 2001).
This basic understanding of structural walls has dominated our thinking for decades.
Granted, the physical picture is beautiful and makes sense, and it should be essentially
correct provided two basic conditions are satisfied. At a qualitative level, the Ginzburg–
Landau models supporting it should contain all the order parameters (distortion fields)
that can play a relevant role at the walls. At a quantitative level, the model coefficients—
typically obtained from experimental information that essentially captures the behavior
of the domains, not the walls—should give a fair description of the boundary region.
38 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
These are perfectly sound hypotheses that have allowed our community to work on this
problem for decades (Chen 2008; Vasudevan et al. 2013). Nevertheless, today we are in
the position to ask ourselves: Is this all we need, or are we missing important effects?
The question of how to treat structural domain walls theoretically is a no-brainer,
at least in principle. We need methods that permit an unbiased description of unusual
(typically, unknown) high-energy structures, with atomic resolution, and not relying on
experimental information that we often lack. This is a perfect problem for modern first-
principles simulation techniques (Martin 2004), one where they are likely to make a
difference in terms of both gaining understanding and revealing unexpected behaviors.
Indeed, while traditional model-Hamiltonian approaches are limited by our imagination
(they are defined by a set of hand-picked variables and interactions among them),
first-principles simulations allow us, like experiments, the luxury of serendipity, that
is, of being surprised by results we would not have anticipated. Further, when it
comes to problems involving defects or, more generally, differences in chemical bonding
between domains and walls, a quantum-mechanical first-principles approach is all but
irreplaceable.
Here I discuss representative first-principles studies of structural domain walls in
ferroics, focusing on the compounds that have received the most attention by the
simulations community so far: perovskite oxides. It is not my intention to be exhaustive.
Rather, I describe in some detail a reduced number of case studies that come in handy
to illustrate different effects and to highlight the added value of the first-principles
investigations. As regards the simulation methods, I focus on applications of density
functional theory (DFT) (Hohenberg and Kohn 1964; Kohn and Sham 1965), typically
employing an approximation for an effective treatment of ionic cores (Pickett 1989;
Blöchl 1994); since these are standard and abundantly reviewed methods, I do not discuss
them here. I also include a section on the application to domain-wall problems of first-
principles-based methods for large-scale simulations of ferroelectrics and ferroelastics,
where our community has followed an original and productive path. Finally, I briefly
comment on the opportunities and challenges for first-principles research in this field.
yield various low-symmetry phases. For example, Figure 3.1(a) displays a characteristic
pattern of atomic displacements that, at a local level, yield an electric dipole. When
repeated homogeneously in a large region of the lattice, this distortion results in sizable
movement of bound charges and an associated electric polarization. Then, if such polar
distortions are oriented differently in different regions, domains appear. Indeed, there
are six symmetry-equivalent orientations of the local dipole shown in Figure 3.1(a),
parallel or antiparallel to the three principal axes of the cubic lattice, implying that six
equivalent ferroelectric domains may form in this case. The homogeneous replication of
the distortion in Figure 3.1(a) reduces the symmetry of the perovskite lattice from cubic
(Pm3m space group) to tetragonal (P4mm); this is exactly the case of PbTiO3 (Lines
and Glass 1977).
Figure 3.1(b) shows a second lattice instability that is ubiquitous in perovskites (Lines
and Glass 1977). Locally it involves a rotation of the O6 octahedron in the unit cell,
and the tilts are modulated in anti-phase as one moves between neighboring lattice
cells. This kind of distortion corresponds to an order parameter sometimes termed
antiferrodistortive. There are six symmetry-equivalent ways to have it, differing in the
axis and phase of the tilts, which yield six equivalent domains. Antiferrodistortive
instabilities like that of Figure 3.1(b) occur spontaneously in materials as SrTiO3 ,
where the symmetry of the cubic lattice is reduced to I 4/mcm. Also, first-principles
simulations have revealed that they are latent in materials like PbTiO3 , where the
antiferrodistortive mode competes (and losses) against the dominant ferroelectric order
parameter (Ghosez et al. 1999; Wojdeł et al. 2013).
In the following, I will refer to the polarization of a domain by the three-dimensional
vector Pα , where α = x, y, z labels the principal (pseudocubic) axes of the perovskite
lattice. Similarly, I will denote the antiferrodistortive order parameter of anti-phase tilts by
Rα , a three-dimensional (axial) vector whose direction and magnitude represent the axis
and angle of the octahedral tilts, respectively. In the above examples, I have assumed that
the symmetry-breaking distortions P and R are parallel to 100, but this does not need to
be the case. Indeed, similar ferroelectric or antiferrodistortive distortions oriented along
a 110 pseudocubic direction would yield orthorhombic low-symmetry phases, while
rhombohedral ones would be obtained if the order parameters lay along a 111 axis. (In
the following, all directions and planes are given in the pseudocubic setting.)
These are self-explanatory examples of symmetry-breaking structural transitions
where the condensation of a (soft) phonon of a high-symmetry phase drives a transfor-
mation into a low-symmetry structure (Cowley 1980; Bruce 1980). However, structural
domains can appear in cases that to do not match this picture exactly. An all-important
example is that of multiferroic perovskite BiFeO3 (Catalan and Scott 2009). In BiFeO3
the ferroelectric transition does not involve the cubic perovskite phase. Instead, the role
of the paraelectric phase is played by a strongly distorted structure (Arnold et al. 2010)
featuring a complex—although extremely frequent (Lufaso and Woodward 2001; Chen
et al. 2018b)—combination of octahedral tilts. This paraelectric phase is itself a low-
symmetry state that holds no group–subgroup relationship with the ferroelectric phase;
hence, it would be misleading (and incorrect) to talk about “symmetry breaking” in
BiFeO3 ’s ferroelectric transformation. Nevertheless, we can still postulate BiFeO3 ’s cubic
40 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
where C, C , and C are material-dependent constants, and the vector Q can stand for
either P or R. (The mathematical expressions for the strain are formally equivalent for
both P or R, but the respective C coupling coefficients will, of course, differ.) Hence, the
strain change across a ferroelectric or antiferrodistortive domain wall is fully determined
by the change in the primary order parameter.
An important point to notice is that ferroelectric and antiferrodistortive domain
walls may or may not involve a strain change. Consider for example the anti-phase
antiferrodistortive wall separating domains with RI [1, 0, 0] and RII [−1, 0, 0]; the strain
components involved here are ηxx ∝ Rx2 and ηyy = ηzz ∝ Rx2 , which do not depend on
the sign of Rx ; hence, this antiferrodistortive wall is not ferroelastic. In contrast, the 90◦
ferroelectric wall separating domains with PI = (P, 0, 0) and PII = (0, −P, 0) does involve
First-Principles Studies of Structural Domain Walls 41
Can we say something a priori about the preferred orientation of such a wall? Yes
indeed: Irrespective of the atomistic details of the ferroic distortions, walls will orient
themselves so as to minimize two large contributions to the energy of any material,
namely, electrostatic and elastic. (Electrostatic effects are all but restricted to insulating
and semiconducting compounds, which are the cases of interest here.)
In order to minimize electrostatic energy, ferroelectric walls will tend to be neutral,
implying that P · n̂ = 0, where P = PII − PI and n̂ is the unit vector normal to the wall
plane (see Figure 3.2(c)). Conversely, if P · n̂ = σwall = 0, the wall is said to be charged,
meaning that it would need to receive a σ wall charge density in order to be electrically
compliant. In this example, we have P = (−P, −P, 0), which implies that walls in planes
(1, −1, 0) or (0, 0, 1) will be neutral; in contrast, walls in planes (1, 1, 0) or (1, 0, 0) would
42 Domain Walls: From Fundamental Properties to Nanotechnology Concepts
(a) (b)
Q/Qmax
R
P
–P DW +P DW –P DW
(c) (d)
P –P
–P P –P
Figure 3.2 Domain wall features. Panel (a) illustrates the possible occurrence of a weaker order
parameter (P in this case) at the walls of a stronger order (R), as discussed by Houchmandzadeh et al.
(1991). Panel (b) illustrates the symmetry breaking at a ferroelastic wall involving a shear strain η; at
the walls, there is no symmetry between up and down, and an improper wall polarization (Pwall ) is
inevitable. Panel (c) shows two 180◦ ferroelectric walls, one neutral and the other charged. This charged
wall configuration corresponds to the maximum | P · n̂ | that is possible, and can be expected to be
unstable against distortions that reduce the wall charge. Panel (d) shows sketches of walls with Ising,
Nèel, and Bloch character, attending to the different ways in which the order parameter may evolve
across the wall.
⎞⎛
−1 0 0
η = ηII − ηI = C − C P 2 ⎝ 0 1 0⎠.
(3.3)
0 0 0
First-Principles Studies of Structural Domain Walls 43
To minimize its elastic energy, the wall (twin plane) between PI and PII must be such
that both domains lead to the same spontaneous strains within that plane. Mathematically,
this means that the application of the difference strain tensor to any vector u within the
twin plane must render either a null result or, at most, a distortion normal to the wall.
Hence, for example, for u = (0, 0, 1) we have η u = 0, implying that this direction is
equally
strained
by both ferroelastic variants. Additionally, for u = (1, 1, 0) we get η u =
C − C P 2 (−1, 1, 0); that is, the difference distortion is along a perpendicular direction.
Combined, these two conditions imply that the elastic energy is minimized if the twin
boundary lies in the (1, −1, 0) plane. As mentioned above, this plane is also electrically
I II
neutral, which makes it a clear candidate
to2 host the wall between P and P . In contrast,
for u = (0, 1, 0) we get η u = C − C P (0, 1, 0), implying that planes containing this
direction are not elastically compliant. For example, while electrically neutral, the (0, 0,
1) plane is unfavorable from the point of view of the elastic energy; thus, we do not expect
walls between PI and PII to be oriented in this way.
Finally, let me note that walls can further optimize their internal domain-wall structure
by implementing the discontinuity of the order parameter in different ways. The most
usual patterns are called Ising, Bloch and Nèel, respectively, as defined in Chapter 1. As
it is essential for the following discussion, sketches of these different boundaries are also
included in Figure 3.2(d). (For experimental studies on Ising, Bloch and Nèel walls, see
Chapter 7.)
Px (C/m2)
c 0.4
x domain I domain II
0
a PBEsol (I)
DW DW DW –0.4 x
z PBEsol (II)
–0.8
domain I DW domain II
0.8
Py (C/m2) 0.4
0
x DW
–0.4
PbO planes TiO2 planes y
–0.8
z
0 5 10 15 20
DW DW Pb Ti O
DW Cell number along z direction
Figure 3.3 Ferroelectric domain walls in PbTiO3 . Panel (a) shows periodically repeated supercells
typically considered to investigate walls of 180◦ (top) and 90◦ (bottom) in PbTiO3 . In the former case,
PI and PII are both along the [001] pseudocubic direction, and the wall is in the (1,0,0) plane. In the
latter, always in the pseudocubic setting, we have PI = (P, 0, 0) and PII = (0, P, 0), and the wall plane
is (1, 1, 0). Note that we have symmetry-equivalent walls at the center and boundaries of the simulation
supercells. Panel (b) summarizes the results by Wojdeł and Íñiguez (2014) for the internal structure of
180◦ walls: The domain polarizations are along x, and the wall is in a (001) plane. A wall polarization
along y develops—as observed in first- (lines) and second- (symbols) principles simulations—which
confers the wall a Bloch character. Details of the atomic distortions at domains and the wall are also
shown. Panel (c) sketches the first-principles result for the multidomain structure of PbTiO3 layers
inserted between SrTiO3 layers in a PbTiO3 /SrTiO3 superlattice.
Source: (a) Adapted from Meyer and Vanderbilt 2002. (c) Taken from Aguado-Puente and Junquera 2012.
wall motion, albeit within drastic approximations (the whole wall plane was assumed
to move rigidly, and the transformation path was obtained from a simple interpolation)
that yield a (probably much exaggerated) upper limit. They also discussed the electric
potential associated with the wall discontinuity, revealing that even such idealized walls
constitute shallow carrier traps. A brief summary of their results is included in Table 3.1.
While being the most important and influential, this study (Meyer and Vanderbilt
2002) was not the first first-principles investigation of the walls of PbTiO3 . Three years
earlier, Pöykkö and Chadi (1999) had conducted a work restricted to the 180◦ boundaries
(see Table 3.1), whose conclusions align with those of Meyer and Vanderbilt except in
one aspect: the former authors report an internal polarization confined to the wall plane,
while the latter did not find such a feature, in spite of explicitly looking for it. Thus,
following the usual domain-wall classification depicted in Figure 3.2(d), Pöykkö and
Chadi predict PbTiO3 ’s walls to have a Bloch character, while Meyer and Vanderbilt’s
results (as well as others’ (Behera et al. 2011)) suggest they have an Ising character.
More recent DFT investigations (Wojdeł and Íñiguez 2014; Wang et al. 2014; Liu
and Cohen 2017b) have confirmed the Bloch polarization at the 180◦ boundaries (see
Figure 3.3(b)). These works also discuss the energetics of the wall polar instability,
showing that it has essentially the same nature as the polar distortion within the domains
(i.e. it is dominated by the off-centering of the Pb cations). They also emphasize
how the polar instability prevails in spite of the factors against it (Wojdeł and Íñiguez
2014); for example, the confinement to the wall plane implies the truncation of many
Exploring the Variety of Random
Documents with Different Content
ceremony, M. Lichtenstein, who is remarkable for his modesty, left
Berlin for Trieste, from whence he was to proceed to Alexandria.
This year, 1852, the Royal Academy of Sweden has caused its annual
medal to be struck to the memory of the celebrated Swedenborg,
one of its first members. The medal, which has already been
distributed to the associates, has, on the obverse, the head of
Swedenborg, with, at the top, the name, Emanuel Swedenborg; and
underneath, Nat. 1688. Den. 1772. And on the reverse, a man in a
garment reaching to the feet, with eyes unbandaged, standing
before the temple of Isis, at the base of which the goddess is seen.
Above is the inscription: Tantoque exsultat alumno; and below: Miro
naturæ investigatori socio quond. æstimatiss. Acad. reg. Scient.
Soec. MDCCCLII.
In Sweden during the year 1851 there were 1060 books published,
and 113 journals. Of the books, 182 were theological, 56 political,
123 legal, 80 historical, 55 politico-economical and technical, 45
educational, 40 philological, 38 medical, 31 mathematical, 22
physical, 18 geographical, 3 æsthetical, and 3 philosophical. Fiction
and Belles-Lettres have 259; but they are mostly translations from
English, French, and German. Of these details we are tempted to
say, remarks the Leader, what Jean Paul's hero says of the lists of
Errata he has been so many years collecting—"Quintus Fixlein
declared there were profound conclusions to be drawn from these
Errata; and he advised the reader to draw them!"
Among the few French books worthy of notice, says the Leader, let
us not forget the fourth volume of Saint Beuve's charming Causeries
du Lundi, just issued. The volume opens with an account of
Mirabeau's unpublished dialogues with Sophie, and some delicate
remarks by Sainte Beuve, in the way of commentary. There are also
admirable papers on Buffon, Madame de Scudery, M. de Bonald,
Pierre Dupont, Saint Evremont et Ninon, Duc de Lauzun, &c.
Although he becomes rather tiresome if you read much at a time,
Sainte Beuve is the best article writer (in our Macaulay sense)
France possesses. With varied and extensive knowledge, a light,
glancing, sensitive mind, and a style of great finesse, though
somewhat spoiled by affectation, he contrives to throw a new
interest round the oldest topics; he is, moreover, an excellent critic.
Les Causeries du Lundi is by far the best of his works.
Dramatic literature is lucrative in France. The statement of finances
laid before the Dramatic Society shows, that during the years 1851-
52, sums paid for pieces amount to 917,531 francs (upward of
£36,000). It would be difficult to show that English dramatists have
received as many hundreds. The sources of these payments are thus
indicated. Theatres of Paris, 705,363 francs; the provincial theatres,
195,450 francs (or nearly eight thousand pounds; whereas the
English provinces return about eight hundred pounds a year!)—and
suburban theatres, 16,717 francs. To these details we may add the
general receipts of all the theatres in Paris during the year—viz., six
millions seven hundred and seventy-one thousand francs, or
£270,840.
Comicalities, Original and Selected.
Pg 356, word "upon" removed from sentence "...attack upon [upon] Mr. Dutton's purse..."
Pg 378, sentence "(TO BE CONTINUED.)" added to the end of article.
Pg 386, word "of" added to sentence "...the wish of the son..."
Pg 416, word "is" removed from sentence "Here [is] is a very amusing picture..."
*** END OF THE PROJECT GUTENBERG EBOOK HARPER'S NEW
MONTHLY MAGAZINE, NO. XXVII, AUGUST 1852, VOL. V ***
Updated editions will replace the previous one—the old editions will
be renamed.
1.D. The copyright laws of the place where you are located also
govern what you can do with this work. Copyright laws in most
countries are in a constant state of change. If you are outside the
United States, check the laws of your country in addition to the
terms of this agreement before downloading, copying, displaying,
performing, distributing or creating derivative works based on this
work or any other Project Gutenberg™ work. The Foundation makes
no representations concerning the copyright status of any work in
any country other than the United States.
1.E.6. You may convert to and distribute this work in any binary,
compressed, marked up, nonproprietary or proprietary form,
including any word processing or hypertext form. However, if you
provide access to or distribute copies of a Project Gutenberg™ work
in a format other than “Plain Vanilla ASCII” or other format used in
the official version posted on the official Project Gutenberg™ website
(www.gutenberg.org), you must, at no additional cost, fee or
expense to the user, provide a copy, a means of exporting a copy, or
a means of obtaining a copy upon request, of the work in its original
“Plain Vanilla ASCII” or other form. Any alternate format must
include the full Project Gutenberg™ License as specified in
paragraph 1.E.1.
• You pay a royalty fee of 20% of the gross profits you derive
from the use of Project Gutenberg™ works calculated using the
method you already use to calculate your applicable taxes. The
fee is owed to the owner of the Project Gutenberg™ trademark,
but he has agreed to donate royalties under this paragraph to
the Project Gutenberg Literary Archive Foundation. Royalty
payments must be paid within 60 days following each date on
which you prepare (or are legally required to prepare) your
periodic tax returns. Royalty payments should be clearly marked
as such and sent to the Project Gutenberg Literary Archive
Foundation at the address specified in Section 4, “Information
about donations to the Project Gutenberg Literary Archive
Foundation.”
• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.
1.F.
1.F.4. Except for the limited right of replacement or refund set forth
in paragraph 1.F.3, this work is provided to you ‘AS-IS’, WITH NO
OTHER WARRANTIES OF ANY KIND, EXPRESS OR IMPLIED,
INCLUDING BUT NOT LIMITED TO WARRANTIES OF
MERCHANTABILITY OR FITNESS FOR ANY PURPOSE.
Please check the Project Gutenberg web pages for current donation
methods and addresses. Donations are accepted in a number of
other ways including checks, online payments and credit card
donations. To donate, please visit: www.gutenberg.org/donate.
Most people start at our website which has the main PG search
facility: www.gutenberg.org.
ebookbell.com