100% found this document useful (1 vote)
27 views90 pages

Domain Walls From Fundamental Properties To Nanotechnology Concepts Dennis Meier Jan Seidel Marty Gregg Ramamoorthy Ramesh Instant Download

The document discusses the book 'Domain Walls: From Fundamental Properties to Nanotechnology Concepts' which explores the significance of ferroelectric domain walls in advancing nanotechnology and information technology. It highlights the unique properties of domain walls, such as their mobility and potential applications in devices like racetrack memory and domain wall logic devices. The book aims to provide a comprehensive overview of recent research and emerging applications in the field, emphasizing the transition of domain walls from scientific curiosity to technological relevance.

Uploaded by

dejahaua
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
100% found this document useful (1 vote)
27 views90 pages

Domain Walls From Fundamental Properties To Nanotechnology Concepts Dennis Meier Jan Seidel Marty Gregg Ramamoorthy Ramesh Instant Download

The document discusses the book 'Domain Walls: From Fundamental Properties to Nanotechnology Concepts' which explores the significance of ferroelectric domain walls in advancing nanotechnology and information technology. It highlights the unique properties of domain walls, such as their mobility and potential applications in devices like racetrack memory and domain wall logic devices. The book aims to provide a comprehensive overview of recent research and emerging applications in the field, emphasizing the transition of domain walls from scientific curiosity to technological relevance.

Uploaded by

dejahaua
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 90

Domain Walls From Fundamental Properties To

Nanotechnology Concepts Dennis Meier Jan Seidel


Marty Gregg Ramamoorthy Ramesh download

https://ebookbell.com/product/domain-walls-from-fundamental-
properties-to-nanotechnology-concepts-dennis-meier-jan-seidel-
marty-gregg-ramamoorthy-ramesh-22122314

Explore and download more ebooks at ebookbell.com


Here are some recommended products that we believe you will be
interested in. You can click the link to download.

Ferroelectric Domain Walls Statics Dynamics And Functionalities


Revealed By Atomic Force Microscopy 1st Edition Jill Guyonnet Auth

https://ebookbell.com/product/ferroelectric-domain-walls-statics-
dynamics-and-functionalities-revealed-by-atomic-force-microscopy-1st-
edition-jill-guyonnet-auth-4662864

Kinks And Domain Walls An Introduction To Classical And Quantum


Solitons 1st Edition Tanmay Vachaspati

https://ebookbell.com/product/kinks-and-domain-walls-an-introduction-
to-classical-and-quantum-solitons-1st-edition-tanmay-vachaspati-905548

Topological Structures In Ferroic Materials Domain Walls Vortices And


Skyrmions 1st Edition Jan Seidel Eds

https://ebookbell.com/product/topological-structures-in-ferroic-
materials-domain-walls-vortices-and-skyrmions-1st-edition-jan-seidel-
eds-5354744

Domainspecific Development With Visual Studio Dsl Tools Steve Cook

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

Bioenergy 1st Edition Demain Arnold L Wall Judy D Harwood

https://ebookbell.com/product/bioenergy-1st-edition-demain-arnold-l-
wall-judy-d-harwood-5307800

Domainspecific Conceptual Modeling Concepts Methods And Adoxx Tools


Dimitris Karagiannis

https://ebookbell.com/product/domainspecific-conceptual-modeling-
concepts-methods-and-adoxx-tools-dimitris-karagiannis-46346498

Domaindriven Design With Golang Use Golang To Create Simple


Maintainable Systems To Solve Complex Business Problems 1st Edition
Matthew Boyle

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

Domainspecific Languages Effective Modeling Automation And Reuse


Andrzej Wsowski

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

R. J. Nicholas University of Oxford


H. Kamimura University of Tokyo

Series on Semiconductor Science and Technology


1. M. Jaros: Physics and applications of semiconductor microstructures
2. V.N. Dobrovolsky and V. G. Litovchenko: Surface electronic transport phenomena in
semiconductors
3. M.J. Kelly: Low-dimensional semiconductors
4. P.K. Basu: Theory of optical processes in semiconductors
5. N. Balkan: Hot electrons in semiconductors
6. B. Gil: Group III nitride semiconductor compounds: physics and applications
7. M. Sugawara: Plasma etching
8. M. Balkanski, R.F. Wallis: Semiconductor physics and applications
9. B. Gil: Low-dimensional nitride semiconductors
10. L. Challis: Electron-phonon interactions in low-dimensional structures
11. V. Ustinov, A. Zhukov, A. Egorov, N. Maleev: Quantum dot lasers
12. H. Spieler: Semiconductor detector systems
13. S. Maekawa: Concepts in spin electronics
14. S. D. Ganichev, W. Prettl: Intense terahertz excitation of semiconductors
15. N. Miura: Physics of semiconductors in high magnetic fields
16. A.V. Kavokin, J. J. Baumberg, G. Malpuech, F. P. Laussy: Microcavities
17. S. Maekawa, S. O. Valenzuela, E. Saitoh, T. Kimura: Spin current
18. B. Gil: III-nitride semiconductors and their modern devices
19. A. Toropov, T. Shubina: Plasmonic Effects in metal-semiconductor nanostructures
20. B.K. Ridley: Hybrid phonons in nanostructures
21. A.V. Kavokin, J. J. Baumberg, G. Malpuech, F. P. Laussy: Microcavities, Second edition
22. S. Maekawa, S. O. Valenzuela, E. Saitoh, T. Kimura: Spin current, Second Edition
23. M. M. Glazov: Electron and nuclear spin dynamics in semiconductor nanostructures
24. D. Meier, J. Seidel, M. Gregg, R. Ramesh: Domain walls: from fundamental properties to
nanotechnology concepts
Domain Walls: From Fundamental
Properties to Nanotechnology Concepts

Dennis Meier, Jan Seidel, Marty Gregg, and


Ramamoorthy Ramesh

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

G. Catalan and N. Domingo


Catalan Institute of Nanoscience and Nanotechnology (ICN2), CSIC and BIST, Campus
UAB, Bellaterra 08193, Barcelona, Catalonia.
ICREA-Institucio Catalana de Recerca i Estudis Avançats, Barcelona, Catalonia.

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

(a) ((b) (c) (d)

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.

The field of domain wall nanoelectronics (Catalan et al. 2012a) is predicated on


the premise that the distinct physical properties of domain walls offer new conceptual
possibilities for devices. The first part of this chapter will deal with the basic physics of
domain wall properties, and in particular the cross-coupling that allows domain walls to
display properties and order parameters different from those of the parent bulk material.
The second part will deal with scanning probe techniques for measuring some of these
domain wall properties, and specifically atomic force microscopy (AFM). Together with
transmission electron microscopy, discussed in Chapter 10, AFM is one of the most
important tools we currently have to probe and manipulate the individual position and
physical properties of domain walls; although this book contains many chapters that
discuss AFM probes, e.g. to inject domain walls and control their motion (Chapter
13), here we will focus on two recent developments that allow investigating hitherto
overlooked properties of domain walls: their magnetotransport and their mechanical
response.

1.2 Domain Wall Structure and Thickness


1.2.1 Domain Wall Thickness
Domain walls are not the ground state of any material (with the possible exceptions
of incommensurate materials and relaxors), so they cost energy. The minimization of
this energy cost dictates their thickness and structure. First, there is a nearest neighbor
interaction; in ferromagnets, this is called “exchange energy,” but the concept applies
to any ferroic. The exchange energy penalizes differences between adjacent unit cells.
In other words: it penalizes gradients of the order parameter. The associated energy cost
of the gradient of the order parameter, Ugrad , for a generic ferroic with an order parameter
 = (x) (which can be polarization, or spontaneous strain, or magnetization) is
 
k ∂ 2
Ugrad = (1.1)
2 ∂x
Physical Properties inside Domain Walls 3

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

and the correlation length λ


 
−1 2k 2k
λ = 20 = . (1.6)
b −a

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

the domain wall thickness becomes



2k 
δ=2 = 4 χk. (1.8)
−a

An intuitive simplification that brings in quantitatively similar results consists in replacing


the hyperbolic tangent by a linear polarization profile across the wall (Catalan et al.
2012a):

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:

Uquadratic = 1 2 χ −1 (x)2 . (1.11)

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

Minimizing this energy with respect to the domain wall thickness,

∂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)

All in all, the aforementioned equations tell us that:


Physical Properties inside Domain Walls 5

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.

1.2.2 Internal Symmetry of Domain Walls


The internal symmetry of domain walls can be deduced by group theory analysis (Fousek
and Janovec 1969; Fousek 1971), by thermodynamic/Landau theory (Marton et al.
2010), or by first-principles calculations (Padilla et al. 1996; Lubk et al. 2009; Dieguez
et al. 2013) (see also Chapters 3 and 4, respectively). Symmetry-based analyses, though
quantitatively moot, are powerful for predicting the structurally allowed properties of
the domain walls, and in this respect the seminal works of Privratska and Janovec
(1997, 1999) should be credited as the first to point out the possibility of emergent
magnetization being present inside the ferroelectric domain walls of magnetoelectric
multiferroic materials. An update to those studies has been provided by Privratska
(2007). Quantitative analyses of domain wall structure use the free energy as the starting
point, and impose boundary conditions to obtain the domain wall profile and properties
(Mitsui and Furuichi 1953; Zhirnov 1959). Daraktchiev et al. (2010) and Marton et al.
(2010) solve the problem of calculating local symmetry inside the domain wall by first
calculating the free energy of the bulk ferroic state, and then calculating the least energy
path connecting ferroic minima, i.e. the potential wells of the free energy surfaces. The
trajectory of this connecting path in phase space defines the state of the domain wall,
and the symmetry in the middle of this path defines the state at the center of the domain
wall. By way of illustration, the trajectory that connects the <100> minimum with the
6 Domain Walls: From Fundamental Properties to Nanotechnology Concepts

0.4 0.4
(a) (b)

Pt(C m–2) 0.2 0.2

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.

<010> minimum in a tetragonal ferroelectric such as BaTiO3 corresponds to a 90◦


domain wall, and it is straightforward to see that this trajectory passes through a <110>
point. In the center of the domain wall, then, the local polar symmetry of the material
must be orthorhombic or monoclinic (polarization along <110> instead of tetragonal).
Figure 1.2 provides equivalent examples for domain walls in the orthorhombic and
rhombohedral phases of BaTiO3 , extracted from the work of Marton et al. (2010, 2011).
Another physically intuitive approach is to treat the domain wall as a slab of
paraelectric phase epitaxially sandwiched between the two opposite domains on either
side (Catalan 2012b). Let us take, for example, a tetragonal ferroelectric such as BaTiO3 ,
polarized along z and with a 180o domain wall in the y-z plane with thickness parallel to
the x direction. In this case, the spontaneous strain in the domains is

sx = sy = Q12 P02 , sz = Q11 P02 . (1.15)

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.

1.3 Order Parameter Coupling


The previous analysis shows how the suppression of the main order parameter (e.g.
polarization) inside the wall affects its properties and other parameters. Let us dwell
further on this concept with respect to multiferroic (or quasi-multiferroic) materials.
The phenomenological treatment of this problem starts with a Landau thermodynamic
potential with two coupled order parameters (polarization P and magnetization M for
magnetoelectrics) (Daraktchiev et al. 2008, 2010). Depending on the symmetry of
the material, there can be lower-order couplings, but one that is universally allowed is
biquadratic. In this case, the free energy is given by

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

magnetic systems by Lajzerowicz and Niez (1970) and is supported by the


first-principles calculations of Wojdel and Iñiguez (2014) and the thermodynamic
calculations of Stepkova et al. (2012).

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).

1.4 Physical Properties of Domain Walls


1.4.1 Measuring Physical Properties at Domain Walls by AFM
If domain walls are predicted to have intrinsic physical properties distinguishing them
from the contiguous domains, the experimental challenge is to measure them inde-
pendently: they are too thin to be accessible by optical microscopy, and their volume
fraction is too small for bulk spectroscopic techniques. Transmission electron microscopy
is a suitable instrument for measuring the domain wall structure (Bursill et al. 1983;
Bursill and Peng 1986; Floquet et al. 1997; Foeth et al. 1999; Jia et al. 2008) and
chemical composition with atomic scale precision (Chapter 10), and even monitor
the real-time dynamics of domain switching (Nelson et al. 2011); however, sample
preparation is delicate and destructive. By contrast, AFM has emerged as a mainstream
nanotool playing the role of an authentic nanolaboratory to characterize domain walls
morphologically and functionally in a non-destructive manner, as well as to move them
by applying localized electric fields or mechanical stresses (Lu et al. 2012).
The proper combination of cantilever- and tip-designs operated in appropriate modes
enables the detection of forces below the piconewton and permits the characterization of
physical and chemical properties. The tip of an atomic force microscope can be used as a
top mobile electrode for polarization and transport measurements; a magnetic sensor for
magnetostatic measurements; and a mechanical deformation sensor, able to operate from
the low interacting force regime to a high indentation regime, to examine mechanical
properties.
Given the versatility of AFM, it is also possible to combine several modes simulta-
neously or sequentially to investigate the coupling between different order parameters.
Going beyond its multifunctional characterization options with nanoscale resolution,
another notable distinguishing feature of AFM is its ability to manipulate physical
magnitudes to deliver voltage, stress, magnetic fields, or heat to the domain wall to
make it react. Because of these possibilities, AFM has become a unique tool to visualize,
measure, and move domain walls at will, achieving nanometric lateral resolution (Shilo
et al. 2004) and sub-microsecond dynamic sensitivity via pump-probe techniques
(Gruverman et al. 2011).
AFM-based techniques that can be applied, alone or in combination (see Figure1.5)
to study domain walls in ferroic materials are, for example, piezoresponse force mi-
croscopy (PFM), which allows the ferroelectric polarization to be measured; magnetic
10 Domain Walls: From Fundamental Properties to Nanotechnology Concepts

(a) Topo (b) PFM


300 K

3 μm
10 nm

(c) MFM (d)


5.5 K

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

(a) 500 (b) 500

400 pm 400 Deg


100
200
300 80 300
150
nm

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.

1.4.2 Charge Transport


The switching of polarization across a ferroelectric domain wall can either happen
through progressive change of the magnitude of the polarization (Ising walls), or through
rotation—either within the plane of the wall (Bloch walls) or perpendicular to it (Néel
walls) (Lee et al. 2009), as depicted in Figure 1.8.
Ising walls and Bloch walls are electrostatically neutral, but Néel walls are not,
because they present a discontinuity of the displacement field perpendicular to the wall.
According to Poisson’s equation,∇D = ρf , where ρ f is the free charge density, and D is
the displacement field, which in a ferroelectric is D = ε0 E + P. Any Néel-type wall in a
ferroelectric must therefore accrue a free charge density that is equal to the polarization
discontinuity perpendicular to the wall plane.
Poisson’s free charge will also accumulate at the walls when there is a head-to-head
or tail-to-tail component of the polarization, as this also implies a discontinuity of the
displacement field (Xiao et al. 2005; Gureev et al. 2011). Such head-to-head or tail-to-tail
configurations may appear spontaneously in so-called “improper” ferroelectrics (where
the ferroelectric order parameter is a “slave” to a dominant primary order parameter—
Choi et al. 2010; Meier et al. 2012) (see Chapter 6 for details). Charged domain walls
can also be forced to appear even in proper ferroelectrics, such as BaTiO3 , by pitting the
strain order parameter against the polar order parameter (frustrated poling; Sluka et al.
2013). The trick in either case is to force the appearance of electrostatically unfavorable
charged domain walls by acting on a structural order parameter, such as strain, that is
coupled to the polarization.
Increased charge density, however, does not imply increased conductivity. Fer-
roelectrics are semiconductors. For n-type semiconductors, electrons increase
conductivity, while holes neutralize the majority carriers, thus decreasing conductivity.
Accordingly, for n-type ferroelectrics, head-to-head domain walls will be more
Physical Properties inside Domain Walls 13

(a) (b) (c)


θN
θB

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

Figure 1.9 Upward out-of-plane written ferroelectric square domain on an as-grown


downward-polarized BiFeO3 (111) thin film. The domain was written by applying a negative voltage
with an AFM tip. The polarization structure is obtained from (a) the amplitude and (b) the phase PFM
images. The bottom images from (c) to (f) show the evolution of the domain wall conductivity over
different sequential scans. Note that with each consecutive scan, the conductive region becomes broader
and more “diffuse,” indicative of the sideways diffusion of the conductive defects (likely oxygen vacancies).

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

necessarily imply increased domain wall conductivity, as the reverse—increased domain


wall resistivity—has also been observed (Bartels et al. 2003). Likewise, it is also worth
mentioning that domain wall decoration is sometimes intentional, but it can also happen
spontaneously. In oxides, domain walls are known to attract oxygen vacancies (He and
Vanderbilt 2003), and these will act as charge donors inside the wall (Farokhipoor and
Noheda 2011).
The coupling of ferroelasticity and strain gradients at the domain walls can also
impact their conductivity properties. Strain at the domain walls in strongly corre-
lated electron oxides can deform the octahedral tilting in ABO3 perovskite structures
(B-O-B bond angle distortion) and bond lengths (B-O bond length distortion), changing
the orbital overlapping. Since orbital overlapping is the structural mechanism governing
the bandwidth of these materials, this is observed to enhance electrical conductivity, as
observed in Figure 1.10.

1.4.3 Magnetism at Domain Walls


The strain contained in ferroelastic domain walls affects the orbital overlapping in
ferroelastic magnetic materials and can enhance the magnetic moment at the domain
walls, as observed in LaSrMnO3 and shown in Figure 1.10 (Balcells et al. 2015). In
this case, the magnetic tip of an MFM can be used to monitor specifically the magnetic
moment of the domain walls, which tends to orient following the magnetic field signal
emitted by the MFM tip. This leads to the observation of magnetic domain walls as bright
lines in MFM images when the tip is scanned sufficiently close to the surface.
The coupling of magnetism and ferroelectricity in multiferroics is a promising source
of magnetoelectric coupling that can be exploited to manipulate spins with electric fields.
Most multiferroics are antiferromagnets with a vanishing magnetic moment, but the
symmetry breaking at the domain walls can lead to confined magnetic moments in these
topological structures. For example, spin rotation across a magnetic Néel wall can induce,
via Dzyaloshinskii–Moriya interaction, an electric polarization (Logginov et al. 2008).
Also, strain gradients can cause polarization via flexoelectricity (Zubko et al. 2013), and
in turn this polarization may cause spin canting via magnetoelectricity.
An example of magnetization emerging inside the domain walls of an antiferromag-
netic multiferroic is hexagonal ErMnO3 , an improper ferroelectric showing a collective
domain wall magnetism over its vortex structure (Geng et al. 2012). Low-temperature
MFM images shown in Figure 1.5 are able to map the net magnetization at the
domain walls.

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

(a) (c) (e)

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).

(a) (b) (c)


145° 10 pA

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

10,0 HDW1 10,0 HDW2


I (pA)

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

Transport measurements of domain walls can be achieved by different methods.


The simplest way is to place microelectrodes on top of a crystal and monitor the
conductivity while controlling the number and orientation of the domain walls between
the electrodes. This measurement leads to an average of the transport properties over a
certain distribution of domains. In contrast, C-AFM can measure the local transport
properties at the nanoscale, and enjoys an extra degree of freedom to play with the
geometrical symmetry of the transport measurement. Magnetotransport measurements
at the nanoscale can be achieved by a combination of C-AFM and the in situ application
of an external magnetic field. This allows to capture subtle coupling effects that might
otherwise get cancelled when averaged over microscopic dimensions. In the example
shown in Figure 1.11, this allows distinguishing the local effect of the orientation of
each domain wall with respect the applied magnetic field; this, it turns out, is important,
because the sign of the magnetoresistance is orientation-dependent, and it averages out
to zero in macroscopic measurements containing many randomly oriented walls.

1.4.5. Mechanical Response


The mechanical response of domain walls can also be locally probed by scanning force
microscopy. Among all the available operation modes, a good choice to accomplish this
is CRF measurements based on tracking the mechanical resonance frequency of the
cantilever in contact with the sample. By applying an oscillatory mechanical force to the
cantilever and monitoring the resonance frequency, we can detect relative differences
in stiffness between the domain and the domain wall. Our measurements for different
ferroelectrics (Stefani et al. 2020) consistently show that the resonance frequency of 180◦
domain walls is slower (and thus the domain wall is softer) than the domains themselves
(Figure 1.12). This contrast in mechanical response may seem surprising, given that
there is no mechanical difference between 180◦ domains themselves. (Note that there
is also a weak CRF contrast between domains due to flexoelectric coupling (Cordero-
Edwards et al. 2017))

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.

1.5 Summary and Conclusions


Domain walls behave as embedded layers whose properties can be understood as those
of a thin film of paraphase sandwiched coherently between the domains. Thus, the
effect of cancelling the order parameter can be modeled by standard thermodynamic
arguments. Perhaps the key message from the theoretical section is that the switching off
of the order parameter has consequences not only in terms of new symmetry-allowed
(or symmetry-forbidden) properties, but also it has a knock-on effect on secondary
parameters that might be coupled to or suppressed by the primary ordering. Thus, for
example, the suppression of polarization means that the inside of ferroelectric domain
walls is not piezoelectric, and this implies meaning that straining them will not generate
any polarization—so their mechanical response will be different (softer) than that of the
domains. In magnetoelectrics, the emergence or suppression of ferroelectric polarization
inside a domain wall also has consequences for the wall’s magnetization. In turn, changes
in magnetization inside the wall can affect resistivity via magnetoresistance.
The resistivity itself is also affected by the manner in which polarization switches
from one side of the wall to the other; hence, head-to-head polarization or tail-to-
tail polarization (which, despite being energetically costly, can happen in multiferroic
domain walls where the polarization is a “slave” to a primary order parameter) must
result in charge accumulation or depletion at the wall, and thus, in dramatically modified
electronic properties. Though we have not covered them here, the change in electronic
properties will also affect related properties such as the photovoltaic response at the wall
(see Chapter 9). The bottom line is that the panoply of properties that domain walls
can display is as varied—or more—as that of the multifunctional materials where they
reside, but this complexity of behavior can be understood with simple physics: treat the
wall as a very thin film, and write the thermodynamic potential including gradient and
coupling terms.
The rich variety of domain wall behaviors is matched by the versatility of AFM,
arguably the most important tool in any domain wall research laboratory. The possibility
of using functionalized cantilevers in different measurement modes and in different
in situ environments allows measuring—and manipulating—domain wall responses to
different stimuli.
Physical Properties inside Domain Walls 19

The measurement possibilities of AFM are as varied as the functionalities displayed


by domain walls, but perhaps the best-known and most-used one is PFM, which allows
identifying the exact location of domain walls, and/or creating them by delivering a high
voltage capable of switching a ferroelectric domain. PFM is by now a well-established
technique that is described in more detail in other chapters, so in this chapter we
have described in more detail a couple of more innovative measurement modes that
allow, respectively, the measurement of domain wall magnetoresistance and domain
wall mechanical properties. These are just two examples that illustrate the experimental
possibilities offered by the creative combination of domain walls and alternative scanning
probe microscopy modes. We hope that the readers will be inspired to try their own ideas.
It is a fun game.

..........................................................................................
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,

Departamento de Física de la Materia Condensada, Pedro Cerbuna 12, 50009 Zaragoza,


Spain
3 Instituto de Nanociencia y Nanotecnología, Comisión Nacional de Energía Atómica and

Consejo Nacional de Investigaciones Científicas y Técnicas, Gral. Paz 1499, San Martín,
Argentina

In addition to the distinct internal symmetry of domain walls discussed in Chapter 1,


local stresses can induce new physical properties at domain walls. In this chapter, we
discuss how the intense stress fields generated at the ferroelastic domain walls in some
epitaxially strained oxide layers introduce selective chemical modifications, giving rise
to novel 2D crystal structures at the walls. In the case of the strained TbMnO3 films
presented here, the substitution of Tb by Mn creates a net magnetic moment at each
domain wall, thus making them distinct from the bulk-like antiferromagnetic domains.
The possibility of tuning the domain wall density with the film thickness makes it possible
to tailor the material’s response at the nanoscale.

2.1 Introduction to TbMnO3


The perovskite structure, with the general formula ABO3 , has cubic symmetry with
the space group Pm3m. Oxygen anions are arranged in a corner-shared, face-centered
fashion, forming a connected net of octahedral cages throughout the crystal structure in
three dimensions. The smaller-size cation (B), in 6-fold coordination, is placed inside the
octahedral cage, whereas A is a 12-fold coordinated cation located in the voids between
the octahedra. At room temperature, most of the interesting perovskite-like materials
are slightly distorted with respect to the structure described above, with one or both
cations and/or the oxygens shifted along particular directions, lowering the symmetry.
Among perovskites, manganites present a rich physics with interesting magnetic and

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

Rochelle salt (ferroelectric)


107 Rochelle salt (ferroelectric)
106 Co (ferromagnetic)
PbTiO3 (ferroelectric)
105 PbTiO3 (ferroelectric)
w2/δ(nm)

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

a universal behavior, independent of the physics of the interactions, is found (Catalan


et al. 2009; Catalan et al. 2012).
In a similar manner, for ferroelastic domains, the domain width (w) is determined by
the competition between the elastic energy in the domains and the formation energy of
the domain walls and is a function of the thin film thickness (d). For the cases when
d > w, the Roitburd scaling law applies (Roitburd 1976). Interestingly, this square root
dependence holds for both epitaxial and freestanding layers. In the very thin film limit
(for d < w), Roitburd’s approximations are not valid, and a rigorous calculation of the
elastic energy of the system leads to a nearly linear relation between the domain width
and the film thickness (Pompe et al. 1993; Pertsev and Zembilgotov 1995). Thus, above
the critical thickness for domain formation, the thinner the film, the larger the number
of domain walls and the smaller the size of the domains.
The presence of a net magnetic moment has been detected in several antiferromag-
netic manganites, such as TbMnO3 , YbMnO3 and YMnO3 , when grown in thin film
form (Marti et al. 2008, 2009, 2010; Rubi et al. 2008, 2009). It has been proposed that
the ferromagnetism originates in the strain-induced modification of the balance between
the different magnetic exchange interactions (Dong et al. 2009), which produces a
canting of Mn spins (Marti et al. 2008, 2009, 2010). Interestingly, the magnetic moment
seems to be induced independently of the ground state magnetic structure of the bulk
material: TbMnO3 displays a cycloidal spin structure, whereas YbMnO3 and YMnO3
have an E-type collinear spin structure. In both cases, the induced magnetization has
been reported to follow a similar trend as the unit cell volume (Marti et al. 2010).
This raises the question of whether the magnetic moment arises from strain effects
linked to the modification of the bond lengths and angles, as proposed in the first
instance, or whether it originates from a more general feature of epitaxial orthorhombic
manganites (Fontcuberta 2015), such as the domain microstructure (Dong et al. 2009;
Fontcuberta 2015).
In epitaxial orthorhombic TbMnO3 , the origin of the macroscopic net magnetic
moment was reported to be the result of uncompensated Mn3+ spins upon the formation
of a particular type of domain wall (Farokhipoor et al. 2014). Other examples of
confinement of the magnetic moment in 2D can be found, such as the interface
between LaAlO3 and SrTiO3 (Brinkman et al. 2007), or the ferromagnetically coupled
interface between CaMnO3 and CaRuO3 , due to the competition between antifer-
romagnetic superexchange and ferromagnetic double exchange between Mn4+ and
Ru4+ . In none of these cases are the oxides ferromagnetic as single layer systems
(Takahashi et al. 2001). The existence of a magnetic moment in a two-dimensional
non-magnetic organic system (graphene) has also been reported (Ma et al. 2012;
Yang et al. 2013).
In the following, we discuss the effects of epitaxy and the presence of domain walls
on the magnetic properties of thin films of TbMnO3 under a large strain induced by
epitaxy on SrTiO3 substrates. The presence of intense local stresses leads to selective
atomic substitution, giving rise to the formation of a unique phase, which is responsible
for the distinct magnetic properties at the domain walls of TbMnO3 thin films, as
mentioned above.
28 Domain Walls: From Fundamental Properties to Nanotechnology Concepts

2.2 Formation of a Novel Phase at the Domain Walls


of TbMnO3
It has been shown that orthorhombic manganites grown on cubic SrTiO3 substrates
form crystallographic twins (Marti et al. 2008). Twinning is mainly determined by the
symmetry relationships between the film and substrate materials, but it is affected by
the growth kinetics and can differ depending on the growth conditions. TbMnO3 thin
films grown on (001)-SrTiO3 with low deposition rates (approaching thermodynamic
conditions) are reported to form four types of twin domains, all with the c-axis
perpendicular to the substrate interface (Daumont et al. 2009). Despite the very large
mismatch strain (+4.1% along [100]o and −5.7% along [010]o ), this microstructure
allows the film to maintain partial coherence with the substrate, either along the [100]c
or the [010]c (substrate) directions and, importantly, it determines the evolution of the
lattice parameters with increasing thickness (Daumont et al. 2009): the partial coherence
with the substrate and the crystal twinning are able to maintain the unit cell in-plane
area constant. Thus, the out-of-plane lattice parameter and the unit cell volume remain
basically unchanged for a large range of thicknesses, from 5 to 70 nm (Daumont et al.
2009). This domain/twin configuration prevents the accumulation of large shear stress
with increasing thickness and allows slow relaxation from a fully coherent lattice toward
the bulk orthorhombic structure by modifying ao and bo in opposite directions and by
the same amounts. It can be rationalized that this is an efficient way of minimizing the
elastic energy of the system in the presence of shear stress (Venkatesan et al. 2009).
This phenomenon is very common as it takes place in the case of heteroepitaxy between
cubic and orthorhombic lattices, the latter consisting of non-orthogonal pseudocubic
units, but also between cubic and rhombohedral structures, where similar effects have
been observed (Daumont et al. 2010).
The direct observation of the properties of these domain walls below the ordering
temperature (~41 K) is very challenging due to their atomic-scale width. This is compli-
cated by the fact that TbMnO3 thin films are semiconducting, hampering nanoscale
magnetic imaging using scanning tunneling microscopy (Bode et al. 2006). Instead,
a combination of transmission electron microscopy (TEM), magnetometry, and first-
principles calculations (see Chapter 3) was shown to be decisive for understanding the
formation of two-dimensional magnetic sheets at the structural domain walls, as observed
in the sketch shown in Figure 2.3.
In this way, the inverse dependence of the remanent magnetic moment with the film
thickness (for thicknesses between 5 and 80 nm) could be explained. The evolution
of the density of domain walls as a function of the thickness, evidenced by TEM, was
consistent with the increased net magnetic moment for decreasing thickness; see Figure
2.4 (Daumont et al. 2009; Venkatesan et al. 2009). This phenomenon is attributed to the
presence of chemical reconstruction at the structural domain walls. These domain walls
are (100)o or (010)o planes through the A-cations, separating orthorhombic domains
with opposite “zigzags,” as shown in Figure 2.3. Interestingly, quantitative electron
energy loss spectroscopy in scanning transmission electron microscopy mode (STEM-
Novel Phases at Domain Walls 29

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)

Inverse domain area


1/Sdomain (10–2 nm–2)

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.

2.3 The Oxygen Stoichiometry of Off-Wall and On-Wall


Areas in TbMnO3 Thin Films Grown on SrTiO3 Substrate
The chemical substitution at the domain walls is likely to affect the oxygen lattice, to
accommodate to the different ionic radii and valence of the metal cations. Oxygen
stoichiometry of the domain walls in TbMnO3 can be explored by STEM-EELS,
even with atomic resolution (Pennycook and Nellist 2011). Individual spectra of the
specific EELS edge for a chemical element can be obtained by conventional background
subtraction. Subsequent edge intensity integration and quantification in terms of the
Novel Phases at Domain Walls 31

(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

(a) (b) (c) (d) (e)


1.4
1.3 O-K 3.5 Mn-L2,3
1.2 1 eV

Intensity (counts × 104)

Intensity (counts × 104)


1.1 3.0
1.0 2.5
0.9
0.8 2.0
0.7
0.6 1.5
0.5 1.0
0.4
0.3 0.5
0.2 Top Top
Defect 0.0 Defect
0.1 Bottom Bottom
0.0 –0.5
530 540 550 640 650 660
+2.5 +3 +3.5 11.0 11.4 11.8
Energy Loss (eV) Energy Loss (eV)
Mn Valence Energy (eV)

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,

L-4422 Belvaux, Luxembourg

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) (c)

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.

3.2 Basic Background


For the benefit of newcomers to the field, and to make this chapter reasonably
self-contained, let me start by addressing some basic issues and terminology pertaining
to structural walls and their physics. For the sake of concreteness, I consider the case of
perovskite oxides, taking PbTiO3 and SrTiO3 as representative examples, noting that
most of the ideas presented here are directly applicable to other materials families.

3.2.1 Structural Instabilities, Order Parameters,


and Symmetry Breakings
Figure 3.1 shows sketches of the ABO3 perovskite structure, and how the condensation
of different structural instabilities breaks the symmetry of the prototype cubic phase to
First-Principles Studies of Structural Domain Walls 39

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

phase as a convenient high-symmetry reference to understand the ferroelectric domains


occurring in the material (see Chapter 10 for details on the domains and domain walls in
BiFeO3 ). Hence, as regards the analysis of structural domains and walls, our symmetry
arguments are a useful theoretical tool to interpret and predict behaviors, even if some of
the ingredients (e.g. the high-symmetry reference) may be hypothetical in some cases.

3.2.2 The Role of Strain: Ferroelastic Features


As emphasized in Chapter 5, multi-domain structures may result from elastic boundary
conditions acting on our materials (Salje 1993). Further, we know that strain effects are
very important in perovskite oxides, “strain engineering” of thin films being one of the
most powerful strategies to tune their properties (Schlom et al. 2007). In fact, many of
the materials featured in this chapter (e.g. SrTiO3 and CaTiO3 ) are often described as
being “ferroelastic.” Thus, it may seem surprising that I have not included strain among
the key structural variables in the previous section. This point warrants a clarification.
As regards domain walls, most of the materials of interest today—in particular, all
compounds discussed in this chapter—are not proper ferroelastics. In none of these
materials is strain the primary order parameter; further, strain by itself cannot drive a
structural phase transition in any of them. (This statement relies on first-principles results
showing that the high-symmetry reference structure is stable against pure homogeneous
strains (King-Smith and Vanderbilt 1994; Chen et al. 2018b).) Rather, strain is a
secondary distortion that follows the primary one. More precisely, in the ferroelectric (like
PbTiO3 ) and antiferrodistortive (like SrTiO3 ) materials of interest here, the relationship
between the symmetric strains {ηαβ } and the P or R vectors has a simple analytical form.
As shown by several authors (King-Smith and Vanderbilt 1994; Gu et al. 2012; Chen
et al. 2018b), we have
     
ηxx = CQx2 + C  Qy2 + Qz2 , ηyy = CQy2 + C  Qz2 + Qx2 , ηzz = CQz2 + C  Qx2 + Qy2 ,
(3.1)

ηxy = C  Qx Qy , ηyz = C  Qy Qz , ηzx = C  Qz Qx , (3.2)

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

a change in strain: we have ηxx I = CP 2 → ηII = C  P 2 and ηI = C  P 2 → ηII = CP 2 .


xx yy yy
Hence, this wall is ferroelastic.
A word of caution is in order here. We often refer to PbTiO3 as a ferroelectric, and to
SrTiO3 as a ferroelastic, which seems to imply there is a fundamental difference between
the two compounds in regard to their ferroelastic properties. As should be clear from the
discussion above, this is not correct. PbTiO3 and SrTiO3 present essentially equivalent
improper-ferroelastic behavior.
Finally, ferroelastic walls have an interesting property: as illustrated in Figure 3.2(b),
they are polar by construction. Indeed, the symmetry breaking caused by the strain
discontinuity at a ferroelastic wall necessarily yields a polar distortion. Since ferroelastic
walls are obviously characterized by a strain gradient, it is natural to think of this effect
as a flexoelectric one; indeed, the magnitude of the improper polarization occurring at
ferroelastic walls will depend on the flexoelectric tensor of the material, which quantifies
the polarization appearing in response to a strain gradient (Tagantsev and Yudin 2016).
Yet, the emphasis on flexoelectricity may be misleading, as it seems to suggest that a
polarization will occur in any wall having strain gradients associated with it, which is
not correct. To understand this, first note that any wall (ferroelastic or not) associated
with any order parameter presents strain gradients. Indeed, the local (isotropic) strain is
always coupled to the (square of the) local order parameter; the volume of a cell at the
wall, where Q = 0, cannot be exactly the same as the volume of a cell within a domain,
where Q = 0, and thus, we always have strain gradients. The key is to realize that, in
non-ferroelastic walls, such wall-related strain gradients integrate to zero, yielding no net
electric polarization (instead, one has some sort of antipolar distortions originating from
the corresponding flexoelectric couplings). What is specific to ferroelastic walls is that
this integral is not zero, and a net polarization does appear.

3.2.3 Electric and Elastic Compatibility Conditions


Let us consider again the ferroelectric domains characterized by PI = (P, 0, 0) and
PII = (0, −P, 0), respectively. From the formulas above, we can deduce for the first
domain ηxxI = CP 2 and ηI = ηI = C  P 2 , while for the second one ηII = ηII = C  P 2 and
yy zz xx zz
ηyy = CP . Thus, the wall separating these domains is both ferroelectric and ferroelastic.
II 2

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

Domain I (+η) Domain II (-η)

(c) (d)

P –P
–P P –P

Ising wall Nèel wall


(1) (2)

Neutral wall Charged walls


ΔP ∙ n̂ = 0 |ΔP ∙ n̂| = 2P
(1) tail-to-tail (2) head-to-head Bloch wall

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.

be charged. A finite domain-wall charge might lead to a redistribution of mobile carriers


and, hence, anomalous electronic transport as discussed, for example, in Chapter 6.
Additionally, in order to minimize the elastic energy, ferroelastic walls will tend to
occupy planes that are similarly strained by both domains (Streiffer et al. 1998; Tagantsev
et al. 2010). To understand this better, let us consider again the domains defined by
PI = (P, 0, 0) and PII = (0, −P, 0). These domains cause symmetry-equivalent, but
different, spontaneous strains, and the dissimilarity can be captured by a difference strain
tensor given by

⎞⎛
 −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.)

3.3 Case Studies of Ideal Domain Walls


In the following, I review a selection of representative works on structural walls in
prototypic ferroic materials. I emphasize major milestones and discoveries, but do not
attempt a detailed recollection of the works. The reader is strongly encouraged to cosult
the original articles for more details.

3.3.1 Ferroelectric Materials: Perovskites PbTiO3 and BaTiO3


In my view, a review on this topic has to start with PbTiO3 , because this was the first
material studied, remains the best-studied one and, as we will see in Section 3.5, its walls
attract great attention today.
In 2002, Meyer and Vanderbilt published the article that set the standard for first-
principles investigations of structural walls, tackling the ferroelectric boundaries of
PbTiO3 (Meyer and Vanderbilt 2002). These authors studied ideal (perfectly planar and
neutral) walls associated with polarization changes from PI = (0, 0, P) to PII = (0, 0, −P)
[180◦ wall within a (1, 1, 0) plane] and to PII = (0, P, 0) [90◦ wall within a (1, 1, 0) plane].
To do so, they made some clever choices of supercell (see Figure 3.3(a)) that allowed
them to study the corresponding multi-domain structures with a minimal computational
burden. (Finding such optimal supercells poses an excellent quiz for atomistic-simulation
beginners!) They found that the walls are centered in PbO planes, are very narrow, and
have an Ising character (see Figure 3.2(d)). They also computed activation energies for
44 Domain Walls: From Fundamental Properties to Nanotechnology Concepts

(a) (b) (c)


z
0.8

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.

Görgey's Memoirs of the Hungarian Campaign have been


confiscated, and forbidden throughout Austria. Exceptions, however,
are made in favor of individuals.

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!"

Another eminent and honorable name is added to the list of victims


to the present barbarian Government of France. M. Barthélemy St.
Hilaire has refused to take the oath of allegiance—and he will
accordingly be deprived of the chair which he has long filled with so
much ability at the Collège de France. The sacrifice which M. St.
Hilaire has made to principle is the more to be honored, since he has
no private fortune, and has reached a time of life when it is hard to
begin the world anew. But the loss of his well-earned means of
subsistence is, we know, a light evil in his eyes compared to the loss
of a sphere of activity which he regarded as eminently useful and
honorable, and which he had acquired by twenty-seven years of
laborious devotion to learning and philosophy.

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.

MR. JOHN BULL'S IDEAS ON THE


MUSQUITO QUESTION.

Young Ladies (both at once).—"Why, Mr. Bull! how terribly you


have been bitten by the Musquitoes!"
Mr. Bull (a fresh importation).—"I can't hunderstand 'ow it
'appened. I did hevery thing I could think of to keep them hoff.
I 'ad my window hopen and a light burning hall night in my
hapartment!"
STARVATION FOR THE DELICATE.
That exquisite young officer, Captain Gandaw, was reading a
newspaper, when his brilliant eye lighted on the following passage in
a letter which had been written to the journal by Mr. Mechi, on the
subject of "Irrigation."

"I may be thought rather speculative when I anticipate that


within a century from this period, the sewage from our cities
and towns will follow the lines of our lines of railway, in gigantic
arterial tubes, from which diverging veins will convey to the
eager and distant farmer the very essence of the meat and
bread which he once produced at so much cost."

"Fancy," remarked the gallant Captain, "the sewage of towns and


cities being the essence of owa bwead and meat—and of beeaw too,
of cawse, as beeaw is made from gwain! How vewy disgasting! Mr.
Mechi expects that his ideas will be thought wathaw speculative.—He
flatters himself. They will only be consida'd vewy dawty. The wetch!
I shall be obliged to abjaw bwead, and confine myself to Iwish
potatoes—which are the simple productions of the awth—and avoid
all animal food but game and fish. And when fish and game are not
in season, I shall be unda the necessity of westwicting my appetite
to

"A scwip with hawbs and fwuits supplied,


And wataw fwom the spwing."
YOUNG NEW YORK HARD UP.

Tender Mother.—"A hundred Dollars! why, what can you want a


hundred dollars so soon for?"
Young New York.—"Why, Mother, I'm deucedly hard up. I'm
almost out of Cologne and Cigars. Besides, the fellows are going
to run me for President of the St. Nicholas Club, and I must
pony up my dues, and stand the Champagne."
A VICTIM OF THE TENDER PASSION.

Young Lady.—"Now then, what is it that you wish to say to me


that so nearly concerns your happiness?"
Enamored Juvenile.—"Why, I love you to the verge of distraction,
and can't be happy without you! Say, dearest, only say that you
will be mine!"
A STRIKING EXPRESSION.

Roguy.—"See that girl looking at me, Poguy?"


Poguy.—"Don't I? Why, she can't keep her eyes off you."
Roguy (poking Poguy in the waistcoat).—"What women care for,
my boy, isn't Features, but Expression!"
SCENE IN A FASHIONABLE LADIES'
GROGGERY.

Young Lady "couldn't take any thing—only a Pine-apple Ice"—but


the ice once broken, she makes such havoc upon pies, tongue,
Roman punches, tarts, Champagne, and sundry other potables
and comestibles, as to produce a very perceptible feeling in the
Funds.
RATHER A BAD LOOK-OUT.

Young Sister.—"Oh, Mamma! I wish I could go to a party."


Mamma.—"Don't be foolish. I've told you a hundred times that
you can not go out until Flora is married. So do not allude to the
subject again, I beg. It's utterly out of the question."
THE ATTENTIVE HUSBAND IN AUGUST.

Edward.—"There, Dearest, do you feel refreshed?"


Angelina.—"Yes, my Love. A little more upon the left cheek, if
you please. That's much nicer than fanning one's self. Now a
little higher, on my forehead."
Fashions for Summer.

Figures 1 and 2.—Bride's Toilet and Walking Dress.


Fig. 1.—Bride's Toilet.—Hair in bands very much puffed. Back hair
tied rather low; the wreath of white iris flowers, with foliage. Behind
this, and rather on one side, is the crown of orange flowers that
holds the vail, which is placed very backward, and is of plain tulle,
with a single hem. Dress of taffeta, with bayadères, or, rather, velvet,
with rows of velvet flowers, appearing like terry velvet. The body,
almost high behind, opens very low in front, and is trimmed with a
double plain berthe, that follows its cut. The waist is lengthened in
front, but not pointed. The bouquet decorates the bottom of the
body, and spreads in the form of a fan. The sleeve pagoda-shaped,
half-wide, and plain at top, terminated by two trimmings worked like
the edge of the berthes; a wide lace under-sleeve covers the arm.
The habit shirt is square at the top, composed of lace, the upper row
raised at the edge and four or five other rows below.
Fig. 2.—Walking Dress.—Bonnet of taffeta and blond. The brim, high,
narrow, and sitting close to the chin, is of taffeta, gathered from the
bottom of the crown to the edge; on the sides of the crown an
ornament is placed, cut rather round at the ends, and consisting of
three rows of taffeta bouillonnes, fastened together by a cross-piece
of taffeta. The crown is not deep, falls back, and has a soft top. The
curtain, of taffeta, cut cross-wise, is not gathered in the seam. The
blond that covers the lower part is gathered, and ends in vandykes
that hang below the curtain. A like blond is sewed full on the cross-
piece that borders the ornament, and the points also reaching
beyond the edge are fastened to those of the other blond, so that
the edge of the brim is seen through them. Toward the bottom the
blond above separates from that below, and sits full near the edge of
the ornament. A blond forming a fanchon on the calotte is laid also
under the other edge of the ornament. Lastly the curtain itself is
covered with blond. Inside are white roses, mixed with bows of
ribbon. Dress of taffeta. Body high, buttoning straight up in front.
Two trimmings are put up the side of the body. These trimmings,
made of bands resembling the narrow flounces, get narrower toward
the bottom. They are pinked at the edges, and shaded. The sleeve is
plain, and terminated by two trimmings, pinked and shaded. The
skirt has five flounces five inches wide, then a sixth of eight, pinked
and shaded.
Figure 3.—Bonnet.
Figure 5.—Bonnet.
Figure 4.—Bonnet.
Fig. 3.—Drawn Bonnet, of taffeta and blond; the brim, which is four
inches wide, is of taffeta doubled, that is, the inside and outside are
of one piece. It has several gathers. The side of crown, three inches
and a quarter wide, is of the same material, puffed at the sides for
about an inch, and there are also fourteen ribs in the whole circuit.
The top of crown is soft; a roll along the edge of the crown. The
ornaments consist of small rolls of taffeta, to which are sewed two
rows of blond three-quarters of an inch wide. These same rolls
ornament the brim, being placed on the edge, and inside as well as
outside. There are seventeen of these ornaments on the brim, with
an inch and a half of interval between them. The curtain is trimmed
in the same manner, and has ten of them. The top of crown has five
rolls, trimmed with blond. The inside is ornamented with roses,
brown foliage, and bouclettes of narrow blue ribbons mixing with the
flowers.
Fig. 4.—Drawn Bonnet of white tulle and straw-colored taffeta, edged
with a fringed guipure and bouquets of Parma violets. The taffeta
trimming is disposed inside and outside the brim, in vandykes, the
points of which are nearly three inches apart. In each space
between them is a bouquet of Parma violets. The points of the
fanchon lie upon the crown.
Fig. 5.—Drawn Bonnet, of tulle, blond, taffeta, and straw trimmings,
with flowers of straw and crape. The edge of the brim is cut in
fourteen scollops. The inside is puffed tulle, mixed with blond. The
scollops of the edge are continued all over the bonnet, and are
alternately tulle and white taffeta, with a straw edging.

For morning and home costume, organdie muslins will be in great


favor, the bodies made in the loose jacket style, and worn either with
lace or silk waist coats. Silks, with designs woven in them for each
part of the dress, are still worn; those woven with plaided stripe, à-
la robe, are very stylish.
White bodies will be worn with colored skirts they will be beautifully
embroidered, and will have a very distinguée appearance.
Dress bodies are worn open; they have lappets or small basquines:
for all light materials, such as organdie, tarlatane, barège, &c., the
skirts will have flounces. In striped and figured silks, the skirts are
generally preferred without trimming, as it destroys the effect and
beauty of the pattern. Black lace mantillas and shawls will receive
distinguished favor; those of Chantilly lace are very elegant. Scarf
mantelets are worn low on the shoulders.
A novelty in the form of summer mantelets has just been introduced
in Paris, where it has met with pre-eminent favor. It is called the
mantelet echarpe, or scarf mantelet; and it combines, as its name
implies, the effect of the scarf and mantelet. It may be made in
black or colored silk, and is frequently trimmed simply with braid or
embroidery. Sometimes the trimming consists of velvet or
passementerie, and sometimes of fringe and lace.
FOOTNOTES:
[1] Spelled variously, by different authors, Caïpha, Kaïfa, Caiffa,
and in other ways.
[2] The charts, as executed by the engineers, were on a still
larger scale than is here represented. It was necessary to reduce
the scale by one-fourth, in order to bring the portion to be copied
within the limits of a page.
[3] A striking example of this occurs at Long Branch in New
Jersey, where a stream crosses the beach in entering the sea, at a
point about half a mile to the southward of the hotels resorted to
on that coast in summer by bathers. The visitor who walks along
the shore in that direction, sometimes at a certain point finds
himself upon an elevated sandy ridge, with the surf of the sea
rolling in upon one side of it, and what appears to be a large
inland pond lying quietly on the other. A few days afterward, on
visiting the spot, he observes, perhaps, that the pond has
disappeared; and a wide chasm has been made across the ridge
of sand that he walked over before in safety, through the centre
of which a small stream is flowing quietly into the sea. Neither of
these views are of a nature to awaken any very special interest,
except when they are considered in connection with each other:
but if the observer should chance to come upon the ground when
the pond is nearly full, he may witness a very extraordinary
spectacle in the rushing out of the torrent by which the barrier is
carried away. The boys of the vicinity often find amusement in
hastening the catastrophe, by digging a little channel in the sand
with their hands, when the water has risen nearly to the proper
level. The stream that flows through this opening is at first
extremely small, but it grows wider, deeper, and more rapid every
moment, as the opening enlarges, and soon becomes a roaring
torrent, spreading to a great width, and tossing itself into surges
and crests as it rushes down the slope into the sea, in the most
wild and tumultuous manner.
The spectacle is almost equally imposing when, after the pond
has emptied itself, and the tide begins to rise, the surf of the sea
engages in its work of reconstructing the dam.
[4] It is somewhat doubtful whether the very first discovery of the
art of making glass, took place here or not, as learned men have
noticed a considerable number of allusions in various writings of a
very high antiquity, which they have thought might possibly refer
to this substance. An example of this kind is found in the book of
Job, where a word, translated crystal, is used. The writer,
speaking of wisdom, says, "It can not be equaled with the gold of
Ophir, with the precious onyx, or the sapphire. The gold and the
crystal can not equal it." It has been considered doubtful whether
the word crystal, in this connection, is meant to denote a glass or
some transparent mineral.
[5] See 1 Kings xviii. 17-46. For other passages of Scripture
referring to Mt. Carmel see 2 Kings ii. 25; iv. 25; xix. 23. 2 Chron.
xxvi. 10. Isa. xxxv. 2. Jer. xlvi. 18. Amos i. 2; ix. 3. Micah vii. 14.
[6] 1 Kings xviii. 4
[7] Continued from the July Number.
[8] Continued from the July Number.
[9] A gentleman, after hearing one of Mr. Clay's magnificent
performances in the Senate, thus describes him: "Every muscle of
the orator's face was at work. His whole body seemed agitated,
as if each part was instinct with a separate life; and his small
white hand, with its blue veins apparently distended almost to
bursting, moved gracefully, but with all the energy of rapid and
vehement gesture. The appearance of the speaker seemed that of
a pure intellect, wrought up to its mightiest energies, and brightly
shining through the thin and transparent vail of flesh that invested
it." It is much to be lamented that no painting exists of the
departed statesman that really does him justice. What a treasure
to the country, and to the friends of the "Great Commoner," would
be a portrait, at this time, from the faithful and glowing pencil of
our pre-eminent artist, Elliott! But it is now "too late".
[10] Nicholas Dean, Esq., President of the Croton Aqueduct Board,
a life-long friend of Mr. Clay.
[11] They were reduced to writing immediately afterward.
Transcriber's Notes:
Obvious printer's errors have been repaired, other inconsistent spellings have been kept,
including variation in:
- use of accent (e.g. "Léonard" and "Leonard" in p. 413-414);
- use of hyphen (e.g. "archway" and "arch-way");
- capitalisation (e.g. "Vice-president" and "Vice-President").

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.

Creating the works from print editions not protected by U.S.


copyright law means that no one owns a United States copyright in
these works, so the Foundation (and you!) can copy and distribute it
in the United States without permission and without paying
copyright royalties. Special rules, set forth in the General Terms of
Use part of this license, apply to copying and distributing Project
Gutenberg™ electronic works to protect the PROJECT GUTENBERG™
concept and trademark. Project Gutenberg is a registered trademark,
and may not be used if you charge for an eBook, except by following
the terms of the trademark license, including paying royalties for use
of the Project Gutenberg trademark. If you do not charge anything
for copies of this eBook, complying with the trademark license is
very easy. You may use this eBook for nearly any purpose such as
creation of derivative works, reports, performances and research.
Project Gutenberg eBooks may be modified and printed and given
away—you may do practically ANYTHING in the United States with
eBooks not protected by U.S. copyright law. Redistribution is subject
to the trademark license, especially commercial redistribution.

START: FULL LICENSE


THE FULL PROJECT GUTENBERG LICENSE
PLEASE READ THIS BEFORE YOU DISTRIBUTE OR USE THIS WORK

To protect the Project Gutenberg™ mission of promoting the free


distribution of electronic works, by using or distributing this work (or
any other work associated in any way with the phrase “Project
Gutenberg”), you agree to comply with all the terms of the Full
Project Gutenberg™ License available with this file or online at
www.gutenberg.org/license.

Section 1. General Terms of Use and


Redistributing Project Gutenberg™
electronic works
1.A. By reading or using any part of this Project Gutenberg™
electronic work, you indicate that you have read, understand, agree
to and accept all the terms of this license and intellectual property
(trademark/copyright) agreement. If you do not agree to abide by all
the terms of this agreement, you must cease using and return or
destroy all copies of Project Gutenberg™ electronic works in your
possession. If you paid a fee for obtaining a copy of or access to a
Project Gutenberg™ electronic work and you do not agree to be
bound by the terms of this agreement, you may obtain a refund
from the person or entity to whom you paid the fee as set forth in
paragraph 1.E.8.

1.B. “Project Gutenberg” is a registered trademark. It may only be


used on or associated in any way with an electronic work by people
who agree to be bound by the terms of this agreement. There are a
few things that you can do with most Project Gutenberg™ electronic
works even without complying with the full terms of this agreement.
See paragraph 1.C below. There are a lot of things you can do with
Project Gutenberg™ electronic works if you follow the terms of this
agreement and help preserve free future access to Project
Gutenberg™ electronic works. See paragraph 1.E below.
1.C. The Project Gutenberg Literary Archive Foundation (“the
Foundation” or PGLAF), owns a compilation copyright in the
collection of Project Gutenberg™ electronic works. Nearly all the
individual works in the collection are in the public domain in the
United States. If an individual work is unprotected by copyright law
in the United States and you are located in the United States, we do
not claim a right to prevent you from copying, distributing,
performing, displaying or creating derivative works based on the
work as long as all references to Project Gutenberg are removed. Of
course, we hope that you will support the Project Gutenberg™
mission of promoting free access to electronic works by freely
sharing Project Gutenberg™ works in compliance with the terms of
this agreement for keeping the Project Gutenberg™ name associated
with the work. You can easily comply with the terms of this
agreement by keeping this work in the same format with its attached
full Project Gutenberg™ License when you share it without charge
with others.

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. Unless you have removed all references to Project Gutenberg:

1.E.1. The following sentence, with active links to, or other


immediate access to, the full Project Gutenberg™ License must
appear prominently whenever any copy of a Project Gutenberg™
work (any work on which the phrase “Project Gutenberg” appears,
or with which the phrase “Project Gutenberg” is associated) is
accessed, displayed, performed, viewed, copied or distributed:
This eBook is for the use of anyone anywhere in the United
States and most other parts of the world at no cost and with
almost no restrictions whatsoever. You may copy it, give it away
or re-use it under the terms of the Project Gutenberg License
included with this eBook or online at www.gutenberg.org. If you
are not located in the United States, you will have to check the
laws of the country where you are located before using this
eBook.

1.E.2. If an individual Project Gutenberg™ electronic work is derived


from texts not protected by U.S. copyright law (does not contain a
notice indicating that it is posted with permission of the copyright
holder), the work can be copied and distributed to anyone in the
United States without paying any fees or charges. If you are
redistributing or providing access to a work with the phrase “Project
Gutenberg” associated with or appearing on the work, you must
comply either with the requirements of paragraphs 1.E.1 through
1.E.7 or obtain permission for the use of the work and the Project
Gutenberg™ trademark as set forth in paragraphs 1.E.8 or 1.E.9.

1.E.3. If an individual Project Gutenberg™ electronic work is posted


with the permission of the copyright holder, your use and distribution
must comply with both paragraphs 1.E.1 through 1.E.7 and any
additional terms imposed by the copyright holder. Additional terms
will be linked to the Project Gutenberg™ License for all works posted
with the permission of the copyright holder found at the beginning
of this work.

1.E.4. Do not unlink or detach or remove the full Project


Gutenberg™ License terms from this work, or any files containing a
part of this work or any other work associated with Project
Gutenberg™.

1.E.5. Do not copy, display, perform, distribute or redistribute this


electronic work, or any part of this electronic work, without
prominently displaying the sentence set forth in paragraph 1.E.1
with active links or immediate access to the full terms of the Project
Gutenberg™ License.

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.

1.E.7. Do not charge a fee for access to, viewing, displaying,


performing, copying or distributing any Project Gutenberg™ works
unless you comply with paragraph 1.E.8 or 1.E.9.

1.E.8. You may charge a reasonable fee for copies of or providing


access to or distributing Project Gutenberg™ electronic works
provided that:

• 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 provide a full refund of any money paid by a user who


notifies you in writing (or by e-mail) within 30 days of receipt
that s/he does not agree to the terms of the full Project
Gutenberg™ License. You must require such a user to return or
destroy all copies of the works possessed in a physical medium
and discontinue all use of and all access to other copies of
Project Gutenberg™ works.

• You provide, in accordance with paragraph 1.F.3, a full refund of


any money paid for a work or a replacement copy, if a defect in
the electronic work is discovered and reported to you within 90
days of receipt of the work.

• You comply with all other terms of this agreement for free
distribution of Project Gutenberg™ works.

1.E.9. If you wish to charge a fee or distribute a Project Gutenberg™


electronic work or group of works on different terms than are set
forth in this agreement, you must obtain permission in writing from
the Project Gutenberg Literary Archive Foundation, the manager of
the Project Gutenberg™ trademark. Contact the Foundation as set
forth in Section 3 below.

1.F.

1.F.1. Project Gutenberg volunteers and employees expend


considerable effort to identify, do copyright research on, transcribe
and proofread works not protected by U.S. copyright law in creating
the Project Gutenberg™ collection. Despite these efforts, Project
Gutenberg™ electronic works, and the medium on which they may
be stored, may contain “Defects,” such as, but not limited to,
incomplete, inaccurate or corrupt data, transcription errors, a
copyright or other intellectual property infringement, a defective or
damaged disk or other medium, a computer virus, or computer
codes that damage or cannot be read by your equipment.

1.F.2. LIMITED WARRANTY, DISCLAIMER OF DAMAGES - Except for


the “Right of Replacement or Refund” described in paragraph 1.F.3,
the Project Gutenberg Literary Archive Foundation, the owner of the
Project Gutenberg™ trademark, and any other party distributing a
Project Gutenberg™ electronic work under this agreement, disclaim
all liability to you for damages, costs and expenses, including legal
fees. YOU AGREE THAT YOU HAVE NO REMEDIES FOR
NEGLIGENCE, STRICT LIABILITY, BREACH OF WARRANTY OR
BREACH OF CONTRACT EXCEPT THOSE PROVIDED IN PARAGRAPH
1.F.3. YOU AGREE THAT THE FOUNDATION, THE TRADEMARK
OWNER, AND ANY DISTRIBUTOR UNDER THIS AGREEMENT WILL
NOT BE LIABLE TO YOU FOR ACTUAL, DIRECT, INDIRECT,
CONSEQUENTIAL, PUNITIVE OR INCIDENTAL DAMAGES EVEN IF
YOU GIVE NOTICE OF THE POSSIBILITY OF SUCH DAMAGE.

1.F.3. LIMITED RIGHT OF REPLACEMENT OR REFUND - If you


discover a defect in this electronic work within 90 days of receiving
it, you can receive a refund of the money (if any) you paid for it by
sending a written explanation to the person you received the work
from. If you received the work on a physical medium, you must
return the medium with your written explanation. The person or
entity that provided you with the defective work may elect to provide
a replacement copy in lieu of a refund. If you received the work
electronically, the person or entity providing it to you may choose to
give you a second opportunity to receive the work electronically in
lieu of a refund. If the second copy is also defective, you may
demand a refund in writing without further opportunities to fix the
problem.

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.

1.F.5. Some states do not allow disclaimers of certain implied


warranties or the exclusion or limitation of certain types of damages.
If any disclaimer or limitation set forth in this agreement violates the
law of the state applicable to this agreement, the agreement shall be
interpreted to make the maximum disclaimer or limitation permitted
by the applicable state law. The invalidity or unenforceability of any
provision of this agreement shall not void the remaining provisions.

1.F.6. INDEMNITY - You agree to indemnify and hold the Foundation,


the trademark owner, any agent or employee of the Foundation,
anyone providing copies of Project Gutenberg™ electronic works in
accordance with this agreement, and any volunteers associated with
the production, promotion and distribution of Project Gutenberg™
electronic works, harmless from all liability, costs and expenses,
including legal fees, that arise directly or indirectly from any of the
following which you do or cause to occur: (a) distribution of this or
any Project Gutenberg™ work, (b) alteration, modification, or
additions or deletions to any Project Gutenberg™ work, and (c) any
Defect you cause.

Section 2. Information about the Mission


of Project Gutenberg™
Project Gutenberg™ is synonymous with the free distribution of
electronic works in formats readable by the widest variety of
computers including obsolete, old, middle-aged and new computers.
It exists because of the efforts of hundreds of volunteers and
donations from people in all walks of life.

Volunteers and financial support to provide volunteers with the


assistance they need are critical to reaching Project Gutenberg™’s
goals and ensuring that the Project Gutenberg™ collection will
remain freely available for generations to come. In 2001, the Project
Gutenberg Literary Archive Foundation was created to provide a
secure and permanent future for Project Gutenberg™ and future
generations. To learn more about the Project Gutenberg Literary
Archive Foundation and how your efforts and donations can help,
see Sections 3 and 4 and the Foundation information page at
www.gutenberg.org.

Section 3. Information about the Project


Gutenberg Literary Archive Foundation
The Project Gutenberg Literary Archive Foundation is a non-profit
501(c)(3) educational corporation organized under the laws of the
state of Mississippi and granted tax exempt status by the Internal
Revenue Service. The Foundation’s EIN or federal tax identification
number is 64-6221541. Contributions to the Project Gutenberg
Literary Archive Foundation are tax deductible to the full extent
permitted by U.S. federal laws and your state’s laws.

The Foundation’s business office is located at 809 North 1500 West,


Salt Lake City, UT 84116, (801) 596-1887. Email contact links and up
to date contact information can be found at the Foundation’s website
and official page at www.gutenberg.org/contact

Section 4. Information about Donations to


the Project Gutenberg Literary Archive
Foundation
Project Gutenberg™ depends upon and cannot survive without
widespread public support and donations to carry out its mission of
increasing the number of public domain and licensed works that can
be freely distributed in machine-readable form accessible by the
widest array of equipment including outdated equipment. Many
small donations ($1 to $5,000) are particularly important to
maintaining tax exempt status with the IRS.

The Foundation is committed to complying with the laws regulating


charities and charitable donations in all 50 states of the United
States. Compliance requirements are not uniform and it takes a
considerable effort, much paperwork and many fees to meet and
keep up with these requirements. We do not solicit donations in
locations where we have not received written confirmation of
compliance. To SEND DONATIONS or determine the status of
compliance for any particular state visit www.gutenberg.org/donate.

While we cannot and do not solicit contributions from states where


we have not met the solicitation requirements, we know of no
prohibition against accepting unsolicited donations from donors in
such states who approach us with offers to donate.

International donations are gratefully accepted, but we cannot make


any statements concerning tax treatment of donations received from
outside the United States. U.S. laws alone swamp our small staff.

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.

Section 5. General Information About


Project Gutenberg™ electronic works
Professor Michael S. Hart was the originator of the Project
Gutenberg™ concept of a library of electronic works that could be
freely shared with anyone. For forty years, he produced and
distributed Project Gutenberg™ eBooks with only a loose network of
volunteer support.
Project Gutenberg™ eBooks are often created from several printed
editions, all of which are confirmed as not protected by copyright in
the U.S. unless a copyright notice is included. Thus, we do not
necessarily keep eBooks in compliance with any particular paper
edition.

Most people start at our website which has the main PG search
facility: www.gutenberg.org.

This website includes information about Project Gutenberg™,


including how to make donations to the Project Gutenberg Literary
Archive Foundation, how to help produce our new eBooks, and how
to subscribe to our email newsletter to hear about new eBooks.
Welcome to our website – the perfect destination for book lovers and
knowledge seekers. We believe that every book holds a new world,
offering opportunities for learning, discovery, and personal growth.
That’s why we are dedicated to bringing you a diverse collection of
books, ranging from classic literature and specialized publications to
self-development guides and children's books.

More than just a book-buying platform, we strive to be a bridge


connecting you with timeless cultural and intellectual values. With an
elegant, user-friendly interface and a smart search system, you can
quickly find the books that best suit your interests. Additionally,
our special promotions and home delivery services help you save time
and fully enjoy the joy of reading.

Join us on a journey of knowledge exploration, passion nurturing, and


personal growth every day!

ebookbell.com

You might also like