Amr2013 065 06 060802
Amr2013 065 06 060802
Amr2013 065 06 060802
net/publication/260633648
CITATIONS READS
11 2,999
6 authors, including:
Some of the authors of this publication are also working on these related projects:
All content following this page was uploaded by Faxin Li on 02 December 2014.
Fig. 1 Schematic depictions of typical domain structures observed in ferroelectric (a) single
crystal and (b) polycrystal ceramic
Fig. 3 Variations of (a) Er and (b) Eh on the rim of the hole subjected to tension at infinity.
Reprinted from Ref. [32] with the kind permission of Elsevier Science Ltd.
Fig. 4 Variations of (a) Dh and (b) Eh on the rim of the hole subjected to electric displacement
loading at infinity. Reprinted from Ref. [32] with the kind permission of Elsevier Science Ltd.
h ¼ 0 deg and h ¼ 180 deg are the local maxima and minima, is relatively simple, and Eh takes its minima and maxima at
respectively. h ¼ 0 deg and h ¼ 180 deg.
Figure 3 shows the curves of the tension-induced electric field The case of applied electric displacement at infinity along the
versus the angle h. The radial component Er has a maxima at Y-direction is also investigated. The curve of the electric displace-
h ¼ 45 deg and h ¼ 135 deg and a minima at h ¼ 90 deg, and ment versus angle h is shown in Fig. 4(a). Dh reaches its maxima
these three points are just the points at which Dh changes its direc- at h ¼ 0 deg and its minima at h ¼ 180 deg with Dh max ¼ 2:19D0
tion. The curve of the circumferential component Eh versus angle and Dh ¼ 0 at h ¼ 90 deg. Figure 4(b) gives the distribution of Eh
Du ¼ N Dae (2.2.9)
K Dae ¼ DP (2.2.11)
where
X
K¼ GT Ke G (2.2.12)
e
ð
Ke ¼ BT CBdve (2.2.13) Fig. 5 Stable electric field versus electric displacement hyster-
ve esis loops (the thin line is the simulation result and the bold
ð X ð line the experimental result). Reprinted from Ref. [34] with the
DP ¼ NT DTds þ GT BT CDCr dve kind permission of Elsevier Science Ltd.
St þSD e ve
X ð X ð
þ GT BT DCðC Cr Þdve GT BT DRr dve
e ve e ve
(2.2.14)
correspond to each image related to the local domain states in other states (see Figs. 8 and 9). With the above notion, it is known
Figs. 8 and 9. Figure 8 shows the whole process of the domains that switching between 1 and 2, 3 and 4, and 5 and 6 is 180 deg
switching from the initial unpoled state subjected to the bipolar switching and other types of switching are 90 deg switching.
electric field. Figure 9 demonstrates the particular process of do- It can be seen from Figs. 8 and 9 that there is no direct 180 deg
main switching when the applied electric field is close to the coer- domain switching, but two successive 90 deg switchings occur
cive electric field. In Fig. 9, three states of the domains with when the applied electric field reverses, which is caused by the
continuous loading steps are displayed in detail (i.e., (a), (d), and fact that the energy barrier of the 90 deg switching used here is
(g) correspond to the point F in Fig. 7; (b), (e), and (h) correspond about 1/4 of the barrier of 180 deg switching [34]. In Fig. 9, it can
to point G; and (c), (f), and (i) correspond to point G’. be found that the domains near the free boundaries can switch
As denoted in Ref. [34], in Figs. 8 and 9, the initial state of more easily than those inside the material and those near the
each domain is marked 1, and its 180 deg reorientation state is strongly constrained boundaries. In Figs. 9(a) and 9(b), when the
marked 2. Four possible 90 deg reorientation states are marked 3, macroscopic switching begins, several domains at or near the top
4, 5, and 6, respectively. The state 3 can be reached from the state face boundary switch, while the domains inside the material
4 by a 180 deg switching and vice versa, and the same for state 5 remain unchanged (see Figs. 9(d) and 9(e)). The domains near the
and state 6. Each state is specified by a single color different from bottom face boundary do not switch at all, even after two loading
steps (see Figs. 9(g)–9(i)). All these may indicate that the 90 deg
Fig. 8 The whole process of domain switching under uniaxial Fig. 9 Domain-switching process when the applied electric
electric loading. Reprinted from Ref. [34] with the kind permis- field is close to the coercive field. Reprinted from Ref. [34] with
sion of Elsevier Science Ltd. the kind permission of Elsevier Science Ltd.
simplified and stable at the sacrifice of numerical accuracy to As to the case of 90 deg domain switching illustrated in Fig.
some extent. 13(b), besides the large depolarization field Ed by the spontaneous
polarization change, due to the spontaneous strain change during
Domain-switching instability in FE models based on domain- domain switching, a large mechanical field (i.e., a tensile stress r1
switching criteria. As indicated by Li and Rajapakse [75], for the and a compressive r2 ) will be generated by the adjacent domains.
nonlinear FE models based on micromechanical domain- r1 and r2 are both on the order of Dc=s (1 GPa, where Dc is the
switching criteria [13,33,34,74,96,97], the equivalent nodal load magnitude of the spontaneous strain change and s is the elastic
induced by the change of spontaneous strain and spontaneous compliance constant) and also at least one order higher than the
polarization during domain switching (i.e., the second and the last corresponding coercive stress, making the switched domain
item on the right side of Eq. (2.2.14)) can be rather large com- switch back. Therefore, in an unmodified, nonlinear FE formula-
pared to a moderate electromechanical loading. For the case of tion, such as Eq. (2.2.14), domain switching will always be
180 deg domain switching without change of spontaneous strain, unstable.
as shown in Fig. 13(a), a large depolarization field Ed will be gen- Actually, the simulated domain-switching instability has been
erated [75]. If the charge screening effect is not included (this is noticed by several groups [13,33,34,74,106], while a profound ex-
the usual case in conventional simulations based on FEM), the planation is still lacking. To weaken the rather high electric/me-
depolarization field Ed will be on the order of DP=k (1050 kV/ chanical fields induced by ferroelectric/ferroelastic switching,
mm), where DP is the magnitude of the change of spontaneous Hwang et al. [13] employed a much smaller Yong’s modulus
polarization during switching and k is the dielectric constant of (about 20% of the true value) and a rather larger dielectric con-
the materials. Such a large (usually higher than ten times of the stant (about 14 times of the true value) in their simulations. Chen
coercive field) depolarization field will make the domain switch and Lynch [74] introduce a factor kð0 < k < 1Þ in the last item on
back to its original state; thus, domain-switching instability arises. the right side of Eq. (2.2.14). Kessler and Balke [106] discussed
the unstable switching in their simulations (not FE model), and
they suggested that the switching instability can be removed by
introducing ferroelectric “hardening” in the material model, which
is somewhat equivalent to the simplified bilinear D–E curve
assumption by Kamlah and Bohle [73].
Recently, to overcome the problem of domain-switching insta-
bilities, Li and Rajapakse [75] proposed a finite element model for
ferroelectrics based on constrained domain switching. In their
model, a ferroelectric polycrystalline is made up of numerous
random-oriented grains, each of which contains multidomains and
is represented by a finite element. Charge-screening effect in real
ferroelectrics is taken into account in the model; thus, 180 deg do-
main switching will not be constrained, while the internal stress
induced by non-180 deg switching cannot be compensated in a
similar way and is considered naturally by the FE method. The
fraction of non-180 deg switching in each grain is obtained
through a dichotomy approximation. In this way, the domain-
switching instability problem can be effectively avoided.
Ku0 ¼ f 0 (2.3.19)
Fig. 17 The distribution of stress ryy near the crack tip [76]
Fig. 16 Schematic diagram of interface crack
PZT-5H PZT-4
2
C11 12:6 1010 Nm 13:9 1010 Nm2
C12 5:5 1010 Nm2 7:78 1010 Nm2
C13 5:3 1010 Nm2 7:43 1010 Nm2
C33 11:7 1010 Nm2 11:3 1010 Nm2
C44 3:53 1010 Nm2 25:6 1010 Nm2
e31 6.5 Cm2 6.98 Cm2
e33 23.3 Cm2 13.84 Cm2
e15 17.0 Cm2 13.44 Cm2
d11 151 1010 CV1m1 60:0 1010 CV1m1
d33 130 1010 CV1m1 54:7 1010 CV1m1
the film is larger than that in the substrate; namely, in Fig. 19, the
electric intensity illustrates more concentration near to the crack
than that at the crack tip.
The distribution of hoop stress rh , when various electric
mechanical loadings are applied to the specimen, is illustrated in Fig. 18 The distributions of electric field Ey around the crack
pffiffiffiffiffiffiffiffiffi
ffi tip [76]
Fig. 20, where the coefficient k ¼ 2r=a and a and r are the
crack length and radial distance of the point from the crack tip,
respectively. For low values of the electrical-to-mechanical load and its magnitude is getting larger. The asymmetry of rh along
ratio, the hoop stress is tensile on the whole and does not reach its the crack surface becomes more and more evident. Because of the
maximum value at h ¼ 0 . With the increase of electromechanical large thickness difference between the thin film and substrate, the
load ratio, the hoop stress rh has the tendency to be press stress, stresses in the film are greater than that of in the substrate.
Fig. 21 The engineering domain structures with both 90 deg and 180 deg domain
walls: (a) calculated by the phase field simulation and (b) observed by experiments
in BaTiO3 single crystal. The arrow represents the polarization direction. Reprinted
from Ref. [45] with the kind permission of American Physical Society.
vacancies are uniformly distributed initially with random noises, the redistribution of oxygen vacancies (Fig. 22(b)). It can be found
the resulting distribution of the electrostatic potential and oxygen that the oxygen vacancies are attracted to different sides of the 90
vacancy density are shown in Fig. 22. The electric potential deg domain wall, but no such accumulation occurs near the 180
changes abruptly at the 90 deg domain walls, but not at the 180 deg domain wall. This can be used to explain the domain wall pin-
deg domain walls. Thus, a large depolarization field exists at the ning and the resulted polarization fatigue.
90 deg domain wall, with the maximum magnitude as high as 5.1 To examine the influences of the oxygen vacancies on the
MV/m. The inhomogeneous electrostatic field furthermore drives electromechanical responses of the BaTiO3 single crystal, an
Fig. 23 The electric hysteresis loops (a) and butterfly loops (b) of rank-2 domain structures with different oxygen vacancy den-
sities, and the domain evolutions for Nd 5 1.2 3 1024 m–3 under a positive alternating field (c). Domains indexed by (I) and (II) in
(c) correspond to the two states (I) and (II) in (a) and (b). Reprinted from Ref. [45] with the kind permission of American Physical
Society.
where E and r are the applied electric field vector and applied
stress tensor, respectively. k is the isotropic dielectric constant and
landv are the isotropic shear modulus and Poisson’s ratio. e is the
mean strain tensor of the whole material system or the matrix. ld
is the equivalent shear modulus during domain switching and
should be calibrated experimentally using the stress-strain curve.
As the material will become “softer” when domain switching
occurs, ld < l always hold true.
Pr and er are the remnant polarization and remnant strain of a
Fig. 26 The size-dependent out-of-plane polarization and grain, which can be expressed as linear functions of the volume
domain pattern. The domain exhibits a zigzag pattern with fractions of domains as follows:
eight variants coexisting when h > 7.6 nm, a zigzag pattern with
four variants coexisting when 6.8 nm < h < 8.0 nm, a vortex X
N X
N
pattern with four variants coexisting when 4.4 nm < h < 7.2 nm, Pr ¼ fi Pi ; er ¼ fi ei (3.2.2)
and a stripe pattern with two variants coexisting when
i¼1 i¼1
2.4 nm < h < 4.8 nm; when the thickness is smaller than 2.8 nm,
the domain disappears, indicating that the ferroelectricity is
suppressed. Reprinted from Ref. [46] with the kind permission where Pi and ei are the spontaneous polarization and spontaneous
of The American Institute of Physics. strain tensor of the ith domain.
The six types of tetragonal domains and eight types of rhombo-
hedral domains in quasicubic coordinates are illustrated in
Fig. 28.
vanishes. The mechanical interactions between grains, however, For the tetragonal ferroelectric crystals, we have
cannot be treated similarly and are considered in a self-
consistent Eshelby inclusion manner [155]. That is, each grain is Pð1Þ ¼ Pð2Þ ¼ P0 ð0; 0; 1ÞT ; Pð3Þ ¼ Pð4Þ ¼ P0 ð1; 0; 0ÞT ;
regarded as an inclusion surrounded by an infinite large matrix
with the materials properties the same as the whole material, as Pð5Þ ¼ Pð6Þ ¼ P0 ð0; 1; 0ÞT (3.2.3)
shown in Fig. 27. As the spherical inclusion case is the simplest 2 3
1
and no evidence indicates that other shapes of inclusion are more S0 6 7
accurate, we use the spherical inclusion case in this model. Fur- eð1Þ ¼ eð2Þ ¼ 4 1 5;
3
thermore, to grasp the major responses of domain switching and 2
to make the computational complexity affordable, we also 2 3 2 3
assume that a ferroelectric ceramic is dielectrically and elasti- 2 1
S0 6 7 S0 6 7
cally isotropic and shows linear dielectric and elastic behavior eð3Þ ¼ eð4Þ ¼ 4 1 5; eð5Þ ¼ eð6Þ ¼ 4 2 5
3 3
unless domain switching occurs (i.e., all the nonlinear items are 1 1
caused by domain switching).
(3.2.4)
Free energy of a single grain. In the present model, as For rhombohedral crystals, we have
the interactions between grains have been considered in an
Eshelby inclusion manner, we use the free energy of each grain P0 P0
Pð1Þ ¼ Pð2Þ ¼ pffiffiffi ð1; 1; 1ÞT ; Pð3Þ ¼ Pð4Þ ¼ pffiffiffi ð1; 1; 1ÞT ;
as the optimization objective and sweep the optimization process 3 3
over all grains to get the domain structures and properties of the P0 P0
whole material system. The free energy of a specific grain is P ¼ P ¼ pffiffiffi ð1; 1; 1Þ ; P ¼ P ¼ pffiffiffi ð1; 1; 1ÞT
ð5Þ ð6Þ T ð7Þ ð8Þ
3 3
expressed as
(3.2.5)
Fig. 27 2D Illustration of material model for morphotropic fer- Fig. 28 (a) Six types of tetragonal domains and (b) eight types
roelectric ceramics of rhombohedral domains in ferroelectric crystals
Convergence and computational complexity. The simulation Ferroelastic domain switching under mechanical loading. We
results on tetragonal and rhombohedral PZT ceramics have shown also simulated the ferroelastic domain-switching process in
that both the P–E hysteresis loops and strain curves will stabilize unpoled tetragonal and rhombohedral PZT ceramics under uniax-
only after 1.5 cycles of electric loading (see Fig. 29 for the case of ial tension/compression loading using the proposed model [52].
tetragonal PZT ceramics) [163]. The hysteresis loops and strain The material constants used here are the same as listed in Table 4,
curves for subsequent cyclic loading completely overlap the for- except that the energy barrier for ferroelastic domain switching is
mer curves, which indicate good convergence of this computa-
tional model.
It should be mentioned that, by focusing on the volume frac-
tions of domains instead of the detailed domain patterns in grains,
this optimization-based model has a much smaller computational
complexity compared to other methods, such as the phase field
model (PFM). Therefore, it is feasible to study the 3D cases with
a very large number of grains using this model. For the case of a
polycrystalline ceramic with 10,000 grains and 61 loading steps,
the computation time of this model on a Pentium 3.0 GHz perso-
nal computer is only 1.5 h for tetragonal PZT and about 2.5 h for
rhombohedral PZT ceramics.
Ferroelectric switching under electric loading. Figure 30
shows the simulated polarization and longitudinal strain curves
under cyclic electric loading up to 5EC for both tetragonal and
rhombohedral PZT ceramics. Both the remanent polarization and
strain of the rhombohedral PZT ceramics are obviously larger
than that of the tetragonal PZT ceramics. The rather small strain
in the tetragonal ceramics indicates that little 90 deg domain
switching occurs, although the 180 deg switching is almost
complete, which can be estimated from the polarization curve in
Fig. 30(a). In comparison, the large remanent strain of rhombohe-
dral PZT ceramics shows that a considerable amount of non-180
deg switching has occurred during electric poling. From Fig. 30, it
can also be seen that, in both ceramics, domain switching is a
gradual process with the applied electric loading.
Figure 31 shows the domain textures evolution process in both
tetragonal and rhombohedral PZT ceramics corresponding to the
electric loading points indicated in Fig. 30(a). It can be seen that,
during electric loading, the domain textures of tetragonal PZT
ceramics change little from the initial unpoled state, which is con-
sistent with the small remanent strain shown in Fig. 29(b). As
expected, the domain textures of rhombohedral PZT ceramics
change a lot from the unpoled state during electric poling
(R1!R2), as shown in Fig. 31(b). It also shows that, upon remov-
ing the electric field, partial domain switching reverses, which
causes domain texture variations (R2!R3). Nevertheless, in Fig. 30 Simulated (a) polarization and (b) strain curves of tet-
rhombohedral PZT ceramics, after electrical poling, the domain ragonal and rhombohedral PZT ceramics. Reprinted from Ref.
textures approach the theoretically saturated state [161] and can [53] with the kind permission of The American Institute of
never return to the initial unpoled state in subsequent loading. Physics.
Fig. 31 (a) (001) pole figure of tetragonal PZT and (b) (111) pole
figures of rhombohedral PZT ceramics during cyclic electric
loading. Reprinted from Ref. [53] with the kind permission of
The American Institute of Physics.
Fig. 32 Switching strain curves of tetragonal and rhombohe- Fig. 33 (a) (001) pole figures of tetragonal PZT and (b) (111)
dral PZT ceramics under uniaxial tension/compression. pole figures of rhombohedral PZT ceramics during uniaxial ten-
Reprinted from Ref. [52] with the kind permission of Elsevier sion/compression. Reprinted from Ref. [52] with the kind per-
Science Ltd. mission of Elsevier Science Ltd.
Fig. 34 (a) Switching strain-stress curves and (b) the cumula- Fig. 35 Comparison of switching strain versus stress curves
tive pole figures of (001) and (111) axes in morphotropic PZT of tetragonal, rhombohedral, and MPB PZT ceramics (the nomi-
ceramics under uniaxial tension and compression. Reprinted nal coercive stress rC 5 70 MPa). Reprinted from Ref. [52] with
from Ref. [52] with the kind permission of Elsevier Science Ltd. the kind permission of Elsevier Science Ltd.
Fig. 37 (a) D-E hysteresis loops, (b) butterfly curves, and (c)
reversed butterfly curves of a morphotropic PZT ceramic under
electric loading with different levels of precompression.
Reprinted from Ref. [54] with the kind permission of The Ameri-
can Institute of Physics.
grain is visited 102 times during one Monte Carlo step. The simula- r < 1 describes the reorientation of polar axis of domains toward
tion continues until a stable domain configuration emerges, when poling direction during electric poling, and r > 1 describes the
all the switchings are rejected, and thus, no switching occurs during reorientation of polar axis away from poling direction during me-
one Monte Carlo step. It should be noted that a physical switching chanical depoling. Since ODF is usually measured from X-ray or
step is accomplished through many computational switching steps neutron diffraction, only non-180 deg domain switching is
in Monte Carlo simulation. To speed up the computation, a large reflected in the change of March coefficient, due to the inherent
initial ckT ¼ 103 J=m3 is chosen at beginning and then is reduced centrosymmetry of the diffraction technique [51].
to 95% of the previous value after every Monte Carlo step. This is The variation of March coefficient versus electric field is shown
similar to a physical annealing process. Also note that the simula- in Fig. 40, where in all PZT phases before poling, the March coeffi-
tion results of this model are not sensitive to the initial temperature. cient is close to unit. After applying the electric field, very little
change in March coefficient is observed in single-phase tetragonal
Electric Poling. We consider the electric poling process of
PZT, which is consistent with the simulated domain textures evolu-
three different PZT ceramics, including the tetragonal, the rhom-
tion using the optimization-based model. In comparison, the March
bohedral, and the morphotropic, in which all possible variants
coefficient for tetragonal phase at MPB decreases to about 0.2 after
exist in grains with equal volume fractions. A bipolar poling elec-
poling, suggesting extensive 90 deg switching that reorients polar-
tric field of 3 MV/m is applied to the ceramic with an increment
izations toward the poling direction, which contributes to large
of 0.1 MV/m. The obtained D–E hysteresis loops and butterfly
strain observed in PZT near the MPB during poling. In single-
loops are shown in Fig. 39, in which EC ¼ 1MV=m is used for all
phase rhombohedral PZT, the March coefficient decreases to about
the PZT ceramics. It can be seen that the difference in polarizabil-
0.63 after poling, suggesting modest extent of 71 deg and 109 deg
ity across MPB is evident. Tetragonal PZT has the largest coerciv-
domain switchings that are responsible for the observed modest
ity and is least polarizable, with remanent polarization calculated
strain. Also, the rhombohedral phase at MPB shows more extensive
as 0.21 C/m2. This is only 44% of the single crystalline spontane-
71 deg and 109 deg switchings, as seen from larger change in its
ous polarization and is close to 0.17 C/m2 measured in experiment
March coefficient. The differences in macroscopic behavior of PZT
[171]. The calculated switching strain is only 0.037%, much
ceramics across MPB during poling thus merely reflect the differ-
smaller than single crystalline value of 2.0% and comparable to
ences in their microscopic domain switching. To further validate
the measured value of 0.047% in tetragonal BaTiO3 ceramics,
this Monte-Carlo model, simulations were also done on a rhombo-
which have similar single crystalline transformation strain [172].
hedral, soft PLZT ceramics under bipolar electric loading, and the
Rhombohedral PZT shows improved polarizability, with remanent
results can fit well with the experiments, as shown in Fig. 41 [51].
polarization calculated as 0.31 C/m2, about 64.6% of single
crystalline spontaneous polarization, and agrees well with that of Mechanical loading. The behavior of mechanical depoling of
0.31 C/m2 measured in experiment [171]. Also consistent with the poled PZT ceramics is also investigated by using the Monte Carlo
experiments, the calculated switching strain in rhombohedral model. The starting point is a ceramic poled by an electric field of
ceramics is 0.14%, much larger than tetragonal ceramic. As 3 MV/m, and then the poling field is removed. The stable domain
expected, PZT at MPB has much larger remanent polarization configuration is obtained by Monte Carlo simulation as discussed
(0.36 C/m2) and switching strain (0.375%), both in excellent agree- previously. Then, a compressive stress up to 300 MPa is gradually
ment with experimental values (0.36 C/m2 and 0.37%) [171,172]. applied to the poled ceramic along the poling direction, and then
To understand the difference in polarizability of PZT ceramics it is gradually removed. During both loading and unloading, the
across MPB, we examine the texture evolution of PZT during increment of stress is 10 MPa. The simulated strain-stress curves
electric poling. Here, the orientation distribution function (ODF) and stress-depolarization curves for tetragonal, rhombohedral, and
PðhÞ of domains is employed, from which the March coefficient r MPB PZT ceramics are shown in Fig. 42. It can be seen that PZT
of tetragonal and rhombohedral phases can be fitted according to at MPB shows very large reduction in remanent polarization and
the following equation [164]: is almost completely depoled under the maximum compressive
stress of 300 MPa. This indicates that extensive non-180 deg do-
PðhÞ ¼ ðr 2 cos2 h þ r 1 sin2 hÞ3=2 (3.3.11) main switching occurs, which leads to large nonlinear strain as
high as 0.62% that is not recovered after stress unloading. In com-
As indicated by Tang et al. [51], the March coefficient r is particu- parison, tetragonal PZT only exhibits a polarization reduction of
larly convenient to describe the texture evolution in ferroelectric 0.08 C/m2 and rhombohedral PZT has a modest polarization
ceramics due to domain switching. For example, r ¼ 1 describes reduction of 0.24 C/m2, suggesting modest extent of 71 deg and
random orientation distribution of domains in an unpoled ceramic, 109 deg switchings that lead to modest nonlinear strain.
Fig. 41 Comparison between simulation and experiment to hysteresis (left) and butterfly (right) loops of soft
PLZT during electric poling. Reprinted from Ref. [51] with the kind permission of Elsevier Science Ltd.
The change of March coefficient during mechanical depoling is model, the Monte Carlo model also overestimates the coercive
also calculated and shown in Fig. 43 [51]. In all three PZT, the stress, and thus, a more accurate model addressing the softening
March coefficient increases beyond unit, suggesting that polariza- during domain switching is required in the future.
tions are reoriented away from the poling direction by the com-
pressive stress. As expected, the increase is largest for PZT at
MPB and smallest in single-phase tetragonal PZT, which agrees 4 Atomic-Level Electromechanically Coupled
well with the neutron diffraction data [164]. As indicated by Tang Computational Methods
et al. [51], if we set the spontaneous polarization in all the variants
to be zero, very little changes in March coefficient and strain are 4.1 First-Principles Method of Ferroelectrics. There has
observed in single-phase tetragonal PZT. These results suggest been rapid development in the atomistic modeling of materials in
that the strain compatibility plays a dominant role in domain the last decades, driven by the fast increase of the computational
switching in ferroelectric ceramics and polarization compatibility power and important progresses in the development of efficient
is less influential [51]. algorithms. Nowadays, it is possible to describe very accurately
To further compare the simulation with the experimental results, the properties of materials by using methods directly based on
we also conducted simulations on a soft PLZT ceramic under me- solving the fundamental equations of quantum mechanics. Even if
chanical loading and compared the simulated depolarization curve it may require some approximations for the study of complex sys-
and stress-strain curve with the experiments, as shown in Fig. 44 tems, these methods are free of empirically adjustable parameters.
[51]. It can be seen that, overall, the simulated results can fit well Therefore, they are referred to as “first-principles“or “ab initio”
with the experimental data, although they differ in some details methods. In recent years, first-principles computations have
near the transition regions. Similar to the optimization-based become ubiquitous in materials science, and it is hard to find a
Fig. 43 Evolution of March coefficient of PZT ceramics in tetragonal (left) and rhombohedral (right) phases dur-
ing mechanical depoling. Reprinted from Ref. [51] with the kind permission of Elsevier Science Ltd.
material of scientific interest that has not been studied increases but piezoelectric constant decreases as thickness reduces
computationally. [180]. And for BaTiO3 nanowires, the critical size is 1.2 nm,
Ferroelectric materials have been intensively studied from first- determined from first-principles [69], which agrees with the data
principles since the 1990s, when the modern theory of polariza- extrapolated from experiment [181]. Recent first-principles calcu-
tion was developed [173,174]. The accuracy of density functional lations show that some novel polarization patterns appear in ferro-
theory for characterizing the ground-state structures of ferroelec- electric nanodots and nanowires, which have never been found in
trics was first determined for BaTiO3 and PbTiO3 [66] and later ferroelectric bulk before [68,70,177,182], and all these patterns
extended to other ferroelectric materials [175]. First-principles are found to correspond to the topological point defects with dif-
methods have been remarkably successful in modeling many of ferent winding numbers [177]. Bousquet et al. [179] found an un-
the physical properties of ferroelectric perovskite materials, usual improper ferroelectricity in short-period PbTiO3/SrTiO3
including the ground state structure, the spontaneous polarization superlattices due to the octahedra rotation in the interfaces, which
and the related dynamical effective charges, the dielectric and pie- stimulates the research for the interface engineering in superlatti-
zoelectric response, etc. ces and hybrid improper ferroelectricity in multiferroics [183].
As computational power and algorithmic efficiency have In this section, we first briefly review the modern theory of
improved, it has become possible to study increasingly complex polarization and then review its applications in topological polar-
systems, such as surfaces [176], ultrathin films [67], nanowires ization in nanostructures, flexoelectricity, and size-dependence
[69,177], and superlattices [178,179]. As the size reduces, the fer- properties of films.
roelectric nanostructures show some novel properties, which the
bulk materials do not have. The first-principles calculations show 4.1.1 Polarization Calculation From First-Principles. The
that the ferroelectricity disappears below 2.4 nm for BaTiO3 films macroscopic polarization is a fundamental concept in the theory
at short circuit boundary conditions [67], and its elastic stiffness of electrostatics, particularly in describing the response of systems
to applied electric fields. The presence of a spontaneous (and changes in polarization are rigorously defined and can be calcu-
switchable) macroscopic polarization is the defining property of a lated quantum mechanically using electronic structure methods.
ferroelectric material, and the macroscopic polarization is thus For example, in Fig. 45 the polarization (dipole moment per
central to the whole physics of ferroelectrics. Despite its primary unit length for one-dimensional case) is 1=2 and 1=2 (in unit jej)
role in electrostatics theories and its overwhelming importance, if choosing different unit cells for centrosymmetry (nonpolar)
the macroscopic polarization has long evaded microscopic under- case (top panel). For the distortion (polar) case (bottom panel, cor-
standing, not only at the first-principles level, but even at the level responding to the ferroelectric case) with cation displacement of
of microscopic models. A typical definition of polarization is that D, the polarizations are 1=2 þ D=a and 1=2 þ D=a for different
the macroscopic polarization of a solid is the electric dipole unit cells. It can be seen that the change in polarization between
moment per unit volume. However, such a quantity is neither polar and nonpolar cases is the same (D=2), which is independent
measurable nor model-independent: the dipole of a periodic of the choice of unit cells. Actually, the changes in polarization
charge distribution is in fact ill-defined [184]. For example, it can also correspond to experimentally measurable observables. For
be seen from the top panel of Fig. 45, without any calculation, example, the spontaneous polarization is measured by the half of
that two equal unit cells show completely opposite polarizations the difference between up- and down-polarized states.
based on this definition. The cation and anion are considered as point charges in the pre-
This difficulty is solved thanks to the so-called modern theory vious simple ionic model. However, in a real solid, there is more
of polarization, which was developed in the 1990s [173,174,185]. physics to take care of, such as the electronic contribution to the
It was realized that, even at the theoretical level, polarization dif- polarization. In real materials, the charge (nuclear and electronic)
ferences are conceptually more fundamental than the “absolute” can be decomposed into localized contributions whose dipoles
polarization. This change led to the development of a new theoret- determine polarization. The polarization can be obtained by the
ical understanding, involving formal quantities, such as Berry sum over localized charges multiplied by their positions. This is
phases and Wannier functions, that have come to be known as the straightforward for the ions, which can be still treated as point
“modern theory of polarization”. According to this theory, the charges. For the electrons, this procedure also works through
where the sum over is for all the electronic charges centered at the 4.1.2 Topological Polarization in Ferroelectric Nanowires.
Wannier centers of each occupied Wannier function. As the size of ferroelectric nanostructures continues to decrease, it
Therefore, the polarization is was initially assumed that the polarization would vanish because
ferroelectricity is a collective phenomenon. However, seminal
! works by Fu and Bellaiche [182] and Naumov et al. [68] showed
1 X ions
Xocc
WFs that the polarization may first form a ferroelectric vortex before
P ¼ Pion þ Pel ¼ ðqi ri Þ þ ðqn rn Þ
X i n disappearing. This has attracted much theoretical [69,70,190,191]
occ ð
and experimental work [181,192] to establish critical diameters
1X ions 2ie X ikR
@unk
¼ ðqi ri Þ dke unk
for ferroelectricity and understand the characteristics of ferroelec-
X i ð2pÞ3 n @k tric vortices. First-principles method plays an important role in
(4.1.5) determining critical size of ferroelectrics and exploring novel
properties in ultrathin films and nanowires. Recent study shows
that some novel topological polarization patterns appear in thin
Using the Wannier function picture, the polarization can be nanowires which do not exist in the bulk [177].
written as the sum over the charges times their positions: the con- The two-dimensional (2D) topology of the polarization field in
tribution from the positively charged ion cores and the contribu- ferroelectric nanostructures can be mapped onto the topology of
tion from the negatively charged valence electrons and the ðPx ; Py Þ for wires if the field remains constant along z, the axial
“position” of each valence electron is its Wannier center. direction of the wire. Topological defects in such a 2D field [193]
The polarization of Eq. (4.1.5) is referred to as the “formal are reviewed here. The winding number is defined as n ¼ /=2p,
polarization” to distinguish it from the “effective polarization,” / being the total angle the 2D field vector rotates when going
which is defined as the difference between two states, around a closed circuit, and n must be an integer for a continuous
field. Some novel polarization textures with n ¼ 0, 1, 1, and 3
Peff ¼ Pð1Þ Pð0Þ (4.1.6) in ultrathin BaTiO3 nanowires are found from first-principles
calculations.
where Pð1Þ and Pð0Þ are the formal polarization for state 1 and state The Ti displacement from the center of its coordination octahe-
0. Usually, the state 1 is the low-symmetry state and state 0 is the dron was taken to define PðrÞ from the relaxed structures as a
high-symmetry state (reference state). Effective polarization is measure of the local dipole. For Ti atoms on surfaces or edges, the
well-defined, and it corresponds to the experimentally measurable off-center displacement was determined by the remaining coordi-
polarization. nating O atoms. Four kinds of nanowires with different surface
Actually, the integrals in Eq. (4.1.5) are the Berry phase devel- terminations have been considered, as shown in Fig. 46. Figure 47
oped by the wavefunction unk ðrÞ as it evolves along the path k. (left column) shows the polarization patterns of the wire types of
As a result, the formalism for calculating polarization using this Fig. 46, with qualitative sketches of the corresponding topological
method is often called the Berry phase theory of polarization. defects for a 2D vector field (right column). It shows that, even
Interested readers are referred to the original papers [173–175] for the small size wires, the field textures appear very clearly. All
and reviews [186–188] for details. textures preserve inversion symmetry at the center of the wire if it
Next, we introduce the Born effective charge Zj;ab , which is is not previously broken by the surface termination.
defined as the change in polarization along a direction (Pa ) due to The inward and outward radial patterns in the first two cases in
the displacement of atom j along b direction (uj;b ), Fig. 47 correspond to the same kind of topological defect of
winding number 1. They are thus homotopical with each other but
also with the vortex structures shown in Naumov et al. [68] and
Pilania et al. [70]. Figures 47(c) and 47(d) correspond to different
winding numbers as yet not observed or proposed in this field,
namely 1 and 3. In Fig. 47(c), the defect is displaced from the
center. Figure 47(d) shows the n ¼ 3 case of the BaO-terminated
wire. Figure 47(e) is for the same BaO termination as 47(d) but
smaller thickness. Although a different discretization, the winding
around the outer circuit clearly shows the same n ¼ 3: a central
defect with n ¼ þ1 plus four other defects of n ¼ 1, as illustrated
in Fig. 47(j).
These textures are induced by surface and edge effects. The
phenomenology described can be understood as originated by the
following three tendencies: (i) Ti-terminated surfaces and edges
induce inwards Ti off-centering patterns, (ii) Ba surface termina-
tion induces outward Ti off-centering patterns, and (iii) Ba edge
termination induces an inward tendency. The competition of (ii)
and (iii) in the Ba-terminated wires gives rise to the rich n ¼ 3
pattern. By removing the Ba edges (Fig. 47(b)), the competition
disappears and the fully outward pattern is established. Similarly,
when removing the Ba edge in the stoichiometric case, the polar- Fig. 47 Left column (a)–(d): Ti off-center displacements for the
four respective nanowires in Fig. 46. (e) is similar to (d) but for a
ization field becomes topologically homogeneous (n ¼ 0), with the thinner wire. The right column sketches the corresponding 2D
field lines crossing the wire diagonally pointing left and down field lines around topological point defects of winding numbers
(not shown here). Special chemical tendencies at edges and surfa- n 5 1, 1, 21, and 23, respectively, (j) showing the n 5 23
ces have been previously reported [181]. decomposition into a central n 5 11 and four n 5 21 defects.
The respective largest arrows correspond to (a) 21 pm, (b) 35
pm, (c) 23 pm, (d) 6.8 pm, and (e) 6.6 pm off-centering displace-
ment. Reprinted from Ref. [177] with the kind permission of
4.1.3 Strain Gradient Effects in Ferroelectrics. Inhomogene- American Physical Society.
ous strains can induce dielectric polarization in all insulators due
to flexoelectricity (FxE), which linearly couples polarization to caused a recent explosion of experimental and theoretical interest
strain gradient [194]. FxE is quite different from piezoelectricity, in FxE [195–207].
which arises only in noncentrosymmetric materials. Although it is FxE refers to the linear response of electric polarization to an
a universal property of all dielectrics, FxE is nevertheless a rela- applied strain gradient,
tively small effect compared with piezoelectricity and it is nor-
mally negligible on conventional length scales. However, it may @ejk
Pi ¼ lijkl jkl ¼ lijkl (4.1.9)
become very strong at the nanoscale, where huge strain gradients @xl
can significantly affect the functional properties of dielectric
nanostructures. The possibility of large FxE effects in ferroelec- where Pi is the FxE-induced polarization, jkl ¼ @ejk =@xl is the
trics at the nanoscale with application to functional devices has strain gradient, and lijkl are the flexoelectric coefficients (FECs).
In materials with simple cubic symmetries, such as cubic perov- the fixed ones but also that the difference between the two values
skites, there are only three independent tensor coefficients: longi- decreases with increasing the height of supercell, suggesting that
tudinal l1111 and transverse l1122 and l1212 . such a difference was the artifact of an insufficiently big supercell.
In order to keep the periodic boundary conditions required by The strain gradients of the fixed and relaxed atoms thus converge,
the first-principles calculations, a special supercell is chosen in and an extrapolation of the FECs allows estimating l3333
which the strain gradient is periodic. Specifically, the strain profile 0.37 6 0.03 nC/m for BaTiO3 and l3333 1.38 6 0.65 nC/m
along the z direction is chosen in the cosine form, for SrTiO3.
z The FECs obtained here are all in the order of nC/m, which is
eðzÞ ¼ emax cos 2p (4.1.10) consistent with the theoretical estimation for a simple dielectric
h [194] and also with recent measurements in SrTiO3 single crystals
and ceramics [198]. These values are also on the same order of
where h is the height of the supercell and emax is the maximum magnitude as the effective Hamiltonian calculations of Maran-
strain in the supercell. The supercell is shown in Fig. 48. The posi- ganti and Sharma [208]. Our calculated FEC of BaTiO3 agrees in
tion of h/4 of the supercell is chosen to calculate FECs where the magnitude with the calculations of Maranganti and Sharma [208]
strain gradient is maximum, and the piezoelectric polarization in and are comparable also to those measured for SrTiO3. However,
this unit cell is zero, because the average strain is zero by con- the longitudinal FEC of BaTiO3 has not yet been directly
struction. Thus, choosing the unit cell at h/4 not only provides the
maximum FxE but also eliminates the piezoelectric contributions.
Given a strain profile (Eq. (4.1.10)), the displacement of each
atom in the supercell is given by
ðz
emax h z
dðzÞ ¼ eðnÞdn ¼ sin 2p (4.1.11)
0 2p h
VX 6 X 6
DE ¼ Cij ei ej (4.1.12) Fig. 50 Size effect of elastic stiffness C11 for BaTiO3 films.
2 i¼1 j¼1 Reprinted from Ref. [180] with the kind permission of Taylor &
Francis.
where V is the volume of the initial unit cell undistorted, DE is the
energy increment between the unit cell energy with and without
the strain e ¼ ðe1 ; e2 ; e3 ; e4 ; e5 ; e6 Þ, and Cij is the elastic stiffness
tensor.
The stoichiometric BaTiO3 film is chosen in the study, with one
face terminated with BaO and the other face terminated with
TiO2. The film thickness is specified by an integer n that corre-
sponds to the number of unit cells along the z direction (out of
plane). Periodic replicas of the films are separated by 16 Å of vac-
uum. For the tetragonal BaTiO3, Cij can be conveniently obtained
by applying different strains ei on an initial unit cell and calculat-
ing the energy difference. The density functional theory calcula-
tion shows that the in-plane lattice constants are size-dependent,
which are 0.3945 nm (n ¼ 2), 0.3968 nm (n ¼ 3), 0.3981 nm
(n ¼ 4), and 0.3989 nm (n ¼ 5) for different films. The in-plane lat-
tice constants increase with increments of film thickness, and it is
very close to that of bulk (0.3999 nm) when film has thickness of
five unit cells.
BaTiO3 is brittle material, and its tensile stiffness is different
from compressive stiffness. Therefore, the tensile and compres-
sive stiffnesses (C11) are fitted separately through Eq. (4.1.12) for Fig. 51 The piezoelectric properties of BaTiO3 bulk and thin
films of n ¼ 2–5, which are shown in Fig. 50. It shows that the ten- films, e33 for bulk and e11 for films of n 5 2 and 4. Reprinted
sile stiffness is smaller than compressive stiffness for a film, but from Ref. [180] with the kind permission of Taylor & Francis.
they are both size-dependent. The elastic stiffness C11 of BaTiO3
film reduces as film thickness increases. It should be noted that environment and thus are effective in characterizing the surface
the compressive stiffness is very close to that of BaTiO3 bulk effect and the behaviors of nanoscale ferroelectrics.
when the thickness reaches n ¼ 5 (i.e., 2.0 nm). Molecular dynamics (MD) method is a widely used atomic-
Piezoelectric stress coefficients are calculated for BaTiO3 bulk level simulation method that has been playing an important role in
and films of n ¼ 2 and 4, as shown in Fig. 51. For the bulk, e33 is solid-state and material science. Since the motion of the particles
17.9 C/m2, which agrees with the previous result of 18.6 C/m2 in are governed by the Newtonian equations, this method is easy to
Ref. [212]. The piezoelectric properties of BaTiO3 films also implement and more efficient than the first-principles method.
show size-dependence: e11 decreases with reduction of film Given the expressions characterizing interatomic interactions, the
thickness. variations of particle positions can be easily determined. A key
point of this method is to find an appropriate potential model to
4.2 Molecular Dynamics Method of Ferroelectrics. With reflect the macroscopic material properties. For a ferroelectric,
rapid progress of nanoscale manufacturing technology, low- one defining characteristic is that the material should have a spon-
dimensional ferroelectrics, such as ultrathin films, superlattices, taneous polarization that can be switched under an electromechan-
nanowires, nanoparticles, and nanorings, have attracted increasing ical load [3]. From the microscopic point of view, ferroelectricity
attention [67,68,181,213–216], due to their intriguing properties stems from the delicate balance between the long-range Coulomb
and prospective applications. In these nanostructures, the role of and short-range repulsive interactions [66]. Such complicated
surface atoms becomes outstanding, and the induced surface behavior is not easy to characterize using the traditional intera-
effect, either in the physical or chemical aspects, usually cannot tomic potential model. To overcome this difficulty, the shell
be fully understood using continuum modeling approach. The model [55–57,59,217,218] was introduced in ferroelectric perov-
atomic-level simulation methods are capable of reflecting the skites, which decomposes each atom into a core and a shell, mod-
reconstruction of surface atoms as well as their different atomic eling an atomic nucleus and electron shell, respectively. This
in the film (Fig. 53(c)). The critical thickness is therefore 2.8 nm, dependent on film thickness. Figure 56(a) shows that the average
which is very close to the results (2.5 nm) derived from experi- polarization decreases as the temperature increases, because the
ments [230] and the first-principles calculations (2.4 nm) [67]. atoms become more active at higher temperature. Figure 56(b)
The electric-field-driven domain evolution and the size- shows an almost linear relation between the Curie temperatures
dependent hysteresis loops were also obtained for BaTiO3 ultra- and the reciprocal of film thicknesses, as indicated by MD simula-
thin film. The domain evolutions in the poling and depoling proc- tions, Landau–Ginzburg theory, and experiments.
esses are illustrated in Figs. 54(c)–54(k) for the 4.4-nm-thick film, The in-plane strain, induced by the mismatch between the sub-
and the variations of polarization are plotted in Figs. 54(a) and strate and the film or defects formed during film deposition, is an
54(b). In the spontaneous state, an 180 deg strip domain, consist- important factor that obviously influences the ferroelectric proper-
ing of periodically alternating positive and negative domains with ties of a thin film. In the MD simulations, the strain loading is usu-
2 to 3 unit cells in width, is clearly shown in the top view of the ally applied by controlling the x-y plane boundary displacements
film (Fig. 54(c)). Such strip domain structure is in reasonable of the films [64,152,220]. Meanwhile, the computational system is
agreement with the experimental observations on ultrathin allowed to expand or contract freely in the z-direction to keep the
BaTiO3 and PbTiO3 films [214,231]. In the electrical poling pro- out-of-plane stress zero.
cess, as shown in Fig. 54(a), three stages are distinguished. Stage The out-of-plane polarization profiles were analyzed for
I: When the electric field is less than 0.4 108 V/m, it is too small BaTiO3 film with thickness h ¼ 4.4 nm under different in-plane
to make any negative domain switch from –z to þz direction, as strains. The results show that in-plane compression strains
shown in Figs. 54(c) and 54(d). As a result, the average polariza- strengthen the polarization of thin films along the z-direction,
tion of the film increases very slowly in Fig. 54(a). Stage II: while in-plane tensile strains decrease it. More interestingly, there
When the electric field reaches the value between exists a “critical strain,” above and below which the polarization
0.41.6 108 V/m, domain switching occurs (Figs. 54(e)–54(h)), distributions present two different stable patterns, similar to Figs.
leading to a sharp increase of the system polarization, as shown in 53(b) and 2(c). The dependence of the hysteresis loop on the in-
stage II of Fig. 54(a). At the end of this stage, the strip domains plane strain was also investigated. For films under the in-plane
have merged into a single domain, as shown in Fig. 54(i). Stage compression strain, the restriction along x- and y-direction makes
III: As the electric field further increases above 1.6 108 V/m, the domain switching difficult, so the area of the corresponding
there is no domain switching anymore and the system polarization hysteresis loop is larger as compared with the free-standing ones.
increases slowly with the electric field again. When the external In other words, a larger in-plane compression strain induces a
electric field decreases to zero, the entire polarization is still finite greater remanent polarization and coercive field. By contrast, the
positive, which is just the remanent polarization Pr. The polariza- tensile strain has an opposite effect. The external strain also influ-
tion falls down quickly when the electric field decreases further. ences the Curie temperature as predicted by the MD simulations,
Once the electric field reaches the coercive field Ec (about such that a tensile in-plane strain decreases the Curie temperature
0.29 108 V/m), the entire polarization becomes zero again, as while a compressive one increases it.
shown in Figs. 54(a) and 54(k).
The size-dependent hysteresis loops of the ultrathin BaTiO3 Ferroelectric nanowires. Different from the ultrathin film, the
film are shown in Fig. 55(a). For the 3.6-nm-thick film, the hyster- nanowire has free surfaces in two principle directions and thus
esis loop is obviously smaller than that of 4.4-nm-thick film, may have more evident surface effect than the ultrathin film. In
which is also reflected by the value of corresponding remanent order to explore the size dependence of ferroelectric behaviors,
polarization and the coercive field. Moreover, for the film thinner the ferroelectric nanowires with different diameters (i.e., consist-
than the critical thickness, such as 2.8 nm, there is no hysteresis ing of n n 7 unit cells (2 n 10)) were numerically studied.
loop and the polarization has nearly linear relation with the exter- Periodical boundary conditions are applied in the axial direction,
nal electric field. Interestingly, a quasilinear relation of remanent and a surface-free boundary condition is applied in the other two
polarization Pr and coercive field Ec with the reciprocal film thick- directions. For a nanowire with a small diameter, the polarization
ness was revealed, as shown in Figs. 55(b) and 55(c). An extrapo- distributions calculated by shell model MD simulations were care-
lation of remanent polarization at 1/h ¼ 0 gives 17.5 lC/cm2 that fully tested by comparison with the ab initio results, which dem-
is in reasonable agreement with the experiment measurements of onstrated relatively good accordance both qualitatively and
remnant polarization for rhombohedral BaTiO3 bulk material. quantitatively [61].
Besides the spontaneous polarization and hysteresis loop, The size dependence of polarization distribution was studied by
another important quantity (i.e., the Curie temperature) is also the MD method, and the relation between the average polarization
components and the nanowire diameter is shown in Fig. 57. The lengths, the stable polarization in the x-y plane emerges, while the
polarization distribution significantly depends on the wire diame- axial component is still near zero (i.e., Pz 0). The x-y plane
ter, and three scenarios can be distinguished. Scenario I: When polarization exhibits a quasiaxisymmetric radial distribution. Sce-
the diameter is only one cell length, the nanowire is too thin to nario III: When the diameter is larger than two cell lengths, the
support any ferroelectric distortion along any directions. There- magnitude of polarization along the x-, y-, and z-directions are on
fore, the spontaneous polarization is almost zero everywhere (i.e., the same order and not negligible. The z-component of polariza-
Px ; Py ; Pz 0). Scenario II: When the diameter is two cell tion Pz over the cross section is along the same direction,
indicating that a monodomain is formed, which is consistent with between dcr1 and dcr2. When the nanowire is thicker than dcr2, the
the experimental observation [192]. The axial polarization distri- polarization has both radial and axial components and the axial
bution is similar to that of scenario II. polarization is larger at the surface region and smaller at the inner
According to the three scenarios of different polarization distri- region of the nanowire. Such nonuniform polarization distribution
butions, two critical diameters of the BaTiO3 nanowire, can be understood by a lattice mismatch mechanism [232].
dcr1 ¼ 0.8 nm and dcr2 ¼ 1.2 nm, can be identified. No polarization As shown in Fig. 58, both the size and the shape of the hystere-
exists when the diameter is less than dcr1. The radial polarization sis loop vary with the diameter of the nanowires. When the diame-
appears while no axial polarization exists when the diameter is ter d 0:8nm, there is no loop, since nearly no axial remnant
These two limitations can be overcome if the atomic-scale finite Etot is composed of two parts: the inner potential Uinner character-
element method (AFEM) [233,234] is adopted to determine the izing the atomic interactions between different particles and the
atomic positions at the equilibrium state, instead of MD method. external potential Uextenner induced by the applied physical field,
AFEM is a molecular statics method that solves the energy mini- such as electric field or stress field. The total energy Etot can be
mization problem in the finite element framework. Thus, the parti- given by
cle mass is not needed in the numerical calculation, and more
importantly, the CPU time of this method increases almost line- X
N
arly with the atom number [233,234]. Etot ðxÞ ¼ Uinner ðxÞ þ Uexterner ðxÞ ¼ Uinner ðxÞ i xi ;
F
The combination of AFEM and shell model leads to an effec- i¼1
tive atomic-level computational method of ferroelectric perov- (4.3.1)
skites, as we will introduce in this section. In Sec. 4.3.2, we first
present the formulation of electromechanical AFEM and demon-
strate its applicability in simulations of ferroelectrics and its where x is an integrated coordinate vector of all particles and xi is
the coordinate vector of particle i. F is the generalized force
advantage in computational efficiency. This developed electrome-
chanical AFEM is further utilized to analyze the nanoscale do- induced by an external conserved field, such as the electrome-
main configurations of BaTiO3 and the electric field–driven chanical field. The stabilization of the atomic system requires
evolutions in Sec. 4.3.3. minimization of the total energy, which yields
theoretical and experimental studies [237–239], the presented do- the polarization of variant i to the electric field (E) direction anti-
main wall thickness is nearly an order larger than 0.4 nm. In order clockwise (180 deg hi 180 deg). The domain evolution
to obtain a larger domain wall thickness from AFEM simulations, results are demonstrated in Fig. 62 for different electrical loading
a more reasonable set of potential parameters is required. In the directions. When h1 ¼ 0 deg or 45 deg, the polarization always
vortex domain structure, the clockwise and anticlockwise vortexes remains unchanged, no matter how large the electric field is,
are distributed alternatively along the 645 deg directions with because domain switching would increase the angle between the
respect to the x-axis, similar to the experimental results of Tsou polarization and electric field and thus enhances the system
et al. [240], who observed vortex domain pattern in tetragonal energy. When h1 ¼ 90 deg, as the electric field approaches the
BaTiO3. Interestingly, besides the vortex polarization structure, coercive field, a new domain parallel with E first nucleates auto-
the streamline polarization structure also exists in the vortex do- matically in the original domain, and then the domain wall moves
main configuration. Since the polarizations along opposite direc- laterally until the domain completely switches to the direction of
tions counteract each other in the center of the streamline electric field. When h1 ¼ 135 deg, the domain evolution process is
structure, this domain configuration also satisfies the continuity similar to that of h1 ¼ 90 deg. Interestingly, when the domain is
condition of normal polarization at the domain wall, thus avoiding applied to an opposite electric field, the domain undergoes a con-
the emergence of large depolarization field. In the other ab initio secutive two-step, 90 deg domain-switching process. The original
calculations, single vortex structure is revealed in the PZT nano- domain first switches to variant 4 via nucleation and movement of
wire [68] and PbTiO3 nanowire [241]. Hong et al. [177] initially horizontal 90 deg domain wall (or switches to variant 2 with
reported the existence of streamline polarization pattern in nucleation and movement of vertical 90 deg domain wall), and
BaTiO3 nanowire by first-principles method. Different from then the whole domain switches simultaneously to variant 3.
these polarization structures, the domain structure as shown in Notice that a similar two-step domain-switching process has been
Fig. 61(c) has a coexistence of both vortex and streamline polar- observed in the in situ experiment on tetragonal BaTiO3 single
ization configurations. crystal using polarized light microscopy [242]. From the simula-
By inputting the stable atomic coordinates of different domain tions, it can be found that, when |h| > 45 deg, the domain switch-
configurations as the initial state and applying electric fields along ing can be triggered to reduce the system energy as long as the
different directions, the electric field–driven domain evolutions external electric field surpasses the coercive field.
can be simulated. As an example, we introduce the various evolu- For the 90 deg and vortex domains, the electric field–driven
tion paths of single domain structure. The atomic systems ranging evolution diagrams are more complicated than that of single do-
from 6 6 6 to 20 20 6 are considered, and the simulations main and can be found in Zhang et al. [88]. Besides the qualitative
show that the variation of domain size almost has no influence on evolution patterns, the variations of overall polarizations with the
the domain evolution sequences. Thus, only the simulation results electric field can be also obtained. The P–E curves during the
for the atomic system of 12 12 6 are exemplified to demon- electrical poling and depoling processes were calculated for vor-
strate the domain evolution diagrams. tex domain as the initial pattern. The calculations indicate that, for
Define the unit cell with two positive polarization components two different loading directions (horizontal and diagonal),
(Px > 0 and Py > 0) as variant 1, the cell with Px > 0 and Py < 0 as although the microscopic domain evolution processes are differ-
variant 2, the cell with Px < 0 and Py < 0 as variant 3, and the cell ent, the variation trends of macroscopic polarization responses are
with Px < 0 and Py > 0 as variant 4. Denote hi the angle rotating similar. During the poling process, when the electric field is small
enough so that the domain switching cannot be activated, the [81,120–123], where the misfit strain due to lattice mismatch
polarization increases slowly and almost linearly with the electric between film and substrate is related to the lattice constants of
field, due to the dielectric effect. As the electric field exceeds the both materials. The elastic energy is generally expressed as a
coercive field, the domain wall starts moving and the polarization function of the stiffness tensor for anisotropic materials (depend-
increases quickly until the domain is poled into the saturated state. ing on the phases). Since some components of the stiffness tensor
During the depoling process, when the magnitude of negative are not available or not easy to measure via experiments, first-
electric field is smaller than the coercive field, the domain config- principles calculations could provide an effective alternative route
uration remains unchanged and the polarization decreases rela- to determine the entire stiffness tensor. Besides, the double-well
tively slowly. As the electric field reaches the coercive value, the free energy function (in terms of polarization or electric displace-
domain undergoes a drastic change. ment), as adopted in phase field method, can also be fitted from
first-principles calculations with the recently developed fixed
5 Connections Among Electromechanically Coupled polarization method [243] and fixed electric displacement method
[244–246]. The output from first-principles method can also be
Computational Methods at Different Scales used as input for MD method to obtain the proper potential. As we
The various computational methods introduced above are only know, good potential is essential for the accurate MD simulation,
applicable within their own scales, because of the different and for different materials, the potential has different forms or dif-
assumptions, formulations, and the computational complexity. ferent coefficients. After assuming the potential form for specific
However, in some special cases, simple interfaces or connections materials, its coefficients can be fitted by comparing the properties
between the computational methods at different scales are possi- (free energy, spontaneous distortion, etc.) at different phases
ble (e.g., by passing on the material properties calculated at a and strained states from first-principles calculations and MD
lower-scale method to serve as input parameters for a higher-scale simulations.
method). In the following, we take some representative examples The mobility coefficient of domain wall movement required in
to illustrate such connections among the computational methods phase field method can be possibly obtained by MD simulations.
of three different scales. McCash et al. [247] used classical molecular dynamics with the
The lattice constant, elastic modulus, polarization, piezoelectric force field derived from a first-principles–based effective Hamil-
coefficient, flexoelectric coefficient, and some other fundamental tonian to investigate the intrinsic dynamics of ultrafast polariza-
material properties can be obtained from first-principles calcula- tion reversal in ferroelectric nanowires. They traced the time
tion and passed to the mesoscale methods. For example, the phase evolution of polarization under applied electric field and found
field method was widely adopted to explore the domain patterns that, in defect-free nanowires, the polarization reversal can occur
of ferroelectric ultrathin films grown on a heterogeneous substrate within picoseconds. Similar analyses can be extended to
DownloadedViewFrom:
publicationhttp://appliedmechanicsreviews.asmedigitalcollection.asme.org/
stats on 01/06/2014 Terms of Use: http://asme.org/terms