Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Acta Materialia 55 (2007) 233–241

www.actamat-journals.com

Simulation of hexagonal–orthorhombic phase transformation


in polycrystals
a,*
Y.-W. Cui , T. Koyama b, I. Ohnuma a, K. Oikawa a, R. Kainuma c, K. Ishida a

a
Department of Materials Science, Graduate School of Engineering, Tohoku University, Sendai 980-8579, Japan
b
Computational Materials Science Center (CMSC), National Institute for Materials Science (NIMS), 1-2-1 Sengen, Tsukuba, Ibaraki 305-0044, Japan
c
Institute of Multidisciplinary Research for Advanced Materials (IMRAM), Tohoku University, Sendai 980-8577, Japan

Received 20 April 2006; received in revised form 21 July 2006; accepted 31 July 2006
Available online 12 October 2006

Abstract

The microstructural evolution during hexagonal–orthorhombic (HO) phase transformation in polycrystals was modeled by means of
a modeling technique combining the strain-only theory and phase-field modeling. Specifically, the strain-only theory is used to describe
the HO transformation, whereas phase-field modeling is employed to capture the evolution of the grain orientation in polycrystals. The
modeling appropriately describes important aspects of the HO transformation behavior of polycrystals, such as the phase transition
sequence, the lamellar-like twins in two orthogonal directions, streamers and striated variants. Furthermore, it predicts the accommo-
dation of the HO polycrystal, in which the domain morphologies are rearranged in distinct manners for variously oriented grains as a
response to uniaxial loading, thereby demonstrating the fact that the morphological evolution in polycrystals is a correlated collective
process.
 2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Computer simulation; Phase transformation; Polycrystals; Microstructure; Hexagonal–orthorhombic transformation

1. Introduction exist in three orientation variants with identical energy.


During the transformation, the strain field in the vicinity
Over the years, increasing interest has been focused on of the intersection of twin boundaries is generally relaxed
the modeling of microstructural evolution during such via the creation and motion of dislocations. However,
diffusionless displacive phase transitions as the martens- the low dislocation mobility, occurring in certain ordered
itic transformation in oxides [1], shape memory alloys alloys and minerals, hardly results in relaxation, and as a
[2–6] or ferroelastics [7–9], and the order–disorder trans- result, a linear defect, termed a disclination, is introduced
formation in ordered alloys [10–12]. Among such phe- at the intersection [13,14]. The intricate HO domain
nomena, hexagonal–orthorhombic (HO) transformation, pattern is believed to be closely related to this linear
as well as similar transformations which lead to intricate defect; examples are Ti3Al–Nb [15,16], Mg–Cd alloys
and unique domain structures of the product variants in [13,14,17,18], mineral Mg-cordierite [19], Ta4N [20] and
certain materials, appear to be uncommon. For symme- others [21,22].
try reduction from a hexagonal to an orthorhombic Due to the fact the properties are isotropic in the basal
structure, a low symmetry orthorhombic structure can plane of a hexagonal structure, HO material is ideal for
two-dimensional (2D) modeling of morphological evolution
without losing the essential physics. Fig. 1 illustrates that
*
Corresponding author. Tel.: +81 22 7953650; fax: +81 22 7957323. the HO martensitic transformation produces three product
E-mail address: ycui@material.tohoku.ac.jp (Y.-W. Cui). variants on the basal plane, whereas for the illustration of

1359-6454/$30.00  2006 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2006.07.026
234 Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241

Martensitic transformation formation driven by ordering [23] or martensitic transfor-


y mation driven by distortion [9]. One reason for this
choice is that experiments offer no real applicable
+ + parameters.

x 2. Model
habit plane variant 1 variant 2 variant 3
orientation +15o,-75o +75o,-15o +45o,-45o 2.1. Free energy
Fig. 1. Schematic illustration of the hexagonal-to-orthorhombic martens-
itic transformation on the basal plane producing three different variants. To reflect the symmetry breaking in the HO phase trans-
formation in 2D, derivation of the Landau-type expansion
of the elastic strain energy in the strain order parameter
the HO order–disorder transformation, refer to Ref. [23]. (OP) is required. The strain OPs are the deviatoric strain,
Various types of modeling have been conducted to under- e2 = (exx  eyy)/2, and the shear strain, e3 = exy; the non-
stand the HO by either the strain-only theory or the OP (or secondary strain OP) is the bulk dilatation strain,
phase-field method in the overdamped limit [8,11,23–26]. e1 = (exx + eyy)/2, all of which are components of the sym-
The strain-only theory is thought to be particularly applica- metry-adapted elastic tensors. The appropriate free energy
ble to group–subgroup phase transformations, which can be functional is written as
displacive, order–disorder or mixtures, but it requires that Z
the domain structure be dominated by elastic strain energy. F ¼ dr½fElastic þ fGrain þ fLoad  ð1Þ
A phase-field method can deal with both improper and
proper HO phase transformations; however, the free energy where fElastic is the elastic free energy, fGrain is the free en-
expansion and numerical solution are complex in practice ergy due to the grain orientation and fLoad is the free energy
[11,23], and furthermore, applying them to polycrystals is contribution due to the applied load. Keeping only essen-
not straightforward. tial terms and describing a triply degenerate product phase,
Most crystalline materials of commercial interest are the free energy due to the elastic strain is yielded as
polycrystalline and consist of a very large number of dif-   
A1 2 A2 2 B2
ferently orientated grains separated by grain boundaries. fElastic ¼ e1 þ ðe2 þ e23 Þ  ðe32  3e2 e23 Þ
2 2 3
The evolution of domain morphology within an individ-  n h io
C2 2 2 je 2 2
ual grain is constrained by the neighboring grains; there- þ ðe2 þ e23 Þ þ ðre2 Þ þ ðre3 Þ ; ð2Þ
fore, grain structure plays an important role in the 4 2
development of phase transformation, microstructural where the linear elastic constants, A1 and A2, are the bulk
evolution and macroscopic parameters. The phase-field elastic modulus with 2D hydrostatic constraint and the
method is capable of modeling most of key factors shear modulus, respectively, both of which vary with tem-
involved in polycrystals [27,28], but, as recently pointed perature [8]. B2 and C2 reflect the non-linearity of the sys-
out [6,29,30], it fails to capture the character of the meta- tem, and C2 must be positive to ensure global stability. je is
stability of a polycrystalline state. Thus, a new modeling the gradient energy coefficient. Based on kinematic consid-
approach has been developed to investigate the cubic– erations, three components of the strain tensors are con-
tetragonal transformation of shape memory polycrystals strained to the displacement field, and thus not all the
by integrating the strain-only theory with the phase-field components are independent when they vary spatially.
model [6,29]. The proposed approach allows the orienta- Applying the 2D Saint–Venant compatibility relation in
pffiffi
tion of the grains to evolve. As the elastic energy has k 2 k 2 8k k
the Fourier space, ^e1 ¼ x2 y2 ^e2 ð~
kÞ þ 2 x 2y ^e3 ð~
k x þk y
kÞ, the non-
k x þk y
been theoretically verified to be essential for the domain
OP e1 can be eliminated by expressing it in terms of e2
patterns during the coherent HO transformation [23,26],
and e3 [4].
it is reasonable to use this approach to understand the
The orientational contribution to the free energy, quan-
behavior of the HO polycrystals. More specifically, the
tifying the excess free energy due to spatial inhomogeneities
strain-only theory is used to describe the HO transforma-
tion, whereas the phase-field model is employed to cap- in crystal orientation, can be written as a function of the
ture the evolution of the grain orientation. The purpose multicomponent grain orientation OPs, gi:
Q h
of the present study was, therefore, to extend application X a1 a2 3 a3 4 i a4 X X 2 2
of the proposed approach to description of more complex fGrain ¼ g2i þ g þ gi þ gi gj
i¼1
2 3 i 4 2
heterogeneous configurations in the HO polycrystals and X jg
to solve the problems involved by using a new numerical þ ðrgi Þ2 ð3Þ
scheme. As was done by Wen et al. [11,23] and Curnoe 2
and Jacobs [8], rather than working with a specific mate- with the coefficients a1, a2 < 0 and a3, a4 > 0 to ensure a
rial, we used a generic model material which undergoes a potential with Q degenerate minima, (ge, 0, . . . , 0),
proper HO transformation, either order–disorder trans- (0, ge, . . . , 0), etc. up to (0, 0, . . . , ge) (ge > 0), corresponding
Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241 235

to Q grain orientations [29]. The grain orientation angle verting the PDE in Eq. (7) to a dimensionless form and
can accordingly be evaluated by then taking the Fourier transform with respect to space
P
Q variable, we discretize the system in space with spectral
½ði  1Þ=ðQ  1Þgi accuracy as follows:
hðg; rÞ ¼ hm i¼1 ; o^lð~
-; sÞ  
P
Q
¼ ~-2 l
^ð~
-; sÞ  N ^ e;g ½lðrÞ ¼ L^
l þ N ð^
l; sÞ ð8Þ
gi os
i¼1
As such, the right-hand side is split into a linear part L^ l
where hm is the maximum p grain
ffiffiffi orientation angle. Defining
pffiffiffi and a non-linear part N ð^ l; tÞ. We introduce a fast and effi-
e1 = e1, e2 ¼ e2 cosð2hÞ þ 2e3 sinð2hÞ and e3 ¼ ðe2 = 2Þ
cient scheme, termed the fourth-order Runge–Kutta ETD
sinð2hÞ þ e3 cosð2hÞ, the elastic energy describing the HO
time scheme (ETDRK4) [33–35], to solve Eq. (8). This
transformation in polycrystals is given by
  new ETDRK4 scheme was developed particularly to over-
A1 2 come the stiffness and non-linearity emerging from spec-
fElastic ¼ e
2 1 trally spatial discretization, thus providing a fourth-order
  temporal discretization.
A2 2 B2 C2 2
þ ðe2 þ e23 Þ  ðe32  3e2 e23 Þ þ ðe22 þ e23 Þ A2 is a temperature-like parameter, and A2 = 0 is the
2 3 4
nj h io stability limit of the hexagonal phase. Throughout this
e 2 2
þ ðre2 Þ þ ðre3 Þ paper A2 = 20, i.e. the quench depth is fixed. Other fixed
2
parameters are B2 = 3 · 103, C2 = 2 · 106, a2 = 1, a3 = 1
ð4Þ
and a4 = 2, and a1 is first given as 1 but is abruptly chan-
with a modified constraint by ged to 7.5 after time 100. For simplicity, we assume all
pffiffiffi kinetic coefficients to be constant, Ll = 1. We choose
k 2x  k 2y 8k x k y ^ ~
^e1 ¼ 2 ^ ~
U ðkÞ þ 2 U3 ðkÞ; Q = 5 and hm = p/6, thus describing five grain orientations.
2 2
kx þ ky k x þ k 2y The remaining parameters, A1, je, and jg, are adjustable to
where U ^ 2 and U ^ 3 are the Fourier transforms of produce appropriate configuration descriptions. The initial
pffiffiffi pffiffiffi
e2 cosð2hÞ  2e3 sinð2hÞ and ðe2 = 2Þ sinð2hÞ þ e3 cosð2hÞ, configurations for es = 2, 3 and gi are generated by using a
respectively [29]. random initial condition subject to periodic boundary con-
Note that, since we are interested in uniaxial loading, the ditions. In all simulations, we take 400 · 400 systems into
free energy contribution due to the applied load rex in the account and use a time step Ds = 0.4 in the simulations
[10] direction can be yielded as for a single crystal and 0.2 for a polycrystal.
pffiffiffi
fLoad ¼ rex ðe1 þ e2 Þ= 2 ð5Þ 3. Numerical results and discussion
for a single crystal and
 pffiffiffi  pffiffiffi 3.1. Tweed formation
fLoad ¼ rex e1 þ e2 cosð2hÞ  2e3 sinð2hÞ = 2 ð6Þ
Quenching the system toward a low temperature, the
for a polycrystal. local inhomogeneity of the elastic strain triggers the onset

2.2. Dynamics and parameters

The rate of the OP change can be approximated by the


functional derivative of the free energy of the system with
respect to the OP, namely by time-dependent Ginzburg–
Landau (TDGL) dynamics:
olðr; tÞ dF
¼ Ll ; ð7Þ
ot dlðr; tÞ

where l is either of the OPs, i.e. gi or es = 2, 3. Ordinary


TDGL dynamics are essentially an overdamped description
of the dynamics, corresponding to a correct limit of full
underdamped dynamics [9,31]. It should be emphasized
that the overdamped displacement equation of motion
developed by Curnoe and Jacobs [8,32] was later proved
to be essentially the same as TDGL-like dynamics [9].
Over a periodic domain, the spectral method is believed
to enable very fast and accurate spatial accuracy of the
TDGL-type partial differential equations (PDE) [33]. Con- Fig. 2. The simulated tweed pattern with A1 = 100 and A2 = 20.
236 Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241

Fig. 3. Morphological evolution of single crystals under uniaxial tension in the [1 0] direction with (a) and (b) A1 = 10; (c) and (d) A1 = 100; (e) and (f)
A1 = 500. Left column: the microstructure prior to loading. Right column: loading for a time of 400.

of a fine-scale assembly of non-uniformly distorted ation of tweed to a self-accommodated twinning state, as


domains, termed tweed. Mathematically, it is reflected by KAlSi1 was revealed to stabilize the tweed texture in a
the expression of e1 in terms of e2 and e3, such that the natural cordierite [36].
anisotropy of long-range elastic interaction is induced
[3,4], thus generating a typical cross-hatched tweed pattern, 3.2. Domain evolution of a single crystal under an applied
as seen in Fig. 2. The tweed subjected to Mg-cordierite dur- loading
ing the HO transformation was imaged experimentally by
Putnis et al. in Fig. 2(b)–(e) [19]. During this process, both As the system continues to evolve, the tweed progres-
the e2 and e3 strains remain small, and thus the system sively gives way to an orthorhombic product. In the left
retains its hexagonal structure as a whole. The tweed for- column of Fig. 3, the spatial distributions of the shear
mation provides a partial reduction of the overall elastic strain for single crystals with A1 = 10, 100 and 500 at times
strain. However, if there is a substantial concentration of 600, 1000 and 2000 from top to down are presented. The
a certain stabilizer, such as an impurity or defect, it may snapshots make it possible to discern the characteristic
act as a pinning site, thus preventing or delaying the relax- domain morphology, which consists of three orthorhombic
Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241 237

variants. Each of the three variants has two orientations,


the optimal orientations of which lie in different directions:
p/12 and 5p/12 for the red variant1, 5p/12 and p/12 for
the blue one, and p/4 and p/4 for the green one. These
snapshots also show that the domain morphology and its
evolution in a single crystal vary with the stiffness of the
system (see Refs. [8,23] for more details). In this subsection,
we provide an insight into the morphological evolution
subjected to applied stress and explore its mesoscopic
origin.
In the very soft system with A1 = 10 (see Fig. 3(a)),
when applying moderate uniaxial tension in the [1 0]
direction (or equivalently a compression in the [0 1] direc-
tion), the system accommodates itself by ruling out the
green variant to minimize the free energy of system,
thereby developing into a ‘‘two-variant structure’’ (or
lamellar twin) around time 80 (see Fig. 3(b)). The two
favored variants are orthogonal and lie approximately
along the x- and y-axes of the system, respectively. This
structure is in good agreement with the experimental
observation in Figs. 3 and 7 of Ref. [37]. In contrast, uni-
axial compression in the [1 0] direction suppresses two of
the three variants, but does not affect the green variant.
The reason why tension and compression manipulate
the orthorhombic domain in distinct manners can be
readily understood by studying the change of the overall
structure of the potential in the e2–e3 plane by an applied
load. Note that adding Eq. (5) to Eq. (2) gives the latter
a term linear to both the deviatoric strain and the applied
load. When the applied load is sufficiently large, the
potential as a function of e2 and e3 changes its overall
structure. In the case when tension is applied, three glo-
bal minima, evenly spaced at the same distance from the
center in the e2–e3 plane (Fig. 4(a)), are reduced to two,
and there is a shift in positions (see Fig. 4(b)). From a
microstructural perspective, a two-variant structure will
accordingly be built. In the case of compression, how-
ever, only one minimum remains, as shown in Fig. 4(c),
and thus the system will develop into a single variant
pattern.
Fig. 3(c) and (d) shows that the soft system with
A1 = 100 nearly evolves into a lamellar twin structure in
response to the same [10] tension at time 80, but a very
small amount of unfavored variant persists, thus indicating
that complete removal requires a longer time. In the stiff
system with A1 = 500 (see Fig. 3(f)), more unfavored vari-
ants survive the same loading. As revealed by the previous
modeling for HO ferroelastics [8], the domain walls in a
very soft system could apparently deviate from the optimal
directions without generating remarkable energy from
additional dilatation strain, whereas the domain walls are Fig. 4. Contour plots of the elastic energy: (a) in the absence of external
highly constrained in a stiff system. For exactly the same loading; (b) subjected to uniaxial tension; (c) subjected to uniaxial
compression.
reason, the different systems will exhibit their own

responses to applied load, and the stiffer the system, the lar-
1
For interpretation of the references to colour in this figure, the reader ger the applied tension and the longer the time required to
is referred to the web version of this article. induce a lamellar structure.
238 Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241

3.3. Domain morphology of polycrystals icant domain morphology change of polycrystals may
result from changing A2. However, with the parameters
Fig. 5 shows the snapshots of morphology of the HO we used, A2 is not so important in determining the domain
polycrystals with increasing stiffness from top to bottom. morphology as A1. Future work on this topic will be done
The snapshots in the left column are encoded by a combi- soon.
nation of the shear strain e3 and the grain orientation
angle, whereas those in the right column are simply repre- 3.4. Effect of applied load on polycrystals
sented by the spatial grain orientation angle. Fig. 5(a) and
(b) shows the morphological snapshots for a very soft sys- To gain further insight into the effect of applied load
tem at time 600. The parameters are A1 = 10, je = 0.15 and on morphological evolution, in this subsection we exam-
jg = 0.2, and a1 is suddenly decreased to 7.5 at time 100 ine the effect of applied load on an HO polycrystal. When
to stop rapid grain growth. Note that the characteristic applying moderate compression in the [1 0] direction to
domain morphology in the grains with orientation h = 0 the very soft system shown in Fig. 5(a), the system accom-
remains unchanged from that presented for the single crys- modates it by rearranging the domain morphology in a
tal, referring to the grains with the darkest contrast in different way from that in the single crystal case. Namely,
Fig. 5(a). However, in those grains with orientation the lamellar twins of the green and blue variants are fairly
h > 0, the orthorhombic variants self-accommodate by well formed in the grains with h = p/6, whereas the green
rotating their optimal orientation at certain angles to align variants grow at the expense of the unfavored red and
with their respective grain orientations, in particular, by blue variants in the grains with h = 0 (see Fig. 6(a)). As
p/6 in the grains with orientation h = p/6 (the grains with seen from Fig. 6(c), a larger compression gives rise to a
the brightest contrast in Fig. 5(a)). More specifically, the well-formed lamellar twin structure in the former grains,
red variants1 change from p/12 (5p/12) to p/4 (p/4) while the latter grains take on a single variant pattern.
and the blue ones change from 5p/12 (p/12) to 5p/12 In the case when uniaxial tension is loaded, the system
(p/12), whereas the green ones change from p/4 (p/4) to gives an opposite response, namely, a single red-variant
5p/12 (p/12). This behavior can be readily recognized state in the former grains, and lamellar twins of the red
by a comparison of the three encircled star disclinations. and blues variants in the latter grains (see Fig. 6(b) and
Meanwhile, the lamellar twins in two orthogonal directions (d)).
can be identified in those regions having the lower amounts In the stiff system, either uniaxial compression or ten-
of the red variant. Note that this predicted structure closely sion manifests itself by selective growth in the differently
resembles a transmission electron microscopy (TEM) oriented grains similar to those just observed in the soft
observation of the polycrystalline cordierite in Fig. 2(h) system (see Fig. 6(e)–(h)). However, closer inspection sug-
of Ref. [19], which was also considered to be a soft material gests that the stiff system responds weakly as compared
[8]. with the soft system, and that the complex patterns are
In the soft system with A1 = 100, the optimal directions more apparent in the former.
of the orthorhombic variants vary from grain to grain as in
the previous case, equally suggested by three star disclina- 4. Summary and conclusion
tions highlighted in Fig. 5(c). The snapshots of Fig. 5(c)
and (d) were taken at time 1000, with the other parameters Use of integrated modeling between the strain-only the-
being je = 0.4 and jg = 0.2. The stiff systems evolve to ory and phase-field model enables prediction of the hetero-
relaxed configurations very slowly. The two snapshots, geneous configuration emerging in the HO polycrystal and
Fig. 5(e) and (f), were obtained from the stiff system with its response to applied load. The modeling appropriately
A1 = 500 at time 2000 by using the parameters je = 0.5 describes important factors involved in the HO transfor-
and jg = 0.15. The result is more complicated domain mor- mation behavior of polycrystals.
phologies. For example, the large domains penetrated by
streamers and striated variants form off the optimal direc- 1. The commonly seen tweed formation is observed in the
tions to reduce the larger dilatation stress produced in the very early stage of the modeling, and during the process,
stiff systems. As the stiffness increases further, the systems both OP strains remain small, thus indicting that the
behave in a similar manner. See Fig. 5(g) and (h) for the system retains its hexagonal structure as a whole.
very stiff system with A1 = 1000, je = 0.6 and jg = 0.15 2. The single crystal gives its response to uniaxial tension
at time 2000. No complicated structures are observed in in the [1 0] direction by developing into orthorhombic
the very soft system; therefore, as in the case of a single lamellar twins in two orthogonal directions, while it
crystal [8], they could be transitory. Nevertheless, the poly- responds to uniaxial compression in the [1 0] direction
crystals require a longer time to attain the relaxed by developing a single variant structure. A stiffer system,
configuration. a greater tension and a longer time are required to
Since the relative stiffness, i.e. the ratio of A1 to the cur- induce a lamellar twin structure. The mesoscopic origin
vature of the orthorhombic minima, also depends on the is interpreted to be the change of the overall structure of
temperature-like A2, just like the single crystal [8], a signif- the elastic energy by an applied load.
Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241 239

Fig. 5. Morphological snapshots of polycrystals during the HO transformation with (a) and (b) A1 = 10, je = 0.15, jg = 0.2 at time 600; (c) and (d)
A1 = 100, je = 0.4, jg = 0.2 at time 1000; (e) and (f) A1 = 500, je = 0.5, jg = 0.15 at time 2000; (g) and (h) A1 = 1000, je = 0.6, jg = 0.15 at time 2000. Left
column: the deviatoric strain encoded with the grain orientation. Right column: the spatial distribution of the grain orientation angle. (For interpretation
of the references to colour in this figure legend, the reader is referred to the web version of this article.)
240 Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241

Fig. 6. Strain morphology of polycrystals under the applied stress loading for a time of 5 with (a)–(d) A1 = 10; (e)–(h) A1 = 500. The stress is applied in the
[1 0] direction. Left column: uniaxial compression. Right column: uniaxial tension.

3. The morphological evolution of the polycrystals is a cor- systems, lamellar-like twins in two orthogonal directions
related collective process. The optimal directions of the are present in those regions having lower amounts of the
orthorhombic variants vary from grain to grain to align red variants. The stiff systems evolve to relaxed configu-
with the corresponding grain orientations. In very soft rations very slowly. The result is more complicated
Y.-W. Cui et al. / Acta Materialia 55 (2007) 233–241 241

structures, e.g. large domains penetrated by streamers [5] Artemev A, Jin Y, Khachaturyan AG. Acta Mater 2001;49:1165.
and striated variants, which form off the optimal direc- [6] Ahluwalia R, Lookman T, Saxena A. Phys Rev Lett 2003;91:
055501.
tions to reduce the strong dilatation stress in the stiff sys- [7] Jacobs AE. Phys Rev B 2000;61:6587.
tems. All the findings indicate that the heterogeneous [8] Curnoe SH, Jacobs AE. Phys Rev B 2001;63:094110.
domain morphology strikingly depends on the stiffness [9] Lookman T, Shenoy SR, Rasmussen KO, Saxena A, Bishop AR.
of the system. Phys Rev B 2003;67:024114.
4. When subjected to uniaxial compression in the [1 0] [10] Wang Y, Banerjee D, Su CC, Khachaturyan AG. Acta Mater
1998;46:2983.
direction, the polycrystal accommodates the compres- [11] Wen YH, Wang Y, Chen LQ. Acta Mater 1999;47:4375.
sion by rearranging the domain morphology in such a [12] Wen YH, Chen LQ, Hazzledine PM, Wang Y. Acta Mater
different way from the single crystal case that the lamel- 2001;49:2341.
lar twins of the green and blue variants are well formed [13] Kitano Y, Kifune K, Komura Y. J Phys (Paris) 1988;49:C5–201.
in the grains with h = p/6, whereas the green variants [14] Kitano Y, Kifune K. Ultramicroscopy 1991;39:279.
[15] Bendersky LA, Boettinger WJ. J Res Natn Inst Stand Technol
grow at the expense of the unfavored red and blue vari- 1993;98:585.
ants in the grains with h = 0 toward a single variant pat- [16] Pierron X, De Graff M, Thompson AW. Phil Mag A 1998;77:
tern. When uniaxial tension is loaded, the system gives 1399.
an opposite response, namely, a single red-variant state [17] Kitano Y, Kifune K, Yanagi M, Komura Y. Trans JIM
in the former grains and lamellar twins of the red and 1986;27(Suppl.):181.
[18] Sinclair R, Dutkiewicz J. Acta Metall 1977;25:235.
blues variants in the latter grains. Closer inspection sug- [19] Putnis A, Salje E, Redfern SAT, Fyfe CA, Stobl H. Phys Chem Miner
gests that the stiff system responds weakly compared 1987;14:446.
with the soft system, the complex transitory patterns [20] Vicens J, Delavignette F. Phys Status Solidi A 1976;33:497.
being more apparent in the former. [21] Manolikas C, Amelinckx S. Phys Status Solidi A 1980;60:607.
[22] Manolikas C, Amelinckx S. Phys Status Solidi A 1980;61:179.
[23] Wen YH, Wang Y, Chen LQ. Phil Mag A 2000;80:1967.
Acknowledgements [24] Wen YH, Wang Y, Bendersky LA, Chen LQ. Acta Mater
2000;48:4125.
This work was supported by CREST, Japan Science and [25] Wen YH, Wang Y, Chen LQ. Acta Mater 2001;49:13.
Technology Agency. Y.C. gratefully acknowledges to the [26] Wen YH, Wang Y, Chen LQ. Acta Mater 2002;50:13.
financial support of the 21st century COE program. We [27] Jin YM, Artemev A, Khachaturyan AG. Acta Mater 2001;49:2309.
[28] Artemev A, Jin Y, Khachaturyan AG. Phil Mag A 2002;82:1249.
are grateful to Prof. A.E. Jacobs for discussion on the [29] Ahluwalia R, Lookman T, Saxena A, Albers RC. Acta Mater
strain-only theory. 2004;52:209.
[30] Bhattacharya K, Suquet PM. Proc R Soc A 2005;461:2797.
References [31] Lookman T, Shenoy SR, Rasmussen KO, Saxena A, Bishop AR.
J Phys IV France 2003;112:195.
[1] Wang Y, Khachaturyan AG. Acta Mater 1997;45:759. [32] Curnoe SH, Jacobs AE. Phys Rev B 2001;64:064101.
[2] Kartha S, Krumhansl JA, Sethna JP, Wickham LK. Phys Rev B [33] Cox SM, Matthews PC. J Compt Phys 2002;176:430.
1995;52:803. [34] Kassam A-K, Trefethen LN. SIAM J Sci Comput 2005;26:1214.
[3] Shenoy SR, Lookman T, Saxena A, Bishop AR. Phys Rev B [35] Krogstad S. J Comput Phys 2005;203:72.
1999;60:R12537. [36] Capitani GC, Doukhan JC, Malcherek T, Carpenter M. Eur J Miner
[4] Kerr WC, Killough MG, Saxena A, Swart PJ, Bishop AR. Phase 2001;13:921.
Transitions 1999;69:247. [37] Bendersky LA. Scr Met Mater 1993;29:1645.

You might also like