You are on page 1of 8

Journal of Crystal Growth 457 (2017) 52–59

Contents lists available at ScienceDirect

Journal of Crystal Growth


journal homepage: www.elsevier.com/locate/jcrysgro

Surface rippling during solidification of binary polycrystalline alloy:


Insights from 3-D phase-field simulations
Kumar Ankit a,n, Hui Xing b, Michael Selzer a,c, Britta Nestler a,c, Martin E. Glicksman d
a
Institute of Applied Materials, Karlsruhe Institute of Technology, Haid-und-Neu Str. 7, 76131 Karlsruhe, Germany
b
Key Laboratory of Space Applied Physics and Chemistry, Shaanxi Key Laboratory for Condensed Matter Structure and Properties, Northwestern
Polytechnical University, Xi'an, China
c
Institute of Materials and Processes, Karlsruhe University of Applied Science, Moltkestr. 30, 76133 Karlsruhe, Germany
d
College of Engineering, Florida Institute of Technology, 150 West University Blvd., Melbourne, FL 32901-6975, United States

art ic l e i nf o a b s t r a c t

Communicated by Falini Giuseppe The mechanisms by which crystalline imperfections initiate breakdown of a planar front during direc-
Available online 20 May 2016 tional solidification remain a topic of longstanding interest. Previous experimental findings show that the
Keywords: solid-liquid interface adjacent to a grain boundary provides a potential site where morphological in-
Phase-field modeling stabilities initiate. However, interpretation of experimental data is difficult for complex 3-D diffusion
Solidification fields that develop around grain multi-junctions and boundary ridges. We apply a phase-field approach
Morphological instabilities to investigate factors that induce interfacial instabilities during directional solidification of a binary
Growth models polycrystalline alloy. Using 2-D simulations, we establish the influence of solid-liquid interfacial energies
Diffusion-limited patterns on the spatial localization of initial interfacial perturbations. Based on parametric studies, we predict that
grain misorientation and supersaturation in the melt provide major crystal growth factors determining
solute segregation responsible for surface rippling. Subsequent breakdown of boundary ridges into
periodic rows of hills, as simulated in 3-D, conform well with experiments. Finally, the significance of
crystal misorientation relationships is elucidated in inducing spatial alignment of surface ripples.
& 2016 Elsevier B.V. All rights reserved.

1. Introduction respectively. Several works discussing theoretical aspects of


boundary grooving have been reported since the 1960s [3–6].
The mechanism by which interfacial free energy is minimized Bolling and Tiller [3] described stationary groove profiles of pure
by grain boundary grooving at triple junctions is well known. By materials analytically with solid-liquid interfacial energy, thermal
definition, thermal grooves pertain to solid-vapor, whereas liquid gradient, and dihedral angle. It is reported that the preferred
grooves correspond to solid-melt interfaces. In the present article, growth directions depend upon the anisotropy in surface energy.
we focus on the influence of liquid grooves in instigating surface Coriell and Sekerka [4,5] elucidated grain boundary-mediated
instability during solidification. morphological instabilities by applying linear stability theory for
The dihedral angle, θd, at the triple point of a grain boundary unidirectional solidification of pure materials and binary alloys.
groove develops at equilibrium as a result of solute diffusion. The The spreading of instabilities from initial perturbations can be
equilibrated groove exhibits a pair of convex ridges that are described by a Green's function [7]. Bokstein et al. [6] propose a
symmetrically disposed about the line of intersection of the solid- model of liquid grooving to investigate the kinetic (dissolution)
liquid interface and the grain boundary, provided that the inter- and diffusion regimes at grain boundaries with zero (or nearly
facial energy of the adjacent grains does not depend on the local zero) dihedral angles.
grain orientation to the melt. When constitutional supercooling Grain boundary grooving has been observed and reported in
overcomes the interface-stabilizing thermal gradient, instability numerous experiments on alloy solidification. By experimentally
arises and protruding ridges progressively amplify from the analyzing the isothermal grain boundary grooving in the Sn–Pb
groove. Mullins [1,2] established the kinetic law for grain bound- alloy, Hardy et al. [8] establish synergies with Mullins' power-law.
1 1 Independent experimental investigations on solidification of dilute
ary grooving with corresponding power-law exponents of 4 and 3
Pb–Sb [9] and Al–Cu [10] alloys also reveal that grain boundary
for surface and volume diffusion-controlled kinetic regimes, grooves provide potent sites for initial perturbations; all these
findings agree with theoretical predictions. Further, it is suggested
n
Corresponding author. that crystal orientation has a strong influence on morphological
E-mail address: kumar.ankit@kit.edu (K. Ankit). amplification of ridges and patterns evolving at later stages [9].

http://dx.doi.org/10.1016/j.jcrysgro.2016.05.033
0022-0248/& 2016 Elsevier B.V. All rights reserved.
K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59 53

In order to expose and analyze the mechanisms by which in- describe the methodology, by using constructed free energy den-
terfacial morphological instability initiates adjacent to grain sities, to approximate the variation of the grand-potentials of the
boundary grooves, optically transparent analogues, such as succi- solid and liquid phases as functions of their chemical potentials. In
nonitrile-based alloys (SCN) and CBr2–C2Cl6, were frequently em- Section 3, a validation is presented of phase-field grain boundary
ployed for such directional solidification experiments. The pio- groove simulations checked against classical analytic solutions.
neering study on morphological instabilities adjacent to crystalline The importance of surface energy and grain misorientation in
imperfections by Schaefer and Glicksman [11] shows that grain initiating morphological instability is elaborated using 2-D simu-
boundaries play a key role by initiating morphological instability lations in Section 4. Section 5 is devoted to 3-D analysis of the
of primary ridges. As the initial ridges amplify, parallel secondary mechanisms by which primary ridges amplify, along with a dis-
ridges appear beside them. The primary ridges progressively break cussion of the significance of crystal misorientation relationships
down into periodic rows of hillocks and knobs, some of which inducing spatial alignment of surface ripples. Section 6 contains
eventually evolve to ramified dendritic patterns. A detailed ela- conclusions and the outlook from this study.
boration of these concomitant events is provided in Section 5.
Noël et al. [12,13] investigated the influence of grain bound-
aries and natural convection on microstructure formation in cel- 2. Thermodynamics
lular directional solidification of dilute SCN alloys. Their findings
confirmed that the dynamic grooving of grains occurs by the The grand-chemical potential phase-field model used for our
propagation of ridges and depressions along grain boundaries. numerical investigation is outlined in Appendix A. The free energy
Experimental results by Xing et al. [35] accentuate the potency of of a phase α, in a 2-phase model binary alloy (A-B), are described
grain boundaries in initiating morphological instabilities. The in- by the following form,
itial perturbations observed adjacent to the grain boundary are
reported to be larger than others. Further, it is suggested that ( ) ( ) ( ) ( ) ( )
fα cA, cB, T = Aα T c A2 + B α T c B2 + D α T cA + E α T cB + K α T , ( ) (1)
preferred grain orientations dominate during growth competition
where cA and cB are the concentrations of components A and B,
as the complex patterns evolve. Fig. 1(a–h) shows an exemplary
respectively, and Aα , B α , D α , Eα and Kα are the temperature-de-
evolution of the solid-liquid interface near the grain boundary of
pendent coefficients for phase α. On imposing the binary alloy
succinonitrile-0.9 wt% p-dichlorobenzene (SCN-DCB) alloy for an
constraint cA + cB = 1, one may rewrite the Equation (1) as follows:
applied thermal gradient of 6.2 K/mm and a pulling velocity of
2.80 m/s. ( ) ( ) ( )
fα cA, T = O α T c 2A + P α T cA + Q α T , ( ) (2)
The phase-field simulation method, long established in the
material science community, provides a superlative numerical where O α = Aα + B α , P α = D α − 2Aα − Eα and Q α = B α + Eα + Kα . The
approach to describe microstructural evolution during solidifica- corresponding grand-potential densities are calculated algor-
tion, solid-state and electrochemical transformations [14–18]. ithmically as outlined by Choudhury et al. [21]. We point out that
Two-dimensional phase-field simulations of morphological in- the present generalization of free energy density by a polynomial
stability adjacent to liquid grooves were reported earlier by Yeh facilitates straightforward coupling with CALPHAD databases, as
et al. [19] and by Miller et al. [20]. Both of these works capture the exemplified by previous numerical studies on binary alloys [22–
critical onset of morphological instability near the intersecting 27] and ternary alloys [28,29,21]. As the present investigation is
grain boundary that typically is characterized by “hump amplifi- not limited to a specific alloy, we construct the parabolic free
cation.” However, the influence of lateral flux-assisted solute seg- energy densities such that the equilibrium mole fractions of B in
regation results in a progressive breakdown of the solid-liquid the solid grains and liquid phase are conveniently calculated to be
interface—a process termed “rippling.” For further aspects of this 0.8 ( cBs, eq ) and 0.2 ( cBl, eq ), respectively. The relevant simulation
process, which, heretofore, have not been addressed in detail, see parameters are listed in their dimensional form in Table 1. Since
Fig. 19 in Morris and Winegard [9]. Indeed, further investigation is we are primarily interested in generic features of ripple formation,
warranted to achieve a more precise understanding of defect-in- the reported times are normalized by l02/Dl , where l0 = σ /( RT /Vm ) is
duced interfacial breakdown during directional solidification. the capillary length, T denotes the isothermal transformation
The remainder of this study is organized as follows: we first temperature, s, interfacial energy, and Vm is the molar volume.

Fig. 1. Temporal evolution of solid-liquid interface near the grain boundary of SCN-0.9 wt.%DCB (succinonitrile-p-dichlorobenzene) alloy. The time elapsed during evolution
(a–h) is 7 s.
54 K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59

Table 1 3.2. Grooving kinetics: Mullins's theory


Numerical parameters used for the phase-field simulations.
A grain boundary groove interface in 3-D may be divided into
Parameter Values (Jm  3) Parameter Values Units
three local regions:
Oα −4.4 × 103 Δx = Δy = Δz 1 × 10−7 m
Pα 1.7 × 103 ⎛ c l, eq − c l ⎞ 0.5, 0.6 – 1. Junction of more than two grains, where a groove pit develops.
u ⎜⎜ = A A ⎟
2. The solid–liquid interface far away from the grain boundaries
l, eq s, eq ⎟
⎝ cA − cA ⎠
that may be approximated by an infinite (periodic) surface.
Qα 1.7 × 102 σs, l 0.5, 0.6, 0.7 Jm  2
3. The solid-liquid interface around a grain boundary but away
Oβ −8.9 × 103 sgb 1.0 Jm  2
⎛ σs, l ⎞ 0.5, 0.6, 0.7 –
from the junctions of more than two grains. The grooving ki-
Pβ 1.4 × 104
r ⎜⎜ = ⎟⎟ netics of the third region can be studied efficiently with 2-D
⎝ σgb ⎠
phase-field simulations reported earlier by [31–33].
Qβ 5.7 × 103 DAl = DBl 1 × 10−9 m2s  1
δαβ (Eq. (A.4)) 0.1 Vm 2.24 × 10−4 m3 mol  1
We know from Mullins's theory [2] that d3g increases linearly
with time, t, i.e.
3. Validation with theory dg = d 0 t α , (4)

where d0 is a prefactor and the power-law exponent, α, equals 3


1
3.1. Dihedral angle: Young's law
for volume diffusion-controlled grooving. In order to validate our
The initial condition chosen for validation with Young's law 3-D phase-field model, we choose an initial arrangement of
consists of two grains, each of numerical grid width and thickness 4 equiaxed grains in equilibrium with melt surrounding both sides.
of 2000Δx and 100Δx respectively. Periodic boundary conditions A representative snapshot corresponding to a dihedral angle of
are applied on the left and right-hand edges of the numerical 120° is shown in left inset of Fig. 2 d. The temporal variation of
domain. The dihedral angle pertains to the local equilibrium state groove depths, dg, corresponding to root and pit (illustrated by
resulting from the force balance at the groove's triple-junction, yellow and white arrows in Fig. 2d) are also plotted. The power
given by law exponent α obtained by fitting the simulated dg is found to be
0.3334 in both the cases, which is a result in good agreement with
⎛ σgb ⎞ Mullins's kinetic law, Eq. (4). The pre-factor d0 corresponding to
θd = 2 cos−1⎜ ⎟,
⎝ 2σs, l ⎠ (3) the pit is larger compared to that for the root, which qualitatively
complies with the previous finding of Zhang and Gladwell [34].
where the grain boundary energy, sgb, and the solid-liquid inter-
facial energy, σs, l , can be analytically derived. See [30,19] for de-
tails. We assume for simplicity that the interfacial energies are 4. Numerical investigation of grain rippling
isotropic, and the solid-liquid surface energies for both grains are
identical. In accordance with Eq. (3) the simulated dihedral angle A polycrystalline solidification front is never free of imperfec-
decreases with an increase of grain boundary energy. cf. Fig. 2a–c. tions, especially grain boundary grooves that potentially act as

4
dg

2
Groove pit (numerical)
Groove root (numerical)
100 1000
0 1 2 Time (t)
Fig. 2. Simulations of isotropic grain boundary grooves: (a) σs, l = 0.5σgb , (b) σs, l = 0.6σgb and (c) σs, l = 0.7σgb . The illustrated snapshots of the groove region correspond to a
numerical domain size of 2000Δx × 200Δx , with periodic boundary conditions on the left and right edges. (d) Comparison of Mullins's theory with numerically simulated
grain-boundary grooving kinetics for volume-diffusion controlled regime. Data points represent the simulated values (log-log scale) whereas the lines are power law fits
(growth exponent ¼ 0.3334). The inset picture (on the left) shows an example phase-field simulation with periodic boundary conditions. The grains are plotted with offset to
reveal inner curvatures (see legend) in the vicinity of the triple-junction (right inlet), groove pits (white arrow) and roots (yellow arrow). (For interpretation of the references
to color in this figure legend, the reader is referred to the web version of this article).
K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59 55

sites for initiating morphological instabilities during directional interfacial energy allows the initiation of instabilities close to the
growth. As outlined in Section 1, such observations have been groove cusp.
reported in independent experimental works [11,13,35] on the The initial profile of the two grains chosen for the present phase-
directional solidification of SCN-based alloys. In particular, it is field simulations is elliptical, which corresponds to a groove dihedral
suggested that pulling velocity and alloy concentration also angle of zero. From a numerical perspective, such a choice optimizes
influence the growth of morphological instabilities. As our step the simulation run time, as the onset of morphological instabilities
towards a precise understanding of factors that lead to morpho- relies upon the local gradient of interfacial curvature. Apparently, the
logical instabilities of the solid-melt interface near grain bound- lateral flow of solute along the solid-liquid surface for an initially
aries, we numerically investigate the influence of surface energy planar solidification front is zero, corresponding to a groove angle of
and grain misorientation on the spatial localization of incipient 180° as was the case in the works of Yeh et al. [19] and Miller et al.
perturbations. In this section, the reported phase-field simulations [20]. As grain boundary grooving develops, the interfacial curvature
correspond to the case of an isothermal solidification with a pre- gradients promote the necessary flow of solute along the interface.
scribed melt supersaturation, u, proceeding from a preexisting Obviously, such an initial condition is not favored from a numerical
symmetric arrangement of 2 grains, as shown in Fig. 3a. perspective, as it prolongs simulation run-times.

4.1. Symmetric grooves: influence of surface energy 4.2. Influence of anisotropy in surface energy

In Fig. 3, we illustrate the influence of the ratios, r, between the The correlation between grain boundaries and the formation of
interfacial energy, σs, l , and the grain boundary energy, sgb, both of ordered cellular arrays is reported in the experimental work of
which are assumed to be isotropic. We observe a spatial shift in Noël et al. [13]. It is suggested that undulations and cell alignment
the location of incipient perturbations depending upon this ratio. along the grain boundary, typically observed during the initial
The temporal evolution shown in Fig. 3b, where the ratio r ¼0.5, stages, gives way to flat, regular polygonal arrays of cells. In the
suggests that the initial perturbations develop along the curved following section, we incorporate a smooth form of anisotropy in
portion of the solid-liquid interface, and then continue to amplify. solid-liquid energy (see Eq. A.4) to study the influence of grain
At a larger ratio, where r ¼0.7, the ability of the grain boundary misorientation and melt supersaturation on the morphological
groove to initiate a perturbation large enough to amplify appears stability of the solid-liquid interface. We will establish the influ-
to be hampered, as is shown in Fig. 3c. Such a regime is char- ence of these parameters on 3-D ordering based on insights de-
acterized by the concurrent amplification of perturbations loca- rived from these numerical simulations and reported in Section 5.
lized along the curved surface away from the groove. It is note- Here, we numerically simulate the evolution of grains at dif-
worthy that the flatter portion of solid-liquid interface, i.e., the ferent supersaturations in the melt, u, proceeding from an initial
region around the grain center, is unable to initiate perturbations configuration illustrated in Fig. 3a. In Fig. 4, we compare grain
irrespective of the value of r. The interfacial energy-dependent contours for grain misorientations of 30° and 45°, respectively. At a
localization of initial perturbations surrounding grain boundary supersaturation of u¼ 0.6, Fig. 4a, with solidification proceeding,
grooves reported in these simulations is corroborated by the ex- the primary ridge instability that corresponds to the more favor-
perimental findings reported by Noël et al. [13], who observed the ably-oriented grain is found to amplify at a faster rate compared to
onset of similar instability behavior which depended on the its counter-part, which is overgrown at a later stage. Drawing si-
pulling velocity for their SCN-based alloy. Apparently, such a ca- milarities with the case-study reported in Section 4.1, the ob-
pillary-mediated localization is implicitly related to the ratio of served rippling of solid-liquid surface in the present case is also
driving force (which for the present case is supersaturation in the found to be dependent on grain misorientation. At a lower su-
melt) and the interfacial energy. In the present simulations where persaturation in the melt, u ¼0.5, the overall rate of rippling is
the magnitude of melt supersaturation is held constant, a smaller reduced (Fig. 4b). Interestingly, the left ridge and the perturbation
located away from the groove, corresponding to the more favor-
0.82
t = 12.0 ably-oriented grain, are coarser as compared to Fig. 4a. The 2-D
studies reported in this section signify an intricate correlation
r = 0.5

0.17 between the grain misorientation and the observed evolution of


0.81
t = 8.0 incipient ripples.

200
Δθ=30ο
0.17 (a)
0.76
Y grid-points

ο
(b) t = 2.4 Δθ=45
100

0.16
0.4
(a) t = 0.0 0
500 600 700 800 900 1000 1100 1200 1300 1400 1500
300
ο (b)
Δθ=30
0.2
0.80 Δθ=45 ο
(c) t = 15.2
Y grid-points

200

0.16
0.82 100
t = 55.2
r = 0.7

0
0.15 500 600 700 800 900 1000 1100 1200 1300 1400 1500
0.81 X grid-points
t = 153.6
Fig. 4. Comparison of grain instability contours for assigned relative misorienta-
0.16
tions of 30° and 45°, respectively. The non-dimensional supersaturation, u, equals
(a) 0.6 and (b) 0.5. A representative time-step is selected for comparison In both
Fig. 3. Temporal evolution of the phase contours plotted over the corresponding cases, when temporally amplifying the initial ridge on the right-hand side of the
composition profile for (a) r ¼0.5 and (b) r ¼0.7, where r = σs, l/σgb . groove cusp that overgrows the one on its left.
56 K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59

5. Grain rippling in 3-D periodic hillocks in the third dimension that follow the peak of the
primary ridges. Hillock formation is initiated where a grain
The factors which determine whether or not a planar crystal boundary groove terminates at an intersection with another
surface growing in the melt will break down into a cellular or groove, or where the grain touches the container. The secondary
dendritic structure remain topics of interest. However, the influ- ridges frequently disappear as the primary ridges extend. Where
ence of interface imperfections, such as grain boundary grooves, two grain boundary grooves happen to intersect, a deep pit de-
on the stability of planar surfaces was reported in several experi- velops, which, in turn, often forms several “pre-dendritic disks.”
mental works, but has never been numerically investigated in 3-D. These pre-dendritic disks extend rapidly as protuberances into the
We use the 3-D phase-field modeling in the following section, undercooled or supersaturated melt to form a trio of highly-ra-
outlined in Appendix A, to investigate the influence of grain mis- mified dendrites that grow in preferred crystallographic directions
orientation as one of factors that trigger instability of an advancing set by the orientation of each of the three participating grains.
solid-liquid interface.
5.2. Rippling simulations
5.1. Earlier experiments
In order to validate our numerical technique with the in-
Before discussing our 3-D phase-field simulation results, the stability mechanism suggested above, we started with a sym-
seminal work of Schaefer and Glicksman (See Fig. 3 in [11]) that metric arrangement of 4 equiaxed grains, similar to Fig. 2d, from
underpins the concept of crystalline defect-mediated surface rip- which we simulated the early-stage evolution of the solid-liquid
pling is well worth revisiting. Specifically, Schaefer and Glicksman interface. The supersaturation in the melt contacting the 4 grains
[11] showed experimentally that grain boundary and sub-bound- is u ¼0.5, with misorientations in range of 10° to 20°. The temporal
ary grooves intersecting a nominally planar solid-liquid interface evolution of the solid-liquid interface adjacent to the boundary
play essential roles in initiating morphological instability. An in- groove is illustrated in Fig. 5, shows initial formation of parallel
tersecting straight grain boundary forms a persistent groove-like ridges and then periodic modulation in the third dimension. As the
perturbation on the interface, the lateral dimensions and depth of parallel ridges amplify, secondary ridges form beside them. Once
which depend inversely on the square-root of the local thermal the primary ridges develop more height, they quickly modulate
gradient acting normal to the interface. Morphological instability into hillocks that continue to amplify and form a ramified pattern.
initiates by the localized growth of a pair of ridges that form It is worth noting that the simulated progression in 3-D is in close
parallel to the grain boundary groove, and which approach each agreement with experiment.
other and merge along the triple line formed by the grain Next, we investigated the influence of the grain quadruple
boundary and the planar solid-liquid interface. Ridge growth along junction and its spatial arrangement of secondary ridges and ripples
the grain boundary groove is immediately followed by the for- in 3-D. The initial setting comprises a supersaturated melt, with
mation of weaker, parallel, secondary ridges, with a shallow in- u¼0.5, in contact with a polycrystalline seed consisting of 4 grains
tervening valley connecting the primary and secondary ridges. symmetrically arranged in an array. The grains are misoriented with
Primary ridges then progressively break down into a chain of respect to the most-preferred-orientation in the order Grain 1

(a) (b)

t=0 t = 8.0
(c) (d)

t = 12.8 t = 19.2
Fig. 5. 3-D phase-field simulation showing the morphological evolution of an advancing crystal-melt interface in the vicinity of a grain boundary groove. (a) Initial grain
boundary groove appears on the interface. (b) Ridges form and bulge out towards the melt. The primary ridges formed near the groove are taller in comparison to the weak
secondary ridges adjacent to them. (c) Modulations are observed along the ridges which continue to amplify. Secondary ridges die-out. (d) Hillocks form as periodic ripples
along the ridges as the solid-liquid interface advances to form more complex patterns.
K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59 57

Grain 2 Grain 3

Grain 1 Grain 4

(a) t = 0.0 (b) t = 40.0

(c) t = 28.8 (d) t = 32.0 (e) t = 35.2


Fig. 6. (a) Morphological evolution of an initially flat polycrystalline ‘seed’ in contact with supersaturated melt, illustrated at representative time-steps. The corresponding
axial tilts (θ) for grains 1, 2, 3 and 4 are 15.5,°, 25.5°, 31.6° and 18.1°, respectively. (b) Magnified view of grain 1 /melt surface close to quadruple junction.

(θ = 15.5°) < Grain 4 (θ = 18.1°) < Grain 2 (θ = 25.5°) < Grain 3 (θ = 31.6°), Towards grain center
where θ represents axial-tilt (See Fig. A8c.) Fig. 6 illustrates pro-
gressive amplification of primary and secondary ridges at re-
presentative time-steps.
As the melt supersaturation was chosen to be large enough to
initiate morphological instabilities near the groove, as well as at
the grain center, primary ridges develop adjacent to the groove as
do ripples located close to the grain centers. Another interesting
characteristic is that the primary ridge close to the grain quadruple
junction is found to amplify at a much larger rate as compared to
those along the triple line regions. The influence of grain mis-
orientation is also apparent; the fastest amplifying ridges corre-
spond to the most-favorably oriented grain, which is grain 1 in the (a)
present setting. At a given time, the amplitude of ripples closer to
Grain Boundary
multi-junctions is much larger as compared to those near grain
centers. Before the amplifying ripples evolve to a complex pattern,
the following order of amplification rate is observed: (1) Primary
ridges next to the quadruple junction; (2) primary ridges along the
triple junction; (3) secondary ridges near the quadruple junction;
(4) secondary ridges adjacent to the triple junction, (5) secondary
ridges closer to the grain centers.
Finally, we comment upon the influence of grain misorienta-
tion on the spatial ordering of primary ridges and ripples found in
the experiments reported by Noël et al. [13]. In Fig. 7a, we illus-
trate surface ripples adjacent to the grain boundary interface be-
tween grain 3 (θ ¼ 31.6°) and grain 2 (θ ¼ 25.5°). The primary rid- (b)
ges that are tilted along θ with respect to the most-preferred-or-
ientation (chosen to be the same as the viewing angle) are found Fig. 7. Top-view illustrating the spatial arrangement of primary ridges and ripples
across the grain boundary interfaces of grains (a) grain 3 (left) and grain 2 (right),
to be ordered along the grain boundary. However, the extent of and (b) grain 1 (left) and grain 4 (right).
ordering progressively declines on moving away from the grain
boundary interface. The ordering of primary and secondary ridges, 6. Conclusions and outlook
shown in Fig. 7b, in the form of oblong cells is observed to be more
pronounced across the interface between grain 1 (θ ¼15.5°) and We explored persistent perturbations and supersaturation—
grain 4 (θ ¼18.1°). factors that initiate morphological instability of a solidifying
58 K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59

planar solid-liquid interface by using phase-field simulation. We N

focused primarily on the morphological evolution occurring near Ψ (T , μ, ϕ) = ∑ Ψα (T , μ) hα (ϕ),


α=1
grain boundaries at sites where perturbations develop. First, the
K− 1
required force balance at the equilibrium dihedral angle was es-
Ψα (T , μ) = fα (c α ( T , μ), T ) − ∑ μi ciα (T , μ),
tablished for different ratios of surface and grain boundary en- i=1 (A.2)
ergies, evidenced by agreement between simulated dihedral an-
gles and Young's law. Next, simulated grooving kinetics was found where hα ( ϕ) is an interpolation function of the form
to obey kinetics in close agreement with Mullins's groove growth ( )
hα ϕ = ϕα2 3 − 2ϕα . ( )
power-law. The validated model was then used to conduct a sys- The gradient energy density, a ( ϕ, ∇ϕ), is given by
tematic numerical investigation to decompose the influences of
N, N
grain misorientation and melt supersaturation on the critical onset
σαβ ⎡⎣ a c qαβ ⎤⎦ qαβ ,
2 2
of interfacial ripples. For the first time, we demonstrate that the
a (ϕ, ∇ϕ) = ∑ ( )
α, β = 1
ratio of solid-liquid interface and grain boundary energies is one of ( α < β) (A.3)
the factors which cause localization of incipient perturbations.
Using 3-D phase-field modeling, we established the role of ( )
where ac qαβ defines the form of the surface energy anisotropy of
grain misorientation on influencing the spatial alignment of in- the evolving phase boundary and σαβ is the interfacial free energy
terface ripples. We also showed that the evolution of primary per unit area of the α /β phase boundary, which may additionally
ridges along straight grain boundaries is similar to the instability depend on the relative orientation of the interface. The vector
phenomena observed in directional solidification experiments on quantity qαβ = ϕα ∇ϕβ − ϕβ ∇ϕα is a generalized gradient vector
SCN-based alloys [11,13]. The present work, whilst providing va- normal to the α /β interface.
luable insights, falls short of a fully detailed parametric study. The To assign an isotropic surface energy to the α /β phase bound-
phase-field model could be extended to investigate the influences ( )
ary, ac qαβ is chosen to be 1. Weakly anisotropic crystals with an
of the thermal gradient and melt convection to facilitate a more underlying cubic symmetry are modeled in three dimensions by
quantitative comparison of numerical results with experiments. the expression
The topic of complex pattern evolution, typically observed at later
stages as the initial perturbations progressively amplify, has not ⎛ 4 ⎞
⎜ qαβ ⎟
4
been addressed in the present article. Complementary numerical a c (qαβ ) = 1 − δαβ ⎜ 3 ∓ 4 4 ⎟,
studies addressed to these interesting topics are currently in ⎜ qαβ ⎟
⎝ ⎠ (A.4)
progress.
d ⎛
4 ⎞ ⎡ d ⎤2
with q
4 ⎝ ⎠ ( )
= ∑i = 1 ⎜ qi4 ⎟ and q 4 = ⎢⎣ ∑i = 1 qi2 ⎥⎦ . The anisotropic

Acknowledgments strength of an α /β interface is given by the parameter δαβ . A polar


plot of function A.4 is shown In Fig. A8a. An alternative formula-
tion, corresponding to a strongly anisotropic crystal of faceted type
KA thanks GEOLAB for providing financial support and a mul-
tidisciplinary research platform. HX acknowledges his support was employed elsewhere [37–40]. The evolution equation for the
from National Natural Science Foundation of Shaanxi Province in N phase-field variables can be written as,
China (No. 2015JQ5125). BN acknowledges the financial support of ∂ϕα ⎛ ∂a ( ϕ, ∇ϕ) ∂a ( ϕ, ∇ϕ) ⎞ 1 ∂w ( ϕ) ∂Ψ ( T , μ, ϕ)
τϵ = ϵ ⎜⎜ ∇· − ⎟⎟ − − − Λ,
DFG in the framework of Sino-German research effort on “Phase ∂t ⎝ ∂∇ϕα ∂ϕα ⎠ ϵ ∂ϕα ∂ϕα (A.5)
transformations under extreme conditions” (No. NE 822/22-1).
MEG acknowledges his support from the Allen S. Henry Dis- where Λ is the Lagrange parameter that maintains the constraint
N
tinguished Professorship at Florida Institute of Technology. ∑α = 1 ϕα = 1. The equilibrium crystal shape simulated (corre-
sponding to function A.4) using the volume-preservation techni-
que [41] is illustrated in Fig. A8b.
Appendix A. Phase-field model The concentration fields are obtained by a mass conservation
equation for each of the K  1 independent concentration variables
In this appendix, for the sake of completeness, we briefly re- ci. The evolution equation for the concentration fields can be de-
count the multiphase-field equations used to study micro- rived as,
structural evolution. The reader is referred to our prior studies ⎛ K− 1 ⎞
∂ci
[36,22,24,28,25] for a more detailed description of the simulation = ∇·⎜⎜ ∑ Mij (ϕ)∇μj ⎟⎟,
methods employed. ∂t ⎝ j=1 ⎠ (A.6)
The evolution of phases is governed by the phenomenological
minimization of the grand potential functional Ω, and
⎡ ⎛ ⎞⎤ N
Ω (T , μ, ϕ) = ∫V ⎢⎣ Ψ (T , μ, ϕ) + ⎜⎝ ϵa (ϕ, ∇ϕ) + 1ϵ w (ϕ) ⎟⎠ ⎥⎦ dV . (A.1)
Mij (ϕ) = ∑ Mijα gα (ϕ),
α=1 (A.7)
Here T is the temperature, μ is a chemical potential vector con- where each represents the mobility matrix of the phase α
Mijα
sisting of K  1 independent chemical potentials, ϕ is the phase- (related to solute diffusivities, Dl and Dg in the melt and grains,
field vector containing the volume fractions of the N-phases and ϵ respectively).
is the length scale related to the interface. a ( ϕ, ∇ϕ) an w ( ϕ) re- The removal of artificial, thin-interface solute trapping is
present the gradient and obstacle potential type energy density, achieved by incorporating the form of anti-trapping current pro-
respectively, and finally V denotes the domain volume. posed by Choudhury and Nestler [36] based on a double-obstacle
The grand potential density Ψ ( T, μ, ϕ), which is the Legendre potential formulation. The model equations are discretized using a
transform of the free energy density of the system, f ( T, c , ϕ), is finite-difference method. The spatial derivatives are subdivided
written as an interpolation of individual grand potential densities, into left-sided, right-sided, and centered differences, such that
namely, grid anisotropy is minimized. The time derivative obeys an explicit
K. Ankit et al. / Journal of Crystal Growth 457 (2017) 52–59 59

0
Rotation axis
Z Ax
ial
tilt

45
Y

Small Curvature Large


(a) (b) (c) X
Fig. A8. (a) Interfacial energy plotted in polar coordinates and (b) the corresponding equilibrium shape simulated using the volume preservation algorithm of Nestler et al.
(c) The axial tilt with respect to the most-preferred-orientation aligned along the z-axis and rotation are used to define the crystal misorientation for the phase-field
simulations of directional solidification.

Euler scheme. The grand-chemical potential model is im- [19] S. Yeh, C. Chen, C. Lan, J. Cryst. Growth 324 (2011) 296–303.
plemented in the Pace3D software package version 2.1.1. As a test [20] W. Miller, A. Popescu, G. Cantù, J. Cryst. Growth 385 (2014) 127–133.
[21] A. Choudhury, M. Kellner, B. Nestler, Curr. Opin. Solid State Mater. Sci. 19
case, we used identical simulation parameters, as reported in Fig. 3 (2015) 287–300.
with a smaller grid size of 0.1Δx to observe the influence of in- [22] D. Molnar, R. Mukherjee, A. Choudhury, A. Mora, P. Binkele, M. Selzer,
terfacial energies on rippling. B. Nestler, S. Schmauder, Acta Mater. 60 (20) (2012) 6961–6971.
[23] E. Wesner, A. Choudhury, A. August, M. Berghoff, B. Nestler, J. Cryst. Growth
On comparing the evolution of phase contours (assuming iso-
359 (2012) 107–121.
tropic interfacial energies) corresponding to 1.0Δx with 0.1Δx , it [24] K. Ankit, A. Choudhury, C. Qin, S. Schulz, M. McDaniel, B. Nestler, Acta Mater.
becomes apparent that the simulation results are independent of 61 (11) (2013) 4245–4253.
grid size. [25] K. Ankit, R. Mukherjee, T. Mittnacht, B. Nestler, Acta Mater. 81 (2014) 204–210.
[26] K. Ankit, R. Mukherjee, B. Nestler, Acta Mater. 97 (2015) 316–324.
[27] K. Ankit, T. Milttnacht, R. Mukherjee, B. Nestler, Comput. Mater. Sci. 108B
(2015) 342–347.
References [28] R. Mukherjee, A. Choudhury, B. Nestler, Model. Simul. Mater. Sci. 21 (7) (2013)
075012.
[29] J. Hötzer, M. Jainta, P. Steinmetz, B. Nestler, A. Dennstedt, A. Genau, M. Bauer,
[1] W. Mullins, J. Appl. Phys. 28 (1957) 333–339. H. Köstler, U. Rüde, Acta Mater. 93 (2015) 194–204.
[2] W. Mullins, Trans. Metall. Soc. AIME 218 (1960) 354–361. [30] J. Warren, R. Kobayashi, A.E. Lobkovsky, W.C. Carter, Acta Mater. 51 (2003)
[3] G. Bolling, W. Tiller, J. Appl. Phys. 31 (1960) 1345–1350. 6035–6058.
[4] S. Coriell, R. Sekerka, J. Cryst. Growth 19 (1972) 90–104. [31] M. Bouville, S. Hu, L.-Q. Chen, D. Chi, D. Srolovitz, Model. Simul. Mater. Sc. 14
[5] S. Coriell, R. Sekerka, J. Cryst. Growth 19 (1973) 285–293.
(2006) 433–443.
[6] B. Bokstein, L. Klinger, I. Apikhtina, Mater. Sci. Eng. A 203 (1995) 373–376.
[32] M. Bouville, D. Chi, D. Srolovitz, Phys. Rev. Lett. 98 (2007) 1–4.
[7] R. Sekerka, J. Cryst. growth 10 (1971) 239–250.
[33] M. Bouville, D. Chi, D. Srolovitz, Solid State Phenom. 129 (2007) 89–94.
[8] S. Hardy, G. McFadden, S. Coriell, J. Cryst. Growth 114 (1991) 467–480.
[34] W. Zhang, I. Gladwell, J. Cryst. Growth 277 (2005) 608–622.
[9] L. Morris, W. Winegard, J. Cryst. Growth. 5 (1969) 361–375.
[35] H. Xing, C. Wang, J. Wang, C. Chen, Sci. China Ser. G Phys. Astron. 54 (2011)
[10] R. Sharp, A. Hellawell, J. Cryst. Growth 6 (1970) 334–340.
2174–2180.
[11] R. Schaefer, M. Glicksman, Metall. Trans. 1 (1970) 1973–1978.
[36] A. Choudhury, B. Nestler, Phys. Rev. E 85 (2) (2012) 21602.
[12] M. Noël, H. Jamgotchian, B. Billia, J. Cryst. Growth 181 (1997) 117–132.
[37] B. Nestler, H. Garcke, B. Stinner, Phys. Rev. E 71 (2005) 041609.
[13] M. Noël, H. Jamgotchian, B. Billia, J. Cryst. Growth 187 (1998) 516–526.
[38] K. Ankit, B. Nestler, M. Selzer, M. Reichardt, Contrib. Miner. Pet. 166 (2013)
[14] L.-Q. Chen, Ann. Rev. Mater. Res. 32 (2002) 113–140.
1709–1723.
[15] W. Boettinger, J. Warren, C. Beckermann, A. Karma, Ann. Rev. Mater. Res. 32
[39] K. Ankit, J. Urai, B. Nestler, J. Geophys. Res. Solid Earth 120 (2015) 3096–3118.
(2002) 163–194.
[40] K. Ankit, M. Selzer, B. Hilgers, C. Nestler, J Petrol. Sci. Res. 4 (2015) 79–96.
[16] K. Thornton, J. Ågren, P. Voorhees, Acta Mater. 51 (2003) 5675–5710.
[41] B. Nestler, F. Wendler, M. Selzer, B. Stinner, H. Garcke, Phys. Rev. E 78 (2008)
[17] I. Singer-Loginova, H. Singer, Rep. Prog. Phys. 71 (2008) 106501.
[18] B. Nestler, A. Choudhury, Curr. Opin. Solid State Mater. Sci. 15 (2011) 93–105. 011604.

You might also like