Está en la página 1de 12

Current Opinion in Solid State and Materials Science 20 (2016) 25–36

Contents lists available at ScienceDirect

Current Opinion in Solid State and Materials Science


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

Atomistic to continuum modeling of solidification microstructures


Alain Karma a,⇑, Damien Tourret b
a
Department of Physics and Center for Interdisciplinary Research on Complex Systems, Northeastern University, Boston, MA 02115, USA
b
Los Alamos National Laboratory, Materials Science and Technology Division (MST 6), Los Alamos, NM 87545, USA

a r t i c l e i n f o a b s t r a c t

Article history: We summarize recent advances in modeling of solidification microstructures using computational meth-
Received 11 March 2015 ods that bridge atomistic to continuum scales. We first discuss progress in atomistic modeling of equilib-
Revised 3 September 2015 rium and non-equilibrium solid–liquid interface properties influencing microstructure formation, as well
Accepted 19 September 2015
as interface coalescence phenomena influencing the late stages of solidification. The latter is relevant in
Available online 26 September 2015
the context of hot tearing reviewed in the article by M. Rappaz in this issue. We then discuss progress to
model microstructures on a continuum scale using phase-field methods. We focus on selected examples
Keywords:
in which modeling of 3D cellular and dendritic microstructures has been directly linked to experimental
Multiscale solidification modeling
Atomistics
observations. Finally, we discuss a recently introduced coarse-grained dendritic needle network approach
Interface pattern to simulate the formation of well-developed dendritic microstructures. This approach reliably bridges the
Grain structure well-separated scales traditionally simulated by phase-field and grain structure models, hence opening
new avenues for quantitative modeling of complex intra- and inter-grain dynamical interactions on a
grain scale.
Published by Elsevier Ltd.

1. Introduction On the other hand, phase-field simulations have been exten-


sively used to model the formation of monophase and polyphase
The quantitative prediction of solidification microstructures has solidification microstructures in a wide range of dendritic, eutectic
been a long standing computational challenge [15,5]. It generally and peritectic alloys [16,5]. Progress has been made to formulate
requires accurate modeling of the complex dynamics of the quantitative phase-field models of alloy solidification for a wider
solid–liquid interface during the entire solidification process. The range of phase diagrams [113] and diffusive transport properties
past two decades have witnessed major progress in modeling this [19,20] within the theoretical framework of the thin interface limit
dynamics through atomistic- and continuum-scale simulations, [79,38,77]. Results of atomistic modeling have been incorporated
and a handshake between those methods. into phase-field simulations to model complex observations. For
On the one hand, atomistic simulations have yielded accurate example, the quantitative characterization of interface free-
predictions of fundamental anisotropic properties of the solid– energy anisotropy derived from atomistic simulations has been
liquid interface for various metallic systems, such as the interface useful to model changes of dendrite growth orientation [63,35].
free-energy [65,66,5,48] and kinetic coefficient [69,66,130,53,101] Moreover, boosted by steady increases in computing speed and
that affect microstructural pattern formation over a wide range of massively parallel code implementations on multicore architec-
solidification conditions. Those simulations typically use a million tures such as Graphic Processing Units (GPUs), phase-field
atoms and interatomic potentials modeled by the embedded- simulations have become increasingly 3D and able to access
atom method (EAM). They have also yielded useful quantitative experimentally relevant length and time scales [124,12].
insights into solute-trapping in rapidly solidified alloys [158,70], In the following section, we review recent advances in atomistic
as well as interface coalescence phenomena that control crystal modeling of solid–liquid interface properties and coalescence,
cohesion during the late stages of solidification [67,46,111] and focusing on the prediction of key parameters influencing
are relevant to hot tearing, reviewed in the article by Rappaz in this microstructure formation or hot tearing on a larger scale. In Sec-
issue [118]. tion 3, the core section of this article, we review progress in
phase-field modeling focusing on cellular and dendritic
microstructures. We leave out two-phase microstructures,
reviewed in the separate article by Plapp in this issue [114]. In
⇑ Corresponding author. Section 4, we discuss recent progress to bridge the gap between
E-mail addresses: a.karma@neu.edu (A. Karma), dtourret@lanl.gov (D. Tourret).

http://dx.doi.org/10.1016/j.cossms.2015.09.001
1359-0286/Published by Elsevier Ltd.
26 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

phase-field modeling on the microstructure scale and grain struc- and undercooling) and its anisotropy for various pure metals
ture modeling [118]. This gap stems from the fact that, while 3D [68,69,66,130,53,101]. Simulations for fcc (e.g. Ni) and bcc (e.g.
volumes up to about a mm3 are within reach of today’s large scale Fe) have predicted the ordering l100 > l110 > l111 that favors den-
phase-field computations, those volumes are still minute on the drite growth along h1 0 0i directions, as commonly, albeit not
scale of a casting. In contrast, Cellular Automata coupled with uniquely, observed in deeply undercooled melt experiments. This
Finite Elements (CAFE) models [119,51,118] can access those much ordering was predicted by a theoretical study of bcc crystallization
larger scales and predict grain structures of castings [26]. However, kinetics that exploits a Ginzburg–Landau (GL) expansion of the
those models do not resolve dynamical interactions between excess solid–liquid free-energy [156]. There, the GL expansion
branches of hierarchical dendritic networks, which can strongly was expressed in terms of the amplitudes of density waves corre-
influence both intra-grain microstructure selection and the growth sponding to principal reciprocal lattice vectors of the crystal lattice.
competition of different grains [139]. Section 4 describes a A similar expansion had been used to predict csl and its anisotropy
Dendritic Needle Network (DNN) approach designed to quantita- [155] in equilibrium. In its extension outside of equilibrium, GL
tively model those interactions on the grain scale, beyond the theory is closely related to the linear density wave theory of l by
current reach of phase-field modeling. While this approach is not Mikheev and Chernov (MC) [97,30]. The anisotropy of l arises nat-
a substitute for grain structure models, it opens new avenues to urally in both theories from the dependence of density wave pro-
test and improve those models by investigating complex branch files across the solid–liquid interface on the interface normal
interactions that shape the grain structure [138]. direction n^ . Both theories also predict that l is inversely propor-
tional to the relaxation time sL of density waves in the liquid, itself
2. Atomistic modeling related to the liquid diffusivity. Using a direct computation of sL by
MD simulations for an EAM model of bcc Fe [156], GL theory was
2.1. Solid–liquid interface properties found to predict much more accurately than the MC theory the
magnitude of l and its anisotropy obtained in free-solidification
The advent of microscopic solvability theory during the 1980s MD simulations for the same EAM potential [53]. This improve-
[86,81] led to the understanding that anisotropic properties of ment can be attributed to the fully nonlinear description of density
the solid–liquid interface have a profound influence on dendrite wave profiles across the solid–liquid interface in GL theory. The
growth. Those properties generally determine the selection of both prediction that l  1=sL is also consistent with free-solidification
the dendrite tip operating state and growth direction. Predictions MD simulations showing that, for different EAM potentials and
of solvability theory were validated by phase-field simulations small interface undercooling, l correlates well with the liquid dif-
during the 1990s [78,79] and the first decade of this millennium fusivity [96].
witnessed major progress in characterizing interfacial properties In addition to determining l in pure metals, MD simulations
by atomistic modeling [65,66,5]. Those studies yielded the novel have also probed the strong departure from chemical equilibrium
insight that at least two parameters a1 and a2 (also commonly at the solid–liquid interface during rapid alloy solidification for
denoted by 1 and 2 [66]) are necessary to describe the anisotropy both a Lennard-Jones binary system and an EAM model of Ni–Cu
of the solid–liquid interfacial free-energy csl , as opposed to a single [158,70]. The results have confirmed the expected transition to
parameter a1 previously considered [86,81,78,79]. Since this aniso- complete solute trapping at high growth rates predicted by differ-
tropy is weak in metallic systems forming fcc or bcc crystal struc- ent analytical models [7,50,75], and demonstrated the existence of
tures, the variation of csl with the direction normal to the interface, solute drag previously examined using both sharp-interface [7] and
^ , can be parameterized by an expansion in Kubic harmonics
n phase-field models [2]. They have also shown solute drag to be ani-
sotropic [158], potentially affecting both dendrite growth rate and
csl ðn^Þ ¼ c0 ½1 þ a1 K 1 ðn^ Þ þ a2 K 2 ðn^ Þ þ    ð1Þ orientation [70].

where the K i ’s are combinations of spherical harmonics with cubic


symmetry. The first harmonic K 1 ðn ^ Þ favors dendrite growth along 2.2. Interface coalescence
h1 0 0i directions for a1 > 0, while the second K 2 ðn ^ Þ favors h1 1 0i
directions for a2 < 0. Molecular dynamics (MD) simulations of sev- While crystal cohesion is energetically favored when solid–
eral fcc metals (e.g. Cu, Ni, Al, Au, Ag) and alloys (e.g. Ni–Cu) liquid interfaces from the same grain impinge upon each other
reviewed in [5] have predicted that a1 > 0 and a2 < 0 corresponding during the late stages of solidification, it can become energetically
to the ordering c100
sl > c110
sl > c111
sl ; the magnitude of csl was reason-
disfavored when the two grains are sufficiently disoriented. If the
ably well predicted by the Turnbull relation [143] csl ¼ aL.2=3 with a grain-boundary (GB) free-energy cgb exceeds twice csl , i.e.
Turnbull coefficient a  0:55 for fcc (L is the latent heat of melting Dc  cgb  2csl > 0, the formation of a thin intergranular liquid-
per atom and . is the solid atomic density). A study of a model alloy like film becomes energetically favored at temperatures below the
system with nearly ideal solution thermodynamics has also pre- melting point. This GB premelting phenomenon has been widely
dicted that alloying tends to make a2 more negative, thereby poten- studied experimentally [28,8,71,92,145,152,36,73,37,91,59] and
tially promoting growth towards h1 1 0i [10]. This knowledge has theoretically using lattice models [82], phase-field models
motivated detailed phase-field modeling and experimental studies [89,134,135,99,151,121], MD simulations [24,146,154,67,46,111],
of changes of dendrite growth directions in Al–Zn alloys, discussed and phase-field-crystal simulations that resolves the atomic-scale
in the next section. Those changes are explained by the assumption crystal density field [14,95,1]. In a solidification context, intergran-
that Zn addition changes the csl anisotropy so as to induce a transi- ular films delay crystal cohesion, which occurs at a bridging temper-
tion from h1 0 0i to h1 1 0i directions, but atomistic modeling ature below the liquidus temperature when Dc > 0, thereby
remains needed to predict a1 and a2 for this alloy system to validate reducing shear resistance. For this reason, Dc is a key parameter in
this scenario. continuum models of hot tearing [120,150,5] reviewed in Ref. [118].
While dendrite growth is controlled by csl and its anisotropy for Atomistic modeling has been used to characterize quantita-
low solidification rate, it becomes strongly influenced by anisotro- tively the structural forces that control the intergranular layer
pic interface kinetics for large solidification rates [22,72,21]. MD width w [95,67,46,1]. This characterization has been carried out
simulations have been used to compute the interface kinetic by defining a disjoining potential WðwÞ via the excess Gibbs free-
coefficient l (proportionality constant between interface velocity energy per unit of GB area
A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36 27

Gexc ðw; TÞ ¼ DGðTÞw þ WðwÞ ð2Þ important role in other systems (e.g. ceramic materials [32]). How-
ever, in metallic systems, dispersion forces can be estimated to
where DGðTÞ ¼ GS ðTÞ  GL ðTÞ is the bulk Gibbs free-energy differ- yield a contribution to WðwÞ on the order of mJ/m2 that is negligi-
ence between solid and liquid and WðwÞ  2csl represents the part ble on the scale of Fig. 1. Similar GB premelting behaviors are also
of this excess Gibbs free-energy due to the overlap of spatially dif- predicted by PFC simulations [95,1], which have been compared
fuse solid–liquid interfaces from the two impinging grains. Mini- quantitatively to MD simulations for classical models of pure Fe
mization of Gexc ðw; TÞ with respect to w using the simple [1]. PFC [95] and MD [111] studies have also yielded the insight
analytical form WðwÞ ¼ 2csl þ Dc expðw=dÞ [153,120], where d is that the condition Dc ¼ cgb  2csl > 0 can only accurately predict
proportional to the solid–liquid interface width, predicts that w GB premelting if one takes into account the fact that cgb is signifi-
diverges at the melting point T m for Dc > 0 and remains zero up cantly reduced at the melting point (compared to zero tempera-
to a temperature exceeding T m for Dc < 0. ture) due to the elastic softening of the material. Those studies
MD simulations that exploit fluctuations of layer width to com- have also revealed the existence of novel transitions between dif-
pute WðwÞ [67,46] have shown that high-energy GBs exhibit a ferent GB structures mediated by pairing of dislocations [95,111].
purely repulsive disjoining potential that is consistent with the Atomistic modeling has also been used recently to characterize
exponential form of WðwÞ for Dc > 0, as illustrated in Fig. 1 for a the response of high-temperature GBs to a shear stress
pure Ni twist GB. However, for lower-energy GBs, WðwÞ exhibits [25,74,142,141,80]. An exhaustive MD study of ½1 0 0 symmetric
a shallow minimum that corresponds to a finite layer width at tilt GBs in pure Ni [25] has shown that the range of GB misorienta-
the melting point, as illustrated for a pure Ni tilt GB in Fig. 1. This tion where GB sliding is observed in response to a shear stress
nanometer-scale width represents a compromise between attrac- increases close to the melting point, consistent with the view that
tive and repulsive parts of the disjoining potential that are both the formation of an intergranular liquid-like film favors sliding.
structural in nature, since they are mediated by the overlap of However, a recent combined PFC and MD study [141] also showed
the tails of the crystal density fields and crystal-defect formation that asymmetrical GBs in fcc metals can exhibit sliding unrelated
energy, respectively. The disjoining potential WðwÞ also generally to premelting. While those results have not been incorporated to
contains an attractive part due to London dispersion forces date in mesoscale models of mechanical behavior of the mushy
that are not accounted for in MD or PFC simulations, but play an zone and hot tearing, they should be relevant for understanding

9 9 Tilt

Average
Width

Largest
Width

Smallest
Width

Fig. 1. Snapshots of MD simulations of pure Ni modeled with an EAM potential at an undercooling of 2 K below the melting point with solid (liquid) atoms colored in red
(blue). Snapshots on the top left illustrate different configurations for a wet R9 twist boundary that exhibits large fluctuations of the liquid-layer width w [67], while those on
the top right are for a dry R9 tilt boundary that exhibits smaller width fluctuations [46]. The bottom figure shows the disjoining potentials for those two boundaries. The twist
boundary exhibits a purely repulsive potential, corresponding to a diverging liquid layer width at the melting point, while the potential for the tilt boundary exhibits a
minimum corresponding to a finite width at the melting point.
28 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

those complex phenomena. Finally, while multiphase-field model- matters the most, in the present case close to the solid–liquid
ing has been used to investigate interface coalescence in simple interface and the dendritic tips. Dynamic (adaptive) remeshing
binary alloys [151], MD studies remain limited to pure metals. techniques for PF calculations [117] – also applicable to atomistic
Those studies still need to be extended to metallic alloys to under- phase-field-crystal calculations [6] – allow efficient simulations
stand more quantitatively the influence of solute addition on hot up to the scale of an entire dendritic array. Resulting simulations
cracking sensitivity. have shed light on several microstructure selection mechanisms
[3,61,4,109,60,102], and have been coupled to larger-scale heat
transfer models and probabilistic nucleation models to predict
3. Phase-field modeling
microstructure distribution in processes such as laser powder
deposition [44] and welding [103].
3.1. Quantitative phase-field modeling
From a hardware perspective, the parallel architecture of GPUs,
first developed for graphics in video games and movies, has
Due to their simplicity and potential in handling complex free-
recently emerged as a powerful and cost-effective tool for inten-
boundary problems, phase-field (PF) models have become ubiqui-
sive computations. With the development of GPUs dedicated to
tous in materials science [16,29]. They are well adapted to study
high performance computing, these massively parallel architec-
solid–liquid interface dynamics during solidification. Quantitative
tures – e.g. 2880 cores in one NvidiaÒ TeslaÒ K40 – have been
predictions have long been limited by the requirement of a diffuse
increasingly used in scientific computing. The most advanced use
interface width of comparable size to the physical interface, i.e. a
of GPUs for solidification simulation is the work of Aoki, Takaki,
few atomic planes. However, the development of asymptotic anal-
Yamanaka, and co-workers [157,124], reaching the petaFLOPS cal-
ysis of PF equations, now referred to as the thin interface limit, has
culation on the Japanese TSUBAME 2.0 supercomputer, using a
made it possible to use a numerical interface thickness several
hybrid CPU–GPU algorithm for a three-dimensional competitive
orders of magnitude larger than the physical interface thickness
grain growth simulation – 4000 GPUs along with 16,000 CPU cores,
[79,77,38]. For alloy solidification, a quantitative PF formulation
for 4096  6500  10,400 grid points [124]. While such intensive
was originally formulated for isothermal [77] and non-isothermal
calculations require a significant investment in terms of imple-
[38] solidification by introducing an anti-trapping current
mentation, the simplest single GPU implementations already allow
[77,38]. This current compensates for the excess solute trapping
3D simulations that are two orders of magnitude faster than using
associated with a mesoscale interface thickness and yields the
a single CPU core [157,139].
desired thin-interface limit for low solidification rate, where the
Alongside with adaptive remeshing, parallelization techniques
interface can be assumed to be in local thermodynamic equilib-
on GPUs or other architectures – e.g. using MPI [54,106] or OpenMP
rium. However, this formulation is only valid for dilute alloys
[106,56] – allow simulations at experimentally relevant length and
and vanishing solute diffusivity in the solid. Recent progress has
time scales [12,56,139]. The following subsection highlights a few
been made to develop quantitative PF models of binary alloy solid-
examples of how the combination of theoretical approaches and
ification that also utilize an anti-trapping current, but remain valid
advanced computational techniques can lead to a deeper under-
for arbitrary phase diagrams [113] or for a finite solid diffusivity
standing of growth dynamics and microstructure selection in solid-
[23,19,20]. The generalization to arbitrary phase diagram is based
ification processing.
on deriving the PF equations variationally from a grand-potential,
instead of a free-energy, functional, and concomitantly using the
chemical potential as dynamical variable instead of the solute con-
3.3. Selected applications
centration [113]. This approach also unifies different physical
interpretations of alloy PF models (see [113] and earlier references
3.3.1. Comparison to directional solidification experiments
therein). The extension to finite solid diffusivity, in turn, exploits
The thin interface approach, extended to non-isothermal solid-
additional freedom in model formulation by introducing a cross
ification and combined with the anti-trapping current, has shown
coupling between the nonconserved phase field and the conserved
quantitative agreement with directional solidification experiments
concentration field within the formal theoretical framework of
[38]. It has provided an exact benchmark for analytical models of
irreversible thermodynamics and the Onsager relations [23,19,20].
tip undercooling selection, and highlighted complex mechanisms
such as the existence of stability gaps in pattern selection at low
3.2. Computationally intensive phase-field implementations interfacial anisotropies [62,12]. The long-standing problem of den-
dritic sidebranching (SB) origin was confirmed to be noise amplifi-
A major contribution to bridging scales with phase-field model- cation [86], while a bifurcation to a limit-cycle oscillatory branch of
ing has been the fast growth of computational capabilities. Numer- deterministic SB was shown to exist at large spacings [39]. Inter-
ous approaches have been proposed to take advantage of the estingly, a higher thermal gradient was found to reduce the onset
available computational power including adaptive mesh refine- velocity and wavelength of SB, due to the dendrite tip becoming
ment [116], optimized implementations on parallel architectures blunter, thus promoting noise amplification.
[124], hybrid finite-difference and random-walk algorithms This PF model, implemented on GPUs, yields quantitative pre-
[115], implicit time stepping schemes and multi grid approaches dictions on complex nonlinear dynamics of spatially extended
[58,17], and methods that exploit up-scaling techniques and sym- solid–liquid interfaces. This is exemplified in Fig. 2 by a recent
metry [13]. We highlight two of these approaches that yielded sig- comparison of 3D simulations to directional solidification experi-
nificant insight into microstructure selection, namely mesh ments with in situ imaging of the dynamics of spatially extended
refinement and massive parallelization facilitated by the growing cellular arrays under microgravity conditions where fluid flow is
use of Graphic Processing Units (GPUs). suppressed [12]. There, PF simulations on time scales of several
Phase-field remains the method of choice for simulating com- hours, reproduce oscillatory breathing modes observed in cellular
plex interface patterns in solidification. However, a quantitative arrays with a striking agreement on the oscillation period. The sim-
prediction of growth dynamics with PF requires an accurate ulations have also shown that the absence of long-range spatial
description of the dendrite tip curvature. This limit may be circum- coherence of those oscillations is due to tip splitting events that
vented using a non-uniform discretization of the simulation maintain a disordered cellular array structure, unlike in two-
domain, with an increased accuracy, i.e. a finer mesh, where it dimensions where coherent oscillations are observed [55].
A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36 29

Microgravity experiment t t+ /3 t+2 /3


A EC
D G
≈ 45.6 min B F
0.05
B D G

Area of cell (mm2)


300 µm

0.04

0.03

0.02

0 0.5 1 1.5 2 2.5

Phase-field simulation A E
D G C
≈ 48.1 min B F

Normalized area of cell


1.1 200 µm

0.9 B D G
0 0.5 1 1.5 2 2.5
Time t (h)
Fig. 2. Oscillations in three-dimensional cellular patterns [12]. Locally ordered regions of the array exhibit coherent breathing-mode oscillations. Microgravity experiments
(top) and phase-field simulations (bottom) are in excellent quantitative agreement. In both, cells belonging to three sublattices of the hexagonal array oscillate 2p=3 out-of-
phase with a period s of 45 min.

3.3.2. Orientation transition and twinning in Al–Zn alloys 3.3.3. Intragrain microstructure selection
Atomistic simulations, reviewed in Section 2, have shown that At the scale of a dendritic array, phase-field simulations have
alloying may give rise to substantial changes in interface been shown to reproduce experimentally established scaling laws
free-energy anisotropies [10]. Phase-field simulations of equiaxed for the steady-state selection of primary spacing as a function of
dendritic growth have revealed the existence of a continuous processing conditions and crystal orientation [139]. As the growth
transition of growth direction from h1 0 0i to h1 1 0i as a function velocity is increased, PF simulations can predict the transition of
of anisotropy parameters [63]. A qualitatively similar transition the cell/dendrites growth direction from the orientation of the
was observed to occur as a function of increasing Zn concentration temperature gradient to the orientation of the crystal, as well as
in the fcc Al phase in directionally solidified Al–Zn alloys [63], the morphology transition to degenerate structures at high tem-
suggesting that anisotropy parameters are changing as a function perature gradient and/or high crystal misorientation [4,56,139].
of Zn concentration. This transition has been further character- However, while two-dimensional simulations qualitatively repro-
ized both experimentally by directional solidification of Al–Zn duce realistic scaling laws, full three-dimensional simulations
alloys [47] and numerically [35] by phase-field simulations that remain necessary to reach a quantitative agreement with experi-
model quantitatively those experiments. Those simulations, ments [56].
illustrated in Fig. 3, show a continuous transition of dendrite Provatas and co-workers have also shown that two-dimensional
growth directions over a narrow range of anisotropy parameters, PF simulations can reproduce transient spacing evolution in direc-
producing hyperbranched structures in the form of textured tional solidification [3]. This evolution is characterized by
seaweeds [35]. plateau-like regions of constant primary spacing that are observed
Further investigations on microstructure selection in Al–Zn experimentally [90], which can be related to the hysteretic nature
alloys have focused on the growth morphology of twinned den- of the dynamical selection of the spacing as a function of pulling
drites. Experimental [122] and phase-field [123] studies have sug- velocity [61]. Adding the effect of repulsive grain boundaries
gested the presence of a groove at the tip of twinned dendrites, [109], they investigated grain boundary coalescence at high solid
evolving into a doublon morphology, associated with the forma- fraction, the formation of highly segregated liquid pools, and the
tion of small pockets of highly concentrated liquid at the doublon evolution of solid connectivity during late stage solidification
root for relatively high solidification rates. Recently, new experi- [60,102].
ments have brought strong evidence that twinned dendrites may
nucleate on the faces of a fivefold symmetric phase, such as icosa- 3.3.4. Intergrain microstructure selection
hedral quasicrystals or Al45Cr7. This nucleation mechanism pro- Much attention was recently given to grain growth competition
vides a natural explanation for the significant increase in in polycrystalline directional solidification [40,41,87,133,132,139].
twinned dendrites and the subsequent grain refinement in Al–Zn A conceptual framework to understand this competition was orig-
alloys when minute amounts of Cr are added [84]. A similar mech- inally proposed by Walton and Chalmers [147]. They assumed that
anism was reported in gold alloys when adding small amounts of Ir a grain growing at a higher misorientation with respect to the tem-
into the melt [85]. perature gradient direction cannot overgrow a better-oriented
30 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

Fig. 3. Phase-field prediction of the dendrite orientation transition from h1 0 0i to h1 1 0i in directional solidification of Al–Zn alloys, as a function of the two anisotropy
parameters a1 and a2 defined in Eq. (1) [35].

grain because the latter grows with a smaller undercooling and process, and treated statistically in order to obtain relevant quan-
hence ahead of the misoriented grain. However, observations titative data, which makes the task all the more computationally
[164] and phase-field simulations [40,41,133] have suggested challenging.
more complex mechanisms of grain-boundary evolution during With the exception of a few large scale 3D simulations
solidification. [40,41,133], most recent systematic PF studies of columnar growth
Using 2D phase-field, several groups have confirmed the occur- competition were achieved using 2D simulations with one grain
rence of overgrowth of the better-oriented grain at the converging aligned along the temperature gradient direction [87,132,139]. In
GB [87,132,139]. The underlying mechanism was shown to be dif- this configuration, grain elimination is relatively slow [139]. Even
fusive interactions between dendrite tips. As misoriented dendrites in two dimensions, the combination of orientations of both crystals
impinge upon the well-oriented grain, the local spacing progres- has a strong influence on GB orientation selection [40,41], and a
sively decreases in the well-oriented grain close to the GB, and slight misorientation of the better-oriented grain can yield a much
its local undercooling increases, until elimination of the dendrite faster grain elimination, due to the preferential orientation of sec-
closest to the GB [87,132]. However, the frequency of this over- ondary sidebranches [136].
growth mechanism at the converging GB is relatively low, such Furthermore, in order to reproduce realistic experiments, one
that the orientation of the converging GB remains close the tem- needs to consider the 3D orientation of the two (or more) crystals.
perature gradient orientation [139]. This is illustrated in Fig. 4 with 3D simulations reproducing the
In contrast, grain competition at the diverging GB is controlled mechanisms exposed by Gandin et al. in thin sample experiments
by the growth of tertiary branches and appears more complex [52]. When the two grains have their crystalline axes parallel or
[40,41,139]. Since dendritic sidebranches originate from selective normal to the thin sample walls in (a), tertiary branching occurs
amplification of noise, microscopic thermal fluctuations may have preferentially from the misoriented grain and grain elimination is
significant macroscopic consequences on the selection of GB orien- limited. When the misoriented crystal exhibits a second angle with
tation. This was evidenced by running series of 2D simulations of the walls, e.g. 45° in (b), sidebranching from that grain is inhibited,
grain competition with similar control parameters, but different tertiary branches originate from the best oriented grain, and grain
random number seeds in PF calculations [139]. Therefore, orienta- elimination is much faster (see Fig. 12 in [139] for a schematic
tion selection of the diverging GB has to be regarded as a stochastic description).
A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36 31

(a) (b)

x
y

x
y
y
z

Fig. 4. Dendritic grain growth competition in three-dimensional thin-sample directional solidification. Experimental pictures (top) are from [52]. The arrow in each phase-
field simulation (bottom) follows a unique dendrite in time [125].

3.4. Other advances in phase-field modeling recently introduced by Steinbach and co-workers replaces the
interface equilibrium partitioning condition by coupling two con-
3.4.1. Mixed transport centration fields linked by a kinetic equation, describing the
Since the introduction of the no-slip condition at the solid– exchange of components between phases [128]. This approach
liquid interface to model melt flow in phase-field simulations introduces an effective interface permeability parameter, which
[11], several studies have shed light on microstructure selection can be adjusted by comparison to smaller scale thin-interface cal-
mechanisms in the presence of convection. The solid–liquid culations for rapid solidification [49,161]. This method has also
density change in free growth was shown to affect the dendritic been extended to multicomponent alloys [162]. A similar model,
tip selection parameter [131]. However, a reformulation account- based on thermodynamic extremal principle was also recently pro-
ing for density changes [43] provides a good estimation of the tip posed [149,148]. Such approaches were found to successfully pre-
selection parameter independent of the relative density change. dict solute trapping effects experimentally measured in Si–As
Two-dimensional simulations have shown that modulated flow alloys [128,149].
with a given pulse duration can yield deterministic sidebranching Additionally, using a coupled thermal–solutal thin-interface
events [108], much like with a short thermal perturbation of a den- phase-field model, Mullis and co-workers have recently discussed
drite tip in diffusive growth [39]. However, dimensionality plays a the validity of the anti-trapping current in the context of rapid
critical role in the study of convective flows around dendrites [160] solidification, when solute trapping is expected to occur, and pro-
and three-dimensional simulations appear mandatory to achieve posed strategies to adapt the strength of the corrective current to
conclusive studies on fluid flow effect on dendritic microstructure retrieve quantitative predictions [105]. They have also shown an
selection. evolution of the tip selection parameter at high undercooling
An alternative approach to fluid flow consists in replacing the [104], and suggested it as a possible explanation for grain refine-
full resolution of Navier–Stokes equations by Lattice Boltzmann ment in deeply undercooled melt [105,104].
(LB) equations. Whether using a solidification model based on LB
equations [98,129], coupling with phase-field models [93,107], or
with cellular-automaton based models [159], this simple and com- 3.4.3. Multi-phase and multicomponent systems
putationally efficient approach provides a powerful scale-bridging In order to model industrially-relevant alloys and structures,
tool and may yield valuable insight into convective crystal growth one needs quantitative PF models for multi-grain, multi-phase,
in the future. multicomponent systems [114].
For polycrystalline structures, the orientation-field PF model has
been developed that uses only one additional scalar field corre-
3.4.2. Rapid solidification sponding to the grain orientation in 2D [83]. This model naturally
Rapid solidification is of particular interest in the context of describes grain rotation owing to the rotational invariance of the
microstructure selection in processes such as welding, alongside underlying free-energy and dynamical equations. A different for-
with the growing interest in additive manufacturing. A model mulation of this model was recently proposed in an attempt to
32 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

minimize spurious long-range GB interactions associated with the needles that interact with each other through the long-range diffu-
method [64]. However, limitations of those models have been sion field (Fig. 5b) [76,138]. The solid–liquid interface along each
recently emphasized [112]. The multi-phase-field approach, with needle is assumed to be in local equilibrium. The growth dynamics
each grain attached to an individual phase field [127] remains a of each individual needle can then be uniquely prescribed by two
practical solution in common situations where grain rotation, independent conditions. The first condition stems from the fact
which is not described by this approach, is absent. The method is that the concentration gradient normal to the solid–liquid inter-
more computationally demanding, and requires advanced algo- face is largest at the needle tip as depicted schematically in
rithms to avoid a drastic increase in memory usage [57,144]. In Fig. 5c. Based on the form of the diffusion field in the near-tip
addition, polycrystalline solidification can also be modeled with a region on an intermediate scale much larger than q but much
single phase-field and an integer orientation field in situations, smaller than D=V, it is possible to define a solutal flux intensity fac-
such as the growth competition of columnar grains [139], where tor F ðtÞ that represents the instantaneous strength of this gradient
the GB groove is situated deep into the mushy zone and need not averaged over the tip region. F ðtÞ can be calculated using a condi-
be modeled accurately. tion of solute conservation expressed as a contour (surface) inte-
Several approaches have been proposed in recent years to gral around the tip in 2D (3D) [140], which yields the result that
handle quantitative PF simulations of multicomponent alloys F ðtÞ is proportional to the product qV 2 in 2D and qV in 3D
solidification [42,31,100,110,162]. One approach exploits a grand- [138,140]. The second condition is identical to the standard micro-
potential functional [31] that uses the chemical potential as scopic solvability condition that relates the product q2 V (also the
dynamical variable instead of the solute concentration, similarly selection parameter r ¼ 2Dd0 =q2 V, with d0 the solute capillarity
to the approach introduced for binary alloys [113]. This grand length at the tip) to the strength of the interfacial energy aniso-
potential formulation, which has also been extended to rapid solid- tropy [86,9]. While this condition was originally derived for
ification [34], appears to be a promising candidate to handle alloys steady-state growth conditions [86,9], subsequent PF simulations
with arbitrary phase diagrams, with the potential to be coupled demonstrated that q2 V remains approximately constant in time
with CalPhaD calculations (see e.g. [42,18,45,163]). during transient growth conditions where both q and V are tempo-
rally varying [115]. This finding reflects the fact that the solvability
4. Towards the grain scale condition that determines q2 V is obtained by a solution to the
interface growth equations on the tip scale (Fig. 5d), and hence
In concentrated alloys, solidification usually produces grains in should be expected to remain valid in situations where the diffu-
the form of hierarchical networks of sharp needle-like branches, sion field is slowly evolving on a scale much larger than q, as seen
like in Fig. 5a. Thus, several orders of magnitude separate the scale in PF simulations [115]. This key property allows us to assume that
of the tip radius q and the diffusion length D=V where D is the qðtÞ2 VðtÞ remains constant in time in the DNN model and hence,
solute diffusivity and V the growth velocity. Such morphologies
together with the other condition F ðtÞ  qðtÞV 2 ðtÞ in 2D or
are challenging to model with PF simulations that need to resolve
F ðtÞ  qðtÞVðtÞ in 3D, to determine qðtÞ and VðtÞ independently
the solid–liquid interface of each branch. To overcome this limita-
for each needle in the dendritic network. This approach rigorously
tion, we recently developed a multiscale Dendritic Needle Network
bridges the dynamics of the dendrite tip and diffusion scales.
(DNN) model [138,140] that is briefly reviewed in this section.
Sidebranches are generated periodically on the side of each
The DNN model represents the hierarchical dendritic network
dendrite, including fluctuations to represent the selective amplifi-
(e.g. Fig. 5a) by its ensemble of branches approximated as thin
cation of thermal noise at the origin of dendritic sidebranching
[138]. The time evolution of qV 2 (in 2D) or qV (in 3D) as a function
of the flux intensity factor F ðtÞ captures the evolution of the solute
(b) (a) field surrounding each dendrite tip and the resulting evolution of
VðtÞ, hence making it possible to model branches interactions
and their unsteady growth competition.
The DNN model reproduces the Ivantsov steady-state solutions
for a single free dendrite in 2D and 3D [140], as well as analytical
laws for early-stage transient growth of idealized cross-shaped
equiaxed crystals in 2D [138] and 3D [140]. (The sharp needle for-
mulation in [138] actually reproduces an approximation of the
Ivantsov solution at low supersaturation.) The three-dimensional
(c) (d) DNN model applied to unsteady competitive sidebranch growth
[137] yields a good agreement with experimentally measured scal-
ing laws for dendrite envelopes [88,94]. First applications to direc-
tional solidification experiments show an excellent agreement on
primary spacing selection [140]. For instance, in Fig. 6, the direc-
tional solidification of an Al–12 at.%Cu alloy at a temperature gra-
dient G ¼ 62 K/cm and a growth velocity V ¼ 115 lm/s yields an
average measured spacing Kexp :  258 lm, while DNN simulations
predict an average spacing K3D  237 lm.
Fig. 5. The multiscale Dendritic Needle Network (DNN) model represents a
dendritic grain – such as the ammonium-bromide crystal in (a) from [33] – by a
network of thin needle crystals that interact through the solute field, as illustrated
in (b). The growth dynamics of each individual needle branch is obtained by 5. Outlook
combining two independent conditions at two distinct length scales: (c) a solute
conservation equation at a length scale much larger than the needle tip radius q but The last two decades have witnessed major progress in atomis-
much smaller than the diffusion length D=V, and (d) a solvability condition at the
scale of the needle tip radius q [138]. The u field is a dimensionless measure of the
tic simulations methods to compute key interfacial properties
departure of the concentration field from its equilibrium value ui for a reference flat impacting solidification microstructures. However, quantitative
interface. predictions of those properties remain scarce for alloys. Atomistic
A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36 33

Experiment DNN 3D 500 µm

(a) (b)

Fig. 6. Spacing selection during directional solidification of an Al–12 at.%Cu alloy at G ¼ 62 K/cm and V ¼ 115 lm/s in the direction of the crystal growth (tilted by 30° with
respect to gravity and thermal gradient). The X-ray image in (a) and the 3D simulation in (b) are represented at the same scale [140].

modeling is needed to predict interfacial energy anisotropies in the of Materials Sciences and Engineering and Los Alamos National
Al–Zn system in order to link more directly continuum scale mod- Laboratory, operated by Los Alamos National Security, LLC under
eling to experimentally observed changes of dendrite growth Contract No. DE-AC52-06NA25396 for the U.S. Department of
directions as a function of Zn concentration. Current atomistic Energy.
modeling predictions also fall short explaining the full spectrum
of dendrite growth directions and morphologies observed in rapid
References
alloy solidification [27]. In particular, existing atomistic computa-
tions of interface kinetic anisotropy predict that dendrite growth [1] A. Adland, A. Karma, R. Spatschek, D. Buta, M. Asta, Phase-field-crystal study
along h1 0 0i directions should be favored at large growth rate of grain boundary premelting and shearing in bcc iron, Phys. Rev. B 87 (2)
while other directions h1 1 1i, and disordered growth modes that (2013) 024110.
[2] N. Ahmad, A. Wheeler, W. Boettinger, G. McFadden, Solute trapping and
may underly grain refinement, are also observed. In addition, ato- solute drag in a phase-field model of rapid solidification, Phys. Rev. E 58 (3)
mistic modeling studies of interface coalescence, which have so far (1998) 3436.
focused on pure metals, remain to be extended to binary alloys. [3] M. Amoorezaei, S. Gurevich, N. Provatas, Spacing characterization in Al–Cu
alloys directionally solidified under transient growth conditions, Acta Mater.
The results of such studies, used in conjunction with mesoscale 58 (18) (2010) 6115–6124.
modeling, could potentially help explain how hot cracking sensi- [4] M. Amoorezaei, S. Gurevich, N. Provatas, Orientation selection in solidification
tivity depends on the choice and concentration of alloying ele- patterning, Acta Mater. 60 (2) (2012) 657–663.
[5] M. Asta, C. Beckermann, A. Karma, W. Kurz, R. Napolitano, M. Plapp, G. Purdy,
ments. The increased availability of EAM interatomic potentials
M. Rappaz, R. Trivedi, Solidification microstructures and solid-state parallels:
for different alloy systems should facilitate such studies. Recent developments, future directions, Acta Mater. 57 (4) (2009) 941–971.
At the continuum scale, PF modeling has been used successfully [6] B.P. Athreya, N. Goldenfeld, J.A. Dantzig, M. Greenwood, N. Provatas, Adaptive
mesh computation of polycrystalline pattern formation using a
to model a wide range of solidification microstructures, with the
renormalization-group reduction of the phase-field crystal model, Phys.
last few years witnessing the most rapid advances to simulate Rev. E 76 (5) (2007) 056706.
experimentally relevant length and time scales in 3D. However, [7] M.J. Aziz, T. Kaplan, Continuous growth model for interface motion during
despite advances in computing technologies, quantitative PF mod- alloy solidification, Acta Metall. 36 (8) (1988) 2335–2347.
[8] R. Balluffi, R. Maurer, On rotating sphere-on-a-plate experiments and the
eling still relies on formulating models with the desired thin- question of whether high angle grain boundaries melt below bulk melting
interface limit. While there has been significant recent advances temperatures, Scr. Metall. 22 (5) (1988) 709–713.
in formulating new PF models for a wider range of transport prop- [9] A. Barbieri, J. Langer, Predictions of dendritic growth rates in the linearized
solvability theory, Phys. Rev. A 39 (10) (1989) 5314.
erties, as well as concentrated and multi-component alloys, exten- [10] C. Becker, D. Olmsted, M. Asta, J. Hoyt, S. Foiles, Atomistic underpinnings for
sions of those models and more extensive benchmarking of their orientation selection in alloy dendritic growth, Phys. Rev. Lett. 98 (12) (2007)
predictions remain needed. 125701.
[11] C. Beckermann, H.-J. Diepers, I. Steinbach, A. Karma, X. Tong, Modeling melt
At the grain scale, multiscale approaches like the dendritic nee- convection in phase-field simulations of solidification, J. Comput. Phys. 154
dle network model should pave the way to quantitative studies of (2) (1999) 468–496.
dendritic microstructure selection mechanisms at length and time [12] N. Bergeon, D. Tourret, L. Chen, J.-M. Debierre, R. Guérin, A. Ramirez, B. Billia,
A. Karma, R. Trivedi, Spatiotemporal dynamics of oscillatory cellular patterns
scales out of the reach of phase-field models. Large scale simula-
in three-dimensional directional solidification, Phys. Rev. Lett. 110 (22)
tions of dendritic structures should yield significant insight into (2013) 226102.
complex mechanisms such as grain growth competition [139], [13] M. Berghoff, M. Selzer, A. Choudhury, B. Nestler, Efficient techniques for
bridging from atomic to mesoscopic scale in phase-field simulations, J.
the long range effect of fluid flow on microstructure selection
Comput. Methods Sci. Eng. (2013).
[126], and the columnar-to-equiaxed transition [165]. [14] J. Berry, K. Elder, M. Grant, Melting at dislocations and grain boundaries: a
phase field crystal study, Phys. Rev. B 77 (22) (2008) 224114.
[15] W. Boettinger, S. Coriell, A. Greer, A. Karma, W. Kurz, M. Rappaz, R. Trivedi,
Solidification microstructures: recent developments, future directions, Acta
Acknowledgements Mater. 48 (1) (2000) 43–70.
[16] W. Boettinger, J. Warren, C. Beckermann, A. Karma, Phase-field simulation of
A.K. acknowledges support of grant DEFG02-07ER46400 from solidification, Annu. Rev. Mater. Res. 32 (1) (2002) 163–194.
[17] P. Bollada, C. Goodyer, P. Jimack, A. Mullis, F. Yang, Three dimensional
the US Department of Energy, Office of Basic Energy Sciences, for
thermal-solute phase field simulation of binary alloy solidification, J. Comput.
support of atomistic-scale phase-field-crystal modeling of Phys. 287 (2015) 130–150.
interfacial properties, and NASA for continuum-scale phase-field [18] B. Böttger, J. Eiken, M. Apel, Phase-field simulation of microstructure
and dendritic-needle-network modeling of cellular/dendritic formation in technical castings—a self-consistent homoenthalpic approach
to the micro–macro problem, J. Comput. Phys. 228 (18) (2009) 6784–6795.
microstructures. D.T. is supported by Amy Clarke’s Early Career [19] G. Boussinot, E.A. Brener, Achieving realistic interface kinetics in phase-field
award from the U.S. DOE, Office of Basic Energy Sciences, Division models with a diffusional contrast, Phys. Rev. E 89 (2014) 060402.
34 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

[20] G. Boussinot, E.A. Brener, C. Hueter, R. Spatschek, Elimination of Surface [52] C.-A. Gandin, M. Eshelman, R. Trivedi, Orientation dependence of primary
Diffusion in the Non-diagonal Phase Field Model, 2014. Available from: dendrite spacing, Metall. Mater. Trans. A 27 (9) (1996) 2727–2739.
<1411.0523>. [53] Y.F. Gao, Y. Yang, D.Y. Sun, M. Asta, J.J. Hoyt, Molecular dynamics simulations
[21] J. Bragard, A. Karma, Y.H. Lee, M. Plapp, Linking phase-field and atomistic of the crystal–melt interface mobility in hcp Mg and bcc Fe, J. Cryst. Growth
simulations to model dendritic solidification in highly undercooled melts, 312 (21) (2010) 3238–3242.
Interface Sci. 10 (2–3) (2002) 121–136. [54] W.L. George, J.A. Warren, A parallel 3d dendritic growth simulator using the
[22] E. Brener, V. Mel’Nikov, Pattern selection in two-dimensional dendritic phase-field method, J. Comput. Phys. 177 (2) (2002) 264–283.
growth, Adv. Phys. 40 (1) (1991) 53–97. [55] M. Georgelin, A. Pocheau, Oscillatory instability, limit cycle, and transition to
[23] E.A. Brener, G. Boussinot, Kinetic cross coupling between nonconserved and doublets in directional solidification, Phys. Rev. Lett. 79 (14) (1997) 2698.
conserved fields in phase field models, Phys. Rev. E 86 (6) (2012) 060601. [56] J. Ghmadh, J.-M. Debierre, J. Deschamps, M. Georgelin, R. Guérin, A. Pocheau,
[24] J. Broughton, G. Gilmer, Thermodynamic criteria for grain-boundary melting: Directional solidification of inclined structures in thin samples, Acta Mater.
a molecular-dynamics study, Phys. Rev. Lett. 56 (25) (1986) 2692. 74 (2014) 255–267.
[25] J.W. Cahn, Y. Mishin, A. Suzuki, Coupling grain boundary motion to shear [57] J. Gruber, N. Ma, Y. Wang, A. Rollett, G. Rohrer, Sparse data structure and
deformation, Acta Mater. 54 (19) (2006) 4953–4975. algorithm for the phase field method, Model. Simul. Mater. Sci. Eng. 14 (7)
[26] T. Carozzani, C.-A. Gandin, H. Digonnet, M. Bellet, K. Zaidat, Y. Fautrelle, Direct (2006) 1189.
simulation of a solidification benchmark experiment, Metall. Mater. Trans. A [58] Z. Guo, J. Mi, P. Grant, Phase field simulation of multi-dendrite growth in a
44 (2) (2013) 873–887. coupled thermal-solute-convective environment, IOP Conference Series:
[27] E.G. Castle, A.M. Mullis, R.F. Cochrane, Evidence for an extended transition in Materials Science and Engineering, vol. 33, IOP Publishing, 2012, p. 012101.
growth orientation and novel dendritic seaweed structures in undercooled [59] V.K. Gupta, D.-H. Yoon, H.M. Meyer, J. Luo, Thin intergranular films and solid-
Cu–8.9 wt% Ni, J. Alloy. Compd. 615 (2014) S612–S615. state activated sintering in nickel-doped tungsten, Acta Mater. 55 (9) (2007)
[28] S.-W. Chan, J. Liu, R. Balluffi, Test for a possible ‘‘melting” transition in grain 3131–3142.
boundaries in aluminum near the melting point, Scr. Metall. 19 (10) (1985) [60] S. Gurevich, M. Amoorezaei, D. Montiel, N. Provatas, Evolution of
1251–1255. microstructural length scales during solidification of magnesium alloys,
[29] L.-Q. Chen, Phase-field models for microstructure evolution, Annu. Rev. Acta Mater. 60 (8) (2012) 3287–3295.
Mater. Res. 32 (1) (2002) 113–140. [61] S. Gurevich, M. Amoorezaei, N. Provatas, Phase-field study of spacing
[30] A. Chernov, Notes on interface growth kinetics 50 years after Burton, Cabrera evolution during transient growth, Phys. Rev. E 82 (5) (2010) 051606.
and Frank, J. Cryst. Growth 264 (4) (2004) 499–518. [62] S. Gurevich, A. Karma, M. Plapp, R. Trivedi, Phase-field study of three-
[31] A. Choudhury, B. Nestler, Grand-potential formulation for multicomponent dimensional steady-state growth shapes in directional solidification, Phys.
phase transformations combined with thin-interface asymptotics of the Rev. E 81 (1) (2010) 011603.
double-obstacle potential, Phys. Rev. E 85 (2) (2012) 021602. [63] T. Haxhimali, A. Karma, F. Gonzales, M. Rappaz, Orientation selection in
[32] D.R. Clarke, On the equilibrium thickness of intergranular glass phases in dendritic evolution, Nat. Mater. 5 (8) (2006) 660–664.
ceramic materials, J. Am. Ceram. Soc. 70 (1) (1987) 15–22. [64] H. Henry, J. Mellenthin, M. Plapp, Orientation-field model for polycrystalline
[33] Y. Couder, J. Maurer, R. González-Cinca, A. Hernández-Machado, Side-branch solidification with a singular coupling between order and orientation, Phys.
growth in two-dimensional dendrites. I. Experiments, Phys. Rev. E 71 (3) Rev. B 86 (5) (2012) 054117.
(2005) 031602. [65] J. Hoyt, M. Asta, A. Karma, Method for computing the anisotropy of the solid–
[34] D.A. Danilov, V.G. Lebedev, P.K. Galenko, A grand potential approach to phase- liquid interfacial free energy, Phys. Rev. Lett. 86 (24) (2001) 5530.
field modeling of rapid solidification, J. Non-Equilib. Thermodyn. 39 (2) [66] J. Hoyt, M. Asta, A. Karma, Atomistic and continuum modeling of dendritic
(2014) 93–111. solidification, Mater. Sci. Eng.: R: Rep. 41 (6) (2003) 121–163.
[35] J. Dantzig, P. Di Napoli, J. Friedli, M. Rappaz, Dendritic growth morphologies [67] J. Hoyt, D. Olmsted, S. Jindal, M. Asta, A. Karma, Method for computing short-
in Al–Zn alloys – Part II: Phase-field computations, Metall. Mater. Trans. A 44 range forces between solid–liquid interfaces driving grain boundary
(12) (2013) 5532–5543. premelting, Phys. Rev. E 79 (2) (2009) 020601.
[36] J. Dash, H. Fu, J. Wettlaufer, The premelting of ice and its environmental [68] J. Hoyt, B. Sadigh, M. Asta, S. Foiles, Kinetic phase field parameters for the Cu–
consequences, Rep. Prog. Phys. 58 (1) (1995) 115. Ni system derived from atomistic computations, Acta Mater. 47 (11) (1999)
[37] S. Divinski, M. Lohmann, C. Herzig, B. Straumal, B. Baretzky, W. Gust, Grain- 3181–3187.
boundary melting phase transition in the Cu–Bi system, Phys. Rev. B 71 (10) [69] J.J. Hoyt, M. Asta, A. Karma, Atomistic simulation methods for computing the
(2005) 104104. kinetic coefficient in solid–liquid systems, Interface Sci. 10 (2–3) (2002) 181–
[38] B. Echebarria, R. Folch, A. Karma, M. Plapp, Quantitative phase-field model of 189.
alloy solidification, Phys. Rev. E 70 (6) (2004) 061604. [70] J.J. Hoyt, M. Asta, A. Karma, Atomistic simulations of solute trapping and
[39] B. Echebarria, A. Karma, S. Gurevich, Onset of sidebranching in directional solute drag, in: D.M. Herlach, D.M. Matson (Eds.), Solidification of
solidification, Phys. Rev. E 81 (2) (2010) 021608. Containerless Undercooled Melts, Wiley, 2012, pp. 363–380.
[40] J. Eiken, Dendritic orientation selection and growth texture evolution during [71] T. Hsieh, R. Balluffi, Observations of roughening/de-faceting phase transitions
directional solidification of magnesium-based alloys investigated by phase- in grain boundaries, Acta Metall. 37 (8) (1989) 2133–2139.
field simulations, in: Second International Conference on Advances in [72] T. Ihle, Competition between kinetic and surface tension anisotropy in
Solidification Processes (ICASP-2), June 17–20, 2008, University of Leoben, dendritic growth, Eur. Phys. J. B – Conden. Matter Complex Syst. 16 (2) (2000)
Austria, 2008. 337–344.
[41] J. Eiken, Dendritic growth texture evolution in Mg-based alloys investigated [73] F. Inoko, T. Okada, T. Muraga, Y. Nakano, T. Yoshikawa, Strain induced grain
by phase-field simulation, Int. J. Cast Met. Res. 22 (1–4) (2009) 86–89. boundary premelting in bulk copper bicrystals, Interface Sci. 4 (3–4) (1997)
[42] J. Eiken, B. Böttger, I. Steinbach, Multiphase-field approach for 263–272.
multicomponent alloys with extrapolation scheme for numerical [74] V. Ivanov, Y. Mishin, Dynamics of grain boundary motion coupled to shear
application, Phys. Rev. E 73 (6) (2006) 066122. deformation: an analytical model and its verification by molecular dynamics,
[43] H. Emmerich, The Diffuse Interface Approach in Materials Science: Phys. Rev. B 78 (6) (2008) 064106.
Thermodynamic Concepts and Applications of Phase-field Models, vol. 73, [75] K.A. Jackson, K.M. Beatty, K.A. Gudgel, An analytical model for non-
Springer Science & Business Media, 2003. equilibrium segregation during crystallization, J. Cryst. Growth 271 (3)
[44] V. Fallah, M. Amoorezaei, N. Provatas, S. Corbin, A. Khajepour, Phase-field (2004) 481–494.
simulation of solidification morphology in laser powder deposition of Ti–Nb [76] A. Karma, Dendritic growth, in: V. Fleury, J.F. Gouyet, M. Lonetti (Eds.),
alloys, Acta Mater. 60 (4) (2012) 1633–1646. Branching in Nature, Springer, 2001, pp. 365–402.
[45] J.-L. Fattebert, M. Wickett, P. Turchi, Phase-field modeling of coring during [77] A. Karma, Phase-field formulation for quantitative modeling of alloy
solidification of Au–Ni alloy using quaternions and CALPHAD input, Acta solidification, Phys. Rev. Lett. 87 (11) (2001) 115701.
Mater. 62 (2014) 89–104. [78] A. Karma, W.-J. Rappel, Numerical simulation of three-dimensional dendritic
[46] S.J. Fensin, D. Olmsted, D. Buta, M. Asta, A. Karma, J. Hoyt, Structural growth, Phys. Rev. Lett. 77 (19) (1996) 4050.
disjoining potential for grain-boundary premelting and grain coalescence [79] A. Karma, W.-J. Rappel, Quantitative phase-field modeling of dendritic growth
from molecular-dynamics simulations, Phys. Rev. E 81 (3) (2010) 031601. in two and three dimensions, Phys. Rev. E 57 (4) (1998) 4323.
[47] J. Friedli, J. Fife, P. Di Napoli, M. Rappaz, Dendritic growth morphologies in Al– [80] A. Karma, Z.T. Trautt, Y. Mishin, Relationship between equilibrium
Zn alloys - Part I: X-ray tomographic microscopy, Metall. Mater. Trans. A 44 fluctuations and shear-coupled motion of grain boundaries, Phys. Rev. Lett.
(12) (2013) 5522–5531. 109 (9) (2012) 095501.
[48] T. Frolov, Y. Mishin, Solid–liquid interface free energy in binary systems: [81] D.A. Kessler, J. Koplik, H. Levine, Pattern selection in fingered growth
theory and atomistic calculations for the (110) CuAg interface, J. Chem. Phys. phenomena, Adv. Phys. 37 (3) (1988) 255–339.
131 (2009) 054702. [82] R. Kikuchi, J.W. Cahn, Grain-boundary melting transition in a two-
[49] P. Galenko, E. Abramova, D. Jou, D. Danilov, V. Lebedev, D. Herlach, Solute dimensional lattice-gas model, Phys. Rev. B 21 (5) (1980) 1893.
trapping in rapid solidification of a binary dilute system: a phase-field study, [83] R. Kobayashi, J.A. Warren, W.C. Carter, A continuum model of grain
Phys. Rev. E 84 (4) (2011) 041143. boundaries, Physica D 140 (1) (2000) 141–150.
[50] P. Galenko, S. Sobolev, Local nonequilibrium effect on undercooling in rapid [84] G. Kurtuldu, P. Jarry, M. Rappaz, Influence of Cr on the nucleation of primary
solidification of alloys, Phys. Rev. E 55 (1) (1997) 343. Al and formation of twinned dendrites in Al–Zn–Cr alloys: Can icosahedral
[51] C.-A. Gandin, Modeling of solidification: Grain structures and segregations in solid clusters play a role?, Acta Mater 61 (19) (2013) 7098–7108.
metallic alloys, C.R. Phys. 11 (3) (2010) 216–225. [85] G. Kurtuldu, A. Sicco, M. Rappaz, Icosahedral quasicrystal-enhanced nucleation
of the fcc phase in liquid gold alloys, Acta Mater. 70 (2014) 240–248.
A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36 35

[86] J.S. Langer, Chance and matter, in: Lectures on the Theory of Pastern [120] M. Rappaz, A. Jacot, W. Boettinger, Last-stage solidification of alloys:
Formation, Les Houches, Session XLVI, North-Holland, Amsterdam, 1987, pp. theoretical model of dendrite-arm and grain coalescence, Metall. Mater.
629–711. Trans. A 34 (3) (2003) 467–479.
[87] J. Li, Z. Wang, Y. Wang, J. Wang, Phase-field study of competitive dendritic [121] V. Sai Pavan Kumar Bhogireddy, C. Hüter, J. Neugebauer, I. Steinbach, A.
growth of converging grains during directional solidification, Acta Mater. 60 Karma, R. Spatschek, Phase-field modeling of grain-boundary premelting
(4) (2012) 1478–1493. using obstacle potentials, Phys. Rev. E 90 (1) (2014) 012401.
[88] Q. Li, C. Beckermann, Evolution of the sidebranch structure in free dendritic [122] M. Salgado-Ordorica, P. Burdet, M. Cantoni, M. Rappaz, Study of the twinned
growth, Acta Mater. 47 (8) (1999) 2345–2356. dendrite tip shape ii: Experimental assessment, Acta Mater. 59 (13) (2011)
[89] A.E. Lobkovsky, J.A. Warren, Phase field model of premelting of grain 5085–5091.
boundaries, Physica D 164 (3) (2002) 202–212. [123] M. Salgado-Ordorica, J.-L. Desbiolles, M. Rappaz, Study of the twinned
[90] W. Losert, B. Shi, H. Cummins, J. Warren, Spatial period-doubling instability of dendrite tip shape i: Phase-field modeling, Acta Mater. 59 (13) (2011) 5074–
dendritic arrays in directional solidification, Phys. Rev. Lett. 77 (5) (1996) 889. 5084.
[91] J. Luo, V. Gupta, D. Yoon, H. Meyer III, Segregation-induced grain boundary [124] T. Shimokawabe, T. Aoki, T. Takaki, A. Yamanaka, A. Nukada, T. Endo, N.
premelting in nickel-doped tungsten, Appl. Phys. Lett. 87 (23) (2005) 231902. Maruyama, S. Matsuoka, Peta-scale phase-field simulation for dendritic
[92] R. Masumura, M. Glicksman, C. Vold, Absolute solid–liquid and grain solidification on the TSUBAME 2.0 supercomputer, in: 2011 International
boundary energies of bismuth, Scr. Metall. 6 (10) (1972) 943–946. Conference for High Performance Computing, Networking, Storage and
[93] D. Medvedev, K. Kassner, Lattice Boltzmann scheme for crystal growth in Analysis (SC), IEEE, 2011, pp. 1–11.
external flows, Phys. Rev. E 72 (5) (2005) 056703. [125] Y. Song, A. Karma, unpublished work, 2015.
[94] A. Melendez, C. Beckermann, Measurements of dendrite tip growth and [126] I. Steinbach, Pattern formation in constrained dendritic growth with solutal
sidebranching in succinonitrile–acetone alloys, J. Cryst. Growth 340 (1) buoyancy, Acta Mater. 57 (9) (2009) 2640–2645.
(2012) 175–189. [127] I. Steinbach, F. Pezzolla, B. Nestler, M. Seeßelberg, R. Prieler, G. Schmitz, J.
[95] J. Mellenthin, A. Karma, M. Plapp, Phase-field crystal study of grain-boundary Rezende, A phase field concept for multiphase systems, Physica D 94 (3)
premelting, Phys. Rev. B 78 (18) (2008) 184110. (1996) 135–147.
[96] M. Mendelev, M. Rahman, J. Hoyt, M. Asta, Molecular-dynamics study of [128] I. Steinbach, L. Zhang, M. Plapp, Phase-field model with finite interface
solid–liquid interface migration in fcc metals, Modell. Simul. Mater. Sci. Eng. dissipation, Acta Mater. 60 (6) (2012) 2689–2701.
18 (7) (2010) 074002. [129] D. Sun, M. Zhu, S. Pan, D. Raabe, Lattice Boltzmann modeling of dendritic
[97] L. Mikheev, A. Chernov, Mobility of a diffuse simple crystal–melt interface, J. growth in a forced melt convection, Acta Mater. 57 (6) (2009) 1755–1767.
Cryst. Growth 112 (2) (1991) 591–596. [130] D.Y. Sun, M. Asta, J.J. Hoyt, Crystal–melt interfacial free energies and
[98] W. Miller, S. Succi, D. Mansutti, Lattice Boltzmann model for anisotropic mobilities in fcc and bcc Fe, Phys. Rev. B 69 (17) (2004) 174103.
liquid–solid phase transition, Phys. Rev. Lett. 86 (16) (2001) 3578. [131] Y. Sun, C. Beckermann, Effect of solid–liquid density change on dendrite tip
[99] Y. Mishin, W. Boettinger, J. Warren, G. McFadden, Thermodynamics of grain velocity and shape selection, J. Cryst. Growth 311 (19) (2009) 4447–4453.
boundary premelting in alloys. I. Phase-field modeling, Acta Mater. 57 (13) [132] T. Takaki, M. Ohno, T. Shimokawabe, T. Aoki, Two-dimensional phase-field
(2009) 3771–3785. simulations of dendrite competitive growth during the directional
[100] N. Moelans, A quantitative and thermodynamically consistent phase-field solidification of a binary alloy bicrystal, Acta Mater. 81 (2014) 272–283.
interpolation function for multi-phase systems, Acta Mater. 59 (3) (2011) [133] T. Takaki, T. Shimokawabe, M. Ohno, A. Yamanaka, T. Aoki, Unexpected
1077–1086. selection of growing dendrites by very-large-scale phase-field simulation, J.
[101] J. Monk, Y. Yang, M.I. Mendelev, M. Asta, J.J. Hoyt, D.Y. Sun, Determination of Cryst. Growth 382 (2013) 21–25.
the crystal–melt interface kinetic coefficient from molecular dynamics [134] M. Tang, W.C. Carter, R.M. Cannon, Diffuse interface model for structural
simulations, Modell. Simul. Mater. Sci. Eng. 18 (1) (2010) 015004. transitions of grain boundaries, Phys. Rev. B 73 (2) (2006) 024102.
[102] D. Montiel, S. Gurevich, N. Ofori-Opoku, N. Provatas, Characterization of late- [135] M. Tang, W.C. Carter, R.M. Cannon, Grain boundary transitions in binary
stage equiaxed solidification of alloys, Acta Mater. 77 (2014) 183–190. alloys, Phys. Rev. Lett. 97 (7) (2006) 075502.
[103] D. Montiel, L. Liu, L. Xiao, Y. Zhou, N. Provatas, Microstructure analysis of az31 [136] D. Tourret, unpublished work, 2015.
magnesium alloy welds using phase-field models, Acta Mater. 60 (16) (2012) [137] D. Tourret, Karma, Three-dimensional Dendritic Needle Network Model for
5925–5932. Alloy Solidification, in preparation.
[104] A. Mullis, Prediction of the operating point of dendrites growing under [138] D. Tourret, A. Karma, Multiscale dendritic needle network model of alloy
coupled thermosolutal control at high growth velocity, Phys. Rev. E 83 (6) solidification, Acta Mater. 61 (17) (2013) 6474–6491.
(2011) 061601. [139] D. Tourret, A. Karma, Growth competition of columnar dendritic grains: a
[105] A. Mullis, J. Rosam, P. Jimack, Solute trapping and the effects of anti-trapping phase-field study, Acta Mater. 82 (2015) 64–83.
currents on phase-field models of coupled thermo-solutal solidification, J. [140] D. Tourret, A. Karma, A. Clarke, P. Gibbs, S. Imhoff, Three-dimensional
Cryst. Growth 312 (11) (2010) 1891–1897. dendritic needle network model with application to Al–Cu directional
[106] B. Nestler, A 3d parallel simulator for crystal growth and solidification in solidification experiments, in: IOP Conference Series: Materials Science and
complex alloy systems, J. Cryst. Growth 275 (1) (2005) e273–e278. Engineering, vol. 84, issue no. 1, 2015, p. 012082.
[107] B. Nestler, A. Aksi, M. Selzer, Combined lattice Boltzmann and phase-field [141] Z. Trautt, A. Adland, A. Karma, Y. Mishin, Coupled motion of asymmetrical tilt
simulations for incompressible fluid flow in porous media, Math. Comput. grain boundaries: molecular dynamics and phase field crystal simulations,
Simul. 80 (7) (2010) 1458–1468. Acta Mater. 60 (19) (2012) 6528–6546.
[108] H. Neumann-Heyme, K. Eckert, A. Voigt, S. Odenbach, Growth of a free [142] Z. Trautt, Y. Mishin, Grain boundary migration and grain rotation studied by
dendrite in pure substances under modulated flow conditions, IOP molecular dynamics, Acta Mater. 60 (5) (2012) 2407–2424.
Conference Series: Materials Science and Engineering, vol. 33, IOP [143] D. Turnbull, Formation of crystal nuclei in liquid metals, J. Appl. Phys. 21 (10)
Publishing, 2012, p. 012106. (1950) 1022–1028.
[109] N. Ofori-Opoku, N. Provatas, A quantitative multi-phase field model of [144] L. Vanherpe, N. Moelans, B. Blanpain, S. Vandewalle, Bounding box algorithm
polycrystalline alloy solidification, Acta Mater. 58 (6) (2010) 2155–2164. for three-dimensional phase-field simulations of microstructural evolution in
[110] M. Ohno, Quantitative phase-field modeling of nonisothermal solidification polycrystalline materials, Phys. Rev. E 76 (5) (2007) 056702.
in dilute multicomponent alloys with arbitrary diffusivities, Phys. Rev. E 86 [145] C. Vold, M. Glicksman, Behavior of grain boundaries near the melting point,
(5) (2012) 051603. in: The Nature and Behavior of Grain Boundaries, Metallurgical Society of
[111] D.L. Olmsted, D. Buta, A. Adland, S.M. Foiles, M. Asta, A. Karma, Dislocation- AIME, Plenum Press, New York, 1972, pp. 171–183.
pairing transitions in hot grain boundaries, Phys. Rev. Lett. 106 (4) (2011) [146] S. Von Alfthan, K. Kaski, A. Sutton, Molecular dynamics simulations of
046101. temperature-induced structural transitions at twist boundaries in silicon,
[112] M. Plapp, Remarks on some open problems in phase-field modelling of Phys. Rev. B 76 (24) (2007) 245317.
solidification, Philos. Mag. 91 (1) (2011) 25–44. [147] D. Walton, B. Chalmers, The origin of the preferred orientation in the
[113] M. Plapp, Unified derivation of phase-field models for alloy solidification columnar zone of ingots, Trans. Am. Inst. Min. Metall. Eng. 215 (3) (1959)
from a grand-potential functional, Phys. Rev. E 84 (3) (2011) 031601. 447–457.
[114] S. Akamatsu, M. Plapp, Eutectic and peritectic solidification patterns, Curr. [148] H. Wang, W. Kuang, X. Zhang, F. Liu, A hyperbolic phase-field model for rapid
Opin. Solid State Mater. Sci. 20 (1) (2016) 46–54. solidification of a binary alloy, J. Mater. Sci. 50 (3) (2015) 1277–1286.
[115] M. Plapp, A. Karma, Multiscale random-walk algorithm for simulating [149] H. Wang, F. Liu, G. Ehlen, D. Herlach, Application of the maximal entropy
interfacial pattern formation, Phys. Rev. Lett. 84 (8) (2000) 1740. production principle to rapid solidification: a multi-phase-field model, Acta
[116] N. Provatas, N. Goldenfeld, J. Dantzig, Efficient computation of dendritic Mater. 61 (7) (2013) 2617–2627.
microstructures using adaptive mesh refinement, Phys. Rev. Lett. 80 (15) [150] N. Wang, S. Mokadem, M. Rappaz, W. Kurz, Solidification cracking of
(1998) 3308. superalloy single-and bi-crystals, Acta Mater. 52 (11) (2004) 3173–3182.
[117] N. Provatas, M. Greenwood, B. Athreya, N. Goldenfeld, J. Dantzig, Multiscale [151] N. Wang, R. Spatschek, A. Karma, Multi-phase-field analysis of short-range
modeling of solidification: phase-field methods to adaptive mesh refinement, forces between diffuse interfaces, Phys. Rev. E 81 (5) (2010) 051601.
Int. J. Mod. Phys. B 19 (31) (2005) 4525–4565. [152] T. Watanabe, S.-I. Kimura, S. Karashima, The effect of a grain boundary
[118] M. Rappaz, Modeling and characterization of grain structures and defects in structural transformation on sliding in <1 0 1 0>-tilt zinc bicrystals, Philos.
solidification, Curr. Opin. Solid State Mater. Sci. 20 (1) (2016) 37–45. Mag. A 49 (6) (1984) 845–864.
[119] M. Rappaz, C.-A. Gandin, Probabilistic modelling of microstructure formation [153] B. Widom, Structure of the ac interface, J. Chem. Phys. 68 (8) (1978) 3878–
in solidification processes, Acta Metall. Mater. 41 (2) (1993) 345–360. 3883.
36 A. Karma, D. Tourret / Current Opinion in Solid State and Materials Science 20 (2016) 25–36

[154] P. Williams, Y. Mishin, Thermodynamics of grain boundary premelting in [161] L. Zhang, E.V. Danilova, I. Steinbach, D. Medvedev, P.K. Galenko, Diffuse-
alloys. II. Atomistic simulation, Acta Mater. 57 (13) (2009) 3786–3794. interface modeling of solute trapping in rapid solidification: predictions of
[155] K.-A. Wu, A. Karma, J.J. Hoyt, M. Asta, Ginzburg-Landau theory of crystalline the hyperbolic phase-field model and parabolic model with finite interface
anisotropy for bcc-liquid interfaces, Phys. Rev. B 73 (9) (2006) 094101. dissipation, Acta Mater. 61 (11) (2013) 4155–4168.
[156] K.-A. Wu, C.-H. Wang, J.J. Hoyt, A. Karma, Ginzburg-Landau theory of the bcc- [162] L. Zhang, I. Steinbach, Phase-field model with finite interface dissipation:
liquid interface kinetic coefficient, Phys. Rev. B 91 (1) (2015) 014107. extension to multi-component multi-phase alloys, Acta Mater. 60 (6) (2012)
[157] A. Yamanaka, T. Aoki, S. Ogawa, T. Takaki, GPU-accelerated phase-field 2702–2710.
simulation of dendritic solidification in a binary alloy, J. Cryst. Growth 318 (1) [163] L. Zhang, M. Stratmann, Y. Du, B. Sundman, I. Steinbach, Incorporating the
(2011) 40–45. calphad sublattice approach of ordering into the phase-field model with
[158] Y. Yang, H. Humadi, D. Buta, B.B. Laird, D. Sun, J.J. Hoyt, M. Asta, Atomistic finite interface dissipation, Acta Mater. 88 (2015) 156–169.
simulations of nonequilibrium crystal-growth kinetics from alloy melts, [164] Y. Zhou, A. Volek, N. Green, Mechanism of competitive grain growth in
Phys. Rev. Lett. 107 (2) (2011) 025505. directional solidification of a nickel-base superalloy, Acta Mater. 56 (11)
[159] H. Yin, S. Felicelli, L. Wang, Simulation of a dendritic microstructure with the (2008) 2631–2637.
lattice Boltzmann and cellular automaton methods, Acta Mater. 59 (8) (2011) [165] G. Zimmermann, L. Sturz, B. Billia, N. Mangelinck-Noël, D. Liu, H. Nguyen Thi,
3124–3136. N. Bergeon, C.-A. Gandin, D. Browne, C. Beckermann, et al., Columnar-to-
[160] L. Yuan, P.D. Lee, Dendritic solidification under natural and forced convection equiaxed transition in solidification processing of AlSi7 alloys in microgravity
in binary alloys: 2D versus 3D simulation, Modell. Simul. Mater. Sci. Eng. 18 the CETSOL project, Materials Science Forum, vol. 790, Trans Tech Publ, 2014,
(5) (2010) 055008. pp. 12–21.

También podría gustarte