Top Banner

of 28

SOLIDIFICATION MICROSTRUCTURES

Apr 09, 2018

Download

Documents

glen_beck
Welcome message from author
This document is posted to help you gain knowledge. Please leave a comment to let me know what you think about it! Share it to your friends and learn new things together.
Transcript
  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    1/28

    SOLIDIFICATION MICROSTRUCTURES: RECENT

    DEVELOPMENTS, FUTURE DIRECTIONSp

    W. J. BOETTINGER1, S. R. CORIELL 1, A. L. GREER 2, A. KARMA 3, W. KURZ 4{,

    M. RAPPAZ 4 and R. TRIVEDI 5

    1NIST, Gaithersburg, MD 20899, USA, 2Department of Materials Science & Metallurgy, University ofCambridge, Cambridge CB2 3QZ, UK, 3Department of Physics, Northeastern University, Boston, MA

    02115, USA, 4Department of Materials, Swiss Federal Institute of Technology Lausanne, 1015Lausanne EPFL, Switzerland and 5Iowa State University & Ames Lab. USDOE, Ames, IA 50011,

    USA

    (Received 1 June 1999; accepted 15 July 1999)

    AbstractThe status of solidication science is critically evaluated and future directions of research in thistechnologically important area are proposed. The most important advances in solidication science andtechnology of the last decade are discussed: interface dynamics, phase selection, microstructure selection,

    peritectic growth, convection eects, multicomponent alloys, and numerical techniques. It is shown howthe advent of new mathematical techniques (especially phase-eld and cellular automata models) coupledwith powerful computers now allows the following: modeling of complicated interface morphologies, takinginto account not only steady state but also non-steady state phenomena; considering real alloys consistingof many elements through on-line use of large thermodynamic data banks; and taking into account naturaland forced convection eects. A series of open questions and future prospects are also given. It is hopedthat the reader is encouraged to explore this important and highly interesting eld and to add her/his con-tributions to an ever better understanding and modeling of microstructure development. # 2000 ActaMetallurgica Inc. Published by Elsevier Science Ltd. All rights reserved.

    Keywords: Solidication; Microstructure; Theory and modeling (kinetics, transport, diusion); Casting

    1. INTRODUCTION

    Microstructures are at the center of materials

    science and engineering. They are the strategic link

    between materials processing and materials beha-

    vior. Microstructure control is therefore essential

    for any processing activity. One of the most import-

    ant processing routes for many materials, especially

    metals and alloys, is solidication. Over the last

    decade, important advances have been made in our

    fundamental understanding of solidication micro-

    structures. Three main ingredients have contributed

    to this progress: (i) the development of rigorous

    analytical models that have focused on both steady-

    state and non-steady-state microstructure evolution

    with the inclusion of nucleation for the selection of

    phases; (ii) the emergence of accurate simulation

    methods, and in particular phase-eld and cellular

    automata approaches, which have permitted a vali-

    dation of analytical theories as well as enabling pre-

    dictions on grain structure and morphological

    evolution; and (iii) the development of more rened

    experimental techniques that have led to a better

    visualization and characterization of microstructural

    development. The combination of these advances

    now makes it feasible to address long standing

    microstructure formation questions with a higher

    level of scrutiny and rigor, and thus to end this mil-

    lennium in a renaissance period where solidication

    ``science'' is ourishing and solidication technology

    is leading to a better control of materials proces-

    sing. We highlight in this paper the theoretical andexperimental progress made in understanding basic

    aspects of microstructure formation, emphasizing

    especially the critical questions that remain to be

    examined in this scientically highly interesting and

    technologically important area.

    A decade ago, an extensive overview was given

    on the topic which was based on presentations and

    discussions of the rst 1988 Zermatt Workshop

    Acta mater. 48 (2000) 4370

    1359-6454/00/$20.00 # 2000 Acta Metallurgica Inc. Published by Elsevier Science Ltd. All rights reserved.

    PII: S 1 3 5 9 - 6 4 5 4 ( 9 9 ) 0 0 2 8 7 - 6

    www.elsevier.com/locate/actamat

    p

    The Millennium Special Issue A Selection of Major

    Topics in Materials Science and Engineering: Current

    status and future directions, edited by S. Suresh.

    { To whom all correspondence should be addressed.

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    2/28

    dedicated to solidication microstructures [1]. In the

    present overview the most important ndings of the

    second 1998 Zermatt Workshop on Solidication

    Microstructures are presented by the seven keynote

    speakers. (Contributions to this workshop have

    been published in the form of a CD [2].)

    The paper is organized as follows: Section 2

    describes interface pattern formation models;

    Section 3 considers nucleation and growth of a new

    phase during the growth of an existing phase;

    Section 4 emphasizes the action of uid ow on

    microstructures; Section 5 addresses the appli-

    cations of the models to industrially interesting

    alloys containing several solutes; Section 6 includes

    dierent numerical techniques and their potential

    for solving complex problems in which several

    phenomena must be considered simultaneously to

    predict the microstructure.

    2. INTERFACE DYNAMICS

    Microstructures are formed at moving solid

    liquid interfaces. In this section, the evolution of

    interface morphologies of a single phase solid grow-

    ing into a liquid is presented. The growth in an

    undercooled melt of equiaxed dendrites is rst

    described. Directional solidication with planar, cel-

    lular and dendritic interfaces is then considered.

    Some reference is also given to recent work on two-phase growth, such as eutectic and peritectic

    growth.

    2.1. Equiaxed dendritic growth

    During the 1980s, the study of simplied models

    that incorporate surface tension in a consistent way

    led to the novel insight that dendritic growth is con-

    trolled not only by the balance between diusion

    and capillarity, but also in a subtle way by crystal-

    line anisotropy [3, 4]. This insight led to the advent

    of microscopic solvability theory to predict the

    selected dendrite tip velocity and tip radius [5, 6].

    Over the last decade, this theory has been extended

    to three dimensions [7] and it has even been vali-

    dated quantitatively by fully time-dependent simu-

    lations of dendritic growth in both two dimensions

    [810] and three dimensions [11] (Fig. 1), with the

    added insight that in three dimensions the non-axi-

    symmetric tip morphology inuences the selection

    for large enough anisotropy.

    Beyond the understanding of steady-state growth

    of the tip, the main new concept that has emerged

    over the last few years, is that complex pattern for-

    mation processes occurring on the much larger

    scale of an entire dendrite grain structure can be

    described by remarkably simple ``scaling laws''.These processes include growth transients [13, 14]

    that lead to steady-state growth and the highly non-

    linear competition of secondary branches behind

    the tip [1517]. In addition, a deeper understanding

    of the role of anisotropy has come from the discov-

    ery of new steady-state growth structures (doublons

    [8] and triplons [18]). Following the morphological

    instability of a small spherical grain, the primary

    branches of an equiaxed grain emerge along h100i

    directions in cubic crystals but do not immediately

    reach a steady state. These branches are much thin-

    Fig. 1. Three-dimensional equiaxed dendrites calculatedwith the phase-eld method: (a) thermal dendrite withh100i growth directions [9]; (b) solutal NiCu dendrite

    when the preferred growth directions are h110i instead ofh100i [12].

    44 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    3/28

    ner and thus grow much faster initially than in

    steady state, such that the instantaneous tip velocity

    V(t ) [tip radius r(t )] is a monotonically decreasing

    (increasing) function of time during a transient of

    duration HDaV2ss

    , where Vss is the nal steady-state

    growth velocity{. An analytical treatment of this

    transient has been possible in two dimensions (plate

    dendrites) in the limit of vanishingly small under-

    cooling where the problem is analogous to anisotro-

    pic HeleShaw ow and can be treated rigorously

    by the conformal mapping technique [13]. The main

    result is that the length and width of primary

    branches, and the total area of the plate, obey

    simple power laws given, respectively, by LtHt3a5,

    WtHt2a5 and AtHt for t ` DaV2ssX Moreover, the

    transient interface shape is described by a unique

    scaling shape. Another important feature of this

    transient is that although V(t ) and r(t ) can vary in

    time by one order of magnitude or more, the tip

    selection parameter, s 2Dd0art2Vt, remains

    constant in time and xed at its value determined

    by solvability theory. Physically, this follows fromthe fact that s is determined by the diusion eld

    in the tip region. Thus, at low undercooling, its

    value is established quasi-instantaneously on the

    time scale where the interface moves one tip radius

    since rt2aDrtaVtX Phase-eld simulations in

    two dimensions show a good quantitative agree-

    ment with these predictions at very low undercool-

    ing [14]. In three dimensions, no analytical theory is

    yet available to describe this transient but simu-

    lations reveal the existence of some approximate

    scaling behavior at short time with dierent power

    laws than in two dimensions [14].

    The results of two-dimensional growth transientshave immediate implications for understanding the

    large-scale structure of three-dimensional dendrites,

    since the mean cross-sectional shape of a three-

    dimensional steady-state dendrite (perpendicular to

    the growth axis) can be assumed to evolve with dis-

    tance z behind the tip as a two-dimensional branch-

    less plate dendrite evolves in time with t zaVss[15]. This assumption becomes exact far enough

    behind the tip since the heat (or solute) ux along z

    becomes negligibly small, and it yields the mean

    shape xHz3a5 for z large compared with the tip

    radius rss but small compared with the diusion

    length D/Vss

    , and xH

    z for distances larger than thislength as conrmed by three-dimensional phase-

    eld simulations [11]. Translated in terms of the

    projection area fraction f

    xz dz, the above

    result gives fHz1X6 for z ` DaVss, which is in

    reasonably good agreement with the scaling law

    fHz1X7 obtained by detailed measurements [17] of

    the morphology of pure SCN dendrites grown in a

    diusive regime in space [19]. Actually, on theoreti-

    cal grounds one would expect a time-varying expo-

    nent slightly larger than 1.6, which is a strict lower

    bound valid in the limit of vanishing undercooling.

    A scaling law has also been derived that describes

    how the length of ``active'' sidebranches that sur-

    vive the growth competition behind the tip and the

    spacing l between them depend on the distance z

    behind the tip [16]. The main prediction is thatboth l and increase linearly in z. The morphology

    measurements on SCN crystals yield a good quanti-

    tative agreement with this linear law but only if it is

    interpreted in parabolic coordinates [17], i.e. with

    the length of active sidebranches measured from a

    parabola tted to the tip and plotted vs distance

    along this parabola. This change of coordinate in-

    corporates the fact that sidebranches tend to grow

    perpendicularly to the isotherms [17] and thus eec-

    tively incorporates the eect of the heat ux along z

    that is neglected in the analysis of Ref. [16].

    Finally, the basic concept that sidebranches are

    driven by small perturbations of the tip region [20],

    which originated from the work of Zel'dovich et al.

    on ame fronts [21], has been further developed

    theoretically [16, 22] and validated by phase-eld

    simulations that consistently yield branchless den-

    drites (needle crystals) if numerical noise is kept

    small by using ne meshes [911, 2325].

    Furthermore, when noise is purposely added in a

    quantitatively controlled way, phase-eld simu-

    lations yield sidebranching characteristics (initial

    amplitude and spacing behind the tip) that are in

    good overall agreement with the predictions of the

    analytical theory of noise amplication in two

    dimensions [26]. These simulations presently need to

    be extended to three dimensions in order to test theprediction [16] that thermal noise is responsible for

    the experimentally observed sidebranching activity.

    The new steady-state growth structures that have

    been identied are the so-called ``doublons'' in two

    dimensions [8], rst observed in the form of a doub-

    let cellular structure in directional solidication [27],

    and the ``triplon'' in three dimensions [18]. Both

    structures have been shown [8, 18] to exist without

    crystalline anisotropy unlike conventional dendrites.

    The doublon has the form of a dendrite split in two

    parts about its central axis with a narrow liquid

    groove in between these two parts, and triplons in

    three dimensions are split in three parts. For a niteanisotropy, however, these structures only exist

    above a critical undercooling [28] (or supersatura-

    tion for the isothermal solidication of an alloy),

    such that standard dendrites growing along h100i

    directions are indeed the selected structures in

    weakly anisotropic materials at low undercoolings,

    in agreement with most experimental observations

    in organic and metallic systems. From a broad per-

    spective, the existence of doublons and triplons is

    of fundamental importance since it has provided a

    basis to classify the wide range of possible growth{ See list of symbols in the Appendix.

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 45

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    4/28

    morphologies that can form as a function of under-

    cooling and anisotropy [29].

    2.2. Directional solidication

    Signicant progress has been made over the last

    decade in understanding fundamental aspects of

    interface dynamics in directional solidication of

    alloys. The onset of morphological instability in

    directional solidication has been modeled by the

    classic MullinsSekerka instability [30], which pre-

    dicts the instability wavelength of a steady-state pla-

    nar interface. In a typical directional Bridgman set-

    up, however, the planar interface does not become

    unstable in steady state, but during the transient

    build-up of the solute boundary layer after solidi-

    cation is started. By analyzing the morphological

    stability of the planar interface during this transi-

    ent, and by taking into account that the instability

    takes time to grow from natural modulations until

    it becomes observable, it has been possible to

    obtain for the rst time an accurate prediction of

    the instability wavelength [31]. This prediction

    agrees well quantitatively with experiments on the

    onset wavelength and diers signicantly from the

    wavelength predicted assuming steady-state growth

    [3234] (Fig. 2).The critical role of crystalline anisotropy in inter-

    face dynamics has been demonstrated experimen-

    tally in directional growth [35]. This study exploited

    the ability to control the orientation of the crystal

    grown in a thin sample. With the h100i direction

    oriented (nearly) parallel to the axis of the thermal

    gradient, the typically observed stable cellular/den-

    dritic array structures are obtained [Fig. 3(a)]. In

    contrast, with the h111i direction oriented normal

    to the glass plates, there is no second-order aniso-

    tropy in the plane of the sample [i.e. d 2gyady2 0

    where g(y ) is the surface energy and y is the polar

    angle in this plane]. Thus growth in this plane is

    rendered ``eectively isotropic'' by this judicious

    choice of grain orientation. In this case, a ``sea-

    weed'' structure [35] [Fig. 3(b)] whose underlying

    building block is the theoretically expected doublon

    Fig. 2. Wavelength of morphological instability of plane front during initial transient [31].

    Fig. 3. Role of crystalline anisotropy on interface shape indirectional growth of thin transparent samples [35].

    46 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    5/28

    [8] is formed in agreement with numerical simu-

    lations also presented in Ref. [35]. This nding is

    also consistent with the numerical nding that stan-

    dard cellular structures are linearly unstable in the

    absence of crystalline anisotropy, except in a very

    narrow range of velocity near onset of instability

    [36].

    The formation of doublons and triplons has also

    been suggested to play an important role in the for-

    mation of ``feathery'' grains in aluminum alloys.

    This peculiar morphology, known since the 1940s,

    which appears as a succession of lamellae separated

    by straight and wavy boundaries, was associated

    with the formation of twins parallel to the lamellae

    but the appearance mechanisms remained unclear.

    Recent electron back-scattered diraction (EBSD)

    observations combined with detailed optical and

    scanning electron microscope (SEM) observations

    [2, 37] have clearly shown that feathery grains are

    made of h110i columnar dendrites [Fig. 1(b)], whose

    primary trunks are aligned along and split in their

    center by a (111) coherent twin plane. The impinge-ment of secondary h110i side arms gives rise to

    incoherent wavy twin boundaries. The switch from

    h100i to h110i growth morphologies was attributed

    to the small anisotropy of the solidliquid inter-

    facial energy of aluminum which can be changed by

    the addition of solute elements such as Zn, Mg or

    Ti and possible attachment kinetics eects. More

    details of this kind of growth may be found in Ref.

    [12].

    A new experimental technique has been devel-

    oped in which a brief spatially periodic u.v. laser

    pulse is applied to the solidliquid interface in a

    transparent organic system (succinonitrilecou-

    marin), to force a desired wavelength of the mor-

    phological instability [38]. These experiments have

    made it possible to investigate systematically the

    dynamical selection and stability of cellular struc-

    tures by varying the instability wavelength, and

    thus accessing cell spacings that are not normally

    accessible from a planar interface.

    A stability analysis of dendritic arrays [39] has

    been carried out in the limit where the primary spa-

    cing is larger than the diusion length. The basic

    instability found to limit the array stability at small

    spacing corresponds to a mode where one out of

    every two dendrites in the array is eliminated in

    agreement with experiments [40, 41]. The samemode is found numerically to limit the array stab-

    ility of cells [36] such that its existence appears

    rather universal. A range of interdendritic spacings

    is therefore stable, in agreement with experimental

    observations [42, 43], but experiments with the

    same ``history'' lead to a reproducible spacing [44,

    45]. As an elaboration of this work, a model of

    ``history-dependent'' selection of the primary spa-

    cing has been developed. This model is based on

    the picture that dendrites are eliminated continu-

    ously (subject to this instability) during the long

    transient that follows the initial morphological

    instability of the planar interface and leads to

    steady-state growth of the array with a nal selected

    spacing [31]. The predictions of this model agree

    reasonably well with one set of experiments [44].

    Moreover, at a more qualitative level, it has been

    demonstrated experimentally that the initial

    instability wavelength does indeed inuence the

    steady-state interdendritic spacing [46].

    Cellular/dendritic arrays have also been modeled

    numerically based on the traditional view that the

    structure with the lowest undercooling is selected

    within some stability band of spacings [47]. This

    model has had some success in explaining exper-

    imental data. It does not, however, control the

    strength of crystalline anisotropy, which is now

    understood to crucially inuence the cellular array

    stability both numerically [36] and experimentally

    [35]. Therefore, its validity remains to be further

    investigated. An analytical approach to the primary

    spacing problem by summation of the Ivantsov

    elds and application of the minimum undercoolingcriterion has also been developed recently [48].

    A detailed experimental study has brought new

    insights into the onset of sidebranching in direc-

    tional solidication [49]. In these experiments, the

    cell spacing was made uniform along the array and

    varied by exploiting the history dependence of

    wavelength selection in this system. This technique

    was used to characterize the onset of sidebranching

    systematically and shows that branched and non-

    branched cells in these experiments belong to the

    same branch of steady-state growth solutions.

    Furthermore, it has revealed that the thermal gradi-

    ent plays a destabilizing role (i.e. increasing G

    causes non-branched cells to branch). Theoretical

    models remain to be developed to explain this role

    as well as to characterize the onset of sidebranch-

    ing.

    2.3. High velocity microstructures

    Signicant experimental studies on microstructure

    formation under rapid solidication conditions have

    been carried out in the last decade using the laser

    scanning technique (a type of directional growth

    process) and levitational techniques (undercooled

    solidication).At high rates oscillatory behavior of the solid

    liquid interface (banding) has been analyzed by sev-

    eral authors [5052]. Band formation in the velocity

    regime of strong variation of the distribution coe-

    cient, k(V) [53], was shown to depend strongly on

    the coupling of non-steady-state heat and solute

    transport phenomena [52]. Experiments on the ab-

    solute stability of SCN have been undertaken [54]

    and it was shown that close to this limit, cells fol-

    low a l1X5V constant relationship [55].

    In highly undercooled levitated melts systematic

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 47

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    6/28

    measurements have been undertaken under others

    by the group of Herlach [56]. These authors have

    shown that over a substantial range of undercool-

    ings good agreement may be obtained between the

    measurements in a large number of metals and

    alloys and the analytical model using the transport

    solution of Ivantsov together with Marginal

    Stability arguments including solute trapping eects

    (a theory which we call the IMS model) [5759],

    with the stability parameter s 4p21X Further it

    has been recently shown that excellent agreement

    with no adjustable parameters can be obtained with

    this theory [60]. Note that this agreement should

    only be interpreted to mean that marginal stability

    arguments, although not fundamentally correct, are

    still useful to make quantitative predictions of den-

    drite growth rates in rapidly solidied binary alloys.

    A detailed comparison of solvability and marginal

    stability theory for rapid dendrite growth has

    recently been carried out [61].

    One of the important observations at high under-

    cooling is the formation of very ne-grained struc-

    tures over a range of large undercoolings. This ne-

    grained structure has been explained by dendrite

    fragmentation. At very high undercooling, as the

    dendrite trunk diameter becomes very ne, the ten-

    dency to undergo Rayleigh instability increases. A

    theory has been developed which shows that frag-

    mentation can occur when the characteristic time

    for dendrite break-up is shorter than the post-reca-

    lescence or plateau time in overall agreement withexperiments [62, 63].

    2.4. Coupled and simultaneous growth

    Even if most of the recent modeling was con-

    cerned with single-phase growth phenomena, there

    has also been some work on coupled or simul-

    taneous growth of two phases. A detailed numerical

    survey of the morphological instabilities of lamellar

    eutectics has been carried out in two dimensions by

    the boundary integral method for the transparent

    organic system CBr4C2Cl6 [64] (Fig. 4). In parallel,

    a detailed experimental survey of these instabilities

    has been carried out in the same system [65]. There

    Fig. 4. Calculated stability diagram for two-dimensionalcoupled eutectic growth in CBr4C2Cl6 [64] in excellentquantitative agreement with the experimentally measureddiagram in the same alloy [65]. Z: reduced concentrationwith the eutectic point at Z 0X3; L: lamellar spacing nor-malized by the spacing corresponding to minimum under-cooling. The basic axisymmetric state is stable within thecenter region. Other states include steady-state tilted pat-

    terns (T), 2lO (spatial period doubling oscillations), 1lO(spatial period preserving oscillations), where both 1lOand 2lO oscillations can be either axisymmetric or tilted.Blank regions of the diagram are those in which the

    dynamics is not yet fully understood.

    Fig. 5. Quenched liquidsolid interface of simultaneous two-phase growth in peritectic FeNi alloy [66].

    48 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    7/28

    is a remarkably good quantitative agreement

    between simulations and experiments concerning

    the regions of stability of both non-tilted and tilted

    steady states in the plane of composition and eutec-

    tic spacing, and the oscillatory instabilities that

    limit these regions. This understanding of eutectic

    stability, however, is restricted to two dimensions

    and presently needs to be extended to three dimen-

    sions. Eutectic cells and dendrites forming in multi-

    component alloys have also been studied

    theoretically and experimentally and are presented

    in Section 5.

    Simultaneous growth of two phases in the form

    of oriented bers and lamellae has been observed in

    some peritectic alloys. Figure 5 shows an example

    from a FeNi alloy. For this to happen, the compo-

    sition has to be between the two solid phases and

    the G/V ratio close to the limit of constitutional

    undercooling for the stable phase with the smaller

    distribution coecient [6668]. This interesting in

    situ growth phenomenon still waits for a theoretical

    interpretation, although recent phase-eld calcu-lations have shown the formation of such a struc-

    ture [69].

    3. PHASE AND MICROSTRUCTURE SELECTION

    A microstructure is dened by the morphology,

    size, distribution, crystal orientation, and corre-

    lation (texture), and number of phases. Phase and

    microstructure selection describes the variety of

    phases and microstructures that develop under

    given growth conditions and growth geometries.

    This section treats mainly transformations of phases

    and microstructures from one structure or mor-

    phology into another. It is not so much the for-

    mation of a single growth form itself which is of

    interest in this part of the paper (this has been

    treated in Section 2) but the mechanisms of change

    from one phase and morphology into another. A

    detailed theory of the mechanisms responsible for

    this selection is only at its very beginning. A well-

    known empirical approach that is consistent with

    many experimental results uses extremum criteria,

    such as the highest growth temperature in direc-

    tional growth. In undercooled solidication proces-

    sing the highest nucleation temperature and the

    highest growth rate control the nal appearance ofmicrostructures and phases.

    In many materials, additional phase transform-

    ations take place in the solid state which lead to the

    nal microstructure. In this review only solidica-

    tion will be discussed. In general all solidication

    processes start with nucleation and continue with

    growth. The nal phases may be controlled by

    nucleation, by growth, or by a combination of

    both. In all three cases that will be treated separ-

    ately in the following, much progress has been

    made in recent years.

    3.1. Nucleation control

    If suciently large undercoolings can be attained

    through hindrance of heterogeneous nucleation,

    then there may be access to a variety of metastable

    phases, such phases having lower melting points

    and liquidus temperatures. The importance of

    nucleation is seen when dealing with phase selec-

    tion. A typical process where nucleation plays a

    dominant role is solidication processing of under-

    cooled melts such as is observed in droplets [56].

    There may be a spread in nucleation temperatures

    even under nominally identical conditions, and con-

    sequently the results are best displayed on ``micro-structure-predominance maps''. Such maps have

    been constructed for binary alloys, with alloy com-

    position and droplet diameter as coordinates [70

    72]. It is found that: (i) microstructure correlates

    very strongly with droplet diameter (which deter-

    mines the availability of nucleant sites and the cool-

    ing rate); (ii) the eects of processing conditions

    (e.g. gas purity in atomization) can be taken into

    account; and (iii) correlation with undercooling can

    be found through comparison with controlled

    undercooling experiments and growth modeling [71,

    73] (Fig. 6). It is clear, though, that we are very far

    from being able to predict nucleation undercoolings,the diversity of potential heterogeneous nucleants

    being a key impediment to quantitative modeling of

    most real situations.

    Under given conditions it is usual for one phase

    to dominate, but the primary phases can also be

    mixed. A well-analyzed example is the duplex parti-

    tionless solidication of b.c.c. and f.c.c. phases in

    the NiV system [74]. There are many examples in

    which the phase competition is between b.c.c. and

    f.c.c. phases, and this has been most closely exam-

    ined for FeNi. In undercooled levitated droplets,

    Fig. 6. Phase predominance map (drop diameter vs com-position) for undercooled growth in FeNi alloys [73].

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 49

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    8/28

    the prevalence of one phase or the other can be

    fairly well predicted by which phase has the lower

    work of nucleation (Fig. 6) (e.g. Ref. [75]). The

    b.c.c. phase is easier to nucleate than would be

    expected from its relative thermodynamic stability;

    it has a lower solidliquid interfacial energy. In the

    FeNiCr system, it has been shown that the f.c.c.

    or b.c.c. phase can be selected through the use of

    an appropriate substrate put in contact with the

    droplet to trigger solidication [76].

    The existence of the true primary phase is some-

    times revealed by a double recalescence phenom-

    enon [75]. In this way, for example, the transient

    existence of a hitherto unknown metastable f.c.c.

    phase of rhenium has been inferred [77].

    Observations of this kind, found also in some alloy

    systems [78], may be crucial in analyzing nucleation

    kinetics.

    Particular interest has centered on analyzing het-

    erogeneous nucleation kinetics. Basic treatments of

    heterogeneous nucleation (taking into account vari-

    ations of potency and population) have had successin predicting primary phase selection [79]. Recent

    advances have been made by studying liquid dro-

    plets entrained in solid matrices; when large under-

    coolings b 50 K are required for heterogeneous

    nucleation, the classical spherical cap model seems

    to work well, but for more potent nuclei it breaks

    down [80]. In that case, some success has been

    achieved with a model that considers the thermo-

    dynamics, if not yet the kinetics of adsorption [81].

    Further heterogeneous nucleation studies have been

    undertaken on liquid droplets in an emulsifying or-

    ganic liquid; taking a more classical approach, roles

    have been identied for dierent types of stationary

    and moving surface steps [82]. Yet another advance

    has been transmission electron microscopy of het-

    erogeneous nuclei formed in a glassy matrix [83];

    this study, of relevance for commercial grain rene-

    ment of Al, shows the importance of crystallogra-

    phy and chemistry in nucleation, even though

    quantication of these roles remains elusive [84].

    So far, it has been natural to concentrate on the

    primary stage of solidication, yet there are cases

    where the interest is in the formation of secondary

    phases in the nal stages of solidication. An

    example of considerable industrial interest is the

    DC casting of very dilute Al alloys. In cases of

    practical importance there is a range of intermetal-lics which can nucleate, the phase selection being

    sensitive to many parameters including solidication

    velocity. Directional solidication can reveal

    changes in intermetallic selection and be a basis for

    understanding fundamental mechanisms [85, 86].

    Attempts to analyze phase selection have focused

    on comparative eutectic growth kinetics [87], but

    solid-state changes and nucleation eects have also

    been considered. It appears that a quantitative

    analysis of the phase selection may depend on the

    geometry of the liquid in which the rival intermetal-

    lics nucleate and grow; such an analysis has yet to

    be attempted. Considerations so far suggest that the

    wide range of conditions over which mixtures of

    phases are obtained is indicative of a growth com-

    petition [88].

    3.2. Growth control

    Growth-controlled phase and microstructure

    selection has been successfully treated by comparing

    the steady-state interface response of competing

    phases. Calculating the interface response, i.e. the

    growth behavior of plane front, cells and dendrites,

    for all possible phases one can determine the

    growth form which develops the highest interface

    temperature for a given growth velocity and tem-

    perature gradient [89] or the highest growth velocity

    for a given undercooling{. Growth of eutectic struc-tures can also be included in this treatment. The

    extremum criterion is a strong indication of the

    structure to be formed. It assumes that: (i) the

    microstructure selection is not nucleation controlled

    (i.e. nucleation undercooling is suciently small);

    (ii) interaction between competing growth forms is

    negligible; and (iii) steady-state theory can be

    applied. Despite its simplicity this approach is of

    great help in determining microstructure maps for a

    more rational alloy development. Several cases of

    recent modeling of microstructure selection will be

    presented in the following.

    3.2.1. Stable to metastable phase transition. Using

    the above-mentioned maximum temperature cri-

    terion, the stable to metastable transition for direc-

    tional dendrite growth has been analyzed for

    peritectic systems [9092]. This allows us to ration-

    alize why at high velocities a transition from a

    stable to a metastable peritectic phase is often

    obtained. For example in FeNi or FeNiCr

    steels, high weld speeds lead to the formation of

    austenite dendrites [9395], even if at low velocities

    ferrite is the primary phase, with important conse-

    quences for the integrity (solidication cracking) of

    the weld. No nucleation is needed for this transition

    to occur as the metastable phase (austenite) growsinbetween the stable phase due to microsegregation

    and its growth is accelerated with velocity until the

    metastable phase becomes the leading one. For the

    reverse case (metastable to stable phase) this is not

    true and nucleation is a necessary requirement for

    the transition to happen (see under mixed control).

    Similar stable to metastable phase transitions have

    been analyzed in detail for FeC alloys [96].

    Under conditions of rapid solidication, solute

    and disorder trapping become signicant in the kin-

    etics [97]. By including such eects it has been poss-{ In this approach cells and dendrites are treated as one

    entity with one growth equation.

    50 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    9/28

    ible quantitatively to model the growth kinetics in

    undercooled NiAl alloys [98, 99]. Growth kinetics

    alone can be used to follow the competition

    between the ordered and disordered version of one

    phase. However, prediction of which basic structure

    (b.c.c. vs f.c.c. again in this case [98]) will beobserved requires a modeling of the nucleation

    which does not yet exist.

    3.2.2. Dendrite to eutectic transition (CZ). The

    limits of the so-called ``Coupled Zone'' represent

    the transition between fully eutectic structure and

    primary dendrites or cells with interdendritic eutec-

    tic. This transition which is also hysteretic in nature

    is inherently dicult to model. Karma has made

    progress in this matter and his results on the stab-

    ility of eutectic growth are discussed in Section 2.

    Using steady-state growth theory and the extre-

    mum criterion for the interface temperature, this

    transition can be calculated for directional growth

    (Bridgman, laser treatment, etc.) and is found to be

    in good agreement with experimental evidence

    [100]. In this way, following the early work of

    Boettinger et al . [101], a series of solidication

    microstructure selection maps has been obtained in

    recent years which allows a more rational approach

    to the solidication processing of technically im-

    portant alloys: AlCu [102, 103], AlFe [104], AlSi

    [105], AlCuSi [106], NiAl [99], and even cer-

    amics such as Al2O3ZrO2 [107]. These maps have

    been used as a tool for analyzing and predicting the

    microstructures of laser surface treated materials. A

    similar approach, but for undercooled melts with a

    corresponding maximum velocity criterion, has also

    been developed [108].

    Another way of this type of microstructure mod-

    eling is the ``inverse modeling'' which starts with in-

    formation about the microstructure and optimizes

    the input parameters such as the phase diagram[103, 109111] (Fig. 7). This new approach to deter-

    mine stable and especially metastable phase equili-

    bria is useful in cases where conventional

    techniques do not work.

    3.3. Mixed control

    Mixed control is always found when both nuclea-

    tion and growth play a controlling role in the

    microstructure selection, such as in the columnar to

    equiaxed transition of dendritic or eutectic struc-

    tures or in low velocity microstructures in the two-

    phase region of peritectic systems.

    3.3.1. Columnar to equiaxed transition (CET).

    Hunt's classic approach to model the CET [112] has

    been applied to welding [113] and has been

    extended by using more recent dendrite models

    [114, 115]. In this way critical growth conditions for

    the single crystalline welding of single crystal gas

    turbine blades could be established and a poten-

    tially interesting process for lifetime extension of

    these expensive components developed (Fig. 8)

    [116].

    The transition from the outer equiaxed zone to

    the columnar region (ECT), often observed in cast-ings can be understood in terms of the same CET

    criterion. Such considerations were made for the

    shape of grains continuously nucleating and grow-

    ing in a thermal gradient [117]. When the ratio G/V

    increases up to a critical value, the shape factor of

    the grains becomes innite, meaning that the

    equiaxed grains become columnar. In order to

    explain the ECT, it is necessary to consider the heat

    transfer coecient between the casting and the

    mould (which changes strongly when solidication

    starts) and the superheat of the melt. Such model-

    Fig. 7. Calculated microstructure selection map (VCodiagram) for NiAl alloys (b), and optimized phase dia-gram (a) [109]. (The dierence of the eutectic temperatureofbg ' and bg eutectic is less than the width of the line.)

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 51

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    10/28

    Fig.

    8.

    Epitaxiallasermeta

    lformingofasinglecrystalsuperalloy(CM

    SX4)showingthesinglecrystallinenatureofthelasercladdepositedontothecasts

    inglecrystalsub-

    strate[116].

    52 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    11/28

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    12/28

    two phases, and by growth competition between the

    nucleated grains and the pre-existing phase under

    non-steady-state conditions. In this case the simple

    extremum growth criterion does not lead to the

    right answer and nucleation in the constitutionally

    undercooled zone ahead of the growth front has to

    be taken into account in order to determine the

    microstructure selection [119, 120] (Fig. 9).

    A clear understanding of complex microstructure

    formation has come from directional solidication

    studies of binary alloys with compositions in the

    two-phase region of the peritectic phase diagram at

    large G/V ratio to suppress the morphological

    instability of both the parent (a ) and the peritectic

    (b ) phases, i.e. each phase alone would grow as a

    planar front. Even in this simplied case, a rich var-

    iety of microstructures has been identied that

    depend sensitively upon the relative importance of

    nucleation, diusion and convection [121125] as

    shown in Fig. 10. These microstructures can be

    broadly classied into the following groups based

    on geometrical patterns and the underlying trans-port mechanisms: (a) discrete bands of the two

    phases; (b) partial bands or particulates (or islands)

    of one phase in the matrix of the other phase; (c)

    single primary to peritectic phase transition; (d)

    simultaneous growth of the two phases with a pla-

    nar solidliquid interface; (e) dispersed phases due

    to nucleation ahead of the interface; and (f) oscillat-

    ing continuous tree-like structures of the primary

    phase that are surrounded by the peritectic phase

    [122]. Theoretical models and experimental studies

    in very thin samples have shown that structures

    (a)(e), can form under diusive regimes, whereas

    microstructure (f) is a novel microstructure whose

    formation requires the presence of oscillatory con-

    vection in the melt.

    In order to understand the formation of some of

    the complex microstructures in the two-phase

    region of peritectic systems, an analytical model of

    banding in peritectic systems was rst proposed for

    diusive growth in which the change in phases

    occurred when the appropriate nucleation under-

    coolings were reached [126]. According to this

    model, a banding cycle of alternate nucleation and

    growth of primary, a, and peritectic, b, phases may

    continue, leading to an oscillatory behavior of the

    interface and to alternate bands of a and b. The

    major predictions of this diusive banding model

    are: (i) the banding cycle will operate below and

    above the peritectic temperature; and (ii) the band-

    ing window exists only for a narrow range of initial

    alloy composition in the hypoperitectic range.

    In the above one-dimensional model of discrete

    band formation, it was assumed that the nuclei ofthe new phase spread rapidly in the lateral direc-

    tion, so that no appreciable lateral gradients exist.

    However, this is generally not valid and one must

    consider the relative rates of spreading of the new

    phase and the continuing growth of the parent

    phase. The microstructure for this complicated case

    was investigated experimentally as well as by nu-

    merical simulation of a two-dimensional transient

    phase-eld model for a generic peritectic phase dia-

    gram [69]. Several new morphologies were observed

    and predicted depending on the nucleation rate.

    Fig. 10. Fluid ow controlled microstructures in peritectic alloys: (a) discrete bands of the two phases;(b) partial bands or islands of one phase in the matrix of the other phase; (c) single primary to peritec-tic phase transition; (d) simultaneous growth of the two phases with a planar solidliquid interface; (e)dispersed phases due to nucleation ahead of the interface; (f) oscillating continuous tree-like structures

    of the primary phase that are surrounded by the peritectic phase [122].

    54 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    13/28

    The results of the phase-eld model indicate that

    when only a single nucleus is allowed to form on

    the wall of the sample, discrete band formation in a

    diusive regime is only possible for a nite range of

    system sizes, Lmin ` L ` Lmax, where L is the size of

    the sample or the distance between the nuclei.

    Moreover, this range depends on both the compo-

    sition inside the hypoperitectic region, and the

    nucleation undercoolings of the two phases. For

    internuclei distance L ` Lmin, discrete particles of

    the b phase form inside the a matrix, and for

    L b Lmax, discrete particles of a phase form in the b

    matrix form. In addition to the microstructure of

    discrete particles of one of the two phases

    embedded inside the continuous matrix of the other

    phase, more complex microstructures, including two

    simultaneously growing phases form [Fig. 10(d)].

    Simultaneous two-phase growth has been observed

    in several peritectic systems, including SnCd [122],

    AlNi [67], and FeNi [66] (see also Section 2.4).

    The basic model of nucleation and growth con-

    trolled structures shows that dierent microstruc-tures can form only within a narrow band of

    composition in the hypoperitectic region. However,

    several experimental observations of banded struc-

    tures have been made for compositions outside this

    banding composition window, and banding struc-

    tures were reported even for hyperperitectic compo-

    sitions. These observations clearly indicate that the

    observed structures are not controlled by diusion

    but by convection eects, as shall be discussed in

    the following section.

    4. CONVECTION EFFECTS

    Convection eects are of utmost importance in

    the development of solidication microstructures.

    Despite this fact, most microstructure models are

    based on purely diusive transport mechanisms.

    Only recently, modeling of growth in the presence

    of convection has been successfully undertaken. The

    rst step in such an undertaking is modeling of con-

    vection and its instabilities before coupling of con-

    vection and microstructure formation is done.

    4.1. Convection instabilities

    For the simple problem of an innite layer withvertical temperature and solutal gradients, it is well

    known that convective instabilities can occur even if

    the net density of the liquid decreases with height

    [127]. Similar behavior also occurs in a porous med-

    ium [128]. However, usually the temperature and

    solute concentration in the mushy zone of a binary

    alloy are coupled by the phase diagram and this

    prohibits double-diusive behavior, i.e. a density

    inversion is necessary for the onset of convective

    instability. Worster [129] has reviewed recent work

    on convection in mushy zones. The critical

    Rayleigh numbers for the onset of convection in a

    binary alloy have been calculated for three dierent

    models of the mushy zone during directional solidi-

    cation [130]. In general, there are two modes of

    instability: a mode in the mushy layer and a bound-

    ary-layer mode in the melt; the wavelength of the

    mushy-layer mode is small compared with the

    wavelength of the boundary-layer mode. In addition

    to these non-oscillatory modes, there are modes

    that are oscillatory in time.

    The radial segregation due to solutal convection

    during the directional solidication of leadthallium

    alloys with a planar crystalmelt interface has been

    calculated using pseudo-spectral methods [131].

    Solutal Rayleigh numbers for the calculations ran-

    ged from very near the onset of convective instabil-

    ity to a factor of ten above the instability onset. In

    general, the ows and segregation are asymmetric,

    although for special conditions axisymmetric ows

    can occur.

    A sudden change in ow conditions is correlated

    with the interface concentration during directionalsolidication of a tinbismuth alloy [132]. The

    interface concentration was monitored by Seebeck

    measurements using the MEPHISTO furnace

    during the USMP-3 space ight. Numerical calcu-

    lations of the uid ow and solute redistribution

    due to sudden gravitational accelerations caused by

    thruster activation were in good agreement with the

    observed Seebeck signals.

    4.2. Field eects

    It is well known that a uniform magnetic eld

    can damp convective motions in an electrically con-

    ducting uid. However, when a gradient in the

    Seebeck voltage exists in the presence of a magnetic

    eld and temperature gradients, there can be a

    resulting ow [133, 134]. This can occur at a crys-

    talmelt interface when there is a temperature gra-

    dient along the interface; for example, in a binary

    alloy with a non-uniform concentration. Since this

    thermoelectric magnetohydrodynamic ow occurs

    in the vicinity of the interface, it can play a signi-

    cant role in solute redistribution. It can also be

    used to counteract buoyancy-driven ow in the

    mushy zone during horizontal directional solidica-

    tion [135]. Freckle formation in coppersilver andaluminumcopper alloys have been examined under

    dierent magnetic elds. The observed larger den-

    drite spacings agree with the observation in space

    experiments where the ow is signicantly reduced

    [136].

    There have been a number of studies of crystal

    growth in very high gravitational elds using a cen-

    trifuge [137139]. For germaniumgallium alloys,

    solute segregation exhibits a minimum as a function

    of rotation rate. This behavior can be understood

    by considering non-axial temperature gradients; at

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 55

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    14/28

    low rotation rates, the uid velocity is decreased by

    the action of Coriolis forces, while at large rotation

    rates the buoyancy force due to the centrifugal

    acceleration, increasing as the square of the rotation

    rate, becomes dominant [139]. Thus, the ow and

    segregation are reduced at intermediate rotation

    rates.

    4.3. Eect of ow on interface morphology

    The eect of simple ows on the shape of par-

    ticles growing from a supersaturated solution has

    been calculated [140]. The concentration and uid

    ow elds are solved numerically by a mapping

    technique in the Stokes ow approximation. Simple

    base ows such as a uniform streaming ow or a

    biaxial straining ow lead to non-spherical shapes.

    The particle shape, as function of the ow magni-

    tude and the anisotropy of the crystaluid surface

    tension, has been studied.

    The inuence of convection on morphologicalinstability and interface structure during directional

    solidication was examined theoretically [141].

    There have also been observations on massive

    transparent specimens [142] which have revealed

    that convection results in a gradient of microstruc-

    ture along the interface, from a smooth interface to

    dendrites. Fluid ow eects at the very scale of the

    microstructure have been seen during solidication

    of faceting transparent systems (e.g. salol-based

    alloys) where saw-tooth patterns of millimeter size

    form [143]. Surface tension-driven convection due

    to the presence of uiduid interfaces and its inu-

    ence on the morphology of the growth front

    deserves thorough investigation, e.g. coupled

    growth of bubbles of dissolved gas and monotectic

    alloys in which a second liquid phase forms either

    as rods in a solid matrix or droplets in the melt

    [144].

    Anisotropic interface kinetics stabilizes an inter-

    face with respect to the onset of morphological

    instability [145, 146]. Such anisotropic kinetics

    arises naturally when growth is by step motion and

    the crystalmelt interface is near a singular crystal-

    lographic orientation. When a planar interface is

    perturbed with a sinusoidal perturbation, anisotro-

    pic kinetics causes a lateral translation of the sinus-

    oid (traveling wave). In turn, this lateral motioncan strongly interact with shear ows along the

    interface. Flow in the direction of step motion is

    destabilizing while ow opposite to the step motion

    is stabilizing.

    Experiments on the dendritic growth of succino-

    nitrile and pivalic acid from supercooled melts on

    earth and in microgravity show small discrepancies

    from the classic Ivantsov relation between Peclet

    number and dimensionless supercooling (Fig. 11)

    [147]. Under terrestrial processing conditions, con-

    vection in the melt has a major impact on metallic

    solidication, especially at small crystal growth vel-

    ocities [56]. Previous studies of dendrite growth in

    undercooled Ni melts on earth show systematic de-

    viations of experimental data and dendrite growth

    theory at small undercoolings. The discrepancy is

    partly reduced if convection is taken into consider-

    ation. Measurements of the dendrite growth vel-

    ocity as a function of undercooling on pure Ni and

    dilute Ni0.6 at.% C alloys under microgravity con-

    ditions provide a test of dendrite growth models

    [148]. The experiments were performed using the

    electromagnetic levitation facility TEMPUS.

    Excellent growth velocity data were obtained during

    the mission in an undercooling range between 50

    and 310 K. However, no dierences between micro-

    g and 1 g data were detected in this temperature

    range since ow due to electromagnetic forces may

    be signicant.

    The selection of twinned dendrites in the presence

    of uid ow may be explained by a higher growth

    temperature with respect to normal dendrites, in

    particular as a result of doublon formation (see

    Section 2.2). The eect of convection on the alter-

    nating sequence of straight/coherent and wavy/inco-

    herent twin is shown in Fig. 12 [37]. The alloy hasbeen produced by direct chill (DC) continuous cast-

    ing and exhibits, in some regions, ``feathery grains''.

    In Figure 12(a), three feathery grains labeled 13

    are clearly visible: each one is made of parallel

    lamellae showing an alternating sequence of colors

    (green/red, light blue/purple, and yellow/violet

    for grains 1, 2 and 3, respectively) separated by an

    alternating sequence of straightwavy boundaries.

    This corresponds to twinneduntwinned regions

    separated by coherentincoherent twin boundaries

    across rows of primary h110i dendrite trunks.

    Fig. 11. Tip Peclet number vs supercooling for free den-dritic growth of organics under terrestrial and under

    microgravity conditions [147].

    56 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    15/28

    Transverse melt ow was invoked to explain the

    systematic alternating sequence of lamellae and

    boundaries through branching mechanisms [see Fig.

    12(b), in which the twinned dendrites are seen along

    their trunk axis].

    Numerical simulation of microscopic ow in the

    melt during solidication was introduced through aphase-eld model [149151]. The eects of ow on

    free dendritic growth (tip velocities, radii, and tip

    selection as a function of the orientation of the ow

    with respect to the crystal) were investigated.

    Convection during coarsening of an isothermal

    binary liquidsolid mixture has been studied, i.e.

    the eects of convection on coarsening and of coar-

    sening on the permeability were examined. The

    eect of convection on equiaxed dendrite growth

    and associated macrosegregation has also been stu-

    died. New results have been obtained on the size

    evolution and settling velocity of NH4Cl equiaxed

    crystals growing from supercooled NH4ClH2O sol-

    ution [152]. The results have been analyzed with the

    theory for an isolated dendrite growing in an axi-

    symmetric melt ow [153]. In the range of the ex-

    perimental settling velocities (711 mm/s), the best

    t for the stability constant was found to be 3.12

    times greater than the value measured for the

    purely diusive case [154].

    Extension of a Cellular Automaton (CA) tech-

    nique coupled with a nite element (FE) method

    (CAFE model [155]), improves the modeling of den-

    dritic grain structures in the presence of convection.

    The movement of equiaxed crystals in the liquid to

    form a sedimentation cone, as well as the modi-

    cation of the columnar-to-equiaxed transition in the

    presence of convection, are well described qualitat-

    ively by the CAFE model.

    There has been interesting experimental evidence

    on the mushy zone interactions with melt ows in

    transparent organics. Quantitative measurements of

    the ow eld during solidication could be made[156].

    When the JacksonHunt model of eutectic

    growth [157] is applied to the growth of monotectic

    composites, the predicted value ofl 2V is more than

    an order of magnitude smaller than the experimen-

    tal value for aluminumindium monotectic alloys

    which grow with rods of indium-rich liquid in an

    aluminum-rich solid matrix; here, V is the growth

    velocity and l is the interrod spacing. Allowing for

    diusion in the rod phase does not improve the

    agreement between experiment and theory [158].

    While the discrepancy could be due to inaccurate

    values of the thermophysical properties, another

    transport mechanism such as convection could

    account for the discrepancy. Such uid ow could

    arise from surface tension variations along the

    uiduid interface; a pressure-driven ow could

    also occur at the uiduid interface due to the

    requirement of satisfying both the GibbsThomson

    and YoungLaplace equation at this interface [159].

    Convection eects have also been found to give

    rise to new microstructures that are not observed in

    the diusive growth regime. For example, Fig. 10(f)

    shows a novel microstructure whose formation

    requires the presence of oscillatory convection in

    the melt of peritectic systems. A detailed study of

    the three-dimensional shape of the microstructurerevealed that the bands were not discrete, but both

    the a and the b phases were continuous [160]. It

    was shown that the microstructure, which appears

    like discrete bands on a section close to the surface

    of the sample, is in fact a more complex structure

    made up of two continuous interconnected phases

    in three dimensions. In particular, the microstruc-

    ture consists of a large tree-like domain of primary

    a phase that is embedded inside the peritectic b

    phase. The formation of this structure is governed

    by oscillating convection present in a large diameter

    Fig. 12. Twinned dendrites in AlCu alloy. (a) EBSDreconstruction of the microstructure in a transverse sec-tion to the thermal gradient containing three grains. (b)Schematic view of the eect of convection on twinned

    dendrite formation [37].

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 57

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    16/28

    (6.0 mm) sample [124]. Besides the tree-like struc-

    ture, several other new oscillating microstructures

    were observed experimentally, and predicted nu-

    merically, depending upon the intensity and modes

    of convection [160].

    5. MULTICOMPONENT SYSTEMS

    The application of solidication modeling to

    practical technology is closely linked to our ability

    to model microstructural development in multicom-

    ponent alloys (three or more components). Over the

    past ten years signicant progress has been made in

    this area.

    5.1. Thermodynamics

    Solidication models, which use local interfacial

    equilibrium, have been successfully coupled to

    phase diagram information obtained via the

    Calphad method [161]. Examples include analyses

    of Scheil solidication path, dendrite tip kinetics,solid (back) diusion and macrosegregation.

    Commercial databases are available for Al, Fe, Ni

    and Ti base alloys (ThermoTech, Ltd{) and others

    are distributed with the various thermodynamic

    computational codes available: ThermoCalc (KTH,

    Stockholm), MTData (NPL), Chemsage (RWTH,

    Aachen). All of these computational codes can be

    interfaced with solidication models. As an example

    [162] a set of subroutines, LEVER, SLOPE and

    HEAT, have been built on top of a modied ver-

    sion of the Lukas code. LEVER gives the phase

    fractions and phase compositions at equilibrium for

    a specied temperature T and overall composition.

    SLOPE gives the liquidus temperature, the solid

    phase concentrations, and the liquidus slopes for a

    specied liquid composition and solid phase. HEAT

    gives the enthalpy per unit mass for a specied

    phase for a given temperature and phase compo-

    sition.

    This thermodynamic approach naturally enables

    a Scheil analysis of the solidication path; i.e. the

    evolution of the liquid and solid concentrations and

    the phase fractions during cooling under the

    assumption of complete liquid diusion and no

    solid diusion. This approach easily treats the

    appearance of new phases at eutectic reactions, or

    the disappearance and appearance of phases at peri-tectic reactions. A Scheil analysis provides the basis

    for a good estimate of very practical information

    for castings: (a) how the heat of fusion evolves

    during cooling (for coupling to macroscopic heat

    ow analysis); and (b) how the density of the

    mushy zone changes (for coupling to uid ow

    modeling for macrosegregation, porosity and hot

    tearing analysis).

    Under rapid solidication conditions, when local

    interface equilibrium is invalid, thermodynamic cal-

    culations for multicomponent alloys can be used to

    compute the thermodynamic driving ``force'' and

    the energy dissipated due to solute drag (if a model

    of diusion through the interface is prescribed).

    The thermodynamic driving ``force'' is required for

    analysis of the interface response functions for

    rapid solidication. In this area, the AzizKaplan

    model of solute trapping [163] has been extended to

    multicomponent systems for the case when the dif-

    fusive speed for all of the solutes is identical [164].

    An open question remains regarding the impact of

    dierent diusive speeds for dierent solutes in

    multicomponent solute trapping models of rapid

    solidication. Experimental work in this area would

    be useful.

    5.2. O-diagonal diusion terms

    O-diagonal diusion terms are usually neglected

    for multicomponent liquids, yet there is little justi-

    cation. Moreover the diagonal terms are usually

    assumed to be identical. When diusion uxes are

    related to chemical potential gradients through

    appropriate mobilities, the absence of o-diagonal

    mobility terms does not imply the absence of o-di-

    agonal diusion terms. O-diagonal terms tend to

    be strongly concentration dependent. One set of ex-

    periments [165] measured the o-diagonal diusion

    terms in liquid ternary Al alloys. Diusion couples

    with a step change in one component but a con-

    stant value of the second were analyzed. In thealloys tested, there was no detectable change in con-

    centration of the second component; i.e. negligible

    o-diagonal terms.

    It has been shown [166] that analytical models

    for plane front and dendritic growth developed for

    binary alloys can be extended to multicomponent

    alloys by taking into account the o-diagonal terms

    of the diusion matrix. The diusion elds for the

    n solutes are given by linear combinations of the n

    binary solutions using the eigenvalues and eigenvec-

    tors of the diusion matrix instead of the diusion

    coecients (Fig. 13). For the time being however,

    use of these solutions is limited by the lack of

    measured diusion coecients and methods todetermine the o-diagonal terms. A theoretical

    approach to this problem is needed.

    5.3. Fundamental morphological stability issues

    A complete linear stability analysis of planar

    growth under the assumption of local equilibrium

    for a ternary alloy with no o-diagonal diusion

    terms has been performed [167]. When the pertur-

    bation wavelength is not assumed to be small com-

    { Trade names are used for completeness only and do

    not constitute an endorsement of NIST or any other or-

    ganization.

    58 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    17/28

    pared with the liquid diusion lengths for all of the

    solutes Di/V, the perturbation growth rate, e, not

    only depends on the partition coecients, ki, and

    liquidus slopes, mi, but also on the derivatives of

    liquid concentrations with respect to solid concen-

    trations evaluated on the liquidus surface, quantities

    that can be computed from the thermodynamic

    approach. When the perturbation wavelength is

    small, these factors disappear. In this case, the ex-

    pression for e contains a denominator, which can

    vanish due to the fact that miki 1 can be nega-

    tive for one or more of the solutes in multicompo-

    nent systems. This has the potential to lead to

    oscillatory instabilities. Whether this can occur in

    an experimental system is not known.

    Other stability issues, such as the cell to dendrite

    transition, have not been adequately resolved, even

    for binary alloys. One situation peculiar to ternary

    systems is the formation of eutectic cells by the pre-

    sence of a dilute ternary solute. The full stability

    spectrum of a steady-state lamellar interface in the

    presence of a ternary impurity has been calculated

    and an analytical form of this spectrum has been

    derived in the limit where the wavelength of the

    perturbation is large compared with the lamellar

    spacing [168]. Also preliminary phase-eld calcu-

    lations of the growth of eutectic cells to treat large

    amplitude perturbations have been performed as

    shown in Fig. 14 [169]. (Such calculations are de-

    nitely at the limit of what is actually possible; this

    simulation took approximately 60 CPU hours on 32

    processors of a CRAY T3E.)

    5.4. Microstructure prediction

    5.4.1. Dendritic growth and solid diusion. Even

    though many issues remain regarding the funda-

    mental role of anisotropy on dendrite tip radius

    selection (even for pure materials), models used for

    practical materials typically use the Ivantsov/

    Marginal Stability (IMS) approach [5759]. This

    model for binary alloys has been extended to multi-

    component alloys [170, 171]. The equiaxed growth

    model [172] has been generalized to multicompo-

    nent alloys [173] as well as the standard model of

    secondary spacing [173, 174]. The FloodHunt

    method coupling dendrite tip models to the Scheil

    analysis [175] has been modied to treat multicom-

    ponent alloys. The modication also conserves

    solute, but only for the case of diagonal and equal

    liquid diusion coecients [176].

    Modications have been made to the Scheil

    approach to deal with solid diusion for multicom-

    ponent alloys. An approximate treatment of solid

    diusion [177] has been extended to multicompo-nent alloys and coupled to phase diagram calcu-

    lations [162]. This method is convenient for node by

    node coupling to macroscopic heat ow calculations

    because it reduces computation time compared to a

    full solution of the diusion equations. Solution of

    the full diusion equations has been performed

    using DICTRA [178], a diusion analysis code built

    on top of ThermoCalc. An approach to model solid

    diusion during monovariant eutectic solidication

    in addition to primary solidication has also been

    performed [179].

    Fig. 13. Plane front concentration elds for a three-component system with the liquid solute diusioncoecients; D11 6 10

    9, D22 2 109, D12 D21 3 10

    9 [166].

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 59

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    18/28

    5.4.2. Eutectic coupled zone and associated micro-

    structure maps. By computing the competition

    between dendritic and eutectic growth for a speci-

    ed alloy composition, microstructure maps that

    dene the range of solidication speed and tempera-

    ture gradient required to form a specied growth

    form (hence microstructure) have been developedfor ternary alloys [106]. Similarly the code PHASE

    [180], which has been extended to multicomponent

    multiphase alloys [181], computes the dominant

    growth microstructure and the resultant microsegre-

    gation during cooling or during steady-state direc-

    tional growth through a numerical analysis of

    competitive growth and solid diusion. Phase dia-

    gram information is obtained using graphs as input.

    The analysis of the microstructure selection has

    been extended to a ten-component superalloy

    (CMSX4) [116, 182]. In this way processing win-

    dows for laser metal forming with application to

    repair of single crystal superalloy turbine com-

    ponents could be calculated. The processing con-

    ditions required to avoid stray grain formation were

    evaluated using Hunt's columnar-equiaxed tran-

    sition model [112] but with numerical evaluation of

    dendrite tip kinetics using IMS with solute trapping

    modications [115] and with phase diagram infor-

    mation delivered via coupling to ThermoCalc.

    Computation of macrosegregation using the sub-

    routines described in Section 5.1 has also been per-

    formed [183]. Other practical solidication

    problems in multicomponent alloys are being ana-

    lyzed. For example, the relative importance of

    nucleation vs growth competition in understanding

    the identity of the primary phase (f.c.c. or b.c.c.) in

    FeCrNi alloys near the monovariant eutectic line

    of the ternary liquidus surface of the phase diagram

    has been analyzed [184]. The mechanism for the

    formation of austenite dendrites in the so-called

    eutectic region of the microstructure of FeCSi

    spheroidal cast irons [185] is another example.

    6. SIMULATION METHODS

    With the advent of very powerful computers,

    advanced numerical methods and better under-

    standing of the physical phenomena involved in

    solidication, it is not surprising that computer

    simulations are becoming increasingly used for the

    modeling of microstructure formation and associ-

    ated characteristics or defects (e.g. microsegregation

    pattern, porosity formation, etc.). Over recent

    years, three major contributions have emerged: (1)modeling of microstructure formation using phase-

    eld or front-tracking-type methods; (2) modeling

    of solidication processes and microstructural fea-

    tures using averaging methods; (3) modeling of

    grain structure formation using physically based

    Cellular Automata or ``Granular Dynamics''

    methods. All three are important since the macro-

    scopic scale of a solidication process (typically

    cmm), the grain size (typically mmcm) and the

    characteristic length of the microstructure (mm)

    encompass six orders of magnitude and cannot be

    Fig. 14. Colony structure simulated using a phase-eld model for the directional solidication of aeutectic alloy with a dilute ternary impurity [169].

    60 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    19/28

    taken into account simultaneously. Their main

    characteristics are briey discussed hereafter. It

    should be emphasized that the smallest size of the

    microstructure (mm) is still three to four orders of

    magnitude larger than the size of the atoms or mol-

    ecules or the thickness of the solidliquid interface.

    This much ner scale still sets another limit to be

    accounted for in molecular modeling (which will

    not be treated here) or in any realistic phase-eld

    simulation, which precisely intends to model the

    gradual transition from liquid to solid.

    6.1. Modeling of microstructure

    In most metallic alloys solidied under normal

    conditions, microstructure formation is controlled

    by solute diusion and curvature, heat diusion

    occurring over much longer distances (i.e. Lewis

    number of the order of 104 for most metals).

    Simulation at this level normally requires following

    the interface separating the solid and liquid phases(front tracking). This has been achieved successfully

    in simple two-dimensional geometry using either the

    boundary element method (BEM) [186] or the nite

    element method (FEM) [187]. In the rst technique,

    only the interface is enmeshed and the Greens func-

    tions are used to solve the diusion problem. In the

    second method, dynamic remeshing of the domain

    is necessary. These methods are accurate but di-

    cult to implement even in two dimensions.

    Furthermore, topological changes such as coalesc-

    ence (merging of two dendrite arms) cannot be

    handled. They have been of great use to calculate

    the transient from a planar front to cells and the

    growth kinetics of the dendrite tip [186].

    In pseudo-front-tracking techniques [188190],

    the solidliquid interface is spread over only one

    mesh of the nite dierence (FDM) or nite volume

    (FVM) enmeshment and the concept of the volume

    fraction of solid, f (or liquid), is introduced: it is

    equal to unity in the solid, zero in the liquid and in-

    termediate for the ``interface meshes''. Among the

    advantages of such methods, fairly easy implemen-

    tation and computation speed can be mentioned.

    However, the error associated with the estimation

    of curvature from the divergence of the normalized

    gradient of f is large (H1030%). Since preferred

    growth directions and dendrite tip kinetics are gov-erned by the small anisotropy of the interfacial

    energy (H110% in metallic alloys), such methods

    can only give qualitative results.

    In the phase-eld method, the diuse nature of

    the solidliquid interface of metallic alloys is con-

    sidered and f varies continuously from 0 to 1

    over a certain thickness, d. Using a free energy or

    entropy formulation, two equations governing the

    evolution of the phase eld and the evolution of

    either heat or solute can be derived and solved

    using an explicit FDM. No front tracking being

    required, the technique is ecient and capable of

    reproducing most of the phenomena associated

    with microstructure formation (dendrite tip kin-

    etics, preferred growth direction, coarsening, co-

    alescence, etc.). Initially developed for thermal

    dendrites in two dimensions, it has been extended

    to solutal dendrites [24] and three dimensions [9,

    12]. As an example, Fig. 1 shows a thermal den-drite growing along h100i directions [9] and a solu-

    tal NiCu dendrite growing along h110i directions

    [12]. However, the technique also has some disad-

    vantages. The rst one is related to the eective

    thickness of the diuse interface, d, of alloys (H1

    5 nm) which is three to four orders of magnitude

    smaller than the typical length scale of the micro-

    structure. Since it must spread over several points

    of the mesh, this considerably limits the size of the

    simulation domain, even if d is multiplied by some

    arbitrary factor (10100). It is to be noted that

    this upscaling of d biases curvature eects by

    introducing some ``numerical curvature'' and also

    induces coalescence of dendrite arms at a much

    earlier stage of growth. The second problem arises

    from the attachment kinetics term that plays a sig-

    nicant role in the phase-eld equation, unlike

    microstructure formation of metallic alloys at low

    undercooling. These two factors have so far lim-

    ited phase-eld simulations of alloy solidication

    with realistic solid-sate diusivities to relatively

    large supersaturations. Recent mathematical and

    computational advances, however, are rapidly

    changing this picture. Some of the recent advances

    include: (1) a reformulated asymptotic analysis ofthe phase-eld model for pure melts [9, 11] that

    has (i) lowered the range of accessible undercool-

    ing by permitting more ecient computations with

    a larger width of the diuse interface region (com-

    pared with the capillary length), and (ii) made it

    possible to choose the model parameters so as to

    make the interface kinetics vanish; (2) a method to

    compensate for the FDM grid anisotropy [11]; (3)

    an adaptive FEM formulation that renes the

    zone near the diuse interface and that has been

    used in conjunction with the reformulated asymp-

    totics to simulate two-dimensional dendritic growth

    at low undercooling in two dimensions [10]; (4) astochastic Monte Carlo treatment of the large-

    scale diusion eld that provides an alternative to

    adaptive mesh renement that has been im-

    plemented in both two dimensions and three

    dimensions (Fig. 1) at low undercooling [14]; (5)

    the implementation in the method of uid ow

    eects [191193]; and (6) the extension of the tech-

    nique to other solidication phenomena including

    eutectic [194196] and peritectic reactions [69], and

    the interaction of dendrites with surfaces [197].

    BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES 61

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    20/28

    6.2. Modeling of processes and average

    microstructural features

    Modeling of solidication processes and micro-

    structural features has beneted from two main

    contributions: (i) the introduction of averaged con-

    servation equations previously developed for dipha-

    sic media [198200]; and (ii) the coupling of these

    equations with microscopic models of solidicationdescribing grain structure formation and other

    microstructural features (e.g. secondary arm spa-

    cing, microsegregation model, etc.) [200204].

    When conservation equations are averaged over

    the liquid and solid phases, the interfacial continu-

    ity condition automatically vanishes and average

    entities (e.g. mean specic mass, enthalpy or solute

    concentration) appear. Because the constitutive

    equations of the solid and liquid are widely dierent

    (typically the viscosity increases by 20 orders of

    magnitude during solidication), the momentum

    conservation equation is averaged only over the

    liquid phase. This introduces the interfacial bound-

    ary condition in the form of a drag term when the

    solid is supposed to be xed (packed bed).

    Beckermann and co-workers have extended this

    averaged equations formalism in order to encom-

    pass the situation of free moving equiaxed grains

    and make a smooth transition with the packed bed

    limit [205].

    The same authors have also coupled these aver-

    aged conservation equations with microscopic

    models of grain structure formation in a way simi-

    lar to the micromacro approach reviewed in Ref.

    [206] but including the eect of convection [207

    209]. An example of this is shown in Fig. 15. At

    early times (rst panel), the grains nucleate at the

    left wall and are swept by the convection currents

    around the cavity. At later times, when the grains

    have grown, they settle and form a bed of increas-

    ing height (next panels). The nal structure is highlynon-uniform (right panel), with a lower grain den-

    sity (i.e. larger grains) observed near the top due to

    the settling eect (after Ref. [208]). Convection

    modies the growth kinetics of the dendrite tips

    and thus of the grain envelope [210], but also

    entrains free grains. Experimentally based laws

    similar to Stokes drag are now available for high

    Reynolds number, non-spherical and porous grains

    [211]. The diculty in such an approach is to

    include the fragmentation of dendrites induced by

    convection [198, 212], although a preliminary

    attempt was presented in Ref. [212].

    Averaged conservation equations, coupled or not

    with detailed microscopic models of solidication,

    have been applied primarily to the problem of

    macrosegregation induced by thermal or solutal

    convection. Most of the developments in this area

    are based on structured FVM meshes and for two-

    dimensional geometries. These methods have been

    extended recently to small three-dimensional

    domains and FEM [213] and a benchmark compari-

    son between FVM and FEM has shown that these

    methods are quite sensitive to the formulation, in

    Fig. 15. Computed time evolution of the number density of equiaxed dendritic grains during solidica-tion of an Al4% Cu alloy in a cavity cooled from the left side [208].

    62 BOETTINGER et al.: SOLIDIFICATION MICROSTRUCTURES

  • 8/7/2019 SOLIDIFICATION MICROSTRUCTURES

    21/28

    particular regarding the formation of freckles (seg-

    regated channels) [214]. Implementation of a macro-

    segregation model based on averaged equations into

    a full-scale, three-dimensional simulation was

    recently accomplished [215]. The eect of grain

    movement on macrosegregation has been addressed

    in Refs [208, 209].

    6.3. Modeling of grain structures using stochastic

    methods

    Grain structure formation can be modeled suc-

    cessfully with averaging methods (see previous sec-

    tion). Such methods are particularly suitable when

    the grain size is small with respect to the scale of

    the process (e.g. in continuous casting of inoculated

    aluminum alloys) and/or when only one mor-

    phology is present (e.g. columnar or equiaxed).

    They are however unable to predict grain compe-

    tition in the columnar zone and the associated tex-

    ture evolution, and furthermore cannot provide a

    representation of the microstructure. The predictionof morphology transitions (from outer equiaxed to

    columnar and from columnar to equiaxed) [175,

    204, 216] is also quite dicult with averaging

    methods. In order to overcome these shortcomings,

    stochastic methods have been developed over the

    past decade [217225]. It should be pointed out that

    the stochastic aspect is only related to nucleation

    (random location and orientation of nuclei) whereas

    growth is usually treated in a deterministic way.

    Two types of models can be distinguished:

    (i) Cellular Automata (CA) have been developed

    for dendritic grain structures and can treat arbi-

    trary shapes and grain competition [217222]. Inthis technique, the solidication domain is

    mapped with a regular arrangement of cells and

    each grain is described by a set of cells, those

    located at the boundary (i.e. in contact with

    liquid cells) being active for the calculation of

    the growth process.

    (ii) In ``Granular Dynamics'' (GD) techniques,

    the surface of each grain is subdivided into an

    ensemble of small facets [223225]. The growth

    stage of each grain is then described by a set of

    parameters, e.g. the position of its center, the

    radial positions of its facets and their status

    (contact with the liquid or with another grain),

    etc. This latter technique is more appropriate for

    nearly spherical morphologies (e.g. equiaxed

    eutectics or globulitic grains) and can handle the

    transport of equiaxed grains fairly well.

    Although it has been demonstrated that CA can

    also treat the movement of equiaxed grains, it is

    particularly well adapted to describe grain compe-

    tition and texture evolution in dendritic columnar

    specimens, as