Skip to main content
Philosophical transactions. Series A, Mathematical, physical, and engineering sciences logoLink to Philosophical transactions. Series A, Mathematical, physical, and engineering sciences
. 2019 Mar 4;377(2143):20180205. doi: 10.1098/rsta.2018.0205

Thermodynamics of rapid solidification and crystal growth kinetics in glass-forming alloys

P K Galenko 1, V Ankudinov 2,3, K Reuther 1, M Rettenmayr 1,, A Salhoumi 4, E V Kharanzhevskiy 2
PMCID: PMC6460068  PMID: 30827218

Abstract

Thermodynamic driving forces and growth rates in rapid solidification are analysed. Taking into account the relaxation time of the solute diffusion flux in the model equations, the present theory uses, in a first case, the deviation from local chemical equilibrium, and ergodicity breaking. The second case of ergodicity breaking may exist in crystal growth kinetics of rapidly solidifying glass-forming metals and alloys. In this case, a theoretical analysis of dendritic solidification is given for congruently melting alloys in which chemical segregation does not occur. Within this theory, a deviation from thermodynamic equilibrium is introduced for high undercoolings via gradient flow relaxation of the phase field. A comparison of the present derivations with previously verified theoretical predictions and experimental data is given.

This article is part of the theme issue ‘Heterogeneous materials: metastable and non- ergodic internal structures’.

Keywords: fast transformations, rapid solidification, growth of crystals, glass transition

1. Introduction

Rapid solidification can be initiated by fast quenching or slow cooling to avoid premature nucleation and to attain deep undercoolings. Cooling rates may reach 106 (K s−1) for melt spinning, or atomization by gas, impulse or in a drop tube; temperature gradients can reach 109 (K m−1) in laser annealing, and undercoolings of up to 300–500 (K) can be achieved (see refs. [16] for details of techniques and measurements). Such large driving forces lead to solidification rates in the order of several tens of metres per second [5,7]. Considering that the characteristic diffusion speed is in the order of several metres or tens of metres per second in metallic alloys and of several centimetres or tens of centimetres per second in semiconductors [8], it becomes obvious that the solid/liquid interface may exhibit velocities in the order of or even larger than the diffusion speed. In such a case, the rapid solidification front may undergo diffusionless (chemically partitionless) transformation which may proceed in a wide or narrow interval of driving forces [3,4,9,10].

Progress in the formulation of the thermodynamical basis [1116] has led to an extended description [17] going beyond the hypothesis of local equilibrium that is stipulated by broken ergodicity in rapidly solidifying samples [18]. Using a local non-equilibrium model of dendritic growth [19], the kinetics of rapid solidification in slowly and deeply undercooled samples has been described consistently using experimental data on the solidification kinetics of a number of interstitial and substitutional solution phases during dendritic growth [2023]. In these studies, it was shown that the interface velocity monotonically increases with undercooling (even though specific peculiarities in the transition from diffusion limited to thermally controlled growth of dendrites were found). Rapid kinetics with an unusually steep dendrite velocity/undercooling relationship has also been obtained in intermetallic alloys with an order–disorder transition [24,25]. However, the velocity of dendrites in melts of intermetallic phases also exhibited a monotonic increase with increasing undercooling. Another type of velocity/undercooling relationship can be experimentally found in the kinetics of dendritic solidification of glass-forming metals and alloy melts. For glass-forming samples produced from metals, alloys, oxides, organics or chalcogenides [2628], the growth rate has a maximum value at a defined undercooling that lies between the glass-transition undercooling and zero undercooling directly at the melting (or liquidus) temperature. Moreover, for the alloy Cu50Zr50 (concentrations in %), a transition from purely thermodynamic to kinetically controlled growth has been experimentally observed for the first time [27]. Attempts to describe such transitions using common dendrite growth theory [29] lead to a smooth and gradual behaviour of velocity versus undercooling, describing the main part of the experimental data well, but overestimating velocities at the highest undercooling where the so-called ‘abrupt drop in kinetics’ occurs. The thermodynamic and kinetic analysis of the solidification behaviour in glass-forming samples is the main goal of the present paper.

The article is divided into two main parts. In the first part (§§2 and 3), we analyse thermodynamic driving forces and growth rates in dendritic rapid solidification and compare them with previously verified theoretical predictions. Here the solidification can be considered a rapid process if the solid/liquid interface moves with a velocity comparable to the diffusion speed in the bulk liquid. In this case, the relaxation time of the solute diffusion flux is not negligible and should be included in the model equations that characterize deviations from local equilibrium and ergodicity breaking [18]. The second part of the article is devoted to the analysis of dendritic solidification of congruently melting alloys in which the chemical composition does not influence the growth velocity due to the absence of chemical segregation. A dendrite growth model for predicting the solidification kinetics of glass-forming alloys is formulated in §4. The model consists of a system of equations describing (i) the undercooling balance at the dendritic surface, (ii) the stability condition for the dendrite tip at arbitrary growth Péclet numbers, and (iii) the kinetic contribution following on from the travelling wave solution of the phase field model (see appendices). In this case, we also deal with rapid solidification considering that gradient flow relaxation is used in the model equations describing the deviation from local equilibrium and ergodicity breaking [18].

2. Entropy change in rapid solidification

Consider a solidifying sample of a volume v0 of a binary alloy. The concentration distribution of solute is given by C(x, t), and at the interface the concentrations are C*L in the liquid phase and C*S in the solid phase. The temperature T of the sample is constant and the total pressure tensor P is related to the viscous pressure Pν by P = pU + Pν, where p is the equilibrium (static) pressure and U the identity tensor. As usual, Pν is split into two parts, a bulk pressure pν=13tracePν and a deviatoric part P^ν, so that

P=pU+pνU+P^ν. 2.1

The necessity of the pressure introduction is dictated by (i) accumulated stresses around rapidly moving interface and (ii) the difference in density of the liquid and solid leading to shrinkage effects. For example, an increasing dislocation density in the microstructure of alloys with increasing solidification rate [30] directly testifies that rapid solidification results in an accumulation and relaxation of stresses.

(a). The bulk entropy

The temporal evolution of the concentration field follows from the formalism of extended irreversible thermodynamics (EIT) [31,32]. The total entropy of the entire system of the volume vv is described by

TS(t)=vv[μACA+μBCB+αAJA2+αBJB2+αP:P]dvv, 2.2

where CA is the concentration of the A-atoms having the chemical potential μA, CB is the concentration of the B-atoms having the chemical potential μB, JA and JB are the diffusion fluxes of the A- and B-atoms, respectively. The thermodynamic coefficients αA and αB are defined by

αA=τD(A)D(A)μACA=1(VD(A))2μACA,αB=τD(B)D(B)μBCB=1(VD(B))2μBCB, 2.3

where τ(A)D and τ(B)D are the relaxation times of the diffusion fluxes, JA and JB, respectively, D(A) and D(B) the diffusion coefficients of the A- and B-atoms, respectively, and V(A)D = (D(A)/τ(A)D)1/2 and V(B)D = (D(B)/τ(B)D)1/2 are the maximum speeds for diffusion of the A- and B-atoms, respectively.

Using the expression for the pressure (2.1) and the respective expression for αP (see [31,32]), the entropy (2.2) can be re-written as

TS(t)=vv[μACA+μBCB+αAJA2+αBJB2pv+vα0(pν)2+vα2P^ν:P^ν]dvv, 2.4

where the thermodynamic coefficients α0 = ζ−1B and α2 = (2ηS)−1 are defined by the bulk viscosity ζB and shear viscosity ηS, respectively.

We use the concentration definitions C = CB = 1 − CA together with the fluxes JD =  − JA = JB, J2D = J2A = J2B. Then, considering further only a semi-infinite one-dimensional solidification domain, equation (2.2) simplifies to

S(x,t)=0[seq(x,t)+sne(x,t)]dx, 2.5

where the local equilibrium part of the entropy density is

Tseq(x,t)=μA(1C)+μBCpv, 2.6

and the local non-equilibrium part of the entropy density is

Tsne(x,t)=αDJD2+vα0(pν)2+vα2(Pxxν)2, 2.7

with αD = αA + αB and Pνxx the component of viscous pressure along the x-axis.

(b). Entropy produced by the moving interface

Neglecting the diffusion processes in the solid phase, JD(x, t) = 0:x < x*(t), where x*(t) is the position of the solid/liquid interface, consider now the change of entropy in the entire system caused by a small movement ε of the interface in a small step of time Δt, corresponding to an interface velocity of V = εt (figure 1). In the area swept by the interface, x = x*(t), entropy is produced by the moving interface. The local equilibrium part of entropy increase is described by

ΔΔt(TSeq)=[TSeq(t+Δt)TSeq(t)]=ε[(1CS)(μASμAL)+CS(μBSμBL)pΔv], 2.8

where Δv is the volume change and, for isobaric conditions, equation (2.8) is equivalent to the expression given in refs. [11,16]. In this occasion, all diffusion fluxes have vanished in the area swept by the interface, yielding a total entropy increase as

ΔΔt(TSne)=TSne(t+Δt)TSne(t)=x(t0)x(t0+ε)[αDJD2(x,t0)+vα0(pν)2+vα2(Pxxν)2]dx. 2.9

Figure 1.

Figure 1.

Two snapshots of the concentration distribution at a solidification front, separated by a small time step Δt, where the interface has the position x* and moves with a constant velocity V to the right of the figure. (Online version in colour.)

The entropy production is evaluated by applying the differential quotient in the vanishing limit and applying the relation V = εt by

ddt(TS)e=limΔt0TSe(t+Δt)TSe(t)Δt=limΔt0εΔt[(1CS)(μASμAL)+CS(μBSμBL)pΔv]=V[(1CS)(μASμAL)+CS(μBSμBL)pΔv]. 2.10

With no change of volume, Δv = 0, equation (2.10) becomes identical to the classical result obtained in [11,16]. For the local non-equilibrium part, equation (2.9), we simplify the analysis by neglecting the bulk pressure, i.e. pν → 0. Then, a Taylor series for the square flux and the square of the deviatoric part of the viscous pressure near the interface is introduced by

JD2(x+ε,t0)=JD2(x,t0)+εddxJD2(x,t0)+and(Pxxν)2(x+ε,t0)=(Pxxν)2(x,t0)+εddx(Pxxν)2(x,t0)+} 2.11

Using equation (2.11), the integral in (2.9) can be evaluated as

x(t0)x(t0+ε)[αDJD2(x,t0)+vα2(Pxxν)2]dx=αD[εJD2(x,t0)+ε22ddxJD2(x,t0)+]+vα2[ε(Pxxν)2(x,t0)+ε22ddx(Pxxν)2(x,t0)+]. 2.12

This leads to the following expression for the local non-equilibrium part in the entropy production

ddt(TS)ne=limΔt0TSne(t+Δt)TSne(t)Δt=αDlimΔt0[εΔtJD2(x,t0)+ε22ΔtddxJD2(x,t0)+]+vα2limΔt0[εΔt(Pxxν)2(x,t0)+ε22Δtddx(Pxxν)2(x,t0)+]=αDVJD2(x,t0)+αD[limΔt0ΔtV2ddxJD2(x,t0)+]+vα2V(Pxxν)2(x,t0)+vα2[limΔt0ΔtV2ddx(Pxxν)2(x,t0)+]αDVJD2(x,t0)+vα2V(Pxxν)2(x,t0), 2.13

where Pνxx(x*, t0) is the deviator of the viscous pressure at the interface and JD(x*, t0) is the diffusion flux away from the interface. Note that this flux generally does not vanish (except in the special case of diffusionless solidification in which C*L = C*S), leading to the local non-equilibrium contribution for interface velocities V smaller than the solute diffusion speed VD in bulk liquid, i.e. with V < VD, where VDV(B)D is the maximum speed for the diffusion of B-atoms.

Finally, using equations (2.10) and (2.13), the total entropy production per unit time at the interface is given by

d(TS)Interfacedt=V[(1CS)(μASμAL)+CS(μBSμBL)+αDJD2(x,t0)]V[pΔvvα2(Pxxν)2(x,t0)]. 2.14

(c). Gibbs free energy change on rapid solidification

Using the established system of thermodynamic relations [14,16,17], one obtains

d(TS)Interfacedt=dGdt=vm1VΔGanddGdt=JDXD+JCXC+JPXP,} 2.15

where G and ΔG are the Gibbs free energy and the Gibbs free energy change on solidification, respectively. Taking into account that the change in Gibbs free energy ΔG proceeds at a constant volume, i.e. at Δv = 0, then, as follows from equation (2.14), the diffusion flux JD and the crystallization flux JC together with their conjugate driving forces, XD and XC, respectively, are defined in equation (2.15) by

JD=vm1V(CLCS),XD=ΔμAΔμB+αD(CLCS)V2,JC=vm1V,XC=(1CL)ΔμA+CLΔμBandJP=vm1V,XP=vα2(Pxxν)2,} 2.16

where ΔμA = μSA − μLA and ΔμB = μSB − μLB are the differences in the chemical potentials for A- and B-atoms, respectively. In the linear approximation, JDXD, JCXC and JPXP one can obtain the following relations between fluxes and conjugate driving forces:

(CLCS)Vvm=LD[ΔμAΔμB+αD(CLCS)V2],Vvm=LC[(1CL)ΔμA+CLΔμB]andVvm=LPvα2(Pxxν)2,} 2.17

where LD, LC and LP are the kinetic coefficients for diffusion and crystal growth due to the driving force for atom attachment and due to the elastic and plastic relaxation, respectively. Several limiting cases can be outlined from equation (2.17). First, for the diffusionless (chemically partitionless) solidification, the analytical solution [33] gives naturally C*L = C*S at VVD. Then, the first equation gives the expression ΔμA = ΔμB, which indicates the equality of the chemical potential differences for rapid diffisionless solidification. From this follows the second case, which describes the growth kinetics as for a pure one-component substance upon the diffusionless solidification given by the second equation from equation (2.17): V = vmLCΔμA with VVD. The third equation of equation (2.17) predicts that the viscous relaxation of stresses may cause the motion of the solid/liquid interface.

Using the relations from equation (2.16) in equation (2.15), one obtains

ΔG=(1CS)ΔμA+CSΔμB+αD(CLCS)2V2+vα2(Pxxν)2. 2.18

This expression represents the Gibbs free energy change at the solid/liquid interface and describes the driving force for interface motion. Neglecting relaxation of the diffusion flux αD = 0, and relaxation of viscous stresses α2 = 0, equation (2.18) takes its standard form given and analysed in refs. [11,13,14,16]. The additionally included effects in equation (2.18), which may accompany rapid solidification, give the following contribution to ΔG (see also refs. [34,35]): the diffusion flux relaxation ∝V2, and the viscous stress relaxation ∝(Pνxx)2 should increase the driving force for interface motion.

The established thermodynamic relation for chemical diffusion in rapid solidification (2.18) introduces a pressure term, ∝(Pνxx)2, that may have the potential to advance the current understanding of non-equilibrium solidification. More specifically, it may bring about:

  • new insights into undercooled solidification under hydrostatic pressures or in centrifugal casting [36];

  • the analysis of dendrite growth on the inner core of Earth [37];

  • influence of (elastic and plastic) stresses and dislocation motion on crystal growth kinetics [38].

The quantitative effect of the pressure term ∝(Pνxx)2 on solidification will be clarified in the future works.

Finally, one should note that a thermodynamic analysis was used to describe the liquidus slope in kinetic phase diagram with the appropriate methods described in detail in refs. [12,13,15,17]. Using these methods and neglecting the viscous stresses, Pνxx = 0, one can, in particular, find from the driving force, equation (2.18), the velocity-dependent liquidus slope, mv(V ), as [34,35]

mv={me1ke{1kv+ln(kvke)+(1kv)2V2VD2},V<VD,melnkeke1const,VVD, 2.19

where me and ke are the equilibrium liquidus slope and solute partition coefficient (which can be obtained from the phase diagram or calculated using CALPHAD), respectively, and kvC*S/C*L is the velocity-dependent partition coefficient. The function (2.19) differs slightly from the one derived for the approximation of dilute mixtures [17]. For finite concentrations, the liquidus slope, equation (2.19), depends on the square of the interface velocity, mv∝(V/VD)2, explicitly.1 In the theory of diluted mixtures/alloys, the kinetic liquidus was obtained as linearly proportional to the velocity [17], i.e. mvV/VD. However, as we show in the following section, the predictions of the crystal growth kinetics are quantitatively valid using equation (2.19) and ref. [17] due to the fact that the strongest dependence of mv on V is dictated by the nonlinear behaviour of the atomic distribution function at the interface kv(V ). An estimation of the pressure influence (with Pνxx≠0) on the velocity-dependent kinetic liquidus should be specially given. On the one hand, the pressure plays an insignificant role in the crystal growth under usual circumstances (see Chapter 2 in the monograph of Glicksman [39]). On the other hand, the growth of dendrites might be essentially influenced on the inner core of the Earth from the magma's side [37], where the pressure and temperature are important parameters for dynamical equilibrium at the solid/liquid interface. How important the effect of viscous pressure relaxation on the rapidly moving solid/liquid interface is should be clarified in forthcoming works.

3. Dendrite growth model

In the following, we consider two models for rapid dendritic growth. The first model was formulated to describe primary growth developed for dendrites of interstitial and substitutional solutions. The model was tested in numerous cases of binary and ternary alloys solidifying from undercooled melts [1923]. We formulate this model for two reasons. First, we compare kinetic curves for two cases of liquidus line slopes as formulated in §2 and as derived in [17]. The second model appears as an extended advanced first model to describe kinetic curves in glass-forming alloys exhibiting a maximum in the dendrite growth velocity/melt undercooling relationship.

(a). Undercooling balance and stability criterion

Equations of a sharp interface model were formulated for the analysis of rapid solidification taking into account deviations from equilibrium at the solid/liquid interface and in bulk liquid [19,20,33,40,41]. At the solid/liquid interface, a deviation from local thermodynamic equilibrium may arise due to atom attachment kinetics and solute trapping. A deviation from thermodynamic equilibrium in the liquid may exist in the diffusion field of solutes, which has no time to relax to local chemical equilibrium due to the rapidly growing dendrite. This model combines a selection criterion for stable dendritic growth mode and the balance of different undercooling contributions at the dendritic tip.

The total undercooling, ΔT = Tm + meC0 − T0, represents the undercooling balance at the dendrite tip as

ΔT=ΔTT+ΔTC+ΔTN+ΔTR+ΔTK, 3.1

where Tm is the melting temperature of the solvent, C0 and T0 are the initial (i.e. nominal or far-field) composition and temperature of the liquid, respectively. The balance includes the thermal undercooling ΔTT

ΔTT=TQIv(PT), 3.2

the undercooling ΔTC due to solute redistribution at the solid–liquid interface

ΔTC={kvΔvIv(PC)1(1kv)Iv(PC),V<VD,0,VVD, 3.3

with the Ivantsov functions IvT(C)(PT(C)) defined by

IvT(C)(PT(C))={PT(C)exp(PT(C))E1(PT(C)),in the 3D space,πPT(C)exp(PT(C))erfcPT(C),in the 2D space, 3.4

exponential integral function E1(P_T(C)) of the first kind and the thermal/chemical Peclét numbers PT(C),

E1(PT(C))PT(C)exp(u)udu,PT=VRDT,PC=VRDL, 3.5

where TQ = ΔHf/cp is the adiabatic temperature of solidification (hypercooling), ΔHf is the enthalpy of melting, cp is the heat capacity, DT and DL are the thermal and chemical diffusion coefficients, respectively, and V is the velocity of the dendrite tip with radius R.

The velocity-dependent non-equilibrium interval Δv of solidification in equation (3.3) is given by

Δv={mvC0(1kv)kv,V<VD,0,VVD. 3.6

Note that if the dendrite tip velocity V is equal to or greater than the solute diffusion speed VD, the constitutional undercooling ΔTC is equal to zero, corresponding to the transition to the diffusionless regime of solidification.

Finally, the undercooling ΔTN arising due to the shift of the equilibrium liquidus line from its equilibrium position in the kinetic phase diagram of steady-state solidification is given by

ΔTN=(memv)C0. 3.7

The curvature undercooling ΔTR due to the Gibbs–Thomson effect and the kinetic undercooling ΔTK that change the driving force for interface movement due to the atomic attachment from the liquid to the solid phase are defined as

ΔTR=2d0TQRandΔTK=Vμk, 3.8

where d0 is the capillary length and μk is the interface kinetic coefficient.

The velocity-dependent liquidus slope is described by [17]

mv={me1ke{1kv+ln(kvke)+(1kv)2VVD},V<VD,melnkeke1const.,VVD, 3.9

and the velocity-dependent partition coefficient kvC*S/C*L is given by [33]

kv={(1V2/VD2)ke+V/VDI(1V2/VD2)[1(1ke)CL]+V/VDI,V<VD,1,VVD, 3.10

where C*S and C*L are solute concentrations of the solid and liquid at the dendrite tip, respectively, with

CL(PC)={C01(1kv)IvC(PC),V<VD,C0,VVD, 3.11

Here me is the slope of the liquidus line in the phase diagram of coexisting phases, and ke is the equilibrium partitioning coefficient of solute atoms.

Because equation (3.1) is the only equation for two variables, V and R, at the given (and experimentally measurable) undercooling ΔT, a second equation is required to close the problem, and it is given by the stability condition [41]

R={1σd0Δ0TQPTξT2ΔvPCξC,V<VD,1σd0Δ0TQPTξT,VVD, 3.12

with Δ0 the temperature interval of solidification in the phase state defined by

Δ0={me(ke1)C0ke,for alloys,TQ,for pure substance,C0=0. 3.13

Condition (3.12) defines the dendrite tip growing in the stable mode due to the anisotropy εc of the interfacial energy and the stability parameter σ*∝ε7/4c [4245]. The thermal stability function for rapid solidification is given by

ξT(PT)=[1+a1αdPT(1+a0DTβ0d0)]2, 3.14

where a0 and a1 are the asymptotic coefficients of sewing the large thermal Peclét numbers regime and growth kinetics regime, respectively, β0 = 1/(μkTQ) is the kinetic parameter of growth and αd = 15εc is the stiffness. The chemical stability function for rapid solidification is described by

ξC(PC,W)={[1+a2αdPC(1+a0DLβ0d0CD)]2,V<VD,0,VVD, 3.15

where a2 is the asymptotic coefficients of sewing the large thermal Peclét numbers regime, d0CD is the chemical capillary length and PC=PC/1V2/VD2. With the local equilibrium limit, VD → ∞, the stability growth mode by equations (3.12)–(3.15) describes the regime of Fickean diffusion taken into account by Trivedi & Kurz [46] for rapid dendritic growth [47]. Within the limit of small Peclét numbers, PT≪1 and P*C≪1, and with the local equilibrium limits VDI → ∞ and VD → ∞, equations (3.12)–(3.15) transform to the previously obtained conditions of Ben Amar and Pelce [43]. The selection criteria (3.12)–(3.15) are written for a fourfold symmetry of the crystal lattices. It can be generalized to other crystalline symmetries [48,49].

(b). Predictions of the model: dendrite velocity and tip radius

The model equations (3.1)–(3.15) describe (i) solute diffusion limited growth of dendrites (i.e. growth of ‘solutal’ dendrites at low undercoolings), (ii) solute diffusion limited and thermally controlled growth of dendrites (i.e. growth of solutal and thermal dendrites in the intermediate range of undercoolings), and (iii) purely thermally controlled dendritic solidification at high undercoolings. These regimes are shown in figure 2 and were previously described for alloy dendrites (see [22,41,50] and references therein).

Figure 2.

Figure 2.

Predictions of the model given by two main equations (3.1) and (3.12). Calculations were made using the parameters for Ni–0.7 at.%B taken from ref. [41] for a dendrite velocity V (a) and a tip radius R (b). Different growth regimes are separated from each other by the critical undercoolings, ΔT*1 and ΔT*2 such that the solute diffusion limited growth of dendrites proceeds for ΔT < ΔT*1 (range I), the intermediate stage consisting of solute diffusion limited and thermally controlled growth exists in the range ΔT*1 < ΔT < ΔT*2 (range II), and purely thermally controlled dendritic solidification occurs at ΔT > ΔT*2 (range III). The solid line was obtained for the kinetic liquidus slope given by equation (3.9) and the dashed line was calculated using the kinetic liquidus slope given by equation (2.19). The difference between both curves (dashed and solid) are nearly negligible and they are both consistent with experimental data in a whole range of measurable undercoolings and dendrite velocities within the error bars of experiment, see [41].

The transition from solute diffusion limited growth to the thermally controlled regime of growth in the range ΔT*1 < ΔT < ΔT*2 is characterized by the change from slow growth at ΔT < ΔT*1 to the increase of the interface velocity, figure 2a, and increase of the dendrite tip radius up to its maximal value followed by a decrease at further increasing velocities (figure 2b). Such nonlinear behaviour in V and R exists due to the change in the characteristic spatial lengths defining the dendrite velocity and tip radius, particularly the solute diffusion length and the thermal diffusion length.

At the critical undercooling, ΔT = ΔT*2, the velocity V = VD leads to complete solute trapping, C*L = C*S = C0, for the dendrite stems [51]. Beyond this critical point, dendrite growth occurs without solute redistribution and with the initial chemical composition, i.e. in the diffusionless (chemically partitionless) regime. Thus, in the range ΔT≥ΔT*2, the dendrite growth is thermally controlled only. This result has a clear physical meaning: a source of concentrational disturbances at the solid/liquid interface moving at a velocity equal to or higher than the maximum speed of these disturbances cannot disturb the liquid phase ahead of itself. Therefore, when the interface velocity V passes through the critical point V = VD, the solidification mechanism changes qualitatively and the undercooling in liquid will not depend on the solute distribution during diffusionless growth. In a microscopic interpretation, the atoms have no time for diffusion jumps if VVD; they are all stochastically attaching to the interface and instantaneously captured by the solid phase.

Finally note that the transition to diffusionless solidification proceeds sharply at V = VD and ΔT = ΔT*2 with a steep change of slope of the ‘velocity versus undercooling’ and ‘tip radius versus undercooling’ curves (figure 2). Such sharp transition in the growth kinetics at some critical value of non-equilibrium governing parameter (undercooling, supersaturation) is related to the ‘kinetic phase transitions’, as defined by Alexander Chernov [52], that was observed in experiments on rapid solidification [7,8,22] and analysed using different models (see ref. [25] and references therein).

4. Dendritic solidification in glass-forming metals and alloys

(a). Model equations

The curves shown in figure 2a exhibit a monotonic increase of the dendrite tip velocity with the undercooling that is typically observed in experiments on solidifying interstitial and substitutional alloys [7,8,13]. However, there is a wide class of alloys exhibiting a maximum of the dendrite velocity that is situated between liquidus temperature and glass temperature (at which an amorphous phase forms and the building of primary crystalline structures is impossible) [26,28]. This class includes glass-forming alloys and, quite frequently, congruently melting alloys. To describe such behaviour, the model of dendritic solidification in equations (3.1)–(3.15) can be reformulated and extended to describe the kinetic curves in glass-forming alloys.

Consider the solidification of a glass-forming alloy which is congruently melting without chemical segregation. In this case, the system of equations (3.1)–(3.15) is reduced to the balance of undercoolings as

ΔT=ΔTT+ΔTR+ΔTK, 4.1

with the thermal, capillary/curvature and kinetic contributions

ΔTT=TQIv(PT),ΔTR=2d0TQR,ΔTK=Vμk(ΔT), 4.2

where the kinetic coefficient μkT) is a strongly dependent function of undercooling ΔT. The stable growth mode of the dendrite tip is defined from equations (3.12)–(3.15) as

R=1σd0Δ0TQPTξT,σ=σ0εc7/4andξT=[1+a1αdPT(1+a0DTβ0d0)]2.} 4.3

All the coefficients and functions of equations (4.1)–(4.3) are defined in §3.

(b). Kinetic undercooling

The crystal growth models based on the rate kinetic theory give excellent predictions of the molecular dynamics simulation data at small undercoolings (see appendix A, figure 5a). With the increasing undercooling, the crystal growth velocity exhibits clear nonlinearity which may not be predicted by the models based on the rate kinetic theory (see appendix A, figure 5b).

So far, the nonlinearity in the velocity versus undercooling relationships has been obtained similarly to curves with saturation, which are limited by the maximum velocity of the phase field (see [53] as well as the solid curve with saturation in figure 5b). To predict the nonlinear curves with the velocity maximum [5456] exhibiting one of the distinct characteristics of the crystallization of glass-forming metals and alloys [26,28], we use the travelling wave solution (B 17) of the kinetic phase field model for obtaining the dependence of the kinetic undercooling on the interface velocity.

Using the approximations ΔGkBT and the simplest possible expression for the driving force [57]

ΔG=ΔHf(ΔTK)Tm,ΔTK=TmT, 4.4

equation (B 17) takes the form (see appendix B)

V=βK(PFM)ΔTK, 4.5

which is consistent with the approximations of the kinetic equations (A 4) and (A 5) and for which the kinetic coefficient of growth depends on the kinetic undercooling ΔTK as

βK(PFM)=Dϕ(ΔTK)ΔHfγTm1+[(Dϕ(ΔTK)ΔHf/γTmVϕ(ΔTK))ΔTK]2. 4.6

The maximum propagation speed of the phase field depends on the undercooling through the diffusion of the phase field as

Vϕ(ΔTK)=Dϕ(ΔTK)τϕ, 4.7

where τϕ is the relaxation time of the gradient flow, ∂ϕ/∂t, of the phase field (which is taken in the present analysis as a constant, independent of temperature). The diffusion coefficient of the phase field is taken in the form

Dϕ(ΔTK)=Dϕ0exp(EATmΔTKTA), 4.8

where the diffusion factor D0ϕ, the temperature TA and the energetic barrier EA are the parameters of the phase field propagation. Particularly, TA controls the temperature at which a drastic change in the crystal growth kinetics may occur.

Figure 3 shows the dependence of the phase field diffusion coefficient, equation (4.8), on the undercooling (ΔTK > 0) and the superheating (ΔTK < 0). As the figure shows, the change in the barrier EA can lead to a qualitative change in the behaviour of the phase field diffusion. The higher value of EA (see the curve the curve E(2)A) gives a gradual change of the phase field diffusion around the equilibrium melting point and a steep decrease of the phase field diffusion at larger undercoolings far from the melting point. By contrast, for the smaller value of EA the curve E(3)A exhibits a linear change of the phase field diffusion with the change of ΔTK up to its zero value, Dϕ = 0.

Figure 3.

Figure 3.

Phase field diffusion coefficient versus kinetic undercooling. Curve E(1)A corresponds to EA = 42.11 K, fitted to the experimental data on crystallization of a Cu50Zr50 alloy (figure 4). Curve E(2)A corresponds to higher EA = 126.34 K; E(3)A corresponds to EA = 12.63 K. The coefficients D0ϕ = 3.35 × 10−10 (m2 s−1) and TA = 940 K are the same for all calculations.

One has to finally note that the form of the diffusion coefficient, equation (4.8), is similar to the form of the phase field mobility Mϕ(T) given by Novokreshchenova & Lebedev [58]. This form is inversely proportional to the kinematic viscosity described by the Vogel–Fulcher–Tamman expression [59] which predicts a strongly anomalous increasing viscosity as the temperature begins to approach the glass temperature. Function (4.8) shows inverse behaviour: as soon as the undercooling approaches its critical value, the phase field diffusion begins its steep decrease down to its zero value exactly at the critical undercooling.

(c). Kinetics of dendritic growth in Cu50Zr50

Cu50Zr50 is a glass-forming alloy which has been investigated in detail by heating, melting, cooling and solidifying droplets processed in an electrostatic levitation facility [27,60]. The dendrite growth velocity was measured for undercoolings of up to ΔT = 311 K. It was found that the velocity first increases and then decreases as the total undercooling increases. The congruently melting Cu50Zr50 alloy solidifies without chemical segregation and, together with the first measurements on Ag [61], it was the metallic system for which the transition from thermodynamically to kinetically controlled growth was firstly experimentally observed [27].

Wang et al. [29] suggested a model of dendritic growth in which the kinetic undercooling was defined via a dendrite growth coefficient that is dependent on the viscosity behaviour. They tried to describe the experiments of ref. [27] in the whole range of undercoolings ΔT(K) ≤ 311. The dash-dotted line in figure 4 shows that the experimental data are well described up to an undercooling of approximately ΔT = 290K. The interval 290 < ΔT (K) ≤ 311 shows a drastic decrease of the velocity with its a further abrupt drop at the critical undercooling ΔT = 311K that is not described by the smooth and gradually changing VT)-curve of Wang et al. (figure 4). The same abrupt behaviour has also been obtained for the solidification of Cu–Zr-based alloys, particularly Zr50Cu30Ni20 and Zr50Cu45Ni5 alloys [62]. Therefore, the model presented here, equations (4.1)–(4.8), has been applied to describe the experimental data on Cu–Zr-based alloys exhibiting a maximum of a velocity and an abrupt drop behaviour at the highest critical undercooling.

Figure 4.

Figure 4.

Experimental data on the dendrite growth velocity of Cu50Zr50 measured by Q. Wang et al. (open squares, [27]), R. Kobold et al. (open circles, [63]) and P. Paul (solid diamonds, [60]). The theoretical curves display: (1) solid curve, complete model (4.1)–(4.8); (2) dotted curve, model (4.1)–(4.8) with instant relaxation of the gradient flow, τϕ = 0; (3) dash-dotted curve, model of Wang et al. [29]. The insert shows how the theoretical prediction may lie far from experimental data if the gradient flow is not included in the model as an additional thermodynamic variable (see appendix B).

Using the material parameters of the Cu50Zr50 alloy from table 1, the dendrite growth velocity was calculated by the model, equations (4.1)–(4.8). Note, to obtain parameters for the phase field diffusion and relaxation time, we supposed that a part of the velocity versus undercooling curve is described mainly by the kinetic undercooling ΔTK. Therefore, just to obtain optimal fitting for τϕ, D0ϕ, EA and TA we have used the ‘kinetic undercooling shift’, ΔTΔk = 70 (K), which is a shift of the curve V − ΔT from the origin along the ΔT-axis.

Table 1.

Material parameters of the Cu50Zr50 alloy used in the present calculations.

parameter value source
Tm (K), melting temperature 1208 [27]
ΔHf (J kg−1), enthalpy of fusion 8.78 × 108 [27]
TQ (K), adiabatic temperature 204.87 [27]
d0 (m), capillary length 6.83 × 10−10 [27]
γ (J m−2), interface energy 0.6 [27]
εc (–), anisotropy of interface energy 0.15 present work
σ0 (–), constant of selection criterion 12.8 × 10−4 present work
τϕ (s), relaxation time 3.41 × 10−7 present work
D0ϕ (m2 s−1), diffusion factor 3.35 × 10−10 present work
EA (K), energetic barrier 42.11 present work
TA (K), pseudo-glass temperature 940 present work

Figure 4 shows the comparison of the present model with the predictions of Wang et al. and experimental data obtained from samples processed by an electromagnetic levitator on the Ground [27,63] and under microgravity generated during parabolic flights [60,62]. It can be seen that the solid line calculated by the full model (4.1)–(4.8) describes the whole set of experimental data including the abrupt drop in the velocity at the critical undercooling ΔT ≈ 311 K. If we exclude the local non-equilibrium effect of the relaxation of the gradient flow, i.e. τϕ → 0 and Vϕ → ∞ in equations (4.7) and (4.8), the square root will yield unity in equation (4.5) and we get the velocity as plotted by the dotted line in figure 4. More specifically, it is shown in the insert of figure 4 how the predicted velocity lies far from the experimental data if the local non-equilibrium effect in the phase field dynamics is not taken into account.

The necessity to introduce the gradient flow of the phase field as an additional thermodynamic variable (see appendix B) lies in the fact that with the increase of undercooling in glass-forming alloy one can find a transition from single atoms and single clusters to chains of connected clusters (see data of molecular dynamics simulations in [64]). Such chains represent long molecules which lead to a delay in the kinetic processes in the bulk of the liquid phase and at the solid/liquid interface. In the present model, this transition is phenomenologically expressed by the special form of the phase field diffusion coefficient (4.8) which describes the delay in the phase field propagation up to an interface velocity of zero. The critical undercooling at which one may obtain this interface velocity characterizes the impossibility of the phase field propagating due to the specific structure of the undercooled liquid represented by the very long chains of connected clusters. Behind this critical undercooling the remaining liquid may transform into the amorphous phase (see ref. [60] and references therein). Therefore, glass forms at undercoolings larger than the critical undercooling for the zero velocity of the solid/liquid interface. Details and numerical values for these critical undercoolings can be clarified in an atomistic and phenomenological investigation of crystallization and amorphization processes.

Finally, even though we describe experimental data on solidification kinetics of the Cu50Zr50 melt quite well, figure 4, a couple of important issues should be noted as well. First, the primary phase which may solidify from the undercooled Cu50Zr50 melt is the intermetallic phase of B2-type. Solidification of intermetallic compounds may be drastically influenced by the order–disorder transition at the rapidly moving interface [6569], therefore, this effect leading to pronounced disorder trapping should be introduced into the model of the Cu50Zr50 solidification. Second, intensive investigations of the liquid structure [70,71] and crystallizing phase [72] by in-situ high-energy synchrotron X-ray diffraction on electrostatically levitated samples show that Cu50Zr50 always solidifies as the B2-phase over the whole accessible undercooling range up to 310K. However, periodicity of such measurements was about 5 s [72], which might overcome the duration of decomposition of a primary metastable phase. Indeed, the primary crystalline phase can appear as a new metastable phase at high undercooling with its lifetime smaller than the second phase [62]. Therefore, phase selection and sequence of phase precipitation may also affect the solidification kinetics, which has to be considered as a further advancement of theoretical models for dendrite growth in glass-forming melts.

Appendix A. Appendix A. Models based on the rate kinetic theory

The rate kinetic theory (or thermally activated growth theory [73]) in application to melting/crystallization compares two atomic fluxes at the moving crystal/liquid interface [9]: the first flux is from the liquid to the crystal per unit time at a single kink or atomic micro-roughness, and the second one is from the crystal to the liquid. This results in the non-zero interface velocity

V=kT(M)(T)[1exp(ΔGkBT)], A 1

where kB is the Boltzmann constant and T is the absolute temperature. The temperature-dependent kinetic coefficient k(M)T in equation (A 1) can be found from various atomistic theories adapted to the concrete growth mechanism of crystals.

In the present work, we refer to theories describing the collision limited mechanism as formulated by Broughton, Gilmer and Jackson [74] and the diffusion limited mechanism as analysed by Wilson & Frenkel [75,76]. These theories summarize the kinetic coefficients as

kT(M)(T)={kT(CLT)=aλ3kBTmf0,kT(DLT)=D(T)aλ2f0exp(ΔHfkBT), A 2

with the superscript ‘M’ corresponding to ‘CLT’ in the collision limited theory or ‘DLT’ in the diffusion limited theory, λ is the displacement during crystallization which is proportional to the lattice parameter a, m is the atomic mass, f0 the fraction of liquid atom collisions with the solid which result in a crystallization event, ΔHf the melting enthalpy, D(T) the diffusion coefficient taken as [9] D(T)=a2ν~exp[ΔEB/(kBT)], ν~ the frequency of thermal vibrations of an atom (assumed as equal in the crystal and the liquid for the sake of simplicity), EB the activation energy for diffusion, exp[ΔHf/(kBT)] the probability of finding an atom of the liquid in the immediate vicinity of the kink on the crystal surface in the most advantageous activation complex corresponding to the barrier EB.

Using the approximation ΔGkBT

ΔG=ΔHf(ΔTK)Tm,ΔTK=TmT, A 3

which is adopted for a pure substance, equation (A 1) is reduced to its linear form as

V=βK(M)(T)ΔTK, A 4

with the temperature-dependent coefficients (A 2) transforming to the kinetic coefficients of crystal growth as

βK(M)={βK(CLT)=aλf03kBTmΔHfkBTm2,βK(DLT)=D(T)aλ2ΔHfkBTm2f0exp(ΔHfkBT), A 5

where Tm is the melting point and ΔTK is the undercooling.

In equations (A 1)–(A 5), the diffusion limited theory takes into account the diffusion transport coefficient D(T) of the bulk supercooled liquid; therefore, k(DLT)T(T) (as well as β(DLT)K) exhibits a strong temperature dependence which is generally associated with thermally activated processes. Contrary to this, the collision limited theory predicts a kinetic coefficient proportional to the mean particle velocity, that is, kT(CLT)T (and also βK(CLT)T) presenting the barrierless mechanism of growth. Quantitative estimations of kinetic growth coefficients for various crystal growth models can be found in [77,78] as obtained from atomistic simulations and in [79] as obtained from experimental measurements. However, the collision limited mechanism fails in many cases of crystal growth (see, e.g. supplementary material of ref. [54]). The diffusion limited theory also has difficulties quantitatively describing the growth kinetics in a wide range of temperatures [55,77]. Figure 5 shows the results of calculations by equations (A 1) and (A 2) for Ni crystals in comparison with data of molecular dynamics simulations [78,80] (all data for the computations are given in ref. [53]). In the relatively small range of undercoolings, 0 < ΔTK (K) < 50, and superheatings, −55 < ΔTK (K) < 0, figure 5a shows that MD data exhibit a linear behaviour of the interface velocity, both in melting and in crystallization, that is well described by the diffusion limited theory. For the large range of undercoolings, 0 < ΔTK (K) < 710, figure 5b shows that the equation following from the diffusion limited theory leads to inconsistent behaviour in comparison with MD data qualitatively and, as a consequence, quantitatively.

Figure 5.

Figure 5.

Velocity of crystal growth versus kinetic undercooling at a planar solid/liquid interface [53]. (a) Comparison of the predicted interface velocity V by equations (A 1) and (A 2) (dotted line, curve 2) of the diffusion limited growth (‘M=DLT’) and by equation (B 17) (solid line, curve 1) of the Phase Field Model (PFM) with data of molecular dynamics simulations (full circles) obtained by Mendelev et al. [78] for Ni in the case of low undercoolings 0 < ΔT (K) < 50 and superheatings −55 < ΔT (K) < 0, respectively. (b) Comparison of the predicted interface velocity V by equations (A 1) and (A 2) (dashed line, curve 2) of diffusion limited growth (‘M=DLT’) and by equation (B 17) (solid line, curve 1) of the PFM with data of molecular dynamics simulations (full circles) after Hoyt et al. [80] for Ni in the case of large undercoolings 0 < ΔT (K) < 710. The solid curve is limited by the maximum speed of the phase field propagation (dash-dotted line) which is V = Vϕ = 185 (m s−1) for growing Ni crystals.

Appendix B. Kinetic phase field model

Traditional phase field models based on the hypothesis of local thermodynamic equilibrium predict linear behaviour for the velocity, V ∝ΔG, describe data of atomistic simulations at relatively small interface velocity and clearly disagree with these data for large values of driving forces [81]. Owing to today's experimentally reachable large driving forces and high growth velocities [8], a phase field model going beyond the hypothesis of local equilibrium has been formulated [82,83] which predicts linear behaviour at small driving forces and a nonlinear behaviour at large driving forces [58,84]. We reformulate and further develop this model to compare its predictions with the data of atomistic simulations exhibiting a maximum in the velocity/undercooling relationship.

In the present derivation, the so-called ‘kinetic energy approach’ [82,85] as defined in [86] is used. This model describes non-equilibrium effects appearing in rapid solidification, in particular, non-equilibrium solute trapping with a transition to diffussionless crystal growth kinetics [87].

Consider a binary mixture consisting of solvent and solute undergoing a phase transition, particularly solidification/melting, from the undercooled/overheated state, for which the free energy is described by

G=v0[εϕ22|ϕ|2+G(T,C,ϕ,ϕt)]dv0, B 1

where v0 is the volume of the mixture, ε2ϕ is the gradient energy coefficient related to the interface energy γ, ϕ is the phase field variable defined as ϕ = 0 in the liquid phase and ϕ = 1 in the solid phase, ∂ϕ/∂t is the gradient flow for the phase field and C is the solute concentration. Introducing the phase field ϕ and gradient flow ∂ϕ/∂t as independent thermodynamic variables in the Gibbs potential, G(T, C, ϕ, ∂ϕ/∂t) can be considered to be fully analogous to Newtonian mechanics where the initial position and velocity of a particle must be specified to determine their evolution and velocity. Indeed, if inertial effects are sufficiently low in comparison with dissipative effects during phase field propagation, ∂ϕ/∂t will be directly determined by a dynamical equation in terms of ϕ and its gradient. Otherwise, ϕ and ∂ϕ/∂t will be independent and an equation for ∂2ϕ/∂t2 must be found [82].

The Gibbs potential G(T, C, ϕ, ∂ϕ/∂t) is given by [88]

G(T,C,ϕ)=Geq(T,C,ϕ)+Gneq(T,ϕt), B 2

with the local equilibrium contribution

Geq(T,C,ϕ)=[1p(ϕ)]Gl(T,C)+p(ϕ)Gs(T,C)+Wϕ(T,C)g(ϕ), B 3

and the local non-equilibrium contribution

Gneq(T,ϕt)=αϕ(T)2(ϕt)2. B 4

The contributions equations (B 3) and (B 4) include the interpolation function p(ϕ) and the double-well function g(ϕ) defined by [89]

p(ϕ)=(32ϕ)ϕ2andg(ϕ)=(1ϕ)2ϕ2, B 5

the barrier Wϕ(T, C) between the phases and the phenomenological coefficients αϕ(T) being proportional to the relaxation time τϕ of the gradient flow ∂ϕ/∂t. The local non-equilibrium contribution equation (B 4) can be considered a kinetic energy term added to the Gibbs potential in the traditional phase field theory. Attempts to introduce such a kinetic term was made earlier in the theory of adsorption of sound in liquids [90] or in the theory of kink propagation [91]. Using the formalism of EIT [32], the gradient flow ∂ϕ/∂t is introduced as an independent thermodynamic variable that yields the kinetic energy equation (B 4) naturally [92].

A stable evolution of the entire system is given by the Lyapunov condition of a non-positive change of the total Gibbs free energy with time. For the functional equation (B 1), this condition is given by the non-strict inequality [88]

dGdt=ddtv0[εϕ22|ϕ|2+G(T,C,ϕ,ϕt)]dv00, B 6

from which one finds the following phase field equation [84]

τϕ2ϕt2+ϕt=Dϕ2ϕMϕ[ΔGdp(ϕ)dϕ+Wϕ(T,C)dg(ϕ)dϕ], B 7

where the Gibbs free energy difference ΔG is given by

ΔG=Gs(T,C)Gl(T,C){<0,solidification,>0,melting. B 8

The variety of transformations are obtained for different temperature approximations [57], dilute mixtures [93] and thermodynamic functions from data bases [94,95]. The diffusion coefficient of the phase field

Dϕ(T)=εϕ2Mϕ(T) B 9

essentially depends on the temperature, if the phase transition is considered in a wide temperature range [58].

The hyperbolic equation (B 7) describes the relaxation of two variables: relaxation of the slow ϕ-field by the first time derivative and relaxation of the gradient flow ∂ϕ/∂t by the second time derivative. In this sense, due to introducing the relaxation of ∂ϕ/∂t, equation (B 7) describes the evolution of the local non-equilibrium system. In a general case, this equation should be solved numerically using specially developed algorithms [88,96,97]. Further, we show the importance of using the gradient flow relaxation and, consequently, the role of local non-equilibrium in rapid crystal growth kinetics using a travelling wave analytical solution.

In equilibrium, ΔG = 0, Gs(T, C) = Gl(T, C), equation (B 7) allows a single dimensionally steady solution at ∂ϕ/∂t = 0 along the spatial x-axis:

ϕ(x)=12[1tanh(xδI)]withinthediffuseinterfaceδI2<x<+δI2, B 10

with the stationary width of diffuse interface

δI=2εϕWϕ(T,C), B 11

and with the boundary conditions ϕ = 1 as x ≤ δI and ϕ = 0 as xδI. Then, the surface energy γ of the crystal/liquid interface is given by

γ=+[εϕ22(dϕdx)2+Wϕ(T,C)g(ϕ)]dx=εϕWϕ(T,C)32=δIWϕ(T,C)6. B 12

In the dynamics, ΔG = Gs(T, C) − Gl(T, C)≠0, equation (B 7) has one dimensional travelling wave solution [84,87]

ϕ(x,t)=12[1tanh(xVt)], B 13

with the boundary conditions ϕ → 1 as x − V t → − ∞ and ϕ → 0 as x − V t → + ∞, with the constant velocity V limited by Vϕ as a maximum speed of phase field propagation

V=μkΔHfΔG1+(μkΔG/ΔHfVϕ)2, B 14

the velocity-corrected effective interface thickness

=2δI3[1V2Vϕ2]1/2, B 15

and the mobility Mϕ related to the interface kinetic coefficient μk as

μk=18γWϕ(T,C)Mϕ(T)ΔHf. B 16

First, the particular solution equation (B 13) with the hyperbolic tangent function follows from the general set of analytical solutions of Allen–Cahn-type equations [98,99] which is given in the present model by equation (B 7). Second, the interface velocity, V , cannot exceed the maximum speed of disturbance propagation in the phase field, because the phase field itself dictates the interface shape and its velocity, i.e. V < Vϕ in the solutions equations (B 13)–(B 15). Third, with regard to the effective interface thickness (B 15), one has to note two important issues:

  • (i)

    with increasing interface velocity, ℓ should become smaller than the constant interface width δI that has been chosen as a reference for the interface thickness in equilibrium state, equation (B 11);

  • (ii)

    within the limit VVϕ, one gets ℓ → 0, therefore, the phase field variation will be steeper with the tendency to build up a sharp interface as the velocity increases.

Using equations (B 12) and (B 16), the crystal growth velocity (B 14) can be re-written as

V=Dϕ(ΔTK)ΔG(ΔTK)γ1+[(Dϕ(ΔTK)/γVϕ(ΔTK))ΔG(ΔTK)]2, B 17

where the diffusive motion of the phase field is dictated by the temperature-dependent coefficient equation (B 9). At small and moderate driving forces, DϕΔG/γVϕ, the interface velocity is linearly proportional to the difference of the free energy, i.e. V ∝ΔG. At large driving forces, when DϕΔG/γ is of the order of Vϕ, the square root 1+[DϕΔG/(γVϕ)]2 in equation (B 17) causes nonlinearity of the velocity. This square root appears in equation (B 17) due to taking into account the local non-equilibrium effect in the form of the relaxation of the gradient flow, ∂ϕ/∂t, which results in the second derivative ∂2ϕ/∂t2 in the dynamic equation (B 7).

The quantitative comparison, figure 5a, shows that in the range of small undercoolings, 0 < ΔTK(K) < 50, and superheating, −55 < ΔTK(K) < 0, equation (B 17) predicts linear kinetics for growing/melting Ni crystals, V ∝ΔTK. This is well modelled by the molecular dynamics simulations in [78] and is in confluence with the kinetic behaviour described by the diffusion limited theory, equations (A 1)–(A 5) with ‘M=DLT’.

For the large range of undercoolings, 0 < ΔTK (K) < 710, figure 5b shows that equation (B 17) predicts a gradual deviation of the velocity V from the linear behaviour as the undercooling increases. Such behaviour in crystal growth kinetics has been found by molecular dynamics simulations of elemental systems [80,100], qualitatively confirmed in [84] and quantitatively confirmed in [53]. These qualitative and quantitative deviations are due to the existence of the square root 1+[DϕΔG/(γVϕ)]2 in equation (B 17). Because this square root appears as a consequence of the local relaxation to equilibrium, the quantitative comparison shows the importance and usefulness of the local non-equilibrium effect in the dynamics of a phase field at large driving forces.

Footnotes

1

In its implicit form, the kinetic liquidus (2.19) depends also on the velocity-dependent function kv(V), which is the solute partition coefficient, the coefficient of chemical segregation or, in other words, the atoms distribution coefficient at the interface, see also [10,12,13,15,17].

5. Conclusion

The thermodynamic driving force for rapid solidification that takes into account both the difference in chemical potential and the relaxation of stresses has been derived and analysed. The model for rapid dendritic growth including the transition to diffusionless solidification has been tested and quantified. In particular, we compared the influence of the kinetic liquidus derived for the binary diluted and concentrated mixtures on the growth kinetics of alloy dendrites. On the basis of this model, the description of dendritic solidification in glass-forming alloys of alloy systems with congruent melting phases is possible. The solidification kinetics of this class of alloys is well described by the theory which includes local non-equilibrium effects in the form of relaxation of the gradient flow in the phase field. Good comparison with experimental data confirms the initial theoretical assumption about the predominant influence of local non-equilibrium effects in solidification under large driving forces.

Data accessibility

This article has no additional data.

Authors' contributions

All authors contributed equally to the present review paper.

Competing interests

The authors declare that they have no competing interests.

Funding

This work was supported by the Russian Science Foundation (grant no. 16-11-10095) and the German Space Center Space Management under contract no. 50WM1541.

References

  • 1.Lavernia EJ, Srivatsan TS. 2010. The rapid solidification processing of materials: science, principles, technology, advances, and applications. J. Mater. Sci. 45, 287–325. ( 10.1007/s10853-009-3995-5) [DOI] [Google Scholar]
  • 2.Herlach DM, Matson DM. 2012. Solidification of containerless undercooled melts. Weinheim, Germany: Wiley-VCH. [Google Scholar]
  • 3.Jones H. 1973. Splat cooling and metastable phases. Rep. Prog. Phys. 36, 1425–1497. ( 10.1088/0034-4885/36/11/002) [DOI] [Google Scholar]
  • 4.Miroshnichenko IS. 1982. Quenching from the liquid state. Moscow, Russia: Metallurgia. [Google Scholar]
  • 5.Feuerbacher B. 1989. Phase formation in metastable solidification of metals. Mat. Sci. Eng. Rep. 40, 1–40. ( 10.1016/S0920-2307(89)80008-8) [DOI] [Google Scholar]
  • 6.Herlach DM, Cochrane RF, Egry I, Fecht HJ, Greer AL. 1993. Containerless processing in the study of metallic melts and their solidification. Inter. Mater. Rev. 38, 273–347. ( 10.1179/095066093790326267) [DOI] [Google Scholar]
  • 7.Herlach D. 1994. Non-equilibrium solidification of undercooled metallic melts. Mat. Sci. Eng. Rep. 12, 177–272. ( 10.1016/0927-796x(94)90011-6) [DOI] [Google Scholar]
  • 8.Herlach DM, Galenko PK, Holland-Moritz D. 2007. Metastable solids from undercooled melts. Amsterdam, The Netherlands: Elsevier. [Google Scholar]
  • 9.Chernov AA. 1984. Modern crystallography III. Crystal growth. Berlin, Germany: Springer. [Google Scholar]
  • 10.Baker JC, Cahn JW. 1969. Solute trapping by rapid solidification. Acta Metal. 17, 575–578. ( 10.1016/0001-6160(69)90116-3) [DOI] [Google Scholar]
  • 11.Baker JC, Cahn JW. 1971. Thermodynamics of solidification, ch. 2. In Solidification (eds TJ Hughel, GF Bolling), pp. 23–58. OH, US: ASM Metals Park. [Google Scholar]
  • 12.Boettinger WJ, Coriell SR. 1986. Microstructure formation in rapidly solidified alloys. In Science and technology of the undercooled melting (eds PR Sahm, H Jones, CM Adam). Dordrecht, The Netherlands: Martinus Nijhoâ. [Google Scholar]
  • 13.Kurz W, Fisher DJ. 1989. Fundamentals of solidifification, 4th edn Aedermannsdorf, Switzerland: Trans Tech. [Google Scholar]
  • 14.Aziz MJ, Kaplan T. 1988. Continuous growth model for interface motion during alloy solidification. Acta Metal. 36, 2335–2347. ( 10.1016/0001-6160(88)90333-1) [DOI] [Google Scholar]
  • 15.Biloni H, Boettinger WJ. 1996. Solidification. In Physical metallurgy (eds RW Cahn, P Haasen), ch. 8, vol. I. Amsterdam, The Netherlands: Elsevier. [Google Scholar]
  • 16.Hillert M, Rettenmayr M. 2003. Deviation from local equilibrium at migrating phase interfaces. Acta Mater. 51, 2803–2809. ( 10.1016/S1359-6454(03)00085-5) [DOI] [Google Scholar]
  • 17.Galenko P. 2002. Extended thermodynamical analysis of a motion of the solid-liquid interface in a rapidly solidifying alloy. Phys. Rev. B 65, 144103-1–144103-11. ( 10.1103/PhysRevB.65.144103) [DOI] [Google Scholar]
  • 18.Galenko PK, Jou D. 2019. Rapid solidification as non-ergodic phenomenon. Phys. Rep (Unpublished manuscript) [Google Scholar]
  • 19.Galenko PK, Danilov DA. 1997. Local nonequilibrium effect on rapid dendritic growth in a binary alloy melt. Phys. Lett. A 235, 271–280. ( 10.1016/s0375-9601(97)00562-8) [DOI] [Google Scholar]
  • 20.Galenko PK, Danilov DA. 1999. Model for free dendritic alloy growth under interfacial and bulk phase nonequilibrium conditions. J. Cryst. Growth 197, 992–1002. ( 10.1016/s0022-0248(98)00977-4) [DOI] [Google Scholar]
  • 21.Galenko PK, Reutzel S, Herlach DM, Danilov D, Nestler B. 2007. Modelling of dendritic solidification in undercooled dilute NiZr melts. Acta Mater. 55, 6834–6842. ( 10.1016/j.actamat.2007.08.038) [DOI] [Google Scholar]
  • 22.Hartmann H, Galenko PK, Holland-Moritz D, Kolbe M, Herlach DM, Shuleshova O. 2008. Nonequilibrium solidification in undercooled Ti45Al55 melts. J. Appl. Phys. 103, 073509-1–073509-9. ( 10.1063/1.2903920) [DOI] [Google Scholar]
  • 23.Galenko PK, Reutzel S, Herlach DM, Fries SG, Steinbach I, Apel M. 2009. Dendritic solidification in undercooled Ni-Zr-Al melts: experiments and modeling. Acta Mater. 57, 6166–6175. ( 10.1016/j.actamat.2009.08.043) [DOI] [Google Scholar]
  • 24.Hartmann H, Holland-Moritz D, Galenko PK, Herlach DM. 2009. Evidence of the transition from ordered to disordered growth during rapid solidification of an intermetallic phase. Europhys. Lett. 87, 40007-1–40007-6. ( 10.1209/0295-5075/87/40007) [DOI] [Google Scholar]
  • 25.Galenko PK, Nizovtseva IG, Reuther K, Rettenmayr M. 2018. Kinetic transition in the order–disorder transformation at a solid/liquid interface. Phil. Trans. R. Soc. A 376, 20170207 ( 10.1098/rsta.2017.0207) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Skripov VP, Koverda VP. 1984. Spontaneous crystallization of supercooled liquids. Moscow, Russia: Nauka. [Google Scholar]
  • 27.Wang Q, Wang LM, Ma MZ, Binder S, Volkmann T, Herlach DM, Wang JS, Xue QG, Tian YJ, Liu RP. 2011. Diffusion-controlled crystal growth in deeply undercooled melt on approaching the glass transition. Phys. Rev. B 83, 014202-1–014202-5. ( 10.1103/PhysRevB.83.014202) [DOI] [Google Scholar]
  • 28.Orava J, Greer AL. 2014. Fast and slow crystal growth kinetics in glass-forming melts. J. Chem. Phys. 140, 214504 ( 10.1063/1.4880959) [DOI] [PubMed] [Google Scholar]
  • 29.Wang H, Herlach DM, Liu RP. 2014. Dendrite growth in Cu50Zr50 glass-forming melts, thermodynamics vs. kinetics. Europhys. Lett. 105, 36001 ( 10.1209/0295-5075/105/36001) [DOI] [Google Scholar]
  • 30.Boettinger WJ, Coriell SR, Sekerka RF. 1984. Mechanisms of microsegregation-free solidification. Mat. Sci. Eng. 65, 27–36. ( 10.1016/0025-5416(84)90196-4) [DOI] [Google Scholar]
  • 31.Jou D, Casas-Vazquez J, Lebon G. 1999. Extended irreversible thermodynamics revisited (1988–98). Rep. Prog. Phys. 62, 1035 ( 10.1088/0034-4885/62/7/201) [DOI] [Google Scholar]
  • 32.Jou D, Casas-Vazquez J, Lebon G. 2010. Extended irreversible thermodynamics, 4th edn Berlin, Germany: Springer. [Google Scholar]
  • 33.Galenko PK, Danilov DA, Alexandrov DV. 2015. Solute redistribution around crystal shapes growing under hyperbolic mass transport. Int. J. Heat. Mass Trans. 89, 1054–1060. ( 10.1016/j.ijheatmasstransfer.2015.06.011) [DOI] [Google Scholar]
  • 34.Galenko PK, Herlach DM. 2006. Suppression of eutectic precipitation in rapid solidification of a binary system. In Spring-Meeting of the German Physical Society (DPG). Proc. of the DPG-Conf., Dresden, Germany, 27–31 March, p. 563. Dresden, Germany: Deutsche Physikalische Gesellschaft. [DOI] [PubMed]
  • 35.Galenko PK, Jou D. 2010. Thermodynamics and kinetics of fast spinodal decomposition controlled by diffusion in the presence of fluctuations. Solid-solid phase transformations in inorganic materials. Proc. of PTM-Conf. Avignon, France, 6–11 June, p. 37. Zürich-Dürnten, Switzerland: Trans Tech Publications.
  • 36.Masato D, Sorgato M, Lucchetta G. 2017. Prototyping and modeling of the centrifugal casting process for paraffin waxes. Mater. Manufact. Proc. 32, 1823–1830. ( 10.1080/10426914.2017.1317791) [DOI] [Google Scholar]
  • 37.Alexandrov DV, Galenko PK. 2013. Selection criterion for the growing dendritic tip at the inner core boundary. J. Phys. A: Math. Theor. 46, 195101-1–195101-12. ( 10.1088/1751-8113/46/19/195101) [DOI] [Google Scholar]
  • 38.Zhuravlev VA. 2006. Solidification and crystallization of alloys with hetero-trasformations. Izhevsk, Russia: Regular and Chaotic Dynamics. [Google Scholar]
  • 39.Glicksman ME. 2011. Principles of solidification: an introduction to modern casting and crystal growth concepts. New York, NY: Springer. [Google Scholar]
  • 40.Alexandrov DV, Danilov DA, Galenko PK. 2016. Selection criterion of a stable dendrite growth in rapid solidification. J. Heat Mass Trans. 101, 789–799. ( 10.1016/j.ijheatmasstransfer.2016.05.085) [DOI] [Google Scholar]
  • 41.Alexandrov DV, Galenko PK. 2017. Selected mode for rapidly growing needle-like dendrite controlled by heat and mass transport. Acta Mater. 137, 64–70. ( 10.1016/j.actamat.2017.07.022) [DOI] [Google Scholar]
  • 42.Barbieri A, Langer JS. 1989. Predictions of dendritic growth rates in the linearized solvability theory. Phys. Rev. A 39, 5314–5325. ( 10.1103/PhysRevA.39.5314) [DOI] [PubMed] [Google Scholar]
  • 43.Amar MB, Pelcé P. 1989. Impurity effect on dendritic growth. Phys. Rev. A 39, 4263–4269. ( 10.1103/PhysRevA.39.4263) [DOI] [PubMed] [Google Scholar]
  • 44.Brener E, Mel'nikov VI. 1990. Two-dimensional dendritic growth at arbitrary Peclet number. J. Phys. Fran. 51, 157–166. ( 10.1051/jphys:01990005102015700) [DOI] [Google Scholar]
  • 45.Müller-Krumbhaar H, Abel T, Brener E, Hartmann M, Eissfeldt N, Temkin D. 2002. Growth-morphologies in solidification and hydrodynamics. JSME Inter. J. B 45, 129–132. ( 10.1299/jsmeb.45.129) [DOI] [Google Scholar]
  • 46.Trivedi R, Kurz W. 1986. Morphological stability of a planar interface under rapid solidification conditions. Acta Metal. 34, 1663–1670. ( 10.1016/0001-6160(86)90112-4) [DOI] [Google Scholar]
  • 47.Trivedi R, Kurz W. 1994. Dendritic growth. Inter. Mat. Rev. 39, 49–74. ( 10.1179/imr.1994.39.2.49) [DOI] [Google Scholar]
  • 48.Alexandrov DV, Galenko PK. 2017. Selected mode of dendritic growth with n-fold symmetry in the presence of a forced flow. Europhys. Lett. 119, 1600-1–1600-17. ( 10.1209/0295-5075/119/16001) [DOI] [Google Scholar]
  • 49.Alexandrov DV, Galenko PK, Toropova LV. 2018. Thermo-solutal and kinetic modes of stable dendritic growth with different symmetries of crystalline anisotropy in the presence of convection. Phil. Trans. R. Soc. A 376, 20170215-1 ( 10.1098/rsta.2017.0215) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Binder S, Galenko PK, Herlach DM. 2014. The effect of fluid flow on the solidification of Ni2B from the undercooled melt. J. Appl. Phys. 115, 053511-1–053511-11. ( 10.1063/1.4864151) [DOI] [Google Scholar]
  • 51.Galenko PK, Danilov DA. 2000. Steady-state shapes of growing crystals in the field of local nonequilibrium diffusion. Phys. Lett. A 272, 207–217. ( 10.1016/S0375-9601(00)00417-5) [DOI] [Google Scholar]
  • 52.Chernov AA. 1968. Kinetic phase transitions. Soviet Phys. JETP 26, 1182–1190. [Google Scholar]
  • 53.Salhoumi A, Galenko PK. 2017. Analysis of interface kinetics: solutions of the Gibbs-Thomson-type equation and of the kinetic rate theory. IOP Conf. Ser.: Mat. Sc. Eng. 192, 12014 ( 10.1088/1757-899X/192/1/012014) [DOI] [Google Scholar]
  • 54.Tang C, Harrowell P. 2013. Anomalously slow crystal growth of the glass-forming alloy CuZr. Nat. Mat. 12, 507 ( 10.1038/NMAT3631) [DOI] [PubMed] [Google Scholar]
  • 55.Kerrache A, Horbach J, Binder K. 2008. Molecular-dynamics computer simulation of crystal growth and melting in Al50Ni50. Europhys. Lett. 81, 58001 ( 10.1209/0295-5075/81/58001) [DOI] [Google Scholar]
  • 56.Chan WL, Averback RS, Cahill DG, Ashkenazy Y. 2009. Solidification velocities in deeply undercooled silver. Phys. Rev. Lett. 102, 095701 ( 10.1103/PhysRevLett.102.095701) [DOI] [PubMed] [Google Scholar]
  • 57.Thompson CV, Spaepen F. 1979. On the approximation of the free energy change on crystallization. Acta Metal. 27, 1855–1859. [Google Scholar]
  • 58.Novokreshchenova AA, Lebedev VG. 2017. Determining the phase-field mobility of pure nickel based on molecular dynamics data. Tech. Phys. 62, 642–644. ( 10.1134/S1063784217040181) [DOI] [Google Scholar]
  • 59.Garca-Coln LS, Del Castillo LF, Goldstein P. 1989. Theoretical basis for the Vogel-Fulcher-Tammann equation. Phys. Rev. B 40, 7040–7044. ( 10.1103/PhysRevB.40.7040) [DOI] [PubMed] [Google Scholar]
  • 60.Galenko PK, Hanke R, Paul P, Koch S, Rettenmayr M, Gegner J, Herlach DM, Dreier W, Kharanzhevski EV. 2017. Solidification kinetics of a Cu-Zr alloy: ground-based and microgravity experiments. IOP Conf. Ser.: Mat. Sc. Eng. 192, 012028 ( 10.1088/1757-899X/192/1/012028) [DOI] [Google Scholar]
  • 61.Chan WL, Averback RS, Cahill DG, Ashkenazy Y. 2009. Solidification Velocities in Deeply Undercooled Silver. Phys. Rev. Lett. 102, 095701 ( 10.1103/PhysRevLett.102.095701) [DOI] [PubMed] [Google Scholar]
  • 62.Hanke R, Koch S, Kobold R, Paul P, Galenko PK, Herlach DM, Rettenmayr M. 2017. Parabolic flight experiment on TEMPUS 2016 within the MULTIPHAS-project: Measurement of solidification velocity in Zr50Cu30Ni20. Presentation on the Int. Working Group Meeting in German Aerospace Center (DLR-Bonn), Bonn-Bad-Godesberg, Germany, 15–16 March. Bonn, Germany: Deutsches Zentrum für Luft- und Raumfahrt. [Google Scholar]
  • 63.Kobold R. 2016. Crystal growth in undercooled melts of glass forming Zr-based alloys. PhD thesis, Ruhr-Universitat Bochum, Germany, p. 170.
  • 64.Soklaski R, Nussinov Z, Markow Z, Kelton KF, Yang L. 2013. Connectivity of the icosahedral network and a dramatically growing static length scale in Cu-Zr binary metallic glasses. Phys. Rev. B 87, 184203 ( 10.1103/PhysRevB.87.184203) [DOI] [Google Scholar]
  • 65.Boettinger WJ, Aziz MJ. 1989. Theory for the trapping of disorder and solute in intermetallic phases by rapid solidification. Acta Metall. 37, 3379-3391 ( 10.1016/0001-6160(89)90210-1) [DOI] [Google Scholar]
  • 66.Barth M, Wei B, Herlach DM. 1995. Crystal growth in undercooled melts of the intermetallic compounds FeSi and CoSi. Phys. Rev. B 51, 3422–3428. ( 10.1103/PhysRevB.51.3422) [DOI] [PubMed] [Google Scholar]
  • 67.Greer AL, Assadi H. 1997. Rapid solidification of intermetallic compounds. Mater.Sci. Eng. A 226-228, 133–141. ( 10.1016/S0921-5093(97)80026-3) [DOI] [Google Scholar]
  • 68.Hartmann H, Holland-Moritz D, Galenko PK, Herlach DM. 2009. Evidence of the transition from ordered to disordered growth during rapid solidification of an intermetallic phase. Europhys. Lett. 87, 40 007–40 013. ( 10.1209/0295-5075/87/40007) [DOI] [Google Scholar]
  • 69.Yang C, Gao J. 2014. Dendritic growth kinetics and disorder trapping of the intermetallic compound Ni3Sn under a static magnetic field. J. Cryst. Growth 394, 24–27. ( 10.1016/j.jcrysgro.2014.02.015) [DOI] [Google Scholar]
  • 70.Holland-Moritz D, Yang F, Kordel T, Klein S, Kargl F, Gegner J, Hansen T, Bednarcik J, Kaban I, Shuleshova O, Mattern N, Meyer A. 2012. Does an icosahedral short-range order prevail in glass-forming Zr-Cu melts? Europhys. Lett. 100, 56 002–56 000. ( 10.1209/0295-5075/100/56002) [DOI] [Google Scholar]
  • 71.Yang F, Holland-Moritz D, Gegner J, Heintzmann P, Kargl F, Yuan CC, Simeoni GG, Meyer A. 2014. Atomic dynamics in binary Zr-Cu liquids. Europhys. Lett. 107, 46001 ( 10.1209/0295-5075/107/46001) [DOI] [Google Scholar]
  • 72.Gegner J, Shuleshova O, Kobold R, Holland-Moritz D, Herlach DM. 2013. In situ observation of the phase selection from the undercooled melt in Cu-Zr. J. Alloys Compd. 576, 232 ( 10.1016/j.jallcom.2013.04.035) [DOI] [Google Scholar]
  • 73.Christian JW. 1975. The theory of transformations in metals and alloys. Oxford, UK: Pergamon. [Google Scholar]
  • 74.Broughton JQ, Gilmer GH, Jackson KA. 1982. Crystallization rates of a Lennard-Jones liquid. Phys. Rev. Lett. 49, 1496 ( 10.1103/PhysRevLett.49.1496) [DOI] [Google Scholar]
  • 75.Wilson HA. 1900. On the velocity of solidification and viscosity of supercooled liquids. Phil. Mag. 50, 238. [Google Scholar]
  • 76.Frenkel J. 1946. Kinetic theory of solids. New York, NY: Oxford University Press. [Google Scholar]
  • 77.Ashkenazy Y, Averback RS. 2007. Atomic mechanisms controlling crystallization behaviour in metals at deep undercoolings. Europhys. Lett. 79, 26005-1–26005-6. ( 10.1209/0295-5075/79/26005) [DOI] [Google Scholar]
  • 78.Mendelev MI, Rahman MJ, Hoyt JJ, Asta M. 2010. Molecular-dynamics study of solid-liquid interface migration in fcc metals. Mod. Simul. Mat. Sc. Eng. 18, 074002-1–074002-18. ( 10.1088/0965-0393/18/7/074002) [DOI] [Google Scholar]
  • 79.Herlach DM, Simons D, Pichon PY. 2018. Crystal growth kinetics in undercooled melts of pure Ge, Si and Ge-Si alloys. Phil. Trans. R. Soc. A 376, 20170205-1–20170205-13. ( 10.1098/rsta.2017.0205) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 80.Hoyt JJ, Sadigh B, Asta M, Foiles SM. 1999. Kinetic phase field parameters for the Cu-Ni system derived from atomistic computations. Acta Mater. 47, 3181–3187. ( 10.1016/S1359-6454(99)00189-5) [DOI] [Google Scholar]
  • 81.Berghoff M, Selzer M, Nestler B. 2013. Phase-field simulations at the atomic scale in comparison to molecular dynamics. Sci. World J. 2013, 564272 ( 10.1155/2013/564272) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 82.Galenko P, Jou D. 2005. Diffuse-interface model for rapid phase transformations in nonequilibrium systems. Phys. Rev. E 71, 046125-1–046125-13. ( 10.1103/PhysRevE.71.046125) [DOI] [PubMed] [Google Scholar]
  • 83.Galenko P, Danilov D, Lebedev V. 2009. Phase-field-crystal and Swift-Hohenberg equations with fast dynamics. Phys. Rev. E 79, 051110-1–051110-11. ( 10.1103/PhysRevE.79.051110) [DOI] [PubMed] [Google Scholar]
  • 84.Salhoumi A, Galenko PK. 2016. Gibbs-Thomson condition for the rapidly moving interface in a binary system. Physica A 447, 161–171. ( 10.1016/j.physa.2015.12.042) [DOI] [Google Scholar]
  • 85.Galenko P, Jou D. 2009. Kinetic contribution to the fast spinodal decomposition controlled by diffusion. Physica A 388, 3113–3123. ( 10.1016/j.physa.2009.04.003) [DOI] [Google Scholar]
  • 86.Wang H, Galenko PK, Zhang X, Kuang W, Liu F, Herlach DM. 2015. Phase-field modeling of an abrupt disappearance of solute drag in rapid solidification. Acta Mater. 90, 282–291. ( 10.1016/j.actamat.2015.02.021) [DOI] [Google Scholar]
  • 87.Galenko PK, Abramova EV, Jou D, Danilov DA, Lebedev VG, Herlach DM. 2011. Solute trapping in rapid solidification of a binary dilute system: a phase-field study. Phys. Rev. E 84, 041143-1–041143-17. ( 10.1103/PhysRevE.84.041143) [DOI] [PubMed] [Google Scholar]
  • 88.Lebedev V, Sysoeva A, Galenko P. 2011. Unconditionally gradient-stable computational schemes in problems of fast phase transitions. Phys. Rev. E 83, 026705-1–026705-11. ( 10.1103/PhysRevE.83.026705) [DOI] [PubMed] [Google Scholar]
  • 89.Wang SL, Sekerka RF, Wheeler AA, Murray BT, Coriell SR, Braun RJ, McFadden GB. 1993. Thermodynamically-consistent phase-field models for solidification. Physica D 69, 189–200. ( 10.1016/0167-2789(93)90189-8) [DOI] [Google Scholar]
  • 90.Mandel'shtam LI, Leontovich MA. 1937. On the theory of sound absorption in fluids. Zh. Eksperim. i Theor. 7, 438–449. [Google Scholar]
  • 91.Chaikin PM, Lubensky TC. 1995. Principles of condensed matter physics. Cambridge, UK: Cambridge University Press. [Google Scholar]
  • 92.Danilov DA, Lebedev VG, Galenko PK. 2014. A grand potential approach to phase field modeling of rapid solidification. J. Non-Equilib. Thermodyn. 39, 93–111. ( 10.1515/jnetdy-2013-0032) [DOI] [Google Scholar]
  • 93.Echebarria B, Folch R, Karma A, Plapp M. 2004. Quantitative phase-field model of alloy solidification. Phys. Rev. E 70, 061604 ( 10.1103/PhysRevE.70.061604) [DOI] [PubMed] [Google Scholar]
  • 94.Dinsdale AT. 1991. SGTE data for pure elements. CALPHAD 15, 317–425. ( 10.1016/0364-5916(91)90030-N) [DOI] [Google Scholar]
  • 95.Samal S, Phanikumar G. 2015. Phase evolution in hypereutectic Al90Cu10−xNix (x = 0, 5) alloys. Trans. Indian Inst. Met. 68, 1221 ( 10.1007/s12666-015-0709-3) [DOI] [Google Scholar]
  • 96.Galenko PK, Gomez H, Kropotin NV, Elder KR. 2013. Unconditionally stable method and numerical solution of the hyperbolic phase-field crystal equation. Phys. Rev. E 88, 013310 ( 10.1103/PhysRevE.88.013310) [DOI] [PubMed] [Google Scholar]
  • 97.Bueno J, Starodumov I, Gomez H, Galenko PK, Alexandrov D. 2016. Three dimensional structures predicted by the modified phase field crystal equation. Comp. Mater. Sci. 111, 310 ( 10.1016/j.commatsci.2015.09.038) [DOI] [Google Scholar]
  • 98.Nizovtseva IG, Galenko PK, Alexandrov DV. 2016. The hyperbolic Allen-Cahn equation: exact solutions. J. Phys. A: Math. Theor. 49, 435201-1–435201-14. ( 10.1088/1751-8113/49/43/435201) [DOI] [Google Scholar]
  • 99.Nizovtseva IG, Galenko PK. 2018. Travelling-wave amplitudes as solutions of the phase-field crystal equation. Phil. Trans. R. Soc. A 376, 20170202-1–20170202-15. ( 10.1098/rsta.2017.0202) [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Hoyt JJ, Asta M, Karma A. 2003. Atomistic and continuum modeling of denritic solidification. Mat. Sci. Eng. R 41, 121–163. ( 10.1016/S0927-796X(03)00036-6) [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Data Availability Statement

This article has no additional data.


Articles from Philosophical transactions. Series A, Mathematical, physical, and engineering sciences are provided here courtesy of The Royal Society

RESOURCES