Вы находитесь на странице: 1из 15

International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

www.elsevier.com/locate/ijhmt

Average momentum equation for interdendritic flow


in a solidifying columnar mushy zone
P. Bousquet-Melou a, B. Goyeau b,*
, M. Quintard c, F. Fichot a, D. Gobin b

a
Institut de Protection et de S^
uret
e Nucleaire, C.E.A Cadarache, B^
at. 700, 13108 Saint Paul Lez Durance, France
b
Laboratoire FAST, Universit es Paris VI et Paris XI, UMR CNRS 7608, B^ at. 502, Campus Universitaire, 91405 Orsay Cedex, France
c
Institut de M
ecanique des Fluides de Toulouse, Avenue du Professeur Camille Soula, 31400 Toulouse, France
Received 20 November 2001

Abstract
This paper deals with the derivation of the macroscopic momentum transport equation in a non-homogeneous so-
lidifying columnar dendritic mushy zone using the method of volume averaging. One of the originalities of this study lies
in the derivation of an associated closure problem for the determination of the spatial evolution of the effective transport
properties in such a complex situation. In this analysis – where the phase change has been included at the different stages
of the derivation – all the terms arising from the averaging procedure (geometrical moments, phase interactions, in-
terfacial momentum transport due to phase change, porosity gradients, etc.) are systematically estimated and compared
on the basis of the characteristic length-scale constraints associated with the porous structures presenting evolving
heterogeneities. For dendritic structures with ‘‘moderate’’ (but not small) evolving heterogeneities, we show that phase
change and non local effects could hardly affect the determination of the permeability and inertia tensors. Finally, a
closed form of the macroscopic momentum equation is proposed and a discussion is presented about the need to
consider inertia terms and the second Brinkman correction (explicitly involving gradients of the liquid volume fraction)
in such non-homogeneous systems. Ó 2002 Elsevier Science Ltd. All rights reserved.

1. Introduction the quality of the product (homogeneity, reliability, as-


pect, etc.) can be drastically affected.
Fluid mechanics in a multi-component mixture dur- In most solidification models, the mushy zone is de-
ing solidification is a very complex problem due to the scribed as a porous medium (or equivalent continuum)
development of a two-phase columnar mushy zone. This and the momentum equation has been derived using
mushy zone is composed of solid dendrites and inter- different methods reviewed in detail by [3]. Two distinct
dendritic liquid (Fig. 1) and its evolution depends on approaches have been used to represent the coupling
several conditions such as temperature gradient, initial between the mushy and bulk liquid regions. In the multi-
concentration of the melt, cooling rate, etc. It is now domain approach, the Navier–Stokes equation is written
recognized as essential to give the most possible accurate in the fully melted region while the flow in the porous
description of transport phenomena in the mushy zone mushy layer is governed by Darcy’s law or one of its
since the interdendritic liquid flow, mainly induced by extensions, and appropriate boundary conditions are
thermosolutal natural convection [1], strongly influences written at the interface [4–6]. However, this two-domain
heat and mass transfer, and plays a key role on the method is not suitable for predicting irregular interface
micro- and macro-segregation [2]. Under these circum- shapes and, furthermore, interfacial boundary condi-
stances, the solidification rate, the microstructure and tions still remain a controversial subject of intense re-
search activity [7]. For these reasons, multi-domain
models have been replaced by the more convenient one-
*
Corresponding author. Tel.: +33-1-69-15-80-39. domain continuum model constituted of a single set of
E-mail address: goyeau@fast.u-psud.fr (B. Goyeau). equations describing the transport phenomena in the
0017-9310/02/$ - see front matter Ó 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 1 7 - 9 3 1 0 ( 0 2 ) 0 0 0 7 7 - 7
3652 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

Nomenclature

Abe area of entrances and exits for the b-phase r position vector, m
contained within the averaging volume, m2 r0 radius of the averaging volume, m
Abr b–r interface contained within the averaging t time, s
volume, m2 et time scale for ev b , s
Av specific area, m1 t time scale for hvb ib , s
b vector used to represent Peb when micro- tbr unit tangent vector at the b–r interface
scopic inertial effects are negligible, m1 vb velocity of the b-phase, m s1
B second-order tensor used to represent w e br V averaging volume, m3
when microscopic inertial effects are negli- Vb volume of the b-phase contained within the
gible averaging volume, m3
c vector used to represent the contribution of wbr velocity of the b–r interface, m s1
microscopic inertial effects to Peb , m1 wn averaged growth velocity, m s1
C second-order tensor used to represent the x position of the centroid of the averaging
contribution of microscopic inertial effects volume, m
to we br y position vector relative to the centroid of the
F inertia tensor averaging volume, m
g gravitational acceleration, m s2
K permeability tensor, m2 Greek symbols
‘i i ¼ 1; 2; 3, lattice vectors, m d 1  qr =qb , shrinkage parameter
‘b interdendritic length scale, m cb indicator function for the b-phase
L characteristic length for macroscopic quan- eb volume fraction of the b-phase
tities, m lb dynamic viscosity of the b-phase, Pa s
Lv characteristic length for hvb ib ; m qb density of the b-phase, kg m3
Le characteristic length for eb ; m hwb i superficial volume average of w in the b-
m vector used to represent Peb , m1 phase
m_ b melting rate, kg m3 s1 hwb ib intrinsic phase average of w in the b-phase
m_ r solidification rate, kg m3 s1 e
w spatial fluctuation of w in the b-phase
b
M second-order tensor used to represent ev b
nbr unit normal vector directed from the b- Subscripts
phase toward the r-phase b liquid phase
Pb pression in the b-phase, Pa r solid phase

whole domain (melt, mush and solid regions). In this


approach, quasi-steady approximations, remeshing or
coordinate mapping are not necessary anymore, and the
conservation equations can be numerically solved using
a fixed grid.
Two kinds of continuum models have been used to
derive the conservation equations in the context of so-
lidification: the classical mixture theory [8–10] and the
volume averaging method [11–13]. The former treats the
solid and liquid phases as a solid–liquid mixture to
which macroscopic properties are assigned in a purely
phenomenological manner. Conservation equations for
each phase are added to provide a set of mixture con-
servation equations and interactions between phases are
described using semi-empirical relationships. One of the
first complete solidification models for multi-component
systems using the mixture theory has been developed by
Fig. 1. Macroscopic dendritic mushy zone and an associated Bennon and Incropera [9] and later reassessed by Pres-
averaging volume. cott et al. [10]. This mathematical model has been ex-
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3653

tended by some authors [14–16] and have been exten- previously emphasized, phase interaction terms are esti-
sively used in numerous configurations [17–21]. mated by constitutive laws. Furthermore, besides its
The volume averaging technique considers a repre- strong anisotropy, the dendritic columnar region is char-
sentative elementary volume in the domain under study, acterized by non-uniformity of the macroscopic proper-
and the local conservation equations are integrated over ties such as the liquid volume fraction which continuously
this volume providing averaged macroscopic transport varies from zero in the solid to unity in the melt region.
equations valid in the whole domain [22,23]. Phase in- Although these continuous variations or evolving hetero-
teractions at the solid–liquid interface arise from the geneities [32] can obviously modify the convective flows
averaging process and are represented by interfacial area [33], they have been rarely explicitly included in the
integrals. This technique has been used first in solidifi- theoretical models. Finally, one of the most important
cation modeling by Beckermann and Viskanta [24] and limitation of interdendritic flow modelization comes
later extended and used by Ganesan and Poirier [11], Ni from the Kozeny–Carman relationship used to estimate
and Beckermann [12], Beckermann and Viskanta [25] the permeability from the liquid volume fraction. Several
and Schneider and Beckermann [26]. Due to the com- numerical and experimental attempts have been made to
plexity of the dendritic structures, phase interaction in- provide a better representation of the permeability,
tegral terms are not explicitly calculated in these models especially for liquid volume fractions greater than 0.7
and generally are also represented by semi-empirical [34–37] but more data are still necessary to provide a
laws. Recently, a multiphase/multiscale theory has pro- general description.
vided some progress in this issue, by describing finite In order to overcome the above limitations, this pa-
mass exchange in the solid and extradendritic liquid in per addresses a new derivation of the macroscopic mo-
an averaging volume [27]. The method of volume aver- mentum equation in a solidifying columnar dendritic
aging is often preferred due to its ability to include mi- mushy zone using the volume averaging method and
croscopic information at the macroscopic scale thus provides an associated closure problem for the charac-
improving solidification modeling. terization of the spatial evolution of the transport
Regarding the momentum equation, Prescott et al. properties in such heterogeneous structures. The concept
[10] have shown that under the same physical assump- of closure problem has been previously used in homo-
tions, continuum and volume-averaged approaches give geneous [38–40] or heterogeneous [37] porous media but
rise to equivalent macroscopic equations to describe the always in the absence of phase change. In the context of
interdendritic flow. Indeed, in both cases, the macro- solidification, the analysis becomes much more complex
scopic momentum conservation equation is represented and a full solution taking into account all the phenom-
by a modified Navier–Stokes equation including Darcy’s ena at the different scales is still out of reach. For this
term sometimes completed with a Forchheimer correc- reason, in order to quantitatively justify the necessary
tion term where porosity and the permeability are simplifications and to propose a closed form of the
generally related by the classical Kozeny–Carman rela- macroscopic momentum equation, all the micro- and
tionship. Let us note that, in the absence of phase change, macro-contributions to momentum transport due to
the derivation of Darcy’s law for homogeneous systems phase change and geometry are estimated and compared
has also been obtained using the homogenization theory on the basis of the characteristic length-scale constraints
[28,29]. associated with the dendritic-like porous structures
All comparisons between numerical simulations and presenting evolving heterogeneities.
experiments were found in qualitative agreement, but
important discrepancies still subsist between measured
and predicted fields [19,24,30,31]. These differences are
2. The procedure of volume averaging
mainly attributed to the weakness of the model as-
sumptions, such as local thermal and mass equilibria,
In order to derive the macroscopic momentum
uncertainty on physical properties and geometrical
equations for an incompressible flow of a binary mixture
characterization of the columnar dendritic mushy zone.
during solidification, we consider a columnar mushy zone
It is henceforth clear that one significant improvement in
and the local averaging volume V shown in Fig. 1, where
solidification modeling lies in the introduction of mi-
r0 is the radius of V and ‘b stands for the interdendritic
croscopic information both in the macroscopic conser-
characteristic length. All physical properties of the
vation equations and in the representation of the
mixture are assumed to be constant and the Boussinesq
effective transport coefficients such as permeability, dif-
approximation applies. The boundary-value problem
fusion–dispersion coefficient or effective conductivity.
describing the mass and momentum conservation within
Among the limitations of the previous derivations of
the averaging volume is given by:
the macroscopic momentum equation used in solidifi-
cation modeling, we note that dispersive fluxes and in- oqb
þ r  ðqb vb Þ ¼ 0 ð1Þ
ertial effects are generally intuitively neglected and, as ot
3654 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

o where Le is the characteristic length associated with the


ðq vb Þ þ r  ðqb vb vb Þ ¼ rPb þ lb r2 vb þ qb g
ot b ð2Þ macroscopic variations of porosity. For a ‘‘moderate’’
in the b-phase value of s (less than 4%), Le is not strongly greater than
r0 but the scales still remain distinct and the associated
qb nbr  ðvb  wbr Þ ¼ qr nbr  ðvr  wbr Þ at Abr ; ð3Þ length-scale constraint can be written
vb ¼ f ðr; tÞ at Abe ; ð4Þ ‘b < r0 < Le ; ð7Þ

where the boundary condition (3) represents the mass which also leads to
conservation equation at the solid–liquid interface Abr , ‘b Le : ð8Þ
and Abe in (4) is the area of entrances and exits of the
liquid phase in the macroscopic region V (Fig. 1). At this Under this condition, the averaging procedure can still
stage, this boundary condition is unknown and the be used and we will see in the next section that Eq. (8) is
discussion about the function f ðr; tÞ will take place in very useful for comparing of the order of magnitude of
the section about the closure problem. It should be the different terms during the averaging procedure. Fi-
emphasized here that the physical situation we want to nally, when evolving heterogeneities are important
deal with can be much more complicated than the (greater than 4%), scale separation is not respected and a
problem considered in this section. In particular, mass fixed averaging volume may not be adapted to provide
conservation of several species, and heat transfer could average properties. This situation still remains a chal-
be considered. However, the relaxation times associated lenge and an alternative theory could consist in the use
with these other mechanisms are in most situations of deforming averaging volume [41].
much larger than the relaxation time associated with the According to the estimation by Goyeau [37] the de-
viscous flow under consideration. Therefore, we will creasing rate of the geometry s for real dendritic struc-
neglect the possible coupling between the momentum tures observed experimentally during solidification of a
balance and these other transport mechanisms. 26 wt% solution of aqueous NH4 Cl [42] and succino-
nitrile–4 wt% acetone [43] is less than 4%. Therefore, we
2.1. Geometrical considerations will consider in this work columnar dendritic structures
with ‘‘moderate’’ value of s where the scale separation is
In homogeneous structures, one usually consider that described by Eqs. (7) and (8).
the averaging method is applicable for systems in which
the different length scales are constrained by [22,23]: 2.2. Averaged continuity equation

‘b r0 L; ð5Þ The definitions and theorems used in the following


development are summarized in Appendix A. Let us
where L is the macroscopic length scale of the system. consider the microscopic continuity equation of the solid
For a columnar dendritic-like porous structure pre- and liquid phases in the averaging volume V:
senting evolving heterogeneities the situation is much oqk
more complex since the averaged macroscopic proper- þ r  ðqk vk Þ ¼ 0; k ¼ b; r: ð9Þ
ot
ties are space-dependent but also depend on the size of
the averaging volume V. Indeed, in a mushy zone, the According to Prescott and Incropera [11] and Ganesan
liquid volume fraction (porosity) continuously varies and Poirier [12], we assume that, in the context of so-
from unity in the melt to zero in the solid region. In this lidification, the solid and liquid microscopic densities qr
case, the analysis depends on three length-scale con- and qb are uniform in V and that the solid and the liquid
straints which are functions of some geometrical pa- density variations are only significant on the macro-
rameter characteristic of the evolving heterogeneity. scopic scale. This assumption can be written:
This parameter describing the decreasing rate of the ge-
ometry, denoted s, will be defined as the average de- qk ¼ ck hqk ik ; k ¼ b; r; ð10Þ
crease of the thickness of the dendrite in the direction
where ck is the k-phase indicator function, given by
parallel to the primary dendritic arm [32]. First, when
definition (A.6). hqk ik represents the intrinsic volume
the evolving heterogeneities are small (typically
average density of phase k defined by Eq. (A.9). Using
s < 0:5%), the macroscopic properties are quasi-con-
the averaging theorems provided in Appendix A, the
stant whatever the size r0 of the averaging volume V and
averaged macroscopic mass conservation equations for
the scale separation is represented by the classical con-
the liquid and solid phases can be written as
straint
o
‘b r0 Le ; ð6Þ ðeb qb Þ þ r  ðeb qb hvb ib Þ ¼ m_ b ; ð11Þ
ot
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3655

o Averaged convective term:


ðer qr Þ þ r  ðer qr hvr ir Þ ¼ m_ r ; ð12Þ
ot  
r  ðqb vb vb Þ ¼ r  ðeb qb hvb vb ib Þ
Z
where m_ b and m_ r are, respectively, the melting and so- 1
lidification rates defined by: þ ðnbr  vb Þqb vb dA: ð21Þ
V Abr
Z
1
m_ b ¼  q nbr  ðvb  wbr Þ dA; ð13Þ Using the Gray decomposition [44]:
V Abr b
Z vb ¼ hvb ib þ ev b ; ð22Þ
1
m_ r ¼  q nrb  ðvr  wbr Þ dA; ð14Þ where ev b is the spatial velocity deviation, allows to write
V Abr r
the term hvb vb ib of Eq. (21) under the simplified form
where wbr is the velocity of the interface Abr and nbr is hvb vb ib ¼ hvb ib hvb ib þ hev b ev b ib ; ð23Þ
the unit normal vector pointing from the b-phase to the
r-phase. The microscopic mass balance at the solid– where hev b ev b ib is the liquid momentum dispersion. The
liquid interface given by Eq. (3) gives rise to the mac- simplified form (23) requires that the different length
roscopic mass balance scales verify [45]
 r 2 lb
m_ b þ m_ r ¼ 0: ð15Þ 0
: ð24Þ
L L
In a columnar mushy zone and in the absence of solid The above condition resulting from Eq. (5) is always
transport, the local solid velocity vr due to dilatation is satisfied for homogeneous porous media. This is not the
very small compared to the interface velocity wbr and is case for columnar dendritic porous structures presenting
neglected in this analysis. Therefore, Eq. (3) reduces to evolving heterogeneities, where scale separation depends
nbr  vb ¼ dnbr  wbr at Abr ; ð16Þ on the decreasing rate of the geometry s. Using sche-
matic dendritic structures Benihaddadene [13] has shown
where d is the volume change parameter, defined by: that (24) is verified only for very small values of s
(s 0:5%) and that for ‘‘moderate’’ values of s (less
q
d¼1 r: ð17Þ than 4%), Eq. (24) becomes
qb  r 2 l
0 b
< : ð25Þ
Using Eqs. (11), (12) and (15) leads to the averaged mass L L
conservation equation for the mixture
Under these circumstances, Eq. (23) remains valid and
o the left-hand side of Eq. (19) becomes
ðeb qb þ er qr Þ þ r  ðeb qb hvb ib Þ ¼ 0: ð18Þ  
ot o  
ðq vb Þ þ r  ðqb vb vb Þ
ot b
o
2.3. Averaged momentum equation ¼ ðeb qb hvb ib Þ þ r  ðeb qb hvb ib hvb ib Þ
ot
Z
1
In order to derive the macroscopic momentum þ r  ðeb qb hev b ev b ib Þ þ nbr  ðvb  wbr Þqb vb dA:
equation for the liquid phase, we consider the superficial V Abr
average of the Navier–Stokes equation (2): ð26Þ
  Let us consider now the right-hand side of Eq. (19):
o  
ðq vb Þ þ r  ðqb vb vb Þ Averaged pressure term:
ot b Z
        1
¼  rPb þ lb r2 vb þ qb g : ð19Þ  rPb ¼ rðeb hPb ib Þ  nbr Pb dA: ð27Þ
V Abr

Using Eq. (10) and the derivatives theorems (A.14) and If we use the Gray decomposition for pressure in the last
(A.15), each term of Eq. (19) can be developed. Let us term of Eq. (27), the area integral can be written:
Z Z
start with the left-hand side of Eq. (19): 1 1
nbr Pb jr dA ¼ nbr hPb ib jr dA
Averaged accumulation term: V Abr V Abr
  Z
o o 1
ðqb vb Þ ¼ ðeb qb hvb ib Þ þ nbr Peb jr dA: ð28Þ
ot ot V Abr
Z
1 Here, we are confronted to the fact that the first area
 ðnbr  wbr Þqb vb dA: ð20Þ
V Abr integral in the right-hand side of Eq. (28) contains the
3656 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665
Z Z
average quantity hPb ib jr which is evaluated at r instead lb lb
nbr  rvb jr dA ¼ nbr  rhvb ib jr dA
of the centroid x of the averaging volume (Fig. 1). Ac- V Abr V Abr
cording to Quintard and Whitaker [46], we can over- Z
lb
come this difficulty using Taylor series expansions of this þ nbr  rev b jr dA; ð36Þ
V Abr
quantity about the centroid x of the averaging volume
1 where the hvb ib jr in the first area integral of Eq. (36) has
hPb ib jr ¼ hPb ib jx þ y  rhPb ib jx þ yy : rrhPb ib jx þ    to be developed into Taylor series expansions about the
2
ð29Þ centroid x of the averaging volume. Since geometrical
terms (31) are negligible this integral takes the form
where y ¼ r  x. Substitution of expression (29) in (28) Z
yields lb
nbr  rhvb ib jr dA ’ lb reb  rhvb ib jx : ð37Þ
Z V Abr
1
nbr hPb ib jr dA
V Abr Therefore, using (36) and (37) in (35) gives rise to
Z ! Z !
1 1  
¼ b
nbr dA hPb i jx þ nbr y dA lb r2 vb ¼ eb lb r2 hvb ib þ lb reb  rhvb ib
V Abr V Abr Z
! lb
Z þ lb r2 eb hvb ib þ nbr  rev b dA
b 1 1 V Abr
 rhPb i jx þ nbr yy dA : rrhPb ib jx " #
2 V Abr Z
1
þ  ð30Þ þ lb r  nbr vb dA : ð38Þ
V Abr
Using schematic and real dendritic structures, it may be In the absence of phase change, the last area integral of
shown [32,37] that the geometrical moments of Eq. (30) (38) vanishes due to the no-slip and no-penetration
Z Z conditions (vb ¼ 0) at the solid–liquid interface and the
1 1
nbr y dA; nbr yy dA; . . . ð31Þ average of the viscous term (38) takes the classical form
V Abr V Abr
derived in [32].
Since we are dealing with a phase change problem,
are very small compared to unity and therefore can be
according to the mass conservation Eq. (16), the liquid
neglected in the analysis. Furthermore, using the aver-
velocity is a priori not zero at the solid–liquid interface,
aging theorem (A.14) with wb ¼ cb leads to
due to the volume change (d 6¼ 0) upon solidification. If
Z
1 we use again Gray’s decomposition in the last area in-
nbr dA ¼ reb ð32Þ tegral of (38), we obtain
V Abr
" Z #
and expression (30) becomes 1
r nbr vb jr dA
Z V Abr
1 " #
nbr hPb ib jr dA ’ reb hPb ib jx ¼ reb hPb ib : ð33Þ Z
V Abr 1 b
¼r nbr hvb i jr dA ð39Þ
V Abr
Finally the averaged pressure term takes the form: " Z #
Z 1
  1 þr nbr ev b jr dA
 rPb ¼ eb rhPb ib  nbr Peb dA: ð34Þ V Abr
V Abr

Averaged viscous term: and as previously, it is easy to show that


If the liquid viscosity is assumed to be constant, the " Z #
averaged viscous term is given by 1
r nbr hvb ib jr dA ’ reb hvb ib jx : ð40Þ
  V Abr
lb r2 vb ¼ lb r2  ðeb hvb ib Þ
Z !
1 Finally, the macroscopic viscous term can be written
þ lb r  nbr vb dA under the simplified form:
V Abr
Z Z
  lb
lb lb r2 vb ¼ eb lb r2 hvb ib þ nbr  rev b dA
þ nbr  rvb dA: ð35Þ V Abr
V Abr " #
Z
1
Using the Gray velocity decomposition (22) in the last þ lb r  nbr ev b dA : ð41Þ
V Abr
area integral of Eq. (35) provides
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3657

Averaged buoyancy term: interdendritic liquid velocity results from double-diffu-


  sive natural convection but also from shrinkage (d < 0)
qb g ¼ eb hqb ib g ¼ eb qb g: ð42Þ
or expansion (d > 0). This latter phenomena can con-
Finally, the different averaged terms (26), (34), (41) and tribute to morphological instability [47] and macroseg-
(42) provide the following non-closed averaged mo- regation [48]. Volume change can also play a key role on
mentum equation for the liquid flow through the mushy the formation of gas micropores due to strong pressure
zone: gradients [1,49,50] and may have a significant influence
on the velocity field at high solidification rates [3] or in
o
ðeb qb hvb ib Þ þ r  ðeb qb hvb ib hvb ib Þ þ r  ðeb qb hev b ev b ib Þ microgravity systems [51].
ot Z However, volume expansion or contraction generally
1
þ ðnbr  ðvb  wbr Þqb vb dA hardly influences the interdendritic flow when the vol-
V Abr ume-change parameter d is small, which is the case for
¼ eb rhPb ib þ eb lb r2 hvb ib þ eb qb g most metal alloys where jdj 6 10%. Furthermore, if
Z we consider that, in most solidification processes, the
1
þ nbr  ð Peb I þ lb rev b Þ dA growth velocity nbr  wbr is small compared to the aver-
V Abr
" Z # aged liquid velocity hvb ib induced by natural convection
1 we can estimate from Eq. (45) that
þ lb r  nbr ev b dA : ð43Þ
V Abr
jdnbr  wbr j jnbr  hvb ib j: ð46Þ
In order to develop a closed form of equations (11) and
Therefore, the boundary condition (45) reduces to
(43), a closure problem has to be written to express Peb
and ev b deviations in terms of the intrinsic averaged nbr  ev b ¼ nbr  hvb ib at Abr : ð47Þ
pressure and velocity hPb ib and hvb ib .
Moreover, the no-slip boundary condition is given by
3. Closure problem ~ ðvb Þ ¼ P
P ~ ðvr Þ at Abr ; ð48Þ

The derivation of the closure problem is very im- where P~ ðÞ is a projection operator on the tangent plane
portant since it allows to include microscopic aspects to the interface. Since vr ¼ 0, the above condition takes
such as the tortuosity of the microstructure or micro- the form:
scopic inertia effects in the macroscopic model. This can
P ~ ðhvb ib Þ at Abr :
~ ðev b Þ ¼ P ð49Þ
obviously contribute to significantly improve the geo-
metrical description of the mushy zone in terms of mac-
Therefore, Eqs. (47) and (49) provide
roscopic properties and therefore increase the quality of
the representation of the physical phenomena in solidi- ev b ¼ hvb ib at Abr ; ð50Þ
fication modelling.
which corresponds to a no-slip and no-penetration
condition (vb ¼ 0) at the solid–liquid interface. As it will
3.1. Continuity equation
be seen in the following analysis, the boundary condi-
tion (50) will play a key role in the closed form of Eqs.
Using Gray’s decomposition (22) in Eq. (1) and
(11), (43). Indeed, from (50) we deduce that the order of
subtracting the averaged equation (11) leads to the local
magnitude of the velocity deviation ev b can be estimated
continuity equation:
by:
oeb
eb qb r  ev b þ eb ev b  rqb ¼ qb þ qb reb  hvb ib  m_ b : ev b ¼ Oðhvb ib Þ: ð51Þ
ot
ð44Þ Under these circumstances, the continuity equation (44)
can be simplified on the basis of the order of magnitude
The associated boundary condition is also obtained us- estimates of the different terms of this equation. The
ing Eq. (22) in (3): three terms in (44) involving ev b or hvb ib can easily be
evaluated by:
nbr  ev b ¼ dnbr  wbr  nbr  hvb ib at Abr : ð45Þ !
hvb ib
According to the mass balance equation (16), the liquid eb qb r  ev b ¼ O eb qb ; ð52Þ
‘b
velocity at the solid–liquid interface Abr is non-zero
during phase change if the density of the two phases are  qb 
eb ev b  rqb ¼ O eb hvb ib ; ð53Þ
different (d 6¼ 0). Indeed, in solidification processes, the L
3658 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665
!
hvb ib oev b
qb reb  hvb ib ¼ O eb qb ð54Þ eb q b þ eb qb vb  rev b þ eb qb ev b  rhvb ib
Le ot
 r  ðeb qb hev b ev b ib Þ  m_ b hvb ib
Z
and according to the length scale constraint (8), we 1
 nbr  ðvb  wbr Þqb vb dA
conclude: V Abr
¼ eb r Peb þ eb lb r2 ev b
eb ev b  rqb ; qb reb  hvb ib eb qb r  ev b : ð55Þ Z
1
 nbr  ð Peb I þ lb rev b Þ dA
Furthermore, if wn represents the averaged interfacial V Abr
" Z #
growth velocity defined by:
1
Z  lb r  nbr ev b dA ; ð63Þ
1 V Abr
wn ¼ nbr  wbr dA ð56Þ
A Abr
ev b ¼ hvb ib at Abr : ð64Þ
the macroscopic mass balance (15), with vr ¼ 0, pro-
vides the following estimation of the melting rate m_ b : Eq. (63) is extremely complex, but as previously done for
 the continuity equation, it can be simplified by esti-
wn mating the order of magnitude of its different terms. On
m_ b ¼ m_ r qr Av wn ¼ O eb qr ; ð57Þ
‘b the basis of the constraints (8), (51) and (60) let us ex-
amine each term of Eq. (63).
where Av ¼ Abr =V is the specific area estimated by According to Quintard and Whitaker [52], if et and t
Carbonell and Whitaker [45]: represent the characteristic time variations of ev b and
 hvb ib , respectively, we can assume time scales separation:
eb
Av ¼ O : ð58Þ
‘b et t ð65Þ

Regarding the term involving the time derivative of eb in that leads to neglect oev b =ot when solving the coupled
Eqs. (44), (14) gives rise to: equations (43) and (63) . The convective and dispersive
Z  terms are easily estimated by:
oeb 1 wn !
¼ nbr  wbr dA ¼ Av wn ¼ O eb : ð59Þ
ot V Abr ‘b ðhvb ib Þ2
eb qb vb  rev b ¼ O eb qb ; ð66Þ
‘b
Let us recall that we have assumed that the growth !
velocity nbr  wbr is generally small compared to the b ðhvb ib Þ2
eb qb ev b  rhvb i ¼ O eb qb ; ð67Þ
average velocity hvb ib induced by natural convection. Lv
Therefore we have !
ðhvb ib Þ2
jwn j khvb ib k ð60Þ b
r  ðeb qb hev b ev b i ¼ O eb qb ; ð68Þ
L
and Eqs. (52), (57), (59) and (60) give rise to the com-
where L associated to macroscopic spatial variations, is
parison:
assumed to be such that L Lv . Using the length-scale
oeb constraint given by Eq. (8) leads to
qb ; m_ b eb qb r  ev b : ð61Þ
ot
r  ðeb qb hev b ev b ib eb qb ev b  rhvb ib eb qb vb  rev b :
Finally, Eq. (44) reduces to the simple form: ð69Þ
r  ev b ¼ 0: ð62Þ
The two other terms in the left-hand side of Eq. (63)
represent the momentum transfer due to phase change.
3.2. Momentum equation Since the liquid velocity at the solid–liquid interface
induced by volume change has been neglected (we recall
In order to derive the local momentum closure that boundary condition (64) has been derived on the
equation we introduce Gray’s decomposition for the basis of this assumption), the interfacial momentum
velocity and the pressure in (2) and we subtract the av- transfer (area integral) is therefore also negligible
eraged momentum equation (43). After straightforward Z
manipulations, the closure problem, including the sim- 1
qb nbr  ðvb  wbr Þvb dA ’ 0: ð70Þ
plified boundary condition (50), takes the form: V Abr
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3659
Z
Using the estimation of the solidification rate given by 1
eb lb r2 ev b  nbr  ð Peb I þ lb rev b Þ dA
Eq. (57), the second phase change term m_ b hvb ib in (63) is V Abr
estimated by: " Z #
1
!  lb r  nbr ev b dA : ð79Þ
b wn eb hvb ib V Abr
m_ b hvb i ¼ O qr ð71Þ
‘b
Finally, the closure momentum equation (63) reduces to
and according to (60) and (66), we have:
qb vb  rev b ¼ r Peb þ lb r2 ev b
m_ b hvb ib eb qb vb  rev b : ð72Þ Z
1
 nbr  ð Peb I þ lb rev b Þ dA: ð80Þ
V Abr
Finally, eb qb vb  rev b is the only significant term re-
maining on the left-hand side of (63).
Regarding the right-hand side of Eq. (63), the first 3.3. Local closure problem
three terms have already been estimated by [38] and
details can be found in this reference. All these terms We summarize the closure problem constituted by
have the same order of magnitude, for example the in- the equations derived in the previous section associated
terfacial momentum exchange is such that: to the interface condition at Abe :
Z !
1 hv i b r  ev b ¼ 0; ð81Þ
b
nbr  ð Peb I þ lb rev b Þ dA ¼ O eb lb 2 ; ð73Þ
V Abr ‘b qb vb  rev b ¼ r Peb þ lb r ev b 2
Z
1
where it is assumed that the contribution of the pressure  nbr  ð Peb I þ lb rev b Þ dA; ð82Þ
V Abr
Peb in the integral is not greater than the contribution of
the velocity term and where Eq. (73) takes into account ev b ¼ hvb ib at Abr ; ð83Þ
the boundary condition (51) at Abr . This latter condition |ffl{zffl}
source
at the Abr interface allows to write the last term of Eq.
(63) under the form ev b ¼ gðr; tÞ at Abe : ð84Þ
" Z #
1 In the above problem, the boundary condition given by
lb r  nbr ev b jr dA Eq. (84) is not known a priori but according to Eq. (51)
V Abr
" Z # we know that gðr; tÞ is of the order of hvb ib and we do
1 b know that this condition will influence the ev b -field only
¼ lb r  nbr hvb i jr dA ð74Þ in a region of thickness ‘b at the boundary of the aver-
V Abr
aging volume V [53]. That is the reason why, for ho-
and using Eq. (40) leads to mogeneous porous structures, the closure problem is
" Z # generally solved in a representative region (unit cell)
1 with periodicity conditions at the boundary Abe
lb r  nbr ev b jr dA ¼ lb r  ½reb hvb ib : ð75Þ
V Abr
Peb ðr þ ‘i Þ ¼ Peb ðrÞ; ev b ðr þ ‘i Þ ¼ ev b ðrÞ with i ¼ 1; 2; 3
Finally, (75) can be estimated by: ð85Þ
" Z #
1
lb r  nbr ev b jr dA assuming that variations of hvb ib can be neglected within
V Abr the unit cell. In Eq. (85), ‘i represents the lattice vectors.
¼ lb reb  rhvb ib þ lb r2 eb hvb ib ð76Þ In the case of columnar dendritic porous structures,
! ! macroscopic properties (porosity, permeability, etc.) are
hvb ib hvb ib continuously space-dependent (evolving heterogeneities)
¼ O eb lb ; O eb lb 2 ; ð77Þ
Lv Le Le and the periodicity condition in the direction parallel to
the primary dendrite arms seems to be inappropriate.
where the porosity gradients are explicitly present. At Actually, the type of boundary condition at Abe depends
this stage, it is important to recall that the length scale on the geometry of the dendritic structure and therefore
constraint used in the case of mushy zone with moderate on its decreasing rate s. Indeed, [32,37] have shown that
evolving heterogeneities is defined by: for small or moderate values of this parameter, the pe-
‘b Le ð78Þ riodicity condition could be a relevant approximation,
especially if we do remember that due to the small scale
and therefore, from (73), (77) and (78), we obtain of influence of (84), such a condition is mathematically
3660 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

weak in our problem. Finally, the local closure problem eb qb hvb ib  rhvb ib
for a columnar dendritic mushy zone presenting small or macroscopic convective flux;
moderate evolving heterogeneities is given by Eqs. (81)–
(83) and (85). Under this form, this problem takes the r  ½eb qb hvb ib  hMt Mib  hvb ib 
ð87Þ
same form as the closure problem given by [53] in the momentum dispersive flux;
absence of phase change. This important conclu-
 e2b lb K1  F  hvb ib
sion means that the contribution of ‘‘moderate’’ (but not
negligible) evolving heterogeneities and solid–liquid Forchheimer or microscopic inertial effects:
phase change in solidifying mushy zone hardly influence
the determination of the effective transport properties of Since the columnar dendritic mushy zone is character-
this region. However, we will see in Section 4, that these ized by continuous spatial variations of the macroscopic
contributions at the macroscopic scale are not neces- properties (liquid volume fraction and permeability),
sarily negligible. For the sake of clarity, the main steps and if the Reynolds number is greater than 1, two re-
of the treatment of the system (81)–(85) are briefly re- gions have to be considered.
called in Appendix B and all the details can be found First, for small or intermediate porosity values, the
in [53]. estimation:

ev b ¼ Oðhvb ib Þ ð88Þ
4. Closed form of the momentum equation
provides
Using the macroscopic mass conservation equation
(11) within Eq. (43) according to simplifications (70) and M ¼ Oð1Þ ð89Þ
Eqs. (76), (B.1) and (B.5), we obtain the non-conserva-
and therefore
tive closed form of the averaged momentum equation:
h i
r  eb qb hvb ib  hMt Mib  hvb ib
ohvb ib !
eb qb þ eb qb hvb ib  rhvb ib ðhvb ib Þ2
ot b b
eb qb hvb i  rhvb i ¼ O eb qb ; ð90Þ
h i L
þ r  eb qb hvb ib  hMt Mib  hvb ib þ m_ b hvb ib
while the estimation of the Forchheimer correction leads
¼ eb hPb ib þ eb lb r2 hvb ib þ lb reb  rhvb ib
to [53]
þ lb r2 eb hvb ib  e2b lb K1 :hvb ib  e2b lb K1 :F:hvb ib !
ðhvb ib Þ2
lb e2b K1  F  hvb ib ¼ O : ð91Þ
þ eb q b g ð86Þ ‘b

Hence, due to the length scale constraint (8) it is clear


Some terms of Eq. (86) need to be discussed in the
that the most important term describing inertial effects is
context of columnar dendritic-like porous structure since
given by the Forchheimer correction. This phenomenon
their influence depends on the position within the mushy
is attributed to the microscopic drag forces which are
zone.
much more important than the macroscopic inertial
effects [59,60].
4.1. Inertia terms The second case concerns the region near the tip of
the dendrites where the porosity tends towards unity and
It is well known that flows in porous media for small the notion of dispersion does not make sense any more.
pore-scale Reynolds numbers are governed by Darcy’s Since the permeability becomes infinite, the Forchheimer
law which supposes a linear relationship between the correction does not play any role, and the only re-
average pressure gradient and the seepage velocity hvb i maining inertia term is the usual convective term: this
[54,55]. The transition between Darcy’s regime where corresponds to the situation described by the classi-
viscous forces prevail and the inertial regime takes place cal Navier–Stokes equation. Therefore, whatever the
for seepage Reynolds numbers between 1 and 10, and porosity in the columnar dendritic mushy zone, the
the inertial effects are traditionaly represented by Forch- dispersive flux given by the quantity r  ½eb qb hvb ib 
heimer’s correction [56]. The problem of the correct hMt Mib  hvb ib  is negligible and inertia phenomena is
quadratic or cubic form of this term at small Reynolds described by the Forchheimer correction for small and
numbers is not discussed here [57,58]. The macroscopic moderate porosity while they will be represented by the
momentum Eq. (86) derived in this study involves three macroscopic convective term qb hvb ib  rhvb ib in the vi-
inertia terms which need to be compared: cinity of dendrite tips.
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3661

Finally, since the interfacial momentum transfer due problem and its resolution still remains a challenge.
to phase change (70) can be discarded, the final form of However, for moderate decreasing rates of the geometry,
the macroscopic mass and momentum conservation s, Goyeau et al. [32] have shown that even if the porosity
equations written in term of superficial velocity (filtra- gradients are present at the macroscopic scale, their
tion velocity) take the final form: contribution at the local scale can be neglected in the
o closure problem leading to an acceptable approximation
ðeb qb þ er qr Þ þ r  ðqb hvb iÞ ¼ 0; ð92Þ of the spatial variation of the permeability.
ot
All this analysis has been performed for laminar flow
o  
e1 ðqb hvb iÞ þ e1 1 regimes and needs to be extended to turbulent flows
b b r  eb qb hvb ihvb i
ot where fluctuations ev b involving dispersion phenomena
¼ hPb ib þ qb g þ lb e1 2
b r hvb i
can be much more important and therefore non negli-
1 gible [62].
 lb e1 1
b reb :rðeb hvb iÞ  lb K  hvb i
|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Second Brinkman correction term

 lb K1  F  hvb i: ð93Þ 5. Conclusion

A new derivation of the macroscopic momentum


4.2. Brinkman correction terms equation for an interdendritic flow through a solidifying
columnar mushy zone has been carried out using an
Under this form, Eq. (93) includes the second averaging procedure with local closure problem. In order
Brinkman correction term that explicitly involves po- to simplify such a complex problem, all the terms arising
rosity gradients. Using intrinsic velocity, similar terms from the averaging process (micro- and macro-contri-
(actually, three Brinkman correction terms instead of butions to momentum transport) have been estimated
two) have been derived by Whitaker [38] and Quintard and compared on the basis of the characteristic length
and Whitaker [46] in the absence of phase-change and scale constraints associated with dendritic-like porous
for homogeneous structures. They are generally ne- structures presenting evolving heterogeneities. For den-
glected due to the length-scale constraint (5) in such a dritic structures with ‘‘moderate’’ (but not small)
geometry. decreasing rate of the geometry, we have shown that
In the case of columnar dendritic-like porous struc- heterogeneities and phase change hardly influence the
tures the spatial porosity variations are generally pro- description of effective properties (permeability and in-
gressive and two situations have to be considered. First, ertia tensor) of the dendritic layer. Whatever the po-
for small decreasing rates of the geometry (quasi- rosity within the columnar mushy zone, the macroscopic
homogeneous porous structures) where scale separation dispersive flux has been found to be negligible compared
is given by Eq. (6), the two Brinkman terms in Eq. (93) to other inertial terms of the closed form of the momen-
do not contribute to momentum transport except near tum conservation equation. Therefore, inertial phenom-
the tip of the dendrites where the liquid volume fraction ena will be described by the Forchheimer correction for
is close to one and where the first Brinkman term is small and moderate porosity while they will be repre-
essential to account for viscous diffusion phenomena. In sented by the macroscopic convective term near the top
the case of a dendritic mushy zone presenting larger (but of the dendrites where porosity is close to one. Finally,
moderate) evolving heterogeneities, scale separation the closed form of the macroscopic momentum trans-
is represented by the length-scale constraint (7). In this port equation includes a second Brinkman correction
situation, we have shown (assuming Le Lv ) that the term explicitly involving porosity gradients. In some
order of magnitude of the Brinkman terms is not small cases depending on the decreasing rate of the geometry,
compared to the other terms and could have a non- this additional term could significantly modify the rep-
negligible influence on the momentum, heat and mass resentation of the momentum, heat and mass transfer in
transport within the mushy zone. Numerical simulation the mushy zone.
taking into account the porosity gradient are under de- This work is a first step towards the derivation of a
velopment in order to quantify their actual influence. ‘‘complete’’ macroscopic model for the simulation of
Finally, for rapid spatial changes of the porosity at the multi-component solidification systems and a similar
mush–fluid region interface, the second Brinkman cor- work concerning the heat and species transfer conser-
rection could be replaced by a jump condition [61]. vation equations and the determination of effective
Obviously, retaining the porosity gradients in the transport properties in the mushy zone (diffusion–dis-
macroscopic momentum equation would provide a persion, mass exchange and conductivity coefficients) is
much more complex closure problem where these terms presently under development. One of the application we
would be present as source terms for the deviation fields. are involved in, concerns the solidification of a molten
In this case, the closure problem becomes a non-local pool (corium) that can be found in a nuclear reactor
3662 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

vessel during a scenario of an hypothetical severe acci- The superficial volume average of wb is defined as:
dent [63]. Z
1
hwb ijx ¼ w ðx þ yb Þ dVy : ðA:8Þ
V Vb b
Acknowledgements
The intrinsic phase average of wb is defined as:
This work is supported by IPSN (Department DRS/ Z
1
SEMAR) under research contract number: 402048- hwb ib jx ¼ w ðx þ yb Þ dVy : ðA:9Þ
Vb Vb b
B002850. B.G. thanks Professor Stephen Whitaker
(Department of Chemical Engineering, University of The notation dVy is used to indicate that the integration
California at Davis, USA) for fruitful discussions and is done with respect to the variable y (Fig. 1). Note that
comments. Two of the authors (P.B-M. and F.F.) wish both types of averages are calculated at the centroid x of
to gratefully acknowledge Bruno Piar for his availability the averaging volume. However, for convenience, the
and his contribution to the discussion. point x is usually not specified, and the following no-
tations are adopted:
Appendix A. The volume-averaging technique hwb i ¼ hwb ijx ; ðA:10Þ

Here, we briefly recall the main features of the vol- hwb ib ¼ hwb ib jx : ðA:11Þ
ume averaging theory and all the details are provided in
[22,23,44,45,64], and many subsequent papers. These two kinds of averages are related by:
Let us consider a physical property w, continuous in
hwb i ¼ eb hwb ib : ðA:12Þ
each phase of a b–r system:

wb in b-phase; According to [44], wb can be decomposed in two parts:
w¼ ðA:1Þ
wr in r-phase: e ;
wb ¼ cb hwb ib þ w b ðA:13Þ
In (A.1), wb and wr are defined such as: e are the averaged and the deviation
where hwb ib and w b
wb ¼ 0 in r-phase; (fluctuation), respectively.
ðA:2Þ In order to average the microscopic conservation
wr ¼ 0 in b-phase:
equations in each phase, one uses the following theorems
The technique consists in averaging w in a representative [44] which relate the superficial volume average of the
elementary volume (REV) or averaging volume V of a spatial and temporal partial derivatives to the partial
two-phase system (Fig. 1). The following definitions are derivatives of the superficial volume average:
used: Z
1
hrwb i ¼ rhwb i þ nbr wb dA; ðA:14Þ
V : averaging volume; V Abr
Vb : volume of b-phase within V ; ðA:3Þ   Z
owb ohwb i 1
Vr : volume of r-phase within V : ¼  nbr  wbr wb dA; ðA:15Þ
ot ot V Abr
The phase volume fractions are defined as: where wbr is the local interface velocity and nbr is the
Vb Vr unit normal vector at the b–r interface, pointed from the
eb ¼ ; er ¼ ðA:4Þ
V V b-phase towards the r-phase (Fig. 1). For the demon-
stration of these theorems, one can refer to [65,22].
and related by
eb þ er ¼ 1: ðA:5Þ
Appendix B. Closure problem
A better definition of these quantities involves the phase
indicator defined as
 Due to the boundary condition (83) where hvb ib can
1 in b-phase; be considered as a source term in the boundary value
cb ¼ ðA:6Þ
0 in r-phase problem for Peb and ev b , the following solutions are
proposed for the deviation fields:
in the case of the b-phase. Using this definition, the b-
phase volume fraction is simply ev b ¼ M  hvb ib ; ðB:1Þ

eb ¼ hcb i: ðA:7Þ Peb ¼ lb m  hvb ib ; ðB:2Þ


P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3663

where m and M are vector and tensor fields, respectively. The solution of the above problem does not raise major
In order to separate linear and inertia effects [53] de- difficulties since the system given by Eqs. (B.13),(B.17)–
composed m and M in two parts: (B.19) is similar to the classical Stokes boundary value
m ¼ b þ c; ðB:3Þ problem where d and D would represent the pressure
and the velocity, respectively. The resolution provides
M¼BþC ðB:4Þ the D-field which is integrated to provide the perme-
ability using the average condition (B.17).
and the closed form of the interfacial momentum ex- The Forchheimer correction tensor F is calculated by
change can be written under the generic form solving Problem II, which depends both on the geometry
Z
1 of the porous structure but also on the intensity of the
nbr  ð Peb I þ lb rev b Þ dA
V Abr flow through the porous medium.

¼ e2b lb K1  hvb ib  e2b lb K1  F  hvb ib ; ðB:5Þ Problem II:


|fflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflffl} |fflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflffl}
Darcy term Inertia term
r  C ¼ 0; ðB:20Þ
where K is the permeability tensor that only depends on qb
the tortuosity of the porous structure and where F is the vb  rðB þ CÞ ¼ rc þ r2 C
lb Z
Forchheimer correction tensor which accounts for mi- 1
 nbr  ðIc þ rCÞ dA;
croscopic inertia effects [53]. These tensors can explicitly Vb Abr
be written under the form: ðB:21Þ
Z
1 C ¼ 0 at Abr ; ðB:22Þ
eb K1 ¼  nbr  ðIb þ rBÞ dA; ðB:6Þ
Vb Abr
Z cðr þ ‘i Þ ¼ cðrÞ; Cðr þ ‘i Þ ¼ CðrÞ at Abe ; ðB:23Þ
1
e2b K1  F ¼  nbr  ðIc þ rCÞ dA; ðB:7Þ hCib ¼ 0: ðB:24Þ
Vb Abr

where the vector b and the tensor B are specified by the At this stage, a full solution of this problem is necessary
two periodic boundary problems [53]: to discuss seriously the dependence of F on the Reynolds
number and the geometry. For instance, depending on
Problem I: the cases, a quadratic or a cubic dependence of the in-
r  B ¼ 0; ðB:8Þ ertial correction to Darcy’s law at small Reynolds
Z numbers has been found [53,57,58,66]. According to
1 Whitaker [53], the proposed boundary value problem is
0 ¼ rb þ r2 B  nbr  ðIb þ rBÞ dA; ðB:9Þ
Vb Abr equivalent to the Navier–Stokes equations for steady,
incompressible flow in a spatially periodic system. Nu-
B ¼ I at Abr ; ðB:10Þ
merical calculations are presently under development in
bðr þ ‘i Þ ¼ bðrÞ; Bðr þ ‘i Þ ¼ BðrÞ at Abe ; ðB:11Þ order to study the relevance of the periodic boundary
conditions at Abe in the case of dendritic structures and
hBib ¼ 0: ðB:12Þ to describe accurately the dependence of the Forchhei-
mer correction.
After straightforward manipulations, the system (B.8)–
(B.12) can be written under the more convenient form
[39]: References

r  D ¼ 0; ðB:13Þ [1] M. Flemings, Solidification Processing, McGraw-Hill, New


2 York, 1974.
rd þ r D ¼ I; ðB:14Þ
[2] M. Rappaz, V. Voller, Modeling of micro-macrosegrega-
D ¼ 0 at Abr ; ðB:15Þ tion in solidification processes, Met. Trans. A 21 (1990)
749–755.
dðr þ ‘i Þ ¼ dðrÞ; Dðr þ ‘i Þ ¼ DðrÞ at Abe ðB:16Þ [3] P. Prescott, F. Incropera, Convection heat and mass
transfer in alloy solidification, Adv. Heat Transfer 28
hdib ¼ 0; hDib ¼ e1
b K; ðB:17Þ (1996) 231–337.
[4] J. Szekely, A. Jassal, An experimental and analytical study
where D and d are defined by: of the solidification of a binry dendritic system, Met.
Trans. B 9 (1978) 389–398.
D ¼ e1
b ðB þ IÞ  K; ðB:18Þ [5] S. Ridder, S. Kou, R. Mehrabian, Effect of fluid flow on
macrosegregation in axi-symmetric ingots, Met. Trans. B
d ¼ e1
b b  K: ðB:19Þ 12 (1981) 435–447.
3664 P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665

[6] M. Worster, Natural convection in a mushy layer, J. Fluid [24] C. Beckermann, R. Viskanta, Double-diffusive convection
Mech. 224 (1991) 335–359. during dendritic solidification of a binary mixture, Phys-
[7] M. Kaviany, Principles of Heat Transfer in Porous Media, icochem. Hydrodyn. 10 (1988) 195–213.
Springer, Berlin, 1995. [25] C. Beckermann, R. Viskanta, Mathematical modeling of
[8] R. Hills, D. Loper, P. Roberts, A thermodynamically transport phenomena during alloy solidification, Appl.
consistent model of a mushy zone, Quart. J. Mech. Appl. Mech. Rev. 46 (1) (1993) 1–27.
Math. 36 (1983) 505–539. [26] M. Schneider, C. Beckermann, A numerical study of the
[9] W. Bennon, F. Incropera, A continuum model for combined effects of microsegregation mushy zone perme-
momentum heat and species transport in binary solid– ability and flow, caused by volume contraction and
liquid phase change systems – I. Model formulation, Int. J. thermosolutal convection, on macrosegregation and eutec-
Heat Mass Transfer 30 (10) (1987) 2161–2170. tic formation in binary alloy solidification, Int. J. Heat
[10] P. Prescott, F. Incropera, W. Bennon, Modeling of Mass Transfer 38 (18) (1995) 3455–3473.
dendritic solidification systems reassessment of the contin- [27] C. Beckermann, C. Wang, Mutiphase/scale modeling of
uum momentum equation, Int. J. Heat Mass Transfer 34 alloy solidification, Annu. Rev. Heat Transfer 6 (1995)
(9) (1991) 2351–2359. 115–198.
[11] S. Ganesan, D. Poirier, Conservation of mass and mo- [28] A. Bensoussan, J.L. Lions, G. Papanicolaou, Asymptotic
mentum for the flow of interdendritic liquid during Analysis for Periodic Structures, North-Holland, Amster-
solidification, Met. Trans. B 21 (1990) 173–181. dam, 1978.
[12] J. Ni, C. Beckermann, A volume-averaged two-phase [29] E. Sanchez, in: Non-homogeneous media and vibration
model for transport phenomena during solidification, theory, Lecture Notes in Physics, vol. 127, Springer, New
Met. Trans. B 22 (1991) 349–361. York, 1980.
[13] T. Benihaddadene, Modelisation Macroscopique des [30] W. Bennon, F. Incropera, An experimental investigation of
ecoulements et des Transferts dans un milieu poreux binary solidification in a vertical channel with thermal and
heterogene: application a la solidification, Ph.D. Thesis, solutal mixed convection, J. Heat Transfer 111 (1989) 706–
Universite de Paris VI Mai, 1997. 712.
[14] V. Voller, A. Brent, C. Prakash, The modelling of heat [31] P. Prescott, F. Incropera, D. Gaskell, Convective transport
mass and solute transport in solidification systems, Int. J. phenomena and macrosegregation of a binary metal alloy:
Heat Mass Transfer 32 (9) (1989) 1719–1731. II – Experiments and comparisons with numerical predic-
[15] I. Kececioglu, B. Rubinsky, A continuum model for the tions, J. Heat Transfer 116 (1994) 742–749.
propagation of discrete phase-change fronts in porous [32] B. Goyeau, T. Benihaddadene, D. Gobin, M. Quintard,
media in the presence of coupled heat flow fluid flow and Averaged momentum equation for flow through a non-
species transport processes, Int. J. Heat Mass Transfer 32 homogeneous porous structure, Transport Porous Media
(6) (1989) 1111–1130. 28 (1997) 19–50.
[16] P. Prescott, F. Incropera, D. Gaskell, The effects of [33] P. Nandapurkar, D. Poirier, J. Heinrich, Momentum
undercooling, recalescence and solid transport on the equation for dendritic solidification, Numer. Heat Transfer
solidification of binary metal alloys, Transport Phenom. 19 (1991) 297–311.
Mater. Process. Manufact. 196 (1992) 31–39. [34] D. Poirier, Permeability for flow of interdendritic liquid in
[17] W. Bennon, F. Incropera, A continuum model for columnar–dendritic alloys, Met. Trans. B 18 (1987) 245–
momentum heat and species transport in binary solid– 255.
liquid phase change systems – II. Application to solidifi- [35] S. Ganesan, C. Chan, D. Poirier, Permeability for flow
cation in a rectangular cavity, Int. J. Heat Mass Transfer parallel to primary dendrite arms, Mater. Sci. Eng. A 151
30 (10) (1987) 2171–2187. (1992) 97–105.
[18] W. Bennon, F. Incropera, The evolution of macrosegrega- [36] M. Bhat, D. Poirier, J. Heinrich, Permeability for cross
tion in statically cast binary ingots, Met. Trans. B 18 (1987) flow through columnar–dendritic alloys, Met. Mater.
611–616. Trans. B 26 (1995) 1049–1056.
[19] M. Christenson, W. Bennon, F. Incropera, Solidification of [37] B. Goyeau, T. Benihaddadene, D. Gobin, M. Quintard,
an aqueous ammonium chloride solution in a rectangular Numerical calculation of the permeability tensor in a
cavity – II. Comparison of predicted and measured results, dendritic mushy zone, Met. Mater. Trans. B 30 (1999) 613–
Int. J. Heat Mass Transfer 32 (1) (1989) 69–79. 622.
[20] D.G Neilson, F.P. Incropera, Unidirectional solidifica- [38] S. Whitaker, Flow in porous media I: A theoretical
tion of a binary alloy and the effects of induced fluid derivation of Darcy’s law, Transport Porous Media 1
motion, Int. J. Heat Mass Transfer 34 (7) (1991) 1717– (1986) 3–25.
1732. [39] J. Barrere, O. Gipouloux, S. Whitaker, On the closure
[21] P. Prescott, F. Incropera, Convective transport phenomena problem for Darcy’s law, Transport Porous Media 7 (1992)
and macrosegregation during solidification of a binary 209–222.
metal alloy: I–Numerical predictions, J. Heat Transfer 116 [40] M. Quintard, S. Whitaker, Transport in ordered and
(1994) 735–741. disordered porous media III: Closure and comparison
[22] S. Whitaker, Advances in theory of fluid motion in porous between theory and experiment, Transport Porous Media
media, Ind. Eng. Chem. 61 (12) (1969) 14–28. 15 (1994) 31–49.
[23] J. Bear, Dynamics of Fluids in Porous Media, Dover, New [41] J. Cushman, On unifying the concept of scale, instrumen-
York, 1972. tation, and stochastics in the development of multiphase
P. Bousquet-Melou et al. / International Journal of Heat and Mass Transfer 45 (2002) 3651–3665 3665

transport theory, Water Resour. Res. 20 (11) (1984) 1668– [55] D. Nield, A. Bejan, Convection in Porous Media, Springer,
1676. Berlin, 1999.
[42] H. Huppert, The fluid mechanics of solidification, J. Fluid [56] P. Forchheimer, Wasserbewegung durch boden, Z. Ver.
Mech. 212 (1990) 209–240. Deutsch. Ing. 45 (1901) 1782–1788.
[43] R. Trivedi, K. Somboonsuk, Constrained dendritic growth [57] C. Mei, J. Auriault, The effect of weak inertia on flow
and spacing, Mater. Sci. Eng. 65 (1984) 65–74. through a porous medium, J. Fluid Mech. 222 (1991) 647–
[44] W. Gray, A derivation of the equations for multi-phase 663.
transport, Chem. Eng. Sci. 30 (1975) 229–233. [58] M. Firdaouss, J. Guermond, Sur l’homogeneisation des
[45] R. Carbonell, S. Whitaker, in: Fundamentals of Transport equations de Navier–Stokes a faible nombre de reynolds,
Phenomena in Porous Media, Heat and Mass Transfer Comptes Rendus de l’Academie des Sciences 320 (Serie I)
in Porous Media, Martinus Nijhoff, Dordrecht, 1984, (1995) 245–251.
pp. 121–198. [59] S. Hassanizadeh, W. Gray, High velocity flow in porous
[46] M. Quintard, S. Whitaker, Transport in ordered and media, Transport Porous Media 2 (1987) 521–531.
disordered porous media II: Generalized volume averag- [60] H. Ma, D. Ruth, The microscopic analysis of high
ing, Transport Porous Media 14 (1994) 179–206. Forchheimer number flow in porous media, Transport
[47] C. Misbah, Instabilite morphologique et convection solu- Porous Media 13 (1993) 139–160.
tale en solidification directionnelle des melanges binaires [61] J.A. Ochoa-Tapia, S. Whitaker, Momentum transfer at the
dilues, Ph.D. Thesis, Universite Paris VII mars, 1985. boundary between a porous medium and a homogeneous
[48] M. Krane, F. Incropera, Analysis of the effect of shrinkage fluid: I theoretical development, Int. J. Heat Mass Transfer
on macrosegregation in alloy solidification, Met. Mater. 38 (1995) 2635–2646.
Trans. A 26 (1995) 2329–2339. [62] J. Ni, Development of a two-phase model of transport
[49] D. Poirier, K. Yeum, A. Maples, A thermodynamic phenomena during equiaxed solidification, Ph.D. Thesis,
prediction for microporosity formation in aluminium-rich University of Iowa, 1991.
Al–Cu alloys, Met. Trans. A 18 (1987) 1979–1987. [63] R. Wright, Core melt progression status of current
[50] D. Xu, Q. Li, Gravity- and solidification-shrinkage-in- understanding and principal uncertainties, in: J. Rogers
duced liquid flow in a horizontally solidified alloy ingot, (Ed.), Heat and Mass Transfer in Severe Nuclear Reactor
Numer. Heat Transfer A 20 (1991) 203–221. Accidents, Begell House, New York, 1996.
[51] A. Chiareli, M. Worster, Flow focusing instability in a [64] M. Quintard, S. Whitaker, Transport in ordered and
solidifying mushy layer, J. Fluid Mech. 297 (1995) 293–305. disordered porous media volume-averaged equations, clo-
[52] M. Quintard, S. Whitaker, Convection dispersion and sure problems and comparison with experiment, Chem.
interfacial transport of contaminants: homogeneous po- Eng. Sci. 48 (1993) 2537–2564.
rous media, Adv. Water Resour. 17 (1994) 221–239. [65] C. Marle, Ecoulements monophasiques en milieu poreux,
[53] S. Whitaker, The Forchheimer equation: a theoretical Revue de L’Institut Francais du Petrole XXII (10) (1967)
development, Transport Porous Media 25 (1996) 27–61. 1471–1509.
[54] V.D Cvetkovic, A continuum approach to high velocity [66] J. Wodie, T. Levy, Correction non lineaire de la loi de
flow in a porous medium, Transport Porous Media 1 Darcy, Comptes Rendus de l’Academie des Sciences 312
(1986) 63–97. (Serie II) (1991) 157–161.

Вам также может понравиться