You are on page 1of 36

International Journal of Plasticity 17 (2001) 1623±1658

www.elsevier.com/locate/ijplas

A large strain thermoviscoplastic formulation


for the solidi®cation of S.G. cast iron in a
green sand mould
Diego J. Celentano
Departamento de IngenierõÂa MecaÂnica, Universidad de Santiago de Chile, Av. Bdo. O'Higgins 3363,
Santiago de Chile, Chile

Received in ®nal revised form 24 March 2000

Abstract
This paper presents a large strain thermoviscoplastic formulation for the analysis of the
solidi®cation process of spheroidal graphite (S.G.) cast iron in a green sand mould. This for-
mulation includes two di€erent non-associate constitutive models in order to describe the
thermomechanical behaviour of each of such materials during the whole process. The perfor-
mance of these models is evaluated in the analysis of a solidi®cation test. # 2001 Elsevier
Science Ltd. All rights reserved.
Keywords: Cast iron and green sand constitutive behaviours; A. Phase transformation; A. Solidi®cation;
A. Thermomechanical processes; B. elastic-viscoplastic material

1. Introduction

Several thermomechanical models to simulate di€erent casting processes have


been formulated during the last years (see Zabaras et al., 1990; Inoue and Ju, 1992;
Bellet et al., 1996; Celentano et al., 1996; Trovant and Argyropoulos, 1996 and
references therein). In particular, some of them have been used to analyze the soli-
di®cation and subsequent cooling of spheroidal graphite (S.G.) cast iron in green
sand moulds considering relatively simple constitutive models for these materials
(Celentano et al., 1995; Agelet de Saracibar et al., 1999). Although they predicted a
satisfactory agreement between experimental and numerical results, it has been long
recognized that the use of more sophisticated models is necessary to represent in a
more realistic form some particular physical aspects involved in this process
(Hamata, 1992; Azzouz, 1995; Ami Saada et al., 1996; Celentano, 1997).
0749-6419/01/$ - see front matter # 2001 Elsevier Science Ltd. All rights reserved.
PII: S0749-6419(00)00095-4
1624 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

In this context, the aim of this paper is to present a large strain thermoviscoplastic
formulation which includes microstructural liquid±solid and solid±solid phase-
change e€ects, incorporates volumetric expansions due to the metallurgical trans-
formations, deals with temperature-dependent hardening laws and describes the
hygrometric and damage phenomena experienced by the sand. In particular, the
thermomechanical formulation is presented in Section 2 while Sections 3 and 4
describe the S.G. cast iron and sand constitutive models, respectively. These
descriptions start with the choice of the viscoplastic and phase-change internal
variables with their respective evolution equations. Afterwards, a speci®c free energy
function intended to predict the material behaviour during the whole solidi®cation
and cooling process is proposed as a function of the Almansi strain tensor, the
internal variables and the temperature. This de®nition is an extension, considering
large strains, microstructural phase-changes, mixed response for the di€erent phases
and damage e€ects, of the speci®c free energy function developed, used and partially
validated with experiments by Celentano et al. (1995) and Celentano (1997). At this
point, the standard procedures in thermodynamics allow to derive all the con-
stitutive equations involved in the formulation, namely the stress-strain law, the
entropy function, the tangent conjugate of the thermal dilatation tensor, the internal
heat source, the conjugate of the internal variables and the expression of the dis-
sipation. As a consequence of this proposed free energy function, it is possible to
de®ne the elastic contribution of the Almansi strain tensor and thus to recover the
additive decomposition of such tensor which, in turn, it is shown to be consistent
with a particular kinematic decomposition of the deformation gradient tensor. As
reported by Lubliner (1990), this approach is equivalent to other thermomechanical
theories that consider the kinematic decomposition and the speci®c free energy
function written in terms of the elastic part of a strain measure as starting points to
derive the constitutive equations of the formulation (Wriggers et al., 1989; Armero
and Simo, 1993; Levitas, 1998).
Solid±solid phase-transformations are very important processes that have been
extensively analysed by di€erent researchers (see Fischer et al., 1996; Levitas, 1998;
Levitas et al., 1998; Cherkaoui et al., 1998 and references therein) with particular
emphasis to martensitic phase transitions. Some complex aspects of these phenom-
ena, such as transformation-induced plasticity, nucleation criterion, interface propa-
gation, nucleus nondissappearance conditions and displacements discontinuities, have
been recently taken into account in the development of di€erent thermomechanical
formulations (see e.g. Levitas, 1998). In the present work, however, a more simpler
model is adopted due to the fact that low temperatures rates are considered during the
solid-solid phase-change and, therefore, it can be modelled by means of nucleation
and growth laws together with a transformation-induced plasticity equation already
used by Hamata (1992) and Celentano (1997) for ferritic S.G. cast iron.
The motivation to include large deformation concepts in this proposed formula-
tion is based on the invalidity in a strict sense of the in®nitesimal strain assumption
made by some existing models (Celentano et al., 1996, 1999) that may occur when
describing the quasi-incompressible material behaviour in the liquid phase or in the
mushy zone at high temperatures.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1625

Moreover, it should be noted that the consideration of some microstructural


aspects for both the liquid±solid and solid±solid phase-changes in this thermo-
mechanical context in a coupled form is also a relevant feature of the present for-
mulation. This phenomenological approach includes the de®nition of evolution laws
for the S.G. cast iron phase-change internal variables that are assumed to describe
the average microstructure formation occurring in a certain volume at the macro-
scopic level avoiding in this form a microscopic scale modelling of the micro-
mechanisms involved in the phase transformation. Hence, the aim of considering
this assumption in the thermomechanical/microstructural formulation proposed in
this work is to describe in a phenomenological way such complex e€ects in solidi®-
cation problems.
Although little rate-sensitiveness of green sand behaviour could be observed dur-
ing the experiments conducted by Azzouz (1995) and Ami Saada et al. (1996) and
taking into account that the rate-independent plasticity can be considered as a par-
ticular case of viscoplasticity (Lubliner, 1990), a thermoviscoplasticity framework
has been chosen for both the S.G. cast iron and sand models in order to simplify
their presentations.
All these features have been implemented in a coupled thermomechanical/micro-
structural ®nite element model brie¯y described in Section 5. Further, a strain-dis-
placement matrix able to deal with large strain situations avoiding numerical locking
due to the incompressibility of viscoplastic ¯ows is also proposed. Finally, Section 6
presents the analysis of a solidi®cation test where some available experimental
measurements are compared with the numerical results obtained using this proposed
formulation.

2. Thermomechanical formulation

In a general thermomechanical context, the existence of a speci®c Helmholtz free


energy function E ˆ E …E; k ; T† can be assumed as a function of the Green-
Lagrange strain tensor E, the nint -dimensiona1 vector ®eld …k ˆ 1; . . . ; nint ; nint 51†
of phenomenological internal state variables k (usually governed by evolution
equations with zero initial conditions) and the temperature T (see Coleman and
Gurtin, 1967; Lubliner, 1990; Levitas, 1996, 1998 and references therein for this and
other equivalent expressions for the speci®c free energy function). With this de®ni-
tion of E , the Second Piola±Kirchho€ stress tensor S can be derived using the
classical constitutive relation (Coleman and Gurtin, 1967): S ˆ 0 @ E =@E, where 0
is the density at the material (initial) con®guration. As pointed out by Doyle and
Ericksen (1956), an equivalent stress de®nition is obtained in the form:  ˆ 0 @ e =@e
(see also Lubliner, 1985 for further details in the derivation of this expression),
where  is the Kirchho€ stress tensor ( ˆ FSFT , F being the deformation gradient
tensor and T the transpose symbol), e is the Almansi strain tensor and e ˆ
e …e; F; k ; T† is the reformulated speci®c free energy function written in terms of
Eulerian arguments. Therefore, the Cauchy stress tensor , de®ned by  ˆ =0 
(Malvern, 1969), can be computed in this framework as  ˆ @ e =@e, where  is the
1626 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

density at the spatial con®guration. Moreover, by considering the assumptions of


small elastic strains and isotropic material response (usually accepted for metals and
other materials; see Murnagaham, 1937; Garcia Garino and Oliver, 1992; Garcia
Garino, 1993 and references therein), the speci®c free energy function may be writ-
ten as ˆ …e; k ; T†. This last simpli®ed form of is adopted in the formulation to
be described below.
Taking into account the previous considerations, the existence of a speci®c
Helmholtz free energy function ˆ …e; k ; T† leads to the following form of the
governing local equations (all of them valid in
  where
is the spatial con®g-
uration of a body and  denotes the time interval of interest with t 2 ) expressed
by the mass conservation, the equation of motion, the energy balance and the dis-
sipation inequality (Malvern, 1969):

J ˆ 0 …1†

r ‡ bf ˆ u …2†


:
cT rq ‡ r T : d ‡ rint ˆ 0 …3†

qrT ‡ Dint 50 …4†

together with appropriate boundary and initial conditions and the constitutive rela-
tions given by:  ˆ @ =@e is the Cauchy stress tensor,  ˆ @ =@T is the speci®c
entropy function, c ˆ T@2 =@T2 is the tangent speci®c heat capacity, q ˆ krT is
the heat ¯ux vector de®ned according to the Fourier law (k being the conductivity
tensor), b ˆ @2 =@e@T ˆ @=@T is the tangent conjugate of the thermal dilata-

tion tensor, rint ˆ 10 T@qk =@T qk D k =Dt is the speci®c internal heat source
and Dint ˆ qk D k =Dt is the internal dissipation where qk ˆ 0 @ =@ k are the con-
jugate variables of k . According to the nature of each internal variable, the symbols
* and D…†=Dt
 appearing in the previous expressions respectively indicate an appro-
priate multiplication and a time derivative satisfying the principle of material frame-
indi€erence (Malvern, 1969). Further, r is the spatial gradient operator, bf is the
speci®c body force, the superposed dot indicates time derivative, r is the speci®c heat
source, u is the displacement vector, J > 0 is the determinant of the deformation
gradient tensor F(F 1 ˆ 1 !  u, with 1 being the unity tensor) and d is the rate-
:
of-deformation tensor (d ˆ 1=2…r  v ‡ v  r†, where v ˆ u is the velocity vector).
Instead of Eq. (4), an additional more restrictive dissipative assumption reads: q
rT50 and Dint 50 (see Coleman and Gurtin, 1967). The ®rst condition is auto-
matically ful®lled for a semi-positive de®nite conductivity tensor (in the isotropic con-
text assumed in this work, this condition leads to a non-negative conductivity
coecient) while the second imposes restrictions over the constitutive model de®nition.
It is seen that the de®nition of and consequently of D k =Dt are essentials fea-
tures of the formulation in order to describe the thermomechanical/microstructural
behaviour of the materials involved in the solidi®cation process. To this end, the
following split is proposed: nint ˆ nvp pc vp pc
int ‡ nint , where nint and nint refer to the number
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1627

of internal variables related to viscoplastic (non-reversible that may occur in every


material) and phase-change (for S.G. cast iron) or hygrometric (for green sand)
e€ects, respectively. In this context, this assumption leads to rint ˆ rvp vp
int ‡ rint and
vp pc
Dint ˆ Dint ‡ Dint . The evolution equations for the viscoplastic internal variables are
de®ned within the non associate thermoviscoplasticity theory framework (Perzina,
1971) as D k =Dt ˆ gk where gk ˆ k @=@qk (no sum on k) are known functions
de®ned in terms of a function k and a viscoplastic potential  which in turn may
depend (among other variables) on a yield function F ˆ F…e; k ; T†…k ˆ 1; . . . ; nint †
such that no viscoplastic evolutions occur when F < 0 (Perzina, 1971; Hamata,
1992; Arnold and Saleeb, 1994; Azzouz, 1995; Celentano, 1997). On the other hand,
the rate equations for the phase-change internal variables of the S.G. cast iron are
given by microstructural models based on kinetics considerations (Banerjee and
Stefanescu, 1991; Chang et al., 1991; Goettsch and Dantzig, 1994; Celentano and
Cruchaga, 1999; Onsùien et al., 1999) and the hygrometric internal variable
accounts for the water content existing in the sand (Azzouz, 1995). Moreover, the
speci®c free energy function is decomposed into a thermoelastic, thermoviscoplastic
and phase-change (for S.G. cast iron) or hygrometric (for green sand) contributions.
The viscoplastic potential  is also proposed in additive form for the S.G. cast iron
considering a thermoviscoplastic and phase-change terms while a purely thermo-
viscoplastic behaviour is assumed for the sand (Celentano, 1997). Some details of
the S.G. cast iron and green sand constitutive models are given below.

3. S.G. cast iron model

S.G. cast iron (usually called ductile or nodular iron) has become an engineering
material of major importance since this ternary alloy has many of the mechanical
advantages of steel with the processing economies of cast iron. In this type of cast
iron, the adding of magnesium causes the graphite ¯akes to form in shape of spheres
frequently called spherulites. In spite of decades of study, not all the kinetic
mechanisms involved in the S.G. cast iron solidi®cation are actually known (Flem-
ings, 1974). However, several models have been recently proposed in order to predict
the microstructure formation in castings (see Banerjee and Stefanescu, 1991; Chang
et al., 1991; Goettsch and Dantzig, 1994; Celentano and Cruchaga, 1999; Onsùien et
al., 1999 and references therein).
In a hypoeutectic S.G. cast iron, liquid-solid and solid-solid phase-changes take
place during solidi®cation and cooling. As the temperature of the melt decreases
below the liquidus temperature, primary austenite dendrites are formed. Then, gra-
phite and cementite eutectic grains are developed through nucleation and growth as
the respective eutectic temperatures are reached. The solid-solid or eutectoid trans-
formation is also a competitive growth process consisting of two reactions: auste-
nite-ferrite and austenite-pearlite. Below the stable eutectoid temperature, the ferrite
begins to form. If the transformation has not been completed before the metastable
eutectoid temperature is attained, pearlite starts to form and grows competitively
with ferrite. Therefore, the following relation is assumed to hold:
1628 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

fl ‡ fa ‡ fg ‡ fc ‡ ff ‡ fp ˆ 1 …5†

where fl , fa , fg , fc , ff and fp are the liquid, austenite, graphite eutectic, cementite


eutectic, ferrite and pearlite  volumetric fractions, respectively. For the liquid±solid
phase-change ff ˆ fp ˆ 0 , Eq. (5) reads:

fl ‡ fa ‡ fg ‡ fc ˆ 1 …6†

whereas for the solid±solid phase change (fl ˆ 0, fg ˆ cte, i.e. neglecting secondary
and eutectoid graphite precipitations and fc ˆ cte) it becomes:

fa ‡ ff ‡ fp ˆ 1 fg fc …7†

such that only the variable fractions for the respective phase-changes are written on
the left-hand side of these two last expressions. In this context, the full description of
the material behaviour considering the particular response of these constituents in a
uni®ed formulation using standard concepts of the mixing theory is the main feature
of the present constitutive model. Thus, any mixed variable jmx of this composite
material can be de®ned by weighting the di€erent contributions of each constituent
or component as:

jmx ˆ fl jl ‡fa ja ‡fg jg ‡fc jc ‡ff jf ‡fp jp …8†

As usual, in this mixing approach  ˆ jmx and, therefore, di€erent densities can
be considered for each constituent. The density at the initial con®guration is also
de®ned by Eq. (8) but taking the initial volumetric fractions in its calculation.
Moreover, the computation of jcp …cp ˆ l; a; c; f; p† could be carried out by con-
sidering additional mass balances between the di€erent constituents. Note, however,
that in the constitutive equations described below there is no need to compute jcp as
they may only depend on  which is, in fact, obtained through Eq. (1).
It should be mentioned that this approach is a generalization of that de®ned in
mixed form in terms of the following macroscopic phases (Celentano, 1997): 1)
liquid, 2) mushy zone (liquid+austenite+graphite+cementite), 3) solid (austeni-
te+graphite+cementite), 4) (austenite+graphite+cementite+ferrite+pearlite) and
5) solid (graphite+cementite+ferrite+pearlite).
In what follows any phase-change variable denoted as pc will be decomposed as:

pc ˆ pc ja ‡pc jg ‡pc jc ‡pc jf ‡pc jp …9†

where the ``a, g, c'' and ``f, p'' terms are respectively related to the liquid±solid and
solid±solid phase-changes.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1629

3.1. Viscoplastic internal variables

A possible choice for the  viscoplastic internal variables is: 1 ˆ evp , 2 ˆ #vp ,
vp vp
3 ˆ  , 4 ˆ  nint ˆ 4 with q1 ˆ , q2 ˆ C , q3 ˆ Kvp , q4 ˆ T where evp is the
vp vp

Almansi viscoplastic strain tensor, #vp is the viscoplastic isotropic hardening vari-
able, vp is the viscoplastic kinematic hardening tensor, vp is the viscoplastic yield
entropy,  ˆ J is the Kirchho€ stress tensor, Cvp is the viscoplastic isotropic hard-
ening function and Kvp is the back stress tensor. This choice is an adaptation to the
present viscoplastic context of that considered by Celentano et al. (1999) for elasto-
plasticity. The corresponding evolution equations, assuming k ˆ 1 and
gk ˆ @=@qk , are de®ned as:

@
Lv …evp † ˆ …10:1†
@

: @
#vp ˆ vp …10:2†
@C

@
Lv …vp † ˆ …10:3†
@Kvp

: @
& vp ˆ …10:4†
@T

where the symbol Lv denotes the well-known Lie derivative (Marsden and Hughes,
1983). Note that Eq. (10.1) de®nes in fact the evolution of the viscoplastic rate-of-
deformation tensor dvp ˆ Lv …evp †. The evolution equation for vp originally proposed
within a thermoplasticity framework as consistent with the principle of maximum
plastic dissipation (Armero and Simo, 1993) and extensively used in solidi®cation
problems (Celentano et al., 1996, 1999), is also adopted here.
The viscoplastic potential is assumed to be given by (Hamata, 1992; Celentano,
1997):

 ˆ tp jmx ‡pc …11†

where tp jmx and pc are the thermoviscoplastic and phase-change parts of ,
respectively. The ®rst term accounts for classical viscoplastic phenomena and is
written in mixed form through Eq. (8). The phase-change term describes the solid±
solid phase transformation viscoplasticity e€ect during the ferrite and pearlite
transformations occuring in the macroscopic phase 4).
For each component, tp jcp …cp ˆ l; a; g; c; f; p† is written as:

1‡nv jcp
Fjcp
tp jcp ˆ  n j …12†
1 ‡ nv jcp Kv jcp v cp
1630 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

such that Kv jcp ˆ Kv jcp …T† is the viscosity, nv jcp ˆ nv jcp …T† is a material parameter, hi
denotes de MacAuley symbol and Fjcp is the von Mises temperature-dependent yield
function de®ned as:
p
Fjcp ˆ 3J2k Cjcp …13†

where J2K ˆ 1=2‰ Kvp Š0 : ‰ Kvp Š0 is the second invariant of the deviatoric tensor
 0 Kvp 0 and Cjcp ˆ Cjcp …Cvp ; T† is the total hardening function given by:

Cjcp ˆ Cth jcp ‡Cvp …14†

with Cth jcp ˆ Cth jcp …T† being thermal hardening function which, in absence of visco-
plastic e€ects, establishes the elastic domain for a given temperature. It should be
noted that, in general, Cth jcp could be di€erent for each component.
As stated above, the phase change term pc is de®ned according to Eq. (9) where a
simpler model, originally validated with experiments by Hamata (1992) for the aus-
tenite-ferrite transformation, is considered choosing cp jph ˆ 0…ph ˆ a; g; c† while the
solid-solid phase-change contribution pc jph …ph ˆ f; p† takes the form:

p1‡nt jph
J2K :
pc jph ˆ  nt f ph …15†
1 ‡ nt jph Kt jph

where, once more, the viscosity Kt jph and the material parameter nt jph characterize
the coupled solid viscous behaviour during the ferrite and pearlite transformations.
Note that Eq. (15), which leads to a non associate constitutive model, states that this
phase transformation is considered as a perfect viscoplastic e€ect that only takes
place when fph …ph ˆ f; p† evolves and, moreover, it a€ects the evolution laws for evp
and vp given by Eqs. (10.1) and (10.3), respectively. Further, the evolution of evp
due to the solid-solid phase-change is, essentially, very similar to di€erent simple
transformation-induced plasticity models reviewed by Fischer et al. (1996), i.e. such
evolution is proportional to the deviatoric stress plus the rate of the volumetric
fraction.

3.2. Phase-change internal variables

According to the above considerations, the phase-change


 internal variables are:
1 ˆ fa , 2 ˆ fg , 3 ˆ fc , 4 ˆ ff , 5 ˆ fp npc
int ˆ 5 with q1 ˆ qa , q2 ˆ qg , q3 ˆ qc ,
q4 ˆ qf , q5 ˆ qp . As mentioned above, the austenite, graphite and cementite frac-
tions are associated to the liquid phase-change while the ferrite and pearlite evolve
during the solid-solid transformation.
In the liquid-solid phase-change, experimental observations suggest that both the
dendritic and eutectic grains are equiaxed except over a small region located near the
mould walls where columnar grains appear (Chang et al., 1991). In this work all
grains formed are assumed to be equiaxed and spherical in shape for the eutectics or
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1631

enveloped within a sphere for the dendritic grains. Although di€erent microstructure
models for S.G. cast iron have been proposed taking into account particular kinetic
aspects occuring during the solidi®cation and cooling process (Banerjee and Stefa-
nescu, 1991; Chang et al., 1991), a simpler microstructure model to describe both the
liquid±solid and solid±solid phase-changes is considered in this work. This model
includes primary austenite dendrites formation, nucleation and growth laws for the
graphite eutectic, cementite eutectic, ferrite and pearlite fractions.
Considering that during the austenite solidi®cation the carbon di€uses very
rapidly in iron, the austenite volumetric fraction can be directly determined by the
inverse lever rule as (Goettsch and Dantzig, 1994):

1 T Tl
fa ˆ …16†
1 k0 T T 0
:
subject to the condition f a 50, where k0 ˆ k0 (%Si) is the equilibrium distribution
coecient which is the relationship between the slope of the liquidus and solidus
curves of the Fe-C diagram, Tl ˆ Tl (%C, %Si) is the liquidus temperature, T 0 ˆ
T 0 (%C, %Si) is the temperature corresponding to the intersection of the liquidus
and solidus temperature curves and %C and %Si denote the carbon and silicon
contents, respectively. It is important to mention that the evolution of fa directly
derived from Eq. (16) does not take into account neither the in¯uence of the cooling
rate nor the remelting e€ect (Celentano and Cruchaga, 1999).
The eutectic fractions fg and fc can be determined through a nucleation and
growth laws. Assuming a quasi-instantaneous nucleation and a subsequent spherical
growth, the expression for such fractions is (Goettsch and Dantzig, 1994):

4
fph ˆ Njph R3 jph …17†
3

where Njph is the grain density and Rjph is the grain size …ph ˆ g; c†. These variables
are respectively obtained with the nucleation and growth models brie¯y presented
below. For simplicity, the grain impingement e€ect is not included in Eq. (17) (for
more details, see Chang et al., 1991).
The following nucleation law is adopted in the present work:

:
nj 1 D : E
N jph ˆ Ajph njph Tjph T ph T …18†

where Ajph and njph …ph ˆ g; c† are nucleation parameters which depend on %C, %Si
and the inoculant content. In Eq. (18), Tg and Tc denote the graphite (stable) and
cementite (metastable) eutectic temperatures, respectively, which also depend on
%C and %Si. It is well-known that the major e€ect of silicon is to widen the range
Tg Tc (Flemings, 1974). As this range increases, the probability of forming gra-
phite eutectic rather than a cementite eutectic is also increased. Hence, %Si pro-
motes the formation of graphite. Moreover, Eq. (18) clearly states that the grain
1632 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

density is constrained
to increase
and this fact is only possible for positive under-
coolings (indicated by Tjph T ) and negative cooling rates. Between Tg and Tc only
the graphite eutectic may nucleate while below Tc the cementite eutectic may also
nucleate. As can be seen, distinct nucleation rates are obtained for graphite and
cementite eutectics due to the di€erences in the respective undercoolings.
The grain size evolution equation is de®ned as:
:
mj
Rjph ˆ Bjph Tjph T ph …19†

where Bjph and mjph (ph ˆ g; c) are growth parameters. Once again, note that in the
temperature range Tg Tc only the graphite eutectic evolution takes place and,
therefore, Tc can be considered as the transition temperature for the solidi®cation of
gray to white iron. Further, Bjg ˆ 1=10Bjc is chosen in order to represent the
experimental fact showing that the growth rate of graphite is about 1=10 of that of
cementite for equal respective undercoolings (Flemings, 1974). According to Eq. (6),
the liquid-solid phase-change ends once fl ˆ 0, i.e. fa ‡ fg ‡ fc ˆ 1.
In the solid-solid phase-change, the modelling of ferrite and pearlite can be also
described by a law similar to Eq. (17) where, in this case, ph ˆ f; p. For the ferrite
transformation, in particular, the assumptions leading to this expression can be
found in Chang et al. (1991). It should be noted that for low temperature rates
(approximately less than 0.2 C) a simpli®ed model of ferrite, based on the Avrami±
Johnson±Mehl equation, can be considered (Hamata, 1992). Moreover, an Avrami
equation can be also used to compute the fraction of pearlite (Chang et al., 1991).
However, for simplicity, the uni®ed expression given by Eq. (17) is retained.
The nucleation and growth laws for the ferrite and pearlite transformation are
assumed to be respectively given by expressions similar to Eqs. (18) and (19) written,
for these cases, in terms of the ferrite and pearlite undercoolings and using particular
nucleation and growth parameters …ph ˆ f; p†. These undercoolings are de®ned with
the ferrite (stable) eutectoid temperature Tf and the pearlite (metastable) eutectoid
temperature Tp . It should be mentioned that other nucleation and growth laws have
been proposed in the literature where the more relevant are a di€usion-controlled
growth-rate equation for the ferrite (Chang et al., 1991) and exponential laws based
on the eutectoid undercooling for the nucleation and growth of pearlite (Goettsch
and Dantzig, 1994). Finally, the solid±solid phase-change ®nishes, as stated by Eq.
(7), when fa ˆ 0, that is ff ‡ fp ˆ 1 fg fc .

3.3. Speci®c free energy function

The following speci®c free energy function is proposed:

ˆ te jmx ‡ tp jmx ‡ pc …20†

where te jmx ; tp jmx and pc are the thermoelastic, thermoviscoplastic and phase-
change parts of . This partially coupled additive form of is assumed to describe
the isotropic material behaviour in all the above mentioned macroscopic phases
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1633

present in the solidi®cation process. In all the equations to be described below, the
subscript 0 indicates the value of the variables at the initial con®guration and the
superscript s refers to secant thermomechanical properties measured at the spatial
con®guration with respect to the reference temperature Tref . Additionally, e0 ˆ 0 is
considered.
The thermoelastic part te jcp …cp ˆ l; a; g; c; f; p† of each component is:

1 1
te jcp ˆ …e evp † : Cs jcp : …e evp † …e evp † : Cs jcp : eth jcp
20 0
1
‡ c jcp c0 jcp ‡ …e evp † : 0 0 …T T0 † ‡ 0 jcp …21†
0

where Cs jcp ˆ Cs jcp …T† is the secant elastic isotropic constitutive tensor and eth jcp is
the Almansi thermal strain tensor given by:

1h 2=3 i
eth jcp ˆ 1 1 ath jcp 1 …22†
2

where ath jcp ˆ sth jcp …T Tref † sth0 jcp …T0 Tref † with sth jcp ˆ sth jcp …T† being the
secant volumetric thermal dilatation coecient. Note that if ath jcp  1, eth jcp ˆ
ath jcp =31 which is, in fact, the classical expression for the in®nitesimal thermal strain
tensor. The term c jcp is:

…T
c jcp ˆ Ac jcp d …23†
Tref

where Ac jcp is a function that can be de®ned in terms of the secant speci®c heat
capacity cs jcp ˆ cs jcp …T† (Celentano et al., 1996).
As usual, the secant elastic isotropic constitutive tensor can be decomposed into
its deviatoric and volumetric parts (Malvern, 1969). According to Celentano et al.
(1996), a particular de®nition of Cs jl is adopted here which consists of neglecting its
deviatoric contribution assuming, therefore, a purely elastic volumetric behaviour
for this phase.
The thermoviscoplastic part of is …cp ˆ l; a; g; c; f; p†:

1 1 1
tp jcp ˆ hC jcp …#vp †2 ‡ hK jcp vp : vp T& vp …24†
20 20 0

where hC jcp ˆ hC jcp …T† and hK jcp ˆ hK jcp …T† are the viscoplastic isotropic and kine-
matic hardening moduli, respectively.
The phase-change part of is given by Eq. (9) assuming similar expressions for
the liquid-solid and solid±solid phase-changes, that is …ph ˆ a; g; c; f; p†:
1634 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

pc jph ˆ L jph ‡  jph …25†

with

L jph ˆ Ls jph fph …26†

1
 jph ˆ …e evp † : Cs jph : epc jph …27†
0

where Ls jph ˆ Ls jph …T† is the secant speci®c latent heat and epc jph is the Almansi
phase-change strain tensor both associated to the volumetric fraction fph . This
phase-change strain tensor is similarly de®ned as the Almansi thermal strain tensor
[Eq. (22)] in the form:

1h 2=3 i 2=3
epc jph ˆ 1 1 apc jph 1 ath jph 1 …28†
2

where apc jph ˆ spc jph fph and spc jph ˆ spc jph …T† is the secant phase-change volumetric
deformation. As it can be seen, Eqs. (26) and (27) account for latent heat release and
phase-change volumetric deformation, respectively (see Flemings, 1974, for a further
discussion on the physical aspects of these phenomena).
The de®nition of the di€erent contributions of given by Eqs. (21), (24) and (27)
consider the density at the initial con®guration 0 instead of its current value . This
simpli®cation is consistent with the Doyle±Ericksen approach (1956) commented in
Section 2. In the constitutive equations, however, any density change is given by J.
Moreover, as mentioned above, the present de®nition of is an extension, con-
sidering large strains, microstructural phase-changes and mixed response, of the
speci®c free energy function proposed and used by Celentano et al. (1999).
It should be noted that this large strain context allows to predict the external
(surface) shrinkage occuring in some casting situations. A simple internal shrinkage
prediction criterion due to liquid contraction (Celentano, 1998a) could be included
in Eq. (27) of the present formulation but, for simplicity, is not considered in this
work.
The proposed de®nition of allows the derivation of all the constitutive equations
and the internal dissipation described in Section 2. Some details are discussed below.

3.4. Constitutive laws

With the considerations given above, the Cauchy stress tensor is given by:

 ˆ te jmx ‡pc ‡ 0 …29†

where te jcp …cp ˆ l; a; g; c; f; p† is:


D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1635

1 
te jcp ˆ Cs jcp : e evp eth jcp …30†
J

and the phase-change term …ph ˆ a; g; c; f; p† is once more splitted as in Eq. (9) with:

1 s
pc jph ˆ C j : epc jph …31†
J ph

This secant or hyperelastic constitutive law circumvents some usual thermo-


dynamic constraints (see Cassenti and Annigeri, 1989), exclusively depends on the
thermoelastic and phase-change terms of the free energy function and, according to
the de®nition of , is valid to describe the material behaviour during the whole
solidi®cation and cooling process.
In the classical ®nite deformation elastoplasticity theory, the additive decomposi-
tion of the Almansi strain can be deduced from the multiplicative decomposition of
the deformation gradient into the elastic and plastic parts F ˆ Fe Fp where, for this
case, the link between the multiplicative theory and the additive decomposition of e
is the elastic Finger tensor (see Green and Naghdi, 1971; Kleiber, 1975; Simo and
Ortiz, 1985). In the present formulation, however, it should be noted that the addi-
tive decomposition of the Almansi strain is only recovered for each component as:
ee jcp ˆ e eth jcp for cp ˆ 1 and ee jcp ˆ e evp eth jcp epc jph for cp ˆ ph ˆ a; g; c; f; p
where ee jcp is the Almansi elastic strain tensor. These decompositions are respec-
tively deduced from the multiplicative decompositions of the deformation gradient
tensor given by: F ˆ Fe jcp  Fth jcp for cp ˆ 1 and F ˆ Fe jcp  Fth jcp  Fpc jph  Fvp for cp ˆ
ph ˆ a; g; c; f; p considering for such cases the assumption of small elastic strains. In
this context, note that the thermal part of e [Eq. (22)] depends exclusively on the
1=3
temperature through Fth jcp1 ˆ 1 ath jcp 1 while in the de®nition of the phase-
1=3
change contribution epc [Eq. (28)] both Fth jcp1 and Fpc jph1 ˆ 1 apc jph 1 are
involved. Although the split of F into elastic, thermal and viscoplastic contributions
has been already proposed in the literature (Wriggers et al., 1989; Bammann et al.,
1995; Casey, 1998) as well as recent extensions to include the deformation due to the
martensitic transformation in steels (Levitas, 1998), the consideration of the phase-
change e€ects in such de®nition for solidi®cation problems is an original feature of
the present work. Furthermore, the multiplicative decomposition assumed in this
formulation is consistent with the arguments discussed in detail by Levitas (1998)
regarding to the contradictions associated to the di€erent possible combinations to
de®ne F.
It should be noted again that the constitutive law (29) is only valid for small elastic
strains. This constraint, however, is not a strong restriction of the model since either
the viscoplastic, thermal or phase-change parts of e usually predominate over ee in
the deformation of S.G. cast iron occuring during the solidi®cation and cooling
process.
As mentioned above, the deviatoric response in the liquid phase is neglected with
the adopted de®nition of Cs jl . Nevertheless, an additional assumption (see Lubliner,
1636 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

1990; Baldoni and Rajagopal, 1997) consisting of adding the term 2d ( being the
dynamic viscosity of the ¯uid) to the expression given by Eq. (29) for cp ˆ l would
allow to include this e€ect. In this context, the resulting viscous term 2d : d should
be added to the internal dissipation equation. For simplicity, however, this viscosity
e€ect is not considered in the ¯uid phase throughout this work.
Further, the expressions for the speci®c entropy function, the tangent speci®c heat
capacity, the tangent conjugate of the thermal dilatation tensor, the internal heat
source and the conjugate of the internal variables (note that the relations for  and T
are identically ful®lled) can be straightforwardly derived considering the de®nition
of given above. Such expressions can be found in Box 1

Box 1. Variables of the S.G. cast iron model

Speci®c entropy function

 ˆ te jmx ‡tp jmx ‡pc ‡ 0

with (cp ˆ l; a; g; c; f; p and ph ˆ a; g; c; f; p†

1 @Cs jcp 1
te jcp ˆ …e evp † : : …e evp † ‡ …e evp † : Cs jcp : eth T jcp 1
20 @T 0
1 @Cs jcp th
‡ …e evp † : e jcp ‡Ac jcp
0 @T
1 @hC jcp vp 2 1 @hK jcp vp vp 1 vp
tp jcp ˆ …# †  : ‡ &
20 @T 20 @T 0
s
@Ls jph 1 @C j ph 1
pc jph ˆ fph ‡ …e evp † : : epc jph ‡ …e evp † : Cs jph : epc
T jph 1
@T 0 @T 0

where
 
1  1=3 @ s jcp
eth
T jcp ˆ 1 ath jcp asth jcp ‡ th …T Tref †
3 @T
and
s h 2=3 i th
1  1=3 @pc jph j 2=3
epc
T jph ˆ 1 apc jph 1 ath jph fph 1 1 apc jph eT jph
3 @T

Tangent speci®c heat capacity

c ˆ cte jmx ‡ctp jmx ‡cpc


(continued on next page)
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1637

Box 1 (continued)
with …cp ˆ l; a; g; c; f; p and ph ˆ a; g; c; f; p†

T @2 Cs jcp 2T @Cs jcp


cte jcp ˆ …e evp † : : …e evp † ‡ …e evp † :
20 @T2 0 @T
T @eth jcp T @Cs jcp
: eth
T jcp 1 ‡ …e evp † : Cs jcp : T 1 ‡ …e evp † :
0 @T 0 @T2
s
@c jcp
: eth jcp ‡cs jcp ‡ …T Tref †
@T

T @2 hC jcp vp 2 T @2 hk jcp vp vp
ctp jcp ˆ …# †  :
20 @2 T 20 @2 T

@2 Ls jph T @2 Cs jph pc 2T @Cs jph


cpc jph ˆ T fph ‡ …e evp † : : e jph ‡ …e evp † :
@T 2 0 @T 2 0 @T
2T @Cs jph pc
vp T vp s @epc
T jph
:1‡ …e e †: : eT jph 1 ‡ …e e † : C jph 1
0 @T 0 @T

Tangent conjugate of the thermal dilatation tensor

ˆ te jmx ‡ pc

with …cp ˆ l; a; g; c; f; p and ph ˆ a; g; c; f; p†

1 1 @Cs jcp 
te jcp ˆ Cs jcp : eth
T jcp 1 : e evp eth jcp
J J @T

1 @Cs jph pc 1
pc jph ˆ : e jph ‡ Cs jph : epc
T jph 1
J @T J

Speci®c internal heat source


     
1 @Cvp : @Kvp
rvp
int ˆ …T ‡  † : Lv …evp † T C #vp
vp
T K vp vp
: Lv … †
0 @T @T

(continued on next page)


1638 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Box 1 (continued)
"
@Ls jph T @Cs jph pc
rpc
int jph ˆ T ‡ Ls jph …e evp † : : ef jph 1
@T 0 @T
1
‡ …e evp † : Cs jph : epc
f jph 1
0
#
T vp @epc
f jph
s
:
…e e † : C jph : 1 f ph for ph ˆ a; g; c; f; p
0 @T

and

1  1=3
epc
f jph ˆ 1 apcjph spc jph
3

Conjugate of the internal variables

Cvp ˆ Cvp vp vp
tp jmx where Ctp jcp ˆ hC jcp # for cp ˆ l; a; g; c; f; p

Kvp ˆ Kvp vp
tp jmx where Ktp jcp ˆ hK jcp vp for cp ˆ l; a; g; c; f; p

qph ˆ 0 Ls jph ‡…e evp † : Cs jph : epc


f jph 1 for ph ˆ a; g; c; f; p

3.5. Dissipation inequality

In this context, the viscoplastic internal dissipation is written as:


: : vp
Dvp vp vp vp vp vp
int ˆ  : Lv …e † ‡ C # ‡ K : Lv … † ‡ T& …32†

and the corresponding phase-change contribution is …ph ˆ a; g; c; f; p†:


 
pc s 1 vp s
 1=3 s :
Dint jph ˆ 0 L jph ‡ …e e † : C jph : 1 apc jph pc jph 1 f ph …33†
30

such that, as mentioned before, Dvp pc


int ‡ Dint 50. It is worth noting that the volumetric
fraction derivatives of te jmx and tp jmx have not been considered in Eq. (33) as they
are assumed to be negligible in comparison to the latent heat and phase-change
volumetric conjugate variables of fph indicated in the bracketed terms.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1639

It has been shown (Celentano et al., 1996, 1999) that the condition Dvp int 50 is
satis®ed if two sucient conditions related to the isotropic hardening behaviour and
the thermal softening e€ect are assumed: (i) hC 50 and (ii) @Cth jcp =@T50
…cp ˆ l; a; g; c; f; p†. Knowing that the ®rst term in Eq. (33) (with Ls jph 50
…ph ˆ a; g; c; f; p†) is in general more relevant that the second one in many casting
situations, the ful®llment of the condition Dvp pc
int ‡ Dint 50 can be guaranteed.
Moreover, it should be noted that the second term of Eq. (33) contains a trans-
formation (phase-change) work since it can be written, with the help of Eq. (30), as:
  1=3 s
J=3 te jph ‡1=JCs jph : eth jph : 1 apc jph pc jph 1, where the product of mean ther-
moelastic stress and phase-change volumetric deformation for each phase can be
seen. A similar expression of this conjugate variable to fph using di€erent state vari-
ables in the speci®c free energy de®nition has been reported by Levitas (1998) for
martensitic and other transformations.
Finally, a mixed form for the conductivity coecient is assumed, i.e. k ˆ kjmx ,
such that kjcp 50 …cp ˆ l; a; g; c; f; p† in order to verify the thermal dissipation men-
tioned in Section 2.

4. Green sand model

Green sand, as usually prepared and used in foundries, is a ®ne porous mixture of
approximately 87 wt.% of silica grains, 7 wt.% of clay, 3 wt.% of carbon black and
3 wt.% of free water. Thus,

fm ‡ fv ˆ 1 …34†

where fm and fv respectively denote the matter (silica grains+clay+carbon black)


and pore volumetric fractions. Typical compaction (the compaction of sand during
the manufacturing process of the mould is not studied here) corresponds to an
overall density of 1600 kg/m3, an average value of the porosity fv of 0.26 and a water
saturation of 0.18 (Azzouz, 1995). Considering that the pores are partially ®lled with
water, the value fw0 ˆ 0:26 0:18 ˆ 0:0468 is therefore taken as an initial condition
for the free water volumetric fraction fw which approximately leads to a maximum
stress peak in the compression test.
Complex phenomena occur in the sand when it is subjected to thermomechanical
loading imposed during casting (Azzouz, 1995; Ami Saada et al., 1996). As a ®rst
approach to the problem, however, most of them are considered through the
dependence of the constitutive laws upon temperature and only the vaporization of
free water as an e€ect related to the disappearing of initial humidity will be included
in the model. Moreover, an eventual condensation of water steam is not considered.
In this green sand constitutive model, the viscoplastic and phase-change e€ects are
mainly related, respectively, to damage (degradation of the material) and hygro-
metric (matrix suction due to the presence of free water) phenomena occuring in the
material during the process.
1640 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Once more, the speci®c free energy function is additively decomposed into a
thermoelastic, thermoplastic and hygrometric contributions. Moreover, a non
associate form of the viscoplastic potential , assumed to be independent on the
hygrometric e€ect, is considered for the viscoplastic evolution equations (Azzouz,
1995; Celentano, 1997). For simplicity, the e€ect of the free water is neglected in the
de®nition of te and tp and it is only considered in pc . As no mixing law is applied
to those terms, the variations of the thermomechanical properties with fw are simply
taken into account by means of temperature-dependent laws.

4.1. Viscoplastic internal variables

The following set of viscoplastic internal variables is adopted: 1 ˆ evp , 2 ˆ dvp ,


3 ˆ vp nvp vp
int ˆ 3 with q1 ˆ , q2 ˆ Y , q3 ˆ T where the same notation of previous
Sections is used and, additionally, d is the viscoplastic damage variable and Yvp is
vp

the viscoplastic damage function. The corresponding evolution equations, assuming


gk ˆ @=@qk are

@
Lv …evp † ˆ  …35:1†
@

: @
dvp ˆ  vp …35:2†
@Y

: @
& vp ˆ  …35:3†
@T

where, as shown below, both isotropic and kinematic hardening e€ects are not
de®ned by evolution laws because they can be considered as a function of  and dvp .
The function  is de®ned as:

hF in v
ˆ …36†
Kv

such that Kv ˆ Kv …T† is the viscosity, nv ˆ nv …T† is a material parameter and F is the
yield function given by:

F ˆ f1 I21 ‡ f2 J2 ‡ f3 I1 C …37†

where I1 ˆ tr… † (tr being the trace symbol), J2 ˆ 1=2 0 :  0 is the second invariant of
the deviatoric tensor  0 , f1 ˆ 12 K…1 dvp †2 , f2 ˆ 3=G…1 dvp †2 , f3 ˆ hK =…1 dvp † (K,
G and hK being the bulk, shear and kinematic hardening moduli, respectively) and
C ˆ C…dvp ; T† is the total hardening function given by:

C ˆ Cth ‡ hC dvp …38†


D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1641

with Cth ˆ Cth …T† being the thermal hardening function and hC is the isotropic hard-
ening modulus. Note that for small values of Kv , i.e. Kv ! 0, the rate-independent
plasticity is directly recovered. This expression of F corresponds to an ellipsoid in
the Westergaard stress-space. Conversely to a classical Drucker±Prager criterion,
this yield function is able to describe the inelastic behaviour which is induced by the
degradation of the material observed in the case of loadings close to the hydrostatic
pressure. The term linearly dependent on I1 introduces a non-symmetric elastic
domain in tension and compression (Azzouz, 1995).
The viscoplastic potential  is:

 ˆ F ‡ hd Yvp …39†

where hd is a material property associated to the material degradation (Azzouz,


1995).

4.2. Hygrometric internal variable



The vaporization of free water is described by 1 ˆ fw npc
int ˆ 1 with q1 ˆ qw . The
evolution of fw is assumed to depend exclusively on T. Hence,

: H…fw †fw0 D : E
fw ˆ T …40†
Tv T0

where H is the Heaviside function and Tv is the water vaporization temperature. The
irreversible evolution of fw can clearly be noted in this expression.

4.3. Speci®c free energy function

As mentioned above, the proposed speci®c free energy function that governs the
isotropic material behaviour during the process is written as:

ˆ te ‡ tp ‡ pc …41†

where te , tp and pc are the thermoelastic, thermoviscoplastic and hygrometric


parts of .
The thermoelastic part te is:

1 1
te ˆ …e evp † : …1 dvp †Cs : …e evp † …e evp † : …1 dvp †Cs : eth
20 0
1
‡ c c0 ‡ …e evp † : 0 0 …T T0 † ‡ 0 …42†
0

which is, in fact, similar to Eq. (21) for a single-component material taking into
account the degradation factor …1 dvp † in the elastic constitutive tensor.
1642 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

In the absence of viscoplastic hardening e€ects, the thermoviscoplastic part of is


simply written as:

1
tp ˆ T& vp …43†
0

The hygrometric part of is given as:

pc ˆ L L0 ‡  …44†

with

L ˆ Ls fw …45†

1
 ˆ …e evp † : …1 dvp †Cs : epc …46†
0

where the expressions of the variables involved correspond to those presented above
for a single-component material. As in the previous constitutive model, Eqs. (45)
and (46) include the latent heat release and hygrometric volumetric deformation,
respectively. In this case, note that the initial hygrometric e€ect is explicitly con-
sidered in Eq. (44) and, therefore, is not included in the initial stress of Eq. (42).
Further, the consideration of 0 in Eqs. (42), (43) and (46) is based on the same
arguments stated above for the S.G. cast iron model.
Once more, as shown below, with the proposed de®nition of it is possible to
obtain all the constitutive equations and the internal dissipation corresponding to
this model.

4.4. Constitutive laws

Making use of the proposed speci®c free energy function the Cauchy stress tensor
results:

1 
 ˆ …1 dvp †Cs : e evp eth epc ‡ 0 …47†
J

where the thermoelastic and hygrometric terms can easily be identi®ed and, further,
the additive decomposition of the Almansi strain tensor is apparent, that is, ee ˆ
e evp eth epc derived from F ˆ Fe Fp Fth Fpc taking into account the same
arguments commented for the S.G. cast iron described above. Once again, Eq. (47)
is only valid for small elastic strains and, therefore, this restriction is assumed in the
deformation of the sand during the casting process.
The expressions for the speci®c entropy function, the tangent speci®c heat capa-
city, the tangent conjugate of the thermal dilatation tensor, the internal heat source
and the conjugate of the internal variables can be found in Box 2.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1643

Box 2. Variables of the green sand model.

Speci®c entropy function

 ˆ te ‡ tp ‡ pc ‡ 0

with
1 @Cs 1
te ˆ …e evp † : …1 dvp † : …e evp † ‡ …e evp † : …1 dvp †Cs : eth
T1
20 @T 0
1 @Cs th
‡ …e evp † : …1 dvp † : e ‡ Ac
0 @T
1
tp ˆ & vp
0
@Ls 1 @Cs pc 1
pc ˆ fw ‡ …e evp † : …1 dvp † : e ‡ …e evp † : …1 dvp †Cs : epc
T1
@T 0 @T 0

where
 
1 1=3 @ s
eth
T ˆ …1 ath † sth ‡ th …T Tref †
3 @T

and
s h 2=3 i
1  1=3 @pc
epc
T ˆ 1 apc …1 ath † fw 1 1 apc eth
T
3 @T

Tangent speci®c heat capacity

c ˆ cte ‡ ctp ‡ cpc

with
1 @2 C s 2T @Cs
cte ˆ …e evp † : …1 dvp † : …e evp † ‡ …e evp † : …1 dvp †
20 @T2 0 @T
T @eth T @2 Cs
: eth
T1 ‡ …e evp † : …1 dvp †Cs T 1 ‡ …e evp † : …1 dvp †
0 @T 0 @T2
@cs
: eth ‡ cs ‡ …T Tref †
@T

(continued on next page)


1644 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Box 2 (continued)
ctp ˆ 0

@2 L s T @2 Cs pc 2T @Cs
cpc ˆ T fw ‡ …e evp † : …1 dvp † :e ‡ …e evp † : …1 dvp †
@T 0 @T2 0 @T
T @epc
: epc
T1‡ …e evp † : …1 dvp †Cs : T
1
0 @T

Tangent conjugate of the thermal dilatation tensor:

ˆ te ‡ pc

with
1 1 @Cs 
te ˆ …1 dvp †Cs : eth
T1 …1 dvp † : e evp eth
J J @T

1 @Cs pc 1 s pc
pc ˆ …1 dvp † : e ‡ C : eT 1
J @T J

Speci®c internal heat source:

1
rvp
int ˆ …T ‡  † : Lv …evp †
0

@Ls T @Cs pc 1
rpc
int ˆ T ‡ Ls …e evp † : …1 dvp † : e 1 ‡ …e evp †
@T 0 @T f 0
pc 
T s @ef
:
: …1 dvp †Cs : epc
f 1 …e evp † : …1 vp
d †C : 1 fw
0 @T
and
1  1=3 s
epc
f ˆ 1 apc pc
3

Conjugate of the internal variables


1
Yvp ˆ …e evp † : Cs : …e evp † …e evp † : Cs : eth …e evp † : Cs : epc
2
qw ˆ 0 Ls ‡ …e evp † : …1 dvp †Cs : epc
f 1
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1645

4.5. Dissipation inequality

The derivation of the internal dissipation gives:


: :
Dint ˆ  : Lv …evp † ‡ T& vp ‡ Yvp dvp
 
1  1=3 s :
0 Ls …e evp † : …1 dvp †Cs : 1 apc pc 1 f w 50 …48†
30

which clearly shows the viscoplastic and hygrometric contributions. Similar con-
siderations to those commented in Section 3.5 are also valid for the ful®llment of
this dissipation inequality and the transformation work included in the conjugate
variable to fw .

5. Finite element formulation

This section brie¯y describes the ®nite element equations derived from the pro-
posed thermomechanical/microstructural formulation presented above and, further,
some speci®c details of the solution strategy followed to solve the resulting coupled
system of discretized equations.
Assuming standard spatial interpolations for the displacement and temperature
®elds and following the typical procedures within the ®nite element context (Zien-
kiewicz and Taylor, 1989), the global discretized thermomechanical equations (also
including microstructural e€ects) can be written in matrix form for a certain time t
as:

RU  FU ‡ Ff MU F ˆ 0
: :
RT  FT CT KT Lint GU ˆ 0 …49†

where RU and RT are the mechanical and thermal residual vectors, respectively.
Moreover, FU is the external force vector, Ff is the mechanical contact vector, M is
the mass matrix, U is the nodal displacement vector, F denotes the internal force
vector, FT is the external heat ¯ux vector, C is the capacity matrix, T is the nodal
temperature vector, K is the conductivity matrix, Lint is the internal heat ¯ux vector
(which can be decomposed into Lint ˆ Lvp pc
int ‡ Lint according to the de®nition of rint )
and G is the thermoelastic coupling matrix. The element expressions of these matri-
ces and vectors are very similar to those derived by Celentano et al. (1996) in an
in®nitesimal context using di€erent material constitutive models. Nevertheless, the
extension of such expressions to the formulation presented in this work can be
straightforwardly performed.
The integration of the terms containing time derivatives of U and T in system (49)
is carried out with the Newmark method and the generalized mid-point rule algo-
rithm, respectively (Zienkiewicz and Taylor, 1989). The latter has been also used to
1646 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

integrate all the rate equations involved in the S.G. cast iron and green sand models
presented above. Note that in this phenomenological approach the time integration
of both the viscoplastic and phase-change internal evolution laws is performed at the
same integration points.
A detailed description of the mechanical and thermal boundary conditions is also
reported by Celentano et al. (1996). It is important to remark, however, the con-
sideration of a gap and/or pressure dependent heat transfer coecient between the
casting and the mould. This coupling e€ect can be an extremely important fact in
many casting simulations.
The numerical solution of system (49) is achieved using a staggered scheme con-
sidering an improved isothermal split (Celentano et al., 1999). In this context, dif-
ferent ways to solve the mechanical and thermal problems (given by RU and RT ,
respectively) are available. In this work, in particular, the well-known total Lagran-
gean approach under isothermal conditions has been considered for the mechanical
problem while the solution of the thermal problem is done at the spatial con®gura-
tion assuming ®xed elastic con®gurations for each component of the S.G. cast iron
and the sand. This methodology is stable and preserves the coupling degree of the
formulation, i.e., the evaluation of the coupling terms is not shifted in time in order
to obtain ®xed contributions.
Finally, in order to overcome the volumetric locking e€ect on the numerical
solution when incompressible viscoplastic ¯ows are considered, an improved strain-
displacement matrix is proposed in this work. Based on the deformation gradient
multiplicative standard decomposition into deviatoric and volumetric parts, and
assuming a selective (or averaged or smoothed) numerical integration for the volu-
metric part of F, the derived strain-displacement matrix, directly obtained by line-
arization of the Green-Lagrange strain tensor, can be classi®ed within the so-called
B-bar methods originally developed for in®nitesimal strains by Hughes (1980). The
expressions of this matrix for the 2D, axisymmetric and 3D cases are given in Box 3.

Box 3. Strain-displacement matrix B for large strain analysis

For a given node

!2=3
J
B ˆ ‰Bst ‡ Bim Š
J

where:
 
J ˆ det F with F ˆ Fdev F vol such that:
Fdev ˆ J 1=3 F : deviatoric part of F
F vol ˆ J 1=3 1: volumetric part of F numerically integrated in a selective
(or averaged or smoothed) form (continued on next page)
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1647

Box 3 (continued)
Bst : standard strain±displacement matrix for large strain analysis (see Bathe,
1981)
Bim : improved contribution
2 3
C A C11 AV2D
1 4 11 U2D
Bim ˆ C22 AU2D C22 AV2D 5 …2D case†
3
2C12 AU2D 2C12 AV2D

2 3
C11 AUax C11 AVax
16 C 22 AUax C22 AVax 7
Bim ˆ 6 7 …axisymmetric case†
3 4 2C12 AUax 2C12 AVax 5
C33 AUax C33 AVax

2 3
C11 AU C11 AV C11 AW
6 C22 AU C22 AV C22 AW 7
6 7
1 6 2C12 AU 2C12 AV 2C12 AW 7
Bim ˆ 6 7 …3D case†
366 C33 AU C33 AV C33 AW 77
4 2C13 AU 2C13 AV 2C13 AW 5
2C23 AU 2C23 AV 2C23 AW

with

C ˆ FT F: right Cauchy±Green deformation tensor (T : transpose symbol)


1 @J 1 @J
A ˆ  ˆ U; V; W
J @ J @
U; V; W: components of the nodal displacement vector U
A2D ˆ A jF13 ˆF23 ˆF31 ˆF32 ˆ0;F33 ˆ1  ˆ U; V
Aax ˆ A j NU  ˆ U; V (U: radial nodal displacement)
F13 ˆF23 ˆF31 ˆF32 ˆ0;F33 ˆ1‡
X
N: shape function for displacements
X: radial material coordinate

As recently reported by Celentano (1998b), this proposed approach has been suc-
cessfully checked in the analysis of classical large strain benchmark problems.
Moreover, it can be demonstrated that this matrix recovers the corresponding
expression of the in®nitesimal case, based on the additive decomposition of the
strain tensor into deviatoric and volumetric parts, when F ! 1.
1648 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

6. Analysis of a solidi®cation test

The analysis of a cylindrical casting specimen of S.G. cast iron (diameter=70 mm


and height=140 mm) in a green sand mould surrounded by a steel shell (internal
diameter=185 mm, thickness=30 mm and height=260 mm) is performed. This
problem has been extensively studied using simpli®ed in®nitesimal strains con-
stitutive models for the materials involved (Celentano et al., 1995; Celentano, 1997).
The experimental apparatus is schematically shown in Fig. 1. Both temperature and
radial displacement evolutions have been measured during solidi®cation and cooling
approximately at the midheight of the specimen (BRITE/EURAM Synthesis
Report, 1994). Thermocouples were placed on three radial directions at 0, 120 and
240 , starting from the cylinder central axis to the surrounding sand mould in order
to visualize the thermal gradient evolution. Radial displacement were measured at
the same directions on the cylinder external skin using silica rods.

Fig. 1. Solidi®cation test: experimental apparatus.


D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1649

Table 1
Material propertiesa of S.G. cast iron

Young's modulus (MPa):


10 at 1150 C (1) 44,260 at 750 C (a) 500 (g) 163,471 at 20 C (c, f, p)
10 at 1400 C 43,790 at 770 C 163,113 at 100 C
44,360 at 800 C 160,174 at 200 C
45,935 at 830 C 151,650 at 300 C
42,935 at 850 C 135,276 at 400 C
35,435 at 900 C 110,898 at 500 C
28,435 at 1000 C 81,386 at 600  C
5000 at 1100 C 52,021 at 700 C
46,668 at 720  C
Poisson ratio: 0.33 (1, a, g, c, f, p)
Secant thermal dilatation coecient (1 /C):
20.010 6 at 1155 C (1) 13.010 6 at 20 C (a, g, c, f, p)
20.010 6 at 1400 C 12.010 6 at 750 C
12.010 6 at 800 C
17.010 6 at 1000 C
Thermal hardening function (MPa):
0.000 1 (1), 260 at 20 C (a, g, c, f, p)
250 at 200 C
210 at 400 C
60 at 600 C
50 at 700 C
30 at 900 C
20 at 1000 C
2 at 1141 C
0.1 at 1155 C
Viscoplastic parameters:
Viscosity (MPa s): 100.0 (1, a, g, c, f, p)
Exponent: 1.0 (1, a, g, c, f, p)
Solid±solid phase-change parameters:
Viscosity (MPa s): 0.0 (f, p)
Exponent: 1.0 (f, p)
Isotropic hardening modulus (MPa): 0.0 (l, a, g, c), 300 (f, p)
Kinematic hardening modulus (MPa): 0.0 (l, a, g, c, f, p)
Secant phase-change volumetric deformation: 0.01 (a, g, c), 0.005 (f, p)
Density (initial) (kg/m3): 6700 (1)
Secant speci®c heat (J/kg  C):
917 (1) 740 at 732 C (a) 705 (g, c) 540 at 25 C (f, p)
705 at 1136 C 732 at 723 C
917 at 1145 C
Conductivity (J/m s  C):
33 at 1250 C (1) 35 at 750 C 30 (g, c) 27 at 20 C (f, p)
33 at 1300 C 33 at 1140 C 30 at 328 C
92 at 1350 C 30 at 500 C
(continued on next page)
1650 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Table 1 (continued)
186 at 1375 C 25 at 600 C
280 at 1400 C 32 at 700 C
Secant latent heat (J/kg): 263,000 (a), 233,000 (g), 213,000 (c), 16,300 (f) 85,800 (p)
Composition: 3.5%C, 2.0%Si, 0.0% inoculant
Equilibrium temperatures ( C): 1588 ( 0 ), 1179 (1), 1155 (g), 1056 (c), 768 (f), 740 (p)
Partition coecient: 0.46
Nucleation constants:
Parameter (nuclei/m3  C3.61): 7.1210 3
(g, c, f, p)
Exponent: 3.61 (g, c, f, p)
Growth constants:
Parameter (m/s  C2): 1.610 6
(g,f), 1.610 5
(c,p)
Exponent: 2.0 (g, c, f, p)
a
The variations of these properties with the temperature have been assumed to be piecewise linear
within the mentioned temperatures. Below the lowest and above the highest temperatures, the properties
are assumed to remain constant at the same value de®ned for the extreme temperature.

The material thermomechanical properties for the S.G. cast iron and green sand
can be found in BRITE/EURAM Synthesis Report (1994), Hamata (1992) and
Azzouz (1995). The constants involved in the microstructural model for the S.G.
cast iron are equivalent to those used by Chang et al. (1991), Hamata (1992),
Goettsch and Dantzig (1994) and Celentano and Cruchaga (1999). According to
Schmitt at al. (1997), the elastic mechanical properties (Young's modulus and Pois-
son ratio) of cementite, ferrite and pearlite have been assumed to be equal in the
present analysis. For the steel, the same S.G. cast iron model was used considering a
single-component material without, for simplicity, hardening and phase-change
e€ects. All the material properties considered in the simulation are shown in
Tables 1±4.
The axisymmetric numerical computation used 540 four-noded isoparametric ele-
ments and a time step of 50s. The analysis starts with the mould cavity completely
®lled with molten metal at rest at 1250 C (i.e. instantaneous ®lling is assumed) and
22 C for the sand and steel moulds. The mould is simply supported at the bottom
and convection±radiation conditions have been considered between the external face
of the mould and the environment. The boundary conditions and the ®nite element
mesh used are plotted in Fig. 2.
The experimental temperature evolution at di€erent radial positions are plotted in
Fig. 3. The numerical results obtained with the proposed formulation are also
included for comparison. A good overall agreement can be observed where, more
speci®cally, the liquid-solid and solid±solid phase-changes in the casting are reason-
ably well described.
Moreover, the experimental and numerical radial displacement evolutions at the
casting-mould interface are shown in Fig. 4. The di€erent expansion/contraction
behaviours related to the phase-changes occuring during the process can clearly be
seen: (a) contraction till the beginning of the solidi®cation, (b) expansion during
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1651

Table 2
Material propertiesa of green sand

Young's modulus (MPa): 150


Poisson ratio: 0.33
Secant thermal dilatation coecient (1/ C): 13.010 6

Thermal hardening function (MPa): 5.0


Viscoplastic parameters:
Viscosity (MPa s): 0.0
Exponent: 1.0
Isotropic hardening modulus (MPa): 0.0
Secant phase-change volumetric deformation: 0.01
Density (initial) (kg/m3): 1550
Secant speci®c heat (J/kg  C):
917 at 20 C
917 at 200 C
1025 at 400 C
1063 at 700 C
1100 at 1000 C
1100 at 1300 C
Conductivity (J/m s  C):
0.86 at 20 C
0.72 at 200 C
0.52 at 400 C
0.61 at 700 C
0.78 at 1000 C
0.78 at 1300 C
Secant latent heat (J/kg): 3750
Vaporization temperature ( C): 100
Initial free water volumetric fraction: 0.0468
a
The variations of these properties with the temperature have been assumed to be piecewise linear
within the mentioned temperatures. Below the lowest and above the highest temperatures, the properties
are assumed to remain constant at the same value de®ned for the extreme temperature.

solidi®cation (graphite precipitation), (c) contraction from the end of the solidi®ca-
tion up to the beginning of the eutectoid transformation, (d) second expansion dur-
ing the eutectoid transformation and (e) ®nal contraction to room temperature.
Almost identical behaviours have been observed for the three directions mentioned
above. Some dispersion in the experimental results was detected, however, and an
average curve has been included in Fig. 4. The diculties in the measurement tasks
may be presumably attributed to di€erential vertical dilatation between the casting
and sand mould (see Fig. 5) leading to a unacceptable shear force in the silica rods.
Although the numerical ®tting is only qualitative, the response provided by the S.G.
1652 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Table 3
Material properties of steel

Young`s modulus (MPa): 210000


Poisson ratio: 0.3
Secant thermal dilatation coecient (1/ C): 12.010 6

Thermal hardening function (MPa): 210


Isotropic hardening modulus (MPa): 0.0
Kinematic hardening modulus (MPa): 0.0
Density (initial) (kg/m3): 7900

Secant speci®c heat (J/kg  C): 490


Conductivity (J/m s  C): 35

Table 4
Properties at the interfacesa

Heat transfer coecient at the casting-sand interface (J/m2 s  C):


1000 at gn=0.000 m (gn=normal gap)
100 at gn=0.001 m
Heat transfer coecient at the sand±steel interface (J/m2 s  C): 2000
Heat transfer coecient at the steel±air interface (J/m2 s  C): 20
Environmental temperature ( C): 20
Normal asperity at the casting±sand and sand±steel interfaces (MPa/m): 10.010 (high value to avoid
penetration between bodies, see Celentano et al., 1996)
a
The variation of the heat transfer coecient with the normal gap has been assumed to be piecewise
linear within the mentioned normal gaps. Below the lowest and above the highest normal gaps, the
properties are assumed to remain constant at the same value de®ned for the extreme normal gap.

model proposed in this work correctly reproduces the distinct behaviours observed
at di€erent stages of the process. Further, the di€erence between the displacement
curves corresponding to the casting and sand gives the normal gap evolution which,
as mentioned in Section 5, usually a€ects the heat transfer conditions at the casting-
sand interface. This last e€ect can be also seen in Fig. 5 which depicts the deformed
con®gurations at four times of the analysis.
The volumetric fraction evolutions of the casting for two radial distances at the
midheight of the specimen are plotted in Figs. 6 and 7. The liquid±solid and solid±
solid phase-changes are apparent. It should be mentioned that a low cooling rate has
been obtained promoting the formation of graphite at the expense of cementite and,
moreover, completely inhibiting the development of pearlite since all the austenite is
transformed into ferrite.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1653

Fig. 2. Solidi®cation test: (a) boundary conditions and (b) ®nite element mesh.

Fig. 3. Solidi®cation test: temperature evolutions for di€erent radial positions at the midheight of the
specimen.
1654 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Finally, Fig. 8 shows the free water volumetric fraction evolution for two radial
positions in the sand mould at the same height mentioned above. As expected, the
vaporization starts in regions near the casting-mould interface and progressively
advances towards the outer surface.

Fig. 4. Solidi®cation test: radical displacement evolutions at the midheight of the casting±sand interface.

Fig. 5. Solidi®cation test: deforned con®gurations at times (a) 200 s; (b) 500 s; (c) 1000 s and (d) 3500 s
(ampli®cation factor=10).
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1655

Fig. 6. Solidi®cation test: volumetric fractions in the casting for radius=0 mm at the midheight of the
specimen.

Fig. 7. Solidi®cation test: volumetric fractions in the casting for radius=30 mm at the midheight of the
specimen.
1656 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Fig. 8. Solidi®cation test: free water volumetric fraction in the sand for r=45 mm and r=75 mm at the
midheight of the specimen.

7. Conclusions

A large strain thermoviscoplastic formulation for the analysis of the solidi®cation


process of S.G. cast iron in green sand moulds have been presented. This formula-
tion accounts for thermomechanical as well as microstructural behaviours of these
materials in an uni®ed framework allowing, therefore, to analyze the di€erent
coupled phenomena occuring in complex casting problems. Several original aspects
of such formulation have been discussed. For the S.G. cast iron model, in parti-
cular, the incorporation of latent heat release and volumetric expansions due to all
metallurgical transformations existing during the solidi®cation and cooling pro-
cesses are relevant features of the presented formulation. On the other hand, the
consideration of the hygrometric and damage phenomena stands out in the green
sand model.
The corresponding ®nite element model has been also derived and brie¯y pre-
sented. Some strategies to achieve the numerical solution have been also proposed.
This formulation has been used in the analysis of a solidi®cation test of S.G. cast
iron in a green sand mould. The model has been partially validated with some
available experimental measurements where reasonable agreement between numer-
ical and experimental results can be observed. However, the diculties associated to
the full material characterization lead to a further research in the thermomechanical/
microstructural simulation of solidi®cation processes with the sake of constituting a
robust tool for casting design.
D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658 1657

Acknowledgements

The support provided by CONICYT (FONDECYT Project No. 1990588) and


DICYT-USACH are gratefully acknowledged.

References
Agelet de Saracibar, C., Cervera, M., Chiumenti, M., 1999. On the formulation of coupled thermoplastic
problems with phase-change. Int. J. Plasticity 15, 1.
Ami Saada, R., Bonnet, G., Bouvard, D., 1996. Thermomechanical behavior of casting sands: experi-
ments and elastoplastic modeling. Int. J. Plasticity 12, 273.
Armero, F., Simo, J., 1993. A priori stability estimates and unconditionally stable product formula algo-
rithms for non-linear coupled thermoplasticity. Int. J. Plasticity 9, 149.
Arnold, S., Saleeb, A., 1994. On the thermodynamic framework of generalized coupled therrnoelastic-
viscoplastic-damage modeling. Int. J. Plasticity 10, 263.
Azzouz, F., 1995. Modelisation du comportement thermo-hydro-mecanique d'un sable de fonderie, TheÁse
de Doctorat, Universite Paris 6, France.
Baldoni, F., Rajagopal, K., 1997. A continuum theory for the thermomechanics of solidi®cation. Int. J.
Non-Linear Mechanics 32, 3.
Bammann, D., Chiesa, M., Johnson, G., 1995. A state variable damage model for temperature and strain
rate dependent metals. In: Rajendran, A., Batra, R. (Eds.), Constitutive Laws: Theory, Experiments
and Numerical Implementation. CIMNE, pp. 84±95.
Banerjee, D., Stefanescu, D., 1991. Structural transitions and solidi®cation kinetics of SG cast iron during
directional solidi®cation experiments. AFS Transactions 104, 747.
Bathe, K., 1981. Finite Element Procedures in Engineering Analysis. Prentice-Hall.
Bellet, M., Decultieux, F., MeÂnal, M., Bay, F., Levaillant, C., Chenot, J., Schmidt, P., Svensson, I., 1996.
Thermomechanics of the cooling stage in casting processes: three-dimensional ®nite element analysis
and experimental validation. Metall. Trans. B 27B, 81.
BRITE/EURAM Synthesis Report, 1994. Development of a user-oriented CAB system for simulating the
forming of ductile iron parts (Project No. BE-4596).
Casey, J., 1998. On elastic-thermo-plastic materials at ®nite deformations. Int. J. Plasticity 14, 173.
Cassenti, B., Annigeri, B., 1989. Thermodynamic constraints on stress rate formulations in constitutive
models. Comput. Mech. 4, 429.
Celentano, D., 1997. A thermomechanical formulation for the solidi®cation process of S .G. cast iron in a
green sand mould. Proceedings of COMPLAS V, p. 1284.
Celentano, D., 1998a. Shrinkage prediction in the thermomechanical analysis of castings. In: Proceedings
of the VIIIth Engineering Foundation Conference on Modeling of Casting, Welding and Advanced
Solidi®cation Processes, The Minerals, Metals and Materials Society, London.
Celentano, D., 1998b. VULCAN: coupled thermomechanical ®nite element analysis for solidi®cation
problems. User's and Veri®cation Manuals (version 2.3).
Celentano, D., Cruchaga, M., 1999. A thermally coupled ¯ow formulation with microstructural evolution
for hypoeutectic cast-iron solidi®cation. Metall. Trans. B 30B, 731.
Celentano, D., Gunasegaram, D., Nguyen, T., 1999. A thermomechanical model for the analysis of light
alloy solidi®cation in a composite mould. Int. J. Solids Struct. 36, 2341.
Celentano, D., Oller, S., OnÄate, E., 1996. A coupled thermomechanical model for the solidi®cation of cast
metals. Int. J. Solids Struct. 33, 647.
Celentano, D., Visconte, D., Dardati, P., Oller, S., OnÄate, E., 1995. A thermomechanical model for
solidi®cation problems: experimental validation. Proceedings of COMPLAS IV, p. 2385.
Coleman, B., Gurtin, M., 1967. Thermodynamics with internal state variables. The Journal of Chemical
Physics 47, No. 2.
Chang, S., Shangguan, D., Stefanescu, D., 1991. Prediction of microstructural evolution in SG cast iron
from solidi®cation to room temperature. AFS Transactions 99, 531.
1658 D.J. Celentano / International Journal of Plasticity 17 (2001) 1623±1658

Cherkaoui, M., Berveiller, M., Sabar, H., 1998. Micromechanical modeling of martensitic transformation
induced plasticity (TRIP) in austenite single crystals. Int. J. Plasticity 14, 597.
Doyle, T., Ericksen, J., 1956. Nonlinear elasticity. Advances in Applied Mechanics 4.
Fischer, F., Sun, Q., Tanaka, K., 1996. Transformation-induced plasticity (TRIP). Appl. Mech. Rev. 49,
317.
Flemings, M., 1974. Solidi®cation Processing. McGraw-Hill, New York.
Garcia Garino, C., Oliver, J., 1992. A numerical model for elastoplastic large strain problems funda-
mentals and applications. Proceedings of COMPLAS II, p. 117.
Garcia Garino, C., 1993. A numerical model for the analysis of elastoplastic solids subjected to large
deformations. PhD thesis (in Spanish), Universidad PoliteÂcnica de CatalunÄa, Spain.
Goettsch, D., Dantzig, J., 1994. Modeling microstructure development in gray cast irons. Metall. Trans.
A 25A, 1063.
Green, A., Naghdi, P., 1971. Some remarks on elastic-plastic deformation at ®nite strains. International
Journal of Engineering Sciences 9, 1219.
Hamata, N., 1992. Modelisation du couplage entre l'elasto-viscoplasticite anisotherme et la transfor-
mation de phase d'une fonte G.S. ferritique. TheÁse de Doctorat, Universite Paris 6, France.
Hughes, T., 1980. Generalization of selective integration procedures to anisotropic and nonlinear media.
Short Communications, 1413.
Inoue, T., Ju, D., 1992. Simulation of solidi®cation and viscoplastic stresses during vertical semicontinuous
direct chill casting of aluminium alloy. Int. J. Plasticity 8, 161.
Kleiber, M., 1975. Kinematics of deformation processes in material subjected to ®nite elastic-plastic
strains. International Journal of Engineering Sciences 13, 155.
Levitas, V., 1996. Large Deformation of Materials with Complex Rheological Properties at Normal and
High Pressure. Nova Science, New York.
Levitas, V., 1998. Thermomechanical theory of martensitic phase transformation in inelastic materials.
Int. J. Solids Struct. 35, 889.
Levitas, V., Idesmas, A., Stein, E., 1998. Finite element formulation of martensitic phase transitions in
elastoplastic materials. Int. J. Solids Struct. 35, 855.
Lubliner, J., 1985. Thermomechanics of Deformable Bodies. Department of Civil Engineering, University
of California, Berkeley.
Lubliner, J., 1990. Plasticity Theory. Macmillan, New York.
Malvern, L., 1969. Introduction to the Mechanics of a Continuous Medium. Prentice-Hall, Englewood
Cli€s.
Marsden, J., Hughes, T., 1983. Mathematical Foundations of Elasticity. Prentice Hall, Englewood Cli€s.
Murnagaham, F., 1937. Finite deformations of an elastic solid. Am. J. Math. 59, 235.
Onsùien, M., Grong, é., Gundersen, é., Skaland, T., 1999. A process model for the microstructure
evolution in ductile iron: Part I. The model. Metall. Trans. A 30A, 1053.
Perzina, P., 1971. Thermodynamic theory of viscoplasticity. Advances Appl. Mech. 9, 243.
Schmitt, C., Lipinski, P., Berveiller, M., 1997. Micromechanical modelling of the elastoplastic behavior of
polycrystals containing precipitates-application to hypo- and hyper-eutectoid steels. Int. J. Plasticity 13,
183.
Simo, J., Ortiz, M., 1985. A uni®ed approach to ®nite deformation elastoplastic analysis based on the use
of hyperelastic constitutive equations. Computer Methods in Applied Mechanics and Engineering 49,
221.
Trovant, M., Argyropoulos, S., 1996. Mathematical modeling and experimental measurements of
shrinkage in the casting of metals. Canadian Metallurgical Quaterly 35, 75.
Wriggers, P., Miehe, C., Kleiber, M., Simo, J., 1989. On the coupled thermomechanical treatment of
necking problems via ®nite-element-methods. Proceedings of COMPLAS II, p. 527.
Zabaras, R., Ruan, Y., Richmod, O., 1990. Front tracking thermomechanical for hypoelastic-viscoplastic
behavior in a solidifying body. Computer Methods in Applied Mechanics and Engineering 81, 333.
Zienkiewicz, O., Taylor, R., 1989. The ®nite element method, 4th Edition. McGraw-Hill, London (Vols. 1
and 2).

You might also like