Direct Chill Casting of Al Alloys PDF

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7
At a glance
Powered by AI
The document discusses modeling and experiments on direct chill casting of aluminum alloys. It compares a continuum mixture model to experimental results from industrial billets to simulate the effect of grain refiners on macrosegregation and fluid flow.

Macrosegregation is the uncontrolled redistribution of alloying elements during solidification on the scale of the ingot. In DC cast ingots it produces nonuniform properties and distinct regions of positive and negative segregation from the surface inward. Enrichment near the surface is due to shrinkage flow and exudation while normal behavior occurs after solidification.

Experimental studies have proposed that a duplex microstructure and negative segregation at the center results from settling of solute-depleted free-floating dendrites formed near the top of the sump that are swept to the bottom by convection currents, leaving positive titanium segregation as evidence. This has been observed in other studies.

Direct Chill Casting of Aluminum

Christopher J. Vreeman
Capstone Turbine Corporation,
21211 Nordhoff Street,
Alloys: Modeling and Experiments
Chatsworth, CA 91311

J. David Schloz
on Industrial Scale Ingots
Wagstaff Engineering, A continuum mixture model of the direct chill casting process is compared to experimental
3910 North Flora Road, results from industrial scale aluminum billets. The model, which includes the transport of
Spokane, WA 99216 free-floating solid particles, can simulate the effect of a grain refiner on macrosegregation
and fluid flow. It is applied to an Al-6 wt% Cu alloy and the effect of grain refiner on
Matthew John M. Krane macrosegregation, sump profile, and temperature fields are presented. Two 45 cm diam-
Member, ASME eter billets were cast under production conditions with and without grain refiner. Tem-
Assistant Professor perature and composition measurements and sump profiles are compared to the numerical
School of Materials Engineering, results. The comparison shows some agreement for the grain refined case. It is believed
Purdue University, that an incorrect assumption about the actual grain structure prevents good agreement in
West Lafayette, IN 47907 the non-grain refined billet. 关DOI: 10.1115/1.1482089兴

Keywords: Heat Transfer, Manufacturing, Mass Transfer, Processing, Solidification

Introduction positive titanium segregation at the centerline 关2兴, which is present


in grain refining particles that serve as nucleation sites for free-
Direct chill casting is a semi-continuous process in which mol-
floating dendrites. Duplex microstructures accompanied by the
ten aluminum enters the top of a water-cooled mold, is cooled,
negative centerline segregation have also been observed by Chu
and forms a solid ingot drawn out below 共Fig. 1共a兲兲. The molten
and Jacoby 关3兴 and Dorward and Beerntsen 关4兴, and negative cen-
aluminum begins to solidify due to the cooling effect of the mold
terline segregation with positive segregation of titanium was ob-
wall. As the ingot is withdrawn from the bottom of the mold,
water jets impinge on the surface, directly cooling the billet and served in ingots cast by Gariepy and Caron 关5兴.
forming a solid layer around a liquid melt, commonly referred to To model the transport phenomena which control the segrega-
as the sump. tion in DC casting, single domain models have been used that
A particular defect that occurs in a wide range of casting pro- implicitly couple phenomena in the melt, mushy zone, and solid
cesses, including the direct chill 共DC兲 casting of aluminum alloys, regions of a solidifying ingot. Flood et al. 关6兴 were the first to
is macrosegregation, the uncontrolled redistribution of alloying apply a single domain model for the transport phenomena in DC
elements on the scale of the ingot during solidification. Macroseg- casting to predict macrosegregation in Al-4.5 wt% Cu billets.
regation in DC cast ingots can produce nonuniform mechanical They assumed that the mushy zone consisted only of a rigid,
properties 关1兴, which affects the behavior of the metal during permeable dendritic matrix with the solid moving at the casting
downstream forming and heat treat operations. speed and predicted positive segregation of copper at the center of
A commonly observed, surface-to-surface distribution of alloy- the billet. Reddy and Beckermann 关7兴 included the effect of free-
ing elements at a transverse cross-section of a DC cast ingot re- floating dendrites of varying diameter by employing a modified
veals distinct regions of positive 共solute-rich兲 and negative version of the two-phase model developed by Ni and Beckermann
共solute-depleted兲 segregation 关2兴. Enrichment near the ingot sur- 关8兴 to simulate the DC casting of two Al-4.5 wt% Cu billets. In
face is attributed to a combination of shrinkage induced flow of one case, the solid phase was assumed to form a rigid structure
solute-rich liquid toward the mold wall, where solidification rates moving at the casting speed. In the other case, the transport of
are highest, and exudation induced by local remelting of the ingot solute-depleted, free-floating dendrites was modeled assuming the
surface as the solid shell pulls away from the mold. Once a struc- solid phase formed a rigid structure at a solid volume fraction of
turally sound ingot surface has formed and solidification rates 0.637. In the simulation with no free-floating solid, they predicted
have decreased, more normal segregation behavior is observed, subsurface enrichment, followed by a narrow depleted region just
with the solute composition increasing with decreasing radius within the surface. Positive segregation occurred at the centerline
from the solute-depleted region near the ingot surface. While this and was attributed to the buoyancy driven flow of copper-rich
trend may be expected to continue, ingots produced by DC casting liquid to the bottom of the sump. In the simulation with free-
often exhibit a significant decrease in solute composition near the floating dendrites, they predicted significant negative segregation
centerline. at the centerline, which they attributed to the advection of copper-
Experimental studies have provided some insight into the depleted dendrites to the bottom of the sump. Reddy and Becker-
mechanisms that control macrosegregation in DC cast ingots, es- mann were the first to include the transport of free-floating den-
pecially at the centerline. Yu and Granger 关2兴 investigated nega- drites into a fully coupled, single domain, macrosegregation
tive segregation in Al-Cu-Mg slabs and concluded that a duplex model of the DC casting process and to predict the resulting nega-
microstructure and accompanying negative segregation at the cen- tive segregation at the centerline. More recently, the same authors
ter of DC cast ingots resulted from the settling of solute-depleted, 关9兴 conducted further DC casting simulations of Al-4.5 wt% Cu
free-floating dendrites that formed near the top of the sump and billets to examine the effect of the grain density and interdendritic
were subsequently swept to the bottom of the sump by convection fluid flow on macrosegregation in billets devoid of free-floating
currents in the melt. This theory was also supported by significant dendrites. They demonstrated that, if the mushy zone consists of
a rigid dendritic structure without free-floating dendrites, its per-
Contributed by the Heat Transfer Division for publication in the JOURNAL OF
meability significantly influences macrosegregation in DC cast
HEAT TRANSFER. Manuscript received by the Heat Transfer Division September 12, billets.
2001; revision received March 26, 2002. Associate Editor: C. Amon. Vreeman et al. 关10兴 developed a fully coupled, single-domain,

Journal of Heat Transfer Copyright © 2002 by ASME OCTOBER 2002, Vol. 124 Õ 947

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Numerical Modeling
The numerical model is that of Vreeman et al. 关10兴, which
simulates solidification in systems with two-phase 共solid-liquid兲
regions consisting of a slurry of free-floating dendrites above a
rigid, dendritic matrix saturated with interdendritic liquid. Mixture
transport equations for mass, energy, and species conservation
关12兴 are applied in all regions of the DC cast billet as it solidifies,
while the application of separate and distinct momentum equa-
tions depends on the local solid morphology 关10兴. The liquid
metal velocities calculated in previous papers under very similar
conditions were low enough to justify the assumption that the flow
can be treated as laminar. In the liquid and slurry regions, for
which the solid volume fraction is less than the designated pack-
ing fraction (g s ⬍g s,p ), modified versions of slurry momentum
equations developed by Ni and Incropera 关13兴 are employed
共Table 1兲. In regions consisting of a rigid, solid structure and
interdendritic liquid (g s ⭓g s,p ), momentum equations developed
by Bennon and Incropera 关12兴 and Prescott et al. 关14兴 are em-
ployed to account for interdendritic fluid flow in a translating solid
matrix.
The model used here to predict solid particle motion relative to
the liquid flow is much simpler than some found in the literature.
Wang and Beckermann 关15兴 have developed a detailed multiphase
model to account for the presence of free-floating equiaxed den-
drites and their effect on composition and grain structure, linking
the model to the macroscopic equations governing thermosolutal
convection and heat and mass transfer. Calculations have been
made, some of which were compared to experiments in salt-water
solutions 关16,17兴.
The parameter that defines the relative extent of the slurry and
rigid mushy zone regions is the packing fraction, g s,p , at which
the free-floating dendrites coalesce to form a rigid dendritic struc-
ture. In general, packing fractions are not explicitly known and
vary significantly from system to system. In addition to the pack-
ing fraction, a characteristic diameter of free-floating dendrites, d,
used to calculate solid and liquid phase velocity differences, is not
explicitly known 关10兴. In this study, d⫽75 ␮ m was used. Lumped
Fig. 1 „a… Schematic of generic DC casting process; „b… into this single, unchanging parameter is the real particle size
AirSlip™ mold cross-section distribution which varies in time and position during solidifica-
tion. A rigorous determination of dendrite size variations requires
implementation of an additional transport equation and a nucle-
ation model to account for the advection and ‘‘generation’’
共through nucleation and/or fragmentation兲 of dendrites, respec-
mixture model of the DC casting process which accounted for the tively. An example of such a model is found in 关7兴.
transport of free-floating dendrites. Separate and distinct mixture To model the solid-liquid interactions in the rigid mushy zone,
the permeability, K, is assumed to be isotropic and is evaluated
momentum equations were employed to account for the different
momentum transfer mechanisms in two-phase regions character- using the Blake-Kozeny expression, K⫽ ␬ o (1⫺g s ) 3 /g s2 . Evalua-
ized by a slurry of free-floating dendrites and a rigid, permeable tion of the permeability coefficient, ␬ o for an equiaxed mushy
dendritic matrix saturated with interdendritic liquid. In a compan- zone relies on empirical data for flow through granular materials
关18兴. In this study, a permeability coefficient of ␬ o ⫽3.75
ion paper 关11兴, the model was used to predict macrosegregation
distributions in Al-4.5 wt% Cu and Al-6.0 wt% Mg billets and the ⫻10⫺11 m2 is used. This value is based on a dendritic arm spac-
ing of 75 ␮m, which is in the range of measured values found in
predicted surface-to-centerline distribution of macrosegregation
DC cast billets. The thermophysical properties used in the study
was found to be consistent with observations in DC cast ingots
are shown in Table 2 and are based on values found in 关19兴.
discussed above. Negative segregation at the centerline increased
The model is closed by assuming thermodynamic equilibrium
with an increase in the packing fraction at which free-floating on the scale of individual control volumes and using the Al-Cu
dendrites are presumed to coalesce into a rigid dendritic structure. phase diagram, with constant equilibrium partition coefficients
Likewise, negative segregation at the centerline and positive seg- 共Table 2兲 and thermodynamic relations developed by Bennon
regation in the enriched region increased with an increase in the and Incropera 关20兴 to account for both primary and eutectic
characteristic diameter of the free-floating dendrites. solidification.
The objective of this study is to make a direct comparison of In order to obtain a steady-state solution, the energy equation
Vreeman’s model to data taken from an industrial scale direct chill was solved by itself first, with the resulting solid fraction and
process. Two billets of an Al-6wt%Cu alloy were cast in the temperature distribution used as initial conditions for further cal-
foundry at Wagstaff, Inc., one with and one without the use of a culations. Recognizing that converged solutions could not be ob-
grain refiner to generate free-floating solid particles in the sump. tained without the time-dependent terms in the conservation equa-
Measured composition profile, temperature history, and sump pro- tions, transient simulations of the full model were marched
file results for the grain refined and non-grain refined billets are forward in time with convergence rigorously enforced at each
compared to the numerical results with and without free-floating time step. While a steady state solution was obtained for the billet
solids. without free-floating dendrites (g s,p ⫽0), attempts to obtain

948 Õ Vol. 124, OCTOBER 2002 Transactions of the ASME

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Table 1 Mixture model transport equations †10‡

steady state numerical solutions with free-floating dendrites discussion of the numerical techniques used is found in 关22兴. The
(g s, p ⬎0) were unsuccessful because of the unstable nature of the solution was carried out as described further in 关11兴, with the
interaction of the fluid flow and the solid in the region of g s,p 共see exception of the thermal boundary condition. Instead of using the
关11兴 for details兲. As the simulations proceeded forward in time, a correlations found in 关11兴, the temperature of the surface of the
quasi-steady state was reached where transient variations in the billet was determined by the measurements described below.
dependent variables reached a minimum range.
The numerical solution of the conservation equations 共Table 1兲
is implemented through the use of a control volume formulation Experimental Methods
and the SIMPLER algorithm 关21兴. The procedures outlined by To understand the behavior of these billets during an actual
Bennon and Incropera 关20兴 are used with upwind differencing to casting process and to evaluate the performance of Vreeman’s
discretize all terms. Modifications are made to the differencing model, two 0.45 m 共18 in兲 diameter Al-6wt%Cu billets were cast
procedure of the so-called advection-like source terms in the spe- at Wagstaff’s foundry in Spokane, Washington, using the Wagstaff
cies equation. For these terms, the discretization procedure was Airslip™ process 共see Fig. 1共b兲兲. The alloy was induction melted
modified to insure that mass is conserved when both solid and and degassed with 100 percent argon, then transferred from the
liquid simultaneously cross control volume interfaces. A complete tilting furnace via a heated refractory trough to the mold table.

Table 2 Thermophysical properties and phase diagram data †11‡

Journal of Heat Transfer OCTOBER 2002, Vol. 124 Õ 949

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


The AirSlip™ mold technology is a ‘‘hot-top’’ design with a
metal inlet diameter of 0.305 m 共12 in兲. A unique feature of this
technology is the use of a porous graphite ‘‘casting ring’’ that is
fed with both dry air and a synthetic lubricant. This design pro-
vides minimal surface macrosegregation by reducing the degree of
chill obtained within the mold bore. A specially designed cooling
water array provides the necessary upward conductance and high
heat transfer rate to solidify both shell and bulk of the billet.
The procedure for casting was to transfer the alloy by tilting the
induction hearth, allowing the metal to fill a cavity comprised of
the casting mold and a steel starter block. Following a prescribed
hold time, the starter block was slowly withdrawn from the mold.
The casting speed was 1 mm/s. Heat transfer for the initial 0.02 m
共0.8 in.兲 of cast length is obtained by conduction from the molten
metal through the starter block to the casting water. After this
period the water then impinges on the billet surface and provides
a much higher rate of cooling.
Sump depths were measured during the start phase by briefly
inserting an aluminum rod into the liquid through the hot top and
marking the level of the free surface of the Al alloy. When the
sump depth became constant, the solidification was judged to be
steady state and a thermocouple rake was inserted into the sump.
Three thermocouples were mounted on this rake and positioned
near the centerline, the mid-radius, and the surface of the billet.
First, a thin guide wire was introduced to provide a support for the
rake. The thermocouple rake was then introduced and allowed to
stabilize at the casting temperature. Following temperature stabi-
lization the rake was ‘‘captured’’ by locating the rake to the upper-
and-outermost area of the mold cavity and clamping the rake to
the guide wire. The rake then moved at the casting speed and
recorded the three vertical temperature profiles as it descended.
Due to the inherent inaccuracy when placing thermocouples, the
position of each thermocouple was established by post-cast sec-
tioning of the billet and machining to the thermocouple tips. These Fig. 2 Mixture copper composition, streamlines, and solid
final positions were found to be at r⫽10 mm, 106 mm, and 220 fraction for billet with grain refiner, using experimental bound-
mm, the last of these being only a few millimeters from the sur- ary conditions
face. The uncertainty in the thermocouple measurements was es-
timated to be ⫾2°C.
Grain refiners are usually added to DC cast alloys in order to
produce a more uniform grain structure in the billet. Grain refine- measurements were made using X-ray fluorescence on 20
ment is normally accomplished by inoculation with ‘‘seed’’ grains samples, taken from each billet at 10 different radial positions. At
and growth rate restriction. The former effect is provided by ad- each of the radial positions, the two samples measured were from
different axial locations in the region of the billet apparently at
dition of either TiB2 , Tix Cy , or Al3 Ti, while the latter is provided
steady-state.
by adding solute elements such as titanium that provide constitu-
tional undercooling. Relative uniformity in grain size and equi-
axed shape significantly reduces the anisotropy in the mechanical
properties due to columnar grain growth that occurs without a Results and Discussion
grain refiner. The less oriented grain structure increases the ease Predictions were made of the steady state transport phenomena
with which these billets are rolled or extruded after casting. Two in the casting of the two ingots described above. The numerical
ingots were cast for this study, identical in every respect except results for streamlines, composition, and fraction solid for the
the addition of a grain refinement process to one of them and a grain refined case are shown in Fig. 2. The streamlines show the
slight variation in composition 共grain refined billet⫽6.1 wt% Cu, flow being drawn from the constricted inlet almost horizontally
nongrain refined billet⫽5.9 wt% Cu兲. The grain refiner was a towards the location of jet impingement on the surface. Along
combination of 0.02wt% Ti 共added to restrain growth rate兲 and 6 with the enrichment of the liquid with copper as solidification
ppm B 共added for TiB2 inoculation兲 as Al-3% Ti-1%B rod. This proceeds, the high rate of heat extraction under the jets is the
alloy was introduced by continuously feeding a rod into the engine which drives the buoyancy induced flow in the sump. The
trough. flow accelerates as it approaches and passes the chill, and is
Sump profiles were obtained by two separate methods. In the turned back into the billet by the formation of the rigid solid
case of grain-refined billet, this was accomplished by sudden ad- matrix. The flow then races down the interface of the rigid solid
dition of a molten Al-Si mixture to the hot top of the caster during and the slurry region. As the fluid runs down towards the center-
steady state casting. For the nongrain-refined billet, the Al-Si mix- line, it is continuously entrained into the rigid mushy zone.
ture was replaced by molten Al-3 percent Ti-1 percentB grain In these calculations, the packing fraction is set to g s, P ⫽0.3,
refiner rod. These mixtures were chosen over more traditional which is based on observations in salt solutions and the results of
methods 共e.g., zinc兲 due to the minimal disturbance they cause in numerical studies found in 关11兴 and 关23兴. These works also dem-
the solidification profile. The macrograph in Fig. 4 shows an ex- onstrate the sensitivity of the simulation results to the choice of
ample of an Al-Si sump in a grain-refined billet. The sample was packing fraction. A slurry of solid particles is formed and carried
prepared by slicing the billet in half, polishing the exposed sur- along with the flow down the rigid interface. Because the particles
face, and etching it with a solution of 10 percent HNO3 , 10 per- are formed at temperatures just below the liquidus, they are sig-
cent HCl, and 0.1 percent HF in deionized water. Composition nificantly depleted in copper. The movement of these particles

950 Õ Vol. 124, OCTOBER 2002 Transactions of the ASME

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 4 Sump profiles: „a… Grain refined billet; and „b… Compari-
son of predictions and experiments.

colder liquid rushing underneath it, but most of the heat extracted
from the ingot is taken from regions closer to the surface. The
effect of this region is seen in the temperature contours in Fig.
3共b兲. Some correspondence can be found between the end of the
shallow gradients in Fig. 3共a兲 and the predicted beginning of the
rigid mushy zone where the fluid quickly assumes the casting
speed. Both Figs. 2 and 3共a兲 present some evidence for double-
diffusive flow cells in the sump, each with relatively uniform
composition and temperature.
An etched macrograph of the grain refined billet is compared to
predictions in Fig. 4. In Fig. 4共a兲, lines representing the solidus
interface (g S ⫽1.0) and the transition between the rigid, packed
Fig. 3 Results with grain refiner: „a… Comparison of predicted solid and the slurry (g S ⫽g S, P ) are seen. These lines are repro-
and measured temperature profiles, and „b… Calculated tem-
duced in Fig. 4共b兲 and compared to calculations. The total sump
perature and fraction solid contours.
depth is only overpredicted by approximately 5 percent, while the
actual rigid mushy zone thickness at the surface is twice the cal-
culated value near the surface. The general shape of the sump is
from where they form 共near the outer radius兲 to where they finally correct, but the profiles are steeper near the center and shallower
join the rigid mushy zone 共closer to the centerline兲 results in a near the surface than predicted. This result suggests that the rigid
region in the center which is deficient in copper. mushy zone is thicker than predicted and, given the agreements in
A comparison of vertical temperature profiles for experiments the temperature profiles, perhaps the packing fraction is slightly
and simulations is found in Fig. 3共a兲. At all three radial positions, lower than assumed (g S ⫽0.20⫺0.25?). Near the surface, it is
the temperature decreases in the melt due to conduction towards possible that the packing fraction was even lower, due to the more
the chill until the liquidus temperature is reached. At that point, horizontal surface, which would slow the flow and allow more
solid begins to nucleate on the grain refiner and a solid-liquid time for particles to settle.
slurry forms. Near the surface, the large heat extraction rate so- The radial composition profile is shown in Fig. 5. Near the
lidifies the alloy fast enough that the particles pack quickly and a centerline, the measured composition was up to 10 percent below
rigid structure forms. The other two profiles behave somewhat the nominal value for the alloy, while numerical results predicted
differently, with a sudden shallowing of the temperature gradient less than half that level of segregation there. The study of the
around the liquidus temperature. This shallow gradient begins at effect of packing fraction on centerline segregation found in 关11兴
roughly 0.06 m from the inlet and continues to approximately 0.12 suggests that this lower composition at the centerline could be an
m and 0.22 m for the midradius and centerline profiles, respec- indication that the actual packing fraction was picked to be too
tively. At those points the temperature gradient became markedly low a value. However, the results for the sump temperature pro-
steeper. The reason for this behavior is found in the flow patterns files discussed above suggest otherwise. The calculated values in-
described above. The liquid metal is chilled by the impinging jets creased as the radius increased until a peak was reached around
and much of the metal runs down the interface at which the solid r⫽0.15 m. The experiments also increased, but reached a plateau
packs until it is entrained into the rigid mushy zone. This flow around r⫽0.07 m. Figure 4 shows that the observed sump profiles
pattern leaves a triangular zone of weak recirculation in the center are less steep than in the simulation near the billet surface, sug-
of the sump around which most of the fluid flows. This ‘‘dead gesting a greater tendency for solid particles to settle there than
zone’’ does contain some solid particles, as it is cooled by the was predicted. As the radius decreased, the situation was reversed

Journal of Heat Transfer OCTOBER 2002, Vol. 124 Õ 951

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Fig. 5 Comparison of measured and calculated composition
Fig. 7 Comparison of measured and calculated composition
profiles for grain refined billet
profiles for billet with no grain refiner

共predicted sump profiles were less steep than in the real billet兲, so
copper-depleted particles are more likely to end up at the bottom
of the sump at the centerline. It was noted above that the diameter would be about 0.1 wt percent in Fig. 5, which would not account
of the free-floating solid particles was assumed to be fixed at 75 for the discrepancy between the predictions and the measure-
␮m, which was picked based on typical grain sizes in the actual ments. It should also be noted that there are some measured com-
ingot. Using only this average leads to some uncertainty in the
position variations at given radial positions. These data, taken at
results. It should be noted that, in Vreeman and Incropera 关11兴, the
different axial locations, show that the composition did vary along
difference in the level of centerline macrosegregation with a varia-
tion of diameter between 25 and 100 ␮m was found to be approxi- the axis of the billet and call into question the assumption that this
mately 2 percent of the nominal composition. Such a difference process truly reaches a steady-state in the flow and composition
fields.
To understand how well the model simulates the transport phe-
nomena in a non-grain refined process, another numerical case
was run with g S, P ⫽0.0, simulating purely columnar growth. The
results were compared to experimental data from a billet cast in
the same manner as above except the absence of grain refiner.
Figure 6 shows the predicted composition field and streamline and
fraction solid profiles. With g S, P ⫽0.0, the model assumes that all
of the solid is rigid and moving at the casting velocity. The flow
from the inlet to the chill is similar to the previous case, but,
because of the much lower packing fraction, the temperature of
the flow moving down along the interface of rigid mushy zone is
much higher 共above the liquidus temperature兲 than in the grain
refined case. Also, with no Cu depleted solid being swept down
towards the centerline by the buoyancy driven flow, only Cu en-
riched liquid reaches the center. This effect accounts for the pre-
dicted positive segregation at the centerline.
While the grain refined case showed very good agreement be-
tween experiments and modeling, the case with no grain refiner
did not. The radial composition profile in Fig. 7 shows a marked
difference between predicted and measured macrosegregation.
While the predictions suggested a steady increase in composition
as the radius becomes smaller, the experiments show a pattern
more reminiscent of the profile in the presence of grain refiner
共except at the centerline, where there is an increase in copper
content兲. It is well known that the grain structure of these types of
large billets undergo a columnar to equiaxed transition around the
midradius. This change could be brought about by a slurry of solid
particles in the sump generated by fragmentation of the columnar
dendrites. This mechanism of particle generation would produce a
much smaller slurry region and a much lower packing fraction
than the grain refined billet. However, the assumption that all the
solid is rigid and moves at the casting speed is not valid in the
presence of such a particle cloud. At this point, attributing the
Fig. 6 Mixture copper composition, streamlines, and solid discrepancy between the measured and predicted macrosegrega-
fraction for billet with no grain refiner, using experimental tion to a columnar to equiaxed transition is speculative and more
boundary conditions data are needed to confirm this conclusion.

952 Õ Vol. 124, OCTOBER 2002 Transactions of the ASME

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use


Conclusions References
The continuum mixture model of Vreeman et al. 关10兴 has been 关1兴 Finn, T. L., Chu, M. G., and Bennon, W. D., 1992, ‘‘The Influence of Mushy
Region Microstructure on Macrosegregation in Direct Chill Cast Aluminum-
used to simulate the casting of industrial scale, 45 cm 共18 in.兲 Copper Round Ingots,’’ in Micro/Macro Scale Phenomena in Solidification,
diameter billets of Al-6wt percentCu using Wagstaff’s Airslip™ ASME HTD-Vol. 218, C. Beckermann, et al., eds., ASME, New York, pp.
direct chill casting process. In the case with the addition of an 17–26.
Al-Ti-B grain refiner, some agreement was found between Vree- 关2兴 Yu, H., and Granger, D. A., 1986, ‘‘Macrosegregation in Aluminum Alloy
Ingot Cast by the Semicontinuous Direct Chill 共DC兲 Method,’’ in Aluminum
man’s model and the measured vertical temperature profiles, sump Alloys-Their Physical and Mechanical Properties, EMAS, United Kingdom,
profiles, and macrosegregation in the steady state region of the pp. 17–29.
casting process. The validity of some of the assumed model pa- 关3兴 Chu, M. G., and Jacoby, J. E., 1990, ‘‘Macrosegregation Characteristics of
rameters was discussed. It was surmised that the actual packing Commercial Size Aluminum Alloy Ingot Cast by the Direct Chill Method,’’ in
Light Metals 1990, C. M. Bickert, ed., TMS, pp. 925–930.
fraction is less that the 30 percent solid which was used in the 关4兴 Dorward, R. C., and Beerntsen, D. J., 1990, ‘‘Effects of Casting Practice on
calculations, although it must be recognized that it is likely that Macrosegregation and Microstructure of 2024 Alloy Billet,’’ in Light Metals
the packing fraction in the actual billet will vary with position and 1990, C. M. Bickert, ed., TMS, pp. 919–924.
casting parameters. Without the grain refiner, the model and data 关5兴 Gariepy, B., and Caron, Y., 1991, ‘‘Investigation in the Effects of Casting
Parameters on the Extent of Centerline Macrosegregation in DC Cast Sheet
do not agree as well, possibly because the grain structure in the Ingots,’’ in Light Metals 1991, E. L. Rooy, ed., TMS, pp. 961–971.
real billet undergoes a columnar to equiaxed transition which is 关6兴 Flood, S. C., Katgerman, L., and Voller, V. R., 1991, ‘‘The Calculation of
not accounted for in the model. Macrosegregation and Heat and Fluid Flows in the D.C. Casting of Aluminum
Alloys,’’ in Modeling of Casting, Welding and Advanced Solidification Pro-
cesses V, M. Rappaz, et al., eds., TMS, pp. 683– 690.
关7兴 Reddy, A. V., and Beckermann, C., 1995, ‘‘Simulation of the Effects of Ther-
mosolutal Convection, Shrinkage Induced Flow, and Solid Transport on Mac-
Acknowledgments rosegregation and Equiaxed Grain Size Distribution in a DC Continuous Cast
Al-Cu Round Ingot,’’ in Materials Processing in the Computer Age II, V. R.
The authors wish to thank Mr. Robert Wagstaff of Wagstaff Voller, et al., eds., pp. 89–102.
Engineering for funding the experimental portion of this research. 关8兴 Ni, J., and Beckermann, C., 1991, ‘‘A Volume-Averaged Two-Phase Model for
Transport Phenomena During Solidification,’’ Metall. Mater. Trans. B, 22B,
pp. 349–361.
关9兴 Reddy, A. V., and Beckermann, C., 1997, ‘‘Modeling of Macrosegregation Due
to Thermosolutal Convection and Contraction-Driven Flow in Direct Chill
Continuous Casting of an Al-Cu Round Ingot,’’ Metall. Mater. Trans. B, 28B,
Nomenclature pp. 479– 489.
关10兴 Vreeman, C. J., Krane, M. J. M., and Incropera, F. P., 2000, ‘‘The Effect of
c ⫽ specific heat 关J/kg-K兴 Free-Floating Dendrites and Convection on Macrosegregation in Direct Chill
C Cu ⫽ wt percent of copper Cast Aluminum Alloys—I: Model Development,’’ Int. J. Heat Mass Transf.,
d ⫽ characteristic dendrite diameter 关␮m兴 43, pp. 677– 686.
关11兴 Vreeman, C. J., and Incropera, F. P., 2000, ‘‘The Effect of Free-Floating Den-
D ⫽ mass diffusion coefficient 关m2/s兴 drites and Convection on Macrosegregation in Direct Chill Cast Aluminum
f ⫽ mass fraction Alloys—II: Predictions for Al-Cu and Al-Mg alloys,’’ Int. J. Heat Mass
g ⫽ volume fraction; gravitational acceleration 关m/s2兴 Transf., 43, pp. 687–704.
h ⫽ enthalpy 关J/kg兴 关12兴 Bennon, W. D., and Incropera, F. P., 1987, ‘‘A Continuum Model for Momen-
tum, Heat and Species Transport in Binary Solid-Liquid Phase Change
k ⫽ thermal conductivity 关W/m-K兴 Systems—I: Model Formulation,’’ Int. J. Heat Mass Transf., 30, pp. 2161–
kp ⫽ equilibrium partition coefficient 2170.
K ⫽ permeability 关m2兴 关13兴 Ni, J., and Incropera, F. P., 1995, ‘‘Extension of the Continuum Model for
P ⫽ reduced pressure 关N/m2兴 Transport Phenomena Occurring During Metal Alloy Solidification—I: The
Conservation Equations,’’ Int. J. Heat Mass Transf., 38, pp. 1271–1284.
r, z ⫽ axisymmetric coordinates 关m兴 关14兴 Prescott, P. J., Incropera, F. P., and Bennon, W. D., 1991, ‘‘Modeling of Den-
t ⫽ time 关s兴 dritic Solidification Systems: Reassessment of the Continuum Momentum
T ⫽ temperature 关K兴 Equation,’’ Int. J. Heat Mass Transf., 34, pp. 2351–2358.
u ⫽ axial velocity component 关m/s兴 关15兴 Wang, C. Y., and Beckermann, C., 1996, ‘‘Equiaxed Dendritic Solidification
With Convection: Part I. Multiscale/Multiphase Modeling,’’ Metall. Mater.
v ⫽ radial velocity component 关m/s兴 Trans. A, 27A, pp. 2754 –2764.

V ⫽ velocity vector 关m/s兴 关16兴 Wang, C. Y., and Beckermann, C., 1996, ‘‘Equiaxed Dendritic Solidification
With Convection: Part II. Numerical Simulations for an Al-4 wt pct Cu alloy,’’
Greek Symbols Metall. Mater. Trans. A, 27A, pp. 2765–2783.
关17兴 Beckermann, C., and Wang, C. Y., 1996, ‘‘Equiaxed Dendritic Solidification
␤ ⫽ contraction ratio With Convection: Part III. Comparisons With NH4Cl-H2O Experiments,’’
␤ s ⫽ solutal expansion coefficient Metall. Mater. Trans. A, 27A, pp. 2784 –2795.
␤ T ⫽ thermal expansion coefficient 关l/K兴 关18兴 Ocansey, P., Bhat, M. S., and Poirier, D. R., 1994, ‘‘Permeability for Liquid
␮ ⫽ dynamic viscosity 关kg/s-m兴 Flow in the Mushy Zones of Equiaxed Castings,’’ in Light Metals 1994, U.
␳ ⫽ density 关kg/m3兴 Mannweiler, ed., TMS, pp. 807– 812.
关19兴 Smithell’s Metals Reference Handbook, 1992, 7th ed., E. A. Barnes and G. B.
Subscripts Brook, eds., Butterworth-Heinemann, Ltd., Oxford, pp. 14.1–14.14.
关20兴 Bennon, W. D., and Incropera, F. P., 1988, ‘‘Numerical Analysis of Binary
i ⫽ mold inlet condition Solid-Liquid Phase Change Using a Continuum Model,’’ Numer. Heat Trans-
l ⫽ liquid fer, 13, pp. 277–296.
关21兴 Patankar, S., 1980, Numerical Heat Transfer and Fluid Flow, Hemisphere,
liq ⫽ liquidus New York, pp. 113–134.
m ⫽ mixture 关22兴 Vreeman, C. J., and Incropera, F. P., 1999, ‘‘Numerical Discretization of Spe-
o ⫽ reference, nominal value cies Equation Source Terms in Binary Mixture Models of Solidification and
p ⫽ packed, rigid structure Their Impact on Macrosegregation in Semi-Continuous, Direct Chill Casting
Systems,’’ Numer. Heat Transfer, Part B, 36共1兲, pp. 1–14.
s ⫽ solid 关23兴 Vreeman, C. J., 1997, ‘‘Modeling Macrosegregation in Direct Chill Cast Alu-
S ⫽ solutal minum Alloys,’’ M.S. thesis, School of Mechanical Engineering, Purdue Uni-
T ⫽ thermal versity, West Lafayette, IN.

Journal of Heat Transfer OCTOBER 2002, Vol. 124 Õ 953

Downloaded From: https://heattransfer.asmedigitalcollection.asme.org on 12/03/2018 Terms of Use: http://www.asme.org/about-asme/terms-of-use

You might also like