Vous êtes sur la page 1sur 23

Columnar to equiaxed grain transition in

as solidified alloys
J. A. Spittle*
The generally reported observations pertinent to any proposed interpretation of the columnar to
equiaxed transition (CET) in as solidified alloys are initially considered. The review then proceeds
to consider the proposed mechanisms of equiaxed grain formation, the influence of alloy and
processing conditions on the CET, criteria for the termination of columnar growth and, finally,
deterministic/stochastic models for predicting the CET. In conclusion, the present level of
understanding of the CET and current modelling capabilities are summarised and assessed.
Published by Maney Publishing (c) IOM Communications Ltd

Keywords: Columnar to equiaxed grain transition, As solidified alloys, Modelling

Introduction
As solidified metals (shaped castings, ingots, continuously cast alloys, directionally solidified materials,
welds, etc.) consist of grains (single phase or multiphase) formed during solidification by one or more of
the various types of phase transformation that can occur
on cooling, e.g. primary phase freezing from the melt,
eutectic solidification and peritectic transformation.
Subsequent to nucleation, these grains either continue
to grow preferentially in a direction normal to the
liquidus isotherm in the solidifying system (i.e. they
become elongated in one dimension and are termed
columnar) or they grow in suspension in supercooled
liquid (termed equiaxed).
To manufacture useful products, most alloys are
fabricated either by solid state deformation of simple
cast forms such as sheet, billet or slab (wrought alloys)
or by filling complex shaped moulds/dies with molten
metal (casting alloys). The as cast microstructures of
wrought alloys (which are often continuously or semicontinuously cast) consist predominantly of solid solution grains of the initial primary phase. However, this is
not always the case, e.g. in low carbon steels, the initial
primary phase d ferrite is transformed to single-phase c
austenite at a lower temperature. On the other hand,
casting alloys, in most alloy systems, often deliberately
contain large volume fractions of eutectic (e.g. aluminium and zinc-base alloys and cast irons). Therefore, if
the alloy compositions lie close to the eutectic value, the
microstructures will be dominated by eutectic grains
(cells). However, other casting alloys, e.g. hypoeutectic
AlSi alloys, may contain significant amounts of
primary solid solution phase.
In as solidified structures, growth of columnar grains
often terminates with the appearance of an equiaxed
zone or, possibly, a band of equiaxed grains (which then
undergoes a further transition back to columnar
Materials Research Centre, School of Engineering, University of Wales
Swansea, Swansea SA2 8PP, UK
*Email j.a.spittle@swansea.ac.uk

2006 Institute of Materials, Minerals and Mining and ASM International


Published by Maney for the Institute and ASM International
DOI 10.1179/174328006X102493

growth). This is known as the columnar-to-equiaxed


transition (CET) (Fig. 1). In addition, a band of
equiaxed grains (the chill zone) may be seen at the
outside of the casting, the formation of which precedes
the growth of the columnar grains.
When investigating the CET, experimental observations and models of the transition have usually been
confined to the grain structures of primary solid solution
grains (although it can also be observed in the eutectic
grain structures of impure binary alloys or multicomponent alloys). In alloys, these solid solution grains
invariably grow with non-faceted cellular or dendritic
interfaces. Columnar grains freeze with a preferred
orientation such that the long axis is parallel to a specific
crystallographic direction.1 Equiaxed dendritic grains on
the other hand are randomly oriented.
The CET has been examined in wrought and casting
type alloys because of the advantages offered by
equiaxed grain solidification, in both cases, in the
majority of situations. In the case of wrought alloys,
fine-grained equiaxed structures reduce susceptibility to
hot tearing and generally improve structural homogeneity (e.g. prevent the growth of columnar feather
crystals in Al alloys). Casting alloys, for shaped
castings, are usually far less susceptible to hot tearing
and fine grain sizes enhance feeding (e.g. of long freezing
range aluminium alloys), improve the distribution of
shrinkage porosity and increase fatigue lives.
In commercial practice, attempts are made to produce
either wholly columnar structures (this is rare but the
classic example is the production of directionally
solidified turbine blades) or wholly equiaxed structures.
Interest in the CET therefore stems from the wish to
theoretically understand the solidification conditions
that define the transition between these two extremes. It
is anticipated that a mixed columnar-equiaxed structure
would be undesirable in any situation (if it can possibly
be avoided). The majority of the experimental studies/
developed models of the CET have excluded the
addition of deliberately added inoculants to promote
heterogeneous nucleation and grain refinement. This
enables mixed columnar-equiaxed structures to be more

International Materials Reviews

2006

VOL

51

NO

247

Spittle

Columnar to equiaxed grain transition in as solidified alloys

or (ii) the directional solidification of alloys either under


Bridgman or non-steady (solidification from cooled
chills) conditions. Alternatively, because of the opacity
of metals, the freezing of transparent non-metallic
analogues of alloy systems has been studied (e.g.
NH4ClH2O, cyclohexanolphenol red).
Although the CET has been reported and discussed in
many early studies including those of Stead,3 Howe,4
Genders5 and Hensel,6 significant renewed interest arose
from the initial quantitative derivation of the conditions
necessary for the breakdown of a planar interface as a
result of constitutional supercooling (CS) by Tiller
et al.7 [equation (1)]. CS, leading to the instability of
planar growth, will occur when

Published by Maney Publishing (c) IOM Communications Ltd

G=RvmC0 (1k)=kD

248

(1)

where G is the temperature gradient in the liquid, R the


interface velocity (in this review R and V are used interchangeably for interface velocity for ease of presenting
reported data), m the liquidus slope, C0 the initial alloy
composition, k the equilibrium distribution coefficient
and D the solute diffusion coefficient in the liquid.
Equiaxed grain formation and the CET have been
extensively studied over the last half century and the
important aspects that have been examined can be
summarised as follows
(i) investigation of the mechanisms of equiaxed
grain formation (how and where do the
equiaxed grains originate?)
(ii) qualitative and quantitative experimental evaluation of the parameters influencing the CET
(iii) investigation of the conditions causing the
termination of columnar freezing (what physically causes columnar grains to stop growing?)
(iv) development of deterministic and stochastic
models for predicting the CET.
The most recent comprehensive review of the CET,
giving consideration to all of these aspects is that due to
Flood and Hunt.8 In the intervening years, there has
been additional experimental research and further model
development, particularly the use of stochastic models
for simulating the nucleation and growth of individual
columnar and equiaxed grains.
1 Columnar to equiaxed (CET) grain transition in an Al
4 wt-%Cu alloy56

Requirements and features of CET

easily achieved experimentally and modelled since the


number/potency of nucleant particles will be lower.
However, commercially, the use of grain refiners may be
the normal practice, e.g. in the DC casting of Al alloys,
and one recent study has attempted to model the CET
during directional solidification of Al alloys with grain
refiner addition.2
Because of the problems associated with the manipulation of processing parameters and structural examination on a commercial scale, most qualitative and
quantitative studies of the CET have been performed
experimentally on relatively small volumes of low
melting point materials (usually aluminium alloys).
Generally speaking, most studies have involved one of
two approaches either (i) investigation of the influence
of various parameters on the as solidified grain
structures of simple shapes, e.g. cylinders, of either
poured castings or alloys melted and solidified in situ,

Unless otherwise stated, the remainder of this review


relates to the grain structures of solid solution grains in
alloys (usually binary alloys). Before embarking on a
detailed consideration of the CET, it is useful to state the
obvious requirements for a CET and to list some simple
observations pertinent to the discussion of how it is
influenced by different solidification parameters. To
obtain a structure revealing a CET (i.e. showing
columnar growth arrested by equiaxed grains), the
following are required:
(i) the presence of supercooled liquid ahead of or
near the columnar front in which either (a)
equiaxed grains nucleate and grow or (b) to
which grains/fragments of grains, nucleated elsewhere, are transported and then grow. In the
latter case, during the early stages of solidification, conditions must exist in the melt for the
nucleation/fragmentation of grains elsewhere and
their transport and survival

International Materials Reviews

2006

VOL

51

NO

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

(ii) equiaxed grains of sufficient size and/or number


beside/attached to the columnar interface to
arrest columnar growth.
These statements make no comment on the origin of the
supercooling or the mechanism of columnar growth
arrestment.
Observations that are significant to further understanding of the transition are:
(i) pure metals poured into cold moulds have
either mixed chill/columnar or wholly columnar
structures
(ii) a cast solid solution alloy may have structures
ranging from columnar to mixed columnarequiaxed to wholly equiaxed depending on
casting conditions
(iii) in binary alloys, the same amount of different
solutes (atomic or weight %) produces different
grain structures
(iv) equiaxed zones can be observed in ingots with
cellular columnar zones
(v) equiaxed grains are formed in alloys even in the
absence of deliberately added nucleants. This
implies either the heterogeneous nucleation of
equiaxed grains on unknown impurity nucleants
or alternative mechanisms of origin
(vi) solidified weld metals rarely display a CET
(vii) DC cast wrought Al alloy ingots are dominated
by columnar structures in the absence of added
grain refiners
(viii) because of fluctuations in the solidification
conditions, VAR ingots of high melting point
alloys can display alternating columnar and
equiaxed structures.
(i)(iii) emphasise the importance of solute level and
alloy system on the CET and (ii), (vi)(viii) the
importance of solidification conditions.
Because of the difficulty of tracking the columnar
interface under transient growth conditions and of
accurate measurement of shallow temperature gradients,
experimental studies of the multidirectional freezing of
poured castings, or of in situ melted and solidified alloys,
are generally confined to qualitative investigations of the
origins of equiaxed grains and of the relative effects of
different parameters on the CET. Quantitative evaluation of the conditions at the columnar interface
immediately preceding the CET usually requires the
directional freezing of alloys. However, depending on
the manner in which directional freezing is experimentally achieved, one or more of the proposed mechanisms
of equiaxed grain formation may be eliminated.

Mechanisms of equiaxed grain formation


There have been a number of reviews of the proposed
mechanisms of equiaxed grain formation, which include
those of Kisakurek,9 Flood and Hunt8 and Hutt and St.
John.10 These reviews, particularly that of Hutt and St.
John, also include detailed information on experimental
studies designed to try to validate or repudiate
individual mechanisms. As will be seen below, the
proposed mechanisms differ according to how, where
and when the equiaxed grains are thought to originate
during solidification.
It is extremely confusing, for a reader attempting to
gain an insight into the CET, to be presented with these
vast amounts of detailed and often conflicting data.

Columnar to equiaxed grain transition in as solidified alloys

Suffice to say that a variety of techniques, designed to


either (i) remove the possibility of certain mechanisms
ever operating or (ii) provide sufficiently strong evidence
to support the conclusion that a particular mechanism/s
was likely or unlikely to have operated during solidification, have been devised. These techniques include the
pouring of castings into heated moulds, in situ melting
and solidification (i.e. no pouring), the use of magnetic
fields to either enhance stirring or to remove convection
at specific stages of solidification, the use of mechanical
barriers (assumed not to be thermal barriers) positioned
both horizontally and vertically in castings, vibration of
castings, the direct observation of the solidification of
metals and non-metallic analogues of metals and,
thermal analysis. Consideration of the arguments to
validate/repudiate individual mechanisms will therefore
be omitted, but the reader can refer to one of the
suggested reviews. After presenting the proposed
mechanisms, a summary of current thinking regarding
the mechanisms will be given.
As mentioned above, the use of directional solidification to quantitatively examine the CET may automatically preclude certain mechanisms from operating. In
such studies, experimentalists usually conclude which
mechanism/s they believe were operative and directional
solidification modellers may assume that a specific
mechanism is operative, see the section Models for
predicting the CET below.
The proposed mechanisms of equiaxed grain formation in the absence of added nucleants are considered
below. It is normally assumed that the solutes have
equilibrium distribution coefficients less than unity so
that solute is rejected at the solid/liquid interface on
freezing. Solute accumulation at the interface
lowers the liquidus and depresses the interface freezing
temperature.

Constitutional supercooling hypothesis


This hypothesis, originally proposed by Northcott11 and
later by Winegard and Chalmers,12 assumes that solute
accumulation at the tips of the cellular/dendritic
columnar grains depresses the tip temperature leading
to constitutional supercooling ahead of the tips. If a
critical undercooling is exceeded, heterogeneous nucleation is suggested to occur on unknown nucleants in the
melt (Fig. 2).

The Big Bang hypothesis


Proposed by Chalmers,13 this theory argues that the
grains in the equiaxed zone nucleate at the time of
pouring in the thermally supercooled region beside the
mould walls. Grains not attached to the wall will be
swept into the bulk liquid by convection/pouring
turbulence and those that survive remelting will grow
in the constitutionally supercooled liquid ahead of the
interface advancing from the mould walls. The crux of
this theory is that all grains seen in the casting are
heterogeneously nucleated at the onset of solidification.
Another possible aspect of this proposed mechanism,
not dwelt on by Chalmers, is that the grains nucleated in
the thermally supercooled region will initially grow very
rapidly into the supercooled liquid rejecting solute at
their freezing interfaces. This will lead to solute
accumulation, depression of the liquidus and growth
restriction of the grains. This growth restriction may

International Materials Reviews

2006

VOL

51

NO

249

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

250

Columnar to equiaxed grain transition in as solidified alloys

2 Schematic series (ae) illustrating change of liquidus temperature and actual temperature in bulk liquid ahead of
advancing columnar front and origin of constitutional supercooling13

result in a further increase in undercooling and an


increase in size of the supercooled region. Under
particular circumstances (controlled by alloy parameters
and solidification conditions), undercooling may extend
throughout the casting at the time of pouring giving a
wholly equiaxed structure.

Dendrite arm remelting


From observations on the solidification of cyclohexanol
(an analogue of a pure metal) and cyclohexanol
with fluorescein (an analogue of an alloy), Jackson
et al.14 concluded that the dendrites of pure materials
differ from those of an alloy. In a pure material,
the diameters of the main stem and branches are
similar. However, in the alloy case, as soon as the
branch grows through the impurity layer around the
main stem it broadens so that it is attached to the main
stem by a narrow neck. It was also demonstrated, in a
continuously solidifying analogue system, that fluctuations in growth rate could lead to branches remelting off
the main stem and the formation of isolated crystals.
They therefore proposed that dendrite fragments,
resulting from local recalescence owing to fluctuations
in growth rate caused by convective mixing/stirring, are
the nuclei of equiaxed grains. The fragments are carried
into the bulk melt by buoyancy or convection where
they grow as new grains in the constitutionally supercooled melt.

International Materials Reviews

2006

VOL

51

NO

Showering of dendrite particles


Southin suggested, from observations on the solidified
structures of laboratory ingots of Al, Al0.1 and Al
2%Cu, that as heat is lost from the surface of an ingot, a
zone of coarse dendritic grains forms as an upper
layer.15 At some stage, dendrites or dendrite fragments
are dislodged from this layer, by some unspecified
mechanism, and sink until they meet the solid metal
growing from the mould walls. He observed that the
grains in the equiaxed zone are comet shaped, with a
coarse dendritic head and a tail that grows with the same
structure as the columnar zone.

Separation of equiaxed crystals from mould wall


This theory proposed by Ohno et al.16 bears a
resemblance to that of Chalmers in the section The
Big Bang hypothesis above, in that the grains comprising the equiaxed zone are thought to originate in the
very early stages of solidification. Ohno et al. used
optical microscopy to directly observe the start of
solidification at the mould wall for the unidirectional
freezing of SnBi, BiSn, SnPb and SnSb alloys in
horizontal Pyrex tubes. They observed that, in the
presence of solute, growth of granular shaped equiaxed
crystals took place which were attached to the mould
wall by narrow necks. Subsequently, and before the
formation of a complete solid shell at the mould wall, it
was observed that equiaxed crystals separated from the

Spittle

mould wall owing to remelting of the necks caused by


thermal convection. The investigators considered that
these detached equiaxed crystals act as the nuclei for the
formation of the grains in the equiaxed zone.

Published by Maney Publishing (c) IOM Communications Ltd

Overall consideration of proposed mechanisms


of equiaxed grain formation
The various proposed mechanisms of equiaxed grain
formation can be grouped into two categories
(i) those mechanisms that involve direct heterogeneous nucleation of equiaxed grains in the bulk
liquid12,13
(ii) those mechanisms that involve the detachment
of dendrites/dendrite fragments from grains
nucleated and growing from the mould walls/
upper liquidair surface of the ingot.1416
In each category, the differences between the mechanisms are governed by when during solidification and
where in the casting the mechanisms are thought to
operate. Without exception, experimental data exist to
support every one of the five mechanisms under the
conditions prevailing in the individual experiments.
However, the experimental arrangements, often small
cast cylinders of alloys or the freezing of non-metallic
analogues, differ considerably from the shapes and
solidification conditions found in shaped castings, large
ingots and continuously cast alloys. In reality, it is likely
that more than one of the mechanisms may be operative
in a particular casting situation. However, specific
casting conditions are likely to favour different mechanisms. In the case of shaped castings, which often have
thin sections, significant undercooling (both in degree
and depth) and convection/pouring turbulence are
thought to favour those mechanisms that operate at
the onset of solidification.13,16 In the case of large ingots
and continuously cast alloys (associated with long
solidification times and significant interdendritic convective flow), or processes such as vacuum arc remelting
which can be subject to major fluctuations in thermal
solidification conditions, dendrite remelting is likely to
be enhanced. In the absence of hot tops, solidification on
the top surface of large ingots may also contribute to the
formation of equiaxed grains. On the other hand, under
conditions of perfect directional freezing into a positive
temperature gradient, in the absence of convective flow
or any pouring turbulence, if a CET is observed, it must
originate from grains heterogeneously nucleated in
constitutionally supercooled liquid.
All the above mechanisms have also been considered
when examining the CET in other as solidified
structures, e.g. welds. In the latter case, an additional
alternative mechanism termed grain detachment has
been proposed.17 This mechanism involves the detachment of small grains from the partially molten part of
the heat affected zone. This requires fine-grained alloys
with relatively large freezing ranges.

Parameters influencing CET


Many experimental investigations of the CET have been
qualitative in nature, recording the influence of different
parameters, such as superheat and alloy content, on the
relative sizes of the columnar/equiaxed zones, columnar
zone length, grain sizes, etc. for a specific experimental
arrangement. Although, as will be described below,

Columnar to equiaxed grain transition in as solidified alloys

there have been numerous attempts to quantitatively


determine the precise set of conditions that exist at
the location of a CET, these conditions are still not
sufficiently well understood to facilitate the prediction or
avoidance of a CET for different solidifying geometries,
solidification processes, solidification conditions and
alloy systems. Flood and Hunt8 list the important
parameters as alloy factors, superheat, fluid flow, casting
size, mechanical vibration and inoculation grain refinement. Mechanical vibration is not considered at all in
the present review and there have only been a limited
number of studies of the CET in the presence of added
grain refiners either experimentally or by modelling.

Alloy parameters
Omni-directional freezing

Although equiaxed crystals can be found in thermally


undercooled in situ melted and solidified pure metals,18
it is accepted that in castings/welds of pure metals,
columnar grain growth predominates. In castings and
welds, solute/s (soluble in the liquid) is therefore
required to cause a CET. This solute provides the
constitutional supercooling necessary for the survival
and growth of equiaxed grains and, possibly, their
nucleation. A variety of alloy parameters can influence
the transition including m, k, D and C0.
A number of generally accepted conclusions have
been reached from observations on poured or in situ
castings. For a given alloy system and a constant
superheat, increasing C0 tends to decrease the columnar
zone length and reduce the equiaxed grain size, i.e.
promote the CET.15,19,20 Tarshis et al.,21 in a classic
paper, examined the variation in grain size in binary Ni
alloys and Al alloys. For each alloy system, a variety of
solutes were employed in order to vary the magnitudes
of the different alloy parameters. The grain sizes of a
series of poured binary alloy castings were compared for
a fixed superheat and a fixed level of solute addition.
Three series of alloys were examined Ni1 at.-% solute,
Ni5 at.-% solute and Al1 at.-% solute. For each series,
it was found that the relative grain size decreased as the
parameter P increased, where P is given by
P~{mC0 (1{k)=k

(2)

At low P values, the structures were columnar and


changed to columnar-equiaxed and finally equiaxed as P
increased (Fig. 3).
In a given binary system, if the solidus and liquidus
lines are straight, for compositions below the solubility
limit, P (since termed the constitutional supercooling
parameter) is equal to the equilibrium freezing range.
Tarshis et al.21 proposed that since this parameter
predicts grain size, it permits the selection of solutes as
grain refining additions (in the absence of added
inoculants). It should be borne in mind, however, that
the quantitative observations relate to a particular
mould material, casting size and set of casting conditions. Prediction of the CET, if one or more of these is
varied, is impossible. The form of the variation of cast
structure with P has been confirmed for various binary
Al alloy systems.22 From observations on AlZn alloys,
containing up to 85 wt-%Zn, in which P and freezing
range vary independently with composition, Doherty
et al.23 found no direct correlation between the CET and
freezing range but a reasonable correlation with P. In

International Materials Reviews

2006

VOL

51

NO

251

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

Columnar to equiaxed grain transition in as solidified alloys

3 Variation of relative grain size of Al1 at.-% solute alloys as a function of parameter P 21

recent years, when examining the influence of phase


diagram parameters, it has become more common to
relate grain sizes to the growth-restriction parameter,
Q, rather than P, where Q is given by
Q~kP

(3)

Although, as illustrated above, numerous examples exist


of where solute redistribution on solidification promotes
grain refinement and the CET, in some alloy systems,
solute addition can result in grain coarsening. An
important example is seen in castings of AlSi alloys.
a(Al) grain size initially decreases with Si content up to
,23%Si and thereafter continues to increase.24,25 Pb,
Sb and Bi also cause coarsening of the Zn solid solution
grains in Zn and Zn-base alloys.26

breakdown. They suggested that the results supported


the Winegard and Chalmers mechanism of equiaxed
grain formation.12 Elliott used a thermal valve technique, which permits independent control of G and R, to
investigate the directional freezing of PbSn alloys
containing up to 6 wt-%Sn.28 A linear relationship
between G/R and C0 for the CET was obtained in
agreement with the analytical model of Tiller29 for the
transition (see the section Deterministic models below)
(Fig. 4). This model again assumes that equiaxed grains

Unidirectional solidification

Instead of trying to experimentally quantify the influence of different alloy parameters on the CET, for a
given set of casting conditions, during omni-directional
solidification, an extensively used alternative approach
has been to try to determine the conditions existing at
the location of the CET for a given alloy system. Studies
of this type involve unidirectional solidification.
Plaskett and Winegard27 examined the unidirectional
solidification of AlMg alloys under non-steady conditions. The alloys were melted in situ in a graphite
crucible held in a furnace and directionally solidified
from a cooled chill. For each alloy, the values of G and
R were determined at the location of the breakdown of
columnar growth. Over the composition range investigated, 010%Mg, an almost linear relationship was
obtained between C0 and the value of G/R1/2 at

252

International Materials Reviews

2006

VOL

51

NO

4 Experimental
plot
illustrating
linear
relationship
between wt-%Sn and G/R for columnar to equiaxed
transition in PbSn alloys28

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

form by heterogeneous nucleation ahead of the columnar front due to CS. Tiller predicted that the equiaxed
zone would form when the maximum undercooling
exceeded a critical value.
The significant renewed interest in the CET since the
1980s has resulted from the emergence of new modelling
approaches to investigate the transition which, as a
consequence, have prompted further experimental studies. The studies have been performed on a variety of
alloy systems and the data have often been compared to
the predictions of models. Alternatively, modellers have
used experimental data to assess the general validity of
their models.
Mahapatra and Weinberg30 and Ziv and Weinberg,31
using a 1D finite difference heat transfer model,
investigated the CET, under non-steady freezing conditions, for in situ melted and directionally solidified
alloys of Sn containing 5, 10 and 15 wt-%Pb and, Al
3 wt-%Cu, respectively. In the case of the SnPb alloys,
they reported that the CET occurred when the
temperature gradient ahead of the advancing dendrite
tips for each alloy fell below a critical value. The values
(which were determined from a finite difference heat
transfer model of the solidifying ingot) were 1.0, 1.1 and
1.3 K cm1, respectively. The CET could not be directly
related to dendrite tip velocity. For the Al3 wt-%Cu
alloy, the CET occurred when the gradient fell to
0.6 K cm1 which was found to be in agreement with a
gradient prediction from Hunts analytical model of the
CET (see the subsection Analytical models below).32
From adding different amounts of nominally Al5Ti1B
grain refiner to the AlCu alloy, it was found that a
definite amount was required in order to effect the CET.
They suggest this indicates that a critical high density of
nuclei is required to form a fine-grained structure.
Suri et al.33 again using a 1D heat transfer model to
determine the values of G and V at the positions of the
CET, studied Al4.5%Cu alloys directionally solidified
under non-steady conditions for different superheats
and conditions of heat extraction. They reported that
the transition occurred if
:
Gv0:74V 0 64

(4)
34

Kim and Grugel, using a Bridgman type furnace,


examined the CET of the Cu dendrites in PbCu alloys
containing 4, 9 and 14 wt-%Cu. The alloys were
directionally solidified at velocities ranging from 1 to
100 mm s1 and a temperature gradient of 4.5
0.5 K mm1. In this composition range, it was found
that as growth velocity increased, there was a transition
from fully columnar to fully equiaxed. The growth
velocity to effect a fully equiaxed structure was observed
to drop rapidly with decreasing Cu content. The investigators concluded that the results were in qualitative
agreement with Hunts theory of the CET.32
Ledgard and McCartney,35 using a Bridgman type
arrangement, examined the directional solidification
of Al1.8 wt-%Si alloys to which an Al6 wt-%Ti
0.02 wt-%C grain refiner was added at levels of 0.01,
0.03 or 0.05 wt-%Ti. Alloys were prepared from both
superpurity (99.995 wt-%) and commercial purity
(99.7 wt-%) Al. Pull rates of 1, 4, 10, 30 and
60 mm min1 were used and growth velocities were said
to be within 10% of these rates. Temperature gradients
in the liquid were typically 5 K mm1 at the highest

Columnar to equiaxed grain transition in as solidified alloys

growth rate rising to ,10 K mm1 at the lowest growth


rate. It was found that the superpurity alloys did not
produce equiaxed structures for any grain refiner
addition level or growth rate employed. In the case of
the commercial purity alloys, equiaxed structures were
observed for all growth rates at the 0.03 and 0.05%Ti
levels. At the 0.01% level, columnar grains were
observed at the lowest growth rate but equiaxed
structures were obtained at all pull rates in excess of
4 mm min1. The results for the critical velocity for the
CET were considered using the Hunt model.32
Reasonable agreement with the model was found for
an assumed nucleation undercooling of 0.75 K and an
estimated density of nucleating sites of 200 mm3.
Pollock and Murphy,36 using Bridgman type furnaces,
examined the breakdown of single crystal solidification
in high refractory nickel-base alloys. Ten different alloys
were studied under conditions of directional solidification. All 10 alloys were studied using fixed values for the
growth rate and thermal gradient. However, one of the
alloys (Al6.0, Cr4.5, Co12.5, Hf0.16, Re6.3, Ta7.0,
W5.8, Nibal) was investigated for withdrawal rates
ranging from 4.26104 to 1.136102 cm s1 and thermal gradients in the range 0.3140 K cm1. No grain
defects were observed when gradients were greater than
approximately 15 K cm1. For a fixed withdrawal rate,
they observed that, as the gradients decreased below this
critical value, an abrupt transition from single crystal
dendritic to equiaxed solidification was not observed.
Instead, there was a large intermediate range of
conditions where grains nucleate and grow without
completely interrupting single crystal columnar growth.
The primary dendrite arm spacing at which isolated
grains began to nucleate and grow corresponded with
that which marked the onset of freckling. The authors
conclude that thermo-solutal convection promotes the
development of both types of defect and that dendrite
detachment/fragmentation contribute to the development of isolated grains and ultimately to the transition
to polycrystalline equiaxed solidification. This conclusion regarding the origin of the equiaxed grains is in
marked contrast to those in the directional studies
already described where it is assumed that the equiaxed
grains nucleate in supercooled liquid ahead of the
advancing front.
Gandin37 has studied the directional solidification of
AlSi alloys containing 3, 7, and 11%Si under nonsteady freezing conditions. Liquid alloy is initially held
at a uniform temperature within a mould located in a
furnace. The furnace is lowered away from the mould
and directional freezing initiated by raising a watercooled copper chill to contact the bottom of the mould.
In all three alloys, a CET was observed at approximately
the same position two-thirds along the ingot length.
Gandin suggested that the equiaxed grains may have
originated from a dendritic surface layer at the metal/air
interface, by heterogeneous nucleation in the bulk liquid
or by dendrite arm detachment/fragmentation.
Ares and Schvezov38 studied the CET in PbSn alloys
in the range 240%Sn. Again, directional freezing
occurred under non-steady conditions. During solidification, temperatures were measured at five locations
along the ingot length at 10 s intervals. From the
cooling curves, a variety of parameters were calculated.
These included superheat, cooling rates, positions and

International Materials Reviews

2006

VOL

51

NO

253

Spittle

Columnar to equiaxed grain transition in as solidified alloys

Published by Maney Publishing (c) IOM Communications Ltd

5 Temperature gradients in liquid versus velocity of liquidus interface38

254

velocities of the liquidus and solidus fronts, length of the


mushy zone, local solidification times and temperature
gradients. For an alloy containing 2%Sn, their results
showed that there was no correlation between columnar
length and superheat and that the length of the
columnar zone increases with the average cooling rate
of the melt. For a given cooling rate, it was also found
that the length of the columnar zone increases with alloy
composition. From the five thermocouples, values of the
temperature gradient ahead of the liquidus interface and
the velocity of the liquidus interface were calculated for
all the experiments performed over the composition
range 240%Sn. These were used to plot gradient versus
interface velocity, indicating the type of grain structure
present for each pair of values (Fig. 5). When observed,
the CET was not sharp but showed a transition zone
where some equiaxed grains coexisted with columnar
grains. In all the experiments, the gradient for the CET
was calculated to be within 0.8 to 1 K cm1. It was also
found that the CET corresponded to a critical interface
velocity of ,0.0100.005 cm s1. The data in Fig. 5,
the influence of cooling rate on columnar length and the
fact that the critical values of temperature gradient and
interface velocity were independent of alloy composition
and position of the transition, led the investigators to
suggest that the process is mainly controlled by heat
extraction. From calculations, they showed that the
amount of heat flow decreases with time and reaches a
minimum critical value of ,0.100.04 J cm2 s1 at the
CET. A qualitative fit with Hunts model32 was obtained
by adjusting the values of DTN and N0 to fit the
experimental results. However, in order to fit the
experimental data, the patterns of the variations of
these two parameters with Sn content appeared to be
unrealistic.
Martorano and Capocchi39 examined four castings of
Cu8%Sn which were unidirectionally solidified, under
non-steady conditions, by pouring the molten metal into
an insulated mould standing on a copper base. In two of
the experiments, castings were poured at 1110uC onto a
non-cooled base with and without the addition of
0.08%Zr as an inoculant. In the other two, the alloys
were poured onto a water-cooled base, again with and
without the same level of inoculant addition. In the
absence of the inoculant, raising the pouring temperature and increasing the heat flux from the base increased

International Materials Reviews

2006

VOL

51

NO

the columnar length, i.e. delayed the CET. The


inoculated castings were completely equiaxed.
Siqueira et al.40 examined the CET in SnPb alloys
(10 and 30 wt-%Pb) and AlCu alloys (2, 5, 8 and
10%Cu), again under non-steady directional freezing
conditions. The alloys were preheated, in a mould held
in a furnace, to the desired superheat. The furnace was
then switched off and water-cooling applied to the base
of the mould. The CET was analysed for different
superheats and heat-transfer efficiencies at the base. The
mould consisted of a stainless steel cylinder closed at the
bottom with a disc of carbon steel. The inner vertical
surface was coated with a layer of insulating alumina. In
some experiments, the heat-extracting surface of the
mould was coated with an alumina-based wash and in
others, it was polished. Temperatures along the length of
the casting were monitored from a series of thermocouples. These measured temperatures were used in a 1D
finite difference heat transfer programme to determine
the transient heat transfer coefficient representing the
global coolantcasting heat exchange. Experimental
results of the position of liquidus isotherms as a function
of time gave good agreement with those numerically
predicted using the corresponding transient heat transfer
coefficient. The numerical model was therefore used to
calculate certain solidification parameters associated
with the CET transition. These were tip growth rate,
temperature gradient in the liquid and tip cooling rate.
For all test conditions examined, fully columnar
structures were always observed in the SnPb alloys.
From observations on the AlCu alloys, the CET was
observed to occur rapidly on a near horizontal plane and
further from the chill with increasing heat transfer
coefficient and increasing superheat. For these AlCu
alloys, it was reported that the CET occurred at tip
growth rates ranging from 0.28 to 0.88 mm s1 and for
temperature gradients in the liquid at the liquidus
isotherm ranging from 0.28 to 0.75 K mm1. The
investigators therefore concluded that a criterion for
the CET could not be based solely on tip growth rate or
solely on temperature gradient. They suggested that a
more realistic criterion should encompass both tip
growth rate and temperature gradient through the tip
cooling rate. For the 15 tests on the AlCu alloys, the
CET occurred when the cooling rate fell below the
critical value of ,0.2 K s1. By comparing the gradients

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

calculated by the authors for the CET, for the five


experiments conducted with Al5 wt-%Cu, with those
predicted by the Suri et al.33 criterion for the transition,
only two supported the latter criterion.
Vandyoussefi and Greer,41 using a Bridgman technique, investigated the influence of solidification front
velocity on the grain structures of Al4.15 wt-%Mg
alloys, with and without inoculation. The refiner used
was Al3.16 wt-%Ti0.17 wt-%C and the addition level
was varied. After temperature stabilisation at 720uC, the
alloy contained in an alumina tube was lowered through
the furnace at a velocity between 0.05 and 1.0 mm s1.
The temperature gradient in the liquid was fixed at
102 K mm1. Without inoculant addition, the structures were always columnar. With low to medium levels
of addition (2 to 10 parts per thousand), mixed nonequiaxed and equiaxed structures were found.
Solidification of the alloys was also studied using a
commercial CA-FE (cellular automaton-finite element)
package CalcoMOSTM,42 see the section Stochastic
models below.43 Both experiment and CA-FE modelling predict that the CET is gradual with intermediate
stages of short columnar or elongated grains. Microstructural studies of quenched interfaces appeared to
support the contention of Hunt32 that the CET is the
result of the competition between continued growth of
existing grains and the appearance of new grains in the
constitutionally undercooled region ahead of the main
growth front. Experimental results were plotted on an
interface velocitytemperature gradient map showing
the CET and compared with simulation from Hunts
analytical model. Again, there was reasonable qualitative agreement.
Two very recent investigations on aluminium alloys
have also been carried out, both with and without the
deliberate addition of grain refiners, using Bridgman
type furnaces. As part of a European Space Agency
programme on the columnar to equiaxed transition in
solidification processing, Sturz et al.44 have studied the
directional freezing of Al7 wt-%Si alloy rods 10 mm in
diameter and 200 mm in length. The grain refined alloys
contained 215 mg g of titanium and 15 mg g of boron.
During solidification, the temperature gradient was
decreased and the solidification rate increased simultaneously to initiate the CET in a transient experiment.
Experiments were performed with different cooling
rates. From thermocouples located along the samples,
the values for G and V at the CET were determined. A
linear decrease in columnar grain length with increasing
cooling rate was found for non-refined alloys. The
critical experimental values of G and V at the CET were
compared with the models of Hunt32 and Martorano
et al.45 (see section Deterministic models below),
calculated for Al7 wt-%Si, for three different values
of the critical undercooling: 0, 3 and 5 K. The critical
experimental values were found to be in good agreement
with the model of Martorano et al. for a critical
undercooling of about 5 K. Grain refinement resulted
in a lower critical undercooling, a higher critical temperature gradient and higher grain densities in the
columnar and equiaxed regions. The CET was smoother
with refined alloys.
Reinhart et al.46 have used a novel approach to make
direct observations of the solidifying interface, in
particular at the CET. Synchrotron X-radiography has

Columnar to equiaxed grain transition in as solidified alloys

6 Synchroton X-ray images recorded a 42 s, b 63 s, c


87 s and d 111 s after a sharp increase in pull rate;
dashed line marks eutectic front position46

been used to examine vertically solidified refined and


non-refined Al3.5 wt-%Ni alloys in a Bridgman furnace. The samples were about 40 mm in length by 6 mm
in width by 150200 mm in thickness. In any experiment,
solidification was started with a low pulling velocity to
produce a columnar dendritic structure. The CET was
then initiated by suddenly applying a sharp increase in
pull rate, keeping the gradient constant. Figure 6a to d
shows a sequence of images following an increase in pull
rate from 1.5 to 15 mm s1, for a gradient of 2 K mm1,
for a grain-refined alloy with 0.5 wt-% of an Al5Ti1B
grain refiner added. Direct observation of the CET
reveals some interesting features. A short time after the
velocity jump, a band of equiaxed crystals appears in the
supercooled liquid beside the columnar dendritic front.
Some are nucleating around the columnar dendrites but
some are nucleating in an almost horizontal band. When
enough grains have nucleated and grow, they may block
the columnar growth leading to the CET. The fact that
both columnar and equiaxed dendrites are blocked
before the grains are touching each other (Fig. 6c and d)
led the investigators to suggest that the blocking is
mostly solutal as proposed by Martorano et al.45 see
sections Proposed mechanisms/criteria for the termination of columnar growth and Models for predicting the
CET below. Another interesting feature was that some
of the newly nucleated grains fall down either on the
columnar dendrite or towards the eutectic front, i.e. the
liquid area on both sides of the columnar dendrite can
be gradually filled in with equiaxed dendrites. Therefore,
a post-mortem analysis of the sample could lead to the
false conclusion that the structure is mixed columnarequiaxed, whereas it is only caused by sedimentation.
Direct comparison of the data sets from the various
directional solidification studies is virtually impossible.

International Materials Reviews

2006

VOL

51

NO

255

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

Columnar to equiaxed grain transition in as solidified alloys

First, the studies cover a wide range of alloy systems


including AlCu, AlMg, AlNi, AlSi, PbSn, SnPb,
PbCu, CuSn and Ni-base alloys. Second, even where
the studies have been made on the same system,
sometimes a single composition has been examined
whereas in other studies a range of compositions were
investigated. Third, some investigations were carried out
under steady controlled freezing conditions (using a
Bridgman technique) whereas others were performed
under non-steady conditions. Fourth, in some studies,
grain refining inoculants have been deliberately added
whereas in the others equiaxed grains have either been
heterogeneously nucleated on unknown substrates or
formed by other means. Depending on the alloy system
and directional freezing method employed, other major
differences can be recognised. In some alloy systems,
solutes may have partition coefficients, k, less than unity
and, in others, values greater than unity. In some of the
systems, solute redistribution on freezing may result in
the interdendritic liquid becoming less dense than the
bulk liquid leading to thermosolutal convection when
solidifying upwards. This is the case in those k,1
systems where the solutes are less dense than the solvent
or those k.1 systems where the solutes are more dense
than the solvent. Different mechanisms of equiaxed
grain formation may be operating in the different
studies, even in the absence of deliberately added grain
refiners. Under steady controlled freezing conditions
and in the absence of thermosolutal convection, heterogeneous nucleation in the bulk liquid ahead of the
advancing front may be a plausible mechanism. In the
presence of thermosolutal convection, dendrite remelting/fragmentation may be the mechanism as evidenced
in the study by Pollock and Murphy.36 Under nonsteady freezing conditions, loss of heat from the upper
liquid surface may result in the formation of a surface
dendritic layer. Fragmentation of this layer could lead to
showering of dendrite fragments which grow as
equiaxed crystals. From the directional solidification
studies, it is only possible to draw certain generalised
conclusions. As stated by Quested and Greer,2 in many
instances equiaxed microstructures are found to be
favoured by high interface growth velocities, low
temperature gradients ahead of the advancing front,
high solute levels and a larger number of equiaxed
nucleation events. However, the precise relationship,
governing the CET, between the local solidification
conditions at the advancing interface during directional
solidification and alloy parameters is still far from clear.
Attempts to date to compare experimental data with
Hunts analytical model for the CET have only been
qualitative in nature because of the assumptions
required regarding the number of nucleant particles
and the undercooling for nucleation of the equiaxed
grains.

Superheat
Numerous authors have reported that increasing superheat increases columnar grain length, i.e. suppresses the
CET (Fig. 7).13,19,4751 Although this is true for laboratory scale experiments, frequently performed on cylindrical cast shapes of less than 500 cm3 in volume, the
observation cannot be extrapolated to larger volumes.
Morando et al.50 examined the solidification of a
series of Al2%Cu cylindrical ingots, varying in volume

256

International Materials Reviews

2006

VOL

51

NO

from ,50 to 500 to 5000 cm3, poured into graphite


moulds with a fixed ratio of internal diameter to internal
height of 0.7. For each volume, ingots were poured with
a range of superheats from 20uC to 150uC. The
columnar length from the ingot base was used as the
measure. For the 50 cm3 volume, columnar length
increased with superheat. There was also some length
variation in the 500 cm3 ingots. However, in the
5000 cm3 ingots, the columnar length was invariable
with superheat. As discussed by the investigators, for the
smallest volume, solidification times are short and
settling of crystals is unimportant. However, they
obtained evidence that, with increase in size, Southins
mechanism of equiaxed grain formation15 comes into
play and, as solidification time increases, crystal settling
becomes more important.
Because of their low thermal conductivities and long
freezing times, the solidification of small castings of
transparent non-metallic systems, e.g. NH4ClH2O is
often taken as an analogue of the freezing of large ingots.
Such studies have revealed the importance of dendrite
remelting,52 or alternative mechanisms of equiaxed grain
formation53 and the importance of equiaxed grain
settling.
When dealing with ingot shapes, the often reported
influence of superheat on the CET is therefore only valid
for small laboratory scale castings. However, superheat
will also be of importance in larger commercial thin
walled castings, because of the higher surface to volume
ratios. The influence of superheat on primary phase
grain structures in thin section castings has not been
rigorously studied, primarily because a certain minimum
superheat must be exceeded in order to provide
adequate fluidity.
Assuming, for the case of small laboratory ingots
(,500 cm3), that grains nucleated near the mould wall
on pouring are the probable origin of the central
equiaxed zone,13,16 increasing superheat will decrease
the degree and depth of supercooling and increase the
time required for dissipation of the superheat. Fewer
nuclei will therefore initially form and survive remelting.
Some of the non-steady unidirectional solidification
studies described in the section Unidirectional freezing
also report the effect of superheat on the CET.
Generally speaking, as might be anticipated, increasing
superheat increases columnar zone length and delays the
CET.

Fluid flow
Investigations of the significance of fluid flow to the
CET have centred on two aspects. First, determination
of the influence of natural convection/pouring turbulence on the transition and, second, application of forced
flow to promote equiaxed grain formation. Much of this
work, which has been qualitative in nature, was carried
out by Cole and Bolling.19,48,49,54,55
Considering the role of natural convection, a variety
of techniques have been used to reduce/remove convection in the melt.19,47,49,50,56 Cole and Bolling inserted
grids across the mould diameter19 or slowly rotated the
mould about its vertical axis during solidification.49
Others have solidified alloys in a static magnetic
field,47,50,56 either applying the field at all times or, over
selected time periods during solidification in order to
differentiate between potential mechanisms of equiaxed

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

Columnar to equiaxed grain transition in as solidified alloys

a 40uC superheat; b 80uC superheat


7 Effect of superheat on grain structure of Al2 wt-%Cu alloys50

grain formation. Generally speaking, increasing fluid


flow decreases columnar grain length. This is due to one
or more of the following
(i) increase in the number of dendrites/dendrite
fragments transported from the vicinity of the
mould walls/melt surface into the bulk liquid
(ii) increase in the rate of loss of superheat from the
bulk liquid thus increasing the chance of nucleation and/or survival of equiaxed grains ahead of
the advancing front
(iii) increase in the number of nuclei for equiaxed
grains formed by dendrite fragmentation.
The results support the contention that convection
increases the rate of loss of superheat. They also
highlight the significance of equiaxed grain formation
in the early stages of freezing in small ingots and the
probable importance of dendrite fragmentation in larger
ingots.
Regarding forced flow, early studies were made using
Lorentz interaction between a current and magnetic
field48 and oscillation of the ingot mould during
solidification.55 These studies indicated that forced flow
causes grain refinement, primarily, it was thought, by
enhancing the fragmentation of dendrites.
More recently, several studies have been made of the
influence of forced flow on the CET in alloys unidirectionally solidifying under non-steady conditions.5759
Since most of the early experiments associated with

investigating the influence of fluid flow on grain


structure were qualitative in nature, Griffiths and
McCartney57 set out to try to quantify the effect of
melt velocity on structural transitions. A series of AlSi
alloys was directionally solidified downwards, in a
rectangular mould 150 mm long by 65 mm wide by
210 mm deep, by chilling the top surface of the melt.
The alloys, which contained 2.5, 5.6 and 8.5 wt-%Si,
were made from 99.7 wt-%Al and 98.4 wt-%Si. Experiments were also performed with an Al2.5 wt-%Si alloy
prepared from higher purity materials. The mould was
initially preheated to above the alloy liquidus to prevent
nucleation on pouring. Experiments were performed
with and without electromagnetic stirring. A representation of the fluid flow in the mould under conditions of
natural convection and stirring was obtained using the
commercial CFD software package FLUENT. Temperatures were recorded from a series of thermocouples
aligned vertically along the central axis of the mould.
Ingots were sectioned longitudinally along the midplane. The position of the CET was traced and the area
percentage of the equiaxed region calculated. Under
conditions of natural convection, there was no apparent
relationship between Si content and extent of the
equiaxed region. However, when electromagnetic stirring was applied, the extent of the equiaxed zone was
found to increase systematically with both increasing Si
content, at a fixed stirring current, and with increasing

International Materials Reviews

2006

VOL

51

NO

257

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

258

Columnar to equiaxed grain transition in as solidified alloys

stirring current at a fixed Si content. Electromagnetic stirring therefore promotes the CET.
Electromagnetic stirring resulted in the rapid removal
of the bulk liquid superheat and liquid velocities up to
0.25 m s1 were predicted. The investigators concluded
that their results support the contention that forced fluid
flow promotes the equiaxed region and the CET by
fragmentation of the solidifying front. The role of
heterogeneous nucleation was also investigated by
comparing ingots of Al2.5 wt-%Si made from base
materials of different purities solidified under conditions
of natural convection (no forced stirring). The equiaxed
zone in the higher purity alloy (lower Ti and B levels)
contained fewer and larger grains than in the commercial purity alloy. This suggested that, under conditions
of natural convection, the CET is influenced by the
number of heterogeneous nuclei. With stirring, the
number of heterogeneous nuclei becomes less important
because of the creation of dendrite fragments from
which growth of equiaxed dendrites can proceed.
Willers et al.58 carried out an experimental study to
investigate the influence of bulk flow in the melt on the
unidirectional solidification of cylindrical samples of
alloys of Pb and Sn. The bulk flow was generated by a
rotating magnetic field. The alloys studied were Sn
15 wt-%Pb, Sn38 wt-%Pb and Pb25 wt-%Sn. The
second alloy corresponds to the eutectic composition.
These compositions were chosen to provide different
density ratios between interdendritic liquid, bulk liquid
and primary crystals and, also, to provide an alloy
(eutectic) that is not subject to macro solute rejection at
the solidifying front or grows dendritically. The alloys
were preheated in cylindrical moulds 50 mm internal
diameter by 100 mm in height to a superheat of 90 K.
The alloys were then solidified upwards by placing the
mould on a water-cooled copper chill, the whole
assembly being located inside a magnetic inductor.
Stirring of the melt was initiated simultaneously with
the cooling. Temperatures were recorded along the
central axis of the cylinder and experiments were
conducted with variations of the magnetic field strength.
In the absence of a magnetic field, the alloys were
completely columnar. Melt agitation promoted grain
refinement. The present study revealed, in all cases
including the eutectic alloy, that increasing field strength
displaced the CET towards the bottom of the cylinder.
The CET appears for critical values of the cooling rate
of about 0.4 K s1 and for temperature gradients
between 0.6 and 1.0 K mm1. These are about an order
of magnitude greater than those reported for non-stirred
melts.30,40 Measurements revealed the existence of
remarkable temperature fluctuations in the mushy zone
lending support to thermal melting of dendrites. Solute
accumulation at a CET was considered to be responsible
for obstruction of columnar growth, in accordance with
Ref. 45, see the section Proposed mechanisms/criteria
for the termination of columnar growth below. The
origin and the effect of magnetic field strength on the
CET in eutectic alloys are unclear. It appears that
dendrite fragmentation and solutal interaction between
columnar and equiaxed growth cannot be responsible
and the investigators simply suggest that stirring leads to
a multiplication of nuclei.
In a continuation of the research described in the
previous paragraph using the same experimental

International Materials Reviews

2006

VOL

51

NO

arrangement, Eckert et al.59 studied the solidification


of Sn15 wt-%Pb alloys in a rotating magnetic field. The
results were compared with numerical simulations for
the temperature and velocity fields in the liquid phase. In
one set of experiments, the rotating magnetic field was
not switched on at the onset of cooling but after time
delays. These experiments gave no indication of considerable thermal remelting of the solidification front. It
was therefore concluded that the activation of the CET
by stirring is not necessarily based on fragmentation of
the columnar dendrites.
Electromagnetic stirring to promote the CET has been
applied commercially to the continuous casting of steel.
The low thermal conductivity of steel leads to deep
liquid pools inside the solidifying shell, within which the
temperature gradients are shallow. Electromagnetic
stirring has also been applied to the direct chill casting
of wrought aluminium alloys.

Overall consideration of parameters affecting


CET
The influence of the principal phase diagram parameters
m, k and C0 on the CET are now qualitatively well
understood although, in a particular system, they cannot
be varied independently. The CET is encouraged as the
constitutional supercooling parameter P increases.
Solute redistribution on freezing lowers the solid/liquid
interface temperature, promotes CS and restricts grain
growth prompting further nucleation. In the case of
small experimental ingots, columnar length increases as
superheat increases. This is primarily because of the
influence of superheat on the remelting of equiaxed
dendrites nucleated in the early stages of solidification.
The effect of superheat decreases as ingot size increases
because the origin/s of the equiaxed grains changes and
equiaxed crystal settling becomes more dominant in
determining the cast grain structure. Natural and
enhanced fluid flow encourage the CET by transporting
dendrites/dendrite fragments, enhancing dendrite fragmentation and increasing the rate of loss of superheat.
The collective interaction of all the above parameters
is extremely complex which makes quantitative prediction of the CET very difficult.

Proposed mechanisms/criteria for


termination of columnar growth
Various mechanisms have been proposed to try to
explain the local interaction of columnar grain growth
and equiaxed grain formation that gives rise to a CET.
These consider, to varying extents, the solute and
temperature fields, fluid flow and the growth of equiaxed
grain nuclei. Whatever the mechanism, a transition will
only occur if the equiaxed grains are sufficient in size or
number to arrest columnar grain growth. This is
supported by the observation that individual equiaxed
crystals can be found trapped within columnar zones. A
CET will occur if the equiaxed grains cannot be pushed
ahead of the growing columnars and the columnars are
incapable of growing between the equiaxed grains. The
conditions associated with the local columnar/equiaxed
interaction will be determined, as discussed in earlier
sections for the case of castings, by casting size and
geometry, type of casting process, process operation
parameters and alloy characteristics. Most of these

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

proposed mechanisms are only of academic interest,


having little relevance to the control of the CET.
However, some criteria for the transition have been
suggested, e.g. based on equiaxed grain size, volume
fraction of equiaxed grains or solutal interactions
between the growing equiaxed dendrites and the
advancing columnar front, that can be used in CET
models.
As stated by Kisakurek,9 two possibilities for the
transition exist. Either
(i) columnar growth may slow down allowing the
dominance of equiaxed grains or
(ii) equiaxed grains take over and force the columnar
growth to halt.
The proposed mechanisms therefore describe likely
scenarios leading to one or other of these two conditions.
Winegard and Chalmers12 suggested that the CET
occurs by the impingement of the columnar dendritic
interface upon the dendritic skeleton of floating
equiaxed crystals.
Biloni and Chalmers20 proposed that the floating
equiaxed crystals probably grow initially in a predendritic manner and that each crystal is surrounded by its
diffusion field of higher solute content. They suggest
that columnar growth is arrested when the diffusion
fields of the floating crystals, ahead of the columnars,
have impinged. A cellular dendritic sub-structure then
develops in the crystals.
Fredriksson and Hillert60 suggested that the central
equiaxed zone observed in experimental ingots of a Pb
2%Sb alloy was formed by two different mechanisms,
growth of a sedimentation layer on the bottom and
growth of crystals which have adhered to the vertical
solidification front. In the latter case, it was anticipated
that, as solidification progresses, the vertical columnar
solidification front will become jagged and be able to
catch a floating crystal. Roughness therefore increases
and the chance of catching more crystals is increased.
This process may then spread along the whole solidification front. The subtlety of this argument is that the CET
is caused by the development of individual equiaxed
grains. Southin15 has also suggested that dendrites/
dendrite fragments, dislodged from a surface dendritic
layer, sink in the liquid until they are met by solid
growing from the mould walls. In a later paper,
Fredriksson and Olsson61 reiterate that the CET occurs
when the free equiaxed crystals are sufficiently large or
numerous to physically block columnar growth by
adhering to the solidification front. From observations
on a steel ingot, they also proposed a possible criterion
for the CET, namely that it occurs at the time when the
temperature of the bulk liquid reaches a minimum
before recalescence.
Based on a consideration of the relationship between
dendrite tip temperatures and growth rates for different
temperature gradients, Burden and Hunt62 argued that
as solidification progresses, the growth rate of the
equiaxed crystals ahead of the columnar front rapidly
increases and the growth rate of the columnar interface
decreases. A condition is eventually reached where the
columnar front has almost stopped and the equiaxed
grains are growing rapidly leading to the CET.
Witzke et al.63 modelled the composition and
temperature field ahead of a vertical columnar front in
the presence of convection and suggested that the CET

Columnar to equiaxed grain transition in as solidified alloys

occurs only if the liquid reaches a sufficient degree of


constitutional supercooling and, second, if the volume of
the undercooled zone is sufficient. These two conditions
are required to generate the necessary amount of
equiaxed crystals to hinder columnar growth. Lipton
et al.64 have also suggested that the CET is governed by
the undercooling of the columnar interface and the
thickness of the undercooled boundary layer at the
interface. This layer determines the growth rate of
the equiaxed crystals. For an organic analogue, Lipton
et al.65 observed that release of latent heat during
equiaxed growth caused a rise in temperature of the
boundary layer that accompanied the CET. They
consider that the rise in temperature is because of the
overlap of the thermal boundary layers of adjacent
equiaxed grains (considered to be spheres). An alternative criterion for the CET was therefore suggested,
that it will occur when an equiaxed grain has grown to a
critical radius equal to one-tenth of the distance between
two equiaxed grains.
Hunt,32 using a probability approach to consider the
growth interaction of columnar and equiaxed grains,
derived another criterion for the CET. He argued that
equiaxed growth will occur when the volume fraction of
equiaxed grains (again considered as spheres) is greater
than 0.49 (or an extended volume fraction of 0.66).
Mahapatra and Weinberg30 proposed that the columnar dendrite tips may become unstable when the
temperature gradient ahead of the tips falls below a
critical value. This might then lead to solute accumulation at the tips and restricted growth of the columnar
grains. The liquid ahead of the columnars would then
cool and equiaxed grain nuclei would grow.
Gandin66 applied a one-dimensional finite difference
heat transfer model to the studies of the non-steady
directional freezing of 99.99 wt-%Al and Al7 wt-%Si
described elsewhere.37 It was observed that if the ingot
solidified fully columnar, then the velocity of the
columnar dendrite tips increased initially during the
stage of superheat loss, then decreased when no
substantial gradient remained in the liquid ahead of
the growing interface. When applied to the freezing of
AlSi alloys, the maximum velocity was reached when
the dendrite tip interface reached two-thirds of the
length of the ingot. This position corresponded closely
to the position of the experimentally observed CET
which led the author to propose that at this maximum
velocity, dendrite arm remelting occurs resulting in
destabilisation of the macroscopic interface. Since the
predicted position of maximum velocity as a function of
heat extraction rate, the superheat and the alloy
composition is similar to that reported for the CET, he
proposes a CET criterion based on the position of the
maximum velocity of the columnar dendritic interface.
He comments that since the criterion is independent of
the nucleation parameters of the equiaxed grains, it
should only be applied in the absence of added
inoculants.
Martorano et al.45 proposed a new mechanism for
the transition based on solutal interactions between the
equiaxed grains and the advancing columnar front. The
solute blocking is achieved by basing the undercooling
that drives dendrite tip growth on the average solute
concentration of the liquid surrounding the grain
envelopes (extradendritic liquid), Cl, instead of the

International Materials Reviews

2006

VOL

51

NO

259

Spittle

Columnar to equiaxed grain transition in as solidified alloys

initial alloy composition. When the solute rejected from


the equiaxed grains is sufficient to dissipate the undercooling at the columnar front, such that Cl has increased
to C*l (where C*l is the liquidus concentration given by
the phase diagram at temperature T), the CET will
occur.

Overall consideration of termination of


columnar growth

Published by Maney Publishing (c) IOM Communications Ltd

The proposed mechanisms/criteria for the termination of


columnar growth aid understanding of the CET. If all of
the solidification phenomena and length scales associated with the CET could be incorporated into a single
computational model, then the mechanisms of importance would be naturally revealed. Proposed criteria for
the termination have been incorporated into simplified
models of the CET. The most widely used to date has
been that due to Hunt.32

260

Models for predicting CET


Over the past 20 years, modelling has played a
significant role in the continued interest in the CET.
To simplify the problem, the majority of these models
have considered the unidirectional solidification of
alloys. As discussed earlier, if true unidirectional freezing occurs in the absence of fluid flow, the mechanism of
equiaxed grain formation that is usually assumed is
heterogeneous nucleation in the supercooled liquid
ahead of the columnar front. Models can attempt to
predict the critical conditions at the columnar front that
give rise to the CET or, to match the predicted and
experimentally observed location of a CET for a specific
alloy and set of processing conditions or, to examine the
effect of different parameters on the transition.
CET models can be classified as being either
deterministic or stochastic. Stochastic models follow
the nucleation and growth of each individual grain
whereas deterministic models rely on averaged quantities and equations that are solved on a macroscopic
scale.45
Stochastic models simulate the solidification of both
columnar and equiaxed grains and computed micrographs can be created as the grain structure evolves. No
criterion is required in such models for determining the
position at which the CET occurs. A subjective
judgement is made, from examining the micrographs,
as to when columnar growth ceases.
Some reported deterministic models of the CET only
consider the growth of the columnar grains and attempt
to predict the conditions that exist at the locations of
experimentally observed columnar to equiaxed transitions.30,31,33 Many deterministic models involve consideration of the formation of the equiaxed grains and
their competition with the advancing columnar front.
Such models require the inclusion of a criterion for the
termination of columnar growth. They also attempt to
incorporate the growth kinetics of the columnar and
equiaxed grains. Often the growth of the dendrite tips in
both columnar and equiaxed grains is assumed to be
governed by a simple relationship between tip velocity
and undercooling. This undercooling has most often
been expressed in terms of the initial alloy composition
C0. However, as discussed by Martorano et al.,45 this
implies, for a given columnar dendrite tip temperature,

International Materials Reviews

2006

VOL

51

NO

that tip velocity is the same regardless of the presence of


equiaxed grains growing ahead of the columnar front.
They therefore developed a model where the columnar
and equiaxed dendrite growth velocities are a function
of a solutal undercooling proportional to the difference
between the local liquidus concentration and the local
average solute concentration in the extradendritic liquid,
see the section Proposed mechanisms/criteria for the
termination of columnar growth above. Deterministic
models incorporating equiaxed grains most often consider that the grains are static, although models are now
attempting to take into account equiaxed grain movement during solidification. In some models, nucleation
occurs at a fixed undercooling (termed growth models in
this review). In this type of model, the number of
equiaxed grain nuclei is often chosen to give the best
agreement between predicted and experimental data
relating to the CET for the assumed undercooling.
Alternatively, best agreement may be obtained by
determining the value of the undercooling assuming a
fixed number of nuclei. In other models, nucleation
kinetics are taken into account by considering the
influence of undercooling on nucleation rate (nucleation
and growth models).
Some stochastic models also incorporate dendrite tip
growth models.43 However, other stochastic models,
which simulate solidification at a microscopic scale and
allow the solution of solute diffusion equations during
solidification and the application of solid/liquid interface
front tracking methods to the surfaces of the individual
dendritic grains,67,68 have no need to include dendrite tip
growth models.
Although models are able to indicate probable effects
of alloy and process parameters on the CET, at present
they are unable to simulate the various mechanisms of
equiaxed grain formation and modes of interaction
between equiaxed and columnar grains found in omnidirectionally solidified alloys.

Deterministic models
Analytical models

A number of analytical models for the CET have been


proposed over the last half century that try to illustrate
the interrelationship of the alloy and processing conditions governing the transition.29,32,63
Tillers pioneering model29 considered the unidirectional freezing of a semi-infinite liquid from one end and
calculated the solute and temperature distributions in
the liquid with time assuming that the solid/liquid
interface advances at a rate proportional to (time)1/2.
It was assumed that nucleation of equiaxed grains
occurred in the constitutionally supercooled region
ahead of the interface and that this was directly
responsible for the CET. The transition from columnar
to equiaxed growth was assumed to occur at a critical
nucleation frequency corresponding to a critical maximum undercooling value for a particular alloy. He
argued that this critical condition is determined by G/R
alone. An expression was also derived for the ratio of
equiaxed zone length to columnar zone length in a real
finite ingot. This predicts that the columnar zone will
decrease as superheat increases, as the freezing range
increases, as the number and catalytic activity of the
nucleating centres increase and as stirring of the liquid

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

increases. The CET will again occur at a critical ratio of


G/R but will also depend on ingot length.
Witzke et al.63 considered the thermal and chemical
fields ahead of the tips of dendrites growing as a vertical
columnar front ahead of which liquid is flowing under
conditions of thermal laminar flow. They derived an
expression for constitutional supercooling which, as
stated in the section Proposed mechanisms/criteria for
the termination of columnar growth above, they
consider is responsible for the CET. This model again
accounts for the fact that development of the equiaxed
zone is favoured by low superheat, strong undercooling
at the dendrite tips, high values of thermal to chemical
diffusivity ratio and mould height.
Hunt32 developed a steady state directional solidification model for the columnar and equiaxed growth of
both dendrites and eutectic. An analysis was presented
for the growth of equiaxed grains in the supercooled
liquid ahead of the columnar front in order to predict
the conditions necessary for a wholly equiaxed structure.
As mentioned in the section Proposed mechanisms/
criteria for the termination of columnar growth above,
fully equiaxed growth, i.e. the CET, was considered to
occur when the extended volume fraction of equiaxed
grains exceeds 0.66. Hunt calculated that fully equiaxed
growth would occur when the steady state temperature
gradient was below a critical value given by
1=3
Gv0:617N0 f1{(DTN )3 =(DTC )3 gDTC

(5)

where N0 is the number of nuclei/unit volume, DTN is


the critical undercooling for nucleation and DTC is the
undercooling at the columnar front.
The model was applied to Al3 wt-%Cu. As the
volume fraction of equiaxed grains decreases, mixed
columnar-equiaxed and finally fully columnar structures
are anticipated. The model predicts that at low growth
velocities, equiaxed growth depends on the efficiency of
the nucleating substrates for the equiaxed grains
whereas at high gradients the number of nucleation
sites is more important. The columnar front undercooling can be expressed in the form
DTC ~f{A0 m(1{k)C0 V g1=2

(6)

It can therefore be seen that as the so-called growth


restriction factor, m(1k)C0, increases the tendency to
equiaxed growth increases.
Simulation models

In more recent years, a variety of approaches have been


employed to model the conditions giving rise to the
evolution of the CET for specific alloys by simulating
the solidification process. These models attempt to
include, to different extents, the thermodynamics and
kinetics of the transformation process.
Growth models
Lipton et al.64 assume that the columnar front is
.
advancing with a velocity t0 5 and that its undercooling
is governed by a velocity-undercooling relationship.
A specific number of equiaxed particles (necessarily
considered as a free parameter), occupying fixed
positions, are considered to grow as spheres at a rate
determined by the square root of the mean temperature
difference between the particle and its surrounding melt

Columnar to equiaxed grain transition in as solidified alloys

and of the mean dwell time in the melt. The model does
not take into account the latent heat released by the
growing equiaxed grains. As stated in the section
Proposed mechanisms/criteria for the termination of
columnar growth above, they assumed that the criterion
for the CET was that an equiaxed grain had grown to
one-tenth of the spacing between two adjacent equiaxed
grains and that convective flow equalises the temperature in the bath and controls the heat transfer to the
solid shell. The model showed that two parameters
controlled the CET, undercooling of the interface and
the thickness of the undercooled layer.
Fredriksson and Olsson61 have used a similar
.
approach assuming, again, a velocity t0 5 relationship
for the columnar front, that the front undercooling is
dependent on velocity, that convective flow equalises the
temperature in the bath and that there are a specified
number of growing equiaxed crystals. They assume that
the equiaxed grains grow with a spherical shape, having
an internal solid fraction of 0.3, and that the growth rate
is dependent on the square of the undercooling. Their
model accounts for the latent heat evolved by the
equiaxed grains. However, the growth rate and undercooling of the columnar interface are not affected by the
growth of the equiaxed grains. Best values were chosen,
for a growth constant and the number of equiaxed
dendrites per unit volume, to give a best fit between a
measured cooling curve recorded at the centre of a steel
ingot and a curve predicted from the model. Both curves
displayed recalescence and, as stated in the section
Proposed mechanisms/criteria for the termination of
columnar growth above, they predicted that the CET
occurred at the time corresponding to the minimum
temperature preceding recalescence. Their model predicts an increase in the columnar zone the larger the
superheat, the smaller the number of free crystals, the
lower the solute content, the higher the cooling rate and
the greater the height and/or width of the ingot.
Flood and Hunt69,70 were the first to simultaneously
model the growth of both the columnar and equiaxed
grains. This was achieved using a 1D finite difference
thermal model. The model takes into account both
conduction and convection in the bulk, the latent heat
liberated by the equiaxed grains and the thermal
interaction between the solidifying columnar and
equiaxed grains. In this way, the velocity and undercooling of the columnar front are calculated dynamically
throughout the simulation and are not fixed a priori.
Convection is treated with a simple boundary layer
approximation, it being assumed that there is complete
mixing ahead of a conducting boundary layer and that
the bulk is isothermal. It is assumed that the velocities of
the columnar and equiaxed grains are governed by a
parabolic velocity undercooling relationship. The liquid
fraction in the columnar grains is assumed to be related
to temperature by the Scheil equation,71 truncated at
an undercooling determined by the heat flow at the
dendrite tips.69 The spherically shaped equiaxed grains
ahead of the columnar front either grow as soon as some
critical undercooling is exceeded or are nucleated at a
temperature dependent rate. The equiaxed grains are
assumed to be isothermal and to also have a solid
fraction governed by the Scheil equation. Equiaxed
grain impingement is allowed for using an Avrami
treatment and the CET is assumed to occur when the

International Materials Reviews

2006

VOL

51

NO

261

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

262

Columnar to equiaxed grain transition in as solidified alloys

extended volume fraction of equiaxed grains exceeds


the value specified by Hunt32 (see the section Proposed
mechanisms/criteria for the termination of columnar
growth above). Assuming that equiaxed grains grow as
soon as the temperature falls below the liquidus, the
model predicts that the columnar range decreases on
increasing the alloy composition and convection, and on
increasing the initial superheat and rate of heat
extraction. The model also predicts that the equiaxed
growth dominates columnar growth more than in
practice, possibly pointing to factors that influence
equiaxed grain formation in addition to growth, such
as nucleation, convective flow and sedimentation. The
effect of superheat was not as dramatic as that found in
practice, which they conclude is probably as a result
of the model failing to consider the effect of superheat
on the remelting/survival of nuclei, e.g. as associated
with the big-bang mechanism of equiaxed grain formation. The introduction of a temperature dependent
nucleation rate increased the columnar zone, which
could be combated by increasing the cooling rate.
Mahapatra and Weinberg30 used a completely different approach to those so far described, in that a
numerical model was used to evaluate the CET observed
in experimental ingots. However, the model did not
include the formation of the equiaxed grains. Small
cylindrical ingots of SnPb alloys were melted in situ
and directionally solidified from a water cooled chill.
Temperature measurements were taken throughout
solidification from four thermocouples placed in the
melt at different heights. A 1D implicit finite difference
model was then used to simulate the solidification of the
ingots and heat transfer coefficients at the chill interface
were obtained, by iteration, that gave best fit between
modelled and measured temperatures. The finite difference model, with the appropriate heat transfer coefficient, was then used to determine the conditions at the
CET for each ingot. They concluded that the CET
occurred at a critical temperature gradient of 0.108,
0.101, and 0.126 K mm1, respectively for alloys containing 5, 10 and 15 wt-%Pb. They found that the CET
was independent of superheat and suggested a criterion
for the transition based on instability of the dendrite tips
below a critical temperature gradient (see the section
Proposed mechanisms/criteria for the termination of
columnar growth above). In a later paper, Ziv and
Weinberg31 also concluded, for an Al3 wt-%Cu alloy,
that the CET occurred when the temperature gradient
fell to 0.06 K mm1. Applying a similar 1D finite
difference heat transfer analysis to the data of
Mahapatra and Weinberg, Spittle and Tadayon72 could
find no evidence to support the contention that the CET
occurs at a critical gradient.
Suri et al.33 adopting a similar approach to that of
Weinberg and co-workers above, used a 1D heat
transfer model to identify the locations of the CET in
directionally solidified Al4.5 wt-%Cu alloys.
Brown and Spittle73 described a 2D implicit finite
difference model for simulating the solidification of
experimental AlCu alloy castings in dry sand moulds.
The experiments demonstrated, in line with the observations of Tarshis et al.21 that in small ingots, the
transition from wholly columnar to wholly equiaxed
structures occurs with only a very small increase in the
parameter P. In the model, the growth rates of the

International Materials Reviews

2006

VOL

51

NO

columnar and equiaxed grains are assumed to be equal,


the columnar grains grow with a fixed undercooling
of 1.0 K, the equiaxed grains grow as spheres with a
fixed internal solid fraction (as predicted by the Scheil
equation for an undercooling of 0.5 K), equiaxed grains
do not interact with each other or the growing columnar
front and latent heat generated from the equiaxed grains
is distributed uniformly to all nodes ahead of the
columnar front. The equiaxed grain density was
obtained from grain size measurements on the actual
castings. The CET was assumed to occur for a given
volume fraction of equiaxed grains. A best match
between predicted and measured thermal analysis curves
was obtained to provide the metal/mould heat transfer
coefficient. The model was able to simulate mixed
columnar equiaxed structures and indicated that the
CET can occur for a very low equiaxed grain density.
Wang and Beckermann74,75 reported a novel technique for modelling the CET under conditions of diffusion
controlled dendritic growth, i.e. in the absence of melt
convection and transport of solid grains. The technique
involves the multiphase approach described by Wang
and Beckermann76 in which a control volume, containing either columnar or equiaxed grains, is considered to
consist of three phases, solid, interdendritic liquid and
extradendritic liquid. A set of macroscopic equations
governing solute diffusion in the three-phase system is
derived using a volume averaging procedure. These
equations are then coupled with the heat flow equation
which is solved using a two time-step, fully implicit
control volume-based finite difference method. The
model is able to simulate the CET taking into account
nucleation and growth of grains and dendrite morphology (primary and secondary arm spacings). In the model
reported, an average nuclei density was assumed in the
equiaxed region, with nucleation occurring instantaneously at the liquidus temperature. The CET is
determined using the criterion due to Hunt.32 The
equiaxed nuclei density represents the only adjustable
parameter. A 1D model was applied to the simulation
of the experiments performed by Mahapatra and
Weinberg30 and Ziv and Weinberg31 on SnPb alloys
and AlCu (see above). Applying the model to the Sn
Pb alloys, it was observed that columnar zone length
increases as equiaxed grain density decreases, as the heat
transfer coefficient at the chill increases and as solute
concentration decreases. The effect of superheat was
only pronounced at low superheats (,,10 K), columnar length increasing with superheat. Assuming a single
equiaxed nuclei density for all experimental runs of
107 m3 resulted in a match of experimental and
predicted CET positions to within 20%. In the case of
the Al3 wt-%Cu alloy, the CET position and the time
when it occurs were fairly well predicted. Although it is
reasonable to compare the model with the SnPb and
AlCu data where convection would have been minimal,
Wang and Beckermann acknowledge that convection
and crystal fragmentation/transport need to be included
for situations where diffusion is not dominating.
In a similar manner to that used by Mahapatra and
Weinberg30 and Ziv and Weinberg,31 Gandin66 applied a
one-dimensional finite difference heat transfer model to
the non-steady columnar freezing of 99.99 wt-%Al and
Al7 wt-%Si alloys (equiaxed grain formation was not
considered in the model). The model indicated, as

Spittle

Published by Maney Publishing (c) IOM Communications Ltd

8 Plot of growth velocity versus temperature gradient for


Al3 wt-%Cu showing columnar and equiaxed regions
calculated using approximate analysis and more accurate analysis32

discussed in the section Proposed mechanisms/criteria


for the termination of columnar growth above, that the
experimentally observed CET occurred at the position
where the columnar front velocity was a maximum.
Martorano et al.45 used a modified version of the
multiphase/multiscale model of the CET proposed by
Wang and Beckermann.74 The model incorporated a
new criterion for the CET, as described in the section
Proposed mechanisms/criteria for the termination of
columnar growth above, based on solutal interactions
between the equiaxed grains and the advancing columnar front. The model was validated by predicting the
CET in the three AlSi alloys previously experimentally
studied by Gandin.37 Equiaxed grains were assumed to
grow at a fixed undercooling and nucleation undercoolings were determined that provided best agreement
between measured and calculated CET positions. For
each of the three compositions, the nucleation undercooling for the equiaxed grains was found to be about
equal to the maximum tip undercooling deduced by
Gandin.66 This finding was insensitive to the equiaxed
grain density. The authors say their data support the
suggestion by Gandin that the origin of the equiaxed
grains is breakdown or fragmentation of the columnar
dendrites, rather than heterogeneous nucleation.
Nucleation and growth models
Hunt,32 in addition to his analytical model (see the
subsection Analytical models above), also developed a
more rigorous heat flow model. This did not require
equiaxed nucleation to occur at a fixed undercooling and
the temperature gradient to remain constant during the
growth of the equiaxed grains. Heat flow through a unit
area box was considered, within which nucleation of
equiaxed grains took place at a rate dependent on the
amount of supercooling. The relationship between G and
V for the CET was again determined for Al3 wt-%Cu,
assuming that fully equiaxed growth occurred when the
extended volume fraction of equiaxed grains at the
columnar front reached 0.66. As can be seen in Fig. 8,
there was very close agreement with the predictions of
his analytical model.
Cockroft et al.77 reported the development of a
mathematical model based on Hunts CET model. To
calculate the columnar dendrite tip undercooling the
Kurz, Giovanola and Trivedi (KGT) model for a binary
alloy was used.78 This calculates the undercooling in
terms of G, R, C0, m and k. The mathematical model
was extended to multicomponent alloys by assuming

Columnar to equiaxed grain transition in as solidified alloys

that the diffusion fields around the dendrite tips


associated with each alloy species can be superimposed.
An expression due to Rappaz,79 which assumes a
continuous nucleation distribution of the form of a
Gaussian distribution, was used to calculate the density
of grains at a given undercooling. The model was fitted
to experimental data (G and R conditions) for the CET
transition in a land turbine blade alloy CMSX-4 by
adjusting the value of the centre of distribution
nucleation parameter, required in Rappazs expression.
As with the Hunt model,32 an extended volume fraction
of 0.66 for the equiaxed grains was assumed for the
CET. The value calculated was then used for general
application of the model to other casting geometries of
the alloy.
Gaumann et al.80 developed a numerical model as a
modification of Hunts models,32 which again considers
the growth of equiaxed grains ahead of the columnar
front. The CET is assumed to occur for an actual
volume fraction of equiaxed grains of 0.5 ahead of the
columnars. The model takes account of the composition/
liquidus profile and the change in undercooling with
distance ahead of the columnar front, at both low and
high velocities. The KGT dendrite model is used to
determine the columnar dendrite tip temperature. The
nucleation and growth rates of the equiaxed grains are
governed by the local undercooling which is dependent
on the distance from the front. The model was again
applied to Al3 wt-%Cu and the data compared with
those of Hunt (Fig. 9). At low temperatures, the CET
occurs at lower velocities than predicted by Hunt. This is
thought to be due to the underestimate of tip
temperature by Hunt. At high velocities, they predict
an increase in stability of the columnar structure,
thought to be due to rising tip temperatures. The
investigators contend that the model is more appropriate
for predicting the CET in a range of processes, including
the high cooling rates experienced in welding and rapid
solidification processes.
Quested and Greer2 have used a novel approach to
model the grain structures/grain size of directionally
solidified alloys which have been inoculated with a grain
refiner. Previous studies of the CET have primarily
focused on uninoculated melts. Modelling is complicated by the fact that grain refinement appears to be
very inefficient in alloy castings with, in the case of Al
TiB refiners, ,1% of the TiB2 nucleant substrates
actually nucleating grains. In this kinetic model, the
initiation and growth of equiaxed grains are considered
in order to determine final equiaxed grain size under
directional freezing conditions. The authors use the free
growth theory to describe initiation of grains.81 The
term initiation is used since the onset of free growth is
not nucleation controlled. Larger particles become
active nucleants at smaller undercoolings. The onset of
undercooling for free growth of a grain can be calculated
from the diameter of the inoculant particle. The
inoculant particles are assumed to be randomly dispersed and static and are associated with a known
particle size distribution in the master alloy. The
impingement of growing grains on inoculant particles
is thought to be the mechanism by which refinement
efficiency is reduced and a key assumption of the model
is that the fraction of particles rendered inactive at any
instant is equal to the fraction of the system affected by

International Materials Reviews

2006

VOL

51

NO

263

Spittle

Columnar to equiaxed grain transition in as solidified alloys

Published by Maney Publishing (c) IOM Communications Ltd

9 Comparison of data in Fig. 8 with that predicted from numerical model of Gaumann et al., where DTn is undercooling
at heterogeneous nucleation temperature, N0 is total number of heterogeneous nucleation sites per unit volume and
C0 is wt-%Cu80

impingement with growing grains. Grain size was found


to vary with cooling rate, variation of G or V having the
same effect. This contrasts with the opposing effects of G
and V on the CET. The CET locus on a log V versus
log G plot was calculated using the authors dendrite
growth modelling procedure and Hunts criterion32 for
blocking of columnar grain growth. By plotting the
contours of equal grain size on the same axes, it is
predicted that the equiaxed grain size at the CET could
vary over a wide range depending on the solidification
conditions (Fig. 10).
Ludwig and Wu82 have recently reported a threephase Eulerian approach to model the CET, where the
three phases are considered to be solidifying columnar
dendrites, solidifying equiaxed grains and the parent
melt. The phases are considered to be spatially interpenetrating and interacting continua. The nucleation
rate of the equiaxed grains is modelled with a heterogeneous nucleation law. The conservation equations of
mass, momentum, species and enthalpy for the phases
are solved with the commercial CFD software FLUENT
6.1. The model includes mixed columnar and equiaxed
solidification, the occurrence of the CET, melt convection and equiaxed grain sedimentation. The model was

10 Contours of equal grain size shown on plot of solidification velocity versus thermal gradient together with
locus for CET calculated from Hunts analysis2

264

International Materials Reviews

2006

VOL

51

NO

used to carry out an axisymmetric simulation of a Fe


0.34 wt-%C ingot 66 mm in diameter and 170 mm in
height. The solidification sequence, the sedimentation of
the equiaxed grains, the movement of the columnar
front and the final phase distributions were found to fit
generally accepted explanations of experimental findings. However, no quantitative evaluation was made.

Stochastic models
Simulating the solidification of individual grains provides the potential for examining more precisely the
local conditions controlling the CET. Modelling at this
scale has, in more recent models, enabled the solution of
solute diffusion equations and microscopic tracking of
the movement of the solid/liquid interface of the grains.
This has removed the need to include dendrite tip
growth models for the grains and has resulted in the
simulation of dendritic grain morphologies that have a
realistic appearance. However, assumptions are still
required regarding the nucleation of the grains. The
models are also severely limited by the grain number
densities that can be handled in a simulation. The
significant feature of the models is that they are capable
of creating computed micrographs of grain structures
that closely resemble those found in castings, etc.
Various researchers have reported the use of 2D
stochastic models to simulate the evolution of grain
structures and the CET during solidification.
Brown and Spittle,83 using a Monte Carlo computer
simulation technique previously applied to a number of
solid-state processes such as recrystallisation and grain
growth, were the first to use the method for solidification
transformations. The technique employs a lattice of
sites, squares or triangles, which are initially designated
as liquid. The transformation of these sites to solid is
then simulated according to a set of rules governing
nucleation and growth. Individual grains are identified
and tracked during solidification and the final grain
structure can be drawn as a micrograph. Using
appropriate densities and locations of nuclei, realistic
pictures of CETs were generated (Fig. 11). The model
does not use real temperatures, actual phase diagram
parameters or any thermophysical properties data. It
was found that the CET occurred, in agreement with the
probabilistic argument of Hunt,32 when the actual

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

11 Computer generated columnar to equiaxed transition


using a probabilistic numerical model83

volume fraction of equiaxed grains at the columnar


interface was 0.5. In later papers,8486 the investigators
demonstrated the relative effects of superheat and mould
temperature,84 C0 and P85 and equiaxed grain settling86
on the CET. The modelled effects were in agreement
with experimental observations, see the section
Parameters influencing the CET above. These papers
demonstrated that such models are effective tools for
qualitatively examining the CET.
Rappaz and Gandin87 developed a 2D cellular
automaton method for simulating the solidification of
dendritic alloys held at a uniform temperature. Again
the growth of individual grains is tracked but an attempt
is made to incorporate some of the physics of solidification. Nucleation is treated using temperature dependent
nucleation site distributions. The model also accounts
for the growth kinetics of the dendrite tips and the
preferred growth directions in cubic metals. The model
demonstrated that columnar length decreases as composition and cooling rate increases. The model is limited
by the assumption of uniform temperature.
Stochastic modelling routines for tracking and computational illustration of grain structure evolution have
been linked with finite difference/finite element methods
for solving the heat transfer equation and/or the solute
diffusion equation for actual alloys.43,67,68,88,89
Zhu and Smith89 developed an improved Monte Carlo
model, removing some of the limitations of the model
reported by Brown and Spittle83 in order that it could be
applied to a real system. Again the growth of individual
grains was tracked, but the number of nuclei was related
to undercooling and a probability model of crystal
growth was based on a lowest free energy change
algorithm. The Monte Carlo procedures were also
coupled with 2D finite difference formulations for heat
transfer and solute distribution. The model was applied
to the prediction of the multidirectional freezing of a 2D
ingot of Al4.5%Cu. In examining the CET, it was
observed that if the number of growing nuclei ahead of
the columnar interface is not large, then the growing
grains will be trapped by the columnars which can still

Columnar to equiaxed grain transition in as solidified alloys

continue to grow. The CET will only occur if the


number of growing nuclei in the bulk is large enough to
make them spread over the advancing interface. It was
also commented that, in the presence of a steep gradient
ahead of the columnars, equiaxed growth could be
confined to a narrow zone and equiaxed grain formation
would then progress inwards in a gradual manner.
Gandin and Rappaz43 reported a coupled, cellular
automaton-finite element (CA-FE) model for the 2D
simulation of the directional solidification of an Al
7 wt-%Si alloy. The CA algorithm simulates the evolution of the columnar and equiaxed grains during
solidification and, takes into account nucleation and
growth kinetics and the preferred growth directions of
the grains. Temperatures predicted at nodal points using
an FE heat transfer model are used to calculate the
nucleation and growth of grains with the CA model. The
latent heat released during solidification, determined
from the nucleation and growth CA model, is then
incorporated back into the FE heat transfer model. A
1D implicit formulation was used to solve the heat flow
equation. Nucleation in the bulk and at the surface of
the casting is described by Gaussian distributions79 as
undercooling increases. For the bulk liquid, undercoolings randomly generated with a Gaussian distribution are attributed to randomly chosen cells. If the local
undercooling of a cell exceeds the prescribed undercooling, a new grain forms. A different Gaussian
distribution is used for the surface cells. The nucleation
and growth of the individual grains are tracked and their
final shapes in the solidified alloy reproduced in
computed micrographs. An Al7 wt-%Si ingot was
experimentally melted in situ and directionally solidified
from a cooled chill. The temperatures during solidification were measured from seven thermocouples along the
ingot length. The parameters of the Gaussian distribution for the nucleation of grains in the bulk were
adjusted so that model predictions matched the experimental cooling curves and grain structures. Many of the
features of the structure seen in the actual alloy were
reproduced. In the early stages of freezing, some grains
nucleated in the bulk liquid were encapsulated by the
advancing columnar grains, in agreement with earlier
predictions.32,89 At a later stage, the density of grains
growing in the bulk is sufficient to stop columnar growth
but the grains continue to grow in an elongated fashion.
Further up the ingot, as the temperature gradient
decreases, the density of grains in the bulk increases
and their shape becomes equiaxed. This opens up a
debate as to which transition constitutes the CET, the
one to elongated or the one to equiaxed grains. The
model has better descriptions of dendritic growth than
the Zhu and Smith model.89
Nastac67 developed a comprehensive model for
simulating the evolution of dendritic crystals during
the solidification of binary alloys. Stochastic procedures
were used to control the nucleation and growth of the
dendrites and, finite-difference schemes were used to
calculate the temperature and concentration fields in
the solid and liquid phases. The author reported
investigation of the CET, using 2D models, for both
multidirectional and unidirectional solidification of
IN7185 wt-%Nb alloy. In the unidirectional model,
the modelled geometry was 10 mm620 mm and the
number of cells was 50061000 (i.e. a 20 mm mesh size).

International Materials Reviews

2006

VOL

51

NO

265

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

266

Columnar to equiaxed grain transition in as solidified alloys

12 Predicted microstructural evolution for Al3 wt-%Cu alloy68

It was observed from the unidirectional model that the


CET was sharp which was attributed to the faster
growth of equiaxed dendrites that have nucleated in the
undercooled liquid ahead of the columnar front.
Dong and Lee68 have also used a CA-FD model to
simulate the CET during the directional solidification of
AlCu alloys. The CA-FE model of Gandin and Rappaz
[Gan04] only considers the thermal field and assumes
that equiaxed grains are heterogeneously nucleated in
undercooled liquid ahead of the growing columnar
front. The Dong and Lee model, based on an earlier
reported model90 but modified to include nucleation
ahead of the solidifying interface, has been used to
predict the influence of solutal interactions within an
advancing columnar dendritic network on the nucleation of equiaxed grains. The influence of thermal
gradient, pulling velocity, alloy composition, crystallographic orientation of the columnar dendrites and
nucleation parameters on the CET was also investigated.
The modelled results were compared with previous
models and experimental studies. The CA-FD model
does not use an analytical solution78 to determine the
velocity of the solid/liquid interface, common in other
stochastic models, but solves the solute conservation
equation subject to an equilibrium condition at the solid/
liquid interface. The model determines the solute build
up ahead of the interface and the solutally undercooled
region. This undercooling is used to stochastically
determine the nucleation of equiaxed grains using a
Gaussian nucleation distribution relating grain density
increase to increased effective undercooling. Calculated
numbers of nuclei are distributed randomly among cells.
Thermal profiles for directional solidification were
assumed, where a constant temperature gradient with
horizontal isothermals was imposed moving upwards
with a constant velocity. Simulations were performed on
a regular 2D square grid with a cell edge length of 5 mm.
The domain size was 3 mm wide by 6 mm high

International Materials Reviews

2006

VOL

51

NO

(60061200 cells). Simulations were performed for Al


Cu alloys and phase diagram data were generated from
MTDATA.91 Dong and Lee modelled the solidification
of an Al3 wt-%Cu alloy with G53.0 K mm1 and V
varying as a function of solidification time to give values
similar to those in the non-steady studies of Ares and
Schvezov38 and Siqueira et al.40 The nucleation undercooling was assumed to be close to the maximum
columnar tip undercooling and was set to 5 K. The
maximum nucleation density was set at a value of
1.061012 m3, as used in a previous study.40 The
predicted evolution of the structure is shown in
Fig. 12. An interesting prediction of the model is that,
unlike all prior CET models where equiaxed grains are
assumed to nucleate ahead of the advancing columnar
front, equiaxed grains are predicted to nucleate both at
the front and in the grooves between primary dendrites.
The authors predict that the grooves have the highest
solute-adjusted undercooling and are the favoured
nucleation location. As velocity increases, grains nucleate ahead of the front as well as between the dendrite
tips leading to solutal and/or mechanical blocking and
eventually to the CET. The authors defined the CET as
occurring when more than half the columnar dendrites
are blocked by equiaxed grains. To investigate the
influence of incorporating solute diffusion in the CET
model, the investigators compared the model to one
using the KGT model78 that relates growth velocity to
tip undercooling. It was found that by neglecting solute
diffusion, the predicted CET occurs at a lower velocity
than actually required for a given gradient. The
influence of the crystallographic orientation of the
columnar dendrites on the CET was found to be
negligible. Using the value for maximum nucleation
density given above, a log V versus log G map was
determined for the CET in Al3 wt-%Cu. The CA-FD
predictions are shown in Fig. 13 together with Hunts
analytical prediction32 and the experimental results of

Spittle

Published by Maney Publishing (c) IOM Communications Ltd

13 Predicted CET process map; filled circles columnar,


open circles equiaxed, CET predicted from Hunts
model with DTN of 0.75 K, filled diamond and triangle
experimentally determined CET conditions40 for Al
2 and Al4 wt-%Cu, respectively68

Siqueira et al.40 The predictions agree well with Hunts


analytical model at low velocities but differ significantly
at high velocities. The difference between the CA-FD
predictions and the experimental results is less than that
between Hunts predictions and the experimental
observations.
Beckermann and co-workers are currently applying
2D phase-field simulations to coupled columnar and
equiaxed growth. Beckermann et al.92 reported a model
for the unidirectional solidification of a binary alloy
under an imposed temperature gradient. Competition
between columnar and equiaxed dendrites was investigated as a function of nucleation undercooling, equiaxed
grain density, imposed cooling rate and temperature
gradient.

CET in welds
Several reviews have been published that consider the
detailed relationship between solidification parameters
and weld structures9396 and one of these96 concentrates
on the CET. It is recognised that equiaxed grain
structures at the centres of welds would be of benefit
in reducing/eliminating susceptibility to centreline cracking. They would also improve the ultrasonic inspection
of certain alloys, e.g. austenitic stainless steels. In
practice, columnar equiaxed transitions are rarely
observed in welds and obtaining an equiaxed structure
is often not achievable or difficult to optimise.96 As
would be expected, equiaxed regions are found, if at all,
at the centreline where the solidification rates are fastest
and the temperature gradients lowest. The nature of
welding processes and the cooling rates involved (102
107 K s1) make quantitative experimental evaluation of
the CET virtually impossible. Also, because of weld pool
shape, care must be taken to ensure that an apparent
CET is real and that the equiaxed grains are not
simply transverse sections through columnar grains. The
mechanisms proposed to account for the CET (in the
absence of inoculants) are those previously suggested in
the section Mechanisms of equiaxed grain formation
for castings, with the added possibility of grain detachment in the partially molten region (see the section
Overall consideration of proposed mechanisms of
equiaxed grain formation above). Since the fusion zone
boundary in a weld is at the freezing temperature, weld

Columnar to equiaxed grain transition in as solidified alloys

pool solidification usually proceeds by epitaxial growth


from grains of the parent metal. This therefore
eliminates mechanisms of equiaxed grain formation
that, in casting, may originate at the onset of freezing
as a result of heterogeneous nucleation.13,16 Equiaxed
grain formation because of dendrite remelting is often
suggested, owing to the considerable fluid flow and
thermal fluctuations in the weld pool.
Those parameters that are important in determining
the CET in castings, namely G, R, C0 and columnar
front undercooling, will also be of importance to weld
structures. In welds that completely penetrate a thin
sheet, the solidification velocity R is related to the
welding velocity, U, by
R~Ucosh

(7)

where h is the angle between the welding and local


growth directions.96 h decreases towards the centreline.
In the absence of experimental data of the solidification
conditions that exist during the CET, Hunts model,32
see the section Analytical models above, remains the
basis for explaining the CET in welds.96 The model
could only be used to attempt to predict the CET if
the pool thermal conditions were known. However, the
relationship between weld pool thermal conditions and
the controllable weld process parameters are not
established. In order to relate the parameters to the
thermal conditions in the pool, Clarke et al.97 have
reported the development of a 3D finite element
thermofluids model for simulating steady state, fullpenetration, moving gas tungsten arc (GTA) welds.
Temperature predictions were compared with measured
values in AlCu welds and the welding process parameters were incorporated in the simulations. The model
predicts that a CET is favoured by decreasing welding
speed for a given current or using a high current and
welding speed combination.
Kerr96 summarises a number of techniques that have
been used to increase the equiaxed fraction in welds.
Excluding inoculation, these include alloying to promote
constitutional supercooling/dendrite fragmentation, surface nucleation, e.g. use of an argon jet on a pool surface
ahead of the arc in GTA welds of AlMg alloys,
imposed electromagnetic stirring in GTA and GMA
welds, arc oscillation, e.g. using mechanical vibration
and the use of pulsed and modulated currents.

Summary
The probable origins of the equiaxed grains associated
with a CET (in the absence of added inoculants) are now
well defined and accepted. During solidification, one or
more of them will operate and which ones become active
or dominant will depend on factors such as cooling rate,
freezing time, the temperature distribution ahead of the
columnar front, thermal and constitutional supercooling, and the number and efficiency of potential nucleants
in the melt. In turn, these are governed by the alloy
system, the amount of solute present, section size, type
of solidification process (e.g. casting/welding/directional
solidification) and processing conditions (e.g. superheat,
mould temperatures, extent of fluid flow, etc.). It is
evident that the manner of termination of columnar
growth will differ according to the number and rate of
growth of the equiaxed grains and the degree of

International Materials Reviews

2006

VOL

51

NO

267

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

268

Columnar to equiaxed grain transition in as solidified alloys

sedimentation that occurs. In small laboratory ingots, a


large number of equiaxed grain nuclei may be formed
near the mould wall at the time of pouring, owing to
thermal supercooling and growth restriction of the
grains caused by solute redistribution. At low superheats, the supercooling may extend throughout the bulk
liquid almost from the time of pouring. Because of the
high density of rapidly growing equiaxed grains that
may exist throughout the liquid, the CET will be sharp
and occur over a large part of the columnar front at a
similar time. With increasing ingot size and solidification
time, equiaxed grain settling will become more important in determining the CET and the shape of the
equiaxed zone. It is also unlikely that the equiaxed
grains now originate at the mould walls and other
mechanisms such as dendrite fragmentation become
important. The CET will occur gradually over the
advancing columnar front either due to equiaxed grain
settling onto the lower part of the front or attachment of
individual grains to the vertical faces. An insufficient
fraction of grains reaching/attaching to the front may
lead to entrapment of equiaxed grains in the columnar
zone. Other solidification processes such as directional
freezing and welding are characterised by steep gradients
ahead of the front, however, in the former, there may be
little or no flow in the melt whereas in the latter, it can be
considerable.
It is almost impossible to compare the various sets of
experimental data obtained from directional freezing
studies of the CET because of the wide range of alloy
systems and differing compositions investigated, the fact
that some studies have been performed under steady
growth conditions (Bridgman growth) and others under
non-steady conditions, and different mechanisms of
equiaxed grain formation may have been operative in
the different studies. Although certain generalisations
can be made, i.e. that the CET is favoured by high V,
low G, high solute levels and easier nucleation, there is
no consistency regarding the proposed interaction of the
alloy and processing parameters and the criteria for the
transition. The experiments do however provide necessary data required by modellers to validate models.
Most of the models of the CET are 1D or 2D
descriptions of the directional solidification of alloys
that assume that static equiaxed grains heterogeneously
nucleated ahead of the columnar front evolve until they
terminate columnar growth. Except in the case of
stochastic models where the growth of the individual
grains is tracked, some criterion usually has to be
assumed for the point of termination. Of necessity, all
models require assumptions regarding the nucleation of
the equiaxed grains. These cannot accurately account
for the origins of equiaxed grains, their movement due
to fluid flow and settling, and their possible remelting.
Deterministic models permit modelling at the scale of
actual castings, directionally solidified alloys, etc. However,
they do not permit the accurate modelling of those
microstructural factors that are likely to contribute to
the CET such as solute diffusion, interdendritic convection, evolution of equiaxed grain morphology, etc.
Stochastic models offer the opportunity to examine the
effects of microstructural factors but are severely limited
by the sizes of the domains that can be modelled. It is
evident that the CET is governed by a complex set of
interacting phenomena including the local thermal and

International Materials Reviews

2006

VOL

51

NO

solute fields and the number/size of the equiaxed grains.


Recent models have attempted to simulate deterministically45 and stochastically68 the simultaneous growth
and interaction of the columnar and equiaxed grains and
the associated solute and thermal fields leading to the
CET. Future models must continue in this way if
understanding of the CET is to significantly improve.
However, as generally recognised, a significant factor
limiting the accuracy of models is the description of
nucleation undercooling and the number density of
equiaxed grain nuclei.
Although most models are capable of qualitatively
demonstrating the influence of various parameters on
the CET, as stated by Hansen et al.,98 the modelling
procedures must be classified, at present, as curve-fitting
exercises to experimental data. In their paper,98 they
considered whether it is possible to model the CET in a
physically realistic way. They discussed the various
stages that would be involved in such a model by
assuming that the equiaxed grains originate from
dendrite fragments generated in the columnar zone.
These would then be transported into the bulk liquid by
fluid flow where, as a result of growth and sedimentation, they cause columnar blocking.
It is evident that the CET and the factors influencing
it are qualitatively quite well understood. Any improvement in the predictive capabilities of existing models of
the CET will only be achieved by more accurately
simulating the physical aspects of the formation of
equiaxed grains, their movement in the melt and their
interaction with the columnar front.

References
1.
2.
3.
4.
5.
6.
7.
8.
9.

10.
11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.

J. A. Spittle: Mater. Sci. Technol., 2005, 21, 546.


T. E. Quested and A. L. Greer: Acta Mater., 2005, 53, 4643.
J. E. Stead: Engineering, 1906, 82, 4054.
H. M. Howe: J. Iron Steel Inst., 1916, 94, 181.
R. Genders: J. Inst. Met., 1926, 35, 259.
F. R. Hensel: PhD thesis, University of Berlin, 1929.
W. A. Tiller, K. A. Jackson, J. Rutter and B. Chalmers: Acta
Metall., 1953, 1, 428.
S. C. Flood and J. D. Hunt: ASM Metal Handbook, 1990, 15, 130.
S. E. Kisakurek: Formation of the equiaxed zone in ingot
castings, Bogazici University Research Centre Publications
No. 104, Istanbul, 1981.
J. Hutt and D. St. John: Int. J. Cast Met. Res., 1998, 11, 13.
L. Northcott: J. Inst. Met., 193839, 65, 173.
W. C. Winegard and B. Chalmers: Trans. ASM., 1954, 216, 1214.
B. Chalmers: Aust. J. Inst. Met., 1963, 8, 255.
K. A. Jackson, J. D. Hunt, D. R. Uhlmann and T. P. Seward, III:
Trans. AIME, 1966, 236, 149.
R. T. Southin: Trans. Met. Soc. AIME, 1967, 239, 220.
A. Ohno, T. Motegi and H. Soda: Trans. ISIJ, 1971, 11, 18.
B. P. Pearce and H. W. Kerr: Metall. Trans. B, 1981, 12B, 479.
J. L. Walker: in Physical chemistry of process metallurgy, II, 845;
1961, New York, Interscience.
G. S. Cole and G. F. Bolling: Trans. AIME, 1965, 233, 1568.
H. Biloni and B. Chalmers: J. Mater. Sci., 1968, 3, 139.
L. A. Tarshis, J. L. Walker and J. W. Rutter: Metall. Trans., 1971,
2, 2589.
J. A. Spittle and S. Sadli: Mater. Sci. Technol., 1995, 11, 533.
R. D. Doherty, P. D. Cooper, M. H. Bradbury and F. J. Honey:
Metall. Trans. A, 1977, 8A, 397.
M. Abdel-Reihim, N. Hess, W. Reif and M. E. J. Birch: J. Mater.
Sci., 1987, 22, 213.
J. A. Spittle, J. M. Keeble and M. Al Meshhedani: in Light metals
1997, (ed. R. Huglen), 795800; 2001, Warrendale, PA, TMS.
J. A. Spittle: Met. Sci., 1977, 11, 578.
T. S. Plaskett and W. C. Winegard: Trans. ASM, 1959, 51, 222.
R. Elliott: British Foundryman, 1964, 9, 398.
W. A. Tiller: Trans. Met. Soc. AIME, 1962, 224, 448.

Published by Maney Publishing (c) IOM Communications Ltd

Spittle

30. R. B. Mahapatra and F. Weinberg: Metall. Trans. B, 1987, 18B,


425.
31. I. Ziv and F. Weinberg: Metall. Trans. B, 1989, 20B, 731.
32. J. D. Hunt: Mater. Sci. Eng., 1984, 65, 75.
33. V. K. Suri, N. El-Kaddah and J. T. Berry: Trans. AFS, 1991, 99,
187.
34. S. Kim and R. N. Grugel: Metall. Trans. A, 1992, 23A, 1807.
35. L. J. Ledgard and D. G. McCartney: Proc. 4th Decennial Int.
Conf. on Solidification processing, Sheffield, UK, 1997, 227.
36. T. M. Pollock and W. H. Murphy: Metall. Mater. Trans. A, 1996,
27A, 1081.
37. Ch.-A. Gandin: ISIJ Int., 2000, 40, 971.
38. A. E. Ares and C. E. Schvezov: Metall. Mater. Trans. A, 2000, 31A,
1611.
39. M. A. Martorano and J. D. T. Capocchi: Int. J. Cast Met. Res.,
2000, 13, 49.
40. C. A. Siqueira, N. Cheung and A. Garcia: Metall. Mater. Trans. A,
2002, 33A, 2107.
41. M. Vandyoussefi and A. L. Greer: Acta Mater., 2002, 50, 1693.
42. Calcom SA, Parc Scientifique EPFL, Lausanne, Switzerland.
43. Ch.-A. Gandin and M. Rappaz: Acta Metall. Mater., 1994, 42,
2233.
44. L. Sturz, A. Drevermann, C. Pickmann and G. Zimmermann:
Mater. Sci. Eng., 2005, A 413414, 379.
45. M. A. Martorano, C. Beckermann and Ch.-A. Gandin: Metall.
Mater. Trans. A, 2003, 34A, 1657.
46. G. Reinhart, N. Mangelinck-Noel, H. Nguyen-Thi, T. Schenk,
J. Gastaldi, B. Billia, P. Pino, J. Hartwig and J. Baruchel: Mater.
Sci. Eng. A, 2005, 413-414, 384.
47. D. R. Uhlmann, T. P. Seward and B. Chalmers: Trans. AIME,
1966, 236, 527.
48. G. S. Cole and G. F. Bolling: Trans. AIME, 1966, 236, 1366.
49. G. S. Cole and G. F. Bolling: Trans. AIME, 1967, 239, 1824.
50. R. Morando, H. Biloni, G. S. Cole and G. F. Bolling: Metall.
Trans. A, 1970, 1A, 1405.
51. G. S. Cole, K. W. Casey and G. F. Bolling: Metall. Trans. A, 1970,
1A, 1413.
52. R. J. McDonald and J. D. Hunt: Trans. AIME, 1969, 245, 1993.
53. P. M. Thomas and J. A. Spittle: J. Inst. Met., 1971, 99, 167.
54. G. S. Cole and G. F. Bolling: Trans. AIME, 1968, 243, 1568.
55. G. S. Cole and G. F. Bolling: Trans. AIME, 1969, 245, 725.
56. J. A. Spittle, G. W. Delamore and R. W. Smith: The solidification
of metals, Iron and Steel Institute Publication 110, 318; 1968,
London, Iron and Steel Institute.
57. W. D. Griffiths and D. G. McCartney: Mater. Sci. Eng., 1996,
A216, 47.
58. B. Willers, S. Eckert, U. Michel, I. Haase and G. Zouhar: Mater.
Sci. Eng., 2005, A402, 55.
59. S. Eckert, B. Willers, P. A. Nikrityuk, K. Eckert, U. Michel and
G. Zouhar: Mater. Sci. Eng., 2005, A413-414, 211.
60. H. Fredriksson and M. Hillert: Metall. Trans., 1972, 3, 565.
61. H. Fredriksson and A. Olsson: Mater. Sci. Technol., 1986, 2, 508.
62. M. H. Burden and J. D. Hunt: Metall. Trans. A, 1975, 6A, 240.
63. S. Witzke, J. P. Riquet and F. Durand: Acta Metall., 1981, 29, 365.
64. J. Lipton, W. Kurz and W. Heinemann: Concast Technol. News,
1983, 22, 4.
65. J. Lipton, W. Heinemann and W. Kurz: Arch. Eisenhuttenwes.,
1984, 55, 195.
66. Ch.-A. Gandin: Acta Mater., 2000, 48, 2483.

Columnar to equiaxed grain transition in as solidified alloys

67. L. Nastac: Acta Mater., 1999, 47, 4253.


68. H. B. Dong and P. D. Lee: Acta Mater., 2005, 53, 659.
69. S. C. Flood and J. D. Hunt: Modeling of casting and welding
processes II, 207; 1983, American Institute of Mining,
Metallurgical and Petroleum Engineers.
70. S. C. Flood and J. D. Hunt: J. Cryst. Growth, 1987, 82, 552.
71. E. Scheil: Z. Metallkd., 1942, 34, 70.
72. J. A. Spittle and M. R. Tadayon: Cast Met., 1994, 7, 123.
73. S. G. R. Brown and J. A. Spittle: in Modeling of casting, welding
and advanced solidification processes V, 395; 1991, Warrendale,
PA, TMS.
74. C. Y. Wang and C. Beckermann: Metall. Mater. Trans. A, 1994,
25A, 1081.
75. C. Beckermann and C. Y. Wang: JOM, 1994, 42.
76. C. Y. Wang and C. Beckermann: Metall. Trans. A, 1993, 24A,
2787.
77. S. L. Cockcroft, M. Rappaz, A. Mitchell, J. Fernihough and A. J.
Schmalz: in Materials for advanced power engineering, part II,
(ed. D. Coutsouradis et al.), 1145; 1994, Netherlands, Kluwer
Academic Publishers.
78. W. Kurz, B. Giovanola and T. R. Trivedi: Acta Metall., 1986, 34,
823.
79. M. Rappaz: Int. Mater. Rev., 1989, 34, 93.
80. M. Gaumann, R. Trivedi and W. Kurz: Mater. Sci. Eng., 1997,
A226-228, 763.
81. A. L. Greer, A. M. Bunn, A. Tronche, P. V. Evans and D. J.
Bristow: Acta Mater., 2000, 48, 2823.
82. A. Ludwig and M. Wu: Mater. Sci. Eng., 2005, A413-414, 109.
83. S. G. R. Brown and J. A. Spittle: Mater. Sci. Technol., 1989, 5, 362.
84. J. A. Spittle and S. G. R. Brown: J. Mater. Sci., 1989, 23, 1777.
85. J. A. Spittle and S. G. R. Brown: Acta Metall., 1989, 37, 1803.
86. S. G. R. Brown and J. A. Spittle: Cast Met., 1990, 3, 18.
87. M. Rappaz and Ch.-A. Gandin: Acta Metall. Mater., 1993, 41, 345.
88. J. A. Spittle and S. G. R. Brown: in Numerical methods in thermal
problems, volume VI, (ed. R. W. Lewis and K. Morgan), 291;
1989, Swansea, Pineridge Press.
89. P. Zhu and R. W. Smith: Acta Metall. Mater., 1992, 40, 3369.
90. R. C. Atwood and P. D. Lee: Metall. Mater. Trans., B, 2002, 33B,
209.
91. R. H. Davies, A. T. Dinsdale, T. G. Chart, T. I. Barry and M. H.
Rand: High Temp. Sci., 1990, 26, 251.
92. C. Beckermann, A. Badillo and J. C. Ramirez: Invited Lecture,
JOM, 57(2), TMS Annual Meeting and Exhibition Program and
Guide, 2005, 33.
93. G. J. Davies and J. G. Garland: Int. Met. Rev., 1975, 20, 83.
94. S. A. David and J. M. Vitek: Int. Mater. Rev., 1989, 34, 213.
95. S. A. David and J. M. Vitek: in International trends in welding
science and technology, (ed. S. A. David and J. M. Vitek), 147;
1993, Materials Park, OH, ASM International.
96. H. W. Kerr: in International trends in welding science and
technology, (ed. S. A. David and J. M. Vitek), 157; 1993, Materials
Park, OH, ASM International.
97. J. Clarke, D. C. Weckman and H. W. Kerr: in Modeling of
casting, welding and advanced solidification processes VIII, (ed.
B. G. Thomas and C. Beckermann), 391; 1998, Warrendale, PA,
TMS.
98. G. Hansen, A. Hellawell, S. Z. Lu and R. S. Steube: Metall. Mater.
Trans. A, 1996, 27A, 569.

International Materials Reviews

2006

VOL

51

NO

269

Vous aimerez peut-être aussi