You are on page 1of 11

Proceedings of the 22nd

Ris International Symposium


on Materials Science:
Science of Metastable and Nanocrystalline Alloys
Structure, Properties and Modelling
Editors: A.R. Dinesen M. Eldrup D. Juul Jensen S. Linderoth
T.B. Pedersen N.H. Pryds A. Schrder Pedersen J.A. Wert
Ris National Laboratory, Roskilde, Denmark 2001










MICROSTRUCTURE DEVELOPMENT DURING RAPID
SOLIDIFICATION

B. Cantor

Department of Materials, University of Oxford, Parks Road, OX1
3PH, Oxford, UK



ABSTRACT

This paper discusses the development of microstructure during rapid solidification. A range of
different rapid solidification processes are considered, concentrating on melt spinning and spray
forming. A range of different microstructural features are considered, including porosity, grain size
and segregation. The paper concentrates on describing the heat and mass flow conditions during
rapid solidification, relating the processing conditions to the resulting microstructural features
observed in typical rapidly solidified materials, and explaining the physical mechanisms which
control the final rapidly solidified microstructures.


1. INTRODUCTION

Modern manufacturing industry uses increasingly sophisticated

processing methods in order to
produce complex materials architectures and microstructures and to meet stringent component
performance targets (Cantor 1997). Rapid solidification processing technology provides many
examples of these manufacturing trends. A wide variety of novel materials and components are
manufactured by rapid solidification processing techniques, with operating conditions precisely
controlled in order to obtain the desired material microstructure and properties (Cantor 1998).
Understanding the relationships between the processing conditions, the heat and mass flows taking
place during processing, and the resulting material microstructure is only achieved by a combination
of detailed process monitoring, computer modelling and material characterisation.

The objective of this paper is to provide a brief overview of the relationship between processing
conditions and microstructure for a variety of materials manufactured by rapid solidification
processing. The processing methods discussed include melt spinning, spray forming and related
techniques, which can be used to manufacture a wide variety of components such as Al-Si
compressor rotors and piston rings, stainless steel tubes, Al-Li airframe structures, Ni superalloy
aeroengine casings, wear resistant and thermal barrier coatings, Cu and Al bearing alloys, hard
magnets and Fe-based amorphous transformer cores (Cantor 1998, 1999, Cantor et al. 2001).
483
Cantor

The paper concentrates on describing the physical mechanisms by which the heat and mass flow
conditions during rapid solidification control microstructural features such as porosity, grain size
and segregation. This involves finding approximate solutions to the relevant heat and mass balance
equations, subject to boundary conditions involving processing parameters such as melt superheat
and ingot withdrawal rate. The key scientific problem is to understand the underlying physical
mechanisms, and the key applied problem is to relate the microstructure to the operating conditions,
so as to allow effective process control.


2. MELT SPINNING

There are a range of related rapid solidification processing methods in which a liquid is spread into
a long thin strip or sheet by bringing it into contact with a rapidly rotating wheel or drum. These
processing methods include chill-block melt spinning, melt extraction and planar flow casting. They
differ mainly in the method by which the liquid and the rotating drum are brought into contact. We
refer to these processes generically as melt spinning.

2.1 Heat and mass flow. The thin strip cools rapidly in good thermal contact with the drum, with a
thermal balance:

+ =
dt
df
L
dt
dT
C X T T h
d
) ( (1)

where h is the heat transfer coefficient at the strip-drum interface, T and T
d
are the liquid and drum
temperatures at the interface, , C and L are the density, specific heat and latent heat respectively, X
is the strip thickness, -dT/dt is the cooling rate, f is the fraction liquid and -df/dt is the solidification
rate. Strip thicknesses are typically 20-50m, with measured cooling rates of 10
4
-10
6
K/s (Gillen and
Cantor 1985, Vogt and Frommeyer 1985) and solidification times of a few ms, corresponding to
near-Newtonian conditions with heat transfer coefficients of 10
4
-10
5
W/m
2
K and little or no thermal
gradient in the strip (Kim and Cantor 1990, Cantor et al. 1991).

Thermal and momentum models of strip formation show that the strip thickness depends on the
drum velocity and the mass flow rate:


V
dt dm
a X =
2
(2)

where a is a constant, dm/dt is the mass flow rate and V is the drum velocity. As the velocity
increases, gas is increasingly incorporated between the strip and the drum, reducing the heat transfer
coefficient:


V
b
h =
2
(3)

where b is another constant. Excluding the solidification arrest, equation (1) shows that the average
cooling rate is directly proportional to the heat transfer coefficient and inversely proportional to the
strip thickness, so that the cooling rate increases linearly with drum velocity:

484
Microstructure development during rapid solidification
V
dt dm C
T T b a
XC
T T h
dt
dT
d d
4 / 1
) (
) ( ) (


=

= (4)

as demonstrated experimentally (Kim and Cantor 1990, Bewlay and Cantor 1986).

During solidification, the liquid undercools to the nucleation temperature T
N
before solidification is
initiated, followed by recalescence as latent heat is evolved rapidly. During recalescence, dT/dt is
large, solidification is adiabatic and the first term in equation (1) can be ignored. Integrating from
f=1 at T=T
N
:

) 1 ( f
C
L
T T
N
= (5)

i.e. the solid fraction increases linearly with increasing temperature. For a nucleation undercooling
greater than the hypercooling limit L/C, solidification is complete during recalescence. However,
the hypercooling limit is 200-500K for different materials, and measured nucleation undercoolings
are typically 100-200K, so solidification is not usually completed during recalescence (Cantor 1986,
Doherty et al. 1997). Nucleation undercoolings can be measured from the extent of partitionless
solidification during recalescence, and are found to increase linearly with cooling rate and therefore
with drum velocity (Cantor et al. 1991):

V
dt
dT
T
N
(6)

For a nucleation undercooling less than the hypercooling limit L/C, recalescence is followed by a
steady state growth at a constant plateau growth temperature T
G.
During steady state growth,
dT/dt=0 and the second term on the right hand side of equation (1) can be ignored, giving:

(7)
G S G
T L Lv T T h = = ) (

where v = -Xdf/dt is the solidification rate, is the kinetic coefficient for motion of the solid-liquid
interface, and T
G
is the steady state growth undercooling. In other words, extraction of heat into
the substrate drives the solid-liquid interface across the alloy layer with a velocity v. With h 10
5

W/m
2
K, (T
G
-T
S
) 1000 K and L 10
9
J/m
3
, the steady state solid-liquid interface velocity v is
typically 10cm/s, corresponding to a steady state growth undercooling of T
G
1 K (Cantor 1986,
Doherty et al. 1997).

2.2 Grain size. Solidification during melt spinning takes place by heterogeneous nucleation of the
undercooled liquid on the drum surface, followed by partitionless solidification of columnar grains
through the strip thickness during recalescence, and finally cellular breakdown and segregated
solidification near the free surface (Hayzelden et al. 1982, Lee et al. 1998).

Nucleated grains grow laterally across the drum surface until a stable set of columnar grains is
formed and then continue to solidify through the thickness of the melt spun strip. The grain size is
determined by a balance between the nucleation and growth rates on the drum surface. The average
cross sectional area of the columnar grains can be expressed independently in terms of the grain
size, nucleation rate and growth rate:

485
Cantor

Jt
vt d 1
2 2
2 2
=

(8)

where d is the grain diameter, v is the solidification velocity, t is the solidification time (notice that
the average grain grows for approximately half the solidification time), and J is the nucleation rate
per unit area, giving:


) exp(
)) exp( 1 ( 8 8
2
3
N
N
T B n
T A a
J
v
d


= =

(9)

where a is the atomic spacing, n is the nucleation site density, T
N
is the undercooling at which
nucleation takes place and A and B are constants which depend on material properties such as the
diffusion coefficient, surface energy and latent heat (Doherty et al. 1997).

Grains cannot nucleate near the melting point because there is insufficient driving force to
overcome the solid-liquid surface energy barrier, and cannot grow at low temperatures because
there is insufficient thermal energy to drive atomic diffusion. The grain size therefore decreases
rapidly, reaches a minimum and then increases rapidly again with increasing undercooling. Fine
scale microcrystalline and nanocrystalline grain structures depend, therefore, on a high drum
velocity to produce a high cooling rate and liquid undercooling. Typical melt spun grain sizes in
conventional alloys are ~1-2m (Hayzelden et al. 1982).

2.3 Segregation. Solidification structures access a wide range of solid-liquid interface undercoolings
and solid-liquid interface velocities, as the temperature rises during recalescence. This typically
leads to banded structures, as transitions take place between different solidification mechanisms and
different kinds of microstructure with decreasing interface undercooling and velocity.

In single phase alloys, for instance, the solidification mechanism varies with increasing interface
velocity and undercooling as follows (Trivedi 1994):

(1) Planar front solidification with normal macrosegregation, below the constitutional supercooling
limit v
C
at temperatures near the liquidus:


0
T
GD
v v
C

= < (10)

where G is the thermal gradient, D is the diffusion coefficient and T
o
is the freezing range.

(2) Cellular and dendritic solidification with interdendritic microsegregation, above the
constitutional supercooling limit v
C
between the liquidus and solidus:


0
T
GD
v v
C

> (11)

(3) Planar front solidification again with normal macrosegregation, above the absolute stability
velocity limit v
A
near the solidus:

L
A
T k
D T L
v v

= > (12)
486
Microstructure development during rapid solidification

where k is the partition coefficient, is the solid-liquid surface energy and T
L
is the liquidus
temperature.

(4) Partitionless planar front solidification above the solute trapping limit v
S
, which occurs well
below the solidus in narrow freezing range alloys, but above the solidus in wide freezing range
alloys:


a
D
v v
S
= > (13)

where a is the atomic spacing.

Undercoolings in melt spinning are ~100-200K (Cantor 1986, Bewlay and Cantor 1986, Doherty et
al. 1997). For narrow freezing range single phase alloys, therefore, nucleation of the undercooled
liquid on the drum surface is typically followed by partitionless solidification of the columnar grains
during recalescence, and then segregated solidification of the columnar grains during steady state
growth (Hayzelden et al. 1982, Lee et al. 1998). Typical microsegregation profiles are columnar
throughout the upper part of the columnar grains, with cell sizes and dendrite arm spacings of 0.1-
0.2m. In some alloys, however, nucleation is easier, leading to dendritic microsegregated
structures throughout. In other alloys, nucleation is delayed to lower temperatures, leading to
nanocrystalline, metastable crystalline, quasicrystalline and amorphous structures.

Solidification mechanisms and corresponding microstructures and segregation patterns are
considerably more complex in two-phase eutectic and peritectic alloy systems (Kim et al. 1990, Lee
et al. 1998).


3. SPRAY FORMING

There are a range of rapid solidification processes in which a spray of liquid droplets is scanned
across a substrate to deposit as a thin coating, or with multiple scans to form a larger free standing
object. These processing methods include spray deposition, arc spraying and plasma spraying,
which differ mainly in the method by which the liquid droplets are created. We refer to these
processes generically as spray forming.

3.1 Heat and mass flow. Dense droplet sprays are created by atomising a liquid stream. Atomised
droplet size distributions are broad, ranging typically from 5-500m, and can usually be
approximated as log normal (Bewlay and Cantor 1990):


=
2
2
2
) ln (ln
exp
2
1
s
r r
s
m

(14)

where is the frequency of droplets of radius r, r
m
is the mean droplet size and s is the standard
deviation. Sparser droplet sprays are created by melting a particulate feed as it passes through a
flame, electric arc or plasma. The droplet size distribution can then be controlled much more
precisely, and can be close to mono-sized (Zhao et al. 2000). Droplet sprays are highly turbulent,
with droplet velocities decaying exponentially and Gaussian-spreading laterally as they are carried
487
Cantor
to the substrate. Spray plume semi-angles are typically 10-25, with footprints at the deposition
point of 100-300mm radius.

Droplets cool with a thermal balance:

+ =
dt
df
L
dt
dT
C r T T h r
G

3 2
3
4
) ( 4 (15)

where h is the heat transfer coefficient at the droplet surface, T and T
g
are the droplet and
surrounding gas temperatures respectively, f is the droplet liquid fraction, and , C and L are the
density, specific heat and latent heat respectively. Droplet cooling and solidification are variable
depending upon the droplet size and trajectory. Small droplets cool and solidify rapidly whereas
large droplets cool only slowly, and the depositing spray consists of a mixture of small, cold and
fully-solidified particles, medium-sized mushy liquid/solid particles, and large, fully-liquid droplets
(Bewlay and Cantor 1990, 1991, Grant et al. 1991, 1993a).

Spray formed coatings, strip and sheet are manufactured by scanning the spray across the substrate,
and spray formed cylindrical billets and tubes are manufactured by rotating and withdrawing the
substrate. The throughput of material is described by a mass balance equation:

AW
dt
dm
= (16)

where dm/dt is the mass flow rate, and A and W are the cross section and withdrawal rate of the
resulting rapidly solidified material. The cross sectional dimensions of the spray formed product can
therefore be controlled by manipulating the mass flow rate and withdrawal rate.

There are significant spatial variations in deposit surface temperature during spray forming, with the
temperature decreasing radially towards the edge of the spray plume. At each position, the surface
temperature is low at first, because of the chilling effect of the underlying substrate, but then rises
sharply towards a near-steady state temperature distribution, as heat removal becomes dominated by
the gas flow over the deposit top surface. In the steady state, heat arrival from the spray equals heat
removed by convection into the flowing gas and by conduction into the deposit.

A thermal balance at the deposit top surface gives:

(17) { ) ( ) ( ) ( f f T T CW kG T T h
s S G
+ = + }

where h is the heat transfer coefficient at the top surface, T, T
g
and T
s
are the deposit, gas and spray
temperatures respectively, f and f
s
are the deposit and spray liquid fractions respectively, k is the
thermal conductivity and G is the temperature gradient in the deposit. As expected, the surface
temperature increases with increasing spray temperature and decreasing withdrawal rate (Bewlay
and Cantor 1990, 1991, Grant et al. 1991, 1993a).

3.2 Banding. A step of deposited material builds up in the wake of the spray as it scans across the
surface. A mass balance gives the average step height h:


VR
dt dm
h
2
= (18)
488
Microstructure development during rapid solidification

where R is the radius of the spray footprint and V is the scan rate. Typical deposition and scanning
rates are 5 kg/min and 2.5 m/s for gas atomised spraying and 0.05 kg/min and 0.1m/s for plasma
spraying, leading to step heights of 50-100 m in each case (Cantor et al. 1997).

Surface temperatures have been shown to be in the mushy solid plus liquid region, with typical solid
fractions of 0.75-0.9 (Mingard et al. 1998). As the spray scans over the surface, succeeding tracks
are deposited partly adjacent to and partly overlaying previous tracks. Between succeeding tracks,
the first track cools and solidifies before the second track deposits. A homogeneous microstructure
of fine scale equiaxed grains is formed when each layer of depositing material arrives before
adjacent and underlying tracks are fully solid, allowing liquid mixing to remove interlayer
boundaries. Alternatively, a banded microstructure of splatted pancake grains is formed when each
layer of deposited material arrives after adjacent and underlying tracks are fully quenched, so that
each layer solidifies independently (Cantor et al. 1997).

The transition between equiaxed and banded microstructures corresponds to a critical intertrack
time t
*
. For intertrack times shorter than t
*
, arriving layers have sufficient enthalpy to re-heat
adjacent or underlying tracks up to the solidus and allow liquid mixing. For intertrack times longer
than t
*
, arriving layers have insufficient enthalpy to re-heat adjacent or underlying tracks up to the
solidus and intermixing cannot take place. The critical intertrack time is given by:

{ } { }
dt
dm
Lf T T C t VR kG T T h
S G
+ = + ) ( ) (
*
(19)

where T
S
is the solidus temperature. Values of the critical intertrack time are typically ~1 s.

The critical interlayer time increases linearly with deposition rate and decreases inversely with scan
velocity and spray footprint radius. In the initial stages of spraying onto a cold substrate, the thermal
gradient is large and t
*
is small, leading to initial banded structures. As spraying continues, the
substrate and spray formed material become hotter, the thermal gradient decreases and t
*
rises,
with a transition to a homogeneous equiaxed microstructure in the later stages of spraying. Arc and
plasma spray processes are usually limited to the manufacture of thin surface coatings, and spraying
remains within the banded region. Larger scale spray forming of free standing objects continues for
longer, and equiaxed grain structures can be produced after an initial banded transition region with a
splatted grain structure (Cantor et al. 1997).

3.3 Porosity. Excessive liquid in the spray leads to the incorporation of included gas porosity in the
resulting spray formed alloy microstructure. On the other hand, insufficient liquid in the arriving
spray leads to shrinkage and intersplat porosity. The large scale rounded gas pores can be
distinguished clearly from the fine scale angular shrinkage and intersplat pores. The overall porosity
level decreases and then increases again as the temperature on the deposit top surface decreases
through the mushy solid plus liquid region, with minimum porosity at a temperature corresponding
to a surface solid fraction of ~0.9 (Underhill et al. 1995, Mingard et al. 1998). Gas and shrinkage
porosity can therefore be prevented, or at least minimised, by carefully controlling conditions to
balance the rates of heat extraction into the surrounding gas from the droplets during flight and from
the deposit top surface, in order to maintain the liquid fraction on the deposit top surface at ~0.9.

Porosity levels vary with time, similarly to the transition from an initial banded structure to a
homogeneous equiaxed grain structure. In the initial stages of spraying onto a cold substrate, rapid
thermal conduction into the substrate leads to a high rate of cooling, resulting in an initial chill zone
489
Cantor
with a high degree of shrinkage and intersplat porosity. As spraying continues, the surface
temperature increases and the underlying thermal gradient decreases, rapidly approaching a
near-steady state. The porosity then depends on the steady state surface temperature, with residual
shrinkage porosity at low temperatures, gas porosity at high temperatures, and optimum low
porosity at intermediate temperatures, corresponding to a solid fraction of ~0.9 (Underhill et al.
1995, Mingard et al. 1998).

3.4 Grain size. Grain size also shows a strong correlation with surface temperature, increasing as the
liquid fraction on the top surface and therefore the solidification time both increase. Typical
equiaxed grain sizes are ~20-30m in Al and Ni alloys (Grant et al. 1993b, Underhill et al. 1995,
Mingard et al. 1998).

Grain sizes are small because of the seeding effect of a large number of small fully solidified
droplets arriving at the deposit surface. The grains coarsen according to cubic Ostwald ripening
kinetics:

(20) Kt d d =
3
0
3

where d
o
and d are the initial and final grain size respectively, K is the coarsening rate constant and t
is the solidifcation time. The coarsening rate constant increases and then decreases again as the
temperature on the deposit top surface decreases through the mushy solid plus liquid region, with
the minimum coarsening rate at a temperature corresponding to a surface solid fraction of ~0.7
(Manson-Whitton et al. 1998, 2001).

The coarsening rate constant is given by a modified Lifschitz-Slyovitz-Wagner (LSW) theory
(Lifshitz and Slyovitz1961, Wagner 1961, Annavarapu and Doherty 1995, Manson-Whitton et al.
2001):


y c c RT
V Dc
K
eq
m eq
) (
3

=

(21)

where D is the diffusion coefficient, c
eq
and c are the equilibrium and actual solid compositions
respectively, is the solid-liquid surface energy, V
m
is the molar volume and R is the gas constant.

The parameter y depends upon the coarsening geometry at different temperatures in the mushy zone.
At high temperatures near the liquidus, y = 1 and coarsening takes place by solute migration
between widely separated solid particles at low volume fraction. At intermediate temperatures, y<1
and coarsening takes place by solute migration through a liquid film at the solid grain boundaries; y
decreases and K increases as the liquid film becomes thinner at lower temperatures. At low
temperatures near the solidus, y increases again and K decreases again, as the solid grains impinge
and extinguish liquid diffusion (Manson-Whitton et al. 2001).

The overall effect of seeding followed by complex coarsening kinetics during cooling through the
mushy zone on the deposit top surface is to produce a small grain size, which falls gradually with
decreasing surface temperature (Mingard 1998, Underhill et al. 1995). Second phase particles are
very effective in pinning grain boundaries, and small equiaxed grain sizes are surprisingly resilient
to spray forming conditions.

490
Microstructure development during rapid solidification
3.5 Segregation. Microsegregation is very limited in spray deposited materials, because of the
seeded small grain size, high solid fraction and rapid solidification on the deposit surface.

However, inverse macrosegregation through the solid/liquid material on the top surface sucks solute
to the outer edge of thick strip and cylindrical billets (Mingard et al. 2000). The macrosegregation
can be described by a modified Scheil equation (Flemings and Nereo 1967) for solidification on the
deposit surface (Mingard et al. 2000):


=
1
1
0
k
f kc c (22)

where solid of composition c solidifies from a material of composition c
o
, k is the partition
coefficient, f is the liquid fraction and is the shrinkage.

Integration of equation (22) through the different stages of solidification during droplet flight,
deposition and then coarsening on the deposit surface gives excellent agreement with experiments
for spray formed 2xxx Al alloys (Mingard et al. 2000). Equation (22) shows that the severity of the
macrosegregation increases with increasing surface temperature and liquid fraction. In complex
2xxx series Al alloys and Ni superalloys, all solutes with k<1 segregate to the deposit edge, all
solutes with k>1 segregate towards the deposit centre, and the extent of segregation increases as k
increasingly deviates from unity (Mingard et al. 2000).


4. SUMMARY

This paper has provided an overview of the physical mechanisms which control the development of
microstructure during rapid solidification, and the relationship between the processing conditions
and the resulting rapidly solidified microstructures. A range of rapid solidification processes have
been considered, concentrating on melt spinning and spray forming. A range of different
microstructural features have been considered, including porosity, grain size and segregation.

Considerable progress has been made in recent years towards understanding the relationships
between rapid solidification processing conditions, the heat and mass flows taking place during
processing, and the resulting material microstructure, by using a combination of detailed process
monitoring, theoretical and computer modelling, and material characterisation. This has allowed the
effective manufacture of a variety of rapidly solidified materials and components.


REFERENCES

Annavarapu S and Doherty R D (1995) Inhibited coarsening of solid-liquid microstructures in spray
casting at high-volume fractions of solid Acta Metallurgica et Materialia 43, 3207.
Bewlay B P and Cantor B (1986) Cooling rates in melt spun 316L stainless steel International
Journal of Rapid Solidification 2, 107.
Bewlay B P and Cantor B (1990) Modelling of spray deposition: measurements of particle size, gas
velocity, particle velocity and spray temperature in gas atomized sprays Metallurgical
Transactions 21B, 899.
Bewlay B P and Cantor B (1991) The relationship between thermal history and microstructure in
spray deposited Sn-Pb alloys Journal of Materials Research 6, 1433.
491
Cantor
Cantor B, Kim W T, Bewlay B P and Gillen A G (1991) Microstructure-cooling rate correlations in
melt spun alloys Journal of Materials Science 26, 1266.
Cantor B (1997) Modern design meets materials science Materials World 5 386
Cantor B, Baik K-H and Grant P S (1997) Development of microstructure in spray formed alloys
Progress in Materials Science 42, 373.
Cantor B (1998) Material microstructures manufactured by advanced solidification processing in
Processing and Properties of Advanced Engineering Materials ed. T Kobayashi, M Umemoto
and M Morinaga (Japan Society for Promotion of Science, Toyohashi), p. 85
Cantor B (1999) Nanocrystalline materials manufactured by advanced solidification processing
methods Materials Science Forum 307 ,143.
Cantor B, Allen C M, Dunin-Burkowski R, Green M H, Hutchinson J L, OReilly K A Q, Petford-
Long A K, Schumacher P, Sloan J and Warren P J (2001) Applications of nanocomposites
Scripta Materialia, in press.
Doherty R D, Martin J W and Cantor B (1997) Stability of Microstructure in Metals and Alloys
(CUP, Cambridge)
Flemings M C and Nereo G E (1967) Transactions of the Metallurgical Society of AIME 239 1449
Gillen A G and Cantor B (1985) Photocalorimetric cooling rate measurements on a Ni-5wt%Al
alloy rapidly solidified by melt spinning Acta Metallurgica 33, 1813.
Grant P S, Kim W T and Cantor B (1991) Spray forming of Al-Cu alloys Materials Science and
Engineering A134, 111.
Grant P S, Cantor B and Katgerman L (1993a) Modelling of droplet dynamic and thermal histories
during spray forming Acta Metallurgica and Materialia 41, 3097.
Grant P S, Underhill R P, Kim W T, Mingard K P and Cantor B (1993b) Grain growth in spray
formed Al alloys in Second International Conference on Spray Forming ed. J V Wood
(Woodhead, Cambridge) p. 45.
Hayzelden C, Rayment J J and Cantor B (1982) Rapid solidification microstructures in austenitic
Fe-Ni alloys Acta Metallurgica 31, 379.
Kim W T and Cantor B (1990) The variation of grain size with cooling rate during melt spinning
Scripta Metallurgica and Materialia 24, 633.
Kim W T, Cantor B and Kim T H (1990) Microstructure of rapidly solidified eutectic and
hypereutectic Al-Cu alloys International Journal of Rapid Solidification 5, 251.
Kurz W and Fisher D J (1989) Fundamentals of Solidification (Trans Tech, Switzerland)
Lee S M, OReilly K A Q, Hong C P and Cantor B (1998) Effect of process parameters on planar
flow cast Al-Cu ribbons International Journal of Cast Metals Research 10, 181.
Lifshitz I M and Slyozov V V (1961) Journal of the Physics and Chemistry of Solids 19, 35.
Manson-Whitton E D, Stone I C, Grant P S and Cantor B (1998) Solid/liquid coarsening behaviour
of spray formed IN718 in Solidification 1998 ed. S P Marsh, J A Dantzig, R Trivedi, W
Hofmeister, M G Chu, E G Lavernia and J H Chun (TMS, Warrendale) p. 415.
Manson-Whitton E D, Stone I C, Jones J R, Grant P S and Cantor B (2001) Isothermal grain
coarsening of spray formed alloys in the semi-solid state in preparation
Mingard K P, Alexander P W, Langridge S J, Tomlinson G A and Cantor B (1998) Direct
measurement of sprayform temperatures and the effect of liquid fraction on microstructure Acta
Materialia 46, 3511.
Mingard K P, Cantor B, Palmer I G, Hughes I R, Alexander P W, Willis T C and White J (2000)
Macro-segregation in aluminium alloy sprayformed billets Acta Materialia 48, 2435 .
Trivedi R (1994) Materials Science and Engineering A178, 129.
Underhill R P, Grant P S, Bryant D J and Cantor B (1995) Grain growth in spray formed Ni
superalloys Journal of Materials Synthesis and Processing 3, 3.
Vogt E and Frommeyer G (1985) in Rapidly Solidified Materials ed. P W Lee, R S Carbonara and B
C Giessen (Elsevier North Holland, New York), p. 317.
492
Microstructure development during rapid solidification
Wagner C (1961) Z Electrochemie 65, 581.
Zhao Y Y Grant P S and Cantor B (2000) Modelling and experimental analysis of vacuum plasma
spraying Modelling and Simultion in Materials Science and Engineering 8, 497.
493

You might also like