Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

J Mater Sci (2012) 47:3706–3712

DOI 10.1007/s10853-011-6219-8

Density and thermal expansion of liquid Al–Si alloys


Julianna Schmitz • Bengt Hallstedt •
Jürgen Brillo • Ivan Egry • Michael Schick

Received: 9 August 2011 / Accepted: 19 December 2011 / Published online: 14 January 2012
Ó Springer Science+Business Media, LLC 2012

Abstract The density of Al-rich liquid Al–Si alloys was processability, and mechanical strength, which can be fur-
measured contactlessly on electromagnetically levitated ther improved by heat treatment or additive elements [1].
samples using optical dilatometry. Data were obtained for In practice, mainly Al-based alloys are employed, which
samples covering compositions up to 50 at.% Si and in a contain small amounts of the other alloying components and
temperature range between 650 and 1500 °C. The densities additives. However, there are also some investigations
can be described as linear functions of temperature with dealing with Si richer alloys. These studies focus on
negative slopes. Moreover, they increase monotonically solidification [2–5] or thermodynamics [6–12] of the binary
with an increase of Si concentration. In a temperature range or the ternary alloy [13, 14]. Those liquids’ thermophysical
between 1100 and 1400 °C, it can be deduced from the properties have hardly been studied [15, 16].
composition dependence of the density that virtually no To anticipate and optimize casting processes, a funda-
excess volume arises during alloying of the pure elements. mental understanding of solidification is essential. Models
For lower temperatures an excess volume is discussed, with a quantitative prediction capability are needed, based
considering the temperature dependence of Si density lit- on the thermodynamics of these processes. Using a
erature data. The density data were integrated in a ther- CALPHAD approach [17] the Gibbs free energy can pro-
modynamic model description of the Al–Si system. In this vide this basis. If the Gibbs energies were accurately
way volume changes during solidification and changes in known, all other thermophysical properties could be cal-
phase equilibria as function of pressure can be calculated. culated from basic thermodynamic relations. Vice versa
Gibbs energies can be determined from sound thermody-
namic data. However, for this approach to work, it is not
Introduction sufficient to explore the Al-rich part of the phase diagram
only; a strategy for covering the entire phase diagram must
Motivation be developed.
Of particular interest for solidification is the knowledge
Ternary Al–Cu–Si alloys are technologically widely used as of the density of the alloy melt. When these density data
lightweight casting alloys due to their good castability, are available, volume changes during solidification and
cooling can be calculated. This is a prerequisite for realistic
casting simulations and simulation of macro-segregation.
For the fairly small number of binary systems investi-
gated experimentally, the excess volume has been found to
J. Schmitz (&)  J. Brillo  I. Egry
Institut für Materialphysik im Weltraum, Deutsches Zentrum be in the order of a few percent of the total volume and no
für Luft- und Raumfahrt (DLR), 51170 Cologne, Germany correlation has been found so far with e.g., excess enthalpy
e-mail: julianna.schmitz@dlr.de (i.e., enthalpy of mixing). The excess volume is, thus, in
the same order of magnitude as the volume change
B. Hallstedt  M. Schick
Materials Chemistry, RWTH Aachen University, during solidification. Therefore, measured volumes, and in
52074 Aachen, Germany particular excess volumes, are important.

123
J Mater Sci (2012) 47:3706–3712 3707

Since the expected excess volume is only a few percent, P2


M xi  Mi
it is necessary to access as large part of the composition qðxi Þ ¼ ¼ P2 i¼1 M ð4Þ
V i¼1 xi  q þ V
i E
range of the considered system as possible to gain highly i

accurate measurements. In most cases, VE does not vanish. When atoms are of similar
For ternary systems these data are generally sparse and size and have no preferred coordination with either similar
even data of binary alloys are often missing. Hence, our or different atoms, they are stochastically distributed in the
objective is to systematically develop a reliable thermo- solution. In this situation, the temperature and composition
dynamic dataset, first for the binary Al-alloys before dependence of VE can be split by an approximation, which is
addressing the ternaries. symmetric in the alloy composition [20, 21]:
In the liquid Al–Cu system, density and surface tension
data have already been reported [18, 19], viscosities are in V E ðx1 ; x2 ; T Þ ¼ x1 x2 V0 ðTÞ ð5Þ
preparation. With this article, we continue working on the Here, the composition dependence is incorporated in the
densities of Al–Si, which were measured contactlessly in product of the atomic concentrations xi (i = 1, 2) of the
electromagnetic levitation (EML), a method which enables alloy components. Hence, the whole system can be
us to handle high melting and reactive materials. described by a coefficient V0, which is a function of tem-
perature, but independent of composition.
Density and molar volume of binary mixtures
Thermodynamic modeling
Within a temperature interval, (T-TL), which is small
enough, the density, q(T), of a liquid metal can be con- Volume data are usually not included in thermodynamic
sidered as a linear function of temperature, T: databases so far. In the form of Eqs. 3 and 5, they can
qðTÞ ¼ qL þ qT ðT  TL Þ ð1Þ easily be added to existing thermodynamic model
descriptions. To each Gibbs energy parameter in the ther-
Here, TL is the liquidus temperature, qL is the density at TL
modynamic description a term of the form (P–P0)V is
and qT its temperature coefficient qq/qT.
added;
From these quantities, the volume expansion coefficient
b ¼ V1 oV GðP; TÞ ¼ GðTÞ þ ðP  P0 ÞV; ð6Þ
oT can be calculated:
q where G is the Gibbs energy, P is the pressure, P0 the
b¼ T ð2Þ
qL normal pressure (1 bar) and V the volume. The thermo-
dynamic description of the Al–Si system from Feufel et al.
In multicomponent mixtures, the different elements usually
[22] was used. Volume data for solid and liquid Al and Si
establish additional interactions among each other. This is
from [23] and excess volumes from the present work were
the reason, why in alloying not only the sum of the
added to this description. The data for liquid Al were
weighted molar volumes of the pure components, i, has to
adjusted to match the evaluation of Assael et al. [24]. The
be considered, but also an excess volume, which accounts
modified volume used for liquid Al is V = 9.9044 ?
for these interactions.
1.548 9 10-3T cm3/mol, where T is in K.
In a binary liquid system (i = 1, 2) the molar volume, V,
can be written [20].
X
2
Mi Experimental
V¼ xi  þ VE ð3Þ
i¼1
qi
Electromagnetic levitation (EML)
The molar volumes of the pure components are here
written by considering the molar masses, Mi, of each Experiments were performed in a stainless steel high vac-
component and their densities, qi, at a certain temperature. uum chamber described in detail elsewhere [25, 26]. It was
These are weighted by the atomic concentrations xi. evacuated to 10-7 mbar and then filled with 750 mbar Ar
Deviations of V from ideal mixing behavior are included by (purity: 99.9999 vol.%). The pre-alloyed sample was
the excess volume, VE. That means, for VE = 0 Eq. 3 is a positioned in the center of a levitation coil to which an
simple linear combination of pure component molar vol- alternating current of typically 300 kHz and about 200 A is
umes and describes an ideal mixing without volume applied. Its inhomogeneous electromagnetic field induces
change. eddy currents inside the electrically conductive sample.
From Eq. 3 the composition dependence of the binary The forces arising from interaction of these currents with
system’s density is deduced: the field levitate the sample stably against gravity.

123
3708 J Mater Sci (2012) 47:3706–3712

A detailed consideration of the EML-principle is given in volume, when it is integrated over space assuming
literature [27]. rotational z-axis symmetry:
Beneficial effect of the inherent ohmic losses of these Zp
currents is the heating and melting of the sample, which is 2
VP ¼ p \RðuÞ [ 3 sinðuÞdu ð7Þ
realized simultaneously with EML. As levitation and 3
0
heating are coupled, a certain desired temperature can only
be adjusted within narrow limits by changing the levitation Here, R(u) is the fitted average 2D-edge curve and u is the
power. During our measurements, the forced cooling in a polar angle. The sample volume, Vp, is only obtained in
laminar flow of Ar or even He gas could be avoided, since pixel units. It must be calibrated to the real volume, V, by a
reducing the levitation power was sufficient to cool the comparison to reference volumes, which is also described
samples to the desired temperatures. Only for solidification in [25]. From the mass, m, of the processed sample the
additional cooling by He gas flow was needed. density is calculated, provided evaporation during the
The measurement procedure is to start at a certain lev- measurement was negligible.
itation power at high temperatures and reduce the power
stepwise to reach lower temperatures. Measurements are Sample preparation
then performed in thermal equilibrium.
The sample temperature is measured by an infrared Samples with a diameter of about 6 mm were made of Si
pyrometer. As the emissivity is generally not known, the and Al metal with a purity of 99.999% (metals basis). Mass
pyrometer output signal, TP, has to be recalibrated using adjustment and surface cleaning was achieved by pinching
the known liquidus temperature, TL, of each sample as a off pieces and ultrasonic treatment. Despite the contact
calibration point. For most liquid metals, the emissivity with crucible material, the samples needed to be pre-
within a limited wavelength band is almost independent of alloyed within an arc melting furnace, since otherwise the
the temperature [28]. Hence, for calibration a constant solid semiconducting Si could not have been processed in
emissivity within the operating wavelength range of the EML. Alloying was completed in the EML apparatus by
pyrometer can be assumed. The exact calibration procedure heating the samples to about 1500 °C. At the same step,
is described, for instance, in [29]. oxygen contaminations were removed by keeping the
sample at this temperature until no bright oxide spots could
Optical dilatometry be observed on the surface. After this alloying procedure
and the subsequent experiment, the mass loss of the about
Densities of the liquid droplets are determined by taking 300 mg samples was less than 3 mg. If this evaporation
side view shadowgraph images and calculating its volume affected only one component (Al), the change of the alloy
at different temperatures. The technique is thoroughly composition would amount 0.2 at.% at most. For compar-
described in [25]. ison and to check the reproducibility of the data, also
In principle, the levitated sample is illuminated from the samples prepared in vacuum induction melting and in
back side by an expanded HeNe-laser beam. An interfer- electron beam melting were measured. The inductively
ence filter removes thermal radiation of the sample and a molten samples were obtained from our cooperation part-
pinhole eliminates scattered nonparallel light rays from the ners in Jena1, the electron beam samples from Clausthal.2
beam. The images are recorded by a CCD-camera and
analyzed by an edge detection algorithm to determine the
edge curve of the 2D-projection of the sample. In order to Results
evaluate the liquid sample volume, 3D-information must be
gained from these images. Measurements were performed at temperatures, where
The surface of a levitated droplet performs oscillations evaporation was negligible. Levitation became more and
around its rest position, which is axially symmetric to the more unstable, the colder the samples got and the more Si
symmetry-axis of the electromagnetic field providing the they contained. Hence, a limited temperature interval was
levitating force. Since not every oscillation mode is sym- accessible, which, however, is quite broad for most sam-
metric to this axis, a single frame is not necessarily axially ples (800–1400 °C).
symmetric, when it does not incidentally capture the rest
position of the sample. In averaging over 1000 of these 1
M. Rettenmayr, A. Löffler, Institut für Materialwissenschaft und
frames, these oscillations (usually in the order of 50 Hz)
Werkstofftechnologie, Friedrich-Schiller-Universität Jena, 07743
are eliminated and one obtains an edge curve, which is Jena, Germany.
symmetric to the vertical sample axis [25]. This curve is 2
R. Schmid-Fetzer, J. Gröbner, Institut für Metallurgie, TU Claus-
then fitted by Legendre polynomials and gives the sample thal, 38678 Clausthal-Zellerfeld, Germany.

123
J Mater Sci (2012) 47:3706–3712 3709

Table 1 Fit parameters for the


Sample TL, °C qL, g cm-3 qT, 10-4 g cm-3K-1 b, 10-5 K-1
density and thermal expansion
of each of the investigated Al 660 2.36 ± 0.01 -3.0 ± 0.1 12.8 ± 0.1
Al–Si samples. The samples,
denoted by FSU or TUC were Al88Si12 576 2.46 ± 0.01 -2.1 ± 0.1 13.1 ± 0.3
obtained from Jena (see Al80Si20 697 2.44 ± 0.01 -2.3 ± 0.1 8.9 ± 0.1
footnote 1) or Clausthal (see Al70Si30 831 2.37 ± 0.02 -1.6 ± 0.2 6.8 ± 1.0
footnote 2), respectively
Al70Si30 (FSU) 831 2.40 ± 0.02 -1.7 ± 0.2 7.0 ± 0.9
Al70Si30 (TUC) 831 2.39 ± 0.02 -1.8 ± 0.2 7.0 ± 0.7
Al60Si40 (TUC) 952 2.42 ± 0.01 -2.9 ± 0.1 11.9 ± 0.3
Al50Si50 (TUC) 1061 2.46 ± 0.03 -3.2 ± 0.2 12.9 ± 1.0

The measured densities are displayed in Fig. 1. They Measurements were started at high temperatures and
can be fitted linearly by Eq. 1. The obtained densities qL at data were acquired after subsequent cooling steps. This
liquidus temperature, TL, and the slopes of the curves means that at high temperatures evaporation leads to a
(temperature coefficients), qT, are compiled in Table 1 for change of sample mass between different data points. Later
each measured sample. There, also the volume expansion on, at lower temperatures, no evaporation occurs, and the
coefficient, b, is calculated from Eq. 2. sample mass stays constant. Since these data are evaluated
Literature data of the pure components are denoted as using the mass of the processed sample, the lower tem-
dashed lines in Fig. 1. The literature data for Al [23] were perature densities are determined correctly, while the val-
adjusted to match those of [24] in temperature interval ues at higher temperatures are underestimated. For this
660–920 °C. The ones of pure Si [23, 30] remained reason, recalibrated densities of Al up to 1450 °C are
unmodified. The data in [23] (dashed line in Fig. 1) are reliable and only these data are evaluated in the linear
fitted using a linear temperature dependence of the volume. fitting.
The measured Al-densities fit the reported data quite In all the data of the other samples, where the overall mass
well, except for the values measured above 1450 °C. loss was less than 1%, this effect does not occur. In these
There, a steeper slope arises, leading to lower density cases, the major evaporation was during the alloying and
values. This effect must be due to evaporation of sample surface cleaning step at more elevated temperatures. Hence,
material during measurement at these elevated tempera- for these samples the recalibration by the processed sample
tures. In this temperature region, the mass loss of 3% mass is valid throughout the whole measurement range.
disturbs recalibrating sample volumes to densities with the At first glance the value of the Al80Si20 sample seems to be
processed sample mass. Below 1450 °C no evaporation quite high. But the composition dependence depicted in
occurs, so the mass during the lower T measurements stays Fig. 2 visualizes that this scatter is within the small uncer-
constant. tainties of the measurement principle which are of 1% [25].

Fig. 1 Density of liquid Al–Si versus temperature. The solid lines


represent linear fits to the data. The samples denoted by FSU or TUC Fig. 2 Composition dependence of the liquid Al–Si density at
were obtained from Jena (see footnote 1) or Clausthal (see footnote 1200 °C. The pure component data were taken from literature [23,
2), respectively. For comparison, literature data [23] (partly combined 24, 30]. The lines represent fits of Eq. 5 to the data, giving an excess
with [24], see text) of the pure components are also depicted volume, VE, of zero

123
3710 J Mater Sci (2012) 47:3706–3712

The same holds for the Al70Si30 alloys, where we


compared samples created by different methods. Within the
measurement uncertainties, density values are the same.
The completely homogeneous samples, obtained from our
project partners, levitated more stably from the beginning.
The pre-alloyed samples only became stable after they
were heated above the melting temperature of pure Si. For
this reason, only fully alloyed initial samples were used for
Si-contents above 30 at.% .

Discussion

In Figs. 2, 3 and 4, the composition dependence of q is


depicted at different temperatures, together with literature
Fig. 4 Composition dependence of the liquid Al–Si density at
data [23, 24, 30] of the pure components. The density 800 °C with the extrapolated pure component data from literature
increases monotonically with increasing Si-content. The [23, 24, 30]. Considering the different existing datasets of Si, both fits
considered density range between pure Al and Si is quite to the data are equally good. In either case the excess volume, VE,
does only differ little from zero
narrow and scatter of the data is within the measurement
uncertainties of only 1%.
In Figs. 2, 3 and 4, the composition dependence of the
For liquid Si, there are two classes of datasets [23, 30].
alloy density is fitted by Eq. 4. The black lines are reduced
Starting from quite similar density values above TL, one
v2-fits to the data using one Si dataset [23] the grey curves
dataset [23] observes a linear behavior during undercooling
employ the other [30]. The full lines represent fits, where
to about 1100 °C. The other [30] reports a T2-behavior
qAl, qSi, and VE were varied. These fit parameters are
within the same temperature region and extrapolates a
compiled in Tables 2 and 3. As a guide to the eye, VE was
maximum around 900 °C.
set to zero in the dashed line fits.
Since our measurements were performed in this under-
The main difference in the fits originates from the single
cooled temperature range of pure Si and it is not trivial to
Si-datapoints. The fits start to deviate from each other at Si
select between those reported data, we checked the
extrapolations of both datasets with regard to composition- Table 2 Fit parameters for the excess volume and densities of Al and
dependent volume effects during alloying. Si at different temperatures, utilizing pure metal data from Ref. [23]
T, °C V0, cm3 mol-1 qAl, g cm-3 qSi, g cm-3

800 0.5 ± 0.6 2.31 ± 0.03 2.79 ± 0.03


900 0.6 ± 0.6 2.31 ± 0.03 2.74 ± 0.03
1000 0.2 ± 0.5 2.26 ± 0.02 2.70 ± 0.03
1100 0.0 ± 0.5 2.24 ± 0.02 2.66 ± 0.02
1200 -0.1 ± 0.3 2.21 ± 0.02 2.62 ± 0.01
1300 -0.2 ± 0.4 2.19 ± 0.02 2.58 ± 0.02
1414 -0.3 ± 0.4 2.17 ± 0.02 2.54 ± 0.02

Table 3 Fit parameters for the excess volume and densities of Al and
Si at different temperatures, utilizing pure Si data from Ref. [30]
T, °C V0, cm3 mol-1 qAl, g cm-3 qSi, g cm-3

800 -0.6 ± 0.4 2.30 ± 0.03 2.63 ± 0.04


900 -0.1 ± 0.6 2.30 ± 0.3 2.63 ± 0.04
1000 -0.3 ± 0.5 2.26 ± 0.02 2.63 ± 0.03
Fig. 3 Composition dependence of the liquid Al–Si density at
1100 -0.2 ± 0.4 2.24 ± 0.02 2.62 ± 0.02
1000 °C with the extrapolated pure component data from literature
[23, 24, 30]. There, in the undercooled region of Si different behavior 1200 -0.2 ± 0.4 2.21 ± 0.02 2.62 ± 0.02
is discussed and considered in the evaluation of our data. The lines 1300 0.0 ± 0.4 2.20 ± 0.02 2.60 ± 0.02
represent fits of Eq. 5 to each of the datasets. With the second, a slight 1414 0.01 ± 0.4 2.17 ± 0.02 2.58 ± 0.2
deviation of VE of from zero might occur

123
J Mater Sci (2012) 47:3706–3712 3711

concentrations of about 50 at.%. The Al50Si50 data seem to 2.70


be slightly higher than the fits, implying higher values for Al-6Si
2.65
Si-rich alloys. Since we focused on the Al-rich side of the
alloy, no final statement can be given about the Si-rich 2.60
densities. Solid
Comparing fits with adjustable VE and those with 2.55
E
V = 0, one might observe slight differences between

[g/cm3 ]
2.50
them, intensifying with decreasing T. But even at 800 °C
the deviation of the measured molar volume from the ideal 2.45
one is only in the order of 2%, which is more due to the
scatter of the data. In Fig. 5, the temperature dependence of 2.40
Liquid
the fitted VE is evaluated, considering V0 from Eq. 5.
2.35
There, for both of the evaluated Si-densities, V0 virtually
vanishes considering its uncertainty. 2.30
The observed simple additivity of molar volumes in the
Al–Si system implies comparable nearest–neighbor inter- 2.25
0 200 400 600 800 1000
actions regardless of the kind of atoms involved. In literature
[15], at temperatures 20 K above TL and Si contents higher T [°C]
than 40 at.%, a virtually instantaneous change in the average Fig. 6 Calculated density as function of temperature for an Al–6
coordination number of the Al–Si alloy atoms was observed at.% Si alloy
by neutron diffraction. This was explained by formation of
tetrahedrally coordinated Si atoms only in the Si-rich liquid. Using the thermodynamic description for Al–Si includ-
For Si concentrations below 40 at.% the average coordina- ing volume data (with zero excess volume) with the
tion of the atoms stayed almost constant and decreased modified Gibbs energy expressions according to Eq. 6, the
to another constant value upon increasing the Si content by density as function of temperature can be calculated with
10 at.% to 50 at.%. The mean distance of atoms within the standard thermodynamic software such as Thermo-Calc
pure liquids and the alloys was comparable. [31] for any composition, including any phase transfor-
Considering this, for low Si contents, an ideal mixing mations, such as solidification. An example for an Al–6
seems plausible. In case the Si tetrahedral short-range order at.% Si alloy is shown in Fig. 6. Including volume data in
should establish in Si-rich Al–Si alloys, the atomic inter- the thermodynamic description means, that pressure
actions within the system cannot be equal. Against this dependence of the Gibbs energy is incorporated. This can
background the observed ideal mixing is not trivial and be used to calculate phase equilibria as function of pres-
should be examined in more detail by extending density sure. In Fig. 7, the Al–Si phase diagram is calculated at
measurements to alloys with higher Si-contents. different pressures. Such calculations can be considered
reasonably accurate at moderate pressure. High pressure
differences in compressibility, which is currently not
modeled, will lead to deviations from reality of such
‘‘simple’’ calculations. For pressures above about 1 GPa
(10 kbar), depending on material, it is probably necessary
to consider the compressibilities to reach acceptable
accuracy.

Summary

The density and thermal expansion of liquid Al–Si alloys


was determined experimentally as a function of tempera-
ture and composition up to Si-concentrations of 50 at.%.
The density can be described by a linear combination of the
pure metals’ molar volumes and hence, no pronounced
Fig. 5 Temperature dependence of VE, visualized by the coefficient
excess volume was found. On the other hand, there are
V0 fitted with different Si densities [23, 30]. The dashed line is placed experiments [15], suggesting strong interactions in Al–Si
at zero. Within the uncertainties, V0 and VE equal zero melts with Si-contents above 50 at.%. Addressing densities

123
3712 J Mater Sci (2012) 47:3706–3712

750 4. Nikanorov SP, Volkov MP, Gurina VN, Burenkova YuA, Der-
kachenko LI, Kardashev BK, Regel LL, Wilcox WR (2005)
Liquid Mater Sci Eng A 390:63
5. Griffiths WD, Xiao L, McCartney DG (1996) Mater Sci Eng
700 10 kbar (1 GPa) A205:31
6. Kanibolotsky DS, Bieloborodova OA, Kotova NV, Lisnyak VV
(2002) J Therm Anal Cal 70:975
1 bar 7. Bros JP, Eslami H, Gaune P (1981) Ber Bunsenges 85:333
650 8. Körber F, Oelsen W (1937) Mitt Kaiser-Wilhelm Inst Eisenforsch
T [°C]

1 kbar 19:131
9. N.V. Gizenko, B.I. Emlin, S.N. Kilesso, M.I. Gasik and
A.L. Zavyalov (1983) Izv. Akad Nauk SSSR Met. 1:33–35
600 Engl. Transl
10. Berthon O, Petot-Ervas G, Petot C, Desré P (1969) C R Acad Sci
Paris 268C:1939
11. Schaefer SC, Gokcen NA (1979) High Temp Sci 11:31
550
fcc-Al + Si 12. Bonnet M, Rogez J, Castanet R (1989) Thermochinmica Acta
155:9
13. Kanibolotsky DS, Bieloborodova OA, Kotova NV, Lisnyak VV
500 (2004) Thermochimica Acta 412:39
0 2 4 6 8 10 12 14 16 18 20 14. Witusievicz VT, Arpshofen I, Seifert H-J, Aldinger F (2000) J
Alloys Compounds 297:176
Al at.% Si 15. Gabathuler JP, Steeb S, Lamparter P (1979) Z Naturforsch
34a:1305
Fig. 7 Calculated Al–Si phase diagram in the Al-rich part at normal 16. Kéita NM, Steinemann S (1978) J Phys C: Solid State Phys
pressure (1 bar, solid lines), 1 kbar (dotted lines) and 10 kbar (1 GPa, 11:4635
dashed lines). Note that the Al liquidus increases with pressure since 17. Lukas HL, Fries SG, Sundman B (2007) Computational ther-
the volume of Al increases on melting, whereas the Si liquidus modynamics: the calphad method. Cambridge University Press,
decreases with pressure since the volume of Si decreases on melting. Cambridge
This leads to a considerable shift of the eutectic composition 18. Brillo J, Egry I, Westphal J (2008) Int J Mater Res 99(2):162
19. Schmitz J, Brillo J, Egry I, Schmidt-Fetzer R (2009) Int J Mater
of those alloys experimentally could be worthwhile to gain Res 100(11):1529
more insight into this system. 20. Lüdecke C, Lüdecke D (2000) Thermodynamik. Springer,
By including the experimental density data to the ther- Heidelberg
21. Porter AW (1920) Trans Faraday Soc 16:336
modynamic description, the modeling is capable of includ- 22. Feufel H, Gödecke T, Lukas HL, Sommer F (1997) J Alloys
ing moderate pressure dependencies. Compd 247:31
23. Hallstedt B (2007) Calphad 31:292
Acknowledgement Within the framework of PAK 461 this study 24. Assael MJ, Kakosimos K, Banish RM, Brillo J, Egry I, Brooks R,
was financially supported by the ‘‘Deutsche Forschungsgemeins- Quested PN, Mills KC, Nagashima A, Sato Y, Wakeham WA
chaft’’ under grant numbers EG 93/8-1 and HA 5382/3-1. This is (2006) J Phys Chem Ref Data 35:285
gratefully acknowledged. Further, we would like to thank our coop- 25. Brillo J, Egry I (2003) Int J Thermophys 24:1155
eration partners Rainer Schmid-Fetzer, Joachim Gröbner, Markus 26. Brillo J, Egry I, Giffard HS, Patti A (2004) Int J Thermophys
Rettenmayr, and Andrea Löffler for sharing their expertise and for the 25:1881
preparation of high quality samples. 27. Brillo J, Lohöfer G, Schmidt-Hohagen F, Schneider S (2006) Int J
Mat Prod Tech 26:247
28. Krishnan S, Hansen GP, Hauge RH, Margrave JL (1990) High
Temp Sci 29:17
References 29. Brillo J, Egry I, Ho I (2006) Int J Thermophys 27:494
30. Watanabe M, Adachi M, Morishita T, Higuchi K, Kobatake H,
1. Shivkumar S, Wang L, Keller C (1994) Z Metallkd 85(6):394 Fukuyama H (2007) Faraday Discuss 136:279
2. Ge LL, Liu RP, Li G, Ma MZ, Wang WK (2004) Mater Sci Eng 31. Andersson JO, Helander T, Höglund L, Shi P, Sundman B (2002)
A 385:128 Calphad 28:273
3. Pierantoni M, Gremaud M, Magnin P, Stoll D, Kurz W (1992)
Acta Metall Mater 40(7):1637

123

You might also like