You are on page 1of 15

Intermetallics 18 (2010) 92–106

Contents lists available at ScienceDirect

Intermetallics
journal homepage: www.elsevier.com/locate/intermet

Experimental study and thermodynamic re-assessment


of the binary Al–Ta system
V.T. Witusiewicz a, *, A.A. Bondar b, U. Hecht a, J. Zollinger a, V.M. Petyukh b,
O.S. Fomichov b, V.M. Voblikov b, S. Rex a
a
ACCESS e.V., Intzestr. 5, D-52072 Aachen, Germany
b
Frantsevich Institute for Problems of Materials Science, Krzhyzhanovsky Str. 3, 03680 Kyiv, Ukraine

a r t i c l e i n f o a b s t r a c t

Article history: Phase relations in the binary Al–Ta system were re-assessed taking into account recent experimental
Received 21 April 2009 data. Alloys of different composition were prepared and analyzed by means of DTA, DSC, pyrometry after
Received in revised form Pirani–Alterthum, SEM/EDS and SEM/EBSD techniques in order to prove the phase relations in the
26 June 2009
system. Both, as-cast and annealed samples were investigated. A new thermodynamic description of the
Accepted 29 June 2009
Available online 3 August 2009
binary Al–Ta system was obtained by CALPHAD modeling of the Gibbs energy of all accepted individual
phases, taking into account the above experimental data as well as assessed published data. Applying the
results, calculations of the phase diagram and thermodynamic properties were performed with the
Keywords:
A. Aluminides, miscellaneous Thermo-Calc software. They are shown to reproduce properly experimental data.
B. Phase diagrams Ó 2009 Elsevier Ltd. All rights reserved.
B. Thermodynamic and thermochemical
properties
E. Phase diagram, prediction

1. Introduction [12]. Besides, the description of Du and Schmid-Fetzer [13] leads to


some solid phases decomposing at temperatures lower than room
The present paper is the first part of our work dedicated to the temperature (see Fig. 1) and contradicts the standard enthalpies of
thermodynamic description of the ternary system Al–Ta–Ti: it formation of solid alloys measured by synthesis calorimetry in [21]
includes experimental investigations of phase equilibria and the and [22]. Though the description [13] formally loses its validity
thermodynamic description of the constituent binary system Al–Ta. below room temperature, this fact confirms that enthalpy of
The Al–Ta system is of interest for technologies related with films formation of the decomposed phases should be more negative to
[1–5], coatings [6] (oxidation), barrier layers [5], amorphization by some extent which is also indicated by the calorimetric data of
rapid solidification [7,8], and mechanical alloying [9] (high-energy [21,22].
rod-milling), as well as for reinforcing and refining of Al alloys [10,11]. Finally, our own experiments with several as-cast and annealed
Preceding thermodynamic descriptions of the binary Al–Ta Al–Ta alloys do not confirm the eutectic-type phase diagram
system were proposed by Kaufman [12] and Du and Schmid-Fetzer proposed in [13,19,20,23]. A phase diagram with a cascade of invariant
[13]. The phase diagram calculated by Kaufman differs from the peritectic reactions seems to be more appropriate, as the one
experimental one from Subramanian et al. [14], accepted in hand- proposed in [14] and later accepted in the handbooks [15] and [16].
books [15,16], by congruent melting of the compounds TaAl2 and In the present work we have performed a thorough critical
TaAl3. The thermodynamic assessment of Du and Schmid-Fetzer assessment of the published data and a series of key experiments,
[13] is based on a larger number of literature data and includes the followed by the elaboration of a new thermodynamic description
data of Mahne et al. [17–19], those were given a large weight. As for the Al–Ta system.
a result, the phase diagram calculated using the thermodynamic
parameters from [13] (Fig. 1) is close to the preliminary version of 2. Experimental information
Mahne et al. [19] (repeating that of Schuster [20]), and it also differs
from the phase diagrams of Subramanian et al. [14] and Kaufman 2.1. Literature data on crystal structure and phase equilibria

* Corresponding author. Tel.: þ49 241 8098007; fax: þ49 241 38578. The crystallographic data of the Al–Ta phases with phase desig-
E-mail address: v.vitusevych@access-technology.de (V.T. Witusiewicz). nations used in literature sources are listed in Table 1. In literature

0966-9795/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.intermet.2009.06.015
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 93

there are some rather different versions of the phase diagram con-
structed by authors substantially basing on own experimental results
[14,18,20,23] and also using only assessed data [12,13,15,16]. The most
recent phase diagram proposed by Du and Schmid-Fetzer [13] in 1996
(see Fig. 1) was based on CALPHAD modeling of available phase
equilibria data (mainly [19,20,32]) and thermodynamic properties
[35]. It was used in later works for ternary systems [33,34] and in
Landolt–Börnstein encyclopaedic issues [36,37]. An overview of
literature data on phase structure and equilibria is presented below.
Works of Glazov et al. [10,11] were first devoted to phase equi-
libria in the system. Using XRD and microhardness measurements,
they reported the solid Ta solubility to range from 0.07 wt.%
(0.010 at.%) at room temperature to 0.22 wt.% (0.033 at.%) at 913 K.
The temperature of invariant reaction Lp þ 3 4 (aAl), 941 K was
obtained in [10] as thermal arrests in DTA cooling curves. The DTA
thermocouple was calibrated against the melting points of pure Zn
and Al, 692 and 933 K, respectively.
Yeremenko et al. [38] studied the Ta solubility in liquid Al using
a rotating disk method and obtained values from 0.105  0.008 wt.%
(0.016 at.%) to 0.37  0.02 wt.% (0.055 at.%) at 973 and 1173 K,
respectively. The former value is in good agreement with the
observation of [10,11] that the 0.10 wt.% Ta addition (0.015 at.%) acts
as a refinement of Al alloys while at 0.075 wt.% Ta (0.011 at.%) the
effect is rather weak. Yatsenko et al. [39] used soaking of Al melt by
TaAl3 in Ta crucibles (and also in graphite crucibles) in the temper-
ature interval 953–1323 K and gave the dependence

Fig. 1. Al–Ta phase diagram calculated using thermodynamic description of Du and


log xTa ¼ 0:0263  4:56=T;
Schmid-Fetzer [13].
where xTa is a mole fraction of Ta. When calculating, the data of
Yatsenko et al. [39] (0.009 and 0.037 at.% at 973 and 1123 K, respec-
tively) are somewhat lower than those of Yeremenko et al. [38].

Table 1
Crystallographic data of solid phases of the Al–Ta system.

Phase designation Pearson Space group Prototype Struktur bericht Lattice parameters (pm) Comments/References
symbol designation
a,a b,b c,g
(aAl),a fcc_A1 cF4 Fm3m Cu A1 404.96 [15], pure Al at 25  C
405.30 [10], 0.036 at.% Ta
b, (Ta), bcc_A2 cI2 Im3m W A2 330.30 [15], pure Ta at 25  C
329.6 [24], w2 at.%Al
3, TaAl3 tI8 I4/mmm TiAl3(h) D022 383.4 853.6 [25,26]
384.88 859.82 [3]
383.9 853.5 [20,27]
Ta39Al69 cF432 F43m Ta39Al69 1915.3 (1) [18,19], Al-rich
1917.2 [24], Ta-rich
TaAl2 cFw400 F43m Pt3Zn10b 1915.7 (2) [14,28]
Ta12Al17 cF464 1931.5 [29]
Ta2Al3(ht) oP* 1107.15 687.17 479.45 [20]
Ta2Al3 hP230 1277.6 (5) 2704 (4) [30]
Ta2Al3 aPw68 1074.8, a ¼ 90.45 1107.8, b ¼ 97.60 1043.3, g ¼ 63.19 [19]
AluTa
u z 1.5
Ta2Al3(lt) oP* 1556.14 1394.7 1266.28 [20]
Ta5Al7 hPw730 3214(2) 1341 (1) [19]
AlvTa
v z 1.4
4, Ta48Al38 (b) mP86 P21/c 987.07 (1) 987.66 (1), b ¼ 116.478(1) 1635.39 (2) [24]
TaAl mP* 1486.5(3) 986.3 (2), b ¼ 99.98 (2) 986.2(2) [17]
Ta17Al12 cI58 aMn 988 [29]
s, Ta2Al tP30 P42/mnm sCrFe D8b 982.8 523.2 [31]
984.2 (1) 521.7 (1) 60 at.% Ta [24]
984.5 (1) 521.7 (1) 61 at.% Ta [24]
989.1 (1) 519.8 (1) 65 at.% Ta [24]
994.2 (1) 519.8 (1) 72.4 at.% Ta [24]
995.7 (1) 519.9 (1) 74 at.% Ta [24]
1001.4 (1) 521.1 (1) 76 at.% Ta [24]
a
Designations given in bold letters are used throughout the present work.
b
Pt3Zn10 or Mg6Pd, Fe11Zn39, Mg44Rh7, Na6Tl, Cu41Sn.
94 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

The solubility of Al in solid Ta is at the level of some percents as u z 1.5, was reported to be formed below about 1500 K in solid state,
estimated by [32,40], and [24] reported w2 at.% at 1733 K. and almost single-phase samples to be prepared from the metal
Phases 3, Ta39Al69 and s were reliably identified in both as-cast mixture with xTa ¼ 0.40(5) by heating in sealed Ta ampoules at
and annealed alloys. The phases 3 and Ta39Al69 are so easily formed 1420 K for 3 days. The latter, hexagonal Ta5Al7 designated in [19] as
during solidification that melt spun ribbons from alloy Al-36 at.% AlvTa with v z 1.4 (xTa ¼ 0.42), was claimed to be obtained free of
Ta were reported to contain 3 and Ta39Al69 together with an impurities by heating in the similar way (also 3 days exposure at
amorphous phase Ta58–62Al42–38 [41]. The crystal structure of the 1420 K). The X-ray diffraction patterns presented in [19] for AluTa
3-phase, TaAl3 was solved as early as 1938 by Brauer [25,26]. and AlvTa (triclinic Ta2Al3 and hexagonal Ta5Al7) cannot be
Taking into account EPMA measurements in [5,14], which are in completely indexed on the basis of 4 well-known phases, 3 (TaAl3),
good agreement with data for isostructural TiAl3 [42] and NbAl3 Ta39Al69, 4 (Ta48Al38), and s (Ta2Al). There was only one article [20]
[43], the homogeneity range of 3 may be estimated to be about that may be considered as confirmation for the triclinic phase of
0.5–1 at.% Ta. Subramanian et al. [14] proposed that the 3 homo- Mahne et al. [19]. The XRD pattern of orthorhombic l.t.-Ta2Al3,
geneity range is somewhat shifted from the stoichiometry TaAl3 to formed during annealing at 1273 K for 500 h, taken from [20] seems
about 24 at.% Ta. EPMA data of [5] fell within an interval of to be close to the pattern published in [19] for the triclinic AluTa
24.5  0.5 at.% Ta. In the literature there are both variants of (see Table 1). The XRD pattern of orthorhombic l.t.-Ta2Al3 is not
melting for the 3-phase, congruent at 1823 [23] or 1900 K [20] and consistent with that for 4 (TaAl) or a mixture of 3 (TaAl3) and 4, as it
peritectic at 1773 [44] or 1824 K [14]. should be assumed from the microprobe data of [5,14]. In [16] three
The crystal structure of Ta39Al69-phase was resolved in [18] on samples were two-phase 3 þ 4 after annealing at 1423 K for 500 or
single crystalline sample prepared by annealing in the temperature 1000 h, as found by EPMA measurements. In [5] the same phase
interval between 1720 and 1820 K. It is fcc with a large cell struc- constituents were identified by EPMA in the samples Al-30.4 at.% Ta
ture of new type (see Table 1). The results of [18] are in full and Al-40.3 at.% Ta after annealing at 1273 and 1173 K for 35 days
agreement with TEM examinations of [28,45]. It is obvious that the (840 h). It should be noted that no report on a ternary system based
fcc Ta12Al17-phase of [29] is Ta39Al69 (see Table 1). It should be on Al–Ta was found presenting any experimental identification of
noted here that phases reported as orthorhombic h.t.-Ta2Al3 in [20] more than one phase in the range from 35 to 45 at.% Ta. Palm et al.
and hexagonal Ta2Al3 in [30] (see Table 1) are really fcc Ta39Al69, [46] reported that in Al–Ni–Ta alloys annealed at 1273 and 1523 K
because their experimental XRD patterns represented in [20,30] are there is the phase containing 2–3 at.% Ni and 36–39 at.% Ta (Al to
consistent with the theoretical ones calculated on the basis of balance), as found by EPMA, and its XRD peaks were indexed as
atomic coordinates of [18]. EPMA measurements of [5,14,24] orthorhombic h.t.-Ta2Al3 of [20], i.e. the Ta39Al69-phase. Thus, in
revealed a phase composition of 34–36 at.% Ta, which is close to the literature there are data incompatible with the stability of triclinic
stoichiometry of Ta39Al69 [18]. Since the crystal structure of the Ta2Al3 or AluTa (u z 1.5) and hexagonal Ta5Al7 or AlvTa (v z 1.4)
Ta39Al69-phase includes Ta and Al atoms with the same coordina- reported in [19] and later accepted by [13]. It is obvious that addi-
tion number, equal to 14, a certain mutual substitution may be tional experimental studies are needed to clarify this question.
expected giving rise to a noticeable homogeneity range. The phase close to equiatomic TaAl (in the present work
Literature data are fairly consistent concerning the stability of the designated as 4) was reported in [14,17,19,34]. Although Sub-
Ta39Al69-phase at high temperatures, but contradictory with respect ramanian et al. [14] claimed the peritectic reaction Lp þ s 4 4 to
to its low-temperature limit of stability, the decomposition take place at 2043 K, this phase never was found in as-cast alloys
temperature and the decomposition products. Mahne et al. [19] [17]. Furthermore, as reported [17] it decomposed at a temperature
proposed a decomposition temperature of 1445 K, and in their phase close to melting in the course of solid-state reaction and accord-
diagram there are samples shown in the two-phase 3 þ Ta2Al3 field ingly after [19] it is of a peritectoid type. In [17] the powder
at 1430 and 1310 K. At contrary in [14] the samples with 28.8– diffraction pattern of TaAl taken using a Guinier camera and
46.9 at.% Ta contained the Ta39Al69-phase after annealing at 1423 K monochromatic Cu Ka1 radiation was indexed on the basis of
for 500 or 1000 h. According to [5] annealing of Al–Ta alloys with primitive monoclinic unit cell (see Table 1). More precisely the
30.4–40.3 at.% Ta at 1273 and 1173 K for 840 h (35 days) resulted in structure of the 4-phase was determined by Rietveld refinement of
a two-phase 3 þ 4 microstructure. It should be noted that their powder synchrotron diffraction pattern in [24] (analyzed compo-
samples after annealing at 1023 and 1123 K for the same duration sition Ta50.7Al49.3, refined composition Ta52.6(5)Al47.4(5), stoichio-
contained two phases, 3 þ Ta39Al69 or Ta39Al69 þ 4 depending on the metric composition Ta48Al38). The width of the Ta48Al38-phase (4)
alloy composition, and the authors claimed the Ta39Al69-phase to be at 1733 K was identified to range from 50.7 to 56.6 at.% Ta. This
stable at 1123 K and lower. However, the alloy Al-38.5 at.% Ta was phase was originally named b in [24] but to avoid confusion in the
reported in [5] to consist of two phases with phase composition of ternary Al–Ta–Ti system, where b usually denotes the (b-Ti,Ta)
24.7 and 35.8 at.% Ta (i.e. 3 and Ta39Al69). Because the integral alloy solution phase (see [34]) it is designated here as 4.
composition was 38.5 at.% Ta, there had to exist a phase richer in Ta The s-phase was unambiguously identified in the Ta-rich side
to fulfill mass balance. This phase was apparently not detected. All for a wide concentration range including the ideal composition
these facts evidence occurrence of the decomposition of the Ta2Al [6,7,14,19,20,24,29,31,32,44,47]. Lately, the homogeneity
Ta39Al69-phase, which was far from finishing at the annealing range of the s-phase was investigated by Boulineau et al. [24]. They
temperatures 1023 and 1123 K. Thus, based on [5,14] one can used Rietveld refinement of X-ray powder data obtained from
conclude that the decomposition temperature of Ta39Al69-phase is different samples containing from 60 to 76 at.% Ta, which were
between 1273 and 1423 K. Besides, Schuster [20] reported a poly- annealed at 1273 and 1733 K, as well as EPMA examination. At
morphism for the Ta39Al69-phase and a corresponding DTA arrest at 1733 K the s-phase extends from 60.1 to 76 at.% Ta, and at 1273 K
1498 K upon heating, which is not in sharp disagreement with the the Ta-lean limit was found to be at 61 at.% Ta. The data on the Ta-
decomposition temperature in the interval 1273–1423 K. lean limit are in good agreement with those of [14,29], contrary to
Triclinic Ta2Al3 and hexagonal Ta5Al7 phases (see Table 1) were [19,32,40]. In [19] the sample Al-64.2 at.% Ta annealed at w1510 K
solely reported in the work of Mahne et al. [19]. Both phases were was shown in the two-phase field f þ s, while in [14] the alloy Al-
identified using powder XRD diffraction patterns (Guinier camera 60.1 at.% Ta was single-phase s after annealing at both 1423 and
and monochromatic Cu Ka1 radiation) together with electron 1673 K for 500 and 100 h, respectively. It should be noted that
diffraction (TEM). The former phase, designated in [19] as AluTa with according to XRD analysis [20] the as-cast alloy Al-50 at.% Ta was
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 95

Table 2
Composition and phase identification for as-cast and annealed samples; annealing was performed under argon.

No. Composition, at.% Ta XRD and EBSD phase identification

Nominal EDS analysis As-cast alloy Annealed alloy

Primary phase Phase constituents Conditions Phase constituents


1 23.0 20.5 3 (aAl) þ 3
2 33.0 30.8 Ta39Al69 3 þ Ta39Al69 1773 K/2 h 3 þ Ta39Al69
1593 K/6 h þ 1473 K/75 h 3 þ Ta39Al69
3 39.0 37.6 s Ta39Al69 þ s 1833 K/2 h Ta39Al69 þ s
1593 K/6 h þ 1473 K/75 h Ta39Al69
4 45.0 44.1 s Ta39Al69 þ s 1833 K/2 h Ta39Al69 þ s
5 46.5 45.7 s Ta39Al69 þ s 1833 K/10 h Ta39Al69 þ s
1373 K/8 h 3 þ Ta39Al69 þ 4
6 58.5 58.4 s s 2023 K/1.5 h s
1593 K/6 h þ 1473 K/75 h 4þs
7 80 88 b sþb 2273 K/1.0 h sþb

single-phase s. The temperature of the peritectic reaction from [20], the congruent melting of TaAl3 and the existence of
Lp þ b 4 s was reported to be about 2373 [44] or 2273 K [23]. eutectic formed by Ta39Al69 and s-phase.
For equilibria with the liquid phase, the measurements of Sub-
ramanian et al. [14], Kimura et al. [23], Schuster [20], and Wilhelm 2.2. Experimental procedure and results
and Witte [44] are the experimental data available. Mahne et al.
(see Fig. 1 in their articles [17,18]) used the tentative liquidus curve 2.2.1. Preparation of alloys
of [20] to mark the compositions of as-cast alloys prepared for their In order to verify phase equilibria in the Al–Ta system a number
investigation. These artificial points, which were accorded high of 7 different samples with a mass of 15 g each were prepared by
confidence, were entered in the process of modeling in [13]. As arc melting. The sample compositions ranging from 20.5 to 88 at.%
a result, it reproduced distinctive features of the phase diagram Ta are listed in Table 2. Arc melting with a non-consumable

Fig. 2. Microstructure of as-cast samples: (a) 3 primary solidification in sample Al-20.5 at.% Ta, (b) Ta39Al69 primary solidification in sample Al-30.8 at.% Ta, (c) s primary solidi-
fication in sample Al-45.7 at.% Ta and (d) s single-phase solidification in sample Al-58.4 at.% Ta.
96 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

tungsten electrode was carried out on a water-cooled copper 2.2.2. XRD, SEM/EBSD and SEM/EDS analysis of the samples
hearth under purified argon. The alloys were melted several times Microstructure constituent phases were analyzed both in the as-
to insure homogeneity. The initial materials were metallic Al cast state and after annealing at sub-solidus and other tempera-
(99.99 wt.% Al) and Ta (99.9 wt.% Ta). To prevent extensive tures by XRD and SEM (EBSD and EDS). The results are summarized
evaporation of Al the individual melting cycles were short. As in Table 2 and will be discussed in more detail below. Annealing
shown by reducing extraction in a Ni bath followed by chroma- was performed in a resistance furnace with a tungsten heater in
tography, the oxygen content in the samples ranged from 0.01 to argon atmosphere gettered by Ti cuttings. After annealing the
0.04 wt.%, (100–400 wt-ppm) and the contamination by N and H samples were cooled to room temperature at a rate of 180 K min1.
was lower than the threshold of sensitivity (about 0.001 wt.% N XRD-measurements were carried out in a DRON-3 diffractometer
and 0.003 wt.% H). on powdered samples (a step of 2$q was 0.05 , an exposure of 8 s).

Fig. 3. Crystal structure and EBSD data for the Al–Ta core phases: a) 3-phase in the as-cast and annealed samples Al-20.5 at.% Ta and Al-30.8 at.% Ta, b) Ta39Al69-phase in the as-cast
and annealed samples containing 30.8–45.7 at.% Ta, c) s-phase in the as-cast and annealed samples containing 37.6–88 at.% Ta and d) 4-phase in the sample Al-45.7 at.% Ta annealed
at 1373 K for 8 h.
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 97

Fig. 4. Micrographs of annealed samples: (a) Al-30.8 at.% Ta annealed at 1773 K for 2 h (grey – Ta39Al69, black – 3) (b) Al-37.6 at.% Ta annealed at 1833 K for 2 h (grey – Ta39Al69,
white – s); (c) Al-44.1at.%Ta annealed at 1833 K for 2 h (black – Ta39Al69, white-s); (d) Al-45.7at.%Ta annealed at 1373 K for 8 h (white – 4, grey – Ta39Al69, black – 3).

Scanning electron microscopy (SEM), electron backscatter diffrac-


tion (EBSD) and energy dispersive X-ray analysis (EDS) were carried
out in a SEM type Gemini 1550 equipped with INCA and INCA-
Crystal. For the quantitative analysis of all EDX spectra the
composition of the Ta39Al69-phase, e.g. 36.1 at.% Ta [5,19,24], was 70
used as an internal standard. EDS measurements acquired from Al - 58.4 at.% Ta
65 Al - 45.7 at.% Ta
large sample areas were used to identify the integral chemical Al - 30.8 at.% Ta
composition of the samples. 60
Fig. 2a–d shows the microstructure of as-cast samples, with the
55 σ
primary solidification phase being rich in Ta and hence appearing
white or light grey in the electron backscatter images: the primary 50
Ta (at.%)

phase was identified to be 3 in sample Al-20.5 at.% Ta, Ta39Al69 in 45


sample Al-30.8 at.% Ta, s in samples with 37.6–58.4 at.% Ta and b in Ta39Al69+σ
sample Al-88 at.% Ta. These phases were identified by XRD and also 40 Ta39Al69
EBSD. Characteristic Kikuchi patterns obtained for the intermetallic 35
phases are shown in Fig. 3a–c. After annealing at temperatures
30 Ta39Al69+ε
1773–2073 K the same phases as in the as-cast state were detected
(see Table 2 and Fig. 4a–c). Annealing of the as-cast sample Al- 25 ε
45.7 at.% Ta at 1373 K for 8 h led to decomposition of the s-phase (αAl)+ε
20
and formation of a three-phase microstructure composed of 0.0 0.2 0.4 0.6 0.8 1.0
Ta39Al69 þ 3 þ 4 (see Fig. 4d; the EBSD pattern for the 4-phase in
Fraction of solid
this sample is shown in Fig. 3d). This microstructure is considered
to be a non-equilibrium three-phase structure formed from the Fig. 5. The distribution of tantalum in the as-cast samples Al-30.8 at.% Ta, Al-
initially two-phase Ta39Al69 þ s sample due to the fact that the 45.7at.%Ta and Al-58.4 at.% Ta was measured by EDS grid technique and was evaluated
as function of the cumulative fraction of solid. All samples show that the Ta-content
annealing temperature is close to the temperature of decomposi-
decreases as solidification proceeds. The distinct levels correspond to the distinct
tion of the Ta39Al69-phase. The presence of the intermetallic phases phases formed along the solidification paths of the samples. The corresponding phase
Ti2Al3 and Ti5Al7 reported by Mahne et al. [19] in annealed samples regions are labelled accordingly.
98 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

Table 3
Pyrometric and DTA data measured for the Al–Ta samples with a heating/cooling rate of 20 K min1 in comparison with the transformation temperatures calculated using the
present thermodynamic description.

Alloy number Composition (at.% Ta) Phase transition Temperature (K)

Measured Calculated

Heating Cooling
1 20.5 (aAl) þ 3 4 L þ 3 933; 933.8  0.2a 924 933.7
Lþ34L <1848 1804 1810
2 30.8 3 þ 4 4 Ta39Al69 þ 3 1372 – 1371
Ta39Al69 þ 3 4 L þ Ta39Al69 1822 1806 1814
L þ Ta39Al69 4 L 1876 1879 1874
3 37.6 3 þ 4 4 Ta39Al69 þ 4 1387 – 1371
Ta39Al69 þ 4 4 Ta39Al69 þ s – – 1737
Ta39Al69 þ s 4 Ta39Al69 1820 – 1759
Ta39Al69 4 L þ Ta39Al69 1877; 1881  8b 1855 1881
L þ Ta39Al69 4 L þ s 1884 – 1885
Lþs4L 1978 2002 1972
4 44.1 3 þ 4 4 Ta39Al69 þ 4 1353 – 1371
Ta39Al69 þ 4 4 Ta39Al69 þ s 1740 – 1737
Ta39Al69 þ s 4 L þ s 1885; 1895  13b 1862 1885
Lþs4L 2064 2063 2074
5 45.7 3 þ 4 4 Ta39Al69 þ 4 1355 – 1371
Ta39Al69 þ 4 4 Ta39Al69 þ s 1727 1662 1737
Ta39Al69 þ s 4 L þ s 1876 1855 1885
Lþs4L 2139 2159 2097
6 58.4 4þs4s 1725 – 1716
s4Lþs 2111 2063 2099
Lþs4L 2257 2164 2251
7 88.0 bþs4Lþb 2353  40b – 2347
Lþb4L – – 2980
a
Measured value by DSC.
b
Pyrometric measurement of incipient melting temperature.

containing from 30.8 to 45.7 at.% Ta was not confirmed by XRD and In agreement with the observations of Subramanian et al. [14]
EBSD analysis. The alloys Al-30.8 at.% Ta, Al-37.6 at.% Ta and Al- no evidence of eutectic microstructures were observed in as-cast
58.4 at.% Ta annealed at 1593 K for 6 h and then at 1473 K for 75 h samples within the composition range 37.5–46.9 at.% Ta. More-
according to XRD were 3 þ Ta39Al69, single Ta39Al69 and 4 þ s over, EDS analysis of element microsegregation in the as-cast
materials, respectively (see Table 2). samples Al-30.8 at.% Ta, Al-46.7 at.% Ta and Al-58.4 at.% Ta showed
that solidification of the Al–Ta alloys always proceeds with
5 a positive segregation of Al into the liquid. This is true also for the
1355 1727 1876
2139

45.7 at.% Ta
sample Al-30.8 at.% Ta with Ta39Al69 being the primary solidifica-
tion phase. Fig. 5 shows the measured microsegregation profiles.
Accordingly, the eutectic reaction Le 4 Ta39Al69 þ s proposed by
1380 1877 [20,23] and further adopted by [13,19] has not been verified. From
2161

37.6 at.% Ta 1820


Fig. 5 it may also be observed that both phases Ta39Al69 and 3
0 1978
1884 display a narrow homogeneity range instantly below solidus,
extending from 35.9 to 37.6 at% Ta and from 23.8 to 25.0 at.% Ta,
respectively.
1372 1822

30.8 at.% Ta
ΔT (K)

-5 10 K min-1 939.1
1876

4
-1
5 K min 937.4

933 1805 1890 936.0


2 2 K min-1
Heat Flow (mW)

20.5 at.% Ta 1 K min-1 935.4


-10
0

-2

-15
1848
-4
1000 1250 1500 1750 2000
Temperature (K) 934 936 938 940 942
Temperature (K)
Fig. 6. DTA heating curves obtained for alloys Al-20.5 at.% Ta, Al-30.8 at.% Ta, Al-
37.6 at.% Ta and Al-45.7 at.% Ta at heating rates of 20 K min1 using annealed samples. Fig. 7. DSC-curves obtained at different heating rates for the annealed Al-20.8 at.% Ta
For clarity, the curves have been shifted vertically. samples.
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 99

Al-20.8at.%Ta:
939
Tap= (935.10+/- 0.17)+0.405 V

938
Temperature (K)

937

936 Al:
Tap= (934.55 +/- 0.11)+0.378 V
935

934

Al: Tref = 933.28 K


933
0 2 4 6 8 10
Heating rate (K min-1)

Fig. 8. Results of DSC measurements of the melting temperature of pure Al (circles)


and the solidus temperature of Al-20.8 at.% Ta alloy (stars) as function of the applied
heating rates. Lines and expressions represent the linear fit, extrapolated to an infi-
nitely small heating rate, e.g. to 0 K min1.

2.2.3. DTA analysis and incipient melting


Temperatures of solid-state and solid–liquid phase trans-
formations were determined by DTA using Al2O3 and Sc2O3
crucibles and W/W-20Re string thermocouples designed by
Kocherzhinskiy et al. [48,49]. The DTA measurements were per-
formed under high purity He with heating and cooling rates of 20
or 40 K min1. The thermocouple was calibrated using the IPTS-
90 reference points of Al, Au, Pd, Pt, Rd and also additional
reference points of Fe and Al2O3. The solidus temperature for the
samples number 3, 4 and 7 were also determined from the
observation of incipient melting, using the method developed by
Pirani and Alterthum [50]. Bar-shaped specimens, which were
clamped between two water-cooled copper electrodes through
tungsten inserts (tips of electrodes and plate washers), were
resistively heated under high purity Ar (slightly higher than the
atmospheric pressure). The temperature was measured optically
on the background of a black body hole (the diameter to depth
ratios were about 1–4) with a disappearing filament-type
pyrometer ‘‘EOP-68’’ of standard quality level. Its maximum
instrumental errors amount to 2.8 K for the temperature region
1170–1670 K and 4 K for 1670–2270 K. It was calibrated and
certificated by the Ukrainian National Scientific Centre ‘‘Institute
of Metrology’’.1 The temperature of incipient melting was read in
the moment of formation of a spot on the background of the
black body hole. This moment also corresponds to a change of the
Fig. 9. Phase diagram of the binary system Al–Ta calculated with the present ther-
sample shape at the bottom (which becomes half-spherical) and
modynamic description (solid lines) in comparison with critically selected experi-
(or) to a change of the reflection at the sample surface [51]. The mental data from [5,14,23,24,29,32, 40,44] and the experimental results obtained in
latter turns smooth due to filling the surface roughness by melt. A the present work (symbols) as well as pyrometric data of [20] (dotted line).
number of 5 measurements were carried out for each sample and
afterwards the average value as well as the standard deviation
was determined. transition temperatures would not be affected by segregation
The phase transformation temperatures detected by DTA and inherited from casting. As shown in Fig. 6 the sample with
the incipient melting temperatures obtained by Pirani–Alterthum 20.5 at.% Ta starts to melt at 933 K that is related to the (aAl)
method are listed in Table 3, together with the values calculated phase. At temperatures higher than 1805 K the DTA curve
with the proposed thermodynamic description of the Al–Ta demonstrates a sharply defined peak, the shape of which indicates
system. Fig. 6 depicts selected heating curves obtained by DTA for that the slope of the liquidus line at this composition is small and
a series of samples with Ta-content of 20.5, 30.8, 37.6 and that a modest increase of the temperature causes melting of large
45.7 at.%. These samples were given an annealing treatment prior mole fraction of primary 3-phase. The process of melting of this
to DTA processing, in order to make sure that the measured phase phase finishes at a temperature rather lower than the maximal
deviation of the DTA signal from the baseline (1848 K). The DTA
curves for the samples with 30.8–45.7 at.% Ta contain a sluggish
thermal effect, the peak of which starts at w1370 K and finishes at
1
http://www.metrology.kharkov.ua/eng/centres/NC-1.htm. w1530 K, being associated with the eutectoid reaction
100 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

Ta39Al69 4 3 þ 4. Melting of the sample Al-30.8 at.% Ta starts at


1822 and finishes not higher than at 1890 K. The solidus (DTA
results) or incipient melting temperatures (pyrometric data) of the
samples containing 37.6–45.7 at.% Ta are in good agreement (see
Table 3). The well defined peak at 1727 K on the DTA curve of Al-
45.7 at.% Ta sample corresponds to a solid-state reaction, which is
considered as s 4 d þ 4. Thus, this sample is two-phase L þ s
between solidus and liquidus, and it is two-phase Ta39Al69 þ s
below solidus in the temperature interval of w150 K (see Table 3).
Below the temperature of the solid-state reaction at about 1727 K,
the sample is in the two-phase field Ta39Al69 þ 4 existing up to the
decomposition reaction Ta39Al69 / 3 þ 4 (w1373 K).

2.2.4. DSC measurements for alloy Al-20.8 at.% Ta


The temperature of the peritectic reaction L þ 3 4 (aAl) was
reported in [10] to be 941 K. As mentioned in Section 2.2.3, our own
DTA measurements showed this reaction to be at 933 K. In order to
cross-check this, a series of DSC measurements was performed
with alloy Al-20.8 at.% Ta in a Perkin–Elmer PyrisTM DSC-7 calo-
rimeter, using heating rates of 1, 2, 5 and 10 K min1. The calo-
rimeter was calibrated for each of the heating rates by measuring
the melting point of high purity Al supplied by Perkin–Elmer, with
the melting point (Tref) being equal to 933.28 K.
For each DSC-run a sample of 30–35 mg, both from the Al–Ta
alloy and the Al reference was used. After introducing the sample
into the DSC furnace and sealing the furnace tightly, the pressure of Fig. 10. The enlarged portion of the Al–Ta phase diagram in the region of the Ta39Al69-
and 3-phase: solid lines result from the present description; symbols are experimental
the inner gas (Ar with purity 6 N) in the furnace automatically rose
data form [5,14,23] and from the present work; dotted line are pyrometric data of [20].
to about 100 Pa. The solidus temperature of alloy Al-20.8 at.% Ta or
the melting point of pure Al were read at the intersection point metastable structures of the pure elements were adopted from the
between the slope of the melting curve at the inflection point and SGTE database compiled by Dinsdale [52].
the heat capacity baseline. The DSC-curves for the Al-20.8 at.% Ta The model parameters were evaluated by searching for the best
samples in vicinity of solidus are shown if Fig. 7. The results from all fit to the experimental data, e.g. to phase equilibria and thermody-
runs are summarized in Fig. 8. The temperature of the invariant namic properties of various alloys, using the PARROT optimizer
peritectic reaction L þ 3 4 (aAl) and the melting temperature of of Thermo-Calc [53]. During the process of optimization the present
pure Al were determined by linear extrapolation of the obtained DTA and pyrometric data, as well as the DTA data reported by Sub-
results to a heating rate of 0 K min1. The difference between the ramanian et al. [14] were accorded a weight of 2. A weight of 2 was also
measured melting point of Al Tm ¼ 934.55 K and the reference value assigned to the calorimetric data of Meschel and Kleppa [22]. Liquidus
Tref ¼ 933.28 K was applied as correction. The temperature of the data obtained by pyrometric measurements reported by Kimura et al.
peritectic reaction LP þ 3 4 (aAl) was hence determined to be [23] and Schuster [20] were included into the optimization
933.8  0.2 K. procedure with weight of 0.5. All other literature data used
Thus, the existence of 4 intermetallic phases 3, Ta39Al69, 4, and [5,21,10,11,24,29,32,38–40,44,54] were assigned a weight equal 1.
s is reliably established. For the temperature range close to
melting our results are in good agreement with [14], for the
exception of the formation of 4-phase in solid state and a more
pronounced extension of s-phase to lower Ta-content (almost
reaching 50 at.%, just as it was reported by Schuster [20]). The
Ta39Al69-phase has been shown to decompose at temperatures
lower than 1370 K. The triclinic Ta2Al3 and hexagonal Ta5Al7 pha-
ses reported by Mahne et al. [19] were not confirmed and are not
included in our modeling.

3. Thermodynamic models and optimization procedure

For the liquid, fcc_A1 or (aAl), bcc_A2 Ta or b and s phases we


used the thermodynamic models stemming from the previous
thermodynamic descriptions of the Al–Ta system by [13]. For the
nonstoichiometric phases 3, Ta39Al69 and 4 two-sublattice models
were used, respectively: (Al)3: (Al,Ta%)1, (Al%,Ta)69: (Al,Ta%)39 and
(Al,Ta)48: (Al,Ta)38, where different sublattices are separates by
colon and the percent sign denotes which species are considered to
be major constituents in the sublattice.
The expressions for the molar Gibbs energy as function of Fig. 11. The Al-rich portion of the Al–Ta phase diagram: solid lines result from the
temperature and composition for all models mentioned above are present description, dotted lines result from the description proposed by Du and
given in Ref. [13]. The Gibbs energy descriptions for the stable and Schmid-Fetzer [13] and symbols are experimental data from [10,11,38,39].
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 101

The experimental data were sufficient to model the enthalpic homogeneity ranges reported for isostructural TiAl3 [42] and NbAl3
and the entropic terms of the Gibbs free energy for all phases [43] phases. From Fig. 10 one can observe that the 3-phase with
independently. No excess heat capacity terms were introduced. The composition of 24.3 at.% Ta melts incongruently at temperature
gas phase was included, in order to allow extrapolation up to 1819 K by forming liquid, which is only of about 0.5 at.% richer in Al.
temperatures of 6000 K. The assessed parameters and complete A difference between temperatures of this peritectic invariant
thermodynamic database in Thermo-Calc format [53] are included reaction L þ Ta39Al69 4 3 and liquidus for the given above compo-
in the Appendix 1 and Appendix 2, respectively. sition is equivalent only to 4 K. Probably due to such diminutive
differences the congruent melting of this phase was reported in
4. Results of modeling and discussion works of Kimura et al. [23] and Schuster [20]. The given tempera-
ture of melting is in fair agreement with value reported in the
The phase equilibria and thermodynamic properties presented former work (1823  50 K) and deviates to 77 K from the value
in this section were calculated using the software Thermo-Calc [53] (1900  10 K) found by pyrometry in the latter work. Wilheim and
along with the thermodynamic description obtained from the Witte [44] also reported the melting temperature of the Ta39Al69-
above optimization. phase amount to 1773  50 K that within the limit of experimental
uncertainty agrees with the result of present description.
4.1. Phase equilibria The Al-rich part of the Al–Ta phase diagram is presented in
Fig. 11, with focus on the solubility of Ta in (aAl) solid solution and
Fig. 9 shows the calculated phase diagram without and with the the peritectic reaction L þ 3 4 (aAl). The present thermodynamic
experimental data published in [5,10,11,14,20,23,24,29,32,38– description (solid lines) agrees well with the experimental data on
40,44] and determined in the present work. The majority of the Ta solubility from [10,11,38,39] (points), while the previous
experimental data are well reproduced by the present description. description [13] shown with dashed lines fails to reproduce them.
Even the liquidus data reported by Kimura et al. [23], which were The temperature of the peritectic reaction of (aAl) formation is
rejected in the previous optimization [13], are reasonably repro- calculated to be 933.7 K, which agrees with our own DTA and DSC
duced. Experimental data on solid solubility of Al in Ta of [24,32,40] measurements and is lower than the value of 941 K reported in
are equally well reproduced. The s-phase extends from 51 to 81 at.% [10].
Ta. This is in good agreement with our own EDS measurements on In Table 4 the calculated temperature and composition of phases
sample Al-45.7 at.% Ta annealed at 1833 K, with the experimental coexisting in invariant equilibria are listed for each invariant reac-
results of [20,24,29] as well as with theoretical calculations using tion in the system and compared with available experimental data.
Wilson–Spooner sphere-packing models reported in [47]. The only
congruent formation of 4-phase (contrary to the peritectoid reac- 4.2. Thermodynamic properties
tion Ta39Al69 þ s / 4) is well consistent with its considerable
homogeneity range revealed in [14,24] at the temperatures quite Thermodynamic properties of binary Al–Ta alloys were reported
close to the solid-state reaction that is treated as s / Ta39Al69 þ 4 in literature, being obtained by calorimetry [21,22,54], the Knudsen
in the present work. effusion method [35] and atomic absorption spectroscopy [55].
An enlarged view of the phase diagram in the vicinity of the 3 Fig. 12 displays measured values of the standard enthalpy of
(TaAl3) and Ta39Al69-phases is presented in Fig. 10. It shows the formation of solid Al–Ta phases (points) along with calculated
homogeneity range of the Ta39Al69-phase at sub-solidus tempera- values, using the present description (solid lines) and the
tures extending from 35.9 to 37.7 at.% Ta, which matches well with previous description [13] (dotted lines) of the binary system.
the experimental results presented in Fig. 5. The homogeneity The present description well reproduces the data obtained by
range of the 3-phase at 1500 K extends from 24 to 25 at.% Ta. This is direct synthesis calorimetry [21,22], which is usually considered
consistent with EPMA measurements reported in [5,14] and with to be most reliable. The differences between the calorimetric data
our own EDS measurements. It also corresponds to the [21,22] and results of Schmidt and Franzen [35] are unclear. The

Table 4
Comparison between measured and calculated invariant reactions in the Al–Ta system.

Reaction between phases 41/42/43 Type T, K Composition of Ta in phases, at.% Comment/Reference

41 42 43
Lþb4s Peritectic 2373  50 – – – Experiment [44]
2275  50 63 88 66.7 Experiment [23]
2353  40 – – – Experiment [PW]a
2334 63.7 92.0 76.0 Calculated [13]
2347 67.8 93.0 75.9 Calculated [PW]
L þ s 4 Ta39Al69 Peritectic 1883  10 – 51.2 37.6 Experiment [PW]
1885 32.7 51.0 37.8 Calculated [PW]
s44 Congruent 1854 55.2 55.2 – Calculated [PW]
L þ Ta39Al69 4 3 Peritectic 1822  8 – 35.9 24.9 Experiment [PW]
1814 23.7 35.5 24.4 Calculated [PW]
s 4 Ta39Al69 þ 4 Eutectoid 1733  9 51.0 36.1 – Experiment [PW]
1737 51.1 37.6 52.2 Calculated [PW]
Ta39Al69 4 3 þ 4 Eutectoid 1367  16 35.6 25.1 51.2 Experiment [PW]
1371 36.8 25.0 52.4 Calculated [PW]
L þ 3 4 (aAl) Peritectic 941 w2  102 25 3.7  102 Experiment [10,11]
– 1  102 – – Experiment [38]
– 5  103 – – Experiment [39]
933.8  0.2 – – – Experiment [PW]
935 2.5  102 25 0.45 Calculated [13]
933.7 3.8  103 24.7 3.8  102 Calculated [PW]
a
[PW] denotes the present work.
102 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

Fig. 12. Standard enthalpy of formation of solid Al–Ta alloys at 298 K referred to fcc-Al Fig. 13. The integral enthalpy of mixing at 1850 K referred to liquid Al and liquid Ta:
and bcc-Ta in the Al–Ta system: solid lines result from the present description, dotted solid lines result from the present description, dotted lines result from the former
lines result from the former description [13] of the binary system and symbols are description [13] and symbols are experimental data of [54].
experimental data of [21,22,35].
mixing is shifted towards the Al-rich side, as is the trend for solid
authors measured the rate of Al effusion using continual elec- alloys (see Fig. 12). This seems reasonable since, as a rule, the
trobalance weighing of the Knudsen cell and quadrupole mass composition variation of the integral enthalpy of formation of
spectrometry simultaneously. The measurements at 1321–1713 K liquid and solid alloys should behave similar, if a system is char-
were used to derive reaction enthalpies of incongruent sublima- acterized by moderate and strong interatomic interactions.
tion and then standard enthalpies of formation and atomization.
Unfortunately, the original data were not presented. Possibly, the
studied temperature interval of about 400 K was not appropriate 5. Summary
to obtain reliable standard enthalpies of formation by means of
differentiation. The binary system Al–Ta has been revised within the frame of
The calculated standard thermodynamic properties are rep- activities that aim to develop a thermodynamic description for the
resented in Table 5. The enthalpy of formation of the 4-phase ternary system Al–Ta–Ti. Based on literature data and key experi-
is negative enough to make this phase stable down to 0 K (see ments, a new thermodynamic description was elaborated, involving:
Fig. 9).
Fig. 13 displays the experimentally determined values for the  experimental analysis of 7 distinct alloys in both, as-cast and
integral enthalpy of mixing at 1850 K (points) in comparison with annealed conditions using XRD, SEM/EDS, SEM/EBSD, DSC, DTA
calculated values, using the present description (solid lines) and the and Pirani–Alterthum techniques for determination of micro-
previous description [13] (dotted lines), respectively. Though both structure constituents, composition of coexisting phases and
descriptions properly reproduce the calorimetric data of Sudavt- phase transition temperatures;
sova and Batalin [54] the descriptions differ with increasing Ta-  thermodynamic modeling by the CALPHAD method, taking
content. Compared to [13] the position of the minimum enthalpy of into account assessed experimental information from litera-
ture and own experimental results;
 thermodynamic equilibrium calculations of phase equilibria
Table 5
Standard thermodynamic properties of the Al–Ta phases referred to bcc-Ta and fcc-
and selected thermodynamic properties in comparison to
Al at 298.15 K calculated using the present description (values are given for averaged experimental data.
gram-atom).

Phase Composition D298 G D298 H D298 S


Acknowledgements
(at.% Ta) (J mol1) (J mol1) (J K1 mol1)
3 25.0 27,986 29,954 5.688
Ta39Al69a 36.1 24,850 25,900 3.506
The authors deeply appreciate Prof. T.Ya Velikanova for fruitful
4 55.8 23,900 24,654 2.515 discussions of results. They thank L.V. Artyukh, L.-A. Duma, V.V.
s 55.0b 24,005 24,788 2.611 Garbuz, and N.I. Tsyganenko for technical assistance and would
60.0b 22,859 23,622 2.542 like to express their gratitude for financial support from the
65.0 21,619 22,394 2.580
Integrated Project IMPRESS, ‘‘Intermetallic Materials Processing in
66.7 21,129 22,050 3.071
70.0 19,229 19,867 2.122 Relation to Earth and Space Solidification’’ (Contract NMP3-CT-
75.0 16,142 16,556 1.370 2004-500635) co-funded by the European Commission in the
80.0 12,951 13,246 0.969 Sixth Framework Programme and the European Space Agency
a
Phase is metastable at 298.15 K. and the German Federal Ministry of Research (Contract FKZ
b
Given composition is metastable at 298.15 K. 50WM0843).
V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 103

Appendix 1. List of assessed thermodynamic parameters2

2
All data are given in SI units; description is valid for 298.15 < T < 6000 K; excess
model is Redlich–Kister–Muggianu and the unary parameters are from SGTE [52].
104 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

Appendix 2. Complete thermodynamic database for the Al–Ta system


V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106 105

Appendix 2 (continued)
106 V.T. Witusiewicz et al. / Intermetallics 18 (2010) 92–106

References [29] Raman A. Aluminium 1965;41:318 [in German].


[30] Girgis K, Harnik AB. Naturwissenschaften 1970;57:242.
[31] Edshammar LE, Holmberg B. Acta Chem Scand 1960;4:1219.
[1] Schauer A, Roschy M, Juergens W. Thin Solid Films 1975;27:111.
[32] Nowotny H, Brukl C, Benesovsky F. Mh Chem 1961;92:116 [in German].
[2] Rai AK, Bhattacharya RS, Mendiratta MG, Subramanian PR. J Mater Res
[33] Du Y, Wenzel R, Schmid-Fetzer R. Calphad 1998;22:43.
1988;3:1082.
[34] Velikanova T, Turchanin M, Ilyenko S, Effenberg G. Calphad 2009;33:192.
[3] Schuhmacher B, Köster U. Mater Res Soc Symp Proc 1990;167:323.
[35] Schmidt StR, Franzen HF. J Less-Common Met 1986;116:73.
[4] Jiang SS, Zou J, Cockayne DJH, Sikorski A, Hu A, Peng RW. Philos Mag B
[36] Landolt-Börnstein – Group IV: Physical chemistry. Numerical data and func-
1992;66:229.
tional relationships in science and technology, New Series, vol. 12A, Supple-
[5] Chou Ch-Y, Chen S-W. J Electron Mater 2006;35:22.
ment to vol. 5A. Phase equilibria, crystallographic and thermodynamic data of
[6] Kofstad P, Espevik S. J Less-Common Met 1967;12:11.
binary alloys. Berlin–Heidelberg: Springer; 2006.
[7] Willens RH, Buehler E. Trans AIME 1966;236:171.
[37] Landolt-Börnstein – Group IV: Physical chemistry. Numerical data and func-
[8] Singh S, Lele S, Suryanarayana C. Metall Trans A 1987;18:1915.
tional relationships in science and technology, New Series, vol. 19B1. Binary
[9] EI-Eskandarany MS, Aoki K, Suzuki K. J Appl Phys 1992;72:2665.
systems. Part 1. Elements and binary systems from Ag–Al to Au–Tl, Berlin–
[10] Glazov VM, Mal’tsev MV, Chistyakov YD. Izv Akad Nauk SSSR, Otd Tekh Nauk
Heidelberg: Springer; 2002. p. 201.
1956;4:131 [in Russian].
[38] Yeremenko VN, Natanzon YV, Dybkov VI. J Less-Common Met 1976;50:29.
[11] Glazov VM, Lazarev GP, Korol’kov GA. Metalloved Term Obrab Met 1959;10:48
[39] Yatsenko SP, Skachkov MP. Russ J Phys Chem 1977;51:755 [in Russian].
[in Russian].
[40] Harbrecht B. Solution Range of the s-AlTa2 Phase, Internal Report, University
[12] Kaufman L. Calphad 1991;15:243.
of Bonn: Bonn, Germany;1996.
[13] Du Y, Schmid-Fetzer R. J Phase Equil 1996;17:311.
[41] Donnadieu P, Ochin P. J Alloys Compd 2007;434-435:255.
[14] Subramanian PR, Miracle DB, Mazdiyasni S. Metall Trans A 1990;21:539.
[42] Witusiewicz VT, Bondar AA, Hecht U, Rex S, Velikanova TYa. J Alloys Comp
[15] 3 vols.. In: Massalski TB, Subramanian PR, Okamoto H, Kasprzak L, editors. Binary
2008;465:64.
alloy phase diagrams. 2nd ed. Materials Park, OH: ASM International; 1990
[43] Witusiewicz VT, Bondar AA, Hecht U, Rex S, Velikanova TYa. J Alloys Compd
[16] ASM handbook. Alloy phase diagrams, vol. 3. Materials Park, OH: ASM Inter-
2009;472:133.
national; 1992.
[44] Wilhelm HA, Witte JH. Aluminum–tantalum alloy studies, Report No. IS-500,
[17] Mahne S, Krumeich F, Harbrecht B. J Alloys Compd 1993;201:167.
M48 U.S. Atomic Energy Commission; 1962.
[18] Mahne S, Harbrecht B. J Alloys Compd 1994;203:271.
[45] Weaver ML, Kaufman MJ. Scr Metall Mater 1992;26:411.
[19] Mahne S, Harbrecht B, Krumeich F. J Alloys Compd 1995;218:177.
[46] Palm M, Sanders W, Sauthoff G. Z Metallkd 1996;87:390.
[20] Schuster JC. Z Metallkd 1985;76:724–7.
[47] Wilson CG, Spooner EJ. J Mater Sci 1977;12:1653.
[21] Neckel A, Nowotny H. The thermochemistry of aluminides, 5. Int Leichtme-
[48] Kocherzhinsky JuA. Third ICTA, Davos. Therm Anal Proc 1971;1:549.
talltagung Leoben 1969;95:72 [in German].
[49] Kocherzhinsky JuA, Shishkin YeA, Vasilenko VI. In: Ageev NV, Ivanov OS, editors.
[22] Meschel SV, Kleppa OJ. J Alloys Compd 1993;197:75.
Phase diagrams of metal systems. Moskow: Nauka; 1971. p. 245 [in Russian].
[23] Kimura H, Nakano O, Ohkoshi T. Keikhzzoku 1973;23:106 [in Japanese].
  R. J Solid State Chem 2006;179:3385. [50] Pirani M, Alterthum H. Z Elektrochem 1923;29:5.
[24] Boulineau A, Joubert J-M, Cern y
[51] Yeremenko VN, Listovnichiy VYe. Teplofiz Vys Temp 1965;3:234 [in Russian].
[25] Brauer G. Naturwissenschaften 1938;26:710 [in German].
[52] Dinsdale AT. Calphad 1991;15:317.
[26] Brauer G. Z Anorg Allg Chem 1939;242:1 [in German].
[53] Sundman B, Jansson B, Andersson J-O. Calphad 1985;9:153.
[27] Villars P, Calvert LD. Pearson’s handbook of crystallographic data for inter-
[54] Sudavtsova VS, Batalin GI. Ukr Khim Zh 1989;55:144 [in Russian].
metallic phases. Materials Park, OH: ASM International; 1991.
[55] Nikolaev GI, Bodrov NV. Russ J Phys Chem 1978;52:821 [in Russian].
[28] Miracle DB, Scheltens F, Subramanian PR. Philos Mag B 1995;71:94.

You might also like