You are on page 1of 12

Mater. Res. Soc. Symp. Proc. Vol.

1128 © 2009 Materials Research Society 1128-U02-01

An Overview of the Mechanical Properties of FeAl

I. Baker
Thayer School of Engineering, Dartmouth College, Hanover, NH 03755

ABSTRACT

This paper presents an overview of the mechanical properties of the iron aluminide FeAl.
Both the strength and ductility depend on a variety of parameters, including vacancy
concentration, the environment (principally water vapor), alloy stoichiometry, temperature and
grain size. The effects of alloying elements, particularly boron, are also briefly discussed. The
review emphasizes how much of our current understanding of the strength and ductility of iron
aluminides emanated from the discovery of the importance of both water vapor and thermal
vacancies on these properties.

INTRODUCTION

Iron aluminides, i.e. Fe3Al, FeAl, FeAl2 Fe2Al5 and FeAl3, have been of interest for over
75 years [1]. This review focuses on B2 or ordered body-centered cubic compound FeAl, see
Figure 1. The compound is appealing because of its outstanding oxidation resistance, good
corrosion resistance, good strength up to ~700K and low density (~6000 kg m-3). Difficulties
with this material finding commercial application include its lack of strength at high temperature,
where it still shows excellent corrosion and oxidation resistance, and its poor low-temperature
ductility, particularly with increasing aluminum content.
This short review outlines the effects of a number of parameters, i.e. vacancy
concentration, the environment, alloy stoichiometry, temperature, grain size and alloying
elements, on the mechanical properties of FeAl. Additional details can be found in a number of
other reviews [2-14]. Our understanding of the effects of these different parameters was only
possible once the rôles of the environment and of vacancies on ductility and strength were
understood.

CRYSTAL STRUCTURE AND PHASE DIAGRAM

FeAl exists over a wide range of iron-rich compositions at low temperature, e.g. at 473K
from ~36.5-49.5 at. % Al, whereas aluminum-rich compositions exist only at elevated
temperature, see Figure 2. Hyperstoichiometric compositions of the contiguous D03-structured
compound Fe3Al disorder to the B2 structure at elevated temperatures that decrease with
increasing aluminum content. Up to ~46 at. % Al, FeAl itself disorders to b.c.c. at increasing
temperature with increasing aluminum content; binary alloys with greater aluminum
concentrations appear to be ordered up to melting.
The Young's modulus, E, of polycrystalline FeAl is not strongly dependent on
composition, with values of typically 250 GPa for 40-49 Al [15], and decreases linearly with
increasing temperature. As measured by Zener's parameter, A = 2 C44/(C11-C12), where Cij are
the elastic stiffness constants, FeAl is moderately anisotropic: A decreases slightly with
increasing Al concentration from 4.4 at Fe-35Al to 3.8 for Fe-40Al [16].
<1 0 0 >

Al
<1 1 1 >

Fe

Figure 1. The B2 structure adopted by


FeAl, showing the <111> and <100>
slip vectors.

Figure 2. Iron-rich portion of the Fe-Al


phase diagram. [17]

POINT DEFECTS

Anti-site atoms accommodate deviations from the stoichiometric composition in FeAl. In


addition to anti-site Fe atoms, so-called ‘triple defects’ are also present, which consist of an Fe
atom on the Al sublattice and two vacancies on the Fe sublattice. The concentration of these
triple defects increases with increasing Al concentration.
The measured activation enthalpy for vacancy formation (Ef) is quite low, 91-74 kJ mol-1
for 38.5-47 at. % Al, and the activation entropy is 6kB, where kB is Boltzmann’s constant [18,19].
Thus, unusually high thermal-vacancy concentrations are present in FeAl at elevated
temperatures, e.g., 2.5% in stoichiometric FeAl at 1000°C [20].
The vacancies that form at high temperatures are easily retained upon cooling due to the
high activation enthalpy for vacancy migration (Em) which ranges from 152-169 kJ mol-1 for
38.5-47 at. % Al [18,19]. In FeAl, therefore, Em ≈ 2Ef, unlike most metals where Em < Ef. Note
also that with increasing Al concentration, Ef decreases, i.e., the vacancy concentration
increases), whereas Em increases. Thus, it becomes increasingly difficult to anneal out the
increasing numbers of thermal vacancies that are retained as the Al concentration increases.

DISLOCATIONS

Early data suggested that a transition from <111> to <100> slip occurs with increasing
temperature [21,22]. However, later work showed that while <100> slip and, possibly, [110]
slip, may occur at larger strains at elevated temperatures, yielding is due to the multiplication and
glide of <111> dislocations, dissociated into anti-phase boundary (APB) coupled ½<111>
partials, at all temperatures [23,24]. The <100> dislocations form by the decomposition of
<111> dislocations since they have the lowest self-energy [25].
Low temperature anneals (used to remove retained vacancies) produce: numerous faults;
<111> dislocations, often as helices, after short anneals; and <100> dislocations loops after long
anneals. These observations indicate that <111> dislocations are easy to nucleate by vacancy
condensation but eventually decompose into the lower self-energy <100> dislocations [26].

STRENGTH

As outlined below, several factors influence the strength of FeAl. However, vacancies
play a major impact on the strength, and without understanding their role, it is impossible to
understand the effects of other parameters, a problem that bedeviled some early studies. Water
vapor also affects on the strength, an effect that has been less studied than that of vacancies.

Effects of Vacancies

It is not possible to understand the effects of various parameters on the strength without
removing the excess vacancies that are readily retained after elevated temperature anneals [27].
Even slow cooling rates retain excess thermal vacancies, a phenomenon arising because of the
high Em vacancy migration. In order to remove most of these vacancies, many workers have
adopted the 5 day anneal at 673K used by Nagpal and Baker [27].
Chang et al. [20] showed that the microhardness of binary FeAl alloys, irrespective of
aluminum concentration, is linearly related to the square root of the quenched-in thermal vacancy
concentration, a relationship that can be rationalized by vacancy pinning of dislocations on the
slip plane, which causes them to bow out. Later, Yang and Baker [28] showed that the room
temperature yield strength increased with increasing thermal vacancy concentration for both
polycrystals and single crystals of Fe-40Al at first rapidly and then more slowly as the
concentration increased. Thermal vacancy concentrations in FeAl can be very large and a
vacancy concentration of 9 x 10-3 increases the room temperature strength of Fe-40Al single
crystals from ~125 MPa to ~550 MPa.

Effects of Environment

Liu, Lee, and McKamey [29] first noted that water vapor, by producing atomic hydrogen
through the reaction 3H2O + 2Al → Α2O3 + 6H, has a profound effect on the mechanical
properties of iron aluminides around room temperature. It has been shown to result in both a
lower yield stress and lower subsequent flow stress in both FeAl single crystals and polycrystals
when tested in air compared to tests performed in vacuum as long as the strain rate is less than
1 s-1 [30-32], see Figure 3. Conversely, at higher strain rates the effects of the water
vapor/atomic hydrogen are not present, see Figure 3. Details of the mechanism for this strength
reduction are not clear, but it is worth noting that most studies of the effects of grain size, alloy
stoichiometry, alloying, etc on the strength of iron aluminides at room temperature have not
considered this effect. Thus, this “hydrogen” or “water vapor” effect is probably convoluted
with the effects of other parameters on the yield strength and subsequent flow strength.
450
vacuum

400 oxygen Polycrystals

Yield Strength (MPa)


air
350

300

250
Single Crystals
200 vacuum air

150
-7 -4 -1 2
10 10 -1 10 10
Strain Rate (s )
Figure 3. Yield strength as a function of strain rates for polycrystals [30] and single crystals [31]
of FeAl in air and vacuum.

Effects of Alloy Stoichiometry

Ideally, in order to analyze the effects of alloy stoichiometry on the strength of FeAl, the
critical resolved shear strength determined from single crystals of different compositions would
be compared. However, while there have been mechanical tests on FeAl single crystals, these
are insufficient to illustrate how the strength changes with aluminum content. Another way to
establish the effect of aluminum content on the strength is to scrutinize the lattice resistance or
lattice friction term, σo, in the Hall-Petch Relationship relating the yield strength, σy, to the grain
size, d, i.e. σy = σo + kyd -1/2, as a function of aluminum content. (Testing very large grained
polycrystals is another way, but care has to be taken that there are no textural differences
between specimens.) σo shows a minimum at 45 at . % Al, and increases much more rapidly on
the aluminum rich side of the minimum reflecting that hardening from anti-site atoms (iron rich
side) is much less than hardening from vacancies (aluminum rich side) [10], see Figure 4.

1000 1.8
Hall-Petch slope, GPa. µm- 0.5
Lattice Resistance, MPa

800 1.6
1.4
600
1.2
400 1.0

0.8
200
0.6
0 0.4
30 35 40 45 50 55
at. % aluminum
Figure 4. Lattice resistance and Hall-Petch slope for FeAl as a function of at. % Al [33].
This minimum in strength around Fe-45Al is observed at both room temperature and
77K, but the difference between the minimum at Fe-45Al and the maximum strength at the
stoichiometric composition is less at 77K [8], see Figure 5.
In marked opposition to the low temperature behavior, the high-temperature creep
strength of FeAl has been found to be independent of aluminum concentration [34].

Figure 5. Yield strength as a function of Al concentration of FeAl annealed 5 days at 673K [8].

Effects of Grain Size

Grain size strengthening in FeAl, as measured by the ky in the Hall-Petch relationship, is


only weakly dependent on composition at room temperature for aluminum contents up to 45 at.
% Al, but ky for the stoichiometric FeAl is twice that of iron rich compositions, see Figure 4,
indicating the greater difficulty of slip transmission at the stoichiometric composition. Boron has
been shown to increase ky for both Fe-40Al and Fe-45Al [35].

Effects of Temperature

B2 compounds typically show a yield strength that decreases rapidly with increasing
temperature up to room temperature, followed by either a plateau or a peak (the so-called yield
anomaly, YSA), before declining rapidly at temperatures above half the homologous melting
point, see Figure 6. The YSA in many B2 compounds arises from dynamic strain ageing (DSA).
However, while a YSA is observed at conventional strain rates (10-4–10-3 s-1) in iron-rich FeAl,
the negative strain rate sensitivity is not seen [8,10]. For stoichiometric FeAl, the yield peak is
only observed when straining is performed at higher rates (≥1 s-1) [36], and becomes pronounced
at very high strain rates (~2 x 103 s-1) [37].
The YSA is obscured if quenched-in thermal vacancies are not eliminated prior to testing,
see Figure 7 and disappears at very slow strain rates even in iron-rich FeAl [10].
A number of mechanisms have been proposed for the YSA: the <111> to <100> slip
vector transition, cross-slip/pinning, APB relaxation, local dislocation climb-lock, dislocation
decomposition, dislocation core dissociation, dislocation constriction/pinning and vacancy
hardening (see, e.g., [10,12]). Of these only the vacancy hardening mechanism has been
formulated as a theory [38]: the idea is that immediately below the yield strength peak vacancies
are immobile and impede dislocation motion, while above the peak the vacancies are able to
migrate and aid dislocation climb. Although individual vacancies cannot impede dislocation
flow in a metal, in FeAl, a vacancy can cause the leading a/2<111> APB-coupled dislocation
partial to climb. Since the trailing a/2<111> partial does not climb, APB created by this climbed
segment is not removed by the trailing partial, causing a drag on the dislocation.

Figure 6. Yield strength versus temperature for large-grained, low-temperature annealed FeAl of
various compositions strained under tension at a strain rate of 1 x 10-4 s-1 [8].

600

high vacancy
500
Yield Stress, MPa

400

300
low vacancy

200

100
0 200 400 600 800 1000 1200
Temperature, K

Figure 7. Yield Stress as a function of temperature for Fe-43Al-0.12B annealed for 1.25 hr at
1373 K and 2 hr at 973 K, with and without long term anneal of 120 hours at 673 K, which
produce low and high vacancy concentrations respectively [39].
Using the approximation that the increase in strength from the vacancies is proportional
to (vacancy concentration)1/2 [20] and using the fact that the vacancy concentration increases
exponentially with increasing temperature, the strength increase with increasing temperature up
to the yield strength peak is proportional to exp[-Ef/2kT], where Ef is the activation enthalpy for
vacancy formation, k is Boltzmann's constant and T is the absolute temperature [38]. Above the
peak, the strength can be described by a diffusion-assisted deformation mechanism. Fitting
experimental data for polycrystalline Fe-40Al to the model yields a value for Ef of 92 kJ/mol,
which is close to the measured value. A reasonable value of the creep exponent, m (3.8), is
obtained in the dislocation creep regime and the strength increase is reasonable for the vacancy
concentrations present [38]. The model [38] allowed a number of predictions to be made, viz.,
• The yield strength below the peak is independent of strain rate whereas above the peak, in the
dislocation creep regime, the yield strength is strain rate dependent. It follows that the yield
strength peak should itself be strain rate dependent, the peak moving to higher stresses and
higher temperatures with increasing strain rate and to lower stresses and lower temperatures
with decreasing strain rate.
• Following the above argument, at very low strain rates, the yield peak should not be present.
• The model indicates that for short times, the time at the peak yield stress temperature prior to
testing affects the magnitude of peak yield stress. This effect arises because vacancy formation
is not instantaneous but requires time, with increasing times (which result in greater vacancy
concentrations) leading to larger yield stresses.
• No orientation dependence of the yield peak is expected.
Not only have these predictions have been borne out by experiments [see references 8,
and 40 for details], but also some of the experimental observations exclude dynamic dislocation-
based mechanisms, e.g. climb decomposition, cross-slip pinning, glide decomposition, as
possible causes for the YSA [13,41]. Morris and Morris-Muñoz [41], on examining available
data, suggested that the vacancy-hardening model may also explain the YSA in Fe3Al.
FeAl shows a yield stress tension/compression asymmetry at all temperatures, which may
be related to the effect of hydrostatic pressure on the vacancy concentration or the effect of non-
deviatoric stresses on the dislocation core structure. The latter effect could be the reason that the
critical resolved shear stress of FeAl single crystals does not obey Schmid’s law [10].

Alloying

As might be expected, both interstitials, such as boron, and substitutional elements


increase the strength of FeAl both at room temperature and at elevated temperature with the
substitutional strengthening being related to the misfit of the solute atoms, see reference 10 for
details. Care has to be taken when comparing the effects of ternary additions since in addition to
intrinsic strengthening ternary additions can slow the rate of vacancy removal during low
temperature annealing so that vacancy hardening is actually being measured.
Auger electron spectroscopy studies have demonstrated that boron segregates to the grain
boundaries and increases grain boundary strength [42]. However, some B remains in the lattice,
inducing a lattice strain, and produces strengthening at room temperature that depends on both
the aluminum concentration and the presence of vacancies [43]. Boron also shifts the yield
strength peak to higher temperatures.
DUCTILITY

As for the strength, there are several factors that influence the ductility of FeAl. It should
be noted though that the presence of water vapor has a marked effect on the ductility of FeAl,
with vacancies playing a secondary deleterious role on the ductility. Without taking into account
the role of these two parameters, the effects of other variables cannot really be understood.

Effects of Environment

The effects of the water vapor during tensile testing of iron-rich FeAl at slow to medium
strain rates in air are so dominant that the true effects of other metallurgical variables are largely
obscured. As noted above, Liu, Lee, and McKamey [29] first observed the effects of water vapor
on FeAl and showed that polycrystalline Fe-36.5Al exhibited higher elongation in vacuum
(5.5%) than in air (2%), and even greater elongation in dry oxygen (18%). Transgranular
cleavage occurred in the specimens tested in air, while mixed mode failure occurred in vacuum
and intergranular fracture occurred in oxygen. Their suggested mechanism of atomic hydrogen
produced from water vapor diffusing to crack tips and aiding their advance was later supported
by the work of Gleason et al. [44] who studied the reaction of water vapor at the surface of FeAl,
and observed hydrogen formation occurring there together with oxidation of aluminum. Li and
Thus, factors, such as temperature, strain rate, and protective atmosphere, that affect the
kinetics or energetics of the individual steps in the embrittlement of FeAl, such as the
dissociative adsorption of H2O, movement of H into the crack-tip region, etc. also affect its
ductility.
The environmental and strain-rate dependence of the ductility in air has been observed by
a number of workers on several different compositions of single crystals and polycrystals.
Perhaps, the most dramatic effects were observed in tensile tests on single-slip-oriented Fe-43Al
single crystals [32] where crystals strained at 2.5 x 10-3 s-1 exhibited 46% elongation in oxygen,
but only 8-12% elongation in air, while only ~3% elongation was obtained at a slow strain rate of
4.4 x 10-6 s-1 in air. The fracture mode of the crystals tested in air changed from cleavage with
numerous secondary cracks at the slow strain rate to a rougher fracture surface with no
secondary cracks and some ductile dimples at high strain rate. The latter fracture mode was the
same as that for crystals strained in oxygen.
It has been suggested [45] that atomic hydrogen segregation stabilizes the sessile <100>
dislocations that form by the collision of gliding <111> dislocations [46] thereby promoting
cleavage crack nucleation on {100}.

Effects of Vacancies

While the environment is the major cause of brittleness in iron-rich FeAl, vacancies also
have a marked effect on ductility. Yang et al. [47] showed that the room-temperature fracture
strengths of Fe-40Al single crystals in both air and vacuum are significantly increased due to
vacancy hardening, but introducing a high vacancy concentration decreases elongations. The
fracture mode was also found to be affected by the vacancy concentration: with a low vacancy
concentration, fracture was by cleavage on both {100} and a variety of other planes, whereas at
high vacancy concentration, cleavage occurs mainly on {100} and {110}. The latter fracture
mode suggests that vacancies promote fracture along the slip planes by condensing on the {110}
slip plane and forming crack nuclei.

Effects of Alloy Stoichiometry

There have been a number of studies of the effects of alloy stoichiometry on the ductility
of FeAl. Probably the cleanest study is that by Liu and George [48] who performed tensile tests
on polycrystalline binary FeAl under ultrahigh vacuum. The ductility depends strongly on Al
content with high ductility and transgranular fracture at low Al contents, but a rapid drop in
ductility and change to intergranular fracture with increasing Al, see Figure 8. The change to
intergranular fracture as the aluminum concentration increases indicates the grain boundary
weakening. In fact, room temperature intergranular fracture in near-stoichiometric compositions
shows little evidence of plastic flow, failure occurring before yield under tension.

Figure 8. Effect of Al concentration and B doping (300 wppm) on the ‘intrinsic’ ductility and
fracture behavior of FeAl in an ultrahigh vacuum of ~10-10 torr [49].

Alloying

Schneibel et al. [50] studied the room temperature tensile elongation and fracture mode of
Fe-45Al containing 5% of each of the first row transition metals, all of which decreased the
ductility, irrespective of testing environment. Alloys with ternary additions whose atomic
numbers are less than iron exhibited intergranular fracture, whereas additions whose atomic
numbers are greater than iron exhibited transgranular fracture, probably reflecting the site
occupancy of the ternary atoms. Liu and George [48] performed tensile tests on polycrystalline
binary FeAl under ultrahigh vacuum with added B which segregates to the grain boundaries and
suppresses intergranular fracture thereby shifting the ductile-brittle transition observed in
unalloyed FeAl to higher Al levels, see Figure 8. Boron strengthening of grain boundaries in air
is much less effective [13]. Several groups subsequently demonstrated boron strengthening of
grain boundaries in FeAl.
Effects of Grain Size

The beneficial effects of refining the grain size have been observed during tensile testing
of FeAl at high strain rate, ~1 s-1, where the ductility is unaffected by the environment [30].
When tensile tested at ~1 s-1 in air at room temperature Fe-45Al shows an increase in elongation
to fracture from ~6% to ~9% as the grain size is reduced from 80 µm to 28 µm, whereas at strain
rates less than ~1 x 10-2 s-1 there is no difference in ductility for different grain sizes [51], see
Figure 8. The effects of grain size are more dramatic in boron-doped material, where the
elongation to failure increases from ~7% in large-grained Fe-45Al to ~15% in fine grained
material tested in air at 1 s-1, see Figure 9. Again, tests at slow strain rate show no differences.

16
Fe-45Al +B (230 µm)
Fe-45Al +B (50 µm)
12 Fe-45Al (28 µm)
Fe-45Al (80 µm)
Elongation (%)

0
10 -7 10 -6 10 -5 10 -4 10 -3 10 -2 10 -1 10 0 10 1 10 2 10 3

Strain Rate (s -1 )
Figure 9. Elongation versus strain rate for both boron-doped and un-doped low-temperature-
annealed Fe-45Al of different grain sizes tensile tested in air. After reference 50.

Effects of Temperature

For all binary compositions, the elongation of FeAl increases moderately from room
temperature to 500 K, and then increases sharply with further increases in temperature,
accompanied by considerable necking [52]. The reduction in area at the neck shows a sharp
transition from <10% below 500 K, where deformation is homogeneous, to >80% above 600 K,
with reductions in area >95% at 900 K. Compositions that at room temperature show wholly or
partly intergranular fracture exhibit (more) cleavage fracture at elevated temperature. Small
cleavage lips are observable on intergranular facets. At temperatures above 650 K, fracture is by
ductile dimpled rupture, accompanied by grain boundary migration (above 800 K) and dynamic
recrystallization (>900K). Excess quenched-in vacancies decrease the ductility at temperatures
up to about 800K and solutes, such as boron, cause a dramatic ductility drop at high temperature,
accompanied by intergranular fracture and grain boundary cavitation, see reference 10.
STRAIN-INDUCED FERROMAGNETISM

The phenomenon of strain-induced ferromagnetism in heavily-deformed intermetallic


compounds that contain ferromagnetic elements and are fully disordered has been known for at
least forty years and is well explained using the local environment model, which considers the
number of like nearest-neighbor atoms around potentially ferromagnetic atoms in the compound.
Intermetallic compounds, such as FeAl, that are subject to lighter deformation that does not
completely disorder them can also show strain-induced ferromagnetism. This behavior has been
modeled in FeAl by applying the local environment model to the disordered region inside APB
tubes [53]. That the strain-induced ferromagnetism arises from APB tubes is consistent with the
observations that:
• At temperatures where deformation produces strain-induced ferromagnetism, APB tubes are
also present [54];
• APB tubes observed after room-temperature deformation can be annealed at temperatures of
~400 K, at which point the strain-induced ferromagnetism also disappears [55].

ACKNOWLEDGEMENTS

This research was supported by Award DE-FG02-07ER46392 from the Division of Materials
Sciences, U.S. Department of Energy and grant DMR 0552380 from the National Science
Foundation with Dartmouth College. Any opinions, findings, and conclusions or
recommendations expressed in this material are those of the author(s) and do not necessarily
reflect the views of the NSF, DOE or the U.S. Government.

REFERENCES

1. N. Ziegler, Trans AIME 100, 267 (1932).


2. I. Baker, P. Nagpal, in Processing and Fabrication of Advanced Materials for High
Temperature Applications - II, ed.- T.S. Srivistan and V.A. Ravi (TMS, Warrendale, PA,
1993) p. 3-18.
3. I. Baker, P. Nagpal, in Structural Intermetallics, edited by R. Darolia et al. (TMS,
Warrendale, PA, 1993) p. 463.
4. I. Baker, in Processing, Properties and Applications of Iron Aluminides, ed. - J.H. Schneibel
and M.A. Crimp (TMS, Warrendale, PA, 1994) p. 101.
5. N.S. Stoloff, C.T. Liu, Intermetallics 2, 75 (1994).
6. D. Hardwick, G. Wallwork, Rev. High Temp. Mater. 4, 47 (1978).
7. X. Pierron, I. Baker, in Design Fundamentals of High Temp. Composites, Intermetallics and
Metal-Ceramic Systems, ed. R.Y. Lin et al. (TMS, Warrendale, PA, 1996) p. 271.
8. E.P. George, I. Baker, in The Encyclopedia of Materials: Science and Technology, ed.-
K.H.J. Buschow et al. (Elsevier Press, Pergamon, 2001) pp. 4201.
9. U. Prakash, R.A. Buckley, H. Jones, C.M. Sellars, ISIJ Int. 31, 1112 (1991).
10. I. Baker, P.R. Munroe, Int. Mat. Rev. 42, 181 (1997).
11. D.G. Morris, M.A. Morris-Muñoz, Intermetallics 7, 1121 (1999)
12. I. Baker, E.P. George, Proc. MRS 552, KK4.1.1 (1999).
13. C.T. Liu, E.P. George, P.J. Maziasz, J.H. Schneibel, Mater. Sci. Eng. A. 258, 84 (1998).
14. I. Baker, D. Wu, M. Wittmann, E.P. George, in Proc. 4th Pacific Rim Int. Conf. on Advanced
Materials and Processing, Vol. I, (Jap. Inst. Mater., Sendai, Japan, 2001) p. 811.
15. M.R. Harmouche, A. Wolfenden, Mater. Sci. Eng. 84, 35 (1986).
16. H.J. Leamy, E.D. Gibson, F.X. Kayser, Acta Metall. 15, 1827 (1967).
17. Binary Alloy Phase Diagrams, ed - T.B. Massalski (ASM, Metals Park, OH, 1986), p. 112.
18. J.P. Riviere, J. Grihlé, Scripta Metall. 9, 967 (1975).
19. R. Wurschum, C. Grupp, H.-E. Schaefer, Phys. Rev. Letters, 75, 97 (1995).
20. Y.A. Chang, L.M. Pike, C.T. Liu, A.R. Bilbery, D.S. Stone, Intermetallics 1, 107 (1993).
21. M.G. Mendiratta, H.K. Kim, H.A. Lipsitt, Metall. Trans. A, 15A, 395 (1984).
22. I. Baker, D.J. Gaydosh, Mater. Sci. Eng. 96, 147 (1987).
23. B. Kad, J.A. Horton, Mater. Sci. Eng. 118, 239 (1997).
24. K. Yoshimi, S. Hanada, M.H. Yoo, Proc. MRS, 460, 313 (1997)
25. M.G. Mendiratta, C.C. Law, J. Mater. Sci. 22, 607 (1987).
26. M.A. Morris, O. George, D.G. Morris, Mater. Sci. Eng. A258, 99 (1998).
27. P. Nagpal, I. Baker, Metall. Trans. 21A, 2281 (1990).
28. Y. Yang, I. Baker, Intermetallics 6, 167 (1998).
29. C.T. Liu, E.H. Lee, C.G. McKamey, Scripta Metall. 23, 875 (1989).
30. L.M. Pike, C.T. Liu, Scripta Mater. 38, 1475 (1998).
31. D. Wu, I. Baker, Intermetallics 9, 57 (2001).
32. I. Baker, D. Wu, S.O. Kruijver, E.P. George, Mat. Sci. Eng. A. 329-331, 726 (2002).
33. I. Baker, P. Nagpal, F. Liu, P.R. Munroe, Acta Metall. Mat. 39, 1637 (1991).
34. J. D. Whittenberger: Mater. Sci. Eng. 57, 77 (1983); 77, 103 (1986).
35. L.M. Pike, C.T. Liu, Scripta Mater. 25, 2757 (1991).
36. I. Baker, Y. Yang, Mater. Sci. Eng. A239-240, 109 (1997).
37. Y. Yang, I. Baker, R.T. Gray, III, C. Cady, Scripta Mater. 40, 403 (1999).
38. E.P. George, I. Baker, Philos. Mag. 77, 737 (1998).
39. R. Carleton, E.P. George, R.H. Zee, Intermetallics 3, 433 (1995).
40. D. Wu, I. Baker, P.R. Munroe, E.P. George, Intermetallics 1, 103 (2007).
41. D.G. Morris, M.A. Morris-Muñoz, Intermetallics 13, 1269 (2005).
42. C.T. Liu, E. P. George, Scripta Metall. Mater. 24, 1285 (1990).
43. I. Baker, X. Li, H. Xiao, R. L. Carleton, E. P. George, Intermetallics 6, 177 (1998).
44. N.R. Gleason, C.A. Gerken, D.R. Strongin, Appl. Surf. Sci. 72, 215 (1993).
45. J.C.M. Li, C.T. Liu, Scripta Metall. Mater. 33, 661 (1995).
46. P.R. Munroe, I. Baker, Acta Metall. Mater. 39, 1011 (1991).
47. Y. Yang, I. Baker, E.P. George, Mater. Char. 42, 161 (1999).
48. C.T. Liu, E.P. George, Scripta Metall. Mater. 24, 1285 (1990).
49. J.W. Cohron, Y. Lin, R.H. Zee, E.P. George, Acta Mater. 46, 6245 (1998).
50. J.H. Schneibel, E.P. George, I.M. Anderson, Intermetallics 5, 185 (1997).
51. I. Baker, O. Klein, C. Nelson, E.P. George, Scripta Metall. Mater. 30, 863 (1994).
52. I. Baker, H. Xiao, O. Klein, C. Nelson, J.D. Whittenberger, Acta Metall. Mat. 43, 1723
(1995).
53. D. Wu, P.R. Munroe, I. Baker, Philos. Mag. 83, 295 (2003).
54. K. Yamashita, M. Imai, M. Matsuno, A. Sato, Philos. Mag. A 78, 285 (1998).
55. D. Wu, I. Baker, Mater. Sci. Eng., A, 329-331, 334 (2002).

You might also like