Вы находитесь на странице: 1из 13

Acta Materialia 166 (2019) 702e714

Contents lists available at ScienceDirect

Acta Materialia
journal homepage: www.elsevier.com/locate/actamat

Full length article

Interfacial free energies, nucleation, and precipitate morphologies in


Ni-Al-Cr alloys: Calculations and atom-probe tomographic
experiments
Zugang Mao a, Christopher Booth-Morrison a, Chantal K. Sudbrack a, b, Ronald D. Noebe b,
David N. Seidman a, c, *
a
Department of Materials Science and Engineering, Northwestern University, 2220 Campus Drive, Evanston, IL, 60208-3108, USA
b
Structures and Materials Division, NASA Glenn Research Center, 21000 Brookpark Road, Cleveland, OH, 44135, USA
c
Northwestern University Center for Atom-Probe Tomgraphy, 2220 Campus Drive, Evanston, IL, 60208-3108, USA

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

Article history: The effects of Cr additions on the morphologies οf g0 (L12 ordered) precipitates are studied in three Ni-Al-
Received 13 June 2018 Cr alloys utilizing atom-probe tomography (APT) and first-principles calculations. We find that Cr and Al
Received in revised form share the same sublattice sites in the L12-structure: Chromium partitions from the L12 structure into the
6 January 2019
disordered f.c.c. matrix for a driving force of 0.436 eV atom1. Mixing between the Ni3Al(L12) and
Accepted 13 January 2019
Available online 16 January 2019
Ni3Cr(L12) structures is energetically unfavorable based on first-principles quasi-random structure cal-
culations. Interfacial Gibbs free energies of Ni/Ni3Al(L12), Ni/Ni3Cr(L12), Ni/Ni3Cr(DO22), and Nin(AlyCr1-y)/
Ni3(AlxCr1-x)(L12) interfaces are calculated for the {100}, {110}, and {111} planes. The temperature de-
Keywords:
First-principles calculations
pendencies of the interfacial Gibbs free energies are determined by including phonon vibrational en-
Interfacial Gibbs free-energy tropies. The equilibrium interfacial Gibbs free energies are determined using concentration profiles as a
Phonon vibrational entropy function of aging time, obtained using APT. At early aging times, the initial values of the interfacial Gibbs
Phase partitioning free energies are higher with small Cr concentrations in the g’ (L12)-precipitates. With increasing aging
Atom-probe tomography time, the interfacial Gibbs free energies decrease with increasing Cr concentration and achieve equi-
librium values after 16-h of aging at 873 K (600  C) for the three Ni-Al-Cr alloys. From classical nucleation
theory, the nucleation of metastable Ni3Cr(L12) is easier than Ni3Al(L12) at 873 K (600  C). Ni3Cr(DO22) is
more difficult to nucleate than Ni3Cr(L12) and Ni3Al(L12) at 873 K (600  C). The Cr additions to Ni-Al
alloys have significant effects on the critical radius and critical net reversible work of nucleation. With
increasing Cr concentration, Ni3(Al,Cr)(L12) requires the smallest critical net reversible work to form but
has a larger critical size. An embryo of Ni3Cr(L12) only acts as a nucleant for mixed L12 Ni3(AlxCr1-x). The
morphology of Ni3(Al,Cr)(L12) precipitates transforms from cuboidal-to-spheroidal with increasing Cr
concentration, in agreement with APT observations.
© 2019 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

1. Introduction disordered nickel-rich g(f.c.c.)-matrix [3]. It is of great scientific


interest to understand the temporal evolution of the microstruc-
Ni-Al-Cr alloys are model alloys for Ni-based superalloys, which tures of the model alloys as an important step toward under-
are widely used in high-temperature aerospace jet engines and standing more complex commercial alloys. The g(f.c.c.)/g0 (L12)
land-based gas turbine engines, due to their superior strength, heterophase interfacial structure and concentrations play a key role
coarsening, creep, corrosion and oxidation resistance at elevated in determining the mechanisms of nucleation, growth, and coars-
temperatures [1,2]. The microstructures of high-strength Ni-based ening of the g0 (L12)-phase in a disordered g(f.c.c.)-matrix [4,5]. The
superalloys consist of coherent ordered g0 (L12)-precipitates in a g(f.c.c.)/g0 (L12) heterophase interfacial Gibbs free energy, s, for
different crystallographic orientations, affects strongly the
morphology of the ordered precipitates, which enters directly in all
* Corresponding author. Department of Materials Science and Engineering, coarsening theories used to model microstructural stability, and is
Northwestern University, 2220 Campus Drive, Evanston, IL, 60208-3108, USA. needed to predict the reliability of existing superalloys and to
E-mail address: d-seidman@northwestern.edu (D.N. Seidman).

https://doi.org/10.1016/j.actamat.2019.01.017
1359-6454/© 2019 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 703

further improve them. The values of the g(f.c.c.)/g0 (L12) hetero- conventional APT and from 0.04 to 0.08 ion pulse1 for 3-D-LEAP
phase interfacial Gibbs free energy in Ni-Al-Cr alloys are evaluated tomographic analyses. Pulsed-laser APT experiments were per-
experimentally using the Lifshitz-Slyozov-Wagner (LSW) coars- formed for alloy C with a LEAP4000X Si tomograph at a target
ening model and its variants, which yields average values of the evaporation rate of 0.04 ion per pulse, a specimen temperature of
interfacial Gibbs free energy for all the {hkl} facets of g0 (L12)-pre- 40.0 ± 0.3 K, a laser pulse energy of 0.6 nJ per pulse, a pulse repe-
cipitates [6e8]. There are no prior reports on theoretical calcula- tition rate of 200 kHz, and an ambient gauge pressure of
tions for the g(f.c.c.)/g0 (L12) heterophase interfacial Gibbs free <6.7  108 Pa, employing a green laser (wavelength ¼ 535 nm).
energies for specific {hkl} facets of precipitates in Ni-Al-Cr alloys. These evaporation conditions were optimized to provide the
The phonon vibrational entropy was calculated to determine the highest compositional accuracy for this alloy [23]. The detailed
g(f.c.c.)/g’(L12) heterophase interfacial Gibbs free energies at an experimental and analytical information, including data regarding
elevated temperature in Ni-Al alloys in our prior research [9]. the nanostructural properties of the g0 (L12)-precipitates; compar-
Previous experimental studies demonstrated that the morphology isons with classical nucleation theory, growth and coarsening
changes from cuboidal-to-spheroidal with the addition of Cr to Ni- models can be found elsewhere [6,7,24e30]. Concentration profiles
Al alloys [6e8,10e14]. from APT data discussed herein are generated utilizing the prox-
In this research, the morphological evolution of g0 (L12)-pre- imity historgram methodology [31,32] by averaging across the
cipitates in Ni-Al-Cr alloys as well as their compositional evolution g(f.c.c.)/g0 (L12)-heterophase interfaces of tens to hundreds of
as a function of aging time at 873 K were previously studied uti- g’(L12)-precipitates for the different aging conditions employed. All
lizing atom-probe tomography (APT) and the pertinent results are concentrations are reported in at.%, unless otherwise specified.
reviewed [6e8]. To understand theoretically the mechanism and its
effects on the morphological changes, we employ density func- 2.2. Computational details
tional theory (DFT) first-principles calculations to obtain physically
meaningful values of the interfacial free energies of the {100}, {110} In this study, the plane-wave total-energy methodology with
and {111} heterophase interfaces, accounting for the first time for the Perdew-Burke-Ernzerhof parameterization of the generalized
the effect of temperature on both coherency strain-energy and gradient approximation (GGA) is employed [33,34] for exchange-
phonon vibrational entropy. Magnetic spin-polarized calculations correlation in our density functional theory (DFT) first-principles
are performed to evaluate the effects of ferromagnetism, and calculations, as implemented in the Vienna ab initio simulation
phonon vibrational entropies are utilized to determine the tem- package (VASP) [35e39]. We use the projector augmented wave
perature dependencies of the g(f.c.c.)/g0 (L12) heterophase interfa- (PAW) potentials in our calculations [40]. All the structures
cial Gibbs free energies of the {100}, {110} and {111} interfaces. We considered are fully relaxed with respect to the atomic coordinates
calculated the chemical properties and elastic properties of three as well as the volume inside the supercell. We considered and
possible ordered phases in Ni-Al-Cr alloys: Ni3Al (L12), Al3Cr (L12), tested the convergence of results carefully with respect to a range
and Ni3Cr (DO22). The stabilities of these possible ordered phases of energy cutoffs from 200 to 700 eV and 4x4x4 to 16x16x16 k-
has been studied, and mixing between the two L12 ordered phases points. A plane-wave basis set with spin-polarized method was
[Ni3Al (L12) and Al3Cr (L12)] is demonstrated not to be energetically used with an energy cutoff of 400 eV to represent the Kohn-Sham
favored. The site preference and partitioning behavior of Cr atoms wave functions. The summation over the Brillouin zone for the
across g(f.c.c.)/g0 (L12) heterophase interfaces is considered in bulk structures is performed on an optimized 121212 spacing
detail. The predicted equilibrium morphologies of g’(L12)-pre- with a Monkhorst-pack k-point mesh per f.c.c unit cell for all
cipitates, based on Wulff constructions, are in excellent agreement calculations.
with experimental atom-probe tomography (APT) results. Classical We calculated the phonon vibrational entropies to determine
nucleation theory is utilized to determine the critical radius and net the temperature dependencies of interfacial Gibbs free energies
reversible work for all possible nuclei: Ni3Al (L12), Al3Cr (L12), Ni3Cr and nucleation energies of L12 and DO22 ordered precipitates in the
(DO22), and Ni3(Al,Cr)(L12). Ni-Al-Cr alloys. The harmonic vibrational frequency approximation
was used for all vibrational calculations. To determine the phonon
2. Methodology dispersions for all the related structures, either for ordered struc-
tures or the disordered solid-solution phase, we utilized the frozen-
2.1. Atom-probe tomography phonon methodology [41e43] to compute the vibrational ther-
modynamics for ordered Ni3Al (L12), Ni3Cr (L12), Ni3Cr (DO22),
The temporal evolution of g0 (L12)-precipitates was studied Ni3(AlxCr1-x)(L12) and solid-solutions, and for Ni (f.c.c.)/Ni3Al (L12),
experimentally in three different Ni-Al-Cr alloys using APT. The Ni (f.c.c.)/Ni3Cr(L12), Ni(f.c.c.)/Ni3Cr (DO22), and Ni (f.c.c.)/
compositions in at.% of the three alloys studied are: Ni-7.5Al-8.5Cr Ni3(AlxCr1-x) (L12) interfaces. We calculated the vibrational free
(alloy A); Ni-5.2Al-14.2Cr (alloy B); and Ni-6.5Al-9.5Cr (alloy C). energies of these supercells from the phonon density of states
Solution-treated ingot-sections of alloy samples were aged at 873 K (DOSs) using standard thermodynamic expressions.
(600  C) under flowing argon for times ranging from 1/4 to 1024 h. For the solid-solution Nin(Al,Cr), we utilized a 64-atom cubic cell
We performed voltage-pulsed APT on microtip specimens utilizing (2  2  4 conventional f.c.c.) of Al with a Cr atom replacing a
a conventional APT [15,16], and additionally a Cameca Instruments portion of the Ni atoms on the central sites. For Ni3Al (L12), Ni3Cr
(formerly Imago Scientific Instruments, Madison, WI) picosecond (L12), and Ni3(AlxCr1-x) phases, we employed supercells of 3  3  3
green laser assisted local-electrode atom-probe (LEAP) tomograph cubic units cells, comprising 108 total atoms. For the Ni3Cr (DO22)
[17e22] in the Northwestern University Center for Atom-Probe phase, we used supercells with 2  2  4 units cells, comprising 64
Tomography (NUCAPT). Pulsed-voltage APT was performed for al- total atoms.
loys A and B at a specimen temperature of 40.0 ± 0.3 K, a voltage
pulse-fraction (pulse voltage/steady-state dc voltage) of 19%, a 3. Results and discussion
pulse repetition rate of 1.5 kHz (a conventional APT) or a 200 kHz
(LEAP4000X Si tomograph), and at a background gauge pressure of 3.1. Atom-probe tomography
<6.7  108 Pa (<5  1010 Torr). The average detection rates in the
areas of analysis ranged from 0.011 to 0.015 ion pulse1 for Three different Ni-Al-Cr alloys are studied: Ni-7.5Al-8.5Cr (alloy
704 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

A); Ni-5.2Al-14.2Cr (alloy B); and Ni-6.5Al-9.5Cr (alloy C). 3.2. Bulk chemical properties and elastic properties of the possible
Nanometer-sized g0 (L12)-precipitates are experimentally observed ordered phases in Ni-Al-Cr alloys
and studied over the full range of aging times from 1/6 to 1024 h at
873 K (600  C). The temporal evolution of the morphology of the In this article, we present both experimental determinations
precipitate phase in the three alloys is discussed in great detail in and calculations of the interfacial Gibbs free-energies, nucleation
our prior research [6e8,24,29]. The compositions of the g(f.c.c.)- barriers, and morphologies of precipitates in three Ni-Al-Cr alloys.
matrix and the g0 (L12)-precipitate phases of these alloys evolve We calculated the energetic and thermodynamic properties of the
temporally as the g(f.c.c.)-matrix becomes enriched in Ni and Cr bulk phases, as well as the interfacial Gibbs free energies. We first
and depleted in Al, while the g0 (L12)-precipitates become enriched discuss the first-principles results for the bulk properties, including
in Al and depleted in Cr as function of aging time, Tables 1e3. formation energies of the ordered structures, lattice parameters,
During nucleation, solute-rich g0 (L12)-nuclei form with large values elastic constants and strain energies, and discuss subsequently the
of the Al and Cr supersaturations, which decrease with increasing calculated interfacial energies. From these results, we obtain a more
aging time. The first g’(L12)-nuclei detected by APT for alloy A have complete physical picture of the energetics of nucleation of L12
solute-supersaturated compositions of 72 ± 3 Ni, 21 ± 3 Al, and structures in Ni-AL-Cr alloys. These energies permit an under-
6 ± 1 Cr; 71 ± 3 Ni, 19 ± 3 Al and 10 ± 2 Cr for alloy B; and 70 ± 3 Ni, standing of the morphology of L12 precipitates.
21 ± 5 Al and 9 ± 5 Cr for alloy C. The final equilibrium composi- There are three possible basic ordered structures in Ni-Al-Cr
tions of the L12 ordered precipitates are 76.3 ± 0.1 Ni, 17.8 ± 0.2 Al alloys: Ni3Al (L12), Al3Cr (L12), and Ni3Cr (DO22). The calculated
and 5.9 ± 0.1 Cr for alloy A; 76.5 ± 3.0 Ni, 16.2 ± 0.4 Al and 6.8 ± 0.3 bulk properties and elastic properties of the ordered phases are
Cr for alloy B; and 76.4 ± 0.2 Ni, 17.5 ± 0.3 Al and 6.1 ± 0.6 Cr for listed in Table 4. From our calculations, the formation energies of
alloy C. In these three alloys there is a depletion in the Al concen- Al3Cr (L12) and Al3Cr (DO22) have small negative values
tration in the g(f.c.c.)-matrix (0.4, 0.4 and 0.1 at.% in a (0.26 kJ mol1) for Al3Cr (L12) and 1.31 kJ mol1 for Al3Cr
layer z 2 nm thick), respectively, together with an accumulation of (DO22)), which are significantly greater than for Ni3Al (L12)
Cr (þ0.4, þ0.4 and þ 0.2 at.% over the same distance, respectively). (41.88 kJ mol1). The driving force for the transformation of Al3Cr
A small and more localized retention of Ni occurs as a result of the from L12 to DO22 is 1.05 kJ mol1. There is a large driving force for Al
local imbalance between the Al-depletion and Cr-accumulation. to substitute for Cr to form Ni3(AlxCr1-x) (L12). For the purposes of
Such solute concentration changes as a function of aging time predicting the morphology of the precipitate phases in Ni-Al-Cr
affect the interfacial Gibbs free energy, Section 3.5. The morphology alloys we mainly consider Ni3Al (L12), Al3Cr (L12), and mixed
of the precipitating phase becomes more spheroidal with Ni3(AlxCr1-x)(L12) as Ni3Al is the experimentally observed phase in
increasing Cr concentration in all three alloys, whereas the pre- concentrated Ni-Al-Cr alloys [6].
cipitates in Ni-Al binary alloys have a cuboidal morphology. The calculated lattice parameters at 0 K are 3.53 Å for f.c.c. Ni
Experimentally, the interfacial free energy, sg=g , can be deter-
0
(3.54 Å experimentally at 20  C), 3.57 Å for Ni3Al (L12) (3.56 Å
mined using the Kuehmann -Voorhees (KV) coarsening model [44]. experimentally at 20  C), 3.55 Å for Al3Cr (L12) at 0 K. The calculated
The relationship for sg=g for a nonideal, nondilute ternary alloy
0
lattice parameter misfit is 1.2% for Ni(f.c.c.)/Ni3Al(L12) and 0.6% for
consisting of a g-matrix and a g0 -precipitate phase with a finite Ni(f.c.c.)/Al3Cr(L12) at 0 K. These lattice parameter misfits are in
volume fraction of the g0 -phase is given by: agreement with the experimental value of 1.1% at room tempera-
  ture for Ni(f.c.c.)/Ni3Al(L12) [45].
g g g g
ðKKV Þ1=3 ki p2Al GAl;Al þ pAl pCr GAl;Cr þ p2Cr GCr;Cr We now discuss the bulk elastic properties for the ordered
sg=g ¼
0
(1) phases considered: Ni3Al (L12), Al3Cr (L12), and Ni3Cr (DO22). First,
g0
2pi V m we calculated the three elastic constants C11, C12, and C44 of a cubic
g
crystal by either tetragonal or trigonal lattice distortions [46]. The
where KKV and ki are the rate constants for the mean radius, <R(t) bulk modulus and the Voigt averaged shear modulus are deter-
g0
>, of the g0 -precipitates and supersaturation, respectively, V m is the mined from the bulk elastic constants by considering a hydrostatic
molar volume of the g -precipitate phase, pi is the magnitude of the
0
elastic deformation. The bulk modulus, B, and the Voigt averaged
g0 ;eq g;eq g
partitioning coefficient as defined by C i ð∞Þ  C i ð∞Þ, and Gi;j is shear modulus, <Gv>, are given by:
a shorthand notation for the partial derivatives of the molar Gibbs
free-energy of the g-matrix phase with respect to the solute species C11 þ 2C12
i and j. The experimentally determined average interfacial Gibbs B¼ (2)
3
free energy from coarsening data is 23.0 ± 1.2 mJ m2 for alloy A;
24.0 ± 1.6 mJ m2 for alloy B; and 21.0 ± 1.0 mJ m2 for alloy C using 3C44 þ C11  C12
the Kuehmann-Voorhees coarsening model for a ternary alloy at <G>V ¼ (3)
5
873 K (600  C) [6e8].

Table 1
Temporal evolution of the g(f.c.c.) and g0 (L12) phases's compositions as a function of aging time, at 873 K (600  C), as measured by atom-probe tomography for alloy A: Ni-7.5
Al-8.5 Cr at.%.

Aging time (h) g(f.c.c.)-matrix composition (at.%) g0 (L12)-precipitate composition (at.%)


Ni Al Cr Ni Al Cr

0.167 83.96 ± 0.05 7.37 ± 0.03 8.67 ± 0.03 72 ± 3 21 ± 3 6±1


0.25 84.04 ± 0.05 7.09 ± 0.09 8.9 ± 0.1 73 ± 1 21 ± 1 6.2 ± 0.7
1 84.50 ± 0.04 6.63 ± 0.04 8.87 ± 0.04 73.9 ± 0.4 20.0 ± 0.4 6.2 ± 0.2
4 84.75 ± 0.04 6.20 ± 0.03 9.05 ± 0.03 74.8 ± 0.2 19.1 ± 0.2 6.1 ± 0.1
16 84.92 ± 0.06 5.88 ± 0.04 9.20 ± 0.04 75.4 ± 0.2 18.6 ± 0.2 6.1 ± 0.1
64 84.99 ± 0.05 5.76 ± 0.03 9.25 ± 0.04 75.7 ± 0.2 18.3 ± 0.2 6.0 ± 0.1
256 85.09 ± 0.02 5.61 ± 0.02 9.30 ± 0.02 75.92 ± 0.06 18.18 ± 0.06 5.90 ± 0.04
1024 85.13 ± 0.06 5.53 ± 0.07 9.34 ± 0.04 76.11 ± 0.09 18.02 ± 0.09 5.87 ± 0.05
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 705

Table 2
Temporal evolution of the g(f.c.c.) and g0 (L12) phases's compositions as a function of aging time, at 873 K (600  C), as measured by atom-probe tomography for alloy B: Ni-5.2
Al-14.2 Cr at.%.

Aging time (h) g(f.c.c.) -matrix composition (at.%) g0 (L12)-precipitate composition (at.%)
Ni Al Cr Ni Al Cr

0.167 80.59 ± 0.09 5.19 ± 0.05 14.22 ± 0.08 71 ± 3 19 ± 3 10 ± 2


0.25 80.73 ± 0.09 5.07 ± 0.05 14.20 ± 0.08 73 ± 1 18.2 ± 0.9 9.2 ± 0.7
1 80.9 ± 0.1 4.75 ± 0.06 14.36 ± 0.09 73.4 ± 0.8 17.8 ± 0.6 8.8 ± 0.5
4 81.01 ± 0.15 3.97 ± 0.08 15.02 ± 0.14 74.3 ± 0.5 17.7 ± 0.4 8.0 ± 0.3
16 81.10 ± 0.06 3.61 ± 0.03 15.28 ± 0.06 75.5 ± 0.3 17.2 ± 0.2 7.3 ± 0.2
64 81.22 ± 0.07 3.45 ± 0.04 15.33 ± 0.07 75.7 ± 0.3 17.2 ± 0.3 7.2 ± 0.2
256 81.22 ± 0.07 3.30 ± 0.03 15.47 ± 0.07 76.0 ± 0.2 17.0 ± 0.2 7.1 ± 0.1
1024 81.16 ± 0.09 3.27 ± 0.04 15.57 ± 0.09 76.4 ± 0.3 16.7 ± 0.3 6.9 ± 0.2

Table 3
Temporal evolution of the g(f.c.c.) and g0 (L12) phases's compositions as a function of aging time, at 873 K (600  C), as measured by atom-probe tomography for alloy C: Ni-6.5
Al-9.5 Cr at.%.

Aging time (h) g(f.c.c.)-matrix composition (at.%) g0 (L12)-precipitate composition (at.%)


Ni Al Cr Ni Al Cr

0.5 83.05 ± 0.03 7.13 ± 0.08 9.82 ± 0.07 70 ± 3 21 ± 5 9±5


0.75 82.99 ± 0.03 7.22 ± 0.07 9.79 ± 0.07 71.3 ± 0.1 20.4 ± 0.2 8.4 ± 0.2
1 83.73 ± 0.01 6.46 ± 0.03 9.81 ± 0.03 71.5 ± 0.5 20.5 ± 0.8 8.1 ± 0.9
1.5 83.74 ± 0.03 6.38 ± 0.07 9.88 ± 0.06 72.3 ± 0.2 19.6 ± 0.3 8.1 ± 0.4
2 83.74 ± 0.02 6.38 ± 0.04 9.87 ± 0.04 72.8 ± 0.5 19.4 ± 0.8 7.8 ± 0.8
3.5 83.83 ± 0.04 6.27 ± 0.08 10.08 ± 0.08 74.1 ± 0.4 18.7 ± 0.8 7.2 ± 0.9
4 83.72 ± 0.01 6.29 ± 0.03 9.99 ± 0.03 74.53 ± 0.09 18.4 ± 0.2 7.1 ± 0.2
16 83.90 ± 0.01 5.90 ± 0.03 10.20 ± 0.03 75.09 ± 0.08 11.0 ± 0.1 6.9 ± 0.2
64 83.99 ± 0.02 5.80 ± 0.05 10.21 ± 0.05 75.55 ± 0.07 17.9 ± 0.1 6.6 ± 0.2
256 84.06 ± 0.02 5.71 ± 0.05 10.23 ± 0.05 75.82 ± 0.09 17.8 ± 0.2 6.4 ± 0.2
1024 84.11 ± 0.01 5.64 ± 0.04 10.25 ± 0.03 76.05 ± 0.05 17.65 ± 0.09 6.30 ± 0.07
4096 84.16 ± 0.01 5.57 ± 0.03 10.27 ± 0.02 76.24 ± 0.03 17.54 ± 0.06 6.22 ± 0.07

Table 4
The formation energy, lattice parameter, ao, bulk modulus, B, elastic constants, Cij, anisotropy ratio, A, and the Voigt averaged shear modulus, <G>V, from first-principles
calculations at 0 K (generalized gradient approximation (GGA)).

Ni(f.c.c.) Ni3Al(L12) Ni3Cr (L12) Ni3Cr (DO22) Al(f.c.c.)

Formation energy (kJ$mol1) 0 41.88 0.26 1.31 0


Lattice constant, a (nm) 0.352 0.357 0.349 a ¼ 0.342 4.032
c ¼ 0.678
Bulk module B (GPa) 189 175.8 196 201 79
C11(GPa) 239 228 184 192 109
C12(GPa) 165 149 147 156 65
C44(GPa) 102 115 96 105 32
< GV > (GPa) 66 71 73 76 28

_
The calculated elastic constants, C11, C12, and C44, Voigt averaged the superlattice
_
direction, G . The formation energy per atom,
shear modulus < G > V , and bulk modulus, B, of Ni, Ni3Al(L12), DEeq
CS ðx; G Þ, can be defined as the equilibrium value of the
Ni3Cr(L12), Ni3Cr(DO22) and Al are listed in Table 4. From our cal- composition-weighted sum of the energies to deform both bulk
culations, the two Al3Cr crystal structures have the highest bulk disordered Ni-rich phase A and ordered L12 phase B to the epitaxial
moduli, 196 GPa for L12 and 201 GPa for DO22. The ordered phases geometry_with lattice constant as perpendicular to superlattice
all have larger shear moduli than Ni, and Ni3Al(L12) has larger direction G as following [47]:
values of the elastic constants than Ni3Cr(L12).
We next computed the coherency strain energies for different       
_ _ _
strain orientations across the following three different interfaces: DEeq epi epi
CS x; G ¼ min xDEA aS ; G þ ð1  xÞDEB aS ; G
aS
;
Ni/Ni3Al(L12), Ni/Ni3Cr(L12) and Ni/Ni3Cr(DO22). The strain energies
are used as one contribution of the energetic quantities utilized to (4)
determine the driving force for nucleation, the radius of the critical
nucleus, and the eventual stable precipitate structure in Ni-Al-Cr where x is the mole fraction of Ni-rich phase A, DEepi A
is the epitaxial
epi
alloys. We have reported the values for the Ni/Ni3Al(L12) hetero- deformation energies of Ni-rich phase A, and DEB is the epitaxial
phase interface [9]. For the sake of completeness, we describe deformation energies of ordered L12 phase B. The coherency strain
briefly the methodology for the computation of coherency strain. energies for the [100], [110], and [111] directions for the Ni/
The coherency strain can be visualized in terms of the formation Ni3Al(L12), Ni/Al3Cr(L12) and Ni/Al3Cr(DO22) interfaces are dis-
energy of coherent long-period ApBq superlattices (A and B repre- played in Fig. 1.
sent the different phases considered, such as disordered Ni-rich We find that the Ni/Ni3Cr(L12) interface has the smallest strains
phase and ordered L12 phase), which are nonzero and depend on of the three possible interfaces with f.c.c. Ni. The [100] direction has
706 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

between the [100] and [111] directions, which means the pre-
cipitates tend toward a cuboidal morphology, whereas the Ni/
Al3Cr(L12) interface has a very small difference (0.003 eV$atom1)
between the [100] and [111] directions, which implies that the
precipitates tend toward a spheroidal morphology. The strain for
the Ni/Al3Cr(DO22) interface is different from that of the Ni/
Ni3Al(L12) and Ni/Al3Cr(L12) interfaces, where the [111] direction is
the elastically softest and the [100] direction is the elastically
hardest.

3.3. Solute site preference and partitioning pattern of Cr addition


across Ni(g)/Ni3Al(g’) interface

The site preference of Cr and the energetic driving force for the
partitioning of Cr were determined by first-principles calculations
for a system of 12  2  2 unit cells (192 atoms) constructed along
the [100] direction. The supercell was divided by a (100) interface,
and the two halves of the supercell were occupied by the g-Ni(f.c.c.)
and the g0 -Ni3Al(L12) phases, respectively. Every Ni atom on the g-
Ni(f.c.c.) of the interface was treated as a potential substitutional
site for the Cr atoms, while on the Ni3Al side, the Ni and Al sub-
lattice sites were treated as separate substitutional sites. To ensure
coherency of the (100) g(f.c.c.)/g0 -Ni3Al(L12) interface, the struc-
tures on either side of the interface were relaxed within the con-
straints of the Ni3Al crystal structure. The substitutional formation
energies of Cr, as a function of distance from the (100) interface,
were calculated as:
  
ECr/M ¼ Etot
Cr; precipitate þ nM mM  Etot þ nCr mCr (5)

  
Z; matrix þ nNi mNi  E
ECr/Ni ¼ Etot þ nCr mCr
tot
(6)

where M is Ni or Al, Etot is the total energy prior to substitution,


Etot tot
Cr; precipitate and E Cr; matrix are the total energies when Cr partitions
to the matrix or precipitate phases, respectively.The site substitu-
tional energies of Cr are smaller at the Al sublattice-sites than at the
Ni sublattice-sites, Table 5, which confirms that Cr atoms prefer to
occupy the Al sublattice-sites of the Ni3Al(L12) structure. The
calculated substitutional formation energies of Cr are smaller in the
g-Ni(f.c.c.) matrix than in the g0 -Ni3Al(L12)-precipitate, Fig. 2,
providing the energetic driving force, 0.436 eV atom1, for the
partitioning of Cr atoms to the g-Ni(f.c.c.) matrix phase. The addi-
tion of Cr leads to an increase in the total energy of the L12 crystal-
structure.
The average atomic forces and displacements associated with
the local stresses and strains resulting from the substitution of Cr
on the Ni and Al sublattice sites are displayed in Table 5. The sub-
stitution of Cr in the g-matrix phase results in a smaller value of the
atomic force and average atomic displacement than substitution of
Cr at either Ni or Al sublattice sites in Ni3Al(L12) precipitate phase,
thus the partitioning and substitution of Cr in the g-matrix phase is
preferred.
Fig. 1. Density-functional first-principles theory calculated equilibrium coherency
strain-energies at 0 K for the {100}, {110}, and {111} interfacial planes for: (a) Ni(f.c.c.)/
g0 -Ni3Al (L12); (b) Ni(f.c.c.)/Ni3Cr (L12); and (c) Ni(f.c.c.)/Ni3Cr (DO22). 3.4. Mixing between L12 Ni3Al(L12), Ni3Cr(L12) ordered phase

As discussed above, for both the experimental results,


smallest strain and the [111] direction has the largest strain for both Tables 1e3, and the substitutional pathway calculations, the solute
the Ni/Ni3Al(L12) and Ni/Al3Cr(L12) interfaces, whereas the [100] element, Cr, in concentrated Ni-Al-Cr alloys has a tendency to
direction is the elastically softest, and the [111] direction is the partition to the g-matrix solid-solution phase, whereas Cr and Al
hardest. Alternatively, the Ni/Al3Cr (DO22) interface has the largest share the solute sublattice in the Ni3Al(L12) precipitate phase. To
elastic strain energy among the three systems, which is consistent determine the stability of the mixed L12-ordered phases, we
with the largest lattice parameter mismatch of 3.48%. The Ni/ calculate the mixing energy between the ordered structures. To
Ni3Al(L12) interface has the largest difference (0.021 eV$atom1) mimic adequately the mixing statistics of a random distribution of
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 707

Table 5
Average atomic forces and displacements at the first nearest-neighbor distance, from first-principles magnetic calculations for three different structures: solid-solution NinCr,
Ni3(AlyCr1-y), with Cr substituting on the Al sublattice site, and (NixCr1-x)3Al with Cr substituting on the Ni sublattice site.

ECr/Ni, Al (eV atom1) Average Atomic Force (eV Ǻ1) Average Atomic Displacement (Ǻ)

NinCr 0.129 0.0091 0.0289


Ni3(AlyCr1-y) 0.565 0.0108 0.0371
(NixCr1-x)3Al 0.648 0.0149 0.0307

DHmix ¼ EðSQSÞ  xNi3 Al EðNi3 AlÞ  xNi3 Cr EðNi3 CrÞ


(7)
xNi3 Al þ xNi3 Cr ¼ 1

where EðNi3 AlÞ, EðNi3 CrÞ, and EðSQSÞ are the calculated total en-
ergies of fully relaxed L12 ordered Ni3Al, Ni3Cr, and the SQS
Ni3(Al,Cr), respectively. The mixed systems have positive mixing
energies, Fig. 3, indicating that mixing between Ni3Al(L12) and
Ni3Cr(L12) to form Ni3(Al,Cr)(L12) is energetically unfavorable. The
calculated positive mixing energies confirm Cr partitioning to the
g-matrix phase and agree with the positive substitutional energies,
where Al and Cr atoms share the same sublattice sites of the L12
structure.

3.5. Interfacial free energies of Ni/Ni3Al(L12), Ni/Ni3Cr(L12), Ni/


Ni3Cr(DO22) and Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12) interfaces
Fig. 2. The substitutional formation energies of Cr atoms as a function of distance from
the Ni(f.c.c.)/g0 -Ni3Al (L12) heterophase interface from density-functional first-princi-
ples calculations at 0 K: both Ni and Al sublattice sites are considered. The interfacial Gibbs free energies between disordered g and
ordered g0 phase are critical for determining the mechanisms that
lead to the nucleation of g’ precipitates with an ordered structure
from a disordered f.c.c. solid-solution. We calculate the energies of
the three possible interfacial systems in the Ni-Al-Cr system: Ni/
ordered structures in an alloy, with a supercell size that is Ni3Al(L12), Ni/Ni3Cr(L12), Ni/Ni3Cr(DO22) and Nin(AlyCr1-y)/
computationally feasible for DFT, we use the special quasi-random Ni3(AlxCr1-x) (L12). For each system, the (100), (110), and (111)
structure (SQS) approach developed by Zunger et al. [48]. SQSs are interfacies are considered. The interfacial configurations were
specifically designed small-unit-cell periodic structures that mimic constructed with the initial positions on an ideal and uniform f.c.c.
the most relevant nearest-neighbor pair and multisite correlation lattice with a lattice parameter of 3.57 Å for both phases. We fully
functions of random substitutional solute positions in alloys. The relaxed all atomic positions, cell vectors, and the total cell volume
concept of SQS originated from the well-established cluster for all three interfacial systems.
expansion (CE) methodology [5e12], and SQS structures have been Firstly, we determine three basic interfaces: Ni/Ni3Al(L12), Ni/
developed for mixing on several important lattice types: f.c.c. [49], Ni3Cr(L12), and Ni/Ni3Cr(DO22). We calculate the interfacial energy
b.c.c. [50], and h.c.p [51]. While the CE method is more powerful including strain effects [11], where the interfacial internal energies
and is capable of describing both ordered and disordered states, are calculated by subtracting the total energy of the phases on
within a single unified framework, the SQS approach is computa- either side of the interface from the total energy of a two-phase
tionally more efficient if the sole purpose is to obtain reliably the system containing an interface:
properties of random alloys. As a consequence of such atomic ar-
rangements, local environmentally dependent effects, such as
charge transfer and local atomic relaxations can be accurately
captured by the SQS approach [52]. The most straightforward way
to generate the SQSs is to enumerate exhaustively several possible
mixed L12 structures and then calculate their pair and multisite
correlation functions, as is implemented by the gensqs code in the
Alloy-Theoretic Automated Toolkit (ATAT) [53].
Since both Al and Cr atoms prefer the solute sublattice, for
creating a Ni3(Al1-nCrn) substitutional alloy, the substitutional
arrangement of Al and Cr atoms share the solute sublattice-type
lattice sites. In the present study, we generated different SQS-N
structures (with N ¼ 108, 32, and 108 atoms per unit cell) for a
random fcc lattice, with the compositions x ¼ 0.33, 0.5 or 0.67. For
Ni3(Al0.5Cr0.5), we generate the SQS-32 structure with 32 atoms of a
2  2  2 supercell. For Ni3(Al0.33Cr0.67) and Ni3(Al0.67Cr0.33), we
generate the SQS108 structure with 108 atoms of a 3 x 3 x 3
supercell to achieve convergence. The mixing enthalpy of the
randomly distributed solute elements, Al and Cr, with compositions Fig. 3. The mixing energies for Ni3Al (L12)/Ni3Cr (L12) calculated using the density-
xNi3 Al and xNi3 Cr can be calculated from: functional first-principles quasi-random structure (SQS) method.
708 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

1 h tot   i
sa=b ¼ E  Etot tot
a þ Eb  DEcs  T DSvib (8)
2A a=b

where A is the interfacial area, the subscripts a and b denote the


two phases across the interface plane: the possible phases are
Ni(f.c.c.), Ni3Al(L12), Ni3Cr(L12) and Ni3Cr(DO22)); Etot
a=b is the total
internal energy of the relaxed a/b system containing an interface;
a and Eb are the total internal energies of the a- and b-phases,
Etot tot

respectively, each relaxed to their own equilibrium geometry; DEcs


is the coherency strain energy between the a- and b-phases as
discussed above; DSvib is the phonon vibrational entropy calculated
by the first-principles method as discussed below. The technique of
calculating an interfacial energy, including strain effects, was
described by Wolverton and Zunger [11]. The resulting calculated
interfacial internal energies at 0 K for Ni/Ni3Al(L12), Ni/Ni3Cr(L12),
and Ni/Ni3Cr(DO22) are listed in Table 6. Generally, the Ni/
Ni3Cr(DO22) interface has the largest calculated interfacial internal
energy for all (100), (110), and (111) planes, while the Ni/Ni3Cr(L12)
interfacial internal energy has the smallest value for all (100), (110),
and (111) planes. The (100) interface is found to have the smallest
interfacial internal energy for Ni/Ni3Al(L12) and Ni/Ni3Cr(L12) in-
terfaces. The large energy difference between Ni/Ni3Al(L12)(100)
and Ni/Ni3Al(L12)(111), 11 mJ m2, affects the morphology of
Ni3Al(L12) precipitates, which tends to form cuboidal structures.
This finding is in agreement with prior calculations [54,55] and
with high resolution electron microscopy (HREM) results that show
that Ni3Al(L12) precipitates develop large (100) facets and small
(110) and (111) facets as they nucleate, grow and coarsen [56]. The
energy difference between the (100) and (111) facets is small for Ni/
Ni3Cr(L12), Table 6. Ni3Cr(L12) precipitates tend to form spheroidal
ordered precipitates. The Ni/Ni3Cr(DO22) system has a larger
interfacial internal energy than do the Ni/Ni3Al(L12) and Ni/
Ni3Cr(L12) interfaces and the (111) plane has a lower value than
other two planes (110), (111) at 0 K.
Fig. 4. The temporal evolution of the Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12) interfacial energy
Secondly, we consider the effect of Cr on Nin(AlyCr1-y)/
from the density-functional first-principles calculations as function of aging time for
Ni3(AlxCr1-x) interfaces using the three alloy compositions. First, we three Ni-Al-Cr alloys at 0 K: (a) Alloy A: Ni-7.5Al-8.5Cr at.%; (b) Alloy B: Ni-5.2Al-14.2Cr
calculate the interfacial internal energy at 0 K, Fig. 4, according to at.%; (c) Alloy C: Ni-6.5Al-9.5Cr at.%.
the experimental concentration profiles as a function of aging time
displayed in Tables 1e3 The phase compositions for calculated
interfacial structures of both g and g0 -phases are then varied to interfacial systems. As the aging time increases, the calculated
match the experimentally determined values listed in Tables 1e3. interfacial internal energies decrease with increasing Cr concen-
The substitutional process is random and five configurations are tration and they achieve equilibrium plateaus after 16 h of aging for
generated for each aging time and the total energies Etot tot the three alloys. We take the interfacial internal energies at 128 h as
a=b , Ea and
Etot are calculated to obtain the average interfacial internal energy. the equilibrium values displayed in Table 6 for the vibrational en-
b
The individual phases are fully relaxed in our calculations. The tropy calculations to determine the temperature effects discussed
calculated interfacial internal energy at 0 K is plotted as function of below. We find that the calculated equilibrium interfacial free en-
aging time, Fig. 4, for the three alloys. The (100) interface has the ergies for the three alloys are in good agreement with the values
smallest interfacial internal energy of the three alloys, which determined experimentally at 873 K (600  C): 23.0 ± 1.2 mJ m2 for
exhibit the same trends as the Ni/Ni3Al(L12) and Ni/Ni3Cr(L12) alloy A; 24.0 ± 1.6 mJ m2 for alloy B; and 21.0 ± 1.0 mJ m2 for
alloy C.
Now we consider the effect of temperature on the phonon
Table 6 vibrational entropy using the MedeA package with PHONON soft-
Density functional theory (DFT) first-principles calculated g-Ni/Ni3Al (L12), g-Ni/
ware developed by Krzysztof Parlinski [57]. The magnetic spin
Ni3Cr (L12), g-Ni/Ni3Cr (DO22) and three Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12) interfacial
energies using nonmagnetic spin-polarized methods for three different {hkl} planes
polarization calculations were utilized for the phonon calculations.
at 0 K (mJ m2). The effects of temperature on the phonon vibrational entropy are
based on the sharp interface model btween two phases. The
{100} {110} {111}
vibrational entropy is calculated by:
g-Ni/Ni3Al (L12) 25 33 36
g-Ni/Ni3Cr (L12) 28 31 33
g-Ni/Ni3Cr (DO22) 89 81 66
Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12)A 31 36 38 DSvib ¼ DSmatrix=precipitates
vib
 DSprecipitates
vib
 DSmatrix
vib (9)
Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12)B 32 34 36
Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12)C 30 37 40 To calculate the phonon dispersions for the g-matrix phase
Alloy A: Ni-7.5Al-8.5Cr at.%.
(pure Ni and solid-solution Nin(AlyCr1-y)) and g0 (L12)-precipitates,
Alloy B: Ni-5.2Al-14.2Cr at.%. Ni3Al(L12), Ni3Cr(L12), Ni3Cr(DO22) and L12 Ni3(AlxCr1-x)), we used
Alloy C: Ni-6.5Al-9.5Cr at.%. the frozen-phonon method for 3  3  3 supercells for bulk
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 709

supercells for the g(f.c.c.)/g’ interfacial structures, and 2  2  2


supercells for the different (hkl) planes. Using this methodology we
calculated a series of total energy differences between atomic
configurations with and without a single displaced atom. The
harmonic approximation was employed for all our vibrational cal-
culations. The energy for the harmonic approximation [41e43] is
calculated as:

1X 1X T
H¼ _ 2þ
mi ½uðiÞ u ðiÞFði; jÞuðjÞ (10)
2 i 2 i;j

where mi is the mass of atom i; uðiÞ is the displacement away from


its equilibrium position; and Fði; jÞ is the tensor of the force con-
stants from the displacement of atom j to the force on atom i. The
Helmholtz free energy of the system is calculated as [58]:

ð
∞   
hv
F ¼ E* þ NkB T ln 2 sinh dðvÞdv (11)
2kB T
0

where E* is the total internal energy of the supercell at T ¼ 0 K, at its


equilibrium position; kB is Boltzmann's constant; h is Planck's
constant, N is the number of atoms in the supercell; and dðvÞ is the
phonon density-of-states (DOS). The DOS of the phonon vibrations
for bulk pure Ni (solid-solution), Ni3Al(L12), Ni3Cr(L12) and
Ni3Cr(DO22) are displayed in Fig. 5(a), and the phonon dispersion
spectra of these three ordered structures are displayed in Fig. 5 (b).
There are no imaginary frequencies in the phonon dispersion
curves, and the three structures studied are considered dynami-
cally stable. The ordered phases, Ni3Al(L12), Ni3Cr(L12) and
Ni3Cr(DO22), have higher average phonon frequencies, with less
weight in the low-frequency region compared to the g-(f.c.c.)-
phase. The vibrational modes move to higher frequencies when Ni
atoms are replaced by either Al or Cr atoms. Structural differences
in the vibrational entropy are due to changes in the number of Ni-
Ni, Ni-Al, Ni-Cr nearest-neighbor (NN) bonds. In the strongly
intermetallic-forming Ni-Al system Ni-Al bonds are stiffer than Ni-
Ni, Al-Al, Cr-Cr, or Ni-Crbonds, leading to higher phonon fre-
quencies and smallest vibrational entropies [9]. The DOS of the
high-frequency modes in g0 -Ni3Al is at ~10e11 THz, corresponding
to Ni-Al vibrations, resulting in the DOS peaks moving to higher
frequencies due to fewer NN Ni-Ni bonds. The DOS of the high-
frequency mode in Ni3Cr(L12) is at 9.3 THz, corresponding to Ni-
Cr vibrations, but such a mode is much weaker than g0 -Ni3Al.
Interestingly, Ni3Cr(DO22) has strong lower frequency peaks at 0.4
to 0.5 THz than do Ni3Al(L12) and Ni3Cr(L12). A low frequency mode
is related to the formation of magnetic domains, which make
Ni3Cr(DO22) more stable than Ni3Cr(L12). The calculated vibrational
entropies are: 17.13 J mol1 K1 for g0 -Ni3Al; 11.45 J mol1 K1
for Ni3Cr(L12); and 14.32 J mol1 K1 for g0 - Ni3Cr(DO22). The DOS
of the (100) g-Ni/ordered-phase interface is a combination of g-Ni
and ordered phases: g0 -Ni3Al, Ni3Cr(L12), and Ni3Cr(DO22). The
lower frequency phonon modes are very similar to those of g-Ni, Fig. 5. (a) A comparison of density-functional theory calculated phonon vibrational
density of states (DOS) for the following structures: g-Ni(f.c.c.); ordered g0 -Ni3Al (L12)
and the higher frequency modes decrease from 10 to 11 THz to
and g0 -Ni3Cr(L12), and the Ni3Al (DO22). (b) the phonon dispersion spectra of three
9e10 THz due to the distortion of the ordered phases. We observe ordered g0 -Ni3Al (L12) and g0 -Ni3Cr(L12), and the Ni3Al (DO22).
that the higher frequency modes around 9e10 THz move to lower
frequencies for the L12 phase, which indicates that the Ni-Al and Ni-
Cr bonding structures and ferromagnetic fields across the denser
(111) interface are more disrupted than for the (100) interface Fig. 6. As we discussed in our prior article [9], the calculated
because of the absence of atomic symmetry across this interface. interfacial free energies 22, 28, and 29 mJ m2 for the (100), (110)
By accounting for vibrational entropy, the equilibrium interfa- and (111) Ni/Ni3Al(L12) interfaces, respectively, at 873 K are in
cial free energies are calculated as a function of temperature. The agreement with the average values over all facets, calculated by
interfacial free energies of Ni/Ni3Al(L12), Ni/Ni3Cr(L12) and Ni/ Ardell, of 22.33 ± 1.31 and 31.08 ± 1.45 mJ m2, obtained by fitting
Ni3Cr(DO22) interfaces decrease as the temperature increases, the existing and our previously reported experimental data to his
710 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

We next investigated mixing between Ni3Al(L12) and Ni3Cr(L12).


The Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12) interfacial free energy as a
function of Cr concentration, Fig. 7, follows a similar trend as
Ni3Al(L12) and Ni3Cr(L12); with the (100) having a lower interfacial
free energy than the (110) or (111) interfaces. The calculated
interfacial free energies at 873 K (600  C), are: 25.5 mJ m2 for the
(100); 28.7 mJ m2 for the (110); and 29.8 mJ m2 for the (111) for
alloy A (Ni-7.5Al-8.5Cr at.%). Then 26.6 mJ m2 for the (100),
27.6 mJ m2 for the (110), and 28.7 mJ m2 for the (111) for alloy B
(Ni-5.2Al-14.2Cr at.%), and 24.7 mJ m2 for the (100), 29.6 mJ m2
for the (110), and 31.9 mJ m2 for the (111), for alloy C (Ni-6.5Al-
9.5Cr at.%). The difference between the (100) and (111) interfacial
free energies, calculated at 873 K (600  C), decreases as the Cr
concentration increases: 7.2 mJ m2 for alloy C (Ni-6.5Al-9.5Cr
at.%); 4.3 mJ m2 for alloy A (Ni-7.5Al-8.5Cr at.%); and 2.1 mJ m2
for alloy B (Ni-5.2Al-14.2Cr at.%). We, therefore, conclude that the
addition of Cr decreases the differences of the interfacial energies
between different {hkl} planes, which results in the precipitate
morphology transforming from cubic-to-spherical.

Fig. 6. Density functional theory calculated interfacial Gibbs free energies for the
{100}, {110}, and {111} interfacial planes as a function of temperature: (a) g-Ni(f.c.c.)/
g0 -Ni3Al (L12); (b) g-Ni(f.c.c.)/Ni3Cr (L12); and (c) g-Ni(f.c.c.)/g/Ni3Cr (DO22). These
calculations include coherency strain energies and phonon vibrational entropies.

trans-interface diffusion-controlled (TIDC) [59] and LSW-type


coarsening models, respectively [60]; the latter values are, how-
ever, in better agreement with our calculations than Ardell's TDIC
values. Ji-Cheng Zhao et. al. use dual-anneal diffusion multiples to
to generate a composition gradient and adjust the interfacial en-
ergy value in simulations using the classical nucleation and growth
theories as implemented in the Kampmann-Wagner numerical
(KWN) model. They obtained the average Ni/Ni3Al(L12) interfacial
free energy at 873 K of 14 mJ m2 [61], which is 10 mJ m2 lower
than the one we calculated. Metastable Ni/Ni3Cr(L12) interfacial
structures have very similar interfacial free energy values: 20.5,
21.3, and 21.8 mJ m2 for the (100), (110) and (111) planes at 873 K
(600  C), respectively. Theoretically, the preferred morphology of
Ni3Cr(L12) is spherical at 873 K (600  C), unlike the morphology of
the Ni3Al(L12) phase, which is cubic. The Ni/Ni3Cr(DO22) interfacial
free energy at 873 K (600  C) is larger than that of the two L12 or-
dered phases (Ni/Ni3Al(L12) and Ni/Ni3Cr(L12)), with the values,
73.9, 69.1, and 58.9 mJ m2 for the (100), (110) and (111) planes, Fig. 7. Density functional theory calculated Nin(AlyCr1-y)/Ni3(AlxCr1-x) (L12) interfacial
free energies for the {100}, {110}, and {111} interfacial planes as a function of tem-
which yields a cuboidal-shaped precipitate dominated by {111} perature for three different Ni-Al-Cr systems: (a) Alloy A: Ni-7.5Al-8.5Cr at.%; (b) Alloy
planes due to the (111) interface having the lowest interfacial free B: Ni-5.2Al-14.2Cr at.%; and (c) Alloy C: Ni-6.5Al-9.5Cr at.%. These calculations include
energy. coherency strain energies and phonon vibrational entropies.
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 711

3.6. Nucleation properties: first-principles nucleation-barriers and Ni/Ni3Cr(L12) and 4.89 meV$atom1 for Ni/Ni3Cr(DO22), which en-
critical radii for Ni3Al(L12) Ni3Cr(L12), Ni3Cr(DO22) and Ni3(AlxCr1- ables us to calculate R* and WR ðR* Þ, Eqs. (14) and (15). These elastic
x)(L12) internal strain energies represent an 11% contribution to the total
nucleation free energies of Ni3Al(L12), a 10% contribution for
Classic nucleation theory (CNT) can be used to determine the Ni3Cr(L12), and a 28% contribution for Ni3Cr(DO22). The quantity
nucleation barrier and critical size of all the possible ordered phases DFV for Ni3Al(L12), Ni3Cr(L12) and Ni3Cr(DO22) is a function of
in Ni-Al-Cr alloys studed: Ni3Al(L12), Ni3Cr(L12), Ni3Cr(DO22) and temperature because of the entropic term, and the values of DFV
Ni3(AlxCr1-x)(L12). According to CNT, the thermodynamic energy were determined using Eq. (13).
barrier to nucleation is controlled by a three-factor competition The calculated values at 873 K of R* (Table 7) are: 7.36 Å for
[62,63]: (1) a negative bulk free energy contribution, which is the Ni3Al(L12); 14.87 Å for Ni3Cr(L12); and 17.14 Å for Ni3Cr(DO22). And
chemical formation Helmholtz free energy change, DFchem ; (2) a the values of WR ðR* Þ are 2.7 and 1.62 meV (4.31  1022 and
positive elastic strain energy contribution, DEelastic ; and (3) a pos- 2.59  1022 J) for Ni3Al(L12) and Ni3Cr(L12), respectively, 4.4 meV
itive interfacial Gibbs free energy contribution, s. Using CNT, we (7.03  1022 J) for Ni3Cr(DO22), respectively. These results
can understand and explain the thermodynamic mechanism for demonstrate that the formation of Ni3Cr(L12) requires a smaller
nucleation of Ni3Al(L12), Ni3Cr(L12), Ni3Cr(DO22) and Ni3(AlxCr1- DFV than does Ni3Al(L12) at 873 K (600  C), while Ni3Al(L12) has the
x)(L12) in a Ni(f.c.c.) matrix. We described our first-principles cal- smallest critical radius. Ni3Cr(DO22) has a significantly larger DFV
culations above for the bulk internal energy, elastic strain energy, and has a very large critical radius, which makes Ni3Cr(DO22) more
and interfacial Gibbs free energy of these systems, which are difficult to nucleate than either Ni3Al(L12) or Ni3Cr(L12). This result
required for calculating the energy barrier to nucleation. Now we is consistent with our experimental results as Ni3Cr(DO22) isn't
assemble these three quantities into a first-principles prediction of experimentally observed. Experimentally, we only observe
the nucleation thermodynamics of putative Ni3Al(L12), Ni3Cr(L12) Ni3Al(L12) and not Ni3Cr(L12). The reason is that even though
and Ni3Cr(DO22) ordered precipitates. We consider homogeneous Ni3Cr(L12) has the smallest, but it has a much larger critical radius
nucleation of spherical precipitates. The reversible work for the than Ni3Al(L12). An embryo of Ni3Cr(L12) has difficulty to stabilize
formation of a spherical nucleus, WR , as a function of nucleus due to its larger critical size. Also, the Cr diffusivity is much smaller
radius, R, is given by Refs. [64,65]: than Al's diffusivity. Ni-Cr short-range order was only observed
experimentally in the high Cr concentration alloy, B, in the early
4p 3 stage of nucleation [66]. An embryo of Ni3Cr(L12) only acts as the
WR ¼ R DFV þ 4pR2 s (12)
3 nucleant for mixed L12 Ni3(AlxCr1-x). Similar calculations were also
performed for the mixed L12 Ni3(AlxCr1-x) phase using the inter-
Where facial free energies and formation energies for the three alloys
studied, Table 5. The critical precipitate radii are: 9.23 Å for alloy B
DFV ¼ DFchem þ DEelastic (13) (Ni-5.2Al-14.2Cr at.%); 8.45 Å for alloy A (Ni-7.5Al-8.5Cr at.%); and
The critical radius, R*, for nucleation is given by: 8.04 Å for alloy C (Ni-6.5Al-9.5Cr at.%). The values of WR ðR* Þ are:
1.12 meV (1.79  1022 J) for alloy B (Ni-5.2Al-14.2Cr at.%); 1.75 meV
2s (2.81  1022 J) for alloy A (Ni-7.5Al-8.5Cr at.%); and 2.04 meV
R* ¼ (14) (3.27  1022 J) for alloy C (Ni-6.5Al-9.5Cr at.%). We therefore
DFchem þ DEelastic
conclude that the addition of Cr decreases significantly WR ðR* Þ for
And the critical net reversible work required for the formation of nucleating Ni3(AlxCr1-x)(L12), but it has only a small effect on R*,
a critical spherical nucleus, WR ðR* Þ, with radius R*, is given by: ranging from 9.23 to 8.04 Å.

 16p s3
WR R * ¼ (15) 3.7. The effects of Cr additions on precipitation morphology
3 ðDFchem þ DEelastic Þ2

The chemical formation energies were calculated based on the The equilibrium morphology of the g0 (L12)-precipitates trans-
nucleation of Ni3Al (L12), Ni3Cr (L12), or Ni3Cr (DO22) from a dilute forms from cubic-to-spherical, which may be faceted, when adding
solid-solution, Nin M/Ni3 M þ Nin3 , where M ¼ Al or Cr, and Cr to the Ni-Al alloys, studied experimentally utilizing 3-D APT
n ¼ 31 in 2x2x2 supercell. The equation for DFchem is given by: reconstructions for specimens aged at 873 K (600 ); Figs. 8(a) and
(b) [6,24,67]. The morphology of g0 (L12)-precipitates derived from
DFchem ¼ DFðNi3 MÞ þ ðn  3ÞDFðNiÞ  DFðNin MÞ interfacial free energies calculated using first-principles methods
¼ ½DHðNi3 MÞ  DHðNin MÞ (16) involves the {100}, {110} and {111} facets. Figs. 8(c), (d), and (f) were
T½DSðNi3 MÞ  DSðNin MÞ determined from Wulff constructions using our first-principles

The entropy terms DS in Eq. (16) are from the phonon vibra-
tional entropies of the ordered Ni3M and NinM cells in eqs. (9), (10), Table 7
and (11). We consider the enthalpy, H, to be equal to the internal The calculated nucleation parameters (critical radius, R* and net reversible work,
WR ðR* Þ) for classical nucleation theory for four possible nuclei: Ni3Al (L12);
energy, E, because the pressure-volume term in H is negligible
Ni3Cr(L12); Ni3Cr(DO22); and Ni3(Al,Cr)(L12).
compared to E in the solid-state. In a dilute NinZ solution, we did
not consider its contribution because the configurational entropy is R* (Å) WR ðR* Þ (meV) WR ðR* Þ (J)

much smaller than the vibrational entropy. Using Eq. (16) the Ni3Al (L12) 7.36 2.70 4.31  1022
calculated values of DFchem at 873 K are 38.5 meV$atom1 for Ni3Cr(L12) 14.87 1.62 2.59  1022
7.03  1022
Ni3Al(L12), 9.78 meV$atom1 for Ni3Cr(L12), Ni3Cr(DO22) 17.14 4.40
Ni3(Al,Cr)(L12)A 8.45 1.75 2.81  1022
and 13.78 meV$atom1 for Ni3Al(DO22). The values of DEelastic are Ni3(Al,Cr)(L12)B 9.23 1.12 1.79  1022
calculated from the coherency strains as explained above. We use Ni3(Al,Cr)(L12)C 8.04 2.04 3.27  1022
the elastic strains in the corresponding softest directions, (100) for Alloy A: Ni-7.5Al-8.5Cr at.%.
Ni3Al(L12) and Ni3Cr(L12), and (111) for Ni3Al(DO22). The elastic Alloy B: Ni-5.2Al-14.2Cr at.%.
strains are 5.49 meV$atom1 for Ni/Ni3Al(L12), 1.12 meV$atom1 for Alloy C: Ni-6.5Al-9.5Cr at.%.
712 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

Fig. 8. Comparisons of the morphologies of the precipitate phases between experimental 3-D atom-probe tomography reconstructions and calculated Wulff constructions based on
first-principles calculated values of the {100}, {110} and {111} interfacial Gibbs free energies at 873 K (600  C): (a) experimental g0 (L12) phase in a Ni-13 at.% Al alloy. The specimens
were aged in bulk for 4096 h at 823 K (600  C) prior to being studied by atom-probe tomography. The [100]-direction is perpendicular to a cube-face of a g0 (L12)-precipitate. (b)
experimental g0 (L12) phase in Alloy A (Ni-7.5 Al 8.5Cr at.%). The alloy samples were aged at 873 K (600  C) under flowing argon for 1024 h. (c) calculated g0 -Ni3Al (L12) precipitate
phase; (d) calculated g0 - Ni3(Al,Cr)(L12)- precipitate phase; and (f) calculated g0 -Ni3Cr (L12)-precipitate phase.

interfacial free energies at 873 K (600  C) employing the NIST Both Ni3Al(L12) and Ni3Cr(DO22) are thermally stable structures,
program Wulffman (http://www.ctcms.nist.gov/wulffman/) while Ni3Cr(L12) is a metastable structure. The formation en-
[55,68,69]. If the anisotropic interfacial free energies are known, ergies for Ni3Cr(L12) and Ni3Cr(DO22) are significantly smaller
the Wulff construction yields the equilibrium morphology of a than for Ni3Al(L12). The driving force for the transformation of
small precipitate. The smaller is the interfacial free energy the ordered Al3Cr from the L12 structure to the DO22 structure is
larger is the area of an {hkl}-facet. Employing our calculated 1.05 kJ mol1 (0.011eV atom1).
interfacial free energies we determined the equilibrium Wulff  Chromium and Al share the same sublattice site in the L12
morphologies at 873 K (600  C) for Ni3Al(L12), Ni3(AlxCr1-x)(L12) ordered-structure. Chromium partitions from the L12 ordered-
and Ni3Cr(L12), which all exhibit the morphological transformation structure to the disordered f.c.c. matrix with a driving force of
from a cube-to-a-sphere, which may be faceted. The result for alloy 0.436 eV atom1.
A (Ni-7.5Al-8.5Cr at. %) is displayed in Fig. 8(b). For Ni3Al(L12) it is a  It is demonstrated that phase mixing between Ni3Al(L12) and
great rhombicuboctahedron (or a truncated cuboctahedron), while Ni3Cr(L12) is not energetically preferred, utilizing first-principles
Ni3Cr(L12) has the morphology of a perfect sphere. The morphology quasi-random structure (SQS) calculations.
of g0 -Ni3(AlxCr1-x)(L12)-precipitates is spherical covered with  From experimental 3-D APT reconstructions we observed that
smaller {100} facets than the bigger {100} facets that occur for bi- the precipitate morphology with Cr additions to the Ni-Al alloy
nary Ni-Al alloys [9,70e72]. Hence, the addition of Cr changes changes from cubic-to-spherical at a temperature of 873 K
significantly the morphology of g0 (L12)-precipitates. (600  C). The compositions of the g(f.c.c.)-matrix and the or-
dered g0 (L12)-precipitate phases of these alloys evolve tempo-
4. Summary and conclusions rally as the g(f.c.c.)-matrix becomes enriched in Ni and Cr and
depleted in Al, while the g0 (L12)-precipitates become enriched
We have studied systematically the stability and nucleation of in Al and depleted in Cr, Tables 1e3.
putative ordered-precipitates in Ni-Al-Cr alloys based on bulk  The interfacial Gibbs free energies of binary Ni/Ni3Al(L12), Ni/
thermodynamics (static total energies and vibrational free en- Ni3Cr(L12), and Ni/Ni3Cr(DO22) interfaces were calculated for
ergies) and interfacial Gibbs free energies. Additionally, the effects the {100}, {110}, and {111} planes. The temperature de-
of Cr additions on the morphology of the ordered-precipitate pendencies of the interfacial Gibbs free energies is determined
phases has been studied experimentally employing 3-D atom- utilizing the phonon vibrational entropy. The calculated inter-
probe tomography (APT) and first-principles calculations. From facial free energies at 873 K (600  C) are: 25.5 mJ m2 {100},
these studies we reach the following conclusions: 28.7 mJ m2{110}, and 29.8 mJ m2 {111} for alloy A (Ni-7.5Al-
8.5Cr at.%); 26.6 mJ m2 {100}, 27.6 mJ m2 {110}, and
 The formation energy and elastic constants for three possible 28.7 mJ m2 {111} for alloy B (Ni-5.2Al-14.2Cr at.%); and
ordered-precipitate phases, Ni3Al(L12), Ni3Cr(L12) and 24.7 mJ m2 {100}, 29.6 mJ m2 {110}, and 31.9 mJ m2 {111} for
Ni3Cr(DO22), are calculated utilizing first-principles calculations. alloy C (Ni-6.5Al-9.5Cr at.%). The Ni/Ni3Cr(DO22) interface has
Z. Mao et al. / Acta Materialia 166 (2019) 702e714 713

the largest calculated interfacial Gibbs free energy, 73.9 mJ m2 These results are in agreement with our three-dimensional
for {100} plane, while the Ni/Ni3Cr(L12) interfacial Gibbs free atom-probe tomographic observations of the three Ni-Al-Cr al-
energy has the smallest value 20.5 mJ m2 for {100} plane. The loys studied [71].
{100} interface is found to have the smallest interfacial Gibbs
free energy for the Ni/Ni3Al(L12) and Ni/Ni3Cr(L12) interfaces. Acknowledgements
The {111} interface is found to have the smallest interfacial en-
ergy for the Ni/Ni3Cr(DO22) interface. The large interfacial free This research was supported by the National Science Founda-
energy difference between the {100} and {111} interfaces in Ni/ tion, Division of Materials Research grant number DMR-1610367
Ni3Al(L12) affects the morphology of Ni3Al(L12) precipitates, 001 Profs. Diana Farkas and Gary Shiflet, grant officers. The alloys
which tends to be cubic. The interfacial Gibbs free energy dif- were processed at the NASA Glenn Research Center by Dr. Ronald D.
ference between the {100} and {111} interfaces of Ni/Ni3Cr(L12) Noebe. APT measurements were performed at the Northwestern
is very small and therefore Ni3Cr(L12) precipitates tend to form University Center for Atom Probe Tomography (NUCAPT) by Drs.
spherical precipitates. Chantal K. Sudbrack, Kevin E. Yoon, Chris Booth-Moorison, and
 Chromium additions to Ni-Al alloys affect significantly the Yang Zhou during their Ph.D. thesis studies and Ms. Jessica
interfacial Gibbs free energies differences between the {100} Weninger during her undergraduate research studies. We thank
and {111} interfaces. The interfacial Gibbs free energy differ- Prof. C. Wolverton for helpful discussions and suggestions. The
ences are calculated at 873 K (600  C), and become smaller with LEAP tomograph was purchased with initial funding from the NSF-
increasing Cr concentration: 7.1 mJ m2 for alloy C (Ni-6.5Al- MRI program (DMR 0420532, Dr. Charles Bouldin, grant monitor)
9.5Cr at.%), 5.3 mJ m2 for alloy A (Ni-7.5Al-8.5Cr at.%), and and ONR-DURIP program (N00014-0400798, Dr. Julie Christodou-
1.9 mJ m2 for alloy B (Ni-5.2Al-14.2Cr at.%). lou, grant monitor) programs. Additionally, the LEAP tomograph
 The equilibrium interfacial Gibbs free energies in three Ni-Al-Cr was enhanced with a picosecond green laser with funding from the
ternary alloys are determined by utilizing the experimental ONR-DURIP program (N00014-0610539, Dr. Julie Christodoulou,
concentration profiles for both the disordered g(f.c.c.)-matrix grant officer). We gratefully acknowledge the Initiative for Sus-
and ordered g0 (L12)-precipitates as function of aging time from tainability and Energy at Northwestern (ISEN) for grants to upgrade
atom-probe tomography experimental three-dimensional re- the capabilities of NUCAPT. We wish to thank research associate
constructions. The calculated Gibbs interfacial free energies for Prof. Dieter Isheim for managing NUCAPT and discussions.
Ni3(AlxCr1-x)(L12) are in very good agreement with experi-
mentally determined average interfacial Gibbs free energies at References
873 K 600  C). We find that the g(f.c.c.)-matrix/{100}-interface
has the smallest interfacial Gibbs free energy for the three alloys [1] C.T. Sims, N.S. Stoloff, W.C. Hagel, Superalloys II High-Temperature Materials
studied. As the aging time increases, the interfacial Gibbs free for Aerospace and Industrial Power, John Wiley & Sons Inc., New York, 1987.
[2] B.B. Seth, Superalloys-the utility gas turbine perspective, in: T.M. Pollock,
energies decrease with increasing Cr concentrations in the L12- R.D. Kissinger, R.R. Bowman, K.A. Green, M. McLean, S.L. Oson, J.J. Schirra
ordered precipitates and achieve equilibrium plateaus after 16 h (Eds.), Superalloys 2008, Proceedings of the International Symposium on Su-
aging time. peralloys, 11th, Champion, PA, United States, 2000, pp. 16e23, 2008, TMS,
Warrendale, PA.
 Classical nucleation theory demonstrates that nucleation of [3] H. Wendt, P. Haasen, Nucleation and growth of g'-precipitates in Ni-14 at.% Al,
Ni3Al(L12) is easier to stabilize than that of metastable Acta Metall. 31 (10) (1983) 1649e1659.
Ni3Cr(L12) at 873 K (600  C). Ni3Cr(L12) has the smallest [4] Y. Amouyal, Z.G. Mao, C. Booth-Morrison, D.N. Seidman, On the interplay
between tungsten and tantalum atoms in Ni-based superalloys: an atom-
WR ðR* Þ, but it has a much larger critical radius than Ni3Al(L12). probe tomographic and first-principles study, Appl. Phys. Lett. 94 (4) (2009).
The embryo of Ni3Cr(L12) is dffucult to stabilize due to its larger [5] C.L. Jia, M. Lentzen, K. Urban, Atomic-resolution imaging of oxygen in
critical radius and it only acts as an nucleant for mixed L12 perovskite ceramics, Science 299 (5608) (2003) 870e873.
[6] C. Booth-Morrison, J. Weninger, C.K. Sudbrack, Z. Mao, R.D. Noebe,
Ni3(AlxCr1-x). Ni3Cr(DO22) is more difficult to nucleate than
D.N. Seidman, Effects of solute concentrations on kinetic pathways in Ni-Al-Cr
Ni3Cr(L12) and Ni3Al(L12). The calculated critical radius of alloys, Acta Mater. 56 (14) (2008) 3422e3438.
Ni3Al(L12) nuclei (7.36 Å) is smaller than the values for [7] C.K. Sudbrack, R.D. Noebe, D.N. Seidman, Compositional pathways and capil-
lary effects during early-stage isothermal precipitation in a nondilute Ni-Al-Cr
Ni3Cr(L12) (14.87 Å) and Ni3Cr(DO22) (17.14 Å). The critical net
alloy, Acta Mater. 55 (2007) 119e130.
reversible work required for nucleation is 4.31x10-22, 2.59x10- [8] C.K. Sudbrack, K.E. Yoon, Z. Mao, R.D. Noebe, D. Isheim, D.N. Seidman, Tem-
22 and 7.03x10-22 J for Ni3Al(L12), Ni3Cr(L12) and poral evolution of nanostructures in a model nickel-base superalloy: experi-
Ni3Cr(DO22), respectively. Chromium additions to Ni-Al alloys ments and simulations, Proc. TMS Annual Meet. (2003) 43e50.
[9] Z. Mao, C. Booth-Morrison, E. Plotnikov, D.N. Seidman, Effects of temperature
have significant effects on the critical radius and critical net and ferromagnetism on the gamma-Ni/gamma'-Ni3Al interfacial free energy
reversible work. The critical radius of Ni3(Al,Cr)(L12) is 9.23 Å from first principles calculations, J. Mater. Sci. 47 (21) (2012) 7653e7659.
for alloy B (Ni-5.2Al-14.2Cr at.%); 8.45 Å for alloy A (Ni-7.5Al- [10] Y. Mishin, Atomistic modeling of the gamma and gamma '-phases of the Ni-Al
system, Acta Mater. 52 (6) (2004) 1451e1467.
8.5Cr at.%); and 8.04 Å for alloy C (Ni-6.5Al-9.5Cr at.%). The [11] C. Wolverton, A. Zunger, Magnetic destabilization of Ni7Al, Phys. Rev. B 59
critical net reversible work required for Ni3(Al,Cr)(L12) is (19) (1999) 12165e12168.
1.12 meV (1.79x10-22 J) for alloy B (Ni-5.2Al-14.2Cr at.%); [12] D.L. Price, B.R. Cooper, Full-potential LMTO calculation of Ni/Ni3Al interface
energies, Mater. Theory, Simulat. Parallel Algorithm 408 (1996) 463e468.
1.75 meV (2.81x10-22 J) for alloy A (Ni-7.5Al-8.5Cr at.%); and [13] Y. Amouyal, Z.G. Mao, D.N. Seidman, Segregation of tungsten at gam-
2.04 meV (3.27x10-22 J) for alloy C (Ni-6.5Al-9.5Cr at.%). As the ma(')(L1(2))/gamma(fcc) interfaces in a Ni-based superalloy: an atom-probe
Cr concentration increases, from 7.5 to 14.2 at. %, Ni3(Al,Cr)(L12) tomographic and first-principles study, Appl. Phys. Lett. 93 (20) (2008)
201905.
becomes easier to nucleate.
[14] C.K. Sudbrack, K.E. Yoon, Z. Mao, R.D. Noebe, D. Isheim, D.N. Seidman, Tem-
 The predicted morphologies of L12-ordered precipitates is based poral evolution of nanostructures in a model nickel-base superalloy: experi-
on Wulff constructions using the NIST program Wulffman. The ments and simulations, in: J.R. Weertman, M.E. Fine, K.T. Faber, W. King,
g0 -Ni3Al(L12) precipitate has the morphology of a great rhom- P. Liaw (Eds.), Electron Microscopy: its Role in Materials Science, the Mike
Meshii Symposium, TMS (The Minerals, Metals & Materials Society), San
bicuboctahedron (truncated cuboctahedron), while Ni3Cr(L12)’s Diego, CA, 2003, pp. 43e50.
morphology is spherical: the {hkl}-interfacial Gibbs free en- [15] D. Blavette, B. Deconihout, A. Bostel, J.M. Sarrau, M. Bouet, A. Menand, The
ergies were calculated based on first-principles calculations. The tomographic atom-probe - a quantitative 3-dimensional nanoanalytical in-
strument on an atomic-scale, Rev. Sci. Instrum. 64 (10) (1993) 2911e2919.
morphology of the Ni3(Al,Cr)(L12) precipitates transforms from [16] A. Cerezo, T.J. Godfrey, S.J. Sijbrandij, G.D.W. Smith, P.J. Warren, Performance
cuboidal-to-spheroidal as a result of Cr additions to Ni-Al alloys. of an energy-compensated three-dimensional atom probe, Rev. Sci. Instrum.
714 Z. Mao et al. / Acta Materialia 166 (2019) 702e714

69 (1) (1998) 49e58. tantalum site substitution patterns in Ni(3)Al (L1(2)) gamma '-precipitates,
[17] S.S. Bajikar, D.J. Larson, T.F. Kelly, P.P. Camus, Magnification and mass reso- Appl. Phys. Lett. 93 (3) (2008).
lution in local-electrode atom probes, Ultramicroscopy 65 (1e2) (1996) [46] X.Q. Guo, R. Podloucky, A.J. Freeman, 1st principles calculation of the elastic-
119e129. constants of intermetallic compounds - metastable Al3li, J. Mater. Res. 6 (2)
[18] T.F. Kelly, P.P. Camus, D.J. Larson, L.M. Holzman, S.S. Bajikar, On the many (1991) 324e329.
advantages of local-electrode atom probes, Ultramicroscopy 62 (1e2) (1996) [47] V. Ozolins, C. Wolverton, A. Zunger, Cu-Au, Ag-Au, Cu-Ag, and Ni-Au in-
29e42. termetallics: first-principles study of temperature-composition phase dia-
[19] T.F. Kelly, D.J. Larson, Local electrode atom probes, Mater. Char. 44 (1/2) grams and structures, Phys. Rev. B 57 (11) (1998) 6427e6443.
(2000) 59e85. [48] A. Zunger, S.H. Wei, L.G. Ferreira, J.E. Bernard, Special Quasirandom Structures,
[20] D.N. Seidman, Three-dimensional atom-probe tomography: advances and Phys. Rev. Lett. 65 (3) (1990) 353e356.
applications, Annu. Rev. Mater. Res. 37 (2007) 127e158. [49] J. von Pezold, A. Dick, M. Friak, J. Neugebauer, Generation and performance of
[21] D.N. Seidman, K. Stiller, A renaissance in atom-probe tomography (vol 44, pg special quasirandom structures for studying the elastic properties of random
1034, 1973), MRS Bull. 34 (12) (2009), 892-892. alloys: application to Al-Ti, Phys. Rev. B 81 (9) (2010).
[22] D.N. Seidman, K. Stiller, An atom-probe tomography primer, MRS Bull. 34 (10) [50] C. Jiang, C. Wolverton, J. Sofo, L.Q. Chen, Z.K. Liu, First-principles study of bi-
(2009) 717e724. nary bcc alloys using special quasirandom structures, Phys. Rev. B 69 (21)
[23] Y. Zhou, C. Booth-Morrison, D.N. Seidman, On the field evaporation behavior (2004).
of a model Ni-Al-Cr superalloy studied by picosecond pulsed-laser atom- [51] D.W. Shin, R. Arroyave, Z.K. Liu, A.V. de Walle, Thermodynamic properties of
probe tomography, Microsc. Microanal. 14 (6) (2008) 571e580. binary hcp solution phases from special quasirandom structures (vol 74, art
[24] C.K. Sudbrack, R.D. Noebe, D.N. Seidman, Direct observations of nucleation in a no 024204, 2006), Phys. Rev. B 76 (6) (2007).
nondilute multicomponent alloy, Phys. Rev. B 73 (21) (2006), 212101/1- [52] A.M. Saitta, S. de Gironcoli, S. Baroni, Structural and electronic properties of a
212101/4. wide-gap quaternary solid solution: (Zn,Mg)(S,Se), Phys. Rev. Lett. 80 (22)
[25] C.K. Sudbrack, K.E. Yoon, R.D. Noebe, D.N. Seidman, Temporal evolution of the (1998) 4939e4942.
nanostructure and phase compositions in a model Ni-Al-Cr alloy, Acta Mater. [53] A.v.d. Walle, https://www.brown.edu/Departments/Engineering/Labs/avdw/
54 (12) (2006) 3199e3210. atat/https://www.brown.edu/Departments/Engineering/Labs/avdw/atat/.).
[26] K.E. Yoon, C.K. Sudbrack, R.D. Noebe, D.N. Seidman, The temporal evolution of [54] M. Asta, S.M. Foiles, A.A. Quong, First-principles calculations of bulk and
the nanostructures of model Ni-Al-Cr and Ni-Al-Cr-Re superalloys, interfacial thermodynamic properties for fcc-based Al-Sc alloys, Phys. Rev. B
Z. Metallkd. 96 (5) (2005) 481e485. 57 (18) (1998) 11265e11275.
[27] C.K. Sudbrack, Decomposition Behavior in Model Nickel-Aluminum- [55] M. Asta, V. Ozolins, Structural, vibrational, and thermodynamic properties of
Chromium-X Superalloys: Temporal Evolution and Compositional Pathways Al-Sc alloys and intermetallic compounds, Phys. Rev. B 6409 (9) (2001).
on a Nanoscale, Northwestern University, arc.nucapt.northwestern.edu, 2004, [56] E.A. Marquis, D.N. Seidman, Nanoscale structural evolution of Al3Sc pre-
p. 209. cipitates in Al(Sc) alloys, Acta Mater. 49 (11) (2001) 1909e1919.
[28] K.E. Yoon, Temporal Evolution of the Chemistry and Nanostructure of Multi- [57] K. Parlinski, http://wolf.ifj.edu.pl/phonon/).
component Model Nickel-Based Superalloys, Northwestern University, 2004, [58] A. van de Walle, G. Ceder, The effect of lattice vibrations on substitutional
p. 189, arc.nucapt.northwestern.edu. alloy thermodynamics, Rev. Mod. Phys. 74 (1) (2002) 11e45.
[29] C. Booth-Morrison, Y. Zhou, R.D. Noebe, D.N. Seidman, On the nanometer scale [59] A.J. Ardell, Trans-interface-diffusion-controlled coarsening of gamma ' pre-
phase separation of a low-supersaturation Ni-Al-Cr alloy, Philos. Mag. A 90 cipitates in ternary Ni-Al-Cr alloys, Acta Mater. 61 (20) (2013) 7828e7840.
(1e4) (2010) 219e235. [60] A.J. Ardell, V. Ozolins, Trans-interface diffusion-controlled coarsening, Nat.
[30] C. Booth-Morrison, Nanoscale Studies of the Early Stages of Phase Separation Mater. 4 (4) (2005) 309e316.
in Model Ni-Al-Cr Superalloys, PhD. thesis, Northwestern University, June [61] Q.F. Zhang, S.K. Makineni, J.E. Allison, J.C. Zhao, Effective evaluation of inter-
2009. arc.nucapt.northwestern.edu 196. facial energy by matching precipitate sizes measured along a composition
[31] O. Hellman, J. Vandenbroucke, J.B. du Rivage, D.N. Seidman, Application gradient with Kampmann-Wagner numerical (KWN) modeling, Scripta Mater.
software for data analysis for three-dimensional atom probe microscopy, Mat. 160 (2019) 70e74.
Sci. Eng. A-Struct. 327 (1) (2002) 29e33. [62] F. Soisson, G. Martin, Kinetics of decomposition of metastable solid solutions:
[32] O.C. Hellman, J.A. Vandenbroucke, J. Rusing, D. Isheim, D.N. Seidman, Analysis comparison between Monte Carlo simulations and classical theories, J. Jpn.
of three-dimensional atom-probe data by the proximity histogram, Microsc. Inst. Metals 12 (1999) 757e760.
Microanal. 6 (5) (2000) 437e444. [63] F. Soisson, G. Martin, Monte Carlo simulations of the decomposition of
[33] Z.G. Wu, R.E. Cohen, More accurate generalized gradient approximation for metastable solid solutions: transient and steady-state nucleation kinetics,
solids, Phys. Rev. B 73 (23) (2006). Phys. Rev. B 62 (1) (2000) 203e214.
[34] J.P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made [64] L. Ratke, P.W. Voorhees, Growth and Coarsening: Ostwald Ripening in Ma-
simple, Phys. Rev. Lett. 77 (18) (1996) 3865e3868. terials Processing, Springer-Verlag, 2002.
[35] G. Kresse, Ab initio molecular dynamics applied to the dynamics of liquid [65] E. Clouet, M. Nastar, C. Sigli, Nucleation of Al3Zr and Al3Sc in aluminum al-
metals and to the metal-nonmetal transition, J. Non-Cryst. Solids 205e207 (Pt. loys: from kinetic Monte Carlo simulations to classical theory, Phys. Rev. B 69
2) (1996) 833e840. (6) (2004).
[36] G. Kresse, J. Furthmueller, Efficient iterative schemes for ab initio total-energy [66] C.K. Sudbrack, R.D. Noebe, D.N. Seidman, Direct observations of nucleation in a
calculations using a plane-wave basis set, Phys. Rev. B Condens. Matter 54 nondilute multicomponent alloy, Phys. Rev. B 73 (21) (2006).
(16) (1996) 11169e11186. [67] Z. Mao, C.K. Sudbrack, K.E. Yoon, G. Martin, D.N. Seidman, The mechanism of
[37] G. Kresse, J. Furthmuller, Efficiency of ab-initio total energy calculations for morphogenesis in a phase separating concentrated multi-component alloy,
metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci. Nat. Mater. 6 (2006) 210e216.
6 (1) (1996) 15e50. [68] E. Clouet, M. Nastar, C. Sigli, Nucleation of Al3Zr and Al3Sc in aluminum al-
[38] G. Kresse, J. Hafner, Ab initio molecular dynamics of liquid metals, Phys. Rev. loys: from kinetic Monte Carlo simulations to classical theory, Phys. Rev. B 69
B: Condens. Matter Mater. Phys. 47 (1) (1993) 558e561. (6) (2004) 064109.
[39] G. Kresse, J. Hafner, Ab initio molecular-dynamics simulation of the liquid- [69] A. Roosen, R. McCormack, W.C. Carter, Wulffman Package from NIST. http://
metal-amorphous-semiconductor transition in germanium, Phys. Rev. B: www.ctcms.nist.gov/wulffman/.
Condens. Matter Mater. Phys. 49 (20) (1994) 14251e14269. [70] E.Y. Plotnikov, Z.G. Mao, R.D. Noebe, D.N. Seidman, Temporal evolution of the
[40] G. Kresse, D. Joubert, From ultrasoft pseudopotentials to the projector gamma(fcc)/gamma '(L1(2)) interfacial width in binary Ni-Al alloys, Scripta
augmented-wave method, Phys. Rev. B 59 (3) (1999) 1758. Mater. 70 (2014) 51e54.
[41] M.T. Yin, M.L. Cohen, Microscopic theory of the phase transformation and [71] C.K. Sudbrack, K.E. Yoon, Z. Mao, R.D. Noebe, D. Isheim, D.N. Seidman, Tem-
lattice dynamics of silicon, Phys. Rev. Lett. 45 (12) (1980) 1004e1007. poral evolution of nanostructures in a model nickel-base superalloy: experi-
[42] K.M. Ho, C.L. Fu, B.N. Harmon, W. Weber, D.R. Hamann, Vibrational fre- ments and simulations, in: J.R. Weertman, M.E. Fine, K.T. Faber, W. King,
quencies and structural properties of transition metals via total-energy cal- P. Liaw (Eds.), Electron Microscopy: its Role in Materials Research e the Mike
culations, Phys. Rev. Lett. 49 (9) (1982) 673e676. Meshii Symposium, TMS (The Minerals, Metals & Materials Society), War-
[43] P.K. Lam, M.L. Cohen, Ab initio calculation of phonon frequencies of rendale, PA, 2003, pp. 43e50.
aluminum, Phys. Rev. B Condens. Matter 25 (10) (1982) 6139e6145. [72] C.K. Sudbrack, K.E. Yoon, Z. Mao, R.D. Noebe, D. Isheim, D.N. Seidman, Tem-
[44] C.J. Kuehmann, P.W. Voorhees, Ostwald ripening in ternary alloys, Metall. poral evolution of nanostructures in a model nickel-base superalloy: experi-
Mater. Trans.: Phys. Metall. Mater. Sci. 27A (4) (1996) 937e943. ments and simulations, electron microscopy: its role in materials, Science
[45] C. Booth-Morrison, Z.G. Mao, R.D. Noebe, D.N. Seidman, Chromium and (2003) 43e50.

Вам также может понравиться