+ All Categories
Home > Documents > NORTHWESTERN UNIVERSITY Nanoscale Studies of the...

NORTHWESTERN UNIVERSITY Nanoscale Studies of the...

Date post: 06-Nov-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
196
NORTHWESTERN UNIVERSITY Nanoscale Studies of the Early Stages of Phase Separation in Model Ni-Al-Cr-X Superalloys A DISSERTATION SUBMITTED TO THE GRADUATE SCHOOL IN PARTIAL FULFILLMENT OF THE REQUIREMENTS For the degree DOCTOR IN PHILOSOPHY Field of Materials Science and Engineering By Christopher Booth-Morrison EVANSTON, ILLINOIS June, 2009
Transcript
Page 1: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

NORTHWESTERN UNIVERSITY

Nanoscale Studies of the Early Stages of Phase Separation in Model Ni-Al-Cr-X Superalloys

A DISSERTATION

SUBMITTED TO THE GRADUATE SCHOOL IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

For the degree

DOCTOR IN PHILOSOPHY

Field of Materials Science and Engineering

By

Christopher Booth-Morrison

EVANSTON, ILLINOIS

June, 2009

Page 2: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

2

© Copyright Christopher Booth-Morrison 2009

All Rights Reserved

Page 3: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

3

Abstract

Nanoscale Studies of the Early Stages of Phase Separation in Model Ni-Al-Cr-

X Superalloys

Christopher Booth-Morrison

The phase separation of model Ni-Al-Cr-X alloys is studied at the nanoscale employing

atom-probe tomography (APT), electron microscopy and first-principles calculations.

A comparison of the kinetic pathways resulting from the formation of coherent γ’-precipitates in

two Ni-Al-Cr alloys, Ni-7.5 Al-8.5 Cr and Ni-5.2 Al-14.2 Cr at.%, with similar γ’-precipitate

volume fractions at 873 K, is performed. The morphologies of the γ’-precipitates of the alloys are

similar, though the degrees of γ’-precipitate coagulation and coalescence differ. Quantification

within the framework of classical nucleation theory reveals that differences in the chemical

driving forces for phase decomposition result in differences in the nucleation behavior of the two

alloys. The temporal evolution of the γ’-precipitate average radii and the γ-matrix

supersaturations follow the predictions of classical coarsening models. The compositional

trajectories of the γ-matrix phases of the alloys are found to follow approximately the

equilibrium tie-lines, while the trajectories of the γ’-precipitates do not, resulting in significant

differences in the partitioning ratios of the solute elements.

Phase separation in a Ni-6.5 Al-9.5 Cr at.% alloy aged at 873 K occurs in four distinct

Page 4: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

4

regimes: (i) quasi-stationary-state γ’ (L12)-precipitate nucleation; (ii) concomitant precipitate

nucleation, growth, and coagulation and coalescence; (iii) concurrent growth and coarsening,

wherein coarsening occurs via both γ’-precipitate coagulation and coalescence and by the

classical evaporation-condensation mechanism; and (iv) quasi-stationary-state coarsening of γ’-

precipitates, once the equilibrium volume fraction of precipitates is achieved. The predictions of

classical nucleation and growth models are not validated experimentally, likely due to the

complexity of the atomistic kinetic pathways involved in precipitation. During coarsening, the

temporal evolution of the γ’-precipitate average radius, number density, and the γ(FCC)-matrix

and γ’-precipitate supersaturations follow the predictions of classical models.

The γ’(L12)-precipitate morphology of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at

1073 K is found to evolve from a bimodal distribution of spheroidal precipitates, to {001}-

faceted cuboids and parallelepipeds aligned along the elastically soft <001>-type directions. The

phase compositions and the widths of the γ’-precipitate/γ-matrix heterophase interfaces evolve

temporally as the Ni-Al-Cr-Ta alloy undergoes quasi-stationary state coarsening after 1 h of

aging. Tantalum is observed to partition preferentially to the γ’-precipitate phase, and suppresses

the mobility of Ni in the γ-matrix sufficiently to cause an accumulation of Ni on the γ-matrix side

of the γ'/γ interface. Measurements of the γ’-phase composition suggest that Al, Cr and Ta share

the Al sublattice sites of the γ’-precipitates. The calculated substitutional energies of the solute

atoms at the Ni and Al sublattice sites, from first-principles calculations, indicate that Ta has a

strong preference for the Al site, while Cr has a weak Al site preference.

Page 5: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

5

Acknowledgements

I would like to thank my advisor, Professor David N. Seidman, for his support and

encouragement, and for providing a project which was both scientifically challenging, and a

pleasure to work on. I am grateful to Dr. Ronald D. Noebe at NASA Glenn Research Center for

his technical guidance and critical evaluation of the work as it progressed, and for providing

specimens and sample preparation. I would also like to thank Professors Peter W. Voorhees and

David C. Dunand, and Dr. Georges Martin for their advice and technical guidance.

I am very grateful to a number of people who have helped me with both experimental and

analytical work, and whose contributions to the research can be found throughout this thesis,

namely:

• Yang Zhou, for his invaluable assistance with LEAPTM tomography, and for all the lively

debates.

• Dr. Zugang Mao for his computational work, and for all of his help and good cheer.

• Drs. Chantal Sudbrack and Kevin Yoon, who pioneered the work on nickel-based

superalloys in the Seidman group, and provided a great foundation of knowledge and

advice. I would also like to thank Chantal for mentoring me during my first years at NU,

and for her constant support and encouragement.

• Professor Dieter Isheim, for all of his assistance and advice, and for keeping NUCAPT in

top shape.

• Drs. Mark Anderson and Yaron Amouyal for their assistance with SEM and TEM

microscopy, respectively, and for all their other contributions to this work.

• Dr. Carelyn E. Campbell at the National Institute of Standards and Technology, who

Page 6: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

6

graciously calculated diffusivities for a wide variety of alloy compositions and aging

temperatures.

• Prof. G. B. Olson and Dr. H.-J. Jou of QuesTek LLC (Evanston, IL) for use of

PrecipiCalc.

• Dr. Kathleen Stair for her help with metallography, and her patience.

• Undergraduate research assistants Jessica Weninger and Gillian Hsieh, whose work is

included herein.

• The past and present members of the Seidman group for all their help over the years.

This research was sponsored by the National Science Foundation under grants DMR-

9728986, DMR-0241928 and DMR-0804610. I am grateful to the Fonds québécois de la

recherche sur la nature et les technologies for partial support, and to The Graduate School at

Northwestern University for providing a Walter P. Murphy graduate fellowship during my first

year of studies, as well as a B.J. Martin dissertation year fellowship in my last year.

Atom-probe tomographic measurements were performed at the Northwestern University

Center for Atom-probe Tomography (NUCAPT). The LEAPTM tomograph was purchased with

funding from the NSF-MRI (DMR 0420532, Dr. C. Bouldin, grant officer) and ONR-DURIP

(N00014-0400798, Dr. J. Christodoulou, grant officer) programs, and enhanced with a

picosecond laser with funding from the ONR-DURIP (N00014-0610539, Dr. J. Christodoulou,

grant officer). The SEM and TEM studies were performed in the EPIC facility of the NUANCE

Center at Northwestern University. The NUANCE Center is supported by NSF-NSEC, NSF-

MRSEC, the Keck Foundation, the State of Illinois, and Northwestern University.

Page 7: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

7

List of acronyms and symbols

APT atom-probe tomography

DTA differential thermal analysis

GGA generalized gradient approximation

LDA local density approximation

LEAP local-electrode atom-probe

NUCAPT Northwestern Center for Atom-Probe Tomography

PAW projector augmented-wave potentials

PLAP pulsed-laser atom-probe tomography

SEM scanning electron microscopy

TEM transmission electron microscopy

VASP Vienna ab initio simulation package

AKV coarsening rate constant for Nv(t) from the KV model

>< )(, tC ffiγ far-field γ-matrix concentration of atomic species i

)(, ∞eqiC γ equilibrium γ-matrix concentration of each atomic species i

)(,' ∞eqiC γ equilibrium γ’-precipitate concentration of each atomic species i

D diffusivity

f fraction of coagulating and coalescing γ’-precipitates γ

jiG , partial derivatives of the molar Gibbs free-energy of the γ-matrix phase

Jeff effective nucleation rate

Jst stationary-state nucleation rate

kB Boltzmann’s constant

KKV coarsening rate constant for <R(t)> from the KV model

Kiγ’/γ partitioning ratio of each atomic species i

Kγ’ bulk modulus of the γ’-precipitate phase

m/Δm mass resolving power

Page 8: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

8

m/n mass-to-charge state ratio

N0 total number of possible nucleation sites per unit volume

Nv(t) number density of γ’-precipitates

Nv(t0) precipitate number density at the onset of quasi-stationary coarsening

pi magnitude of the partitioning between two phases

R radius of the γ’-precipitate/γ’-nucleus

R* radius of the critical γ’-nucleus

<R(t)> mean γ’-precipitate radius

<R(t0)> average precipitate radius at the onset of quasi-stationary coarsening

S/N signal to noise ratio

t0 time at the onset of quasi-stationary coarsening

tc time required to reach stationary-state coarsening

T absolute temperature in degrees Kelvin γ ′

mV average atomic volume per mole of the γ’-precipitate phase

WR net reversible work required to form a γ’-nucleus *

RW net reversible work required to form a critical nucleus

Z Zeldovich factor

β kinetic coefficient describing the rate of condensation of a single atom on the

critical nucleus

δ lattice parameter misfit

)(tCiγΔ supersaturation of element i in the γ-matrix phase

)(' tCiγΔ supersaturation of element i in the γ’-precipitate phase

ΔFch chemical free energy change on forming a γ’-nucleus

ΔFel elastic free energy change on forming a γ’-nucleus

ΔFnuc bulk free energy change on forming a γ’-nucleus

φ volume fraction of the γ’-precipitate phase

φeq equilibrium volume fraction of the γ’-precipitate phase

Page 9: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

9

γκ KVi, coarsening rate constant of the atomic species i for )(tCi

γΔ from the KV model

<λe-e> mean edge-to-edge interprecipitate spacing

μi chemical potential per atom of the bulk element i

μγ shear modulus of the γ-matrix phase

σ standard deviation

σγ/γ’ γ-matrix/γ’-precipitate interfacial energy

υγ specific volume of an atom in the γ-matrix

υ specific volume of an atom in the γ’-precipitate phase

Page 10: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

10

For my wife Agatha, my parents Donald and Pamela, and my sister Gloria, for all

their love and support.

Page 11: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

11

Table of contents

Abstract………………………………………………………………………………………….. 3

Acknowledgements…………………………………………………………………………...… 5

List of acronyms and symbols………………………………………………………………….. 7

Table of contents…………………………………………………………………………….… 11

List of figures………………………………………………………………………….……..... 15

List of tables…………………………………………………………………………..……..... 21

Chapter 1: Introduction………………………………………………………………………. 23

Chapter 2: Background……………………………………………………………………….. 27

2.1 An introduction to nickel-based superalloys………………………………… 27

2.1.1 Metallurgy of modern nickel-based superalloys…………………………28

2.1.2 Strengthening and deformation mechanisms in

nickel-based superalloys……………………………………………… 31

2.2. Atom-probe tomography for characterizing nanostructures…………….... 34

Chapter 3: Effects of solute concentrations on kinetic pathways in Ni-Al-Cr alloys...…… 35

Abstract……………………………………………………………………………. 35

3.1. Introduction………………………………………………………………….. 36

3.1.1 Classical nucleation theory……………………………………………. 38

3.1.2 Coarsening theory………………………………………………………. 40

3.2. Experimental………………………………………………………………… 42

3.3. Results…………………………………………………………………………. 44

3.3.1 Morphological development………………………………………….. 44

Page 12: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

12

3.3.2 Temporal evolution of the nanostructural properties of γ’-precipitates.. 48

3.3.2.1 Nucleation and growth of γ’-precipitates…………………………... 51

3.3.2.2 Growth and coarsening of γ’-precipitates……………………………. 52

3.3.3 Temporal evolution of the compositions of the γ and γ’-phases……… 54

3.4. Discussion…………………………………………………………………… 64

3.4.1 Effects of solute concentration on nucleation behavior………………. 64

3.4.2 Effects of Solute Concentration on the Coarsening Behavior………… 71

3.5. Summary and conclusions………………………………………………….… 74

Chapter 4: On the field evaporation behavior of a model Ni-Al-Cr superalloy studied by

picosecond pulsed-laser atom-probe tomography……………………………… 78

Abstract………………………………………………………………………….…. 78

4.1 Introduction………………………………………………………………….. 79

4.2 Materials and methods……………………………………………………… 82

4.3 Results and discussion………………………………………………………… 85

4.3.1 Mass spectra……………………………………………………………. 85

4.3.2 Compositional accuracy: Effects of laser energy…………………….. 92

4.3.3 Compositional accuracy: Effects of specimen base temperature……… 94

4.3.4 Analysis of events detected at mass-to-charge state

ratios of 100-300 amu…………………………………………….….. 97

4.4 Conclusions…………………………………………………………………….. 99

Chapter 5: On the nanometer scale phase decomposition of a low-supersaturation

Ni-Al-Cr alloy…………………………………………………………………… 101

Page 13: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

13

Abstract…………………………………………………………………………… 101

5.1 Introduction………………………………………………………………… 102

5.2 Experimental…………………………………………………………………. 104

5.3 Results and discussion……………………………………………………… 106

5.3.1 Morphological evolution…………………………………………….. 106

5.3.2 Temporal evolution of the nanostructural properties

of the γ’-precipitates…………………………………………………. 108

5.3.3 Temporal evolution of the compositions of

the γ and γ’ phase compositions…………………………….………. 116

5.4 Conclusions…………………………………………………………………. 124

Chapter 6: Effects of tantalum on the temporal evolution of a model

Ni-Al-Cr superalloy during phase decomposition……………………………. 127

Abstract…………………………………………………………………………. 127

6.1 Introduction………………………………………………………………… 128

6.2 Experimental……………………………………………….………………… 129

6.3 Results………………………………………………………………………… 132

6.3.1 Morphological development……………………………………...…… 133

6.3.2 Compositional evolution…………………………………………..….. 138

6.3.3 Partitioning of elemental species………………………………….….. 140

6.4 Discussion……………………………………………………………….……. 142

6.4.1 Morphological development…………………………………………... 142

6.4.2 Compositional evolution…………………………………………….… 148

Page 14: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

14

6.4.3 Comparison with other Ni-Al-Cr-X alloys…………………………….. 153

6.5 Summary and conclusions…………………………………………………. 155

Chapter 7: Chromium and tantalum site substitution patterns

in Ni3Al (L12) γ’-precipitates…………………………………………………… 158

Abstract…………………………………………………………………………… 158

7.1 Introduction…………………………………………….…………………….. 159

7.2 Experimental…………………………………………………………………. 160

7.3 Results and discussion……………………………………………………….. 161

7.4 Conclusions………………………………………………………………….. 165

Chapter 8: Future work…………………………………………………………………….. 166

References……………………………………………………………………………………. 169

Appendix 1: Recalculation of the Ni-Al-Cr interfacial free energy from the data of

Schmuck et al. ……………………………………………………………….… 190

Appendix 2: Temporal evolution of the phase compositions of

Ni-Al-Cr alloys at 873 K……………………………………………………… 192

Vita………………………………………………………………………………………….. 195

Page 15: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

15

List of figures

Figure 2.1. Arrangements of Ni (green) and Al (red) in the (a) disordered γ-matrix and (b) the ordered γ’-Ni3Al, L12, crystal structures. (Image courtesy of Dr. Y. Amouyal, Northwestern University)……………………………………………………………………………………. 30

Figure 3.1. A partial ternary phase diagram of the Ni-Al-Cr system at 873 K calculated using the Grand Canonical Monte Carlo simulation technique [113], showing the proximity of both alloys (A) Ni-7.5 Al- 8.5 Cr and (B) Ni-5.2 Al-14.2 Cr to the (γ + γ’) / γ solvus line. The tie-lines are drawn through the nominal compositions of the alloys and between the equilibrium phase compositions determined by extrapolation of APT concentration data to infinite time. Equilibrium solvus curves determined by Thermo-Calc [114], using databases for nickel-based superalloys due to Saunders [115] and Dupin et al. [116], are superimposed on the GCMC phase diagram for comparative purposes……………………………………………………………... 37

Figure 3.2. The temporal evolution of the γ’-precipitate nanostructure in Ni-7.5 Al-8.5 Cr aged at 873 K is displayed in a series of APT parallelepipeds. The parallelepipeds are 10 x 10 x 25 nm3

subsets of the analyzed volume and contain ~ 125,000 atoms. The γ/γ’ interfaces are delineated in red with 10.5% Al isoconcentration surfaces………………………………………………. 45

Figure 3.3. A subset of an APT micrograph of Ni-7.5 Al-8.5 Cr aged at 873 K for 1024 h, containing 350,000 atoms, with Al atoms in red, and Cr atoms in blue; Ni atoms are omitted for clarity. A γ’-precipitate with a radius of ~ 9 nm is delineated by the red 10.5% Al isoconcentration surface, and shows {110}-type superlattice planes with an interplanar spacing of 0.26 ± 0.03 nm……………………………………………………………………………… 46

Figure 3.4. A centered superlattice reflection dark–field image of spheroidal Ni3(AlxCr(1-x)) γ’-precipitates for a Ni-7.5 Al-8.5 Cr sample aged for 1024 h at 873 K, recorded near the [011] zone axis……………………………………………………………………………………………. 47

Figure 3.5. The temporal evolution of the value of the fraction of γ’-precipitates interconnected by necks, f, and the average interprecipitate edge-to-edge spacing, <λe-e>. The maximum value of f is 18±4 % corresponds to the minimum value of <λe-e> of 7±2 nm at an aging time of 1 h for alloy (A) Ni-7.5 Al-8.5 Cr. For alloy (B) Ni-5.2 Al-14.2 Cr, the minimum value of <λe-e> of 5.9±0.8 nm and the maximum value of f of 30±4 % coincide at an aging time of 4 h……….. 48

Figure 3.6. The temporal evolution of the γ’-precipitate volume fraction, φ, number density, Nv(t), and mean radius, <R(t)>, for (A) Ni-7.5 Al-8.5 Cr and (B) Ni-5.2 Al-14.2 Cr alloys aged at 873 K as determined by APT microscopy. The quantity <R(t)> is proportional to t1/3 as predicted by the Umantsev and Olson (UO) and Kuehmann and Voorhees (KV) models for isothermal coarsening in ternary alloys. The temporal dependence of the diminution of the quantity Nv(t) deviates from the t-1 prediction of the UO and KV models for both alloys…….. 50

Page 16: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

16

Figure 3.7. The composition profiles on either side of the heterophase γ-matrix/ γ’-precipitate interface for (A) Ni-7.5 Al-8.5 Cr aged at 873 K for aging times of 1/4, 1, 4, 1024 h. The phase compositions evolve temporally, as the γ-matrix becomes enriched in Ni and Cr and depleted in Al. The values of <R(t)> for these aging times are 1.00 ± 0.11 nm for 1/4 h, 1.24 ± 0.12 nm for 1 h, 1.70 ± 0.25 nm for 4 h and 8.30 ± 2.93 nm for 1024 h……………………………………… 54

Figure 3.8. The compositional trajectories of the γ-matrix and γ’-precipitate phases of alloys (A) Ni-7.5 Al-8.5 Cr and (B) Ni-5.2 Al-14.2 Cr as they evolve temporally, displayed on a partial Ni-Al-Cr ternary phase diagram at 873 K. The tie-lines are drawn through the nominal compositions of the alloys (squares) and between the experimentally determined equilibrium phase compositions. The trajectories of the γ-matrix phases in alloys (A) and (B) lie approximately on the experimental tie-lines. The trajectories of the γ'-precipitate phases do not lie along the tie-line, as predicted by the Kuehmann and Voorhees model of isothermal coarsening in ternary alloys [133]……………………………………………………………………………………. 56

Figure 3.9. The partitioning ratios, Kiγ’/γ

, of Al and Cr demonstrate that both alloys exhibit partitioning of Al to the γ’-precipitates and Cr to the γ-matrix. Partitioning is more pronounced in alloy (B), where the smaller Al and larger Cr concentration results in a smaller value of the γ-matrix Al solubility, and a larger γ-matrix Cr solubility………………………………………. 59

Figure 3.10. The magnitude of the values of the supersaturations, )(tCiγΔ , of Al and Cr in the γ-

matrix are smaller for alloy (A) Ni-7.5 Al-8.5 Cr than for alloy (B) Ni-5.2 Al-14.2 Cr. The magnitude of the )(tCi

γΔ values decrease as t-1/3 in the coarsening regimes for both alloys, as predicted by the Umantsev and Olson (UO) and Kuehmann and Voorhees (KV) models for isothermal quasi-stationary state coarsening in ternary alloys. The values of )(tCCr

γΔ are expressed as an absolute value because they are negative, and reflect a flux of Cr into the γ-matrix with increasing aging time…………………………………………………………….… 62

Figure 3.11. The supersaturation values of the γ’-precipitates, )(tCiγ ′Δ , reflect the chemical

compositions of the two alloys, as the value of )(tCAlγ ′Δ is larger in alloy (A) Ni-7.5 Al- 8.5 Cr,

which contains more Al, than in alloy (B) Ni-5.2 Al-14.2 Cr, while the inverse is true for Cr. The values of )(tCi

γ ′Δ decrease as approximately t-1/3 in the coarsening regimes for both alloys… 63

Figure 4.1. The steady-state dc voltage applied to the tip to maintain a specimen evaporation rate of 0.04 ions per pulse for a Ni-6.5 Al-9.5 Cr at.% alloy is shown as a function of: (a) the laser pulse energy applied in PLAP tomography at a specimen base temperature of 40±0.3 K and a pulse repetition rate of 200 kHz; and (b) the specimen base temperature used in pulsed-voltage APT at a pulse repetition rate of 200 kHz and a pulse fraction of ~20%..................................... 85

Figure 4.2. A comparison of the mass spectra obtained by PLAP tomography with laser pulse energies of: (a) 0.2 nJ per pulse at a tip voltage of 7.40-7.70 kV; (b) 1.0 nJ per pulse at a tip voltage of 8.0-8.3 kV; and (c) 2.0 nJ per pulse at a tip voltage of 6.55-6.85 kV. The frequencies

Page 17: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

17

of detected events are normalized to the number of events detected in the most populous m/n bin. The m/n ratios of the evaporating species change by over four orders of magnitude with increasing laser pulse energy. At a laser pulse energy of 2.0 nJ per pulse, the m/n spectrum is populated by numerous hydrides, oxides and metal ion clusters peaks, which are a result of overheating and are undesirable………………………………………………………………… 87

Figure 4.3. A section of the m/n range showing the Ni2+ peaks, highlighting both the improved mass resolving power of PLAP tomography when compared to pulsed-voltage APT, and the improved mass resolving power with increasing laser pulse energy. The relative frequencies of the Ni2+ peaks from the data set collected at a laser pulse energy of 2.0 nJ per pulse are nearly two orders of magnitude smaller than the relative frequencies of the Ni2+ peaks from the set collected at 0.2 nJ per pulse……………………………………………………………….……. 88

Figure 4.4. The ratio of the ions detected in the 2+ charge state to those detected in the 1+ charge state, as a function of laser pulse energy. The Ni and Al ions appear to have similar evaporation behavior, whereas Cr atoms evaporate exclusively as Cr2+ until the laser energy reaches 1.0 nJ per pulse………………………………………………………………………………………. 89

Figure 4.5. The relative frequencies of the Ni1+ peaks at specimen base temperatures of 125, 150 and 175±0.3 K, at a constant steady-state dc tip voltage of 5.2±0.2 kV. The increase in the relative frequency of Ni1+ ions with increasing specimen base temperature suggests that the evaporation of Ni1+ ions occurs by a different thermally activated mechanism than the evaporation of Ni2+ ions………………………………………………………………………. 92

Figure 4.6. Variation in the detected composition of as-quenched Ni-6.5 Al-9.5 Cr at.% as a function of laser pulse energy (solid line) compared to the concentration measured by pulsed-voltage APT (dashed line) and the nominal composition of the alloy (dotted line). Preferential evaporation of Ni is responsible for the inaccuracy in the concentrations measured by pulsed-laser and pulsed-voltage APT………………………………………………………………….. 95

Figure 4.7. Variation of the detected composition of as-quenched Ni-6.5 Al-9.5 Cr at.% as a function of specimen base temperature by pulsed-voltage APT. At higher temperatures, preferential evaporation of Ni is less severe, while preferential evaporation of Cr is more severe.. ……………………………………………………………………………………………………96

Figure 4.8. The relative frequency of detected events between 100-300 amu, as a function of: (a) laser pulse energy in PLAP tomography; and (b) specimen base temperature in pulsed-voltage APT. The amount of noise measured by PLAP tomography is directly proportional to the tip voltage, and does not increase with increasing laser pulse energy, and thus increasing thermal energy. On the contrary, the noise measured by pulsed-voltage APT increases as the specimen base temperature increases. This result demonstrates that the Ni-Al-Cr tips are not affected by significant steady-state heating during PLAP tomography……………………………………. 98

Figure 5.1. A partial ternary phase diagram of the Ni-Al-Cr system at 873 K calculated using the Grand Canonical Monte Carlo simulation technique [113], showing the proximity of Ni-6.5 Al-

Page 18: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

18

9.5 Cr at.% to the (γ + γ’) / γ solvus line. Two other alloys that have been investigated by APT, Ni-7.5 Al- 8.5 Cr and Ni-5.2 Al-14.2 Cr at.%, are shown for comparative purposes. Equilibrium solvus curves determined by Thermo-Calc [114], using databases for nickel-based superalloys due to Saunders [115] and Dupin et al. [116], are superimposed on the GCMC phase diagram. The tie-lines are determined from the equilibrium phase compositions, determined by extrapolation of APT concentration data to infinite time……………………………………. 103

Figure 5.2. APT reconstructed 3D images of a Ni-6.5 Al-9.5 Cr at.% alloy aged at 873 K for (a) 1 h, (b) 4 h, (c) 64 h and (d) 4096 h. The nanometer-sized γ’-precipitates are delineated with red Al isoconcentration surfaces. Al, which partitions to the γ’-precipitates, is shown in red, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity……. 107

Figure 5.3. The fraction of γ’-precipitates undergoing coagulation and coalescence, f, is significant for aging times of 0.5 to 16 h. After aging for 64 h, only 4±1 % of γ’-precipitates are undergoing coagulation and coalescence, and therefore coarsening proceeds primarily via the evaporation-condensation mechanism. No γ’-precipitate coagulation and coalescence is detected beyond an aging time of 64 h. ……………………………………………………………..…. 108

Figure 5.5. The temporal evolution of the γ’-precipitate volume fraction, φ, number density, Nv(t), and mean radius, <R(t)>, for Ni-6.5 Al-9.5 Cr at.% at 873 K. The quantity <R(t)> is approximately proportional to t1/3 during quasi-stationary state coarsening for aging times of 4 h and longer, as predicted by classical coarsening models. Once the equilibrium volume fraction is achieved after 256 h, the temporal dependence of the quantity Nv(t) achieves the t-1 prediction of the coarsening models…………………………………………………………………………. 109

Figure 5.6. The composition profiles across the γ-matrix/ γ’-precipitate interface for Ni-6.5 Al-9.5 Cr at.% at 873 K for aging times of 1, 4, 4096 h. The phase compositions evolve temporally, as the γ-matrix becomes enriched in Ni and Cr and depleted in Al. The values of <R(t)> for these aging times are 0.65 ± 0.19 nm for 1 h, 1.23 ± 0.43 nm for 4 h and 15.80 ± 3.39 nm for 4096 h………………………………………………………………………………………………. 119

Figure 5.7. The compositional trajectories of the γ-matrix and γ’-precipitate phases of Ni-6.5 Al-9.5 Cr at.%, displayed on a partial Ni-Al-Cr ternary phase diagram at 873 K. The trajectory of the γ-matrix phase lies approximately on the experimental tie-line, while the trajectory of the γ'-precipitate phases does not lie along the tie-line, as predicted by the Kuehmann-Voorhees coarsening model [133]. The tie-lines are determined from the equilibrium phase compositions, determined by extrapolation of APT concentration data to infinite time……………………. 121

Figure 5.8. The partitioning ratios, Kiγ’/γ

, quantify the partitioning of Al to the γ’-precipitates and of Ni and Cr to the γ-matrix. The values of Ki

γ’/γ are constant from 256 to 4096 h, when the alloy is undergoing quasi-stationary state coarsening………………………………………………. 122

Page 19: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

19

Figure 5.9. The magnitude of the values of the γ-matrix (left) and γ’-precipitate (right) supersaturations, )(tCi

γΔ and )t(C 'iγΔ , of Ni, Al and Cr, decrease as t-1/3 in the coarsening

regime, as predicted by classical coarsening models…………………………………………. 123

Figure 6.1. Vickers microhardness measurements for Ni-10.0 Al-8.5 Cr-2.0 Ta and Ni-10.0 Al-8.5 Cr at.% aged at 1073 K. The addition of 2.0 at.% Ta results in a 47±5% increase in the microhardness over the full range of aging times, t = 0 to 256 h, due to increases in solid-solution strengthening and to a dramatic increase in the value of φ.………………………………….. 132

Figure 6.2. APT reconstructed 3D image of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 0 h. The solute elements that partition to the γ’-precipitates, Ta and Al, are shown in orange and red, respectively, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity. The morphology of the γ’-precipitate phase is spheroidal in the as-quenched state, and a bimodal particle size distribution is apparent………………………….134

Figure 6.3. SEM images of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for: (a) 0 h: (b) 0.25 h: (c) 1 h: (d) 4 h: (e) 16 h: and (f) 64 h. The primary γ’-precipitate morphology evolves from spheroidal γ’-precipitates to a cuboidal and parallelepipedic morphology with primary γ’-precipitates aligned along the elastically soft <001>-type directions……………………….. 135

Figure 6.4. APT reconstructed 3D image of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 0.25 h. The primary γ’-precipitates in the APT reconstruction have both spheroidal and cuboidal characteristics. The radius at which the γ’-precipitates undergo the spheroidal-to- cuboidal morphological transformation is measured to be 61±7 nm. Tantalum and Al are shown in orange and red, respectively, Cr is shown in blue, and Ni atoms are omitted for clarity… 136

Figure 6.5. An APT reconstructed 3D image of cuboidal and parallelepipedic primary γ’-precipitates aligned along <001>-type directions from a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 64 h. Tantalum and Al are shown in orange and red, respectively, Cr is shown in blue, and Ni atoms are omitted for clarity………………………………………………….. 137

Figure 6.6. After aging for 256 h, the primary γ’-precipitates have cuboidal and parallelepipedic morphologies with {001}-type facets, and are aligned along the <001>-type crystallographic directions, as confirmed by a selected-area electron diffraction pattern taken along the ]101[ zone axis…………………………………………………………………………………………….. 137

Figure 6.7. The elemental concentration profiles across the primary γ’-precipitate/γ-matrix heterophase interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The phase compositions evolve temporally, and the widths of the concentration profiles decrease with increasing aging time. An accumulation of Ni is observed to develop on the γ-matrix side of the interface, evidence of a kinetic effect associated with the addition of Ta. The concentration profiles shown here represent the average of all of the interfaces in the analyzed volume….. 139

Page 20: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

20

Figure 6.8. The temporal evolution of the widths of the concentration profiles across the primary γ’-precipitate/γ-matrix interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The interfacial widths are observed to decrease with increasing aging time. The species with the largest diffusivity in the γ-matrix, Al and Ta, have the smallest interfacial widths, while the opposite is true for Ni and Cr, the slower diffusing species…………………………………. 140

Figure 6.9. The temporal evolution of the partitioning ratios, γγ /′iK , of the constituent elements i

of: (a) Ni-10.0 Al-8.5 Cr-2.0 Ta; and (b) Ni-10.0 Al-8.5 Cr at.% aged at 1073 K. In both alloys, Al is observed to partition to the γ’-precipitates, while Cr and Ni partition to the γ-matrix. Ta shows a strong preference for the γ’-phase……………………………………………………. 141

Figure 6.10. The values of the γ’-precipitate number density, Nv(t ), radius, <R(t )>, and volume fraction, φ, during the water quench from 1503 K for Ni-10.0 Al-8.5 Cr-2.0 Ta at.%, as modeled by PrecipiCalc. A bimodal distribution of γ’-precipitates is predicted for the as-quenched state with primary γ’-precipitates with <R(t)> and φ values of 30.2 nm and 31.5 %, and secondary γ’-precipitates with <R(t)> and φ values of 0.8 nm and 14.2 %. The predicted values of Nv(t) for the primary and secondary γ’-precipitates are 2.7 x 1021 and 3.4 x 1025 m-3, respectively………. 145

Figure 6.11. The evolution of the values of the bulk driving force for nucleation, ΔFnuc, the effective nucleation rate, Jeff, and the critical radius for nucleation, R*, during the quench. Decreases in the magnitude of ΔFnuc are predicted at 1380 and 1190 K due to the nucleation of primary and secondary γ’-precipitates, respectively. The magnitude of ΔFnuc increases and the values of R* decrease during the quench. This figure was calculated using PrecipiCalc…….. 146

Figure 6.12. The magnitude of the values of the supersaturations, )(tCiΔ , of Ni, Al, Cr and Ta in the γ-matrix and γ’-precipitates for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The magnitudes of the )(tCiΔ values decrease as approximately t-1/3 in the coarsening regimes for both phases, as predicted by the Umantsev and Olson (UO) models for quasi-stationary state coarsening………………………………………………………………………………….….. 150

Figure 7.1. APT reconstruction of a cuboidal γ’-precipitate in a model Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 256 h. The elements that partition to the γ’-precipitates, Al and Ta, are shown in orange and red, respectively, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity…………………………………………………………. 160

Figure 7.2. The elemental concentration profiles across the γ’/γ interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 256 h. Tantalum and Al partition to the γ’-precipitates, while Cr partitions to the γ-matrix…………………………………………………………… 161

Page 21: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

21

List of tables

Table 3.1. Temporal evolution of the nanostructural properties of γ’-precipitates determined by APT for Ni-7.5 Al-8.5 Cr aged at 873 K…………………………………………..…………… 51 Table 3.2. Equilibrium γ’-precipitate and γ-matrix concentrations, as determined by atom-probe tomography (APT), Grand Canonical Monte Carlo (GCMC) simulation, and thermodynamic modeling employing Thermo-Calc for alloy (A) Ni-7.5 Al-8.5 Cr aged at 873 K……………. 58 Table 3.3. Equilibrium γ’-precipitate volume fraction, φeq, as determined by atom-probe tomography (APT), Grand Canonical Monte Carlo (GCMC) simulation, and Thermo-Calc for alloy (A) Ni-7.5 Al-8.5 Cr, and alloy (B) Ni-5.2 Al-14.2 Cr, aged at 873 K………………….. 60 Table 3.4. Curvatures in the molar Gibbs free-energy surface of the γ-matrix phase evaluated at the equilibrium composition with respect to components i and j, γ

jiG , , obtained from ideal solution theory and Thermo-Calc thermodynamic assessments for alloy (A) Ni-7.5 Al-8.5 Cr aged at 873 K…………………………………………………………………………………. 66

Table 3.5. Free-energy of the γ/γ’ interfaces, σγ/γ’, at 873 K in Ni–7.5 Al–8.5 Cr calculated from the experimental values of the Kuehmann-Voorhees coarsening rate constants for the average precipitate radius and the supersaturation of solute species i employing Equation 3.9 with solution thermodynamics described by the ideal solution and Thermo-Calc databases given in Table 3. 4……………………………………………………………………………………… 67 Table 3.6. The interfacial free energy, σγ/γ’, the chemical free energy, ΔFch, and the elastic strain energy, ΔFel, components of the driving force for nucleation used to estimate the net reversible work, *

RW , required for the formation of critical nuclei of size, R*, and the nucleation current, Jst, according to classical nucleation theory……………………………………………………… 69 Table 5.1. Temporal evolution of the nanostructural properties of γ’-precipitates a determined by APT for Ni-6.5 Al-9.5 Cr at.% aged at 873 K. The γ’-precipitate mean radius, <R(t)>, number density, Nv(t), and volume fraction, φ, are given, along with their standard errors…………… 110 Table 6.1. Equilibrium γ’-precipitate volume fraction, φeq, determined by APT, ICP chemical analysis, and thermodynamic modeling employing Thermo-Calc for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% aged at 1073 K…………………………………………………………………………… 133 Table 6.2. Equilibrium compositions (at.%) of the γ’-precipitate, )(C eq,'

i ∞γ , and γ-matrix, )(C eq,

i ∞γ , determined by APT, ICP chemical analysis, and thermodynamic modeling employing Thermo-Calc for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% aged at 1073 K…………………………….. 151 Table 6.3. Tracer diffusivity of element i in the γ-matrix, γ

iD , of Ni-10.0 Al-8.5 Cr and Ni-10.0

Page 22: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

22

Al-8.5 Cr-2.0 Ta at.% calculated with Dictra, and Campbell [208] and Saunders [115] databases at 1073 K…………………………………………………………………………………….. 152

Table 7.1. Total, Etot, and substitutional, EZ→Ni, Al, energies determined by first-principles calculations. (For these calculations x = 0.042, y = 0.125, and Z = Cr, Ta.)……………….…. 164 Table 7.2. Average atomic forces and displacements at the first nearest-neighbor distance from first-principles calculations for four different substitutional structures……………………….. 165 Table A.1.1 Equilibrium γ’-precipitate and γ-matrix equilibrium concentrations, as determined by atom- probe tomography (APT) and Thermo-Calc for Ni-5.2 Al-14.8 Cr at.% aged at 873 K from the data of Schmuck et al. [28] .……………………………………………………….. 191 Table A.2.1. Temporal evolution of the phase compositions in Ni-7.5 Al-8.5 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter for this alloy is estimated to be 0.0027 ± 0.0004, and the equilibrium volume fraction is 16.4 ± 0.6 %.……………………… 192 Table A.2.2. Temporal evolution of the phase compositions in Ni-5.2 Al-14.2 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter misfit for this alloy is estimated to be 0.0006 ± 0.0004 and the equilibrium volume fraction is 15.7 ± 0.7 %.…….. 193 Table A.2.3. Temporal evolution of the phase compositions in Ni-6.5 Al-9.5 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter misfit for this alloy is estimated, from PrecipiCalc, to be 0.00284, and the equilibrium volume fraction from APT, is 9.6 ± 2.9 %.………………………………………………………………………………….. 194

Page 23: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

23

Chapter 1

Introduction

Global warming and the volatility of fuel costs have amplified the urgency for increased

fuel efficiency in high-performance engines. Fuel efficiency is directly related to engine

operating temperature, which is tied to the high-temperature properties of the engine materials.

For example, a 200 K rise in the turbine engine entry temperature of a Boeing 777 yields a 5%

savings in fuel burn, which in turn equals a $12.5 million savings over 15 years. Due to their

excellent strength and resistance to both corrosion and creep-induced damage at temperatures up

to 1373 K, nickel-based superalloys are used for critical components of both aerospace and land-

based turbine engines.

Though the most technologically demanding applications of nickel-based superalloys

remain in the aerospace sector , these alloys are also used in land-based turbines for generating

electricity [1]. The U.S.A. consumes >4 trillion kWh of electrical power per year, of which ca.

70% comes from fossil fuels [2]. Decreasing CO2 emissions to avert global warming is crucial,

given the imminent dangers associated with climate change, and the ongoing industrial

expansion in developing nations. Nickel-based superalloys are also used in the chemical industry

because of their outstanding corrosion resistance, and in the oil and gas industry for coating pipes

and down-hole equipment to enhance resistance to hydrogen embrittlement due to H2S, and to

reduce high-temperature corrosion.

The nickel-based superalloys used in aerospace turbine blades have progressed from

equiaxed to directionally solidified grain-structures, and presently to single-crystal blades that

Page 24: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

24

contain 12 elements [3]. The current generation of nickel-based superalloys take advantage of the

sophisticated physical metallurgy of the γ/γ’ system and of the effects of alloying on the high-

temperature microstructure and physical and mechanical properties [4]. The addition of

refractory elements such as Ta, Mo, W, Re, and recently Ru [5], increases the solvus

temperatures, thereby increasing the maximum operating temperature [6].

The high-temperature strength and creep resistance of modern nickel-based superalloys

are due to the presence of coherent, elastically hard, L12-ordered γ’-precipitates in a γ (FCC)

nickel-rich solid-solution matrix [7, 8]. Efforts to improve these properties by process

optimization, and to develop reliable life-prediction techniques, have created a demand for a

quantitative understanding of the kinetic pathways that lead to phase decomposition at service

temperatures up to 1373 K [3, 9, 10]. The technological importance of commercial nickel-based

superalloys has motivated extensive investigations of the precipitation of the γ’-phase from the

supersaturated γ-matrix of model alloys by conventional [11-15] and high–resolution [16]

transmission electron microscopy (TEM), x-ray analysis [17-19], small-angle and wide angle

neutron scattering [20-23], atom-probe field-ion microscopy (APFIM) [24-27], atom-probe

tomography (APT) [28-30], phase-field modeling [31-36], and lattice kinetic Monte Carlo

simulation (LKMC) [37-39]. Many of these techniques are limited by either, or both, their spatial

and analytical resolutions for composition [40], and thus the early stages of phase decomposition

are still not well understood. This is particularly true of concentrated multicomponent alloys due,

in part, to the complexity of the diffusion processes involved [39].

Due to their excellent mechanical and physical properties at elevated temperatures,

nickel-based superalloys based on Ni-Al-Cr compositions are technologically extremely

Page 25: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

25

important alloys [3, 7, 41]. The addition of Cr to the binary Ni-Al system reduces the lattice

parameter misfit between the γ’-precipitates and the γ-matrix, often leading to γ’-precipitates that

are nearly misfit-free [12], thereby allowing the γ’-precipitates to remain spheroidal to fairly

large dimensions as aging progresses [17, 42]. The formation of coherent, spheroidal γ’-

precipitates with relatively stress-free precipitate/matrix heterophase interfaces in model Ni-Al-

Cr alloys makes them excellent candidates for comparison of experimental data with the

predictions of classical treatments of nucleation, growth and coarsening for ternary alloys, for

which there is only a little detailed quantitative data available.

This Ph.D. thesis focuses on the decomposition behavior of two model Ni-Al-Cr alloys,

and one model Ni-Al-Cr-Ta alloy. This work is part of a systematic investigation of the phase

decomposition of model Ni-Al-Cr alloys at 873 K [39, 42-54], and of the effects of dilute

refractory additions, such as W [43, 55, 56], Re [50, 51, 57-59], and Ta [60-63], on the γ/γ’ phase

decomposition of a model Ni-10.0 Al-8.5 Cr at.% alloy aged at 1073 K. The thesis is divided as

follows:

Chapter 2 gives a short background to nickel-based superalloys, including their basic

physical metallurgy and materials properties.

Chapter 3 presents an investigation of the effects of solute content on the kinetic

pathways that lead to phase decomposition in two Ni-Al-Cr alloys with moderate

solid-solution supersaturations (Ni-7.5 Al-8.5 Cr and Ni-5.2 Al-14.2 Cr at.%) aged at

873 K [52].

Chapter 4 describes the field evaporation behavior of a model Ni-Al-Cr superalloy

(Ni-6.5 Al-9.5 Cr at.%) studied by picosecond pulsed-laser atom-probe tomography

Page 26: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

26

[64].

Chapter 5 provides the details of the phase decomposition of a low-supersaturation

Ni-Al-Cr alloy (Ni-6.5 Al-9.5 Cr at.%) aged at 873 K, and compares the results to the

predictions of the classical models of nucleation, growth and coarsening [53].

Chapter 6 is a description of the effects of a dilute addition of Ta to a model Ni-Al-Cr

alloy (Ni-10.0 Al-8.5 Cr at.%) aged at 1073 K [60].

Chapter 7 is an investigation of the sublattice site preference of Cr and Ta in Ni3Al

(L12) γ’-precipitates employing both atom-probe tomography and first-principles

calculations [61].

Chapter 8 gives possible directions of future research on these subjects.

Page 27: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

27

Chapter 2

Background

2.1 An introduction to nickel-based superalloys

Considerable research and development has led to significant improvements in the

service temperatures of commercial superalloys since their first appearance in the 1940s. The

complex multicomponent alloys in service today are a result of both tremendous progress in

alloy development, and improvements in manufacturing techniques, such as the evolution of

alloys for blade applications from wrought, to polycrystalline to single crystal form.

Alloy development has evolved through four generations of superalloys. First generation

alloys contained high concentrations of the γ’-precipitate hardening elements Al, Ti and Ta, as

well as the grain boundary-strengthening additions C and B. The second and third generation

superalloys were characterized by 3 and 6 wt.% additions of Re, respectively, which provides

solid solution strengthening. Rhenium also significantly decreases the rate of γ’-precipitate

coarsening, extending the temperature capability of directionally solidified, columnar grain, and

single-crystal nickel superalloys. The fourth generation of superalloys, in use since 2000, is

characterized by additions of ruthenium, which inhibits the formation of the deleterious

topologically close-packed (TCP) phases [65-69]. These refractory-rich TCP phases evolve from

localized supersaturations of Cr, Re and W at temperatures above 1273 K [70-74]. The TCP

phases are deleterious to the mechanical properties of commercial superalloys because they

consume refractory additions, reducing both solid-solution and precipitation strengthening [71,

75]. In addition, Ruthenium improves the oxidation and creep resistance of nickel-based

Page 28: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

28

superalloys [76].

2.1.1 Metallurgy of modern nickel-based superalloys

Nickel has emerged as the premiere solvent for high-temperature applications. As a

transition metal, Ni offers both considerable toughness and ductility due to its high cohesive

energy, related to bonding of the outer d electrons. Nickel remains stable in the FCC structure up

to its melting point [77], providing phase stability that is critical to high-temperature

applications. In addition, the small diffusivity of Ni leads to excellent resistance to diffusion

mediated processes such as γ’-precipitate coarsening, creep, and oxidation [3, 78, 79]. Nickel

also offers considerable cost and weight benefits when compared to other suitable FCC metals,

such as the platinum group metals.

While some solid-solution strengthened nickel-based superalloys, such as Hastelloy X,

are used commercially when modest strength is required, the most demanding applications

require a class of superalloys that are strengthened by intermetallic compound precipitation in

the austenitic (FCC) γ-matrix. Generally, alloying additions in nickel-based superalloys can be

characterized as:

(1) γ-matrix formers, elements that partition to the FCC matrix (Ni, Co, Cr, Fe,

Mo, W, Re, Ru)

(2) γ’-precipitate formers (Al, Ti, Ta, Nb)

(3) γ’’-precipitate formers (Nb, V)

Page 29: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

29

(4) Carbide formers (reactive refractory additions such as Ti, Ta, Nb, Hf form

carbides that may decompose at temperature to form complex carbides with

Cr, Mo and W at γ-matrix grain boundaries)

(5) Boride formers (Cr and Mo form borides at γ-matrix grain boundaries)

(6) Grain boundary segregants (C, B, Zr)

(7) TCP phase formers (excessive additions of refractory additions such as W and

Re, as well as Cr, lead to the formation of these highly deleterious phases)

This work focuses exclusively on the γ → γ’ phase transformation in model Ni-Al-Cr-X

alloys, thus emphasis will be placed on these two phases, which provide the strengthening for

nickel alloys used in the most extreme environments, such as the hot sections of gas turbine

engines.

The binary Ni-Al phase diagram exhibits a number of solid phases, all of which show a

significant degree of directional covalent bonding, leading to precise stoichiometric

relationships, and crystal structures in which Ni-Al bonds are preferred, rather than Al-Al or Ni-

Ni bonds [3]. The Ni3Al γ’-phase has the primitive cubic L12 crystal structure, where Ni and Al

have distinct site occupancies, Figure 2.1. The L12 crystal structure can accommodate substantial

solute additions [3], and the mechanical properties of the γ’-phase depend on the sublattice site

substitution behavior of these alloying additions [80]. Considerable experimental and theoretical

work has demonstrated that Al, Ta, Ti, Nb prefer to occupy the Al sublattice sites, and Co, Mo,

Re, Ru, Pt prefer the Ni sublattice sites, while Cr, Fe and Mn have been shown to display mixed

behavior, depending on the alloy composition and temperature of interest [3].

Page 30: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

30

(a) (b)

Figure 2.1. Arrangements of Ni (green) and Al (red) in the (a) disordered γ-matrix and (b) the ordered γ’-Ni3Al, L12, crystal structures. (Image courtesy of Dr. Y. Amouyal, Northwestern University)

The γ/γ’ interface is of critical importance in nickel-based superalloys. Interfacial

coherency is critical in order to minimize γ’-precipitate coarsening. As such, the lattice parameter

misfit between the γ-matrix and γ’-precipitate phases must be controlled to prevent a loss of

coherency. The morphology of the γ’-precipitate phase has been demonstrated to be a function of

the Gibbs interfacial free energy and the elastic energy, which is directly related to the lattice

parameter misfit. The sign of the γ/γ' lattice parameter mismatch is important for the elevated

temperature creep strength; a mismatch of zero improves creep properties by preventing the

evolution of a γ’-rafted morphology [81, 82]. Nuclei of the γ’-precipitate phase have a spherical

morphology to minimize the elastic free energy, however as particle growth occurs, {001}-

faceted cuboids develop, which then form arrays, and eventually solid-state dendrites, as

coarsening proceeds [83]. The crystallographic orientation of the two phases corresponds to the

cube-on-cube relationship, that is, {100}precipitate//{100}matrix, and ⟨010⟩precipitate//⟨010⟩matrix.

Page 31: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

31

2.1.2 Strengthening and deformation mechanisms in nickel-based superalloys

To achieve optimal yield and creep strengths for turbine engine components, modern

commercial nickel-base superalloys contain a high volume fraction, typically 70 %, of coherent

γ'-precipitates having an average edge length of the order of 500 nm [3, 7, 84]. This leads to a

microstructure of closely spaced cuboidal γ'-precipitates aligned along the <001>-type directions,

and separated by narrow, ca. 60 nm, γ-channels. When the applied load is small, plastic

deformation is generally confined to the γ-matrix channels. At higher loads, precipitate-cutting

mechanisms, which entail deformation in both the matrix and precipitate phases, become

dominant. If the temperature is sufficiently high, dislocations can bypass the precipitates by

thermally-activated climb processes.

The high-temperature strength of nickel-based superalloys is a result of the deformation

mechanisms operating in the γ/γ’ system. The slip system of the FCC γ-matrix is a2

11 0 111{ },

and glide occurs by the passage of two Shockley partial dislocations. The dissociation reaction is

of the form:

a2

110 1 11{ }→a6

211 1 11{ }+a6

121 1 11{ } (1.1)

The magnitudes of the Burgers vectors of the Shockley partials are a / 6 , equivalent to the

displacement of neighboring planes in the FCC structure. As such, the passage of the first

Shockley partial dislocation results in an intrinsic stacking fault, which is removed by the

passage of the second Shockley partial. Slip deformation in the γ’-precipitate phase occurs along

the {111} planes, however the preferred slip system at low temperatures is a 11 0 111{ }, thus

Page 32: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

32

the passage of a a2

11 0 111{ } dislocation from the γ-matrix into the γ’-precipitates results in the

formation of an anti-phase boundary (APB). The energy of an APB is substantial, typically on

the order of 100 mJ m-2, although the energy is anisotropic, depending on the plane on which the

APB resides [3, 85, 86]. Therefore, a2

11 0 111{ } dislocations travel in pairs, known as

superdislocations, linked by an APB [87-91]. Dislocations in the γ’-L12 structure dissociate into

partial dislocations just as they do in the γ-matrix phase, though the Burgers vectors of the

dislocations are expected to be longer as a result of ordering in the crystal structure, and the

dissociation reactions are quite complex [3, 92]. Cutting of the γ/γ’ microstructure thus requires a

pair of dislocations, and the formation of an APB, leading to significant particle strengthening,

and hence an increase in the static tensile properties.

Upon deformation, the applied stress, the anisotropy of the APB energy, and

contributions of the elastic anisotropy combine to promote cross-slip of the superpartial

dislocations from the {111} to the {010} planes of the γ’-phase, where the APB energy is lowest.

These cross-slipped segments are then locked in place, as they cannot move without producing a

trail of APBs [93]. Since a component of the cross-slip process is thermally activated, the yield

strength of many nickel-based superalloys increases with increasing temperature, up to ca. 1073

K. For alloy CMSX-10, a single-crystal alloy used in turbine engines, for example, the yield

strength increases from a value of ca. 1000 MPa at a temperature of 473 K, to a peak value of ca.

1200 MPa at 1073 K [3].

Creep strengthening in nickel-based superalloys is a result of both solid-solution

strengthening and precipitation hardening. Because of its electronic structure, nickel has a large

Page 33: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

33

solid-solubility for many other elements, and thus there are many possibilities for solid-solution

strengthening [3, 94]. Studies of the effects of solid-solution strengtheners on the creep

resistance of nickel-based superalloys have determined a strong correlation between the

percentage difference in atomic size between the solute and solvent (Ni) atoms, and the creep-

strengthening increment [95]. Since a large-range of solid solubility is achieved only when the

atomic size of the metals are similar in size, a trade-off between solid solubility and lattice

distortion is mandated for effective solid-solution strengthening. High melting point elements,

such as Mo, Re, Ru, Nb and W, can provide strong lattice cohesion and reduce diffusion,

particularly at high temperatures, and are effective solid-solution strengtheners in nickel-base

superalloys [3, 7, 84]. Atomic clustering and short-range order can also strengthen the matrix,

but the concentration of refractory elements must be such that the deleterious TCP phases are not

formed. Substantial improvements in the creep resistance are achieved by the precipitation of the

γ’-phase, particularly when the grain size is large [96-98]. The prominent form of creep damage

in polycrystalline nickel-based alloys is creep cavitation at γ-grain boundaries, which is

minimized by increasing the grain size, hence the use of single-crystal superalloys for

applications that require creep resistance [3]. In current manufacturing processes, it is common

to cast engine components, such as nozzle guide vanes and turbine blades as single crystals, thus

eliminating the potential for grain boundary embrittlement and the need for grain boundary

strengtheners such as C, B, and Zr.

Page 34: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

34

2.2 Atom-probe tomography for characterizing nanostructures

The temporal evolution of the γ/γ’ microstructure in this thesis is studied from its earliest

stages with subnanometer spatial resolution employing pulsed-voltage and pulsed-laser atom-

probe tomography [99-104]. The atom-probe technique yields the position and chemical identity

of individual atoms in direct lattice space with a lateral spatial resolution of 0.3 to 0.5 nm, and a

depth resolution for composition analyses that may be less than 0.1 nm [25, 105]. The detection

sensitivity is identical for all elements in the periodic table; light elements (H, B, C, N, O) are

detected with the same sensitivity as Ta, Ru, W, Re. Rectangular parallelepipeds with a

maximum cross-sectional area of ca. 200 x 200 nm2 and a length approaching several microns

are obtained and generate 3-D atom-by-atom reconstructions that can contain 100-600 million

atoms. From the 3-D reconstruction, we obtain local compositions and concentration profiles of

each element, precipitate dimensions and morphology, number density, spatial dispersion,

volume fraction, size distributions, and edge-to-edge interprecipitate spacing distributions [106].

Additionally, the Gibbsian interfacial excesses of solutes at the γ/γ’ interfaces [50, 107-111], and

the partitioning of all elements between the γ- and γ’-phases, can be measured. These results, in

concert with simulations employing advanced computational methods such as first-principles

calculations, lattice kinetic Monte-Carlo simulation, and thermodynamic simulations, elucidate

the kinetic pathways that lead to phase decomposition at high-temperatures. The development of

future generations of nickel-based superalloys that are able to withstand higher engine operating

temperatures will rely on a detailed understanding of the γ/γ' phase transformation. These

complex multi-component alloys will serve as the building blocks for advanced turbine engines

that will require less fuel, and produce fewer CO2 greenhouse gas emissions.

Page 35: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

35

Chapter 3

Effects of solute concentrations on kinetic pathways in Ni-Al-Cr alloys

Abstract

The kinetic pathways resulting from the formation of coherent γ’-precipitates from the γ-

matrix are studied for two Ni-Al-Cr alloys with similar γ’-precipitate volume fractions at 873 K.

The details of the phase decompositions of Ni-7.5 Al-8.5 Cr at.%, and Ni-5.2 Al-14.2 Cr at.%,

for aging times from 1/6 to 1024 h are investigated by atom-probe tomography, and are found to

differ significantly from a mean-field description of coarsening. The morphologies of the γ’-

precipitates of the alloys are similar, though the degrees of γ’-precipitate coagulation and

coalescence differ. Quantification within the framework of classical nucleation theory reveals

that differences in the chemical driving forces for phase decomposition result in differences in

the nucleation behavior of the two alloys. The temporal evolution of the γ’-precipitate average

radii and the γ-matrix supersaturations follow the predictions of classical coarsening models. The

compositional trajectories of the γ-matrix phases of the alloys are found to follow approximately

the equilibrium tie-lines, while the trajectories of the γ’-precipitates do not, resulting in

significant differences in the partitioning ratios of the solute elements.

Page 36: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

36

3.1 Introduction

Τhe research of Schmuck et al. [28, 112] and Pareige et al. [37, 38] combined atom-probe

tomography and lattice kinetic Monte Carlo (LKMC) simulation to analyze the decomposition of

a Ni-Al-Cr solid-solution. A similar approach was applied by Sudbrack et al. [42, 45, 46, 55, 56]

and Yoon et al. [51, 58, 59] for studying Ni-5.2 Al-14.2 Cr at.% aged at 873 K and Ni–10 Al-8.5

Cr at.%, Ni–10 Al-8.5 Cr-2.0 W at.% and Ni–10 Al-8.5 Cr-2.0 Re at.% aged at 1073 K, which

decompose via a first-order phase transformation to form a high number density, Nv(t), of

nanometer-sized γ’-precipitates (~1020-1025 m-3). The addition of Cr to the binary Ni-Al system

reduces the lattice parameter misfit between the γ’-Ni3(AlxCr1-x)-precipitates and the γ-matrix,

often leading to γ’-precipitates that are nearly misfit free [12], thereby allowing the γ’-

precipitates to remain spheroidal to fairly large dimensions as aging progresses [17]. At 873 K,

γ’-precipitate radii as large as ~10 nm are measured by APT for concentrated Ni-Al-Cr alloys

[42]. The formation of coherent, spheroidal, γ’-precipitates with relatively stress-free

precipitate/matrix heterophase interfaces in the model Ni-Al-Cr alloys studied herein makes

them excellent candidates for comparison of experimental data with the predictions of classical

treatments of nucleation, growth and coarsening for ternary alloys, for which there is very little

detailed quantitative data available.

The present investigation focuses on comparing the temporal evolution of the

nanostructural and compositional properties at 873 K of alloy (A), Ni-7.5 Al-8.5 Cr at.%, with

those previously reported for alloy (B), Ni-5.2 Al-14.2 Cr at.% [42, 45, 46]; all concentrations

herein are in atomic percent (at.%) unless otherwise noted. A ternary Ni-Al-Cr phase diagram

determined by the Grand Canonical Monte Carlo (GCMC) technique at 873 K, Figure 3.1 [113],

Page 37: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

37

predicts that the values of φeq are 17.5 ± 0.5 and 15.1 ± 0.5 for alloys (A) and (B), respectively.

Since the predicted equilibrium volume fractions of the γ’-phase, φeq, for alloys (A) and (B) are

similar, it follows that any differences observed in the kinetic pathways during decomposition

are principally due to differences in solute concentrations.

Figure 3.1. A partial ternary phase diagram of the Ni-Al-Cr system at 873 K calculated using the Grand Canonical Monte Carlo simulation technique [113], showing the proximity of both alloys (A) Ni-7.5 Al- 8.5 Cr and (B) Ni-5.2 Al-14.2 Cr to the (γ + γ’) / γ solvus line. The tie-lines are drawn through the nominal compositions of the alloys and between the equilibrium phase compositions determined by extrapolation of APT concentration data to infinite time. Equilibrium solvus curves determined by Thermo-Calc [114], using databases for nickel-based superalloys due to Saunders [115] and Dupin et al. [116], are superimposed on the GCMC phase diagram for comparative purposes.

Page 38: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

38

3.1.1 Classical nucleation theory

The early stages of phase decomposition by nucleation have been studied theoretically in

a set of models known as classical nucleation theory (CNT), which have been reviewed

extensively in the literature [40, 117-120]. According to CNT, nucleation is governed by a

balance between a bulk free energy term, which has both chemical, ΔFch, and elastic strain

energy, ΔFel, components, and an interfacial free energy term, σγ/γ’, associated with the formation

of a γ-matrix/γ’-precipitate heterophase interface; F is the Helmholtz free energy. Thus, the

expression for the net reversible work required for the formation of a spherical nucleus, WR, as a

function of nucleus radius, R, is given by:

'/23 43

4)( γγσππ RRFFW elchR +Δ+Δ= . (3.1)

According to CNT, the net reversible work acts as a nucleation barrier that nuclei must

surmount to achieve a critical nucleus radius, R*. The critical net reversible work, *RW , required

for the formation of a critical spherical nucleus is expressed as:

2

3*

)(316 '/

elchR FF

WΔ+Δ

=γγσπ ; (3.2)

and R* is given by:

)(2*

'/

elch FFR

Δ+Δ−=

γγσ . (3.3)

For nucleation to occur, (ΔFch + ΔFel) must be negative. From CNT, the stationary-state

nucleation current, Jst, which is the number of nuclei formed per unit time per unit volume, is

given by:

Page 39: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

39

)exp(**

0 TkWNZJB

Rst −= β ; (3.4)

where Z, the Zeldovich factor, accounts for the dissolution of supercritical clusters, β∗ is a

kinetic coefficient describing the rate of condensation of single atoms on the critical nuclei, N0 is

the total number of possible nucleation sites per unit volume, taken to be the volume density of

lattice sites occupied by Al, the precipitate-forming solute [121, 122], kB is Boltzmann’s constant

and T is the absolute temperature in degrees Kelvin. The standard definitions of Z and β∗ were

employed and are given by [40, 117-120, 122]:

2/1

b2

i2

Tk21

i*)i(W

Z⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂−=

π, (3.5)

40

aDC*R4* π

β = ; (3.6)

where Wi is the net reversible work required for the formation of a spherical nucleus containing i

atoms, D is the diffusion coefficient of the precipitating solute element, C0 is the initial

concentration of said solute, and a is the average of the lattice parameters of the matrix and

precipitate phase.

In their review of nucleation kinetics results for binary alloys, Aaronson and Legoues

[123] note that while there exists some experimental evidence to support the correctness of CNT,

the nucleation currents predicted by the extant theories are often several orders of magnitude

smaller than those measured experimentally, which is most likely due to the presence of

precursor clusters that form between the solutionizing, and quenching and aging treatments. Such

precursor clustering has been detected in both alloys (A) [45] and (B), and thus, it is anticipated

that our calculated values of Jst will be smaller than those measured by APT.

Page 40: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

40

3.1.2 Coarsening theory

The first comprehensive mean-field treatment of Ostwald ripening [124], due to Lifshitz

and Slyozov [125] and Wagner [126], known as the LSW model, is limited to dilute binary

alloys with spatially-fixed spherical precipitates whose initial compositions are equal to their

equilibrium values. The LSW model for a binary alloy assumes: (i) no elastic interactions among

precipitates, thereby limiting the precipitate volume fraction to zero; (ii) precipitates have a

spherical morphology; (iii) coarsening occurs in a stress-free matrix; (iv) the precipitate diffusion

fields do not overlap; (v) dilute solid-solution theory obtains; (vi) the linearized version of the

Gibbs–Thomson equation is valid; (vii) coarsening occurs by the evaporation-condensation

mechanism; and (viii) precipitates coarsen with a fixed chemical composition, which is the

equilibrium composition. These requirements are highly restrictive and difficult to meet in

practice, and while experimental evidence exists to support the prediction of the time

dependency of the mean precipitate radius, <R(t)>, experimentalists have been unable to achieve

the exact stationary-state precipitate size distributions (PSDs) predicted by the LSW model [9,

10, 127]. Researchers have worked to remove the mean-field restrictions by developing models

based on multiparticle diffusion that are able to describe stress-free systems with finite volume

fractions [10, 128, 129].

Umantsev and Olson (UO) [130] were the first to demonstrate that the exponents of the

temporal power-laws predicted for binary alloys by LSW-type models are identical for

concentrated multi-component alloys, but that the explicit expressions for the rate constants

depend on the number of components. The UO model did not allow the composition of the

precipitates to evolve temporally, which is also true for the Morral and Purdy treatment [131].

Page 41: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

41

Kuehmann and Voorhees (KV) [132] considered isothermal quasi-stationary state coarsening in

ternary alloys and developed a model that permits the precipitate composition to evolve, such

that the matrix and precipitate compositions can deviate locally from their equilibrium

thermodynamic values. In the quasi-stationary limit in the KV model, δCi/δ t ≈ 0, the exponents

of the power-law temporal dependencies for <R(t)>, Nv(t), and the γ-matrix

supersaturation, )(tCiγΔ , of each solute species i, are:

);()()( 03

03 ttKtRtR KV −=><−>< (3.7)

);tt(K74.4

)t(N)t(N 0eqKV1

0v1

v −≅− −−

φ and (3.8)

;)()()( 3/1,

,, −=∞−>=<Δ tCtCtC KVieq

iff

iiγγγγ κ (3.9)

where KKV, and γκKVi ,

are the coarsening rate constants for <R(t)> and )(tCiγΔ , respectively;

<R(t0)> is the average precipitate radius and Nv(t0) is the precipitate number density at the onset

of quasi-stationary coarsening, at time t0. The quantity )(tCiγΔ is denoted a supersaturation and

is the difference between the concentration in the far-field γ-matrix, >< )(, tC ffiγ , and the

equilibrium γ-matrix solute-solubility, )(, ∞eqiC γ . The quantity )(, ∞eq

iC γ needs to be calculated

or determined experimentally as it is not available for Ni-Al-Cr multicomponent alloys at 873 K.

Atom-probe tomography of the nanostructures of alloys (A) and (B) provides an in-depth

examination of the compositional and nanostructural evolution of the γ’-precipitate phase as it

evolves toward its equilibrium thermodynamic composition. In this chapter, the decomposition

of the γ-matrix phase, from the earliest stages of solute-rich γ’-nuclei formation, to the

subsequent growth and coarsening of γ’-precipitates, is accessed within the framework of

Page 42: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

42

classical nucleation, growth, and coarsening theories. The effects of varying the solute

concentrations on the temporal evolution of the Ni-Al-Cr alloys are shown to provide a

quantitative understanding of the kinetic pathways that lead to phase separation, and the

achievement of the equilibrium compositions of both phases. We demonstrate that the kinetic

pathways to achieve thermodynamic equilibrium for the two alloys are very different, even

though the equilibrium volume fractions of the γ’-phase are approximately equal.

3.2 Experimental

High-purity constituent elements (99.97 Ni wt.%, 99.98 Al wt.%, and 99.99 Cr wt.%)

were induction-melted and chill cast in a 19 mm diameter copper mold under an Ar atmosphere.

The overall compositions of the two alloys were determined by inductively coupled plasma

(ICP) atomic-emission spectroscopy, which yielded average atomic compositions of 83.87 Ni-

7.56 Al- 8.56 Cr and 80.52 Ni-5.24 Al-14.24 Cr for alloys (A) and (B), respectively. Chemical

homogeneity of the cast ingots was achieved by annealing at 1573 K in the γ-phase field for 20 h.

Next, the ingots were held in the γ−phase field at 1123 K for 3 h to reduce the concentration of

quenched-in vacancies, and then water quenched and sectioned. Ingot sections were then aged at

873 K under flowing argon for times ranging from 1/6 to 1024 h, then water quenched, and

microtip specimens were prepared from each of the aged sections for study by APT. We

performed voltage-pulsed APT with a conventional APT [133, 134] and an Imago Scientific

Instrument’s local-electrode atom-probe (LEAPTM) tomograph [100-103]. APT data collection

was performed at a specimen temperature of 40.0 ± 0.3 K, a voltage pulse fraction (pulse

voltage/steady-state direct current voltage) of 19%, a pulse frequency of 1.5 kHz (conventional

Page 43: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

43

APT) or 200 kHz (LEAPTM tomograph), and a background gauge pressure of < 6.7 x 10-8 Pa.

The average detection rate in the area of analysis ranged from 0.011 to 0.015 ions per pulse for

conventional APT and from 0.04 to 0.08 ions per pulse for the LEAPTM tomographic analyses.

Conventional APT data were visualized and analyzed with APEX a12 software, the successor of

ADAM [135], while LEAPTM data were analyzed employing the IVAS 3.0 software program

(Imago Scientific Instruments, Madison, Wisconsin). The γ/γ' interfaces were delineated with

10.5 at.% and 9 at.% Al isoconcentration surfaces generated with efficient sampling procedures

[136], for alloys (A) and (B), respectively, and in-depth compositional information was obtained

with the proximity histogram method [137]. The volume fraction of the γ’-precipitate phase was

defined as the ratio of the total number of atoms contained within the isoconcentration surfaces

to the total number of atoms collected. Further experimental and analytical details for alloy (B)

can be found elsewhere [42, 46], the same procedures were employed for alloy (A). It is noted

that spatial convolution effects such as ion trajectory overlap and local magnification effects

have been cited as possible sources of misleading results in APT analysis of nickel-based

superalloys [138]. Ion trajectory overlap may cause both interfacial broadening and artificially

small values of the γ’-precipitate Al composition, particularly at early aging times. A comparison

of the composition profiles across the γ/γ’ interfaces measured by APT and simulated by lattice

kinetic Monte Carlo simulations for alloy (B) [39], showed no evidence of artificial interfacial

broadening in the APT data. Additionally, our results indicate that the Al composition is largest

at the shortest aging times due to a supersaturation of Al in the γ’-precipitates, thus trajectory

overlap effects are assumed to be minimal for the APT conditions used herein. Local

magnification effects due to differences in the required evaporation fields of different phases are

Page 44: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

44

not present in Ni-Al-Cr superalloys containing only γ-matrix and γ-precipitate phases because the

evaporation fields of the two phases are essentially identical.

The commercial software package Thermo-Calc [139] was used to calculate the values of

φeq, )(, ∞eqiC γ and the equilibrium γ’-precipitate composition of each solute species i, )(,' ∞eq

iC γ ,

for Ni-7.5 Al-8.5 Cr and Ni-5.2 Al-14.2 Cr at a pressure of 1 atmosphere, using databases for

nickel-based superalloys due to Saunders [115] and Dupin et al. [116]. For comparative

purposes, the equilibrium phase boundaries determined by Thermo-Calc using the Saunders and

Dupin et al. databases are superimposed on the GCMC phase diagram at 873 K, Figure 3.1.

While the generated γ / (γ+γ’) solvus lines show good agreement, the curvatures of the (γ+γ’) / γ’

phase lines differ for each technique. Although the Ni-Al-Cr system was studied extensively by

Taylor and Floyd [12], there is no extant experimental phase diagram at 873 K to compare to.

3.3 Results

3.3.1 Morphological development

Nanometer-sized spheroidal γ’-precipitates are detected in both alloys over the full range

of aging times, from 1/6 to 1024 h. The temporal evolution of the morphology of alloy (A) is

shown in a series of APT micrographs in Figure 3.2, which can be compared with a similar series

for alloy (B) in reference [42]. Figure 3.3 is a projection of a 15 x 15 x 25 nm3 subset of an APT

analysis of alloy (A) aged to 1024 h, showing a spheroidal γ’-precipitate of radius ~ 9 nm,

delineated by a red 10.5% Al isoconcentration surface. Atomic planes are clearly visible within

the γ’-precipitate, and the value of the interplanar spacing is 0.26 ± 0.03 nm, suggesting {110}-

type planes. This APT image demonstrates that the γ’-precipitates remain spheroidal for aging

Page 45: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

45

times as long as 1024 h at an aging temperature of 873 K, as confirmed by the TEM micrograph

in Figure 3.4.

Figure 3.2. The temporal evolution of the γ’-precipitate nanostructure in Ni-7.5 Al-8.5 Cr aged at 873 K is displayed in a series of APT parallelepipeds. The parallelepipeds are 10 x 10 x 25 nm3

subsets of the analyzed volume and contain ~ 125,000 atoms. The γ/γ’ interfaces are delineated in red with 10.5% Al isoconcentration surfaces.

Page 46: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

46

Figure 3.3. A subset of an APT micrograph of Ni-7.5 Al-8.5 Cr aged at 873 K for 1024 h, containing 350,000 atoms, with Al atoms in red, and Cr atoms in blue; Ni atoms are omitted for clarity. A γ’-precipitate with a radius of ~ 9 nm is delineated by the red 10.5% Al isoconcentration surface, and shows {110}-type superlattice planes with an interplanar spacing of 0.26 ± 0.03 nm.

Page 47: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

47

Figure 3.4. A centered superlattice reflection dark–field image of spheroidal Ni3(AlxCr(1-x)) γ’-precipitates for a Ni-7.5 Al-8.5 Cr sample aged for 1024 h at 873 K, recorded near the [011] zone axis.

There is evidence of γ’-precipitate coagulation and coalescence in alloys (A) and (B),

characterized by the formation of necks that interconnect the γ’-precipitates, and exhibit L12-type

ordering. Although the first γ’-nuclei are detected by APT at an aging time of 1/6 h,

interconnected γ’-precipitates first evolve in both alloys after 1/4 h of aging, which coincides

with the end of the quasi-stationary-state nucleation regime. After 1/4 h of aging, the fraction of

coagulating and coalescing γ’-precipitates, f, is 15 ± 4% and 9 ± 3%, for alloys (A) and (B),

respectively. From Figure 3.5, the maximum value of f of 18 ± 4% for alloy (A) occurs at an

aging time of 1 h, while the maximum value for alloy (B) of 30 ± 4% occurs at an aging time of

4 h.

Page 48: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

48

Figure 3.5. The temporal evolution of the value of the fraction of γ’-precipitates interconnected by necks, f, and the average interprecipitate edge-to-edge spacing, <λe-e>. The maximum value of f of 18 ± 4 % corresponds to the minimum value of <λe-e> of 7 ± 2 nm at an aging time of 1 h for alloy (A) Ni-7.5 Al-8.5 Cr. For alloy (B) Ni-5.2 Al-14.2 Cr, the minimum value of <λe-e> of 5.9 ± 0.8 nm and the maximum value of f of 30 ± 4 % coincide at an aging time of 4 h.

3.3.2 Temporal evolution of the nanostructural properties of γ’-precipitates

Figure 3.6 provides a quantitative description of the temporal evolution of the γ’-

precipitate volume fraction, φ, and the quantities <R(t)> and Nv(t), for alloys (A) and (B). The γ’-

Page 49: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

49

precipitate nanostructural properties determined by APT analysis for alloy (A) are summarized in

Table 3.1, while reference [42] contains details for alloy (B). The standard errors for all

quantities, σ, are calculated based on counting statistics and reconstruction scaling errors using

standard error propagation methods [140], and represent one standard deviation from the mean.

The two Ni-Al-Cr alloys were designed to have similar values of φeq of approximately 16%, and

the measured values of φ for alloy (A) and (B) are statistically indistinguishable over the full

range of aging times, from 1/6 to 1024 h. It is worth noting that over the range of aging times

studied, the values of <R(t)> of alloy (A) are, on average, 32 ± 6% larger than those of alloy (B).

From Figure 3.6, the temporal evolution of alloys (A) and (B) can be divided into three regimes:

(i) quasi-stationary-state γ’-precipitate nucleation at early aging times; followed by (ii)

concomitant precipitate nucleation and growth, and finally (iii) concurrent growth and

coarsening once the maximum value of Nv(t) is achieved. The characteristic microstructural

evolution associated with these three regimes is described in detail below.

Page 50: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

50

Figure 3.6. The temporal evolution of the γ’-precipitate volume fraction, φ, number density, Nv(t), and mean radius, <R(t)>, for (A) Ni-7.5 Al-8.5 Cr and (B) Ni-5.2 Al-14.2 Cr alloys aged at 873 K as determined by APT microscopy. The quantity <R(t)> is proportional to t1/3 as predicted by the Umantsev and Olson (UO) and Kuehmann and Voorhees (KV) models for isothermal coarsening in ternary alloys. The temporal dependence of the diminution of the quantity Nv(t) deviates from the t-1 prediction of the UO and KV models for both alloys.

Page 51: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

51

Table 3.1. Temporal evolution of the nanostructural properties of γ’-precipitates a determined by APT for Ni-7.5 Al-8.5 Cr aged at 873 K.

Aging time (h)

Npptb <R(t)> ± σ

(nm) Nv(t) ± σ

(x 1024 m-3) φ ± σ (%)

f ± σ (%)

<λe-e> ± σ (nm)

1/6 8 0.90 ± 0.32 0.26 ± 0.14 0.31 ± 0.11 ND c 17.7 ± 6.3

1/4 101 1.00 ± 0.11 1.89 ± 0.59 1.36 ± 0.14 15 ± 4 7.7 ± 2.7

1 70.5 1.24 ± 0.12 2.21 ± 0.64 2.48 ± 0.25 18 ± 4 7.0 ± 2.5

4 46 1.70 ± 0.25 1.02 ± 0.31 5.98 ± 0.88 16 ± 3 8.9 ± 3.2

16 42 2.80 ± 0.43 0.60 ± 0.18 9.12 ± 1.4 13 ± 3 9.1 ± 3.2

64 76.5 3.59 ± 0.41 0.34 ± 0.10 11.8 ± 1.4 7.4 ± 1.9 10.6 ± 3.7

256 15 5.54 ± 1.43 0.20 ± 0.05 14.6 ± 3.8 2.5 ± 0.9 10.1 ± 3.6

1024 8 8.30 ± 2.93 0.13 ± 0.05 16.0 ± 5.7 ND c 7.9 ± 2.8

a Mean radius of γ’-precipitates, <R(t)>; the number density, Nv(t); precipitated volume fraction, φ; fraction of γ’-precipitates interconnected by necks, f; average edge-to-edge interprecipitate spacing, <λe-e>; and their standard errors, σ, one standard deviation is reported.

b The number of γ’-precipitates analyzed, Nppt, is smaller than the total number of γ’-precipitates detected by APT. Precipitates that intersect the sample volume contribute 0.5 to the quantity Nppt, and are included in the estimates of Nv(t), φ, f, and the phase compositions, and not in the measurement of <R(t)>.

c ND = not detected.

3.3.2.1 Nucleation and growth of γ’-precipitates

Phase decomposition in alloys (A) and (B) begins with the formation of solute-rich nuclei

that grow to form stable γ’-precipitates. Precipitates are first detected by APT in alloys (A) and

(B) after 1/6 h of aging. Alloy (A) forms (2.6 ± 1.4) x 1023 γ’-precipitates m-3 with an <R(t = 1/6

h)> value of 0.90 ± 0.32 nm, corresponding to a φ value of 0.31±0.11 %, after 1/6 h of aging. For

the same aging time, Alloy (B) forms (3.6 ± 1.3) x 1023 γ’-precipitates m-3 with an <R(t = 1/6

Page 52: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

52

h)> value of 0.74 ± 0.24 nm, accounting for a φ value of 0.11±0.04 %. The nucleation of stable

γ’-precipitates in alloy (A) for aging times between 1/6 and 1/4 h results in a sharp linear slope of

the Nv(t) profile of (5.4 ± 1.5) x 1021 m-3 s-1, while the values of <R(t)> of 0.90 ± 0.32 nm and

1.00 ± 0.11 nm, remain constant within statistical error. This slope for Nv(t), although based on

two experimental data points, is taken to be an estimate of the quasi-stationary-state nucleation

current of γ’-precipitates, Jst. In contrast, alloy (B) undergoes nucleation of stable γ’-precipitates

with a constant <R(t)> value of 0.75 ± 0.24 nm for aging times less than 1/4 h. The value of Jst

for alloy (B) is estimated to be (5.9 ± 1.7) x 1021 m-3 s-1, which is statistically indistinguishable

from the value of Jst of (5.4 ± 1.5) x 1021 m-3 s-1

measured for alloy (A).

At an aging time of 1/4 h, alloys (A) and (B) enter a regime of concomitant γ’-precipitate

nucleation and growth, which results in steadily increasing φ and <R(t)> values, and a maximum

value for Nv(t). The peak value in Nv(t) of (2.21 ± 0.64) x 1024 m-3 for alloy (A) occurs at 1 h,

while for alloy (B), the peak value in Nv(t) of (3.2 ± 0.6) x 1024 m-3 is achieved at 4 h. We note

that the prediction of the temporal dependence of <R(t)> of t1/2 for interface reaction controlled

growth [141-143], is observed in neither of the Ni-Al-Cr alloys studied, contrary to the results

for an earlier APT study of an alloy whose composition is essentially identical to that of alloy

(B) [28].

3.3.2.2 Growth and coarsening of γ’-precipitates

Beyond the peak in Nv(t), alloys (A) and (B) enter a quasi-stationary-state of growth and

coarsening characterized by a steady diminution of Nv(t) and increasing values of φ and <R(t)>.

Beyond 1 h, the quantity Nv(t) for alloy (A) displays a temporal dependence of t-0.42 ± 0.03, which

Page 53: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

53

differs significantly from the predicted value of t-1 from the UO and KV models. This is not

surprising, since the quantity φ continues to evolve temporally in this regime, implying that the

system has not achieved a stationary-state [144]. During coarsening in alloy (B), the quantity

Nv(t) displays a temporal dependence of t-0.67 ± 0.01 beyond an aging time of 4 h, which also differs

from the predicted value of t-1, though to a lesser extent. The values of φ for the two alloys

increase steadily in this regime, reaching values of φ of 16.0 ± 5.7 % and 15.6 ± 6.4 %, for alloys

(A) and (B), respectively, which are statistically indistinguishable. In the growth and coarsening

regime, the quantity <R(t)> displays a temporal dependence of t0.29 ± 0.03 for alloy (A) and t0.29 ±

0.05 for alloy (B), which both agree approximately, but not exactly, with the predicted value of

t1/3, indicating that the phase transformation is primarily diffusion-limited.

From Figure 3.5, the maximum value for f, after 1 h of aging, of 18 ± 4% for alloy (A)

coincides with the minimum value of the average edge-to-edge interprecipitate spacing, <λe-e>,

of 7.0 ± 2.5 nm and the peak value of Nv(t) of (2.21 ± 0.64) x 1024 m-3. For alloy (B), the peak

value of Nv(t), at an aging time of 4 h, is (3.2 ± 0.6) x 1024 m-3, corresponding to a minimum

value of <λe-e> of 5.9 ± 0.8 nm at a maximum value of f of 30 ± 4%. The quantity <λe-e> is

calculated using Equation 3.10, which assumes (i) a regular simple cubic array of γ’-precipitates;

and (ii) the unit volume is equivalent to that of a lattice enclosed sphere in order to preserve the

radial symmetry of the 3D structure [42]:

⎥⎥⎦

⎢⎢⎣

⎡><−⎟

⎠⎞

⎜⎝⎛ ⋅>=<

− )()(342

3/1

tRtNvee πλ . (3.10)

Page 54: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

54

3.3.3 Temporal evolution of the compositions of the γ and γ’-phases

The compositions of the γ-matrix and the γ’-precipitate phases of alloys (A) and (B)

evolve temporally as the γ-matrix becomes enriched in Ni and Cr and depleted in Al, as shown in

Figure 3.7 for alloy (A).

Figure 3.7. The composition profiles on either side of the heterophase γ-matrix / γ’-precipitate interface for (A) Ni-7.5 Al-8.5 Cr aged at 873 K for aging times of 1/4, 1, 4, 1024 h. The phase compositions evolve temporally, as the γ-matrix becomes enriched in Ni and Cr and depleted in Al. The values of <R(t)> for these aging times are 1.00 ± 0.11 nm for 1/4 h, 1.24 ± 0.12 nm for 1 h, 1.70 ± 0.25 nm for 4 h and 8.30 ± 2.93 nm for 1024 h.

Page 55: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

55

The compositional trajectories of the γ-matrix and γ’-precipitate phases are shown on a

partial Ni-Al-Cr ternary phase diagram at 873 K in Figure 3.8. The mean-field KV model of

quasi-stationary state coarsening predicts that the slope of the trajectory of the γ-matrix phase

during coarsening lies along the equilibrium tie-line and has a value of pAl/pcr, where pi is the

magnitude of the partitioning of solute species i, defined as pi = ,[ ( )eqiCγ ′ ∞ - , ( )]eq

iCγ ∞ . For a binary

alloy, the compositions at the interface are given by the assumption of local equilibrium,

whereas for a ternary alloy, the interfacial compositions must be further defined by the condition

of flux balance at the interface [128]. As such, according to the KV model, the addition of a third

alloying element alters both the form of the Gibbs-Thompson equations and the predictions of

the temporal evolution of the phase compositions. The compositional trajectory of the γ’-

precipitate phase is predicted to lie on a straight line that is not necessarily parallel to the

equilibrium tie-line, which is the case for both alloys (A) and (B). In order to quantify deviations

from the equilibrium tie-line, the quantities )(/)( tCtC CrAlγγ δδ and )(/)( '' tCtC CrAl

γγ δδ for aging times

of 4-1024 h are compared to the values of pAl/pcr for both alloys, where δCpi is the slope of the

concentration of element i in phase p with aging time. The trajectories of the γ-matrix phase and

the γ’-precipitate phase compositions of alloy (A) have slopes of – 2.33 ± 0.06 and 5.71 ± 1.42,

respectively, while the slope of the equilibrium tie-line is estimated to be -3.52 ± 0.09. Slopes of

-1.36 ± 0.24 and 0.85 ± 0.20 are estimated for the γ-matrix and γ’-precipitate phases,

respectively, from the APT data for alloy (B), while the slope of the equilibrium tie-line is -1.53

± 0.07. From these results, it is absolutely clear that the trajectories of the composition of the γ’-

precipitate phases in alloys (A) and (B) do not lie along the equilibrium tie line of the respective

alloys, which is contrary to all coarsening models except the KV model. Hence, our results

Page 56: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

56

indicate that the temporal evolution of ternary alloys is considerably more complicated than for

binary alloys, and certainly warrants future research.

Figure 3.8. The compositional trajectories of the γ-matrix and γ’-precipitate phases of alloys (A) Ni-7.5 Al-8.5 Cr and (B) Ni-5.2 Al-14.2 Cr as they evolve temporally, displayed on a partial Ni-Al-Cr ternary phase diagram at 873 K. The tie-lines are drawn through the nominal compositions of the alloys (squares) and between the experimentally determined equilibrium phase compositions. The trajectories of the γ-matrix phases in alloys (A) and (B) lie approximately on the experimental tie-lines. The trajectories of the γ'-precipitate phases do not lie along the tie-line, as predicted by the Kuehmann and Voorhees model of isothermal coarsening in ternary alloys [132].

During nucleation, solute-rich γ’-nuclei form with large values for the Al and Cr

supersaturations. The first γ’-nuclei detected by APT for alloy (A) have solute-supersaturated

compositions of 70.9 ± 1.4 Ni, 23.3 ± 1.5 Al and 5.8 ± 0.9 Cr at an <R(t = 1/6 h)> value of 0.90

± 0.32 nm for alloy (A). The γ’-nuclei in alloy (B) have a composition of 71.3 ± 1.6 Ni, 19.1 ±

1.4 Al and 9.7 ± 1.1 Cr, and an <R(t = 1/6 h)> value of 0.74 ± 0.24 nm. As phase decomposition

Page 57: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

57

progresses beyond nucleation, the magnitude of the values of )(tCiγΔ decrease asymptotically

toward a value of zero as the equilibrium γ- and γ’-phase compositions are approached. The

equilibrium γ-matrix and γ’-precipitate compositions are extrapolated by fitting the measured

concentrations from the quasi-stationary coarsening regime to Equation 3.9 for aging times

beyond 1 h for alloy (A) and 4 h for alloy (B). The equilibrium γ’-precipitate composition of

alloy (A) is estimated to be 76.33 ± 0.12 Ni, 17.82 ± 0.15 Al and 5.85 ± 0.12 Cr, while the

equilibrium γ-matrix has a composition of 85.19 ± 0.08 Ni, 5.42 ± 0.09 Al and 9.39 ± 0.09 Cr at

infinite time. The equilibrium γ’-precipitate phase composition of alloy (B) is estimated to be

76.53 ± 0.25 Ni, 16.69 ± 0.22 Al and 6.77 ± 0.15 Cr, while the equilibrium composition of the γ-

matrix phase is 81.26 ± 0.09 Ni, 3.13 ± 0.04 Al and 15.61 ± 0.09 Cr. Summaries of the

equilibrium phase compositions of alloy (A) determined by APT are presented in Table 3.2, and

agree with results obtained from Thermo-Calc and GCMC simulation. Reference [42] contains a

similar table for alloy (B).

Page 58: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

58

Table 3.2. Equilibrium γ’-precipitate and γ-matrix concentrations, as determined by atom-probe tomography (APT), Grand Canonical Monte Carlo (GCMC) simulation, and thermodynamic modeling employing Thermo-Calc for alloy (A) Ni-7.5 Al-8.5 Cr aged at 873 K.

Equilibrium composition of γ'-precipitates Ni (at.%) Al (at.%) Cr (at.%)

Measured by APT at 1024 h: 76.11 ± 0.09 18.02 ± 0.09 5.87 ± 0.05

Extrapolated from APT data: 76.33 ± 0.12 17.82 ± 0.15 5.85 ± 0.12

Modeled by GCMC simulation [113]: 76.3 ± 0.5 17.8 ± 0.5 5.9 ± 0.5

Calculated with Thermo-Calc and Saunders database [115]:

76.40 17.79 5.81

Calculated with Thermo-Calc and Dupin et al. database [116]:

75.44 17.87 6.99

Equilibrium composition of γ-matrix Ni (at.%) Al (at.%) Cr (at.%)

Measured by APT at 1024 h: 85.13 ± 0.06 5.53 ± 0.07 9.34 ± 0.04

Extrapolated from APT data: 85.19 ± 0.08 5.42 ± 0.09 9.39 ± 0.09

Modeled by GCMC simulation [113]: 85.8 ± 0.5 5.2 ± 0.5 9.0 ±0 .5

Calculated with Thermo-Calc and Saunders database [115]:

85.71 5.42 8.86

Calculated with Thermo-Calc and Dupin et al. database [116]:

85.3 5.99 8.70

The partitioning behavior of the elements can be determined quantitatively by calculating

the partitioning ratio, Kiγ’/γ

, defined as ratio of the concentration of an element i in the γ’-

precipitates to the concentration of the same element in the γ-matrix. Figure 3.9 demonstrates

that alloys (A) and (B) both exhibit partitioning of Al to the γ’-precipitates and of Ni and Cr to

the γ-matrix. Partitioning is more pronounced in alloy (B) than (A), as the smaller Al and the

larger Cr concentrations of this alloy result in a smaller solubility of Al, and a larger solubility of

Cr in the γ-matrix, respectively. Nickel partitions to the γ-matrix with a Kiγ’/γ of ~ 0.9 for both

alloys, and is only slightly time dependent.

Page 59: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

59

Figure 3.9. The partitioning ratios, Kiγ’/γ

, of Al and Cr demonstrate that both alloys exhibit partitioning of Al to the γ’-precipitates and Cr to the γ-matrix. Partitioning is more pronounced in alloy (B), where the smaller Al and larger Cr concentration results in a smaller value of the γ-matrix Al solubility, and a larger γ-matrix Cr solubility.

The lever rule is applied to the equilibrium phase compositions for both alloys to estimate

a φeq value of 16.4 ± 0.6 % for alloy (A), and 15.7 ± 0.7 % for alloy (B). The φ values of 16.0 ±

5.7 % and 15.6 ± 6.4 % determined by APT at an aging time of 1024 h for alloy (A) and (B), are

within experimental error of the φeq values estimated by the lever rule. The Thermo-Calc

software package yields φeq values of 16.7 % and 14.9 % for alloy (A) according to the Saunders

Page 60: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

60

and Dupin et al. databases, respectively, thus the Saunders database yields the value closest to

the experimental value of 16.4 ± 0.6%. Alloy (B) is predicted to achieve φeq values of 12.83%

and 12.34% for the same databases, which are both smaller than the experimentally determined

value of 15.7 ± 0.7%. Summaries of the φeq values determined by APT are presented in Table

3.3, and agree with results obtained from Thermo-Calc and GCMC modeling to different

degrees.

Table 3.3. Equilibrium γ’-precipitate volume fraction, φeq, as determined by atom-probe tomography (APT), Grand Canonical Monte Carlo (GCMC) simulation, and Thermo-Calc for alloy (A) Ni-7.5 Al-8.5 Cr, and alloy (B) Ni-5.2 Al-14.2 Cr, aged at 873 K.

Technique used to estimate φeq (A) Ni-7.5 Al-8.5 Cr (B) Ni-5.2 Al-14.2 Cr

Determined by lever rule calculation with phase compositions measured by APT:

16.4 ± 0.6 15.7 ± 0.7

Measured by APT at 1024 h: 16.0 ± 5.7 15.6 ± 6.4

Modeled by GCMC simulation [113]: 17.5 ± 0.5 15.1 ± 0.5

Calculated with Thermo-Calc and Saunders database [115]:

16.69 12.83

Calculated with Thermo-Calc and Dupin et al. database [116]:

14.90 12.34

The values of the solid-solution supersaturations of alloys (A) and (B) are calculated

based on the equilibrium phase compositions from APT data. The initial solid-solution

supersaturation values are estimated to be 2.08 ± 0.02 Al and -0.89 ± 0.01 Cr, for alloy (A), and

2.09 ± 0.04 Al and -1.46 ± 0.02 Cr, for alloy (B). Figure 3.10 shows the temporal evolution of

the Al and Cr γ-matrix supersaturation values for both alloys. The formation of γ’-nuclei during

the early stages of aging results in a decrease in the magnitude of the values of )(tCiγΔ , which in

Page 61: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

61

turn causes the slowing and eventual termination of γ’-precipitate nucleation. Beyond the aging

time corresponding to the peak γ’-precipitate number density, 1 h for alloy (A), and 4 h for alloy

(B), the diminution of the )(tCiγΔ values approximately follow the t-1/3 prediction of the KV

model. From Figure 3.10, alloy (A) demonstrates a temporal dependence of t-0.32 ± 0.03 for

)(tCAlγΔ and t-0.29 ± 0.04 for )(tCCr

γΔ , and alloy (B) exhibits a dependence of t-0.33±0.04 for )(tCAlγΔ

and t-0.34±0.07 for )(tCCrγΔ . The supersaturation values of the γ’-precipitate phases, )(tCi

γ ′Δ , of

alloys (A) and (B) are a reflection of their alloy composition, as the magnitude of )(tCiγ ′Δ is

greater in alloy (A), Ni-7.5 Al-8.5 Cr, which contains more Al, than in alloy (B), Ni-5.2 Al-14.2

Cr, while the inverse is true for Cr. From Figure 3.11, the quantities )(tCAlγ ′Δ and )(tCCr

γ ′Δ for

alloy (A) exhibit temporal dependencies of t-0.34 ± 0.04 and t-0.32 ± 0.05 respectively, which are close

to the predicted value of t-1/3. A dependence of t-0.30 ± 0.05 for )(tCAlγ ′Δ , and t-0.29 ± 0.07 for )(tCCr

γ ′Δ ,

are demonstrated for alloy (B).

Page 62: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

62

Figure 3.10. The magnitude of the values of the supersaturations, )(tCiγΔ , of Al and Cr in the γ-

matrix are smaller for alloy (A) Ni-7.5 Al-8.5 Cr than for alloy (B) Ni-5.2 Al-14.2 Cr. The magnitude of the )(tCi

γΔ values decrease as t-1/3 in the coarsening regimes for both alloys, as predicted by the Umantsev and Olson (UO) and Kuehmann and Voorhees (KV) models for isothermal quasi-stationary state coarsening in ternary alloys. The values of )(tCCr

γΔ are expressed as an absolute value because they are negative, and reflect a flux of Cr into the γ-matrix with increasing aging time.

Page 63: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

63

Figure 3.11. The supersaturation values of the γ’-precipitates, )(tCiγ ′Δ , reflect the chemical

compositions of the two alloys, as the value of )(tCAlγ ′Δ is larger in alloy (A) Ni-7.5 Al- 8.5 Cr,

which contains more Al, than in alloy (B) Ni-5.2 Al-14.2 Cr, while the inverse is true for Cr. The values of )(tCi

γ ′Δ decrease as approximately t-1/3 in the coarsening regimes for both alloys.

Page 64: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

64

3.4. Discussion

While the temporal evolution of the morphology and the volume fraction of the γ’-

precipitates in alloy (A) Ni-7.5 Al-8.5 Cr and alloy (B) Ni-5.2 Al-14.2 Cr are similar, the

nanostructural and compositional results for the two alloys exhibit significant differences. These

differences can be attributed to the effects of solute concentrations on the kinetic pathways

involved in phase decomposition.

3.4.1 Effects of solute concentration on nucleation behavior

The nucleation behavior observed by APT for alloys (A) and (B) can be compared with

the predictions of CNT using Equations 3.2-3.4, given the values of σγ/γ’, ΔFch, and ΔFel for both

alloys. The values of σγ/γ’ are estimated as first shown by Ardell for a binary alloy [145, 146],

and later for a ternary alloy by Marquis and Seidman [147], and as applied to alloy (B) by

Sudbrack et al. [46]. The relationship for σγ/γ’ in a nonideal, nondilute ternary alloy consisting of

a γ-matrix and a γ’-precipitate phase with a finite volume fraction of the γ’-phase is given by

[128]:

)GpGpp2Gp(pV2

)K(Cr,Cr

2CrAl,CrCrAlAl,Al

2Al

im

KVi3/1

KV/ γγγγ

γγγ κ

σ ++= ′′ ; (3.11)

where KKV and γκ KVi, are the rate constants for the quantities <R(t)> and )(tCiγΔ , respectively,

from the KV coarsening model, pi is the magnitude of the partitioning as previously defined,

γ ′mV is the molar volume of the γ’-precipitate phase, calculated to be 6.7584 x 10-6 m3 mol-1,

and γjiG , is shorthand notation for the partial derivatives of the molar Gibbs free-energy of the γ-

Page 65: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

65

matrix phase with respect to the solute species i and j. The quantities KKV and γκ KVi, are

determined by fitting the experimental APT data to Equations 3.7 and 3.9 from the KV

coarsening model. For alloy (A), KKV is (1.84 ± 0.43) x 10-31 m3s-1 and γκ KVAl , and γκ KVCr , are 0.18

± 0.05 at. fr. s1/3 and -0.06 ± 0.01 at. fr. s1/3, respectively, while a value of KKV of (8.8 ± 3.3) x 10-

32 m3s-1, and values of γκ KVAl , and γκ KVCr , of 0.19 ± 0.02 at. fr. s1/3 and -0.14 ± 0.05 at. fr. s1/3, are

found for alloy (B). For the general case described by nonideal and nondilute solution theory, the

values of γjiG , may be calculated using Thermo-Calc employing the extant databases for nickel-

based superalloys. As noted by Sudbrack et al. [46], the calculations of γjiG , from Thermo-Calc

predict a more highly-curved free energy surface than ideal solution theory with respect to all

solute species combinations. The Thermo-Calc assessments take into account the excess free-

energies of mixing and the magnitudes of γjiG , are 1.5 to 13 times larger than those for ideal

solution theory for alloys (A) and (B) as shown in Table 3.4 for alloy (A). See Table 3.4 in

Sudbrack et al. [46] for the values of γjiG , for alloy (B). Note that two values of σγ/γ’, '/ γγσ Al and

'/Cr

γγσ , are obtained from the coarsening data, depending on the choice of the solute element used

to estimate γκ KVi, and pi.

Page 66: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

66

Table 3.4. Curvatures in the molar Gibbs free-energy surface of the γ-matrix phase evaluated at the equilibrium composition with respect to components i and j, γ

jiG , , obtained from ideal solution theory and Thermo-Calc thermodynamic assessments for alloy (A) Ni-7.5 Al-8.5 Cr aged at 873 K.

γjiG , Ideal Solution Theory

( J mol-1) Saunders database [115]

(J mol-1)Dupin et al. database

[116] (J mol-1)γ

AlAlG 142,441.7 271,392.5 306,697.8 γ

CrCrG ,85,821.1 166,601.4 166,708.5

γCrAlG ,

8,521.4 112,567.3 140,244.9

Fortuitously, the Thermo-Calc and ideal solution theory assessments of γjiG , yield

approximately the same value for σγ/γ’ for alloy (B) of 22-23 ± 7 mJ m-2 [46]. For alloy (A), the

values for σγ/γ’, calculated from the ideal solution theory of 14-16 ± 3 mJ m-2 are smaller than

those from the Thermo-Calc assessment of γjiG , of 23-25 ± 6 mJ m-2, Table 3.5. For the purposes

of CNT, values of σγ/γ’ of 24 ± 6 mJ m-2, for alloy (A), and 22.5 ± 7 mJ m-2, for alloy (B) are

used, the averages of the values generated from the Thermo-Calc assessments of γjiG , .

Additionally, estimates of 36.3 ± 3.8 mJ m-2 and 35.7 ± 1.7 mJ m-2 are obtained for alloys (A)

and (B) at 0 K, from first principles calculations, in the framework of density functional theory

(DFT) and the local density approximation (LDA), employing ultrasoft Vanderbilt potentials

(US-PP), using the VASP (Vienna Ab-initio simulation package) code [148-151]. It is noted that

the values of σγ/γ’ from the first principles calculations are larger than those determined

experimentally because the first principles calculations assume a sharp γ/γ’ interface and are

performed at 0 K, and thus do not include entropic effects. The experimental estimates of σγ/γ’ are

for diffuse γ/γ’ interfaces at 873 K, and are free energies because they include entropic effects.

Page 67: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

67

Table 3.5. Free-energy of the γ/γ’ interfaces, σγ/γ’, at 873 K in Ni–7.5 Al–8.5 Cr calculated from the experimental values of the Kuehmann-Voorhees coarsening rate constants for the average precipitate radius and the supersaturation of solute species i employing Equation 3.11 with solution thermodynamics described by the ideal solution and Thermo-Calc databases given in Table 3.4.

Thermodynamic models '/ γγσ Al

(mJ m-2)

'/ γγσ Cr (mJ m-2)

σγ/γ’ (mJ m-2)

Ideal solution theory 14.2 ± 3.1 16.0 ± 3.6 15.1 ± 3.4

Saunders database [115] 21.6 ± 4.0 24.4 ± 6.1 23.0 ± 5.1

Dupin et al. database [116] 23.6 ± 4.9 26.6 ± 6.2 25.1 ± 5.5

A value of σγ/γ’ of 12.5 mJ m-2 was previously determined for a ternary Ni-5.2 Al-14.8 Cr

alloy aged at 873 K by Schmuck et al. [28]. The value of σγ/γ’ for this alloy, which has an overall

composition that is very close to that of alloy (B) Ni-5.2 Al-14.2 Cr, was determined assuming

the LSW model for binary alloys. We reanalyzed their data according to the method developed

by Marquis and Seidman for ternary alloys, and values of 20-21 ± 5 mJ m-2 are estimated

for σγ/γ’, in good agreement with the values of σγ/γ’ estimated for alloys (A) and (B); additional

details are presented in Appendix 1. Gleiter and Hornbogen [152] estimated a value of 13.5 mJ

m-2 for a Ni-5.4 Al-18.7 Cr alloy aged at 750 ºC, though they also applied binary LSW theory to

a ternary alloy, and their results did not include concentration data, making recalculation of their

value of σγ/γ impossible by the method developed by Marquis and Seidman for ternary alloys.

Baldan [9] provides a review of literature values of σγ/γ’ for several different Ni-Al and Ni-Al-Cr

systems over a range of aging temperatures, which need to be evaluated in detail in light of our

work.

The values of the chemical driving force of alloys (A) and (B) are estimated as proposed

Page 68: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

68

by Lupis [153] and applied by Schmuck et al. [28], from thermodynamic data taken from

Thermo-Calc [139] and the extant nickel-based superalloy databases [115, 116]. The values of

ΔFch for alloys (A) and (B) are estimated to be -6.6 x 107 J m-3 and -8.2 x 107 J m-3, respectively.

The values of ΔFel for the two alloys are estimated using [154]:

'

2'

)43()(2

γγγ

γγγγ

a

aael VSB

VVBSF+

−=Δ ′

; (3.12)

where Sγ is the shear modulus of the γ-matrix phase, Bγ’ is the bulk modulus of the γ’-precipitate

phase, and γaV and 'γ

aV are the atomic volumes of the γ-matrix and γ’-precipitate phases,

respectively. No elastic constants are available for this alloy, therefore the value of Sγ of 100.9

GPa, of a similar alloy, Ni-12.69 Al at 873 K [155] is employed, while the value of Bγ’ is taken

to be 175 GPa [156]. The lattice parameters for the equilibrium phases in alloy (A) at 873 K, are

estimated to be 0.3554 ± 0.0001 nm and 0.3544 ± 0.0001 nm for the γ' and γ-phases,

respectively, based on room-temperature x-ray diffraction measurements on similar Ni-Al-Cr

alloys [12, 17]. These lattice parameter values result in a near-zero estimate of the lattice misfit

of 0.0027 ± 0.0004 for alloy (A). The lattice parameter misfit for alloy (B) is estimated to be

0.0006 ± 0.0004 [42]. Substituting these values into Equation 3.12 yields values of ΔFel of 2.5 x

106 J m-3 for alloy (A) and 1.1 x 105 J m-3 for alloy (B). The larger Cr concentration in alloy (B)

is responsible for a smaller value of the lattice parameter misfit and therefore a smaller elastic

strain energy. The quantity ΔFel is also estimated by a simpler technique due to Eshelby [157],

which yields values of 2.80 x 106 J m-3 and 1.38 x 105 J m-3 for alloys (A) and (B), respectively.

The high degree of coherency of the γ’-precipitates in these alloys is such that the bulk

component of the driving force for nucleation is dominated by the ΔFch term, as ΔFel is only

Page 69: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

69

3.9% of the value of ΔFch for alloy (A) and 0.1% for alloy (B). As such, experimentally

determined differences in the nucleation behavior may be described as due primarily, but not

exclusively, to differing values of ΔFch. The tracer diffusivity of Al in the γ-matrix phase of these

Ni-Al-Cr alloys is significantly larger than the diffusivity of Cr. For alloy (A), fccAlD is 9.6 x 10-21

m2s-1 and fccCrD has a value of 2.6 x 10-21 m2s-1, while for alloy (B), fcc

AlD is 11 x 10-21 m2s-1 and

fccCrD is 3.0 x 10-21 m2s-1.

The value of *RW for alloy (A) is estimated from Equation 3.2 to be 35.2 kJ mol-1 or 0.365

eV atom-1, while R* is estimated from Equation 3.3 to be 0.76 nm. The value of *RW for alloy (B)

is found to be 17.0 kJ mol-1 or 0.177 eV atom-1, and *R is calculated to be 0.55 nm. The

nucleation currents for the two alloys are estimated from Equation 3.4 to be 4.0 x 1022 m-3 s-1 for

alloy (A) and 3.2 x 1023 m-3 s-1 for alloy (B). The calculated values of σγ/γ’, ΔFch, ΔFel, *RW , R*,

and Jst for both alloys are summarized in Table 3.6.

Table 3.6. The interfacial free energy, σγ/γ’, the chemical free energy, ΔFch, and the elastic strain energy, ΔFel, components of the driving force for nucleation used to estimate the net reversible work, *

RW , required for the formation of critical nuclei of size, R*, and the nucleation current, Jst, according to classical nucleation theory.

Alloy σγ/γ’ (mJ m-2)

ΔFch (107 J m-3)

ΔFel (106 J m-3)

*RW

(kJ mol-1) R*

(nm) Jst

(m-3 s-1)

(A) Ni-7.5 Al-8.5 Cr 24.0 ± 6 -6.6 2.5 35.2 0.76 4.0 x 1022

(B) Ni-5.2 Al-14.2 Cr 22.5 ± 7 -8.2 0.11 17.0 0.55 3.2 x 1023

Page 70: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

70

The predictions of the value of R* from CNT are verified experimentally for both alloys,

as the first nucleating precipitates detected consistently by APT at an aging time of 1/6 h have

<R(t)> values of 0.9 ± 0.32 nm for alloy (A) and 0.74 ± 0.24 nm for alloy (B), which are both

slightly greater than the calculated R* estimates of 0.76 and 0.55 nm, respectively. The predicted

value of Jst of 4.0 x 1022 m-3 s-1 for alloy (A) is nearly an order of magnitude greater than the

experimental value of (5.4 ± 1.5) x 1021 m-3 s-1. The calculated value of Jst for alloy (B) of 3.2 x

1023 m-3s-1 is 50 times greater than the experimentally measured value of Jst of (5.9 ± 1.7) x 1021

m-3 s-1. Given the evidence of precursor clustering in these Ni-Al-Cr alloys [45], it is surprising

that the experimentally determined values of Jst are significantly less than the predicted values.

Xiao and Haasen [16] performed a similar comparison of experimentally determined nucleation

currents with those predicted by CNT for a binary Ni-12.0 Al at.% alloy aged at 773 K and found

that the predicted value of Jst was a factor of 500 larger than the measured value of Jst. They

attributed this discrepancy to the sensitivity of the predicted value of Jst to the value of R*. Xiao

and Haasen also pointed out that predicted nucleation currents are likely an overestimate due to

the assumption that the value of N0 is equal to the volume density of lattice sites, which is a

commonly made assumption that is not necessarily correct. By assuming that the value of N0 is

equal to the volume density of Al atoms, we seem to have mitigated this error to some extent.

Given that the experimentally determined values of Jst are measured from two experimental data

points, and that the detailed kinetics involved in the formation of γ’-nuclei in these ternary

systems are not completely understood, further analysis is presently not instructive. The

nucleation behavior of Ni-Al-Cr alloys certainly warrants future research, given the

technological importance of these systems.

Page 71: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

71

The formation of stable, growing nuclei at early aging times causes a decrease in the

values of )(tCiγΔ , and results in a reduction in, and the eventual termination of, nucleation. By

an aging time of 1 h, the values of )(tCiγΔ for alloy (A) decrease from the solid-solution values

of 2.08±0.02 Al and -0.89±0.01 Cr to 1.21±0.02 Al and -0.52±0.04 Cr. This decrease in the

magnitude of the )(tCiγΔ values is reflected in a 50% decrease in the quantity ΔFch to a value of

-3.3 x 107 J m-3, and therefore a decrease in the predicted value of Jst of seven orders of

magnitude, after only 1 h of aging. After 4 h of aging in alloy (B), the values of )(tCiγΔ decrease

from 2.09 ± 0.04 Al and -1.46 ± 0.02 Cr, to 0.84±0.04 Al and -0.59 ± 0.09 Cr, while the value of

ΔFch decreases 51% to -4.01 x 107 J m-3. This decrease in the quantity ΔFch results in a decrease

in the predicted value of Jst of three orders of magnitude for alloy (B). From these estimates, and

our APT nanostructural results, it appears that the larger initial value of the chemical driving

force in alloy (B) sustains a significant nucleation current for aging times as long as 4 h, whereas

the nucleation current decreases significantly in alloy (A) after 1 h.

3.4.2 Effects of solute concentration on the coarsening behavior

As alloys (A) and (B) coarsen, the quantities )(tCiγΔ and )(tCi

γ ′Δ evolve temporally as

approximately t-1/3, and <R(t)> grows as approximately t1/3, as predicted by the UO and KV

models. The experimentally measured temporal dependencies of Nv(t) of t-0.42 ± 0.03. and t-0.67 ± 0.01,

for alloys (A) and (B), respectively, differ significantly, however, from the t-1 prediction of both

models. The magnitudes of the quantities )(tCiγΔ for both alloys remain non-zero at an aging

time of 1024 h, thus neither alloy has achieved a true stationary-state, which is a basic

Page 72: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

72

assumption of classical LSW Ostwald ripening behavior. The UO and KV models, however,

assume a quasi-stationary state, which may have been achieved for alloys (A) and (B). At 1024

h, the value of φ, Figure 3.6, is still evolving temporally.

When Ostwald ripening is considered to be a diffusion-limited process, it is limited by

the characteristic length over which diffusion can occur, taken as the average edge-to-edge

interprecipitate spacing in these systems. It is important to note that as Nv(t) decreases with

increasing time, <R(t)> increases and the value of <λe-e> increases concomitantly. The time

required to reach stationary-state coarsening, tc, may be estimated employing [158]:

3

266

2

3 1172964

DKt KV

c ⎟⎟⎠

⎞⎜⎜⎝

⎛−⎟⎟

⎞⎜⎜⎝

⎛≈

φωω ; (3.13)

where KKV is the coarsening rate constant for <R(t)> as defined in section 1.2, ωi is the ith

moment of the precipitate size distribution, ω2 = 1.046 and ω3 = 1.130 for the LSW distribution

[159], and D is the diffusivity of the least mobile atomic species in the alloy, taken to be the

diffusivity of Cr in the γ’-precipitate phase, which has a value of 1.44 x 10-23 m2 s-1 at 873 K

[160]. The estimate of tc for alloy (A) from Equation 3.13 is (8 ± 5) x 106 h, while the predicted

value of tc for alloy (B) is (2 ± 1) x 106 h, both of which are greater than the longest aging time

studied, 1024 h. Our inability to achieve stationary-state coarsening explains the continuously

increasing values of φ and the non-zero values of )(tCiγΔ and )(tCi

γ ′Δ at an aging time of 1024

h, which may explain the deviation from the t-1 prediction for the temporal evolution of Nv(t).

When D is taken to be the diffusivity of Cr in the γ-matrix phase, 3.0 x 10-21 m2s-1, for alloy (A),

and 2.6 x 10-21 m2s-1 for alloy (B), the estimates of tc (1.0 ± 0.1) h and (0.3 ± 0.1) h are generated

for alloy (A) and (B), respectively, are three orders of magnitude less then the longest aging time

Page 73: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

73

studied experimentally, and indicate that the diffusivity inside the γ’-precipitates may be the rate

limiting step in γ’-precipitate coarsening. Additionally, the values of tc for alloy (A) are longer

than those estimated for alloy (B), an indication that the coarsening kinetics are slower in alloy

(A). Thus the rate of the diminution of Nv(t) is significantly slower in alloy (A) (Nv(t) ∝ t -0.42 ±

0.03) than in alloy (B) (Nv(t) ∝ t -0.67 ± 0.01).

A recent study by Mao et. al [39] combined APT and LKMC to study the role of the

precipitation diffusion mechanism on the early-stage precipitate morphology of alloy (B). They

showed that the long-range solute-vacancy binding energies (out to fourth nearest-neighbor

distance) strongly affect the γ’-precipitate coagulation and coalescence process, which occurs

abundantly at early stages, from 0.25 to 64 h. Coagulation and coalescence are shown to result

from the overlap of nonequilibrium concentration profiles surrounding γ’-precipitates that give

rise to nonequilibrium diffuse interfaces. The concentration profiles associated with the

interfacial regions between γ’-precipitates are spread over distances significantly larger than that

of the equilibrium interfacial thickness. This is due to specific couplings between the diffusion

fluxes of the constituent elements toward and away from γ’-precipitates, which are a result of the

finite vacancy-solute binding energies. From this analysis, the coagulation and coalescence of γ’-

precipitates is more likely when <λe-e> has a minimum value, as evidenced experimentally for

alloys (A) and (B), Figure 3.5. Additionally, the larger Al concentration in alloy (A) leads to the

formation of more highly mobile Al clusters than in alloy (B), which explains why the quantities

Nv(t) and f achieve their maximum values, and the quantity <λe-e> reaches a minimum value,

after aging for only 1 h in alloy (A), while this condition is reached at 4 h in alloy (B).

Page 74: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

74

3.5. Summary and Conclusions

We present a detailed comparison of the nanostructural and compositional evolution of

alloy (A) Ni-7.5 Al-8.5 Cr and alloy (B) Ni-5.2 Al-14.2 Cr, during phase separation at 873 K for

aging times ranging from 1/6 to 1024 h, employing atom-probe topography (APT). These ternary

alloys have similar equilibrium γ’-precipitate volume fractions, φeq, of 16.4±0.6% for alloy (A)

and 15.7±0.7% for alloy (B), and were designed to study the effects of solute concentration on

the kinetic pathways in model Ni-Al-Cr alloys, leading to the following results:

• The morphology of the γ’-precipitate phase in both alloys is found by both APT and TEM

to be spheroidal for aging times as long as 1024 h, as a result of a near-zero lattice

parameter misfit between the γ-matrix and γ’-precipitate phases. This high degree of

coherency makes these alloys amenable to a comparison of APT results to predictions of

classical theories of nucleation, growth and coarsening where the chemical free energy

term is dominant. Coagulation and coalescence of γ’-precipitates is observed and is

argued to be a result of the overlap of the nonequilibrium concentration profiles

associated with adjacent γ’-precipitates [39].

• After aging to 1/6 h, precipitation of the γ’-phase is evident for both alloys, as alloy (A)

forms nuclei with a composition of 70.9 ± 1.4 Ni, 23.3 ± 1.5 Al and 5.8 ± 0.9 Cr, at a γ’-

precipitate number density, Nv(t = 1/6 h), of (2.6 ± 1.4) x 1023 m-3, a mean radius, <R(t =

1/6 h)>, of 0.90 ± 0.32 nm, and a volume fraction, φ, of 0.31 ± 0.11 %. Alloy (B) forms

71.3 ± 1.6 Ni, 19.1 ± 1.4 Al and 9.7 ± 1.1 Cr nuclei with a Nv(t = 1/6 h) value of (3.6 ±

1.3) x 1023 m-3 , an <R(t = 1/6 h)> value of 0.74 ± 0.24 nm, and a φ value of 0.11 ±

0.04%.

Page 75: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

75

• Classical nucleation theory (CNT) is applied to demonstrate that the chemical free energy

changes for forming a nucleus, ΔFch, of -6.6 x 107 J m-3 for alloy (A) and -8.2 x 107 J m-3

for alloy (B), provide the primary driving force for nucleation. As such, the differing

solute concentrations of the two alloys are responsible for differences in the nucleation

behavior and in the nanostructural and compositional evolution of the alloys as they

decompose. The high degree of γ’-precipitate coherency with the γ-matrix in these alloys

is such that the elastic strain energy values, ΔFel, estimated to be -2.5 x 106 J m-3 for alloy

(A), and -1.1 x 105 J m-3 for alloy (B), are only a small fraction of the values of ΔFch, for

both alloys.

• Estimates of the γ/γ’ interfacial free energy values, σγ/γ’, from the coarsening data

obtained by APT yield values of 23-25 ± 6 mJ m-2 for alloy (A) and 22-23 ± 7 mJ m-2 for

alloy (B). These values are in good agreement with our recalculated values of σγ/γ’ of 20-

21 ± 5 mJ m-2 for the coarsening data of Schmuck et al. [28] for a ternary Ni-5.2 Al-14.8

Cr alloy aged at 873 K.

• The predictions of the critical radii for nucleation, R*, of 0.76 nm and 0.55 nm for alloys

(A) and (B), show reasonable agreement with the average radii of the first γ’-precipitates

detected by APT at 1/6 h, of 0.90 ± 0.32 nm and 0.74 ± 0.24 nm, respectively. The

predicted value of Jst of 4.0 x 1022 m-3 s-1 for alloy (A) is nearly one order of magnitude

greater than the experimental value of (5.4 ± 1.5) x 1021 m-3 s-1, and the calculated value

of Jst for alloy (B) of 3.2 x 1023 m-3s-1 is 50 times greater than the experimentally

measured value of Jst of (5.9 ± 1.7) x 1021 m-3 s-1. This discrepancy may be due to the

sensitivity of the predicted value of Jst to the value of R*, and due to the assumption that

Page 76: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

76

the number of nucleation sites per volume is equal to the volume density of lattice points

occupied by Al, the precipitate-forming solute element, which may be an overestimate.

Further research is required to measure the value of Jst more accurately for ternary Ni-Al-

Cr alloys to better understand the nucleation kinetics in these concentrated

multicomponent alloys. To date, there is not a generally accepted theory of nucleation in

concentrated multicomponent alloys.

• After 1 h of aging, alloy (A) achieves a peak value of Nv(t) of (2.21 ± 0.64) x 1024 m-3,

which coincides with a minimum value of the average interprecipitate edge-to-edge

spacing, <λe-e>, of 7.0 ± 2.5 nm and a maximum in the fraction of interconnected

precipitates, f, of 18 ± 4 %. At an aging time of 4 h, alloy (B) achieves a peak value of

Nv(t) of (3.2 ± 0.6) x 1024 m-3 at a minimum value of <λe-e> of 5.9±0.8 nm and a

maximum value of f of 30 ± 4 %. The larger peak value of Nv(t) for alloy (B) is a result of

the larger initial solid-solution value of ΔFch in alloy (B), which sustains nucleation for

longer aging times.

• In the growth and coarsening regime that follows the peak in the value of Nv(t), the

quantity <R(t)> displays a temporal dependence of t0.29 ± 0.03 for alloy (A) and t0.29 ± 0.05

for alloy (B), in approximate agreement with the t1/3 prediction of the Kuehmann and

Voorhees (KV) model for isothermal coarsening in concentrated ternary alloys. The

supersaturation of the γ-matrix phase in alloy (A) exhibits a temporal dependence of t-0.32

± 0.03 for Al and t-0.29 ± 0.04 for Cr, while alloy (B) demonstrates a temporal dependence of t-

0.33 ± 0.04 for Al and t-0.34 ± 0.07 for Cr, in good agreement with the t1/3 prediction of the UO

and KV models. The supersaturation values of Al and Cr in the γ’-precipitate phase of

Page 77: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

77

Alloy (A) display a temporal dependence of t-0.34 ± 0.04 for Al and t-0.32 ± 0.05 for Cr, while

the supersaturation values in alloy (B) demonstrate a temporal dependence of t-0.30 ± 0.05

for Al and t-0.29 ± 0.07 for Cr.

• During growth and coarsening, the diminution of the quantity Nv(t) in alloy (A) exhibits a

temporal dependence of t-0.42 ± 0.03, while alloy (B) exhibits a temporal dependence of t-0.67

± 0.01. These temporal exponents deviate from the t-1 prediction of the KV coarsening

model because neither alloy has achieved stationary-state coarsening, characterized by a

constant value of φ and a zero value of the γ-matrix supersaturation. The estimate of the

critical time required for stationary-state coarsening, tc, for alloy (A) is (8 ± 5) x 106 h,

while for alloy (B) the value is (2 ± 1) x 106 h. The estimates of tc for both alloys are well

beyond the longest aging time we studied, 1024 h, and this is most likely the reason that

stationary-state coarsening is not achieved.

• The compositional trajectories of the γ-matrix during phase decomposition lie

approximately along the tie-lines, while the trajectories of the γ’-precipitate phase do not,

as predicted by the KV model for quasi-stationary state isothermal coarsening in ternary

alloys. The addition of a third alloying element alters the Gibbs-Thompson equations

significantly, and as such the KV model predicts that the compositional trajectory of the

γ’-precipitate phase will lie on a straight line that is not necessarily parallel to the

equilibrium tie-line.

Page 78: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

78

Chapter 4

On the field evaporation behavior of a model Ni-Al-Cr superalloy studied by

picosecond pulsed-laser atom-probe tomography

Abstract

The effects of varying the pulse energy of a picosecond laser used in the pulsed-laser atom-probe

(PLAP) tomography of an as-quenched Ni-6.5 Al-9.5 Cr at.% alloy are assessed based on the

quality of the mass spectra and the compositional accuracy of the technique. Compared to

pulsed-voltage atom-probe tomography (APT), PLAP tomography improves mass resolving

power, decreases noise levels, and improves compositional accuracy. Experimental evidence

suggests that Ni2+, Al2+ and Cr2+ ions are formed primarily by a thermally activated evaporation

process, and not by post-ionization of the ions in the 1+ charge state. An analysis of the detected

noise levels reveals that for properly chosen instrument parameters, there is no significant

steady-state heating of the Ni-6.5 Al-9.5 Cr at.% tips during PLAP tomography.

Page 79: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

79

4.1 Introduction

Pulsed-laser atom-probe (PLAP) tomography permits atom-probe tomography (APT)

with increased mass resolving power and pulse repetition rate, and allows the study of materials

with poor electrical conductivity [161-163]. PLAP tomography extends the range of study of the

APT technique, though there are some inherent drawbacks, such as inaccurate quantitative

results which can arise due to the complexity of the evaporation mechanism [164]. Additionally,

surface diffusion can occur if heating of samples is excessive under laser illumination. PLAP

tomography has become an increasingly popular technique for the study of metals and

semiconductors [165-169], thus a quantitative understanding of the effects of pico- or

femtosecond laser pulsing on the evaporation behavior of investigated materials is essential to

obtain physically meaningful results.

The extant theories of field evaporation, the image hump theory [170] and the charge

exchange theory [171-173], were successful in predicting both the correct charge state of

evaporating atomic species, and the fields required for evaporation, though refinement was

required [163, 174]. The ionic model upon which these theories are based treats the temperature

dependence of the evaporation field to be exclusively from the heat of sublimation, which from

Kirchhoff’s equation, displays very little temperature dependence from absolute zero to the

melting point of metals [175]. Hence, the ionic model is incapable of explaining the observed

experimental temperature dependencies of the evaporation field, and the calculated values of the

evaporation field in the literature are essentially 0 K values. While no fully satisfactory

theoretical treatment of field evaporation is currently available [176, 177], modification of the

image hump theory of field evaporation led to an expression for the temperature sensitivity of the

Page 80: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

80

evaporation field at constant evaporation rate [174, 178, 179], which predicted an increase in the

rate of evaporation with increasing temperature and was validated experimentally [174, 176,

180-185]. Currently, the mechanism by which laser pulsing induces field evaporation remains

somewhat controversial, particularly for femtosecond laser pulsing [104, 186-190]. We note that

there is convincing evidence of heating of the tip under laser illumination with picosecond

pulsing in other fields, such as scanning tunneling electron microscopy [191, 192]. The

theoretical and experimental work performed to measure the temperature rise due to laser

illumination in field-ion microscopy has been reviewed in detail [193], though more work

remains to be done, particularly for the new generation of atom-probes using pico- and

femtosecond laser pulsing.

In pulsed-voltage APT, variables such as sample preparation, instrument specifications,

and the instrumentation parameters selected to evaporate ions, contribute to the quality of the

data collected. Material- and specimen-dependent parameters, such as the evaporation behavior

of the phases being studied, and the crystallographic direction being examined, are also of

critical importance and must be considered on a sample-by-sample basis. With the continued

improvement in both instrument capability and sample preparation techniques [99, 100, 104,

165, 166], optimization of data quality must be performed by fine tuning the instrumentation

parameters, specifically: (1) specimen base temperature; (2) pulse fraction, the ratio of the pulse

voltage to the steady-state voltage; (3) the specimen evaporation rate, the number of ions

evaporated per pulse, and (4) the pulse repetition rate. This has been performed extensively for

pulsed-voltage atom-probe microscopy; see for example [194-196] who used a time-of-flight

atom-probe field-ion microscope [197]. Little, however, has been published on the quantification

Page 81: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

81

of the effects of the controllable experimental parameters in PLAP tomography [164],

particularly for the new generation of high-performance instruments employing ultrafast pico-

and femtosecond lasers. The instrumentation parameters for PLAP tomography are the same as

those for pulsed-voltage APT, though the pulse fraction is replaced by the effective pulse

fraction, which is not a directly tunable instrumentation parameter. The effective pulse fraction is

defined as one minus the ratio of the steady-state tip voltage during laser-assisted data

acquisition to the steady-state dc tip voltage required to field evaporate ions from a specimen at a

rate of 103 ions per second [104]. The laser adds a further degree of complexity by introducing

instrumentation variables such as the laser pulse energy, spot diameter, wavelength, and pulse

widths. In the case of the LEAP 3000X Si™ instrument used in this study, these variables have

been set by the instrument manufacturers [104], with the exception of the laser energy, which

remains a variable instrumentation parameter under the control of the experimentalist.

It has been demonstrated that decreasing the laser pulse repetition rate and increasing the

shank angle of specimen tips improves mass resolving power, m/Δm, the signal-to-noise ratio,

S/N, and the compositional accuracy of PLAP tomography for materials with small values of the

thermal diffusivity [104, 186]; the thermal diffusivity is equal to the ratio of the thermal

conductivity to the product of the density and the specific heat at constant pressure. Little work

has been performed to model the field evaporation of atoms in concentrated alloys because the

relevant thermodynamic quantities are generally unknown [176], though some early work was

performed to model the field evaporation of solute atoms in dilute alloys of iron [198]. The

present study determines the effects of the laser pulse energy on the evaporation behavior of a

concentrated model Ni-Al-Cr superalloy. The quality of the mass spectra and the compositional

Page 82: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

82

accuracy obtained for the same specimen with varying picosecond pulse energies are compared

at constant base specimen temperature. These results are compared with those of a test of the

effect of specimen base temperature on the evaporation behavior of the Ni-Al-Cr alloy,

conducted by performing pulsed-voltage APT at different tip temperatures.

4.2 Materials and methods

Ni-6.5 Al-9.5 Cr at.% solid-solutions were homogenized at 1573 K for 20 hours,

solutionized at 1123 K for 3 hours then water quenched to room temperature. The as-quenched

solid-solutions did not contain γ’-precipitates, and the chemical composition of the as-quenched

samples was measured by inductively coupled plasma (ICP) atomic-emission spectroscopy to be

Ni-6.24 Al-9.64 Cr at.%. Pulsed-laser and pulsed-voltage APT were performed employing an

Imago Scientific Instrument’s local-electrode atom-probe (LEAP™) tomograph [101-103], and

APT data were analyzed employing the IVAS® 3.0 software program (Imago Scientific

Instruments, Madison, Wi).

The data presented herein are obtained from two Ni-6.5 Al-9.5 Cr at.% tips; the first was

analyzed by PLAP tomography to study the effects of the laser pulse energy, while the second tip

was analyzed by pulsed-voltage APT to study the effects of specimen base temperature on the

evaporation behavior of the Ni-Al-Cr alloy. Datasets of ~5 million atoms were collected for each

laser pulse energy and specimen base temperature studied, and the first 0.5 million atoms were

discarded due to the initial instability of evaporation upon application of the pulsed-laser or

pulsed-voltage source. The depth of analysis of the datasets collected by pulsed-laser APT was

constant at 21±3 nm, with cross-sectional areas of 3,800±300 nm2, as determined by calculations

Page 83: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

83

based on the theoretical density of this alloy, with the tip radius estimated from the evaporation

field of pure Ni at 73 K. PLAP tomography was performed at a constant specimen base

temperature of 40±0.3 K and ambient pressure of < 6.7 x 10-8 Pa, a pulse repetition rate of 200

kHz, a specimen evaporation rate of 0.04 ions per pulse, and laser energies of 0.2, 0.4, 0.6, 0.8,

1.0, 1.2 or 2.0 nJ. A laser with a wavelength of 532 nm, a pulse duration of ~ 12 ps, and a 5 μm

spot size was used. Pulsed-voltage APT at a specimen base temperature of 40±0.3 K was

performed after the PLAP tomography runs at laser pulse energies of 0.6, 1.0 and 2.0 nJ in order

to study the cumulative long-range effects of laser pulsing on the alloy composition.

Pulsed-voltage APT of the second tip was performed at a constant background pressure

of < 6.7 x 10-8 Pa, a pulse repetition rate of 200 kHz, a voltage pulse fraction (pulse

voltage/steady-state DC voltage) of ~19%, a constant evaporation rate of 0.04 ions per pulse, and

specimen base temperatures of 40, 60, 80, 100, 125, 150 and 175±0.3 K. The depth of analysis of

the datasets collected by pulsed-voltage APT was constant at 27±2 nm, with cross-sectional areas

of 2,200±200 nm2, as determined above. The steady-state tip voltages required to achieve a

specimen evaporation rate of 0.04 ions per pulse for both pulsed-laser and pulsed-voltage APT

are shown in Figure 4.1. We note that the tips had different initial radii of curvature and taper

angles, and hence, under the same field evaporation conditions, the specimens required different

steady-state voltages to evaporate atoms at the same rate. As such, results from the two Ni-Al-Cr

tips studied herein are not comparable in a quantitative manner, though trends in the evaporation

behavior provide insights into the effects of the picosecond laser pulse energy. The differences in

the tip shapes and dimensions were such that the evaporation rates of the two tips differed when

normalized to the effective area subtended by the detector. Normalized evaporation rates of

Page 84: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

84

1.1±0.1 x 10-5 ions pulse-1 nm-2 and 1.8±0.2 x 10-5 ions pulse-1 nm-2 were achieved for pulsed-

laser and pulsed-voltage APT, respectively.

Ni-Al-Cr samples in the as-quenched state were studied in order to avoid γ’(L12)-

precipitation due to the problems associated with differences in the evaporation fields of the

atomic species in the γ(FCC)-matrix and γ’-precipitate phases. No γ’-precipitates are detected in

the APT reconstructions, though partial radial distribution functions (RDFs) [45] of the as-

quenched samples indicate small Ni-Al ordered domains, which evolve into γ’-precipitates upon

subsequent aging. We note that deconvolution of the overlapping 27Al1+ and 54Cr2+ peaks was

performed based on the relative abundances of Cr isotopes. The standard errors, 2σ, for all

quantities are calculated based on counting statistics and reconstruction scaling errors, using

standard error propagation methods [140], and represent two standard deviations from the mean.

Page 85: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

85

Figure 4.1. The steady-state dc voltage applied to the tip to maintain a specimen evaporation rate of 0.04 ions per pulse for a Ni-6.5 Al-9.5 Cr at.% alloy is shown as a function of: (a) the laser pulse energy applied in PLAP tomography at a specimen base temperature of 40±0.3 K and a pulse repetition rate of 200 kHz; and (b) the specimen base temperature used in pulsed-voltage APT at a pulse repetition rate of 200 kHz and a pulse fraction of ~20%.

4.3 Results and discussion

4.3.1 Mass spectra

Figure 4.2 shows a comparison of the mass spectra obtained by PLAP tomography with

laser energies of: (a) 0.2 nJ per pulse and a steady-state tip voltage of 7.40-7.70 kV; (b) 1.0 nJ

per pulse and a tip voltage of 8.0-8.3 kV; and (c) 2.0 nJ per pulse and a tip voltage of 6.55-6.85

Page 86: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

86

kV. The amount of noise increases with increasing laser pulse energy; the percentage of

evaporation events that are detected in expected ranges decreases from 97.24 % at 0.2 nJ, to

95.15 % at 1.0 nJ, to 93.96 % at 2.0 nJ. We note that at a laser pulse energy of 2.0 nJ, the mass-

to-charge state, m/n, spectrum is populated by numerous metal hydride peaks, particularly Ni1+

hydrides, and unexpected peaks at m/n ratios that correspond to hydrides, oxides and metal ion

clusters, which are probably a result of overheating of the tip. Because these peaks affect the

quality of the compositional data obtained by PLAP tomography, laser energies greater than 1.0

nJ per pulse should be avoided for this alloy to minimize background noise and undesirable

peaks. A section of the m/n range displaying the Ni2+ peaks is presented in Figure 4.3, to

highlight both the improved m/Δm values of PLAP tomography when compared to pulsed-

voltage APT, and the improved m/Δm values at higher laser pulse energies.

Page 87: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

87

Figure 4.2. A comparison of the mass spectra obtained by PLAP tomography with laser pulse energies of: (a) 0.2 nJ per pulse at a tip voltage of 7.40-7.70 kV; (b) 1.0 nJ per pulse at a tip voltage of 8.0-8.3 kV; and (c) 2.0 nJ per pulse at a tip voltage of 6.55-6.85 kV. The frequencies of detected events are normalized to the number of events detected in the most populous m/n bin. The m/n ratios of the evaporating species change by over four orders of magnitude with increasing laser pulse energy. At a laser pulse energy of 2.0 nJ per pulse, the m/n spectrum is populated by numerous hydrides, oxides and metal ion clusters peaks, which are a result of overheating and are undesirable.

Page 88: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

88

Figure 4.3. A section of the m/n range showing the Ni2+ peaks, highlighting both the improved mass resolving power of PLAP tomography when compared to pulsed-voltage APT, and the improved mass resolving power with increasing laser pulse energy. The relative frequencies of the Ni2+ peaks from the data set collected at a laser pulse energy of 2.0 nJ per pulse are nearly two orders of magnitude smaller than the relative frequencies of the Ni2+ peaks from the set collected at 0.2 nJ per pulse.

The most striking feature of Figures 4.2 and 4.3 is the dramatic change in the m/n ratios

of the evaporating species with increasing laser pulse energy. The relative frequencies of the Ni2+

peaks from the data set collected at a laser energy of 2.0 nJ per pulse, Figure 4.3, are nearly two

orders of magnitude less than the relative frequencies of the Ni2+ peaks at 0.2 nJ per pulse. The

Ni atoms are observed to evaporate as Ni1+ ions with increasing regularity as the laser pulse

energy is increased. This trend is observed to hold for all the atomic species to varying degrees,

Figure 4.4. While Ni and Al have similar evaporation behavior in terms of the m/n ratio, Cr

evaporates exclusively as Cr2+ until the laser energy reaches 1.0 nJ per pulse, at which point the

trend of the Cr2+/Cr1+ ratio with increasing laser pulse energy is similar to that of Ni2+/Ni1+ and

Page 89: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

89

Al2+/Al1+. Similar trends have been observed previously for other materials such as pure W, Si,

Ni, Rh, Mo, Re, Ir and Pt [162, 183, 186, 199]. High charge states on the order of 5+ and 6+

were detected for 5d transition metals evaporated at very high electric fields [200].

Figure 4.4. The ratio of the ions detected in the 2+ charge state to those detected in the 1+ charge state, as a function of laser pulse energy. The Ni and Al ions appear to have similar evaporation behavior, whereas Cr atoms evaporate exclusively as Cr2+ until the laser energy reaches 1.0 nJ per pulse.

It was previously suggested that field evaporated ions with lower than predicted m/n

ratios were a result of post-ionization of desorbed ions by an electron tunneling mechanism

similar to that involved in field ionization of gas atoms above the surface of the tip [201-205]. It

was reasoned that the increased electric field required to maintain a constant evaporation rate at

smaller laser pulse-energies led to progressively more post-ionization of atoms, and to the

appearance of lower m/n ratios. In Kellogg’s PLAP tomography experiment with tungsten, for

example, the ions field evaporated predominantly as W3+ at voltages greater than 3.6 kV, while

Page 90: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

90

they evaporated as predominantly W2+ at voltages less than 2.9 kV [199]. For the Ni-Al-Cr alloy

studied here, when the laser pulse energy is varied between 0.4 and 1.0 nJ, the Ni2+/Ni1+ and

Al2+/Al1+ ratios both decrease by more than two orders of magnitude. Over this range of laser

pulse energies, the tip voltage required to maintain an evaporation rate of 0.04 ions per pulse

only varies between 8.15 and 8.30 kV. Given that the applied electric field varies by no more

than 2% over this range of laser energies, it is difficult to ascribe the dramatic changes in the

charge state ratios exclusively to post-ionization. Additionally, if post-ionization is responsible

for the higher charge states, the conditions required to post-ionize Ni1+ and an Al1+ ions differ

significantly from those required to post-ionize Cr1+. Thus, it may be that post-ionization is not

solely responsible for the change in the charge state of the evaporated species in the Ni-Al-Cr

alloy. We note that the average depth of the datasets was 21±3 nm, thus the change in the radius

of curvature, which affects the local electric field at the tip surface, is not responsible for a

decrease in the ratio of the Ni2+/Ni1+ and Al2+/Al1+ of more than four orders of magnitude over

the range of pulse energies studied herein.

Previous work with PLAP tomography at constant steady-state tip voltage and varying

specimen base temperatures showed that temperature had no effect on the m/n ratio of

evaporating species [183, 199]. These results were interpreted to mean that the activation

energies of desorption for ions of different charge states are identical, and thus provided

evidence for the post-ionization model. To quantify the importance of post-ionization in our

model ternary alloy, a test at a constant specimen evaporation rate of 0.04 ions per pulse and

varying specimen base temperature was performed for a second Ni-6.5 Al-9.5 Cr at.% tip to

understand the effect of specimen base temperature on the evaporation behavior of this model

Page 91: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

91

Ni-Al-Cr alloy. The mass spectra from pulsed-voltage APT for specimen base temperatures

ranging from 40-125 ±0.3 K do not exhibit any evidence of evaporation of any of the atomic

species in the 1+ charge state. At specimen base temperatures of 150 and 175±0.3 K, however,

when the tip voltage is a constant 5.2±0.2 kV, Ni1+ ions are detected, and increase in relative

frequency of occurrence with increasing temperature, Figure 4.5. Thus, Ni atoms at specimen

base temperatures below 150 K seem to field evaporate as Ni2+ ions, and do not to result from

post-ionization of Ni1+ ions. For the evaporation conditions employed herein, Ni1+ ions field

evaporate by a thermally activated process at temperatures of ~150 K and greater. Neither Al1+

nor Cr1+ ions are detected by pulsed-voltage APT, suggesting that, for the given evaporation

conditions, the base specimen temperatures chosen herein are insufficient for evaporation of

these ionic charge states. From these results, the change in the abundances of the charge states

with increasing laser pulse energy in PLAP tomography, Figure 4.4, may be partly or completely

due to thermally activated field evaporation of ions in the 1+ charge state at the higher tip

temperatures that result from laser illumination.

Page 92: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

92

Figure 4.5. The relative frequencies of the Ni1+ peaks at specimen base temperatures of 125, 150 and 175±0.3 K, at a constant steady-state dc tip voltage of 5.2±0.2 kV. The increase in the relative frequency of Ni1+ ions with increasing specimen base temperature suggests that the evaporation of Ni1+ ions occurs by a different thermally activated mechanism than the evaporation of Ni2+ ions.

4.3.2 Compositional accuracy: Effects of laser energy

The concentrations of Ni, Al and Cr measured by PLAP tomography as a function of

laser pulse energy are presented in Figure 4.6, and clearly demonstrate that the measured

composition of the alloy differs with increasing laser pulse energy. Interestingly, the alloy

composition differs from both the nominal composition obtained by ICP chemical analysis, and

the values measured by pulsed-voltage APT after the PLAP tomography runs at laser pulse

energies of 0.6, 1.0 and 2.0 nJ. The compositions measured by pulsed-voltage APT were found

to be 82.94±0.08 Ni, 6.98±0.02 Al, and 10.07±0.03 Cr at.% and did not vary beyond

experimental error. Thus even at the largest laser pulse energy, 2.0 nJ, the laser power has no

cumulative, long-range effect on the tip composition.

The compositions measured by PLAP tomography are closer to the nominal composition

Page 93: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

93

of the alloy than those measured by pulsed-voltage APT of the same tip at 40±0.3 K, Figure 4.6.

Previous studies of Ni-Al-Cr alloys by pulsed-voltage APT showed evidence of preferential field

evaporation of Ni atoms resulting in differences between the expected and measured Ni

concentration of 2.7 at.% for Ni-5.23 Al-14.77 Cr at.% [112], and 3.0 at.% for Ni-5.2 Al-14.2 Cr

at.% [42, 43, 46, 50]. These differences are greater than the 1.18±0.08 at.% Ni measured here by

pulsed-voltage APT. At a pulse energy of 0.6 nJ, the difference between the nominal Ni

concentration and the value measured by PLAP tomography is 0.22±0.14 at.%, suggesting that

for the experimental conditions used herein, preferential evaporation is less severe in PLAP

tomography than in voltage-pulsed APT, leading to improved compositional accuracy. The value

of the effective pulse fraction at a laser pulse energy of 0.6 nJ per pulse is ~20%, which is close

to the pulse fraction of ~19% that was determined to minimize preferential evaporation in Ni-Al-

Cr systems in a pulsed-voltage APT [43, 50, 112]. In both voltage and PLAP tomography, the

pulse fraction, or effective pulse fraction, respectively, have been recognized as the most

important instrumentation parameter for controlling preferential evaporation; values of 15-20%

are often cited as giving the most accurate compositional results [104]. In the case of PLAP

tomography, an excessive effective pulse fraction can lead to overheating of the specimen tip,

causing diffusion on the tip surface and resulting in a loss of spatial resolution [161, 186, 193].

The most striking feature of Figure 4.6 is the variation of the concentrations of the

constituent elements with laser pulse energy during PLAP tomography. At laser pulse energies of

0.2-0.6 nJ, the measured Ni concentration increases, and the Al and Cr concentrations decrease,

with increasing pulse energy. This suggests that the severity of Ni preferential evaporation

decreases with increasing laser pulse energy, leading to a decrease in the detected concentration

Page 94: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

94

of solute atoms. At laser energies greater than 0.6 nJ per pulse, the Ni and Al concentrations do

not vary beyond the experimental error. The measured Cr concentration increases with

increasing pulse energy at energies of 1.0 nJ per pulse and greater. At a laser pulse energy of 2.0

nJ, the Cr concentration has increased by a statistically significant 0.11±0.08 at.% Cr value over

the value obtained for 0.6 nJ per pulse. The increase in the Cr concentration at 1.0 nJ per pulse

and greater, coincides with the evaporation of Cr1+, Figure 4.4. This effect cannot be ascribed to

post-ionization, but may be a result of a thermally activated evaporation mechanism for Cr1+.

4.3.3 Compositional accuracy- Effects of specimen base temperature

The effect of varying the specimen base temperature on the alloy composition measured

by pulsed-voltage APT is displayed in Figure 4.7. The measured Ni concentration increases with

increasing temperature, reaching the nominal composition at a temperature of 175±0.3 K. Over

the specimen base temperature range of 40–175 ±0.3 K, the measured Al concentration remains

constant, while the Cr concentration decreases by 1.43±0.04 at.% Cr. As the specimen base

temperature increases, and the tip voltage reaches a plateau value, Figure 4.1a, and Ni

preferential evaporation is diminished, while the preferential evaporation of Cr increases. This is

a result of differences in the required evaporation fields of Ni and Cr at different temperatures.

Page 95: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

95

Figure 4.6. Variation in the detected composition of as-quenched Ni-6.5 Al-9.5 Cr at.% as a function of laser pulse energy (solid line) compared to the concentration measured by pulsed-voltage APT (dashed line) and the nominal composition of the alloy (dotted line). Preferential evaporation of Ni is responsible for the inaccuracy in the concentrations measured by pulsed-laser and pulsed-voltage APT.

Page 96: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

96

Figure 4.7. Variation of the detected composition of as-quenched Ni-6.5 Al-9.5 Cr at.% as a function of specimen base temperature by pulsed-voltage APT. At higher temperatures, preferential evaporation of Ni is less severe, while preferential evaporation of Cr is more severe.

Page 97: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

97

4.3.4 Analysis of events detected at mass-to-charge state ratios of 100-300 amu

An analysis of the events detected at m/n ratios between 100-300 amu, ranges where

peaks are neither expected nor observed for this Ni-Al-Cr alloy, provides some insight into

random field evaporation due exclusively to the steady-state dc tip voltage. Figure 4.8a shows

that for PLAP tomography, the extent of random evaporation is directly proportional to the

steady-state dc tip voltage, and does not increase with increasing pulse laser energy, and

concomitantly increasing input thermal energy. The events detected in this same mass range for

pulsed-voltage APT increase in frequency with increasing specimen base temperature, Figure

4.8b, as a result of the decrease in the evaporation fields required for evaporation of the atomic

species. This direct experimental result demonstrates clearly that results from PLAP tomography

with picosecond laser pulsing for this Ni-Al-Cr alloy are not affected by significant steady-state

heating of the tip. This is consistent with previous calculated estimates of the increase in tip

temperature on the order of 200-300 K due to laser illumination [182, 188], and estimates of tip

cooling times that are on the order of hundreds of nanoseconds [206].

The relative frequency of detected evaporation events with m/n ratios of 100-300 amu in

the three pulsed-voltage APT measurements made between pulsed-laser analyses, averaged

1.3±0.1 % of the total spectrum at an average tip voltage of 10.5±1.0 kV. This should be

compared to a value of 1.0±0.1 % of the total spectrum measured for the same tip, run by PLAP

tomography at a steady-state dc voltage of 8.0-8.3 kV, for a laser energy of 0.6 nJ per pulse, and

an equivalent pulse fraction of ~20%. Evaporation between pulses is less severe in PLAP

tomography due to the lower steady-state dc voltage required to achieve a field evaporation rate

of 0.04 ions per pulse, minimizing the preferential evaporation of Ni, as previously discussed.

Page 98: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

98

Figure 4.8. The relative frequency of detected events between 100-300 amu, as a function of: (a) laser pulse energy in PLAP tomography; and (b) specimen base temperature in pulsed-voltage APT. The amount of noise measured by PLAP tomography is directly proportional to the tip voltage, and does not increase with increasing laser pulse energy, and thus increasing thermal energy. On the contrary, the noise measured by pulsed-voltage APT increases as the specimen base temperature increases. This result demonstrates that the Ni-Al-Cr tips are not affected by significant steady-state heating during PLAP tomography.

Page 99: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

99

4.4 Conclusions

The effects of picosecond laser pulse energy on the compositional accuracy and the

quality of the mass spectra obtained for as-quenched Ni-6.5 Al-9.5 Cr at.% solid-solutions

analyzed by PLAP tomography are assessed, leading to the following conclusions:

• For the conditions used herein, it is found that when compared to pulsed-voltage APT,

PLAP tomography yields improved mass resolving power and compositional accuracy

and decreased noise levels.

• The most accurate compositional results are found at a laser pulse energy of 0.6 nJ per

pulse, a specimen base temperature of 40±0.3 K, a pulse repetition rate of 200 kHz, and a

specimen evaporation rate of 0.04 ions per pulse; these conditions correspond to an

effective pulse fraction of ~20%.

• An analysis of the events detected at m/n ratios of 100-300 amu by PLAP tomography

with varying laser pulse energies shows no evidence for steady-state heating of the tips

due to laser illumination.

• The laser pulse energy is found to have an important impact on the observed m/n values

of the evaporated species. The Ni1+, Al1+ and Cr1+ ions are observed to evaporate

predominantly at higher laser energies, while at lower laser pulse energies, Ni2+, Al2+ and

Cr2+ ions are predominantly detected, an effect which had previously been ascribed to

post-ionization. The post-ionization model, however, is unable to account for: (i) the

decrease in the Ni2+/Ni1+ and Al2+/Al1+ ratios by more than two orders of magnitude when

the laser energy is varied between 0.4-1.0 nJ per pulse at steady-state tip voltages ranging

from 8.15-8.30 kV; (ii) the detection of Ni1+ ions in pulsed-voltage APT exclusively at

Page 100: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

100

specimen base temperatures above 150±0.3 K with a constant steady-state dc tip voltage.

The latter indicates that the evaporation of Ni1+ occurs by a thermally activated process at

higher temperatures, and thus it is unlikely that the evaporation of Ni2+, Al2+ and Cr2+

ions is due primarily to post-ionization of ions in the 1+ charge state.

• This study demonstrates the importance of selecting the correct instrumentation

parameters to achieve compositional accuracy for a Ni-Al-Cr alloy, analyzed by both

pulsed-laser and pulsed-voltage APT. A similar investigation is recommended for all

materials systems, particularly those with small thermal diffusivities, prior to in-depth

APT analysis, to achieve optimal evaporation conditions and the highest degree of

compositional accuracy.

Page 101: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

101

Chapter 5

On the nanometer scale phase decomposition of a low-supersaturation Ni-Al-

Cr alloy

Abstract

The phase decomposition of a Ni-6.5 Al-9.5 Cr at.% alloy aged at 873 K is studied by atom-

probe tomography and compared to the predictions of classical precipitation models. Phase

separation in this alloy occurs in four distinct regimes: (i) quasi-stationary-state γ’ (L12)-

precipitate nucleation; (ii) concomitant precipitate nucleation, growth, and coagulation and

coalescence; (iii) concurrent growth and coarsening, wherein coarsening occurs via both γ’-

precipitate coagulation and coalescence and by the classical evaporation-condensation

mechanism; and (iv) quasi-stationary-state coarsening of γ’-precipitates, once the equilibrium

volume fraction of precipitates is achieved. The predictions of classical nucleation and growth

models are not validated experimentally, likely due to the complexity of the atomistic kinetic

pathways involved in precipitation. During coarsening, the temporal evolution of the γ’-

precipitate average radius, number density, and the γ(FCC)-matrix and γ’-precipitate

supersaturations follow the predictions of classical models.

Page 102: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

102

5.1 Introduction

Efforts to improve the fuel efficiency of aviation fuel and natural gas burning turbine

engines have led to the development of complex concentrated multicomponent nickel-based

superalloys that can withstand the extreme environments inside these engines [3, 7]. Nickel-

based superalloys are used in sections of turbine engines where operating temperatures exceed

1073 K, and resistance to chemical and mechanical degradation is critical. The high-temperature

mechanical properties of these alloys depend primarily on the presence of coherent, elastically-

hard, L12-ordered γ’-precipitates that result from decomposition of the γ(FCC) matrix.

Recent investigations of γ’-precipitation in model Ni-Al-Cr superalloys employing atom-

probe tomography (APT) have elucidated the kinetic pathways of phase decomposition for two

alloys, Ni-7.5 Al-8.5 Cr and Ni-5.2 Al-14.2 Cr at.%, both of which have a γ’-phase volume

fraction of about 16% [28, 37, 38, 42, 45, 46, 52, 112]. These studies have, however, been

limited by both the number of γ’-precipitates investigated, and by the number of data points

collected during the γ’-precipitate nucleation and growth regimes. Recent advances in atom-

probe tomography have improved the data collection rate and the field-of-view (200x200 nm2 at

the lowest magnification) of this technique [99, 104, 165], thereby eliminating the problems

associated with an insufficient γ’-precipitate sample size.

We present details of the phase decomposition of a low-supersaturation Ni-6.5 Al-9.5 Cr

at.% alloy aged at 873 K. This alloy was chosen to employ a small chemical driving force for

nucleation, permitting a detailed study of precipitate nucleation, and subsequently growth and

coarsening. The reduced γ-matrix supersaturation also results in a smaller γ’-precipitate volume

fraction, making this alloy more amenable to comparison with the predictions of classical models

Page 103: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

103

of phase decomposition. A ternary Ni-Al-Cr phase diagram determined by the Grand Canonical

Monte Carlo (GCMC) technique at 873 K, Figure 5.1, predicts that the value of the equilibrium

volume fraction, φeq, for Ni-6.5 Al-9.5 Cr at.% aged at 873 K is 10.2±0.5 %. Ni-7.5 Al-8.5 Cr

and Ni-5.2 Al-14.2 Cr at.%, the other Ni-Al-Cr alloys that have been investigated in detail by

APT, are also plotted in Figure 5.1, and are predicted to have values of φeq of 17.5±0.5 and

15.1±0.5 %, respectively [52, 113].

Figure 5.1. A partial ternary phase diagram of the Ni-Al-Cr system at 873 K calculated using the Grand Canonical Monte Carlo simulation technique [113], showing the proximity of Ni-6.5 Al-9.5 Cr at.% to the (γ + γ’) / γ solvus line. Two other alloys that have been investigated by APT, Ni-7.5 Al- 8.5 Cr and Ni-5.2 Al-14.2 Cr at.%, are shown for comparative purposes. Equilibrium solvus curves determined by Thermo-Calc [139], using databases for nickel-based superalloys due to Saunders [115] and Dupin et al. [116], are superimposed on the GCMC phase diagram. The tie-lines are determined from the equilibrium phase compositions, determined by extrapolation of APT concentration data to infinite time.

Page 104: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

104

5.2 Experimental

High-purity Ni, Al and Cr were induction-melted under a partial pressure of Ar and chill

cast in a 19 mm diameter copper mold to form a polycrystalline master ingot. The overall

composition of the master ingot was determined by inductively coupled plasma atomic-emission

spectroscopy to be 84.12 Ni-6.24 Al-9.64 Cr at.%, and was indistinguishable, within

experimental error, from the targeted composition of Ni-6.5 Al-9.5 Cr at.%. Samples from the

ingot were subjected to a three-stage heat-treatment: (1) homogenization at 1573 K in the γ-

phase field for 20 h; (2) a vacancy anneal in the γ−phase field at 1123 K for 3 hours followed by

a water quench; and (3) an aging anneal at 873 K under flowing argon for times ranging from

0.25 to 4096 h, followed by a water quench. Pulsed-laser APT was performed with a 3-D

LEAPTM tomograph [100-104] at a target evaporation rate of 0.04 ions per pulse, a specimen

temperature of 40.0±0.3 K, a pulse energy of 0.6 nJ per pulse, a pulse repetition rate of 200 kHz,

and an ambient gauge pressure of less than 6.7 x 10-8 Pa. These evaporation conditions have been

optimized to provide the highest compositional accuracy for this alloy [64]. APT data were

analyzed with the software program IVAS 3.0 (Imago Scientific Instruments). The γ’-

precipitate/γ-matrix heterophase interfaces were delineated with Al isoconcentration surfaces

employing the inflection point method [55], and compositional information was obtained using

the proximity histogram methodology [136, 137]. The standard errors for all quantities are

calculated based on counting statistics, and represent two standard deviations from the mean

[140]. We note that spatial convolution effects such as ion-trajectory overlap and local

magnification effects, which have been cited as sources of misleading results in APT analysis of

nickel-based superalloys [138], are not significant for low-supersaturation Ni-Al-Cr alloys aged

Page 105: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

105

at 873 K [52]. Further experimental details are provided in references [42, 52]; the same

procedures were employed herein.

For comparative purposes, the equilibrium phase boundaries determined by the software

program Thermo-Calc [139], using the thermodynamic databases due to Saunders [115] and

Dupin et al. [116], are superimposed on the phase diagram at 873 K, Figure 5.1 [48]. While the

generated γ / (γ+γ’) solvus lines show good agreement, the curvatures of the (γ+γ’) / γ’ phase

lines differ for each technique. We note a discrepancy between the (γ+γ’) / γ solvus lines

predicted by Thermo-Calc and the GCMC technique as they approach the binary Ni-Al axis.

Given that GCMC employs changes in chemical potentials to determine the equilibrium tie-line

compositions, it is unlikely that this discrepancy is an artifact of the simulation, and may be due

to the fitting involved in the design of Thermo-Calc databases.

All thermodynamic quantities used herein were obtained from the commercial software

package Thermo-Calc [114] using thermodynamic databases due to Saunders [115] and Dupin et

al. [116]. The tracer diffusivities of the atomic species in the γ-matrix phase were calculated

employing Dictra [207] with the mobility database due to Campbell [208] and employing the

Saunders thermodynamic database.

Page 106: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

106

5.3 Results and discussion

5.3.1 Morphological evolution

The temporal evolution of the γ’-precipitate morphology of the model Ni-Al-Cr alloy is

displayed in a series of 25 x 25 x 200 nm3 3D-APT reconstructions, each containing ca. 5.5

million atoms, Figure 5.2. Nanometer-sized spheroidal γ’-precipitates are detected in Ni-6.5 Al-

9.5 Cr at.% for aging times ranging from 0.5 to 1024 h. The γ’-precipitates at an aging time of

4096 h have commenced a spheroidal-to-cuboidal morphological transformation to minimize

their elastic strain energy as they increase in size [83, 209, 210]. For aging times of 1.5 to 64 h,

γ’-precipitate coagulation and coalescence is evident, characterized by the formation of L12-

ordered necks that interconnect γ’-precipitates. From Figure 5.3, the fraction of γ’-precipitates

undergoing coagulation and coalescence, f, is 25±3% after 4 h of aging, and 12±2% after 16 h of

aging; thus coarsening via this mechanism is significant at these times. After aging for 64 h, only

4±1 % of γ’-precipitates are undergoing coagulation and coalescence, and therefore coarsening

proceeds primarily via the evaporation-condensation mechanism. Precipitate coagulation and

coalescence has been detected previously in two different Ni-Al-Cr alloys [42, 52], and is due to

the overlap of the nonequilibrium γ’-precipitate diffusion fields caused by the specific coupling

among diffusion fluxes of Ni, Al, and Cr toward and away from γ’-precipitates [39].

Page 107: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

107

Figure 5.2. APT reconstructed 3D images of a Ni-6.5 Al-9.5 Cr at.% alloy aged at 873 K for (a) 1 h, (b) 4 h, (c) 64 h and (d) 4096 h. The nanometer-sized γ’-precipitates are delineated with red Al isoconcentration surfaces. Aluminum, which partitions to the γ’-precipitates, is shown in red, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity.

Page 108: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

108

Figure 5.3. The fraction of γ’-precipitates undergoing coagulation and coalescence, f, is significant for aging times of 0.5 to 16 h. After aging for 64 h, only 4±1 % of γ’-precipitates are undergoing coagulation and coalescence, and therefore coarsening proceeds primarily via the evaporation-condensation mechanism. No γ’-precipitate coagulation and coalescence is detected beyond an aging time of 64 h.

5.3.2 Temporal evolution of the nanostructural properties of the γ’-precipitates

The temporal evolution of Ni-6.5 Al-9.5 Cr at.% aged at 873 K is complex and is divided

into four regimes: (i) quasi-stationary-state γ’-precipitate nucleation from 0.5 to 1.5 h; followed

by (ii) concomitant precipitate nucleation, growth, and coagulation and coalescence from 1.5 to 4

h; (iii) concurrent growth and coarsening between 4 and 256 h; and finally (iv) quasi-stationary

state coarsening of γ’-precipitates from 256 to 4096 h. Figure 5.4 displays the temporal evolution

of the γ’-precipitate volume fraction, φ, average radius, <R(t)>, and number density, Nv(t), and

clearly shows these four stages of phase decomposition in this model alloy. The γ’-precipitate

nanostructural properties determined by APT analysis are summarized in Table 5.1.

Page 109: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

109

Figure 5.4. The temporal evolution of the γ’-precipitate volume fraction, φ, number density, Nv(t), and mean radius, <R(t)>, for Ni-6.5 Al-9.5 Cr at.% at 873 K. The quantity <R(t)> is approximately proportional to t1/3 during quasi-stationary state coarsening for aging times of 4 h and longer, as predicted by classical coarsening models. Once the equilibrium volume fraction is achieved after 256 h, the temporal dependence of the quantity Nv(t) achieves the t-1 prediction of the coarsening models.

Page 110: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

110

Table 5.1. Temporal evolution of the nanostructural properties of γ’-precipitates a determined by APT for Ni-6.5 Al-9.5 Cr at.% aged at 873 K. The γ’-precipitate mean radius, <R(t)>, number density, Nv(t), and volume fraction, φ, are given, along with their standard errors.

Aging time (h)

Npptb <R(t)> ± 2σ

(nm) Nv(t) ± 2σ

(x 1023 m-3) φ ± 2σ

(%)

0.5 4 0.57 ± 0.10 0.11 ± 0.05 0.002 ± 0.001

0.75 175 0.67 ± 0.21 0.68 ± 0.05 0.02 ± 0

1 328 0.65 ± 0.19 1.27 ± 0.07 0.05 ± 0

1.5 1599 0.90 ± 0.30 5.53 ± 0.14 0.46 ± 0.01

2 755.5 0.99 ± 0.36 6.84 ± 0.25 0.54 ± 0.02

3.5 109.5 1.16 ± 0.53 8.13 ± 0.78 2.14 ± 0.20

4 905 1.23 ± 0.43 9.84 ± 0.33 2.32 ± 0.08

16 314.5 1.64 ± 0.26 2.09 ± 0.12 6.06 ± 0.34

64 47.5 2.82 ± 0.79 1.21 ± 0.08 8.17 ± 1.19

256 40.5 3.58 ± 0.93 0.45 ± 0.07 8.29 ± 1.30

1024 25 7.10 ± 1.86 0.14 ± 0.08 8.53 ± 1.71

4096 13.5 15.80 ± 3.39 0.03 ± 0.01 8.74 ± 2.38

a The number of γ’-precipitates analyzed, Nppt, is smaller than the total number of γ’-precipitates detected by APT. Precipitates that intersect the sample volume contribute 0.5 to the quantity Nppt, and are included in the estimates of Nv(t) and φ, and the phase compositions, and not in the measurement of <R(t)>.

Phase decomposition by nucleation has been studied theoretically for binary alloys in a

set of models known as classical nucleation theory (CNT), which have been reviewed

extensively in the literature [40, 117-120, 122]. According to CNT, nucleation is governed by a

balance between a bulk free energy term, which has both chemical, ΔFch, and elastic strain

Page 111: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

111

energy, ΔFel, components, and an interfacial free energy term, σγ/γ’; where F is the Helmholtz

free energy. Experimental validation of CNT is quite difficult for solid solutions, due to the short

time and length scales involved in precipitation of second phases, which typically contain less

than 100 atoms [40, 211]. Control of the quenching conditions is important to minimize the

vacancy concentration, while simultaneously insuring that phase decomposition does not occur

during the quench. Additionally, the quantities involved in CNT, such as the interfacial energy,

the driving force, and the solute diffusion coefficient, are difficult to measure, and the

predictions of the models are highly sensitive to uncertainties in these values. As a result, the

discrepancies between the measured and predicted nucleation rates are typically between 3 and 5

orders of magnitude [40, 212].

From CNT, the expression for the net reversible work required for the formation of a

spherical nucleus, WR, as a function of nucleus radius, R, is given by:

'/23 43

4)( γγσππ RRFFW elchR +Δ+Δ= . (5.1)

The critical net reversible work, *RW , required for the formation of a critical spherical

nucleus is given by:

2elch

3*R )FF(3

16W'/

ΔΔσπ γγ

+= ; (5.2)

The critical net reversible work for nucleation acts as a barrier that nuclei must overcome

to achieve a critical nucleus radius, R*, given by:

)(

2*'/

elch FFR

Δ+Δ−=

γγσ . (5.3)

From CNT, the stationary-state nucleation current, Jst, the number of nuclei formed per

Page 112: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

112

unit volume per unit time (m-3 s-1), is given by:

)Tk

Wexp(N*ZJB

*Rst −

= β ; (5.4)

where Z, the Zeldovich factor, accounts for the dissolution of supercritical clusters, β∗ is a

kinetic coefficient describing the rate of condensation of single atoms on the critical nuclei, N is

the total number of possible nucleation sites per unit volume, taken to be the volume density of

lattice sites occupied by Al, the precipitating solute element [122, 213], kB is Boltzmann’s

constant, and T is the absolute temperature. The standard definitions of Z and β∗ were employed

and are given by [40, 117-120, 122]:

2/1

b2

i2

Tk21

i*)i(WZ

⎥⎥⎦

⎢⎢⎣

⎡⎟⎟⎠

⎞⎜⎜⎝

⎛⎟⎟⎠

⎞⎜⎜⎝

⎛∂

∂−=

π, (5.5)

40

aDC*R4* π

β = ; (5.6)

where Wi is the net reversible work required for the formation of a spherical nucleus containing i

atoms, D is the diffusion coefficient of the precipitating solute element, C0 is the initial

concentration of said solute, and a is the average of the lattice parameters of the matrix and

precipitate phase, equal, in this case, to 0.3568 nm. Given that the extant theories of nucleation

are for binary alloys, the calculations of the stationary-state nucleation rate were performed for

both Al and Cr, thereby providing an upper and lower bound for the predicted nucleation rate.

Diffusion coefficients of 9.47 and 2.61 x 10-21 m2 s-1 were employed for Al and Cr, respectively,

along with C0 values of 6.5 and 9.5 at.%.

The value of ΔFch of -6.25 x 107 J m-3 is determined from Thermo-Calc [139] and the

Saunders thermodynamic database [145], and is calculated employing the equilibrium phase

Page 113: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

113

compositions, an assumption which is valid for this alloy, given the small values of the solute

supersaturations [40, 214-216]. The values of ΔFel for the two alloys are estimated using [154]:

'a

2a

'a

el V)S4B3()VV(BS2F γγγ

γγγγ

Δ+

−= ′

; (5.7)

where Sγ is the shear modulus of the γ-matrix phase, Bγ’ is the bulk modulus of the γ’-precipitate

phase, and γaV and 'γ

aV are the atomic volumes of the γ-matrix and γ’-precipitate phases,

respectively. No elastic constants are available for this alloy, therefore the value of Sγ of 100.9

GPa, of a similar alloy, Ni-12.69 Al at 873 K [155] is employed, while the value of Bγ’ is taken to

be 175 GPa [156]. The lattice parameters of the γ'- and γ-phases at 873 K are calculated to be

0.3573 nm and 0.3563 nm, respectively, employing the precipitation simulation software

program PrecipiCalc [217, 218]. These lattice parameter values result in a near-zero value of the

lattice parameter misfit of 0.00284 for Ni-6.5 Al-9.5 Cr at.%. Substituting these values into

Equation 5.4 yields a value of ΔFel of 2.67 x 106 J m-3. The high degree of coherency of the γ’-

precipitates in this alloy is such that the bulk component of the driving force for nucleation is

dominated by the ΔFch term, as ΔFel is only 4.27% of the value of ΔFch. A value of σγ/γ’ of 22±1

mJ m-2 is estimated by a technique that relies on thermodynamic information and experimental

coarsening data, and was first used by Ardell for a binary alloy [145, 146], and extended to the

ternary Al-Sc-Mg and Ni-Al-Cr alloys [42, 52, 147], and is described in detail therein. The value

of σγ/γ’ of 22±1 mJ m-2 estimated for Ni-6.5 Al-9.5 Cr at.% is in excellent agreement with the

values of 20-25 mJ m-2 determined by the same technique for Ni-Al-Cr alloys aged at 873 K

[52]. We note that the value of σγ/γ’ during nucleation may differ significantly from this value,

since the latter was obtained from the coarsening regime, after the γ’-precipitates have undergone

Page 114: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

114

a significant amount of temporal evolution. Unfortunately, there is no extant technique to

estimate the value of σγ/γ’ during nucleation for this ternary alloy aged at 873 K.

Employing the values of ΔFch, ΔFel, and σγ/γ’ given above, CNT predicts nucleation

currents of 1.06 x 1023 and 6.24 x 1022 m-3 s-1, employing the diffusion data for Al and Cr,

respectively. These values are 700 and 285 times greater than the measured rate of 1.5±0.7 x 1020

m-3 s-1, respectively. Given the evidence of precursor clustering in these Ni-Al-Cr alloys [45], it

is surprising that the experimentally determined value of Jst is significantly less than the

predicted value. Predictions of the nucleation rate by CNT have been found to be 50-500 times

greater than the experimental values for other Ni-Al and Ni-Al-Cr alloys [16, 52]. Xiao and

Haasen attributed this discrepancy to the sensitivity of the predicted value of Jst to the value of R*

[16]. They also noted that predicted nucleation currents are likely overestimated due to the

assumption that the value of N0 is equal to the volume density of lattice sites, which is a

commonly made assumption that is not necessarily correct. By assuming that the value of N0 is

equal to the volume density of Al atoms, we have mitigated this error to some extent. The

prediction of R* from CNT, which depends only on thermodynamic quantities, yields a value of

0.74 nm, in agreement with the value of <R(t)> of 0.62±0.17 we measured for aging times of

0.5-1 h.

To date, there is not a generally accepted theory of nucleation in concentrated

multicomponent alloys. Given that the detailed atomistic kinetics involved in the formation of γ’-

nuclei in these ternary systems are not completely understood, further analysis herein is not

instructive. The earliest stages of precipitate formation in Ni-6.5 Al-9.5 Cr at.% are described in

greater detail elsewhere [219], and complemented with results of lattice kinetic Monte Carlo

Page 115: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

115

simulation to develop a deeper understanding of the kinetic pathways of nucleation in this alloy.

Nucleation and growth of γ’-precipitates occurs concurrently for aging times of 1.5 to 4

h, resulting in a nucleation current of 4.1±1.6 x 1019 m-3 s-1, and an increase in the value of φ

from 0.46±0.01 % to 2.32 ±0.08 %. As such, we do not observe a pure growth regime [141-143],

which is not a surprise given the evidence of concomitant γ’-precipitate nucleation, growth and

coagulation and coalescence between 1.5 and 4 h.

For aging times between 4 and 256 h, the values of φ and <R(t)> increase, and the values

of Nv(t) decrease, as expected for quasi-stationary-state coarsening. Previous investigations of

the phase decomposition of Ni-Al-Cr alloys have shown that the predictions of the temporal

evolution of the quantity <R(t)> from classical coarsening models are validated experimentally

during quasi-stationary-state coarsening [28, 37, 38, 42, 45, 46, 52, 112]. The predicted temporal

evolutions of the quantity Nv(t) and of the particle size distributions have not been validated

experimentally. This is because experimental alloys have failed to achieve genuine stationary-

state coarsening at the equilibrium volume fraction of the precipitated phase [9, 10, 127, 144].

The first comprehensive mean-field treatment of coarsening, due to Lifshitz and Slyozov [125]

and Wagner [126], known as the LSW model, is limited to dilute binary alloys with spatially-

fixed spherical precipitates whose initial compositions are equal to their equilibrium values, as

determined by the equilibrium binary phase diagram. These stringent requirements, and other

assumptions inherent to the LSW model, are highly restrictive, and difficult to meet in practice.

Researchers have worked to remove the mean-field restrictions by developing models based on

multiparticle diffusion that are able to describe stress-free systems with finite volume fractions

[10, 128, 129, 131]. Umantsev and Olson (UO) [130] demonstrated that the exponents of the

Page 116: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

116

temporal power-laws for concentrated multi-component alloys are identical with those for dilute

binary alloys, but that the explicit expressions for the rate constants depend on the number of

components. Kuehmann and Voorhees (KV) [132] considered isothermal coarsening in ternary

alloys and developed a detailed model that includes the effects of capillarity on the precipitate

composition, such that the phase compositions can deviate from their equilibrium values. In the

quasi-stationary-state limit of the KV model, the expressions for the temporal dependencies for

<R(t)>, Nv(t), and the γ-matrix supersaturation, )(tCiγΔ , of each solute species i, are:

)()()( 0KV3

03 ttKtRtR −=><−>< ; (5.8)

)()()( 0eqKV1

0v1

v ttK74.4

tNtN −≅− −−

φ; and (5.9)

3/1KV,i

eq,i

ff,ii tCtCtC −=∞−>=< γγγγ κΔ )()()( ; (5.10)

where KKV, and γκKV,i

are the coarsening rate constants for <R(t)> and )(tCiγΔ , respectively;

<R(t0)> is the average precipitate radius and Nv(t0) is the precipitate number density at the onset

of quasi-stationary-state coarsening, at time t0. The quantity )(tCiγΔ is denoted a supersaturation

and is the difference between the concentration in the far-field γ-matrix, >< )(, tC ffiγ , and the

equilibrium γ-matrix solute-solubility, )(, ∞eqiC γ . The temporal dependence of the quantity

<R(t)> for the alloy studied herein is measured to be t0.36±0.06 from the onset of coarsening at 4 h

to 4096 h. As noted previously, the prediction of the temporal evolution of the quantity <R(t)>

has been validated for systems where the equilibrium volume fraction has not been achieved, as

is the case for Ni-6.5 Al-9.5 Cr at.% for aging times of 4-256 h. Thus, its measurement alone is

not a particularly strong test of the LSW model. For this alloy, the temporal dependence of the

Page 117: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

117

quantity Nv(t) is t -0.80±0.06 for aging times of 4 h to 4096 h, and does not achieve the value of t-1

predicted by classical models of coarsening because the equilibrium volume fraction of

precipitates was not achieved for all data points considered.

During aging from 4 to 16 h, the value of Nv(t) drops sharply from the maximum value of

9.84±0.33 x 1023 m-3 at 4 h, to 2.09±0.12 x 1023 m-3 at 16 h as a result of coarsening via both the

classical evaporation-condensation mechanism and the coagulation and coalescence of γ’-

precipitates. The fraction of coagulating and coalescing γ’-precipitates is 25±3% after 4 h of

aging, and 12±2% after 16 h of aging; thus coarsening via this mechanism is significant at these

times. Note that the value of φ increases from 2.31±0.08 to 6.06±0.34 % over this range of time,

thus γ’-precipitate growth is also ongoing. After aging for 64 h, only 4±1 % of γ’-precipitates are

undergoing coagulation and coalescence, and therefore coarsening proceeds primarily via the

evaporation-condensation mechanism. For aging times of 256 to 4096 hours, where no

coagulation and coalescence of γ’-precipitates is detected, and the volume fraction of γ’-

precipitates is approximately constant, the temporal dependence of the quantity Nv(t) is t -1.04±0.10,

as predicted by extant models of coarsening, including the detailed KV model for ternary alloys

[125, 126, 132, 220, 221]. We note that the temporal dependence of the quantity <R(t)> is

measured to be t0.54±0.15 for aging times of 256 to 4096 hours, when volume fraction of γ’-

precipitates is approximately constant.

Page 118: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

118

5.3.3 Temporal evolution of the compositions of the γ and γ’ phase compositions

The compositions of the γ-matrix and γ’-precipitate phases in Ni-6.5 Al-9.5 Cr at.%

evolve temporally, as the γ-matrix is enriched in Ni and Cr and depleted of Al, as shown in

Figure 5.5. The first γ’-nuclei, detected at an aging time of 0.5 h, have a solute supersaturated

composition of 70.1 ± 3.0 Ni, 21.3 ± 4.9 Al and 8.7 ± 5.3 Cr at.%, with a value of <R(t=0.5 h)>

of 0.57 ± 0.10 nm. As phase decomposition proceeds, the values of )(tCiγΔ and the γ’-precipitate

supersaturation, )t(C 'iγΔ , decrease asymptotically as the γ- and γ’-phase compositions approach

their equilibrium values. The equilibrium compositions of the γ-matrix and γ’-precipitate phases

are determined from Equation 5.10 to be 84.19±0.02 Ni-5.49±0.05 Al-10.30±0.05 Cr and

76.40±0.19 Ni-17.53±0.33 Al-6.06±0.36 Cr at.%, respectively. These values are used to estimate

an equilibrium volume fraction of γ’-precipitates, φeq, of 9.6±2.9 %, according to the lever rule.

The values of φ of 8.29±1.30, 8.53±1.71 and 8.74±2.38 % determined by APT for aging times of

256, 1024 and 4096 h, respectively, are all within the experimental error of the φeq value of

9.6±2.9 %. Thus, by an aging time of 256 h, growth of γ’-precipitates is complete, and phase

decomposition proceeds only by quasi-stationary-state coarsening.

Page 119: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

119

Figure 5.5. The composition profiles across the γ-matrix/ γ’-precipitate interface for Ni-6.5 Al-9.5 Cr at.% at 873 K for aging times of 1, 4, 4096 h. The phase compositions evolve temporally, as the γ-matrix becomes enriched in Ni and Cr and depleted in Al. The values of <R(t)> for these aging times are 0.65 ± 0.19 nm for 1 h, 1.23 ± 0.43 nm for 4 h and 15.80 ± 3.39 nm for 4096 h.

The temporal evolution of the phase compositions is shown on a partial ternary Ni-Al-Cr

phase diagram determined by the GCMC technique at 873 K, Figure 5.6 [130]. For comparative

purposes, the compositional trajectories of Ni-7.5 Al-8.5 Cr and Ni-5.2 Al-14.2 Cr at.%, two

alloys that have also been investigated by APT at Northwestern University, are superimposed on

the phase diagram. According to the KV coarsening model, the addition of a third alloying

element alters both the form of the Gibbs-Thompson equations and the predictions of the

Page 120: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

120

temporal evolution of the phase compositions [132]. The slope of the trajectory of the γ-matrix

phase during coarsening is predicted by the KV model to lie along the equilibrium tie-line and

have a value of pAl/pCr, where pi is the magnitude of the partitioning of solute species i, as

defined by the expression pi = ,[ ( )eqiCγ ′ ∞ - , ( )]eq

iCγ ∞ . Alternatively, the compositional trajectory of

the γ’-precipitate phase is predicted by the KV model to lie on a straight line that is not

necessarily parallel to the equilibrium tie-line. To verify these predictions for the ternary alloy

studied, the quantities )(/)( tCtC CrAlγγ δδ and )(/)( '' tCtC CrAl

γγ δδ for aging times of 4-4094 h are

compared to the value of pAl/pcr, where δCpi is the slope of the concentration of element i in

phase p with aging time. The trajectories of the γ-matrix phase and the γ’-precipitate phases have

slopes of -4.29±0.52 and 0.56±0.24, respectively, while the slope of the equilibrium tie-line is

estimated to be -2.84±0.06. From these results, it is absolutely clear that the trajectory of the γ’-

precipitate composition does not lie along the equilibrium tie line. This result is contrary to all

coarsening models, except the KV model for ternary alloys. We note that the KV mean-field

model is unable to predict the correct value of the slopes, which is most likely due to the

omission of the off-diagonal terms in the diffusion matrix in this model. The omission of the off-

diagonal terms, in turn, generates inaccurate predictions of the diffusion fluxes into and out of

the γ'-precipitate phase by suppressing flux coupling [39].

Page 121: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

121

Figure 5.6. The compositional trajectories of the γ-matrix and γ’-precipitate phases of Ni-6.5 Al-9.5 Cr at.%, displayed on a partial Ni-Al-Cr ternary phase diagram at 873 K. The trajectory of the γ-matrix phase lies approximately on the experimental tie-line, while the trajectory of the γ'-precipitate phases does not lie along the tie-line, as predicted by the Kuehmann-Voorhees coarsening model [132]. The tie-lines are determined from the equilibrium phase compositions, determined by extrapolation of APT concentration data to infinite time.

The values plotted on the phase diagram in Figure 5.6 are the far-field plateau

compositions of the two phases, and do not include atoms that lie within the interfacial γ/γ’

region. As such, the rule of mixtures, which dictates that the nominal γ’-precipitate and γ-matrix

compositions at a given aging time lie along a straight line, is not satisfied. Classical models of

phase decomposition often ignore the interfacial region, thus we have chosen a similar approach

for comparative purposes. The inclusion of the interfacial region in the phase compositions (not

displayed) does not alter the compositional trajectories significantly, and does satisfy the rule of

mixtures. Advanced models of phase decomposition are needed to account for the interfacial

region between phases. The interfacial region, which has a definite width on the order of ~ 2 nm

Page 122: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

122

for this alloy, Figure 5.5, has been measured to be on the order of several nanometers for other

model nickel-based superalloys [42, 52, 56, 58, 62, 113, 222-224].

The partitioning behavior of the elements can be determined quantitatively by calculating

the partitioning ratio, Kiγ’/γ

, defined as the ratio of the concentration of an element i in the γ’-

precipitates to the concentration of the same element in the γ-matrix. Figure 5.7 demonstrates

partitioning of Al to the γ’-precipitates, and of Ni and Cr to the γ-matrix. The values of Kiγ’/γ are

constant from 256-4096 h, when the alloy is undergoing quasi-stationary state coarsening.

Figure 5.7. The partitioning ratios, Kiγ’/γ

, quantify the partitioning of Al to the γ’-precipitates and of Ni and Cr to the γ-matrix. The values of Ki

γ’/γ are constant from 256 to 4096 h, when the alloy is undergoing quasi-stationary state coarsening.

Page 123: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

123

The diminution of the γ-matrix supersaturation values approximately follow the t-1/3

prediction of the KV model. From Figure 5.8, the alloy displays a temporal dependence of t-

0.32±0.03 for )t(CNiγΔ , t-0.32±0.04 for )(tCAl

γΔ , and t-0.32±0.04 for )(tCCrγΔ . The temporal dependencies

of the γ’-precipitate supersaturation values also follow approximately the t-1/3 prediction, with

measured dependencies of t-0.32±0.03 for )t(C 'NiγΔ , t-0.29±0.04 for )t(C '

AlγΔ , and t-0.34±0.03 for )t(C '

CrγΔ .

Figure 5.8. The magnitude of the values of the γ-matrix (left) and γ’-precipitate (right) supersaturations, )(tCi

γΔ and )t(C 'iγΔ , of Ni, Al and Cr, decrease as t-1/3 in the coarsening

regime, as predicted by classical coarsening models.

Page 124: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

124

5.4 Conclusions

We present a detailed comparison of the nanostructural and compositional evolution of

Ni-6.5 Al-9.5 Cr at.% during phase separation at 873 K for aging times ranging from 0.5 to 4096

h, employing atom-probe topography (APT). This ternary alloy with a small equilibrium γ’-

precipitate volume fraction of 9.6±2.9 % was designed to study precipitate nucleation, and

subsequently growth and coarsening, leading to the following results:

• The phase decomposition of the model alloy occurs in four distinct regimes: (i) quasi-

stationary-state γ’-precipitate nucleation from 0.5 to 1.5 h; (ii) followed by concomitant

precipitate nucleation, growth and coagulation and coalescence from 1.5 to 4 h; (iii)

concurrent growth and coarsening from 4 to 256 h, wherein coarsening occurs via both

γ’-precipitate coagulation and coalescence and by the classical evaporation-condensation

mechanism; and (iv) finally quasi-stationary-state coarsening of γ’-precipitates from 256

to 4096 h.

• The morphology of the γ’-precipitate phase is found to be spheroidal for aging times as

long as 1024 h, as a result of a near-zero lattice parameter misfit between the γ-matrix

and γ’-precipitate phases. Coagulation and coalescence of γ’-precipitates is observed for

aging times of 1.5 to 64 h, as a result of the overlap of the non-equilibrium concentration

profiles associated with adjacent γ’-precipitates [39]. The γ’-precipitates that have formed

by an aging time of 4096 h have commenced a spheroidal-to-cuboidal morphological

transformation to minimize their elastic strain energy as they increase in size [83, 209,

210].

• It is found that classical nucleation theory accurately predicts the critical radius for

Page 125: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

125

nucleation for the model Ni-Al-Cr alloy, although the prediction of the stationary-state

nucleation current is at least 285 times greater than the value measured experimentally.

These findings indicate that a better understanding of the kinetic pathways of nucleation

in this concentrated ternary alloy is required.

• The predictions of the Umantsev-Olson [130] and Kuehmann-Voorhees [132] models of

coarsening are verified experimentally for this model Ni-Al-Cr alloy. Beyond an aging

time of 4 h, the temporal evolutions of the values of the average precipitate radii and γ-

matrix and γ’-precipitate supersaturations follow the predictions of t1/3 and t-1/3,

respectively. For aging times beyond 256 h, the volume fraction of γ’-precipitates is

constant and the temporal dependence of the quantity Nv(t) is t-1.0±0.1.

• The solute solubility in the γ-matrix phase of Ni-6.5 Al-9.5 Cr at.% is determined by APT

to be 5.49±0.05 Al-10.30±0.05 Cr, while the equilibrium γ’-precipitate composition is

76.40±0.19 Ni-17.53±0.33 Al-6.06 ±0.36 Cr at.%. These values are used to estimate an

equilibrium volume fraction of γ’-precipitates, φeq, of 9.6±2.9 %, according to the lever

rule. The values of φ of 8.29±1.30, 8.53±1.71 and 8.74±2.38 % determined by APT for

aging times of 256, 1024 and 4096 h, respectively, are all within the experimental error of

the φeq value of 9.6±2.9 %. Thus, by an aging time of 256 h, growth of γ’-precipitates is

complete, and phase decomposition proceeds only by quasi-stationary-state coarsening.

• The compositional trajectory of the γ-matrix during phase decomposition lies

approximately along the tie-line, while the trajectory of the γ’-precipitate phase does not,

as predicted by the KV mean-field model for quasi-stationary state coarsening in ternary

alloys. The addition of a third alloying element alters the Gibbs-Thompson equations

Page 126: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

126

significantly, and as such, the KV model predicts that the compositional trajectory of the

γ’-precipitate phase will lie on a straight line that is not necessarily parallel to the

equilibrium tie-line. The KV model, however, is unable to predict the exact value of the

slopes, which is most likely due to the omission of the off-diagonal terms in the diffusion

matrix in this model, which generates inaccurate predictions of the diffusion fluxes into

and out of the γ'-precipitate phase by suppressing flux coupling [39].

• An estimate of the γ/γ’ interfacial free energy, σγ/γ’, from the coarsening data obtained by

APT yields a value of 22±1 mJ m-2. This value for Ni-6.5 Al-9.5 Cr at.% is in excellent

agreement with the values of 20-25 mJ m-2 measured by the same technique for other Ni-

Al-Cr alloys aged at 873 K [52].

Page 127: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

127

Chapter 6

Effects of tantalum on the temporal evolution of a model Ni-Al-Cr superalloy

during phase decomposition

Abstract

The effects of a 2.0 at.% addition of Ta to a model Ni-10.0 Al-8.5 Cr at.% superalloy aged at

1073 K are assessed using scanning electron microscopy and atom-probe tomography. The

γ’(L12)-precipitate morphology that develops as a result of γ(FCC)-matrix phase decomposition

is found to evolve from a bimodal distribution of spheroidal precipitates, to {001}-faceted

cuboids and parallelepipeds aligned along the elastically soft <001>-type directions. The phase

compositions and the widths of the γ’-precipitate/γ-matrix heterophase interfaces evolve

temporally as the Ni-Al-Cr-Ta alloy undergoes quasi-stationary state coarsening after 1 h of

aging. Tantalum is observed to partition preferentially to the γ’-precipitate phase, and suppresses

the mobility of Ni in the γ-matrix sufficiently to cause an accumulation of Ni on the γ-matrix side

of the γ'/γ interface. Additionally, computational modeling, employing Thermo-Calc, Dictra and

PrecipiCalc, is employed to elucidate the kinetic pathways that lead to phase decomposition in

this concentrated Ni-Al-Cr-Ta alloy.

Page 128: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

128

6.1 Introduction

Tantalum is an important refractory addition to commercial nickel-based superalloys as

both a solid-solution strengthener and a precipitate former. Tantalum has been shown to increase

the high-temperature strength and ductility, and to improve the resistance to creep, fatigue, and

corrosion of these high-performance materials used in land-based and aerospace turbine engines

at operating temperatures up to 1373 K [3, 7, 78, 79, 225, 226]. The effects of Ta on the

microstructure and mechanical properties of nickel-based superalloys have been investigated

[226-231], however, little has been done to characterize the morphological and compositional

changes due to the addition of Ta.

Phase decomposition in model Ni-Al-Cr alloys has been investigated in detail by atom-

probe tomography (APT) [138]. Τhe research of Schmuck et al. [28, 112] and Pareige et al. [37,

38] combined APT and lattice kinetic Monte Carlo (LKMC) simulation to analyze the

decomposition of a Ni-Al-Cr solid-solution at 873 K. A similar approach was applied by

Sudbrack et al. [42, 45, 46, 55, 56], Yoon et al. [58, 59], Mao et al. [39] and Booth-Morrison et

al. [52] for studying Ni-5.2 Al-14.2 Cr and Ni-7.5 Al-8.5 Cr at.% aged at 873 K, and Ni-10 Al-

8.5 Cr at.%, Ni-10 Al-8.5 Cr-2.0 W at.% and Ni-10 Al-8.5 Cr-2.0 Re at.% aged at 1073 K, which

decompose via a first-order phase transformation to form a high number density (~1020-1025 m-3)

of nanometer-sized γ’-precipitates. We report on the temporal evolution of a model Ni-10.0 Al-

8.5 Cr-2.0 Ta at.% alloy aged at 1073 K that decomposes to form a microstructure consisting of

γ’(L12)-precipitates in a γ(FCC)-matrix. Chromium is added to the binary Ni-Al system to reduce

the lattice parameter misfit between the γ’-precipitates and the γ-matrix. The addition of Ta

increases the volume fraction of the γ’-precipitate phase, providing significant strengthening.

Page 129: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

129

Tantalum also decreases overall alloy diffusivity, thereby improving phase stability and service

life by retarding diffusion mediated processes such as γ’-precipitate coarsening, creep and

oxidation [78, 79]. The effect of a 2.0 at.% addition of Ta to a model Ni-Al-Cr superalloy is

studied using scanning electron microscopy (SEM) and APT. The experimental results are

complemented by computational modeling employing the commercial software packages

Thermo-Calc, Dictra and PrecipiCalc, to elucidate the thermodynamic and kinetic pathways that

lead to phase decomposition in a concentrated Ni-Al-Cr-Ta alloy.

6.2 Experimental

High-purity constituent elements were induction-melted under a partial pressure of Ar

and chill cast in a 19 mm diameter copper mold to form a polycrystalline master ingot. Samples

from the cast ingot then underwent a three-stage heat-treatment: (1) homogenization at 1573 K in

the γ-phase field for 20 h; (2) a vacancy anneal in the γ−phase field at 1503 K for 3 hours

followed by a water quench; and (3) an aging anneal at 1073 K under flowing argon for times

ranging from 0.25 to 256 h, followed by a water quench. Microtip specimens and metallographic

samples were prepared from each of the aged sections for study by APT and SEM.

Vickers microhardness was measured using a Buehler Micromet™ instrument on

samples polished to 1 μm, with an applied load of 500 g sustained for 5 s, using the average

value of fifteen independent measurements made on several grains. SEM was performed on

samples polished to an 0.02 μm finish and etched in a 100 ml HCl/100 ml deionized H2O/1g

K2S205 mixture, employing a LEO Gemini 1525™ field-emission SEM operating at 5 kV with a

20-30 μm aperture and a working distance of 6 mm. APT microtips were prepared using

Page 130: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

130

standard procedures [176, 232], and analyzed with a local-electrode atom-probe (LEAPTM)

tomograph [99, 101-103, 165] at the Northwestern University Center for Atom-Probe

Tomography (NUCAPT). Pulsed-laser APT data collection was performed at a target

evaporation rate of 0.04 ions per pulse, a specimen temperature of 40.0±0.3 K, a pulse energy of

0.6 nJ, a pulse repetition rate of 200 kHz, and a background gauge pressure of less than 6.7 x 10-8

Pa. Pulsed-laser atom-probe was employed to improve the compositional accuracy of the APT

technique by limiting preferential evaporation [54], and to increase the tip specimen life by

reducing the dc voltage required for evaporation. APT data were analyzed with the IVAS® 3.0

software program (Imago Scientific Instruments). The γ’-precipitate/γ-matrix heterophase

interfaces were delineated with Al isoconcentration surfaces generated by efficient sampling

procedures [136], and detailed compositional information was obtained with the proximity

histogram method [137]. The equilibrium volume fraction of the γ’-precipitate phase, φeq, was

estimated by averaging the values obtained from the lever rule of the extrapolated equilibrium

concentrations of Ni, Al, Cr and Ta.

Spatial convolution effects such as ion trajectory overlap and local magnification effects

have been cited as possible sources of misleading results in the APT analysis of nickel-based

superalloys [138]. A comparison of the composition profiles across the γ/γ’ interfaces measured

by APT to those simulated by the lattice kinetic Monte Carlo technique for a Ni-5.2 Al-14.2 Cr

at.% alloy [39], showed no evidence of artificial interfacial broadening in the APT data due to

ion trajectory overlap. Local magnification effects due to differences in the required evaporation

fields of different phases are unlikely in Ni-Al-Cr superalloys containing only the γ-matrix and

γ'-precipitate phases because the evaporation fields of the phases are essentially identical.

Page 131: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

131

The overall composition of the master ingot was determined by inductively coupled

plasma (ICP) atomic-emission spectroscopy to be 80.01 Ni-9.75 Al-8.21 Cr-2.02 Ta at.%, and

was indistinguishable, within experimental error, from the targeted composition of Ni-10.0 Al-

8.5 Cr-2.0 Ta at.%. ICP chemical analysis was also used to determine the compositions of the γ-

matrix and γ’-precipitate phases of a sample aged at 1073 K for 256 h after phase extraction by

anodic dissolution of the γ-matrix phase with a 1:1 aqueous solution of citric acid and

ammonium nitrate at constant current density. We note that the standard errors for all quantities

are calculated based on counting statistics and reconstruction scaling errors using standard error

propagation methods [140], and represent two standard deviations from the mean.

The commercial software package Thermo-Calc [114] was used to estimate the values of

φeq and the equilibrium compositions of the γ-matrix, )(, ∞eqiC γ , and γ’-precipitate phases,

)(,' ∞eqiC γ , at a pressure of 1 atmosphere, using a database for nickel-based superalloys developed

by Saunders [115]. The tracer diffusivities of the atomic species in the γ-matrix phase were

calculated employing Dictra [207] with the mobility database due to Campbell [208] and

employing the Saunders thermodynamic database. Additionally, precipitation modeling was

performed with the commercial software PrecipiCalc [217, 233], employing the Saunders

thermodynamic database and the mobility database due to Campbell. Precipicalc applies

thermodynamic and kinetic data from Thermo-Calc and Dictra to continuum models of

precipitation for multicomponent, multiphase alloys to provide a unified treatment of nucleation,

growth and coarsening.

Page 132: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

132

6.3 Results

A 2.0 at.% addition of Ta to a model Ni-10.0 Al-8.5 Cr at.% alloy aged at 1073 K results

in a 47±5% increase in the microhardness, Figure 6.1, over the full range of aging times, t = 0 to

256 h. The microhardness of Ni-Al-Cr-Ta varies from 3.3 to 3.6 GPa, with peak microhardness

occurring between t = 1 and 16 h. For Ni-10.0 Al-8.5 Cr at.% aged at 1073 K, the microhardness

varies over a narrow range, 2.15 GPa to 2.5 GPa, with peak microhardness occurring at 4 h [55,

56]. The increase in the microhardness due to the addition of Ta is a result of both the added

solid-solution strengthening provided by Ta, and to a dramatic increase in the value of the γ’-

precipitate volume fraction, φ. At 256 h, for example, the value of φ from the phase extraction

results is 18.9% for Ni-Al-Cr and 36.4% for Ni-Al-Cr-Ta. The values of the equilibrium volume

fractions, φeq, estimated by APT and from Thermo-Calc for Ni-Al-Cr-Ta are presented in Table

6.1, and compared to the values of φ at 256 h, as measured by both APT and phase extraction.

Figure 6.1. Vickers microhardness measurements for Ni-10.0 Al-8.5 Cr-2.0 Ta and Ni-10.0 Al-8.5 Cr at.% aged at 1073 K. The addition of 2.0 at.% Ta results in a 47±5% increase in the microhardness over the full range of aging times, t = 0 to 256 h, due to increases in solid-solution strengthening and to a dramatic increase in the value of φ.

Page 133: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

133

Table 6.1. Equilibrium γ’-precipitate volume fraction, φeq, determined by APT, ICP chemical analysis, and thermodynamic modeling employing Thermo-Calc for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% aged at 1073 K.

Technique used to estimate φeq: φeq (%)

Lever rule with APT compositions, 256 h: 37.2 ± 8.9

Lever rule with APT equilibrium compositions: 37.0 ± 9.0

ICP chemical analysis, 256 h: 36.4

Thermo-Calc using Saunders database [115]: 38.17

6.3.1 Morphological development

The microstructure of the material as-quenched from 1503 K consists of a bimodal

distribution of γ’-precipitates, with primary γ’-precipitates of radii, R(t = 0 h), on the order of 30-

40 nm, and smaller secondary γ’-precipitates with an average radius, <R(t = 0 h)>, of 6.4±1.1

nm. From APT, the number density, Nv(t = 0 h), of the primary γ’-precipitates is 3±2 x 1021 m-3,

occupying a volume fraction of 28±13%. The value of Nv(t = 0 h) of the secondary γ’-

precipitates is 1.3±0.1 x 1023 m-3, with a volume fraction of 11±1%. The average edge-to-edge

distance, λe-e, between secondary γ’-precipitates is measured to be 4.3±1.3 nm, while the value of

λe-e for the primary γ’-precipitates is 109±63 nm, as measured by a method developed by

Karnesky et al. [106]. A solute-depleted, precipitate-free shell with a thickness of 5.7±1.1 nm

surrounds the primary γ’-precipitates. The bimodal distribution is only observed in the as-

quenched state. With aging, the secondary γ’-precipitates are likely consumed by an Ostwald

ripening process to decrease the total γ/γ’ interfacial free energy of the system.

Page 134: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

134

Figure 6.2. APT reconstructed 3D image of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 0 h. The solute elements that partition to the γ’-precipitates, Ta and Al, are shown in orange and red, respectively, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity. The morphology of the γ’-precipitate phase is spheroidal in the as-quenched state, and a bimodal particle size distribution is apparent.

Page 135: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

135

The morphology of the Ni-Al-Cr-Ta alloy evolves from spheroidal to cuboidal γ’-

precipitates with significantly rounded edges, and finally to cuboidal and parallelepipedic γ’-

precipitates aligned along the elastically soft <001>-type directions, Figure 6.3. For aging times

beyond the as-quenched state, the primary γ’-precipitates develop an {001}-faceted cuboidal

morphology with rounded corners, as seen by both SEM, Figure 6.3b, and APT, Figure 6.4, for a

sample aged for 0.25 h. We note that the γ’-precipitates in Figure 6.3b and 4 exhibit both

spheroidal and cuboidal characteristics. The value of R(t = 0.25 h) at which primary γ’-

precipitates in this Ni-Al-Cr-Ta alloy undergo the transformation from a spheroidal-to-cuboidal

morphology is estimated from APT to be 61±7 nm.

Figure 6.3. SEM images of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for: (a) 0 h: (b) 0.25 h: (c) 1 h: (d) 4 h: (e) 16 h: and (f) 64 h. The primary γ’-precipitate morphology evolves from spheroidal γ’-precipitates to a cuboidal and parallelepipedic morphology with primary γ’-precipitates aligned along the elastically soft <001>-type directions.

Page 136: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

136

Figure 6.4. APT reconstructed 3D image of a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 0.25 h. The primary γ’-precipitates in the APT reconstruction have both spheroidal and cuboidal characteristics. The radius at which the γ’-precipitates undergo the spheroidal-to- cuboidal morphological transformation is measured to be 61±7 nm. Tantalum and Al are shown in orange and red, respectively, Cr is shown in blue, and Ni atoms are omitted for clarity.

The cuboidal morphology of the γ’-precipitates persists with aging, and for aging times

beyond of 4 h, the γ’-precipitates align along orthogonal <001>-type directions, Figures 6.2 and

6.5. For aging times of 64 and 256 h, the primary γ’-precipitates have a non-equiaxed,

rectangular parallelepipedic morphology with {001}-type facets. We note that the direction of γ’-

precipitate alignment was verified employing selected-area electron-diffraction (SAED) patterns

using conventional transmission electron microscopy, with an Hitachi HF-8100 at an

accelerating voltage of 200 kV, for a sample aged for 256 h at 1073 K, Figure 6.6.

Page 137: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

137

Figure 6.5. An APT reconstructed 3D image of cuboidal and parallelepipedic primary γ’-precipitates aligned along <001>-type directions from a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 64 h. Tantalum and Al are shown in orange and red, respectively, Cr is shown in blue, and Ni atoms are omitted for clarity.

Figure 6.6. After aging for 256 h, the primary γ’-precipitates have cuboidal and parallelepipedic morphologies with {001}-type facets, and are aligned along the <001>-type crystallographic directions, as confirmed by a selected-area electron diffraction pattern taken along the ]101[ zone axis.

Page 138: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

138

6.3.2 Compositional evolution

The compositional information generated by APT permits the study of the temporal

evolution of the phase compositions, and of the concentration profiles at the γ’-precipitate/γ-

matrix interface, Figure 6.7. The γ’/γ interfaces are defined by employing the inflection-point

approach to determine a threshold isoconcentration surface that represents the average of the far-

field plateau concentrations of the γ’-precipitate and γ-matrix phases [55]. The phase

compositions at, and away from, the interface evolve temporally, and the interfacial widths,

defined as the distance between the plateau concentrations of the two phases, decrease with

aging time, Figure 6.8. Interestingly, an accumulation of Ni is observed ~ 2 nm into the γ-matrix.

We note that there is no evidence of non-monotonic (confined) segregation of Ta at the γ’-

precipitate/γ-matrix heterophase interface.

Page 139: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

139

Figure 6.7. The elemental concentration profiles across the primary γ’-precipitate/γ-matrix heterophase interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The phase compositions evolve temporally, and the widths of the concentration profiles decrease with increasing aging time. An accumulation of Ni is observed to develop on the γ-matrix side of the interface, evidence of a kinetic effect associated with the addition of Ta. The concentration profiles shown here represent the average of all of the interfaces in the analyzed volume.

Page 140: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

140

Figure 6.8. The temporal evolution of the widths of the concentration profiles across the primary γ’-precipitate/γ-matrix interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The interfacial widths are observed to decrease with increasing aging time. The species with the largest diffusivity in the γ-matrix, Al and Ta, have the smallest interfacial widths, while the opposite is true for Ni and Cr, the slower diffusing species.

6.3.3 Partitioning of elemental species

The partitioning behavior of the elements can be determined quantitatively by calculating

the partitioning ratio, Kiγ’/γ; the ratio of the concentration of an element i in the γ’-precipitates, to

the concentration of the same element i in the γ-matrix, Figure 6.9. Aluminum and Ta are

observed to partition to the γ’-precipitates, while Cr and Ni partition to the γ-matrix. The strong

partitioning of Ta to the γ’-precipitates increases the partitioning of Cr to the γ-matrix and of Al

to the γ’-precipitates, when compared to the results obtained for a ternary Ni-10.0 Al-8.5 Cr at.%

alloy [55, 56]. For example, for a sample aged at 1073 K for 256 h, the value of KCrγ’/γ decreases

Page 141: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

141

from 0.63±0.01 to 0.26±0.03 and the value of KAlγ’/γ increases from 2.20±0.02 to 3.17±0.01, due

to the formation of a significantly higher volume fraction of γ’-precipitates. The value of KTaγ’/γ

for the Ni-Al-Cr-Ta alloy at 256 h is 10.25±0.07, and the partitioning of Ni is unaffected by the

addition of Ta.

Figure 6.9. The temporal evolution of the partitioning ratios, Kiγ’/γ, of the constituent elements i

of: (a) Ni-10.0 Al-8.5 Cr-2.0 Ta; and (b) Ni-10.0 Al-8.5 Cr at.% aged at 1073 K. In both alloys, Al is observed to partition to the γ’-precipitates, while Cr and Ni partition to the γ-matrix. Ta shows a strong preference for the γ’-phase.

Page 142: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

142

6.4 Discussion

6.4.1 Morphological development

The morphology of the γ’-precipitates is observed to evolve from a bimodal distribution

of spheroidal γ’-precipitates in the as-quenched state, to cuboids with significantly rounded

edges, and finally to cuboids and rectangular parallelepipeds aligned along orthogonal <001>-

type directions. The γ’-precipitates that form the bimodal distribution in the as-quenched state

nucleate rapidly during the quench due to the small barrier height to nucleation caused by the

large supersaturations of alloying elements. The larger γ’-precipitates nucleate first, and

subsequently undergo high-temperature growth. As the alloy cools, however, the diffusion fields

associated with the primary γ’-precipitates shrink, and growth slows [234]. With additional

cooling, a significant solute supersaturation develops in the interprecipitate spaces, resulting in a

secondary burst of precipitation, Figure 6.2. In a recent study of multimodal precipitation in the

commercial nickel-based superalloy Udimet 720 Li, Radis et al. [234] found that multimodal

precipitation is highly dependent on alloy cooling rate, and noted that this phenomenon has been

observed to occur in other nickel-based alloys during continuous cooling.

The nucleation of γ’-precipitates during the quench from 1503 K was modeled with

PrecipiCalc, which incorporates the results of the classical models of nucleation, growth and

coarsening. These calculations were performed assuming a continuous quench rate of 20 K s-1,

and a γ’-precipitate/γ-matrix interfacial free energy of 25 mJ m-2, based on results from

concentrated Ni-Al-Cr alloys [42, 52] aged at 873 K. The γ/γ’ solvus point was determined

experimentally by differential thermal analysis (DTA) to be 1393 K, which agrees well with the

value of 1412 K predicted by Thermo-Calc and employed in the PrecipiCalc modeling. The

Page 143: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

143

results of the PrecipiCalc modeling are shown in Figure 6.10, and predict a bimodal distribution

of γ’-precipitates. Primary γ’-precipitate nucleation is predicted to begin at a temperature of 1380

K due to a bulk free energy driving force for nucleation, ΔFnuc, of -123 J mol-1, and continues to

a temperature of 1357 K. Primary γ’-precipitate growth follows nucleation, and values of <R(t)>

of 28.2 nm, Nv(t) of 2.7 x 1021 m-3, and φ of 25.7 % are predicted for a temperature of 1194 K.

According to the simulation results, significant secondary γ’-precipitate nucleation begins at a

temperature of 1190 K due to a value of ΔFnuc of -132 J mol-1, and continues throughout the

remainder of the quench. The final as-quenched microstructure is predicted to be a bimodal

distribution of γ’-precipitates, with primary γ’-precipitates with <R(t)> and φ values of 30.2 nm

and 31.5 %, respectively, and secondary γ’-precipitates with <R(t)> and φ values of 0.8 nm and

14.2 %, respectively. The predicted values of Nv(t) for the primary and secondary γ’-precipitates

are 2.7 x 1021 and 3.4 x 1025 m-3, respectively.

The PrecipiCalc predictions of the values <R(t)>, φ and Nv(t) for the primary γ’-

precipitates of 30.2 nm, 31.5 % and 2.7 x 1021 m-3 are in good agreement with the measured

values of 30-40 nm, 28±13% and 3±2 x 1021 m-3. The PrecipiCalc predictions for the secondary

γ’-precipitate properties of 0.8 nm, 14.2 % and 3.4 x 1025 m-3 differ, however, from the

experimental values of <R>, φ and Nv(t) of 6.4±1.1 nm, 11±1% and 1.3±0.1 x 1023 m-3,

respectively. The accuracy of the predictions for the primary γ’-precipitates indicates that the

thermodynamics and kinetics of the γ/γ’ phase transformation are modeled correctly.

Inaccuracies in the predictions of the properties of the secondary γ’-precipitates may be due to

the limitations in of the extant models of nucleation and growth upon which PrecipiCalc is based.

Page 144: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

144

These models are best suited for dilute, misfit-free, binary alloys with near-zero volume fractions

of the secondary phase, and hence very small elemental supersaturations. These conditions are

not satisfied for a concentrated quaternary Ni-Al-Cr-Ta alloy with an equilibrium precipitated

volume fraction of 37.0 ± 9.0 %, and a complex bimodal microstructure in the as-quenched state.

We note that efforts to improve the predictive power of this modeling technique, and to

incorporate the effects of variables such as lattice misfit, are ongoing [217]. The PrecipiCalc

predictions are highly dependent on the choice of the alloy cooling rate, the shape of the cooling

curve, the γ/γ’-interfacial energy, and the kinetics and thermodynamics of the alloy during the

quench. Thus, the modeling predictions can be modified by adjusting any of these parameters.

Page 145: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

145

Figure 6.10. The values of the γ’-precipitate number density, Nv(t ), radius, <R(t )>, and volume fraction, φ, during the water quench from 1503 K for Ni-10.0 Al-8.5 Cr-2.0 Ta at.%, as modeled by PrecipiCalc. A bimodal distribution of γ’-precipitates is predicted for the as-quenched state with primary γ’-precipitates with <R(t)> and φ values of 30.2 nm and 31.5 %, and secondary γ’-precipitates with <R(t)> and φ values of 0.8 nm and 14.2 %. The predicted values of Nv(t) for the primary and secondary γ’-precipitates are 2.7 x 1021 and 3.4 x 1025 m-3, respectively.

Figure 6.11 shows the PrecipiCalc predictions of the values of ΔFnuc, the effective

nucleation rate, Jeff, and the critical radius for nucleation, R*, during the quench. Decreases in the

magnitude of ΔFnuc are predicted at 1380 and 1190 K due to the nucleation of primary and

secondary γ’-precipitates, respectively. For the remainder of the quench, the driving force for

phase separation increases rapidly. This is due to the increasing γ-matrix supersaturation values

that are not sufficiently consumed by primary γ’-precipitation growth, due, in turn, to the

Page 146: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

146

exponentially decreasing diffusivities with decreasing temperature. As expected, the predicted

values of R* decrease rapidly as the magnitude of ΔFnuc increases.

Figure 6.11. The evolution of the values of the bulk driving force for nucleation, ΔFnuc, the effective nucleation rate, Jeff, and the critical radius for nucleation, R*, during the quench. Decreases in the magnitude of ΔFnuc are predicted at 1380 and 1190 K due to the nucleation of primary and secondary γ’-precipitates, respectively. The magnitude of ΔFnuc increases and the values of R* decrease during the quench. This figure was calculated using PrecipiCalc.

The composition of the primary γ’-precipitates in the as-quenched state is measured to be

76.04±0.04 Ni-17.02±0.02 Al-3.37±0.02 Cr-3.57±0.03 Ta at.%, while the secondary γ’-

precipitates have a composition of 76.05±0.05 Ni-16.82±0.03 Al-4.60±0.03 Cr-2.53±0.03 Ta

at.%. We note that the composition of the larger γ’-precipitates is closer to the equilibrium

composition of 75.24±0.04 Ni-16.48±0.07 Al- 3.20±0.06 Cr-5.03±0.07 Ta at.% estimated from

APT data. The primary γ’-precipitates reject Cr to the γ-matrix, and preferentially consume the

Page 147: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

147

supersaturation of Ta, which partitions strongly to the γ’-phase. As such, the secondary γ’-

precipitates nucleate from a Cr-enriched and Ta-depleted γ-matrix, resulting in Cr and Ta

concentrations that are 36% larger and 29% smaller, respectively, than those measured for the

primary γ’-precipitates.

The morphological development of nickel-based superalloys has been shown to proceed

from spheroidal to cuboidal to cuboidal arrays and finally to solid-state dendrites [83]. The γ’-

precipitates of the Ni-Al-Cr-Ta alloy studied herein undergo the first two transitions,

transforming from spheroids-to-cuboids at an aging time of 0.25 h, and forming cuboidal arrays

by an aging time of 4 h. The sizes at which these transitions occur are a function of the

magnitude of the lattice parameter misfit, and are partially diffusion controlled [83]. The elastic

self-energy of a precipitate increases as R3, while the heterophase interfacial free energy

increases as R2. Thus, as growth and coarsening proceed, and the average γ’-precipitate radius

increases, the elastic energy ultimately determines the precipitate morphology, and is decreased

by the formation of cuboidal γ’-precipitates. The evolution of the spheroidal γ’-precipitates into

cuboids in the Ni-Al-Cr-Ta alloy commences at an aging time of 0.25 h, when the γ’-precipitates

have an average radius of 61±7 nm. In contrast, the radius at which the microstructure becomes

cuboidal for the Ni-Al-Cr alloy was estimated to be ~ 88 nm at an aging time of 64 h at 1073 K

[56]. The lattice parameters of the γ-matrix and the γ’-precipitates of the Ni-Al-Cr-Ta alloy are

estimated from PrecipiCalc to be 0.3589 and 0.3603 nm, respectively, resulting in a lattice

parameter misfit, δ, of 0.0039. The value of δ of the Ni-Al-Cr alloy was estimated to be 0.0022 ±

0.0007 [56]. The addition of 2.0 at.% Ta increases the lattice parameter misfit by 78 %, resulting

in a spheroid to cuboid transition that occurs at a smaller γ’-precipitate size, and at an earlier

Page 148: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

148

aging time. For aging times of 64 and 256 h, the γ’-precipitates have a non-equiaxed, rectangular

parallelepiped morphology with {001}-type facets. The parallelepipedic morphology of the γ’-

precipitates further minimizes the elastic energy [209, 210].

The alignment of the γ’-precipitates results from the minimization of the elastic

interactions between the γ'-precipitates, where the elastic interaction energy depends on the

elastic anisotropy, the difference in the elastic constants of the two phases, and the sign and

magnitude of the misfit strain [235-237]. The alignment of γ’-precipitates into cuboidal and

parallelepipedic arrays was found to occur after 64 h of aging at 1073 K for the base Ni-Al-Cr

alloy [56], while the formation of arrays begins at 4 h for the Ni-Al-Cr-Ta alloy. The increased

lattice parameter misfit due to the addition of Ta leads to γ’-precipitate alignment at smaller

precipitate radii and earlier aging times in the Ni-Al-Cr-Ta alloy. The cuboidal γ’-precipitates in

both alloys have a coherent cube-on-cube relationship with the γ-matrix, with clearly defined

{001}-type facets, and are aligned along the elastically soft <001>-type directions, Figure 6.6.

6.4.2 Compositional evolution

From the APT results, Al and Ta partition strongly to the γ’-precipitate phase, while Ni

and Cr partition to the γ-matrix phase, in agreement with past experimental results [226, 229,

230, 238-241]. The equilibrium composition of the γ’-precipitates suggests that Al, Ta and Cr

occupy the Al sublattice sites of the L12-ordered γ’-phase of the Ni-Al-Cr-Ta alloy at 1073 K. A

first-principles study, discussed in detail in Chapter 7, confirms that Ta and Cr substitute

preferentially at the Al sublattice sites of the Ni3Al-L12 structure [61], in agreement with

previous work on this subject [138, 240, 242-257].

Page 149: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

149

The measured γ’-precipitate volume fractions, Table 6.1, are close to their equilibrium

values, suggesting that the Ni-Al-Cr-Ta alloy may be undergoing quasi-stationary state

coarsening. The first comprehensive mean-field treatment of precipitate coarsening, due to

Lifshitz and Slyozov [125] and Wagner [126], is limited to very dilute binary alloys with

spatially-fixed spherical precipitates whose initial compositions are equal to their equilibrium

values. Umantsev and Olson (UO) demonstrated that the exponents of the temporal power-laws

predicted for binary alloys by LSW-type models are identical for concentrated multi-component

alloys, but that the explicit expressions for the rate constants depend on the number of

components [130]. In the UO model, the asymptotic solution for the supersaturation of solute

element i, ΔCi(t), the difference between the concentration in the far-field γ-matrix, >< )(, tC ffiγ ,

and the equilibrium γ-matrix solute-solubility, )(C eq,i ∞γ , is:

, , 1/3,( ) ( ) ( ) ;ff eq

i i i i UOC t C t C tγ γγ κ −Δ =< > − ∞ = (6.1)

where UO,iκ is the coarsening rate constant for )t(CiγΔ and is related to the partial derivatives of

the chemical potentials, the γ/γ’ interfacial free energy, the partitioning ratios of the alloying

elements, and the diffusivities of the elements [130]. The temporal evolution of the

supersaturations of the alloying elements in the γ-matrix and γ’-precipitates are observed to

follow approximately the t-1/3 prediction of the UO model, Figure 6.12. The supersaturations of

Ni, Al, Cr and Ta in the γ-matrix evolve as t-0.33±0.04, t-0.35±0.05, t-0.31±0.04 and t-0.30±0.04, respectively,

while the supersaturations of the same elements in the γ'-precipitates evolve as t-0.33±0.03, t-0.36±0.05,

t-0.28±0.04 and t-0.36±0.03, respectively. Quasi-stationary state coarsening has begun by an aging time

of 1 h, and continues throughout the aging times studied here.

Page 150: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

150

Figure 6.12. The magnitude of the values of the supersaturations, )(tCiΔ , of Ni, Al, Cr and Ta in the γ-matrix and γ’-precipitates for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K. The magnitudes of the )(tCiΔ values decrease as approximately t-1/3 in the coarsening regimes for both phases, as predicted by the Umantsev and Olson (UO) models for quasi-stationary state coarsening.

The values of )(, ∞eqiC γ and )(,' ∞eq

iC γ estimated from APT data employing Equation 6.1

are listed in Table 6.2, and compared to the equilibrium values predicted by Thermo-Calc, and

the concentrations measured by APT and ICP analysis at 256 h. The phase compositions after

Page 151: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

151

256 h of aging are near their equilibrium values, suggesting that the magnitudes of the γ-matrix

supersaturations are small and that the γ’-precipitate growth and coarsening regime is nearly

complete.

Table 6.2. Equilibrium compositions (at.%) of the γ’-precipitate, )(C eq,'i ∞γ , and γ-matrix,

)(C eq,i ∞γ , determined by APT, ICP chemical analysis, and thermodynamic modeling employing

Thermo-Calc for Ni-10.0 Al-8.5 Cr-2.0 Ta at.% aged at 1073 K.

Technique: )(C eq,'Ni ∞γ )(,' ∞eq

AlCγ )(,' ∞eqCrCγ )(,' ∞eq

TaCγ

APT, 256 h: 75.37±0.05 16.56±0.08 3.25±0.09 4.82±0.09

APT, extrapolated: 75.24±0.08 16.48±0.07 3.20±0.06 5.03±0.07

ICP, 256 h: 75.56 16.57 3.28 4.58

Thermo-Calc: 76.47 17.00 2.66 3.88

Technique: )(, ∞eqNiC γ )(, ∞eq

AlC γ )(, ∞eqCrC γ )(, ∞eq

TaC γ

APT, 256 h: 81.65±0.01 5.23±0.03 12.66±0.03 0.47±0.03

APT, extrapolated: 81.68±0.04 5.18±0.08 12.70±0.02 0.44±0.06

ICP, 256 h: 83.17 5.49 10.74 0.60

Thermo-Calc: 81.37 5.68 12.11 0.84

The most striking feature of the concentration profiles across the γ’-precipitate/γ-matrix

heterophase interface, Figure 6.7, is the accumulation of Ni in the γ-matrix, roughly 2 nm from

the interface. This effect was observed at all γ/γ’ interfaces, regardless of direction, and is likely

kinetic in origin, resulting from a decrease in the diffusivity of Ni in the γ-matrix. Estimates of

the tracer diffusivities from Dictra predict a threefold decrease in the calculated γ-matrix tracer

diffusivity of Ni, γNiD , from 6.06 x 10-18 m2 s-1 to 2.02 x 10-18 m2 s-1 due to a 2.0 at.% addition of

Page 152: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

152

Ta, Table 6.3. This decrease in Ni diffusivity is significant because Ni becomes the least mobile

species in the γ-matrix phase. Hence, as phase decomposition proceeds, Ni partitions to the γ-

matrix and accumulates on the γ-matrix side of the γ'/γ interface. The diffusivities of all of the

atomic species decrease due to the addition of Ta, thus the γ’-precipitate coarsening kinetics are

expected to decrease. We note that the calculations of the γ-matrix diffusivities used herein are

dependent on the thermodynamic predictions from the Saunders database. Since the Saunders

database predicts an equilibrium partitioning ratio for Ta of 4.61, while a value of 11.4 is

measured herein, the predictions of the database are not absolutely accurate. As such, the

predictions of the γ-matrix kinetics likely suffer from some inaccuracy, as previously observed in

comparisons of microanalysis results of diffusion couples to Dictra predictions [217].

Table 6.3. Tracer diffusivity of element i in the γ-matrix, γ

iD , of Ni-10.0 Al-8.5 Cr and Ni-10.0 Al-8.5 Cr-2.0 Ta at.% calculated with Dictra, and Campbell [208] and Saunders [115] databases at 1073 K.

Alloy (at.%) γNiD

(10-18 m2s-1)

γAlD

(10-18 m2s-1)

γCrD

(10-18 m2s-1)

γTaD

(10-18 m2s-1)Ni-10.0 Al-8.5 Cr 6.06 13.9 5.13 -

Ni-10.0 Al-8.5 Cr-2.0 Ta 2.02 12.6 4.80 5.14

The widths of the concentration profiles across the γ’-precipitate/γ-matrix interface are

observed to decrease with increasing aging time, Figure 6.8, which is a kinetic effect. As aging

progresses, the two species with the largest predicted diffusivities in the γ-matrix, Al and Ta,

Table 6.3, have the smallest interfacial widths, while the opposite is true for Ni and Cr, the

slower species. This is expected, though the diffusivities of the atomic species in the γ’-

Page 153: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

153

precipitate, for which there are no extant kinetic databases that include Ta, also play an important

role in determining the interfacial widths. The interfacial widths of model nickel-based

superalloys have been measured to range from a single atomic layer to up to 5 nm, depending on

the alloy system and temperature studied [39, 222, 258].

6.4.3 Comparison with other Ni-Al-Cr-X alloys

This work is part of a systematic investigation of the effects of dilute refractory additions

on the γ/γ’ phase decomposition of a model Ni-10.0 Al-8.5 Cr at.% alloy aged at 1073 K. A 2.0

at.% addition of W [55, 56] was previously found to increase the microhardness and decrease the

coarsening kinetics of the γ’-phase. Tungsten was found to partition preferentially to the γ’-

precipitates, with no evidence of non-monotonic (confined) segregation of W at the heterophase

interface. The addition of W led to stronger partitioning of Al to the γ’-precipitates, and of Cr to

the γ-matrix, than was measured for the base Ni-Al-Cr alloy. The Ni-Al-Cr-W alloy, like Ni-Al-

Cr-Ta and the base Ni-Al-Cr alloys, undergoes a γ’-precipitate morphological evolution from

spheroids-to-cuboids, as well as alignment of the cuboids along the elastically-soft <001>-type

directions.

A 2.0 at.% addition of Re [58, 59] was found to result in the strong partitioning of Re to

the γ-matrix, and increased partitioning of Al to the γ′-precipitates, and of Cr to the γ-matrix.

Rhenium was found to decrease the coarsening kinetics, and to stabilize the spheroidal

morphology of the γ’-precipitates, allowing the γ’-phase to remain spheroidal after aging to 256

h. A high degree of γ’-precipitate coagulation and coalescence was detected, in contrast with the

Ni-Al-Cr-Ta and Ni-Al-Cr-W alloys. Additionally, no evidence of confined (non-monotonic)

Page 154: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

154

interfacial Re segregation was detected for Ni-Al-Cr-Re, unlike commercial nickel-based

superalloys, such as René N6, which exhibits significant confined interfacial segregation of Re

of 2.41 atoms nm-2 [110, 111]. The effects of Re and Ta in modern nickel-based superalloys are

dramatically different, though both decrease the kinetics of γ-phase decomposition, leading to

phase stability at high temperature.

The accumulation of Ni on the γ-matrix side of the γ’/γ heterophase interface is unique to

the addition of Ta, due to a decrease in the value of γNiD from 6.06 to 2.02 x 10-18 m2 s-1 for a 2.0

at.% addition of Ta. The value of γNiD was calculated to decrease to 4.92 x 10-18 m2 s-1 [56] and

4.05 x 10-18 m2 s-1 [59] due to the additions of W and Re, respectively, leading to a deceleration

of the coarsening kinetics. The values of γNiD are, however, apparently still sufficient to transport

Ni atoms away from the γ’/γ interface during phase decomposition of the Ni-Al-Cr-W and Ni-Al-

Cr-Re alloys, avoiding Ni accumulation. The bimodal distribution of γ’-precipitates in the as-

quenched state is also unique to Ni-Al-Cr-Ta among the Ni-based alloys we have studied to date.

The bimodal distribution is due to the initially high value of the Ta supersaturation in the γ-

matrix, and to the relatively large diffusivity of Ta in the γ-matrix. The value of γTaD of 5.14 x 10-

18 m2 s-1 is significantly greater than the values of γWD and γ

ReD of 4.93 x 10-19 m2 s-1 and 1.07 x

10-19 m2 s-1, respectively, in the model Ni-Al-Cr alloys.

Page 155: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

155

6.5 Summary and conclusions

The effects of a 2.0 at.% addition of tantalum to a model Ni-10.0 Al-8.5 Cr at.% superalloy aged

at 1073 K for 0 to 256 h are assessed using mainly scanning electron microscopy and atom-probe

tomography (APT), leading to the following conclusions:

• The addition of Ta results in a 47±5% increase in microhardness due to increases in

solid-solution strengthening and to a dramatic increase in the volume fraction of the γ’-

phase. After 256 h of aging, the value of φ from phase extraction results is 18.9% for Ni-

Al-Cr and 36.4% for the Ni-Al-Cr-Ta alloy. The equilibrium volume fraction of the γ’-

phase for the Ni-Al-Cr-Ta alloy is estimated by APT to be 37.0 ± 9.0 %.

• The morphology of the γ’-precipitates evolves from a bimodal distribution of spheroidal

precipitates in the as-quenched state, to primary γ’-precipitate cuboids with significantly

rounded edges, and finally to cuboidal and parallelepipedic γ’-precipitates with clearly

defined {001} facets, aligned along the elastically soft <001> directions.

• The morphology of the γ’-precipitate phase is spheroidal in the as-quenched state, and a

bimodal particle size distribution is apparent. The smaller secondary γ’-precipitates are

observed to have an average radius of 6.4±1.1 nm, and the larger primary γ’-precipitates

have radii on the order of 30-40 nm. A solute-depleted, precipitate-free shell with a

thickness of 5.7±1.1 nm surrounds the primary γ’-precipitates. Precipitation modeling

employing the commercial software program PrecipiCalc predicts that primary and

secondary γ’-precipitate nucleation begin at 1380 and 1190 K, respectively. The smaller

secondary γ’-precipitates are only observed in the as-quenched state and disappear

between 0 and 0.25 h of aging at 1073 K due to coarsening of the primary γ’-precipitates.

Page 156: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

156

• The transition from a spheroidal-to-cuboidal morphology commences at an aging time of

0.25 h, when the primary γ’-precipitates have an average radius of 61±7 nm. In contrast,

the radius at which the microstructure becomes cuboidal for the base Ni-Al-Cr alloy is

estimated to be ~ 88 nm at an aging time of 64 h at 1073 K [56]. The addition of 2.0 at%

Ta increases the lattice parameter misfit, δ, by 78 %, from 0.0022 ± 0.0007 [56] to

0.0039. This increase in δ results in a spheroid-to-cuboid transition that occurs at a

smaller γ’-precipitate radius, and at an earlier aging time, than for the base ternary Ni-Al-

Cr alloys, to minimize the free energy associated with elastic misfit.

• Alignment of primary γ’-precipitates into cuboidal and parallelepipedic arrays occurs

after 4 h of aging at 1073 K for the model Ni-Al-Cr-Ta alloy, while the formation of

arrays was observed to begin after 64 h for the base Ni-Al-Cr alloy. The increased lattice

parameter misfit due to the addition of Ta leads to alignment at smaller primary γ’-

precipitate radii and earlier aging times in the Ni-Al-Cr-Ta alloy.

• Aluminum and Ta partition to the γ’-precipitates, while Cr and Ni partition to the γ-

matrix. The strong partitioning of Ta to the γ’-precipitates increases the partitioning of Cr

to the γ-matrix and of Al to the γ’-precipitates, when compared to the results of the base

Ni-Al-Cr alloy.

• The concentration profiles of the constituent elements of the Ni-Al-Cr-Ta alloy across all

γ’-precipitate/γ-matrix interfaces exhibit an accumulation of Ni on the γ-matrix side of the

interface. This effect is likely due to the diminution of the mobility of Ni in the γ-matrix

caused by the addition of Ta, and was not previously observed in alloys with 2.0 at.%

additions of W or Re.

Page 157: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

157

• The widths of the concentration profiles across the γ’-precipitate/γ-matrix interface are

observed to decrease with increasing aging time. Upon aging, the atomic species with the

largest diffusivities in the γ-matrix, Al and Ta, have the smallest interfacial widths, while

the opposite is true for Ni and Cr, the slower species.

• For aging times of 1 h and longer, the temporal evolutions of the supersaturations of the

alloying elements in the γ-matrix and γ’-precipitates are observed to follow

approximately the t-1/3 prediction of the Umantsev-Olson quasi-stationary state

coarsening model.

Page 158: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

158

Chapter 7

Chromium and tantalum site substitution patterns in Ni3Al (L12) γ’-

precipitates

Abstract

The site substitution behavior of Cr and Ta in the Ni3Al (L12)-type γ’-precipitates of a Ni-Al-Cr-

Ta alloy is investigated by atom-probe tomography (APT) and first-principles calculations.

Measurements of the γ’-phase composition by APT suggest that Al, Cr and Ta share the Al

sublattice sites of the γ’-precipitates. The calculated substitutional energies of the solute atoms at

the Ni and Al sublattice sites indicate that Ta has a strong preference for the Al site, while Cr has

a weak Al site preference. Furthermore, Ta is shown to replace Cr at the Al sublattice sites of the

γ’-precipitates, altering the elemental phase partitioning behavior of the Ni-Al-Cr-Ta alloy.

Page 159: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

159

7.1 Introduction

Modern nickel-based superalloys are used in land-based and aerospace turbine engines at

operating temperatures up to 1373 K. The high-temperature strength and creep resistance of

these alloys are due primarily to strengthening by the precipitated γ’-(L12) phase, which can

accommodate substantial solute additions [3]. The mechanical properties of the γ’-phase depend

on the sublattice site substitution behavior of these alloying additions [80], motivating an

investigation of the site occupancy of Cr and Ta in a model Ni-Al-Cr-Ta alloy that has been

studied by atom-probe tomography (APT) and scanning electron microscopy [60].

The addition of Cr to Ni-Al alloys reduces the lattice parameter misfit between the γ’-

phase and the γ-matrix, often leading to spheroidal, nearly misfit-free γ’-precipitates [12].

Tantalum provides solid-solution strengthening and increases the volume fraction of the γ’-

precipitate phase, thereby providing significant strengthening while improving γ’-phase stability

and service life [3]. Previous APT [240], atom-probe field-ion microscope [242, 243], and x-ray

analysis experiments [244-246], as well as results of first-principles calculations [247-251], the

cluster variation method [252-254], and other techniques [255-257], have indicated that Ta

occupies the Al sublattice sites in ordered Ni3Al. These findings contradict experimental results

from ion-channeling and nuclear-reaction analysis studies [259] and research based on short-

range ordering employing pseudopotential approximations [260], which claim that Ta occupies

the Ni sublattice sites in Ni3Al. The site preference of Cr has generally been found to depend on

the alloy composition and aging treatment [244, 247, 249, 253, 257, 260, 261], though some

studies concluded that Cr occupies the Al sites, regardless of composition [246], or temperature

[262].

Page 160: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

160

7.2 Experimental

We present APT results and first-principles calculations that elucidate the site

substitution behavior of Cr and Ta in the γ’-phase of a Ni-Al-Cr-Ta superalloy aged at 1073 K. A

Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy was homogenized at 1573 K for 20 h, then annealed in the

γ−phase field at 1503 K for 3 h, and water-quenched. Ingot sections were aged at 1073 K under

flowing argon for times ranging from 0.25 to 256 h, then water quenched, and microtip

specimens were prepared for study by APT. Pulsed-laser APT data collection was performed at

an evaporation rate of 0.04 ions pulse-1, a specimen temperature of 40.0±0.3 K, a pulse energy of

0.6 nJ pulse-1, a pulse repetition rate of 200 kHz, and a gauge pressure of < 6.7 x 10-8 Pa. APT

data were visualized and analyzed with IVAS™ 3.0 (Imago Scientific Instruments). Figure 7.1

shows an APT reconstruction that intersected a cuboidal γ’-precipitate in a sample aged for 256

h. The strong partitioning of Al and Ta, shown in orange and red, respectively, to the γ’-phase,

and Cr, in blue, to the γ-phase is demonstrated.

Figure 7.1. APT reconstruction of a cuboidal γ’-precipitate in a model Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 256 h. The elements that partition to the γ’-precipitates, Al and Ta, are shown in red and orange, respectively, while Cr, which partitions to the γ-matrix, is shown in blue; Ni atoms are omitted for clarity.

Page 161: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

161

7.3 Results and discussion

The composition profiles across the γ’/γ interface in Figure 7.1 are displayed in Figure

7.2. The measured composition of the γ’-precipitates is 75.37±0.05 Ni- 16.56±0.08 Al-

3.25±0.09 Cr- 4.82±0.09 Ta at.%, suggesting that Cr and Ta occupy the Al sites of the Ni3Al γ’-

precipitates. The relatively Cr rich γ’-precipitate composition of 76.49±0.13 Ni- 17.47±0.12 Al-

6.04±0.07 Cr at.% measured for the reference Ni-Al-Cr alloy [55, 56], indicates that the addition

of 2.0 at.% Ta leads to the preferential replacement of Cr by Ta in the γ’-precipitates.

Figure 7.2. The elemental concentration profiles across the γ’/γ interface for a Ni-10.0 Al-8.5 Cr-2.0 Ta at.% alloy aged at 1073 K for 256 h. Tantalum and Al partition to the γ’-precipitates, while Cr partitions to the γ-matrix.

Page 162: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

162

First-principles calculations were performed to confirm the site substitution preferences

of the solute atoms in the γ’-precipitates and to determine the thermodynamic driving force for

the replacement of Cr by Ta atoms in the γ’-phase. The calculations employed the plane-wave

pseudopotential total energy method with generalized gradient approximations [263], as

implemented in the Vienna ab initio simulation package (VASP) [149-151], using the projector

augmented-wave (PAW) potentials [264]. A plane wave cutoff energy of 300 eV and 8×8×8

Monkhorst-Pack k-point grids were used, and found to be sufficient to give fully converged

results. A three-dimensional periodic supercell with 2×2×2 unit cells (32 atoms) was employed

to determine the total energies of the cells, which converge to 2 x 10−5 eV atom−1, while residual

forces converge to 0.005 eV nm−1. The Ni3Al (L12) structure was fully relaxed, and the lattice

parameter was determined to be 0.3566 nm, in good agreement with the experimental room

temperature value of 0.3570 nm [3].

The antisite formation energies associated with the Ni3(Al1-yNiy) and (Ni1-xAlx)3Al

structures were calculated and defined as:

)()(313 )( Ni

totAlNiAl

totNiAlNiAlNi EEE

yyμμ +−+=

−→ ; (7.1)

)()(331 )( Al

totAlNiNi

totAlAlNiNiAl EEE

xxμμ +−+=

−→ ; (7.2)

where μi is the chemical potential per atom of the bulk element i assuming same cell symmetry,

and are -5.432 and -3.697 eV atom-1 for Ni and Al, respectively. The antisite formation energies

are ENi→Al = 0.986 eV atom-1 and EAl→Ni = 0.742 eV atom-1, and are energetically unfavorable.

Other first-principles studies have found that the antisite formation energies are similar and range

from 0.51 to 0.99 eV [265-267]. Another investigation calculated the Al and Ni antisite energies

to be -0.92 and 2.04 eV, respectively [251].

Page 163: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

163

Four substitutional structures, (Ni1-xCrx)3Al, Ni3(Al1-yCry), (Ni1-xTax)3Al and Ni3(Al1-

yTay), were modeled by substituting one Cr or Ta atom at one of the Ni or Al sublattice sites.

Table 7.1 summarizes the calculated total energies of the relaxed structures, Etot, and the

substitutional energies, EZ→Ni and EZ→Al, (Z = Cr, Ta), which are defined as follows:

)()(331 )( Z

totAlNiNi

totAlZNiNiZ EEE

xxμμ +−+=

−→ ; (7.3)

)()(313 )( Z

totAlNiAl

totZAlNiAlZ EEE

yyμμ +−+=

−→ ; (7.4)

where the chemical potentials of Cr and Ta are -9.103 and -11.478 eV atom-1, respectively. The

calculated results demonstrate that Ta atoms strongly prefer to occupy the Al sublattice sites of

Ni3Al, as ETa→Al is significantly smaller than ETa→Ni.. Chromium weakly prefers to substitute at

the Al sites, as ECr→Al is only slightly smaller than ECr→Ni. The substitutional energy associated

with the replacement of Cr atoms by Ta atoms on the Al sublattice sites of the Ni3Al structure is

calculated as follows:

)()( )()( 1313 Tatot

CrAlNiCrtot

TaAlNiCrTa yyyyEEE μμ +−+=

−−→ ; (7.5)

and is -1.038 eV atom-1. Thus the replacement of Cr by Ta in the γ’-precipitates is energetically

favorable, and is observed by APT to occur.

To verify that the substitutional patterns of Cr modeled herein result in the lowest energy

configurations, two other substitutional patterns (Ni1-xAlx)3(Al1-yCry) and (Ni1-xCrx)3(Al1-y-

Niy) were modeled. These structures may form by the occupation of a Ni or Al sublattice by a Cr

atom, and the displacement of the Ni or Al atom to an antisite. These structures are also found to

Page 164: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

164

be energetically unfavorable, yielding formation energies of 1.138 and 1.337 eV atom-1 for

(Ni1-xAlx)3(Al1-yCry) and (Ni1-xCrx)3(Al1-yNiy), respectively.

Table 7.1. Total, Etot, and substitutional, EZ→Ni, Al, energies determined by first-principles calculations. (For these calculations x = 0.042, y = 0.125, and Z = Cr, Ta.)

Etot (eV) EZ→Ni, Al (eV atom-1)

Ni3Al -174.944 -

(Ni1-xCrx)3Al -177.966 0.648

Ni3(Al1-yCry) -179.748 0.565

(Ni1-xTax)3Al -180.957 0.033

Ni3(Al1-yTay) -183.198 -0.473

The local stresses and strains due to site substitution result in the average atomic forces

and displacements at the first nearest-neighbor distance displayed in Table 7.2. We note that no

atomic force was measured for relaxed Ni3Al prior to site substitution. The substitution of Ta at

the Al sublattice sites results in smaller values of the atomic force and displacement than

substitution of Ta at the Ni sites. The atomic forces and displacements associated with Cr

substitution at the Ni and Al sites are comparable, providing supporting evidence that Cr only

weakly prefers Al sublattice sites.

Page 165: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

165

Table 7.2. Average atomic forces and displacements at the first nearest-neighbor distance from first-principles calculations for four different substitutional structures.

Average Atomic Force (eV Ǻ-1) Average Atomic Displacement (Ǻ)

(NixCr1-x)3Al 0.0150 0.0294

Ni3(AlyCr1-y) 0.0112 0.0325

(NixTa1-x)3Al 0.0154 0.0479

Ni3(AlyTa1-y) 0.0095 0.0215

From APT results, the addition of Ta to a Ni-Al-Cr alloy leads to the rejection of Cr from

the γ’-precipitates.. The elemental phase partitioning behavior is quantified by the partitioning

ratio, Kiγ’/γ, the ratio of the concentration of an element i in the γ’-precipitates to the

concentration of the same element in the γ-matrix. For example, for a sample aged at 1073 K for

256 h, the value of /CrK γ γ′ decreases from 0.63±0.01 to 0.26±0.03 and the value of /

AlK γ γ′

increases from 2.20±0.02 to 3.17±0.01, due to the formation of a significantly higher volume

fraction of Al-rich γ’-precipitates. The value of /TaK γ γ′ for the Ni-Al-Cr-Ta alloy at 256 h is

10.25±0.07, and the partitioning of Ni is unaffected by the addition of Ta.

7.4 Conclusions

To conclude, Ta and Cr substitute preferentially at the Al sublattice sites in Ni3Al (L12).

The substitution of Ta for Al is energetically favorable, while the substitution of Cr for Al is less

so. This results in the replacement of Cr by Ta at the Al sublattice sites and in the γ’-precipitates,

and alters the partitioning behavior of the elements in the Ni-Al-Cr-Ta alloy.

Page 166: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

166

Chapter 8

Future Work

There are multiple possible extensions of the research presented herein. Given the

success of the atom-probe technique in analyzing the phase decomposition of Ni-Al-Cr-X alloys,

a natural continuation of this work would be more detailed simulation of these results,

employing, for example, molecular dynamics, phase-field modeling, first-principles calculations,

and lattice kinetic Monte-Carlo simulation. These detailed analyses would provide further insight

into the kinetic pathways that lead to phase decomposition in these alloys, and would provide

quantitative data to improve the extant theories of precipitate nucleation, growth and coarsening.

These experimental and simulated results also provide insight into precipitate coagulation and

coalescence, a mechanism that is critical to phase decomposition in these alloys, and has only

recently been examined in detail [39].

The γ/γ’ interfacial region is of critical importance in the phase decomposition of both

model and commercial nickel-based superalloys. This region has been shown to have a width on

the order of several nanometers for other model nickel-based superalloys [42, 52, 56, 58, 62,

222-224, 268]. Classical models of phase decomposition often ignore the interfacial region,

however, given the experimental and simulated evidence of the diffuseness of the γ/γ’ interface,

advanced models of phase separation will need to account for the interfacial region [269].

Page 167: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

167

The energy of the γ/γ’ interface is of critical importance to the properties and structures of

nickel-based superalloys; it remains, however, very difficult to determine the energy directly. In

order to estimate the interfacial energy, one must resort to either indirect measurements, from the

results of the classical models of nucleation or coarsening, or directly through simulation or fist-

principles calculations [222]. Ardell [13, 145] first demonstrated that the value of σγ/γ’ can be

estimated by comparison of independent measurements of the kinetics of γ’-particle growth

during coarsening, and of the γ-matrix supersaturation to the predictions of classical LSW-type

models of Ostwald ripening [125, 126]. Attempts to remove many of the restrictions of the LSW

model have been made, such as efforts to incorporate the effects of finite volume fractions in

these models [128]. Calderon et al. developed a model for binary alloys that incorporates a finite

precipitate volume fraction, and allows for nonzero solubilities and nonideal solution

thermodynamics [270]. The results of the work by Calderon et al. were applied by Ardell to data

from Ni-3.5 and Ni-6.0 wt.% Al alloys aged at 898 and 988 K [145], and a Ni-12.5 at.% alloy

aged at 823-973 K [271], to yield values of σγ/γ’ ranging from 8 to.0 8.7 mJ m-2, assuming non-

ideal solution thermodynamics. Wang et al. estimated values of σγ/γ’ of ~8.5 to 13 mJ m-2 by the

cluster variation method for a Ni-18 Al at.% alloy at temperatures of 873 to 1573 K [272].

Mishin employed an embedded-atom-potential to calculate Ni/Ni3Al interfacial energies of 46,

28 and 12 mJ m-2 for the (100), (110) and (111) planes, respectively [268]. Recent atom-probe

tomographic investigations of Ni-Al-Cr alloys [28, 42, 46, 52] aged at 873 K have led to

estimates of σγ/γ’ of 20-25 mJ m-2, by employing the coarsening model developed by Kuehmann

and Voorhees for ternary alloys [132]. A thorough study of the effect of ternary additions on the

γ/γ’ interfacial energy of the binary Ni-Al system, employing both experimental and simulation

Page 168: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

168

techniques, would provide a valuable tool for the design of complex, concentrated

multicomponent superalloys.

The work on Ni-Al-Cr-Ta is part of a systematic investigation of the effects of dilute

refractory additions, such as W [43, 55, 56], Re [50, 51, 57-59], and Ta [60-62], on the γ/γ’ phase

decomposition of a model Ni-10.0 Al-8.5 Cr at.% alloy aged at 1073 K. A very interesting

extension of this work would be a more detailed study of the interactions between refractory

additions in more complex model alloys containing more than one refractory addition. For

example, the interplay between Ta and W has recently been shown to be significant [63, 273],

thus the next logical step in this work would be a detailed study of a Ni-Al-Cr-Ta-W alloy,

employing both atom-probe tomography and first-principles simulations.

Page 169: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

169

References

1. Seth, B. B., Superalloys-The Utility Gas Turbine Perspective. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., Green, K. A., McLean, M., Oson, S. L. and Schirra, J. J., Editors. TMS, Warrendale, PA, 2000.

2. Energy Information Administration: Electric Power Annual Report 2006, http://www.eia.doe.gov/, (2006).

3. Reed, R. C., The Superalloys: Fundamentals and Applications. (Cambridge University Press, New York, 2006), 372.

4. Nabarro, F. R. N., Rafting in Superalloys, Metall. Mater. Trans. A, 27 (3) (1996), 513-530.

5. Argence, D., Vernault, C., Desvallees, Y. and Fournier, D., MC-NG: A 4th Generation Single-Crystal Superalloy for Future Aeronautical Turbine Blades and Vanes. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., Green, K. A., McLean, M., Oson, S. L. and Schirra, J. J., Editors. TMS, Warrendale, PA, 2000.

6. Caron, P., High γ' Solvus New Generation Nickel-Base Superalloys for Single-Crystal Turbine Applications. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., Green, K. A., McLean, M., Olson, S. L. and Schirra, J. J., Editors. TMS, Warrendale, PA, 2000.

7. Durand-Charre, M., The Microstructure of Superalloys. (Gordon and Breach Science, Amsterdam, 1997),

8. Argon, A. S., Strengthening Mechanisms in Crystal Plasticity. (Oxford University Press, Oxford, 2007),

9. Baldan, A., Review Progress in Ostwald Ripening Theories and their Applications to the Gamma'-precipitates in Nickel-base Superalloys - Part II - Nickel-base Superalloys, J. Mat. Sci., 37 (12) (2002), 2379-2405.

10. Baldan, A., Review Progress in Ostwald Ripening Theories and their Applications to Nickel-base Superalloys. Part I: Ostwald Ripening Theories, J. Mater. Sci., 37 (11) (2002), 2171-2202.

11. Hirata, T. and Kirkwood, D. H., Prediction and Measurement of Precipitate Number Densities in a Nickel-6.05 Wt.% Aluminium Alloy, Acta Metall., 25 (12) (1977), 1425-1434.

12. Taylor, A., Floyd W., The Constitution of Nickel-Rich Alloys of the Nickel-Chromium-Aluminium system, J. Inst. Metals, 81 (1952), 451-464.

13. Ardell, A. J. and Nicholson, R. B., Coarsening of Gamma Prime in Ni-Al Alloys, J. Phys. Chem. Solids, 27 (11-1) (1966), 1793-1804.

14. Kirkwood, D. H., Precipitate Number Density in a Ni-Al Alloy at Early Stages of Ageing,

Page 170: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

170

Acta Metall., 18 (6) (1970), 563-570.

15. Chellman, D. J. and Ardell, A. J., Coarsening of γ' Precipitates at Large Volume Fractions, Acta Metall., 22 (5) (1974), 577-588.

16. Xiao, S. Q. and Haasen, P., HREM Investigation of Homogeneous Decomposition in a Ni-12 at-Percent Al-Alloy, Acta Metall. Mater., 39 (4) (1991), 651-659.

17. Hornbogen, E. and Roth, M., Distribution of Coherent Particles in Nickel Alloys, Z. Metall., 58 (12) (1967), 842-855.

18. Royer, A., Bastie, P. and Veron, M., In Situ Determination of γ' Phase Volume Fraction and of Relations Between Lattice Parameters and Precipitate Morphology in Ni-based Single Crystal Superalloy, Acta Mater., 46 (15) (1998), 5357-5368.

19. Broz, P., Svoboda, M., Bursik, J., Kroupa, A. and Havrankova, J., Theoretical and Experimental Study of the Influence of Cr on the γ+γ' Phase Field Boundary in the Ni-Al-Cr system, Mater. Sci. Eng., A, A325 (1-2) (2002), 59-65.

20. Messoloras, S. and Stewart, R. J., Studies of Precipitation Using Small-Angle Neutron Scattering (SANS), J. Appl. Crystall., 21 (6) (1988), 870-872.

21. Beddoe, R., Haasen, P. and Kostorz, G., Early Stages of Decomposition in Ni-Al Single Crystals Studied by Small-Angle Neutron Scattering. In: Decomposition of Alloys: The Early Stages, Haasen P, G. V., Wagner R, and Ashby MF, Editor. Pergamon Press, Oxford, 1984.

22. Staron, P. and Kampmann, R., Early-Stage Decomposition Kinetics in Ni-Al Alloys - I. Small and Wide-Angle Neutron Scattering Investigation on Ni-13 at.% Al and Clusterdynamic Modelling, Acta Mater., 48 (3) (2000), 701-712.

23. Gilles, R., Mukherji, D., Hoelzel, M., Strunz, P., Toebbens, D. M. and Barbier, B., Neutron and X-Ray Diffraction Measurements on Micro- and Nano-sized Precipitates Embedded in a Ni-based Superalloy and After their Extraction from the Alloy, Acta Mater., 54 (5) (2006), 1307-1316.

24. Wendt, H. and Haasen, P., Nucleation and Growth of γ'-Precipitates in Ni-14 at.% Al, Acta Metall., 31 (10) (1983), 1649-1659.

25. Vanbakel, G. P. E. M., Hariharan, K. and Seidman, D. N., On the Structure and Chemistry of Ni3Al on an Atomic-Scale via Atom-Probe Field-Ion Microscopy, Appl. Surf. Sci., 90 (1) (1995), 95-105.

26. Al-Kassab, T., Wollenberger, H., Schmitz, G. and Kirchheim, R., Tomography by Atom Probe Field ion Microscopy, Springer Ser. Mater. Sci., 50 (High-Resolution Imaging and Spectrometry of Materials) (2003), 271-320.

Page 171: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

171

27. Geber, G. P. and Kirchheim, R., Discontinuous Precipitation in a Ni-In Alloy Studied by Analytical Field Ion Microscopy, Acta Mater., 45 (5) (1997), 2167-2175.

28. Schmuck, C., Caron, P., Hauet, A. and Blavette, D., Ordering and Precipitation of γ' Phase in Low Supersaturated Ni-Cr-Al Model Alloy: An Atomic Scale Investigation, Philos. Mag. A, 76 (3) (1997), 527-542.

29. Chambreland, S., Walder, A. and Blavette, D., Early Stages of Precipitation of γ'-phase in a Nickel Base Superalloy: An Atom-Probe Investigation, Acta Metall., 36 (12) (1988), 3205-3215.

30. Miller, M. K., Contributions of Atom Probe Tomography to the Understanding of Nickel-based Superalloys, Micron, 32 (8) (2001), 757-764.

31. Simmons, J. P., Wen, Y., Shen, C. and Wang, Y. Z., Microstructural Development Involving Nucleation and Growth Phenomena Simulated with the Phase Field Method, Mater. Sci. Eng., A, A365 (1-2) (2004), 136-143.

32. Wang, J., Osawa, M., Yokokawa, T., Harada, H. and Enomoto, M., Phase-Field Modeling with CALPHAD and CVM for Microstructural Evolution of Ni-base Superalloy. In: Proc. Int. Symp. Superalloys, Green, K. A., Pollock, T. M., Harada, H., Howson, T. E., Reed, R. C., Schirra, J. J. and Walston, S., Editors. TMS (The Minerals, Metals and Materials Society), Warrendale, PA, 2004.

33. Chu, Z., Chen, Z., Wang, Y. X., Lu, Y. L. and Li, Y. S., Microscopic Phase-Field Simulation of Competition Mechanism Between L12 and D022 Structure in Ni-Cr-Al alloy, J. Mater. Sci. Tech., 22 (3) (2006), 315-320.

34. Wang, Y., Banerjee, D., Su, C. C. and Khachaturyan, A. G., Field Kinetic Model and Computer Simulation of Precipitation of L12 Ordered Intermetallics from F.C.C. Solid Solution, Acta Mater., 46 (9) (1998), 2983-3001.

35. Wen, Y. H., Wang, B., Simmons, J. P. and Wang, Y., A Phase-Field Model for Heat Treatment Applications in Ni-based Alloys, Acta Mater., 54 (8) (2006), 2087-2099.

36. Grafe, U., Bottger, B., Tiaden, J. and Fries, S. G., Coupling of Multicomponent Thermodynamic Databases to a Phase-Field Model: Application to γ' Growth in a Ternary Ni-Al-Cr Model Superalloy, Scripta Mater., 42 (Pt. 1) (2000), 1179-1186.

37. Pareige, C., Soisson, F., Martin, G. and Blavette, D., Ordering and Phase Separation in Ni-Cr-Al: Monte Carlo Simulations vs Three-dimensional Atom probe, Acta Mater., 47 (6) (1999), 1889-1899.

38. Pareige-Schmuck, C., Soisson, F. and Blavette, D., Ordering and Phase Separation in Low Supersaturated Ni-Cr-Al alloys: 3D Atom Probe and Monte Carlo Simulation, Mater. Sci. Eng., A, 250 (1) (1998), 99-103.

Page 172: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

172

39. Mao, Z., Sudbrack, C. K., Yoon, K. E., Martin, G. and Seidman, D. N., The Mechanism of Morphogenesis in a Phase Separating Concentrated Multi-Component Alloy, Nature Mater., 6 (2006), 210-216.

40. Wagner, R., Kampmann, R. and Voorhees, P. W., Homogeneous Second-Phase Precipitation. In: Phase Transformations in Materials, Kostorz, G., Editor. Wiley-VCH, Weinheim, 2001.

41. Sims, C. T., Stoloff, N. S. and Hagel, W. C., Superalloys II High-Temperature Materials for Aerospace and Industrial Power. (John Wiley & Sons Inc., New York, 1987), 615.

42. Sudbrack, C. K., Yoon, K. E., Noebe, R. D. and Seidman, D. N., Temporal Evolution of the Nanostructure and Phase Compositions in a Model Ni-Al-Cr alloy, Acta Mater., 54 (12) (2006), 3199-3210.

43. Sudbrack, C. K., Decomposition Behavior in Model Nickel-Aluminum-Chromium-X Superalloys: Temporal Evolution and Compositional Pathways on a Nanoscale, Ph.D., Northwestern University, arc.nucapt.northwestern.edu, 2004

44. Sudbrack, C. K., Noebe, R. D. and Seidman, D. N., Temporal Evolution of Sub-Nanometer Compositional Profiles Across the γ/γ' Interface in a Model Ni-Al-Cr Superalloy. In: Proc. Int. Conf. Solid-Solid Phase Transformations, Howe, J. M., Editor. Vol. 2. 2005.

45. Sudbrack, C. K., Noebe, R. D. and Seidman, D. N., Direct Observations of Nucleation in a Nondilute Multicomponent Alloy, Phys. Rev. B, 73 (21) (2006), 212101/212101-212101/212104.

46. Sudbrack, C. K., Noebe, R. D. and Seidman, D. N., Compositional Pathways and Capillary Effects during Early-Stage Isothermal Precipitation in a Nondilute Ni-Al-Cr alloy, Acta Mater., 55 (2007), 119-130.

47. Sudbrack, C. K., Yoon, K. E., Mao, Z., Noebe, R. D., Isheim, D. and Seidman, D. N., Temporal Evolution of Nanostructures in a Model Nickel-Base Superalloy: Experiments and Simulations, Proc. TMS Annual Meeting, (2003), 43-50.

48. Sudbrack, C. K., Yoon, K. E., Mao, Z., Noebe, R. D., Isheim, D. and Seidman, D. N., Temporal Evolution of Nanostructures in a Model Nickel-base Superalloy: Experiments and Simulations. In: Electron Microscopy: Its Role in Materials Science, the Mike Meshii Symposium, J.R. Weertman, M. E. Fine, K. T. Faber, W. King and Liaw, P., Editors. TMS (The Minerals, Metals & Materials Society), San Diego, CA, 2003.

49. Sudbrack, C. K., Yoon, K. E., Noebe, R. D. and Seidman, D. N., The Temporal Evolution of the Nanostructure of a Model Ni-Al-Cr Superalloy, TMS Lett., 1 (2) (2004), 25-26.

50. Yoon, K. E., Temporal Evolution of the Chemistry and Nanostructure of Multicomponent Model Nickel-Based Superalloys, Ph.D., Northwestern University, arc.nucapt.northwestern.edu, 2004

Page 173: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

173

51. Yoon, K. E., Sudbrack, C. K., Noebe, R. D. and Seidman, D. N., The Temporal Evolution of the Nanostructures of Model Ni-Al-Cr and Ni-Al-Cr-Re Superalloys, Z. Metall., 96 (5) (2005), 481-485.

52. Booth-Morrison, C., Weninger, J., Sudbrack, C. K., Mao, Z., Noebe, R. D. and Seidman, D. N., Effects of Solute Concentrations on Kinetic Pathways in Ni-Al-Cr alloys, Acta Mater., 56 (14) (2008), 3422-3438.

53. Booth-Morrison, C., Zhou, Y., Noebe, R. D. and Seidman, D. N., On the Nanometer Scale Phase Decomposition of a Low-Supersaturation Ni-Al-Cr Alloy, Accepted by Phil. Mag., (2008).

54. Zhou, Y., Booth-Morrison, C. and Seidman, D. N., On the Field-Evaporation Behavior of a Model Ni-Al-Cr Superalloy Studied by Picosecond Pulsed-Laser Atom Probe Tomography, Accepted by Microsc. Microanal., (2008).

55. Sudbrack, C. K., Isheim, D., Noebe, R. D., Jacobson, N. S. and Seidman, D. N., The Influence of Tungsten on the Chemical Composition of a Temporally Evolving Nanostructure of a Model Ni-Al-Cr Superalloy, Microsc. Microanal., 10 (3) (2004), 355-365.

56. Sudbrack, C. K., Ziebell, T. D., Noebe, R. D. and Seidman, D. N., Effects of a Tungsten Addition on the Morphological Evolution, Spatial Correlations, and Temporal Evolution of a Model Ni-Al-Cr Superalloy, Acta Mater., 56 (2008), 448-463.

57. Yoon, K. E., Noebe, R. D. and Seidman, D. N., The Role of Rhenium on the Temporal Evolution of the Nanostructure of a Model Ni-Al-Cr-Re Superalloy, TMS Lett., 1 (2) (2004), 27-28.

58. Yoon, K. E., Noebe, R. D. and Seidman, D. N., Effects of Rhenium Addition on the Temporal Evolution of the Nanostructure and Chemistry of a Model Ni–Cr–Al Superalloy. I: Experimental Observations, Acta Mater., 55 (2007), 1145-1157.

59. Yoon, K. E., Noebe, R. D. and Seidman, D. N., Effects of Rhenium Addition on the Temporal Evolution of the Nanostructure and Chemistry of a Model Ni–Cr–Al Superalloy. II: Analysis of the Coarsening Behavior, Acta Mater., 55 (2007), 1159-1169.

60. Booth-Morrison, C., Noebe, R. D. and Seidman, D. N., Effects of a Tantalum Addition on the Morphological and Compositional Evolution of a Model Ni-Al-Cr Superalloy. In: Proc. Int. Symp. Superalloys, Reed, R. C., Green, K. A., Caron, P., Gabb, T. P., Fahrmann, M. G., Huron, E. S. and Woodward, S. A., Editors. TMS (The Minerals, Metals & Materials Society), Warrendale, PA, 2008.

61. Booth-Morrison, C., Mao, Z., Noebe, R. D. and Seidman, D. N., Chromium and Tantalum Site Substitution Patterns in Ni3Al (L12) γ'-Precipitates Appl. Phys. Let., 93 (2008), 033103.

62. Booth-Morrison, C., Noebe, R. D. and Seidman, D. N., Effects of Tantalum on the Temporal Evolution of a Model Ni-Al-Cr Superalloy During Phase Decomposition, Acta Mater., 57 (3)

Page 174: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

174

(2008), 909-920.

63. Amouyal, Y., Mao, Z., Booth-Morrison, C. and Seidman, D. N., On the Interplay between Tungsten and Tantalum Atoms in Ni-based Superalloys: An Atom-Probe Tomographic and First-Principles Study, Appl. Phys. Let., 94 (4) (2009), 041917/041911-041917/041913.

64. Zhou, Y., Booth-Morrison, C. and Seidman, D. N., On the Field Evaporation Behavior of a Model Ni-Al-Cr Superalloy Studied by Picosecond Pulsed-Laser Atom-Probe Tomography, Microsc. Microanal., 14 (6) (2008), 571-580.

65. Hobbs, R. A., Zhang, L., Rae, C. M. F. and Tin, S., Mechanisms of Topologically Close-Packed Phase Suppression in an Experimental Ruthenium-Bearing Single-Crystal Nickel-base Superalloy at 1100 oC, Metall. Mater. Trans. A, 39A (5) (2008), 1014-1025.

66. Sato, A., Harada, H., Yokokawa, T., Murakumo, T., Koizumi, Y., Kobayashi, T. and Imai, H., The Effects of Ruthenium on the Phase Stability of Fourth Generation Ni-base Single Crystal Superalloys, Scripta Mater., 54 (9) (2006), 1679-1684.

67. Yeh, A. C. and Tin, S., Effects of Ru and Re Additions on the High Temperature Flow Stresses of Ni-base Single Crystal Superalloys, Scripta Mater., 52 (6) (2005), 519-524.

68. Volek, A., Pyczak, F., Singer, R. F. and Mughrabi, H., Partitioning of Re Between γ and γ' phase in Nickel-base Superalloys, Scripta Mater., 52 (2) (2004), 141-145.

69. Carroll, L. J., Feng, Q., Mansfield, J. F. and Pollock, T. M., Elemental Partitioning in Ru-Containing Nickel-Base Single Crystal Superalloys, Mater. Sci. Eng., A, A457 (1-2) (2007), 292-299.

70. Walston, W. S., O'Hara, K. S., Ross, E. W., Pollock, T. M. and Murphy, W. H., René N6: Third Generation Single Crystal Superalloy. In: Proc. Int. Symp. Superalloys, Kissinger, R. D., Deye, D. J., Anton, D. L., Cetel, A. D., Nathal, M. V., T.M. Pollock, T. M. and Woodford, D. A., Editors. The Minerals, Metals and Materials Society (TMS), Warrendale, PA, 1996.

71. Murakami, H., Honma, T., Koizumi, Y. and Harada, H., Distribution of Platinum Group Metals in Ni-base Single-Crystal Superalloys. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., Green, K. A., McLean, M., Olson, S. L. and Schirra, J. J., Editors. The Minerals, Metals and Materials Society (TMS), Warrendale, PA, 2000.

72. Tin, S. and Pollock, T. M., Phase Instabilities and Carbon Additions in Single-Crystal Nickel-Base Superalloys, Mater. Sci. Eng., A, A348 (1-2) (2003), 111-121.

73. Rae, C. M. F., Karunate, M. S. A., Small, C. J., Broomfield, R. W., Jones, C. N. and Reed, R. C., Topologically Close Packed Phases in an Experimental Rhenium-Containing Single Crystal Superalloy. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., Green, K. A., McLean, M., Olson, S. L. and Schirra, J. J., Editors. TMS, Warrendale, PA, 2000.

Page 175: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

175

74. Giamei, A. F. and Anton, D. L., Rhenium Additions to a Nickel-Base Superalloy: Effects on Microstructure, Metall. Trans. A, 16A (11) (1985), 1997-2005.

75. Rae, C. M. F. and Reed, R. C., The Precipitation of Topologically Close-Packed Phases in Rhenium-Containing Superalloys, Acta Mater., 49 (19) (2001), 4113-4125.

76. Zheng, Y., Wang, X., Dong, J. and Han, Y., Effect of Ru Addition on Cast Nickel Base Superalloy with Low Content of Cr and High Content of W. In: Proc. Int. Symp. Superalloys, Pollock, T. M., Kissinger, R. D., Bowman, R. R., McLean, M., Olson, S. L. and Schirra, J. J., Editors. TMS, Warrendale, PA, 2000.

77. Dinsdale, A. T., SGTE Data for Pure Elements, CALPHAD, 15 (4) (1991), 317-425.

78. Forde, P. T., Tantalum in Superalloys, Adv. Mater. Processes, 149 (4) (1996), 39-40.

79. Jena, A. K. and Chaturvedi, M. C., The Role of Alloying Elements in the Design of Nickel-Base Superalloys, J. Mater. Sci., 19 (10) (1984), 3121-3139.

80. Rawlings, R. D. and Staton-Bevan, A. E., Alloying Behavior and Mechanical Properties of Polycrystalline Nickel-Aluminum (Ni3Al) (γ' phase) with Ternary Additions, J. Mater. Sci., 10 (3) (1975), 505-514.

81. MacKay, R. A., Nathal, M. V. and Pearson, D. D., Influence of Molybdenum on the Creep Properties of Nickel-Base Superalloy Single Crystals, Metall. Trans. A, 21A (2) (1990), 381-388.

82. Baldan, A., Relationship Between Rafted γ' Morphology and Creep Resistance of a Nickel-Based Superalloy, Proc. Int. Symp. Superalloys, (1992), 647-656.

83. Ricks, R. A., Porter, A. J. and Ecob, R. C., The Growth of γ' Precipitates in Nickel-base Superalloys, Acta Metall., 31 (1) (1983), 43-53.

84. Donachie, M. J. and Donachie, S. T., Superalloys: a Technical Guide, 2nd edition. (ASM International, Materials Park, OH, 2002), 439.

85. Flinn, P. A., Theory of Deformation in Superlattices, AIME Trans., 218 (1960), 145-154.

86. Kruml, T., Conforto, E., Lo Piccolo, B., Caillard, D. and Martin, J. L., Cores to Strength and Work-Hardening: A Study of Binary Ni3Al, Acta Mater., 50 (20) (2002), 5091-5101.

87. Reppich, B., Some New Aspects Concerning Particle Hardening Mechanisms in γ' Precipitating Nickel-base Alloys. I. Theoretical Concept, Acta Metall., 30 (1) (1982), 87-94.

88. Reppich, B., Schepp, P. and Wehner, G., Some New Aspects Concerning Particle Hardening Mechanisms in γ' Precipitating Nickel-base Alloys. II. Experiments, Acta Metall., 30 (1) (1982), 95-104.

Page 176: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

176

89. Nembach, E., Suzuki, K., Ichihara, M. and Takeuchi, S., HVEM In situ Observation of Dislocation Processes in the γ'-hardened Superalloy NIMONIC 105, Philos. Mag. A, 56 (4) (1987), 439-451.

90. Nembach, E., Suzuki, K., Ichihara, M. and Takeuchi, S., In situ Deformation of the γ' hardened Superalloy Nimonic PE16 in High-Voltage Electron Microscopes, Philos. Mag. A, 51 (4) (1985), 607-618.

91. Nembach, E., Particle Strengthening of Metals and Alloys. (John Wiley and Sons, New York, 1996), 320.

92. Pope, D. P. and Ezz, S. S., Mechanical Properties of Nickel-Aluminum (Ni3Al) and Nickel-base Alloys with High Volume Fraction of γ', Int. Met. Rev., 29 (3) (1984), 136-167.

93. Kear, B. H. and Wilsdorf, H. G. F., Dislocation Configurations in Plastically Deformed Polycrystalline Cu3Au Alloys, AIME Trans., 224 (1962), 382-386.

94. Mishima, Y., Ochiai, S., Hamao, N., Yodogawa, M. and Suzuki, T., Solid Solution Hardening of Nickel - Role of Transition Metal and B-Subgroup Solutes, Jpn. Inst. Met. Trans., 27 (9) (1986), 656-664.

95. Dennison, J. P., Llewellyn, R. J. and Wilshire, B., Creep and Fracture Behavior of some Dilute Nickel Alloys at 500 and 600 C, J. Inst. Metals Bull. Met. Rev., 94 (4) (1966), 130-134.

96. Gibbons, T. B. and Hopkins, B. E., Creep Behavior and Microstructure of Nickel-Chromium-base Alloys, Metal Sci., 18 (5) (1984), 273-280.

97. Gibbons, T. B. and Hopkins, B. E., High Temperature Ductility of Nickel-20% Chromium base Alloys, Metal Sci., 8 (7) (1974), 203-208.

98. Gibbons, T. B. and Hopkins, B. E., Influence of Grain Size and Certain Precipitate Parameters on the Creep Properties of Nickel-Chromium-base Alloys, Metal Sci., 5 (Nov.) (1971), 233-240.

99. Seidman, D. N., Three-Dimensional Atom-Probe Tomography: Advances and Applications, Annu. Rev. Mater. Res., 37 (2007), 127-158.

100. Seidman, D. N., Perspective: from Field-ion Microscopy of Single Atoms to Atom-Probe Tomography: a Journey, Rev. Sci. Instrum., 78 (3) (2007), 030901/030901-030901/030903.

101. Bajikar, S. S., Larson, D. J., Kelly, T. F. and Camus, P. P., Magnification and Mass Resolution in Local-Electrode Atom Probes., Ultramicroscopy, 65 (1-2) (1996), 119-129.

102. Kelly, T. F., Camus, P. P., Larson, D. J., Holzman, L. M. and Bajikar, S. S., On the Many Advantages of Local-Electrode Atom Probes. , Ultramicroscopy, 62 (1-2) (1996), 29-42.

Page 177: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

177

103. Kelly, T. F. and Larson, D. J., Local Electrode Atom Probes., Mater. Charact., 44 (1/2) (2000), 59-85.

104. Bunton, J. H., Olson, J. D., Lenz, D. R. and Kelly, T. F., Advances in Pulsed-Laser Atom Probe: Instruments and Specimen Design for Optimum Performance, Microsc. Microanal., 13 (2007), 418-427.

105. Hono, K., Atom Probe Microanalysis and Nanoscale Microstructures in Metallic Materials, Acta Mater., 47 (11) (1999), 3127-3145.

106. Karnesky, R. A., Isheim, D. and Seidman, D. N., Direct Measurement of Two-Dimensional and Three-Dimensional Interprecipitate Distance Distributions from Atom-Probe Tomographic Reconstructions, Appl. Phys. Let., 91 (1) (2007), 013111/013111-013111/013113.

107. Hellman, O. C. and Seidman, D. N., Measurement of the Gibbsian Interfacial Excess of Solute at an Interface of Arbitrary Geometry Using Three-Dimensional Atom Probe Microscopy, Mater. Sci. Eng., A, A327 (1) (2002), 24-28.

108. Krakauer, B. W. and Seidman, D. N., Subnanometer Scale Study of Segregation at Grain Boundary in an Fe(Si) Alloy, Acta Mater., 46 (1998), 6145-6161.

109. Seidman, D. N., Krakauer, B. W. and Udler, D., Atomic Scale Studies of Solute-Atom Segregation at Grain Boundaries: Experiments and Simulations, J. Phys. Chem. Solids, 55 (10) (1994), 1035-1057.

110. Yoon, K. E., Noebe, R. D., Hellman, O. C. and Seidman, D. N., Dependence of Interfacial Excess on the Threshold Value of the Isoconcentration Surface, Surf. Interface Anal., 36 (5/6) (2004), 594-597.

111. Yoon, K. E., Isheim, D., Noebe, R. D. and Seidman, D. N., Nanoscale Studies of the Chemistry of a René N6 Superalloy, Interface Science, 9 (3/4) (2002), 249-255.

112. Schmuck, C., Danoix, F., Caron, P., Hauet, A. and Blavette, D., Atomic Scale Investigation of Ordering and Precipitation Processes in a Model Ni-Cr-Al Alloy, Appl. Surf. Sci., 94-5 (1996), 273-279.

113. Mao, Z., Sinnott, S. B., Martin, G. and Seidman, D. N., Determination of Pair-Wise Interaction Energies and the Calculation of Ternary Phase Diagrams of Ni-Al-Cr Alloys by First-Principles Calculations, In preparation, (2009).

114. Sundman, B., Jansson, B. and Andersson, J. O., The Thermo-Calc databank system., CALPHAD: Computer Coupling of Phase Diagrams and Thermochemistry, 9 (2) (1985), 153-190.

115. Saunders, N., Phase Diagram Calculations for Ni-based Superalloys. In: Proc. Int. Symp. Superalloys, R.D. Kissinger, D.J. Deye, D.L. Anton, A.D. Cetel, M.V. Nathal, T.M. Pollock and

Page 178: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

178

Woodford, D. A., Editors. TMS (The Minerals, Metals & Materials Society), Warrendale, PA, 1996.

116. Dupin, N., Ansara, I. and Sundman, B., Thermodynamic Re-Assessment of the Ternary System Al-Cr-Ni., CALPHAD, 25 (2) (2001), 279-298.

117. Martin, G., The Theories of Unmixing Kinetics of Solid Solutions. In: Solid State Phase Transformation in Metals and Alloys. Les Éditions de Physique, Orsay, France, 1978.

118. Russell, K. C., Nucleation in Solids: The Induction and Steady State Effects, Adv. Colloid Interface Sci., 13 (3-4) (1980), 205-318.

119. Kashchiev, D., Nucleation: Basic Theory and Applications. (Elsevier Science, Oxford, 2000),

120. Brechet, Y. and Martin, G., Nucleation Problems in Metallurgy of the Solid State: Recent Developments and Open Questions, C.R. Phys., 7 (2006), 959-976.

121. Robson, J. D., Modelling the Overlap of Nucleation, Growth and Coarsening during Precipitation, Acta Mater., 52 (15) (2004), 4669-4676.

122. Stowell, M. J., Precipitate Nucleation: Does Capillarity Theory Work?, Mater. Sci. Tech., 18 (2) (2002), 139-144.

123. Aaronson, H. I. and LeGoues, F. K., An Assessment of Studies on Homogeneous Diffusional Nucleation Kinetics in Binary Metallic Alloys, Metall. Trans. A, 23A (7) (1992), 1915-1945.

124. Greenwood, G. W., The Growth of Dispersed Precipitates in Solutions, Acta Metall., 4 (1956), 243–248.

125. Lifshitz, I. M. and Slyozov, V. V., The Kinetics of Precipitation from Supersaturated Solid Solutions, J. Phys. Chem. Solids, 19 (1961), 35-50.

126. Wagner, C., Theorie der Alterung Niederschlagen durch Umlosen, Z. Electrochem, 65 (1961), 581-591.

127. Rowenhorst, D. J., Kuang, J. P., Thornton, K. and Voorhees, P. W., Three-Dimensional Analysis of Particle Coarsening in High Volume Fraction Solid-Liquid Mixtures, Acta Mater., 54 (8) (2006), 2027-2039.

128. Voorhees, P. W., Ostwald Ripening of Two-phase Mixtures, Annu. Rev. Mater. Sci., 22 (1992), 197-215.

129. Ratke, L. and Voorhees, P. W., Growth and Coarsening, Ripening in Material Processing. (Springer-Verlag, Berlin, 2002),

Page 179: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

179

130. Umantsev, A. and Olson, G. B., Ostwald Ripening in Multicomponent Alloys, Scripta Metall. Mater., 29 (8) (1993), 1135-1140.

131. Morral, J. E. and Purdy, G. R., Particle Coarsening in Binary and Multicomponent Alloys, Scripta Metall. Mater., 30 (7) (1994), 905-908.

132. Kuehmann, C. J. and Voorhees, P. W., Ostwald Ripening in Ternary Alloys, Metall. Mater. Trans. A, 27 (4) (1996), 937-943.

133. Blavette, D., Deconihout, B., Bostel, A., Sarrau, J. M., Bouet, M. and Menand, A., The Tomographic Atom-Probe - a Quantitative 3-Dimensional Nanoanalytical Instrument on an Atomic-Scale, Rev. Sci. Instrum., 64 (10) (1993), 2911-2919.

134. Cerezo, A., Godfrey, T. J., Sijbrandij, S. J., Smith, G. D. W. and Warren, P. J., Performance of an Energy-Compensated Three-Dimensional Atom Probe, Rev. Sci. Instrum., 69 (1) (1998), 49-58.

135. Hellman, O. C., Vandenbroucke, J. A., Blatz du Rivage, J. and Seidman, D. N., Application Software for Data Analysis for Three-Dimensional Atom Probe Microscopy, Mater. Sci. Eng., A, 327 (1) (2002), 29-33.

136. Hellman, O. C., Blatz du Rivage, J. and Seidman, D. N., Efficient Sampling for Three-dimensional Atom Probe Microscopy Data, Ultramicroscopy, 95 (2003), 199-205.

137. Hellman, O. C., Vandenbroucke, J. A., Rusing, J., Isheim, D. and Seidman, D. N., Analysis of Three-Dimensional Atom-Probe Data by the Proximity Histogram, Microsc. Microanal., 6 (5) (2000), 437-444.

138. Blavette, D., Cadel, E., Pareige, C., Deconihout, B. and Caron, P., Phase Transformation and Segregation to Lattice Defects in Ni-Base Superalloys, Microsc. Microanal., 13 (6) (2007), 464-483.

139. Sundman, B., Jansson, B. and Andersson, J. O., The Thermo-Calc Databank System., CALPHAD, 9 (2) (1985), 153-190.

140. Parratt, L. G., Probability and Experimental Errors in Science. (John Wiley, New York, 1966),

141. Wagner, R., Kampmann, R. and Voorhees, P. W., Homogeneous Second-Phase Precipitation. (Phase Transformations in Materials, ed. Kostorz, G. Wiley-VCH, Weinheim, 2001), p. 314.

142. Zener, C., Theory of growth of spherical precipitates from solid solution, J. Appl. Phys., 20 (1949), 950-953.

143. Ham, F. S., Diffusion-Limited Growth of Precipitate Particles, J. Appl. Phys., 30 (10)

Page 180: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

180

(1959), 1518-1525.

144. Ardell, A. J., Temporal Behavior of the Number Density of Particles During Ostwald Ripening, Mater. Sci. Eng., A, A238 (1) (1997), 108-120.

145. Ardell, A. J., An Application of Theory of Particle Coarsening - Gamma' Precipitate in Ni-Al Alloys, Acta Metall., 16 (4) (1968), 511-516.

146. Ardell, A. J., Further Applications of Theory of Particle Coarsening, Acta Metall., 15 (11) (1967), 1772-1775.

147. Marquis, E. A. and Seidman, D. N., Coarsening Kinetics of Nanoscale Al3Sc Precipitates in an Al-Mg-Sc Alloy, Acta Mater., 53 (15) (2005), 4259-4268.

148. Kresse, G. and Hafner, J., Ab Initio Molecular Dynamics of Liquid Metals, Phys. Rev. B, 47 (1) (1993), 558-561.

149. Kresse, G. and Furthmueller, J., Efficient Iterative Schemes for Ab Initio Total-Energy Calculations using a Plane-Wave Basis Set, Phys. Rev. B, 54 (16) (1996), 11169-11186.

150. Kresse, G. and Furthmuller, J., Efficiency of Ab-Initio Total Energy Calculations for Metals and Semiconductors using a Plane-Wave Basis Set, Comput. Mater. Sci., 6 (1) (1996), 15-50.

151. Kresse, G. and Hafner, J., Ab Initio Molecular-Dynamics Simulation of the Liquid-Metal-Amorphous-Semiconductor Transition in Germanium, Phys. Rev. B, 49 (20) (1994), 14251-14269.

152. Gleiter, H. and Hornbogen, E., The Formation of Coherent Ordered Precipitates in Ni-Cr-Al Alloy, Z. Metall., 58 (3) (1967), 157-163.

153. Lupis, C. H. P., Chemical Thermodynamics of Materials. (North-Holland, New York, 1983), 581.

154. Christian, J. W., Theory of Transformations in Metals and Alloys, Part 1. (2nd ed. In. Pergamon Press, Oxford, U. K., 1975),

155. Prikhodko, S. V., Carnes, J. D., Isaak, D. G. and Ardell, A. J., Elastic constants of a Ni-12.69 at.% Al alloy from 295 to 1300 K, Scripta Mater., 38 (1) (1997), 67-72.

156. Stassis, C., Kayser, F. X., Loong, C. K. and Arch, D., Lattice-Dynamics of Ni3Al, Phys. Rev. B, 24 (6) (1981), 3048-3053.

157. Eshelby, J. D., The Determination of the Elastic Field of an Ellipsoidal Inclusion and Related Problems, Proc. Roy. Soc. A 241 (1957), 376–396.

158. Karnesky, R. A., Martin, G. and Seidman, D. N., Criteria for Consistent Steady-State

Page 181: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

181

Coarsening, In preparation, (2009).

159. Nembach, E., Particle Strengthening of Metals and Alloys. (John Wiley & Sons, Inc., New York, 1997),

160. Minamino, Y., Jung, S. B., Yamane, T. and Hirao, K., Diffusion of Cobalt, Chromium, and Titanium in Nickel Aluminide (Ni3Al), Metall. Trans. A, 23A (10) (1992), 2783-2790.

161. Kellogg, G. L. and Tsong, T. T., Pulsed-Laser Atom-Probe Field-Ion Microscopy, J. Appl. Phys., 51 (2) (1980), 1184-1193.

162. Kellogg, G. L., Pulsed-Laser Atom Probe Mass Spectroscopy, J. Phys., E, 20 (2) (1987), 125-136.

163. Tsong, T. T., Field Ion Image Formation, Surf. Sci., 70 (1978), 211-233.

164. Smith, G. D. W., Grovenor, C. R. M., Delargy, K. M., Godfrey, T. J. and McCabe, A. R., Direct Comparison of Performance of Atom Probe in Pulsed-Laser and Voltage-Pulsed Modes, IFES Proc., (1982), 283-289.

165. Kelly, T. F. and Miller, M. K., Invited Review Article: Atom Probe Tomography, Rev. Sci. Instrum., 78 (3) (2007), 031101/031101-031101/031120.

166. Cerezo, A., Clifton, P. H., Galtrey, M. J., Humphreys, C. J., Kelly, T. F., Larson, D. J., Lozano-Perez, S., Marquis, E. A., Oliver, R. A., Shab, G., Thompson, K., Zandbergen, M. and Alvis, R. L., Atom Probe Tomography Today, Mater. Today, 10 (12) (2007), 36-42.

167. Larson, D. J., Petford-Long, A. K., Ma, Y. Q. and Cerezo, A., Information Storage Materials: Nanoscale Characterisation by Three-Dimensional Atom Probe Analysis, Acta Mater., 52 (10) (2004), 2847-2862.

168. Gault, B., Mottay, E., Courjaud, A., Vurpillot, F., Bostel, A., Menand, A. and Deconihout, B., Ultrafast Laser Assisted Field Evaporation and Atom Probe Tomography Applications, J. Phys.: Conf. Ser., 59 (2007), 132-135.

169. Kelly, T. F., Larson, D. J., Thompson, K., Alvis, R. L., Bunton, J. H., Olson, J. D. and Gorman, B. P., Atom Probe Tomography of Electronic Materials, Annu. Rev. Mater. Res., 37 (2007), 681-727.

170. Müller, E. W., Field Desorption, Phys. Rev., 102 (1956), 618-624.

171. Gomer, R., Field Desorption, J. Chem. Phys., 31 (1959), 341-345.

172. Gomer, R. and Swanson, L. W., Theory of Field Desorption, J. Chem. Phys., 38 (1963), 1613-1629.

Page 182: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

182

173. Gomer, R., Field Emission and Field Ionization. (Harvard University Press, Cambridge, Mass., 1961), 195 pp.

174. Brandon, D. G., Field Evaporation, Philos. Mag., 14 (130) (1966), 803-820.

175. Zemansky, M. W., Heat and Thermodynamics. 4th ed, pp. 326-327. (McGraw-Hill Book Company, New York, 1957), 490.

176. Miller, M. K. and Smith, G. D. W., Atom Probe Microanalysis: Principles and Applications to Materials Problems. In: Proc. MRS Symp. MRS, Pittsburgh, PA, 1989.

177. Tsong, T. T., Atom-Probe Field Ion Microscopy. (Cambridge University Press Cambridge, UK, 1990), 397.

178. Brandon, D. G., The Structure of Field-Evaporated Surfaces, Surf. Sci., 3 (1964), 1-18.

179. Brandon, D. G., The Analysis of Field Evaporation Data from Field-Ion Microscope Experiments, Br. J. Appl. Phys., 16 (5) (1965), 683-688.

180. Tsong, T. T. and Müller, E. W., Field Evaporation Rates of Tungsten, Phys. Stat. Sol. A, 1 (3) (1970), 513-533.

181. Ernst, N., Experimental Investigation on Field Evaporation of Singly and Doubly Charged Rhodium, Surf. Sci., 87 (2) (1979), 469-482.

182. Kellogg, G. L., Determining the Field Emitter Temperature during Laser Irradiation in the Pulsed Laser Atom Probe, J. Appl. Phys., 52 (8) (1981), 5320-5328.

183. Kellogg, G. L., Measurement of the Charge State Distribution of Field Evaporated Ions: Evidence for Post-Ionization, Surf. Sci., 120 (2) (1982), 319-333.

184. Wada, M., On the Thermally Activated Field Evaporation of Surface Atoms, Surf. Sci., 145 (2-3) (1984), 451-465.

185. Southon, M. J., Field Emission and Field Ionization. In: Field-Ion Microscopy, Hren, J. J. and Ranganathan, S., Editors. Plenum Press, New York, 1968.

186. Cerezo, A., Clifton, P. H., Gomberg, A. and Smith, G. D. W., Aspects of the Performance of a Femtosecond Laser-Pulsed 3-Dimensional Atom Probe, Ultramicroscopy, 107 (9) (2007), 720-725.

187. Gault, B., Vella, A., Vurpillot, F., Menand, A., Blavette, D. and Deconihout, B., Optical and Thermal Processes Involved in Ultrafast Laser Pulse Interaction with a Field Emitter, Ultramicroscopy, 107 (9) (2007), 713-719.

188. Cerezo, A., Smith, G. D. W. and Clifton, P. H., Measurement of Temperature Rises in the

Page 183: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

183

Femtosecond Laser Pulsed Three-Dimensional Atom Probe, Appl. Phys. Let., 88 (15) (2006), 154103/154101-154103/154103.

189. Vurpillot, F., Gault, B., Vella, A., Bouet, M. and Deconihout, B., Estimation of the Cooling Times for a Metallic Tip under Laser Illumination, Appl. Phys. Let., 88 (9) (2006), 094105/094101-094105/094103.

190. Vella, A., Vurpillot, F., Gault, B., Menand, A. and Deconihout, B., Evidence of Field Evaporation Assisted by Nonlinear Optical Rectification Induced by Ultrafast laser, Phys. Rev. B, 73 (16) (2006), 165416/165411-165416/165417.

191. Grafstrom, S., Photoassisted Scanning Tunneling Microscopy, J. Appl. Phys., 91 (4) (2002), 1717-1753.

192. Gerstner, V., Thon, A. and Pfeiffer, W., Thermal Effects in Pulsed Laser Assisted Scanning Tunneling Microscopy, J. Appl. Phys., 87 (5) (2000), 2574-2580.

193. Miller, M. K., Cerezo, A., Hetherington, M. G. and Smith, G. D. W., Atom Probe Field Ion Microscopy. (Clarendon Press, Oxford, 1996), 509.

194. Yamamoto, M. and Seidman, D. N., Quantitative Compositional Analyses of Ordered Platinum-Cobalt (Pt3Co) by Atom-Probe Field-Ion Microscopy, Surf. Sci., 118 (3) (1982), 535-554.

195. Yamamoto, M. and Seidman, D. N., The Quantitative Compositional Analysis and Field-Evaporation Behavior of Ordered Nickel-Molybdenum (Ni4Mo) on an Atomic Plane-by-Plane Basis: an Atom-Probe Field-Ion Microscope Study, Surf. Sci., 129 (2-3) (1983), 281-300.

196. Herschitz, R. and Seidman, D. N., A Quantitative Atom-Probe Field-Ion Microscope Study of the Compositions of Dilute Cobalt-Niobium and Cobalt-Iron Alloys, Surf. Sci., 130 (1) (1983), 63-88.

197. Hall, T. M., Wagner, A. and Seidman, D. N., A Computer-Controlled Time-of-Flight Atom-Probe Field-Ion Microscope for the Study of Defects in Metals, J. Phys., E, 10 (9) (1977), 884-893.

198. Brandon, D. G., Field Evaporation of Dilute Alloys, Surf. Sci., 5 (1) (1966), 137-146.

199. Kellogg, G. L., Experimental Evidence for Multiple Post-Ionization of Field-Evaporated Ions, Phys. Rev. B, 24 (4) (1981), 1848-1851.

200. Müller, E. W. and Krishnaswamy, S. V., High Ionic Charges in Field-Evaporating 5d Transition Metals, Phys. Rev. Lett., 37 (15) (1976), 1011-1014.

201. Forbes, R. G., A Generalized Theory of Standard Field Ion Appearance Energies, Surf. Sci., 61 (1) (1976), 221-240.

Page 184: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

184

202. Nakamura, S. and Kuroda, T., Ionization States of Ions Field-Evaporated from the Refractory Metal Alloys, Jpn. J. Apply. Phys., 16 (8) (1977), 1499-1500.

203. Haydock, R. and Kingham, D. R., Post-Ionization of Field-Evaporated Ions, Phys. Rev. Lett., 44 (23) (1980), 1520-1523.

204. Kingham, D. R., The Post-Ionization of Field Evaporated Ions: A Theoretical Explanation of Multiple Charge States, Surf. Sci., 116 (2) (1982), 273-301.

205. Müller, E. W. and Tsong, T. T., Field Ion Microscopy. (American Elsevier Publishing Company, New York, 1969), 314.

206. Lee, M. J. G., Reifenberger, R., Robins, E. S. and Lindenmayr, H. G., Thermally Enhanced Field Emission from a Laser-Illuminated Tungsten Tip: Temperature Rise of Tip, J. Appl. Phys., 51 (9) (1980), 4996-5006.

207. Borgenstam, A., Engstrom, A., Hoglund, L. and Agren, J., DICTRA, a Tool for Simulation of Diffusional Transformations in Alloys, J. Phase Equilib., 21 (3) (2000), 269-280.

208. Campbell, C. E., Boettinger, W. J. and Kattner, U. R., Development of a Diffusion Mobility Database for Ni-base Superalloys, Acta Mater., 50 (4) (2002), 775-792.

209. Lund, A. C. and Voorhees, P. W., The Effects of Elastic Stress on Microstructural Development: The Three-Dimensional Microstructure of a γ-γ' Alloy, Acta Mater., 50 (10) (2002), 2585-2598.

210. Ardell, A. J., Nicholson, R. B. and Eshelby, J. D., The Modulated Structure of Aged Nickel-Aluminum Alloys, Acta Metall., 14 (3) (1966), 1295-1309.

211. Soisson, F. and Martin, G., Monte Carlo Simulations of the Decomposition of Metastable Solid Solutions: Transient and Steady-State Nucleation Kinetics, Phys. Rev. B, 62 (1) (2000), 203-214.

212. Doherty, R. D., Diffusive Phase Transformations in the Solid State in Physical Metallurgy. In: Physical Metallurgy, Cahn R.W., H. P., Editor. Vol. 2. North-Holland, Amsterdam, 1996.

213. Servi, I. S. and Turnbull, D., Thermodynamics and Kinetics of Precipitation in The Copper-Cobalt System, Acta Metall., 14 (2) (1966), 161-169.

214. Cahn, J. W., Free Energy of a Nonuniform System. II. Thermodynamic basis, J. Chem. Phys., 30 (1959), 1121-1124.

215. Cahn, J. W. and Hilliard, J. E., Free Energy of a Nonuniform System. I. Interfacial Free Energy, J. Chem. Phys., 28 (1958), 258-267.

216. Cahn, J. W. and Hilliard, J. E., Free Energy of a Nonuniform System. III. Nucleation in a

Page 185: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

185

Two-Component Incompressible Fluid, J. Chem. Phys., 31 (1959), 688-699.

217. Olson, G. B., Jou, H.-J., Jung, J., Sebastian, J. T., Misra, A., Locci, I. and Hull, D., Precipitation Model Validation in 3rd Generation Aeroturbine Disc Alloys. In: Proc. Int. Symp. Superalloys, Reed, R. C., Green, K. A., Caron, P., Gabb, T. P., Fahrmann, M. G., Huron, E. S. and Woodward, S. A., Editors. TMS (The Minerals, Metals & Materials Society), Warrendale, PA, 2008.

218. Zhou, S. H., Wang, Y., Zhu, J. Z., Wang, T., Chen, L. Q., MacKay, R. A. and Liu, Z.-K., Computational Tools for Designing Ni-base Superalloys. In: Proc. Int. Symp. Superalloys, Green, K. A., Pollock, T. M., Harada, H., Howson, T. E., Reed, R. C., Schirra, J. J. and Walston, S., Editors. TMS (The Minerals, Metals & Materials Society), Champion, PA, 2004.

219. Zhou, Y., Booth-Morrison, C., Mao, Z., Noebe, R. D. and Seidman, D. N., In preparation., (2009).

220. Kampmann, R. and Wagner, R., In: Decomposition of Alloys: the Early Stages, Haasen, P., Gerold, V., Wagner, R. and Ashby, M. F., Editors. Pergamon Press, Oxford, 1984.

221. Langer, J. S. and Schwartz, A. J., Kinetics of Nucleation in Near-Critical Fluids, Phys. Rev. A, 21 (3) (1980), 948-958.

222. Thornton, K., Agren, J. and Voorhees, P. W., Modeling the Evolution of Phase Boundaries in Solids at the Meso- and Nano-Scales, Acta Mater., 51 (19) (2003), 5675-5710.

223. Zhu, J. Z., Wang, T., Ardell, A. J., Zhou, S. H., Liu, Z. K. and Chen, L. Q., Three-Dimensional Phase-Field Simulations of Coarsening Kinetics of γ' Particles in Binary Ni-Al Alloys, Acta Mater., 52 (9) (2004), 2837-2845.

224. Srinivasan, R., Banerjee, R., Hwang, J. Y., Viswanathan, G. B., Tiley, J., Dimiduk, D. M. and Fraser, H. L., Atomic Scale Structure and Chemical Composition across Order-Disorder Interfaces, Phys. Rev. Lett., 102 (8) (2009), 086101.

225. Kumar, P., Role of Niobium and Tantalum in Superalloys. In: Advances in High Temperature Structural Materials and Protective Coatings, Koul, A. K., Parameswaran, V. R., Immarigeon, J.-P. and Wallace, W., Editors. National Research Council of Canada, Ontario (Canada), 1994.

226. Meng, Z., Sun, G., Li, M. and Xie, X., The Strengthening Effect of Tantalum in Nickel-Base Superalloys, Proc. Int. Symp. Superalloys, (1984), 563-572.

227. Nathal, M. V. and Ebert, L. J., The Influence of Cobalt, Tantalum, and Tungsten on the Elevated Temperature Mechanical Properties of Single Crystal Nickel-Base Superalloys, Metall. Trans. A, 16A (10) (1985), 1863-1870.

228. Brinegar, J. R., Mihalisin, J. R. and VanderSluis, J., The Effects of Tantalum for Niobium

Page 186: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

186

Substitutions in Alloy 713C, Proc. Int. Symp. Superalloys, (1984), 53-62.

229. Nathal, M. V. and Ebert, L. J., The Influence of Cobalt, Tantalum, and Tungsten on the Microstructure of Single Crystal Nickel-Base Superalloys, Metall. Trans. A, 16A (10) (1985), 1849-1862.

230. Janowski, G. M., Heckel, R. W. and Pletka, B. J., The Effects of Tantalum on the Microstructure of two Polycrystalline Nickel-Base Superalloys: B-1900 + Hf and MAR-M247, Metall. Trans. A, 17A (11) (1986), 1891-1905.

231. Kear, B. H. and Pope, D. P., Role of Refractory Elements in Strengthening of γ and γ' Pecipitation Hardened Nickel-Based Superalloys. In: Refractory Alloying elements in Superalloys, Tien, J. K. and Reichman, S., Editors. ASM, Metals Park, OH, 1984.

232. Krakauer, B. W., Hu, J. G., Kuo, S. M., Mallick, R. L., Seki, A., Seidman, D. N., Baker, J. P. and Loyd, R. J., A System for Systematically Preparing Atom-Probe Field-Ion-Microscope Specimens for the Study of Internal Interfaces, Rev. Sci. Instrum., 61 (11) (1990), 3390-3398.

233. Jou, H.-J., Voorhees, P. and Olson, G. B., Computer Simulations for the Prediction of Microstructure/Property Variation in Aeroturbine Disks. In: Proc. Int. Symp. Superalloys, Green, K. A., Pollock, T. M., Harada, H., Howson, T. E., Reed, R. C., Schirra, J. J. and Walston, S., Editors. TMS (The Minerals, Metals & Materials Society), Warrendale, PA, 2004.

234. Radis, R., Schaffer, M., Albu, M., Kothleitner, G., Pölt, P. and Kozeschnik, E., Evolution of Size and Morphology of γ' Precipitates in Udimet 720 Li During Continuous Cooling. In: Proc. Int. Symp. Superalloys, Reed, R. C., Green, K. A., Caron, P., Gabb, T. P., Fahrmann, M. G., Huron, E. S. and Woodward, S. A., Editors. TMS (The Minerals, Metals & Materials Society), Champion, PA, 2008.

235. Khachaturyan, A. G., Semenovskaya, S. V. and Morris, J. W., Theoretical Analysis of Strain-Induced Shape Changes in Cubic Precipitates during Coarsening, Acta Metall., 36 (6) (1988), 1563-1572.

236. William, C. J. and Voorhees, P. W., Elastic Interaction and Stability of Misfitting Cuboidal Inhomogeneities, J. Appl. Phys., 61 (4) (1987), 1610-1619.

237. Johnson, W. C. and Lee, J. K., Elastic Interaction Energy of Two Spherical Precipitates in an Anisotropic Matrix, Metall. Trans. A, 10A (8) (1979), 1141-1149.

238. Reed, R. C., Yeh, A. C., Tin, S., Babu, S. S. and Miller, M. K., Identification of the Partitioning Characteristics of Ruthenium in Single Crystal Superalloys Using Atom Probe Tomography, Scripta Mater., 51 (4) (2004), 327-331.

239. Jayaram, R., Hren, J. J. and Miller, M. K., APFIM/TEM Characterization of Solute Partitioning in a Model Nickel-Molybdenum-Aluminum-Tantalum Superalloy, Surf. Sci., 246 (1-3) (1991), 323-328.

Page 187: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

187

240. Blavette, D. and Bostel, A., Phase Composition and Long-Range Order in γ' Phase of a Nickel-base Single Crystal Superalloy CMSX2: An Atom Probe Study, Acta Metall., 32 (5) (1984), 811-816.

241. Jia, C. C., Ishida, K. and Nishizawa, T., Partition of Alloying Elements Between γ (A1), γ′ (L12), and β (B2) Phases in Ni-Al base Systems Metall. Mater. Trans. A, 25 (3) (1994), 473-485.

242. Miller, M. K., Jayaram, R., Lin, L. S. and Cetel, A. D., APFIM Characterization of Single-Crystal PWA 1480 Nickel-base Superalloy, Appl. Surf. Sci., 76-77 (1-4) (1994), 172-176.

243. More, K. L. and Miller, M. K., Microstructural Characterization of Udimet 720: A nickel-base Alloy, J. Phys. C6 49 Supp. 11, (1988), 391–396.

244. Ochiai, S., Oya, Y. and Suzuki, T., Alloying Behavior of Nickel-Aluminum, Nickel-Gallium, Nickel-Silicon, and Nickel-Germanium (Ni3Al, Ni3Ga, Ni3Si and Ni3Ge), Acta Metall., 32 (2) (1984), 289-298.

245. Karg, A. V., Fornwalt, D. E. and Kriege, O. H., Effect of Alloying Elements on the Intensity of Superlattice Reflections of Ni3Al, J. Inst. Metals, 99 (Oct.) (1971), 301-305.

246. Chiba, A., Shindo, D. and Hanada, S., Site Occupation Determination of Pd in Ni3Al by ALCHEMI, Acta Metall. Mater., 39 (1) (1991), 13-18.

247. Ruban, A. V. and Skriver, H. L., Calculated Site Substitution in Ternary γ'-Ni3Al: Temperature and Composition effects, Phys. Rev. B, 55 (2) (1997), 856-874.

248. Wang, S.-Y., Wang, C.-Y., Sun, J.-H., Duan, W.-H. and Zhao, D.-L., Energetics and Electronic Structure of Re and Ta in the γ'-phase of Ni-based Superalloys, Phys. Rev. B, 65 (3) (2002), 035101/035101-035101/035107.

249. Sluiter, M. H. F. and Kawazoe, Y., Site Preference of Ternary Additions in Ni3Al, Phys. Rev. B, 51 (7) (1995), 4062-4073.

250. Raju, S., Mohandas, E. and Raghunathan, V. S., A Study of Ternary Element Site Substitution in Ni3Al using Pseudopotential Orbital Radii Based Structure Maps, Scripta Mater., 34 (11) (1996), 1785-1790.

251. Jiang, C. and Gleeson, B., Site Preference of Transition Metal Elements in Ni3Al, Scripta Mater., 55 (5) (2006), 433-436.

252. Murakami, H., Koizumi, Y., Yokokawa, T., Yamabe-Mitarai, Y., Yamagata, T. and Harada, H., Atom Probe Microanalysis of Ir-bearing Ni-based Superalloys, Mater. Sci. Eng., A, A250 (1) (1998), 109-114.

253. Enomoto, M. and Harada, H., Analysis of γ'/γ Equilibrium in Nickel-Aluminum-X Alloys by the Cluster Variation Method with the Lennard-Jones Potential, Metall. Trans. A, 20A (4)

Page 188: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

188

(1989), 649-664.

254. Wu, Y. P., Tso, N. C., Sanchez, J. M. and Tien, J. K., Modeling of Ternary Site Occupation in L12 Ordered Intermetallics, Acta Metall., 37 (10) (1989), 2835-2840.

255. Machlin, E. S. and Shao, J., Quaternary Gamma-Prime (L12) Pseudobinary Properties as Revealed by the Ionicity Modified Pair Potential Model, Scripta Metall., 11 (10) (1977), 859-862.

256. Guard, R. W. and Westbrook, J. H., Alloying Behavior of Ni3Al (γ'-phase), AIME Trans., 215 (1959), 807-814.

257. Saito, Y. and Harada, H., The Monte Carlo Simulation of Ordering Kinetics in Ni-base Superalloys, Mater. Sci. Eng., A, A223 (1-2) (1997), 1-9.

258. Blavette, D., Deconihout, B., Chambreland, S. and Bostel, A., Three-Dimensional Imaging of Chemical Order with the Tomographic Atom-Probe, Ultramicroscopy, 70 (3) (1998), 115-124.

259. Lin, H., Seiberling, L. E., Lyman, P. F. and Pope, D. P., Lattice Location of Tantalum in Nickel-Aluminum (Ni3Al) by Ion Channeling and Nuclear Reaction Analysis, Proc. MRS Symp., 81 (High-Temp. Ordered Intermet. Alloys 2) (1987), 165-170.

260. Mekhrabov, A. O., Akdeniz, M. V. and Arer, M. M., Atomic Ordering Characteristics of Ni3Al Intermetallics with Substitutional Ternary Additions, Acta Mater., 45 (3) (1997), 1077-1083.

261. Arbuzov, M. P., Kachkovskaya, E. T. and Khaenko, B. V., X-ray Diffraction Study of the Structure of Ni3Al, Alloyed with Ti, Cr, and W, Fiz. Metall. Metallovd., 21 (6) (1966), 854-857.

262. Jiang, C., Sordelet, D. J. and Gleeson, B., Site Preference of Ternary Alloying Elements in Ni3Al: A First-Principles Study, Acta Mater., 54 (4) (2006), 1147-1154.

263. Perdew, J. P., In: Electronic Structure of Solids. Vol. 11. Akademie Verlag, Berlin, 1991.

264. Kresse, G. and Joubert, D., From Ultrasoft Pseudopotentials to the Projector Augmented-Wave Method, Phys. Rev. B, 59 (3) (1999), 1758.

265. Vogtenhuber, D., Houserova, J., Wolf, W., Podloucky, R., Pfeiler, W. and Pueschl, W., Ab Initio Calculation of Point Defect Energies and Atom Migration Profiles in Varying Surroundings in L12-Ordered Intermetallic Compounds, Proc. MRS Symp., 842 (Integrative and Interdisciplinary Aspects of Intermetallics) (2005), 541-546.

266. Fu, C. L. and Painter, G. S., Point Defects and the Binding Energies of Boron near Defect Sites in Ni3Al: A First-Principles Investigation, Acta Mater., 45 (2) (1997), 481-488.

267. Schweiger, H., Semenova, O., Wolf, W., Puschl, W., Pfeiler, W., Podloucky, R. and Ipser,

Page 189: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

189

H., Energetics of Point Defect Formation in Ni3Al, Scripta Mater., 46 (1) (2002), 37-41.

268. Mishin, Y., Atomistic Modeling of the γ and γ'-phases of the Ni-Al system, Acta Mater., 52 (6) (2004), 1451-1467.

269. Ardell, A. J. and Ozolins, V., Trans-Interface Diffusion-Controlled Coarsening, Nature Mater., 4 (4) (2005), 309-316.

270. Calderon, H. A., Voorhees, P. W., Murray, J. L. and Kostorz, G., Oswald Ripening in Concentrated Alloys, Acta Metall. Mater., 42 (3) (1994), 991-1000.

271. Marsh, C. and Chen, H., An In Situ X-Ray Diffraction Study of Precipitation from a Supersaturated Solid Solution: the γ' Precipitate in a Nickel-12.5 at.% Aluminum Alloy, Acta Metall. Mater., 38 (11) (1990), 2287-2298.

272. Wang, J. C., Osawa, M., Yokokawa, T., Harada, H. and Enomoto, M., Modeling the Microstructural Evolution of Ni-base Superalloys by Phase Field Method combined with CALPHAD and CVM, Comput. Mater. Sci., 39 (4) (2007), 871-879.

273. Amouyal, Y., Mao, Z. and Seidman, D. N., Segregation of Tungsten at γ '(L12)/γ (fcc) Interfaces in a Ni-based Superalloy: An Atom-Probe Tomographic and First-Principles Study, Appl. Phys. Let., 93 (20) (2008), 201905/201901-201905/201903.

Page 190: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

190

Appendix 1

Recalculation of the Ni-Al-Cr Interfacial Free Energy from the Data of

Schmuck et al.

Schmuck et al. [28] determined a value for the interfacial free energy, σγ/γ’, of 12.5 mJ m-

2 for a ternary Ni-5.2 Al-14.8 Cr at.% alloy aged at 873 K. This value is approximately one half

the value determined by Sudbrack et al. [46] for a similar alloy Ni-5.2 Al-14.2 Cr at.% aged at

873 K. The Schmuck et al. approach assumes: (i) the LSW model for binary alloys applies to

their ternary system; (ii) the rate constant for the average γ’-precipitate radius from LSW is

correct, which it is not [270]; and (iii) that the γ-matrix and γ’-precipitate phase compositions

achieved their equilibrium compositions after aging for 64 h. We recalculate the values of σγ/γ’

from the data by Schmuck et al. by applying a relationship for σγ/γ’ in a nonideal, nondilute

ternary alloy proposed by Marquis and Seidman [147], based on the results of Calderon et al.

[270] and the Kuehmann-Voorhees coarsening model [132], and as applied to Ni-5.2 Al-14.8 Cr

at.% by Sudbrack et al. [46]. Schmuck et al. found that the γ’-precipitates had compositions very

close to their equilibrium values after only 1 h, and had achieved their equilibrium compositions

after 64 h. Sudbrack et al. [46], however, found that the γ’-precipitate phase compositions of a

similar alloy, Ni-5.2 Al-14.2 Cr at.% aged at 873 K, continued to evolve temporally at an aging

time of 1024 h. As a result, we use the equilibrium compositions calculated by Thermo-Calc for

Ni-5.2 Al-14.8 Cr at.%, employing the Saunders database [115], Table A.1.1, in our estimates of

the value of σγ/γ’ for the data of Schmuck et al. A coarsening rate constant, KKV, of (2.30 ± 0.13) x

Page 191: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

191

10-31 m3 s-1 and values of γκ KVAl , and γκ KVCr , of 0.13 ± 0.01 at. fr. s1/3 and -0.08 ± 0.02 at. fr. s1/3,

respectively, are found from the data of Schmuck et al. These values should be compared to

values of KKV of (8.8 ± 3.3) x 10-31 m3 s-1 and of 0.19 ± 0.02 at. fr. s1/3 and -0.14 ± 0.05 at. fr. s1/3,

for γκ KVAl , and γκ KVCr , respectively, for Ni-5.2 Al-14.2 Cr at.%. The value of σγ/γ’ calculated from

the data of Schmuck et al. using Equation 3.9 is 20-21 ± 5 mJ m-2, which is in good agreement

with values of 22-23 ± 7 mJ m-2 for Ni-5.2 Al-14.2 Cr at.% and 23-25 ± 6 mJ m-2 for Ni-7.5 Al-

8.5 Cr at.%.

Table A.1.1 Equilibrium γ’-precipitate and γ-matrix equilibrium concentrations, as determined by atom- probe tomography (APT) and Thermo-Calc for Ni-5.2 Al-14.8 Cr at.% aged at 873 K from the data of Schmuck et al. [28].

Equilibrium composition of γ'-precipitates Ni (at.%) Al (at.%) Cr (at.%)

Measured by APT at 64 h [28]: 74 ± 2 18 ± 1 7.6 ± 0.8

Extrapolated from APT data [28]: 74.2 ± 0.9 18.4 ± 0.8 7.4 ± 0.5

Calculated with Thermo-Calc and Saunders database [115]: 74.91 16.08 9.01

Calculated with Thermo-Calc and Dupin database [116]: 75.79 14.44 9.76

Equilibrium composition of γ-matrix Ni (at.%) Al (at.%) Cr (at.%)

Measured by APT at 64 h [28]: 80.2 ± 0.3 4 ± 0.2 15.8 ± 0.2

Extrapolated from APT data [28]: 80.3 ± 0.2 3.94 ± 0.08 15.8 ± 0.2

Calculated with Thermo-Calc and Saunders database [115] : 80.80 3.50 15.70

Calculated with Thermo-Calc and Dupin database [116]: 80.65 3.77 15.58

Page 192: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

192

Appendix 2

Temporal evolution of the phase compositions of Ni-Al-Cr alloys at 873 K

Table A.2.1. Temporal evolution of the phase compositions in Ni-7.5 Al-8.5 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter for this alloy is estimated to be 0.0027 ± 0.0004, and the equilibrium volume fraction is 16.4 ± 0.6 %.

Aging time (h) γ-matrix composition (at.%) γ’-precipitate composition (at.%)

Ni Al Cr Ni Al Cr

0.167 83.96 ± 0.05

7.37 ± 0.03

8.67 ± 0.03

72.14 ± 3.16

21.40 ± 3.01 6.46 ± 1.44

0.25 84.04 ± 0.05

7.09 ± 0.09

8.87 ± 0.10

73.07 ± 1.20

20.70 ± 1.09 6.23 ± 0.65

1 84.50 ± 0.04

6.63 ± 0.04

8.87 ± 0.04

73.86 ± 0.42

19.95 ± 0.38 6.19 ± 0.23

4 84.75 ± 0.04

6.20 ± 0.03

9.05 ± 0.03

74.79 ± 0.19

19.13 ± 0.17 6.08 ± 0.11

16 84.92 ± 0.06

5.88 ± 0.04

9.20 ± 0.04

75.38 ± 0.19

18.57 ± 0.17 6.05 ± 0.11

64 84.99 ± 0.05

5.76 ± 0.03

9.25 ± 0.04

75.73 ± 0.16

18.30 ± 0.16 5.97 ± 0.10

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

∞ 85.19 ± 0.08

5.42 ± 0.09

9.39 ± 0.09

76.33 ± 0.12

17.82 ± 0.15 5.85 ± 0.12

Page 193: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

193

Table A.2.2. Temporal evolution of the phase compositions in Ni-5.2 Al-14.2 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter misfit for this alloy is estimated to be 0.0006 ± 0.0004 and the equilibrium volume fraction is 15.7 ± 0.7 %.

Aging time (h) γ-matrix composition (at.%) γ’-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 ± 3.1 19.1 ± 2.7 9.7 ± 2.1

0.25 80.73 ± 0.09

5.07 ± 0.05

14.20 ± 0.08 72.6 ± 1.1 18.2 ± 0.9 9.2 ± 0.7

1 80.88 ± 0.10

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

75.48 ± 0.26

17.19 ± 0.23

7.33 ± 0.16

64 81.22 ± 0.07

3.45 ± 0.04

15.33 ± 0.07 75.7 ± 0.3

17.17 ± 0.28

7.16 ± 0.19

256 81.22 ± 0.07

3.30 ± 0.03

15.47 ± 0.07

75.95 ± 0.20

16.96 ± 0.17

7.08 ± 0.12

1024 81.16 ± 0.09

3.27 ± 0.04

15.57 ± 0.09 76.4 ±0.3

16.70 ±0.29

6.91 ± 0.20

∞ 81.26 ± 0.18

3.13 ± 0.08

15.61 ± 0.18 76.5 ± 0.5 16.7 ± 0.4

6.77 ± 0.30

Page 194: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

194

Table A.2.3. Temporal evolution of the phase compositions in Ni-6.5 Al-9.5 Cr at.% aged at 873 K, as measured by atom-probe tomography. The lattice parameter misfit for this alloy is estimated, from PrecipiCalc, to be 0.00284, and the equilibrium volume fraction from APT, is 9.6 ± 2.9 %.

Aging time (h) γ-matrix composition (at.%) γ’-precipitate composition (at.%)

Ni Al Cr Ni Al Cr

0.5 83.05 ± 0.03

7.13 ± 0.08

9.82 ± 0.07

70.06 ± 3.02

21.28 ± 4.89

8.66 ± 5.27

0.75 82.99 ± 0.03

7.22 ± 0.07

9.79 ± 0.07

71.25 ± 0.13

20.40 ± 0.21

8.35 ± 0.23

1 83.73 ± 0.01

6.46 ± 0.03

9.81 ± 0.03

71.45 ± 0.49

20.46 ± 0.81

8.08 ± 0.88

1.5 83.74 ± 0.03

6.38 ± 0.07

9.88 ± 0.06

72.28 ± 0.20

19.57 ± 0.33

8.14 ± 0.36

2 83.74 ± 0.02

6.38 ± 0.04

9.87 ± 0.04

72.81 ± 0.46

19.35 ± 0.79

7.84 ± 0.84

3.5 83.83 ± 0.04

6.27 ± 0.08

10.08 ± 0.08

74.06 ± 0.44

18.74 ± 0.81

7.19 ± 0.86

4 83.72 ± 0.01

6.29 ± 0.03

9.99 ± 0.03

74.53 ± 0.09

18.36 ± 0.16

7.11 ± 0.17

16 83.90 ± 0.01

5.90 ± 0.03

10.20 ± 0.03

75.09 ± 0.08

17.98 ± 0.14

6.94 ± 0.15

64 83.99 ± 0.02

5.80 ± 0.05

10.21 ± 0.05

75.55 ± 0.07

17.89 ± 0.13

6.56 ± 0.17

256 84.06 ± 0.02

5.71 ± 0.05

10.23 ± 0.05

75.82 ± 0.09

17.81 ± 0.17

6.37 ± 0.18

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

∞ 84.19 ± 0.02

5.49 ± 0.05

10.30 ± 0.05

76.40 ± 0.19

17.53 ± 0.33

6.06 ± 0.36

Page 195: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

195

Vita

Christopher Booth-Morrison

Education 2006 Arizona State University, Tempe, AZ

High Resolution Electron Microscopy Winter School (January)

2004 B. Eng., Metallurgical Engineering McGill University, Montreal, Canada

Journal Publications

1. C. Booth-Morrison, Y. Zhou, R. D. Noebe, and D. N. Seidman, “On the nanometer scale phase separation of a low-supersaturation Ni-Al-Cr alloy”. (To appear in Phil. Mag. 2009).

2. Y. Amouyal, Z. Mao, C. Booth-Morrison, and D. N. Seidman, “On the interplay between tungsten and tantalum in Ni-based superalloys: An atom-probe tomographic and first-principles study”. Applied Physics Letters. 2009;94;041917.

3. C. Booth-Morrison, R. D. Noebe, and D. N. Seidman, “Effects of tantalum on the temporal evolution of a model Ni-Al-Cr superalloy during phase decomposition”. Acta Materialia. 2009;57;909-920.

4. Y. Zhou, C. Booth-Morrison, and D. N. Seidman, “On the field-evaporation behavior of a model Ni-Al-Cr superalloy studied by picosecond pulsed-laser atom probe tomography” Microscopy & Microanalysis. 2008;14;571-580.

5. Y. Zhou, Z. Mao, C. Booth-Morrison, and D. N. Seidman, “The partitioning and site preference of rhenium or ruthenium in model Ni-based superalloys: An atom-probe tomographic and first-principles study”. Applied Physics Letters. 2008;93;171905.

6. C. Booth-Morrison, Z. Mao, R. D. Noebe, and D. N. Seidman, “Chromium and tantalum site substitution patterns in Ni3Al (L12) γ’-precipitates”. Applied Physics Letters. 2008;93;033103.

7. C. Booth-Morrison, J. Weninger, C. K. Sudbrack, Z. Mao, R. D. Noebe, and D. N. Seidman, “Effects of solute concentrations on the kinetic pathways in Ni-Al-Cr alloys”. Acta Materialia. 2008;56;3422-3438.

Conference Proceedings

1. C. Booth-Morrison, R. D. Noebe, and D. N. Seidman, “Effects of a Tantalum Addition on the Morphological and Compositional Evolution of a Model Ni-Al-Cr Superalloy”. Proceedings of 11th International Symposium on Superalloys. The Minerals, Metals and Materials Society, Warrendale, PA. (2008).

Page 196: NORTHWESTERN UNIVERSITY Nanoscale Studies of the ...nucapt.northwestern.edu/refbase/files/Booth-Morrison...• The past and present members of the Seidman group for all their help

196

Conference Presentations

1. C. Booth-Morrison, R. D. Noebe, and D. N. Seidman, “Effects of Tantalum on the Phase Decomposition of a Model Ni-Al-Cr Superalloy on a Nanoscale”. The Minerals, Metals and Materials Society 2009 Annual Meeting. San Francisco, CA. (2009).

2. C. Booth-Morrison, Y. Zhou, Z. Mao, R. D. Noebe, and D. N. Seidman, “The Temporal Evolution of the Nanostructures of a Low-Supersaturation Ni-Al-Cr Superalloy”. The Minerals, Metals and Materials Society 2009 Annual Meeting. San Francisco, CA. (2009).

3. D. N. Seidman, C. Booth-Morrison, Z. Mao, C.K. Sudbrack, “Effects of Solute Concentrations on the Kinetic Pathways in Ni-Al-Cr Alloys: Experiments and Simulations”. The Minerals, Metals and Materials Society 2009 Annual Meeting. San Francisco, CA. (2009).

4. C. Booth-Morrison, R. D. Noebe, and D. N. Seidman, “Effects of a Tantalum Addition on the Morphological and Compositional Evolution of a Model Ni-Al-Cr Superalloy”. 11th International Symposium on Superalloys, The Minerals, Metals and Materials Society, Champion, PA. (2008).

5. C. Booth-Morrison, R. D. Noebe, and D. N. Seidman, “Nanoscale Studies of the Early Stages of Phase Decomposition in Model Ni-Al-Cr Superalloys”. High Throughput Analysis of Multicomponent Multiphase Diffusion Data Conference, National Institute of Standards and Technologies, Gaithersburg, MD. (2007).

6. C. K. Sudbrack, J. A. Wenninger, C. Booth-Morrison, R. D. Noebe, D. N. Seidman, “A Comparison of the Early-Stages of γ’-Precipitation in Two Ni-Al-Cr Superalloys”. The Minerals, Metals and Materials Society 2006 Annual Meeting. San Antonio, TX. (2006).

7. P. Wanjara, C. Booth-Morrison, E. Hsu, Jahazi M., “Process Optimization for Linear Friction Welding of Ti6Al4V”, The 7th International Conference on Trends in Welding Research, Pine Mountain, GA. (2005).

Awards and Honors

2008-2009 B.J. Martin Terminal Year Fellowship (Northwestern University)

2006-2007 Materials Science and Engineering Teaching Assistant of the Year (Northwestern

University)

2005-2006 Northwestern University Graduate Fellowship

2004-2006 Fonds de recherche sur la nature et les technologies Fellowship

2004-2005 Walter P. Murphy Fellowship (Northwestern University)

2000-2004 McGill University Academic Achievement Scholarships

2001-2004 Canadian Mineral Industry Education Foundation Scholarship

2002-2003 ASM International Student Chapter Award


Recommended