+ All Categories
Home > Documents > AN INTEGRATED APPROACH FOR MICROSTRUCTURE SIMULATION ...

AN INTEGRATED APPROACH FOR MICROSTRUCTURE SIMULATION ...

Date post: 17-Mar-2022
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
185
The Pennsylvania State University The Graduate School Department of Materials Science and Engineering AN INTEGRATED APPROACH FOR MICROSTRUCTURE SIMULATION: APPLICATION TO NI-AL-MO ALLOYS A Thesis in Materials Science and Engineering by Tao Wang 2006 Tao Wang Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2006
Transcript

The Pennsylvania State University

The Graduate School

Department of Materials Science and Engineering

AN INTEGRATED APPROACH FOR MICROSTRUCTURE SIMULATION:

APPLICATION TO NI-AL-MO ALLOYS

A Thesis in

Materials Science and Engineering

by

Tao Wang

2006 Tao Wang

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

August 2006

The thesis of Tao Wang has been reviewed and approved* by the following:

Zi-Kui Liu Professor of Materials Science and Engineering Thesis Co-Advisor

Co-Chair of Committee

Long-Qing Chen Professor of Materials Science and Engineering Thesis Co-Advisor

Co-Chair of Committee

Padma Raghavan Professor of Computer Science and Engineering

Jorge O. Sofo Associate Professor of Physics

Associate Professor of Materials Science and Engineering

James P. Runt Professor of Materials Science and Engineering

Associate Head for Graduate Studies

*Signatures are on file in the Graduate School.

iii

Abstract

The properties and performance of a material are strongly dependent on its

microstructure. For example, the γ ’ precipitate coherently embedded in the γ matrix is

the primary strengthening phase in Ni-base superalloys, and its volume fraction,

morphology and size distribution largely determine the strength, fatigue and creep

properties of an alloy. In the present study, a multiscale computational approach was

proposed to predict the microstructure evolution in Ni-base superalloys. It integrated a

quantitative phase-field model with first-principles calculations as well as the CALPHAD

(CALculation of PHAse Diagram) technique. Fundamental materials property databases

such as lattice parameters and atomic mobility were developed.

A phenomenological model was developed to describe the lattice parameter in solid states

as a function of temperature and composition, and successfully applied to Ni-Al binary

system by evaluating the model parameters using experimental data. An integrated

computational approach was also proposed for evaluating the lattice misfit between the

matrix and precipitates by combining first-principles calculations, existing experimental

data and phenomenological modeling when the experimental data is limited. The lattice

parameters and the local lattice distortions around solute atoms in γ -Ni solutions were

studied using first-principles calculations. The solute atoms considered include Al, Co,

Cr, Hf, Mo, Nb, Re, Ru, Ta, Ti and W. The effects of the atomic size and the electronic

and magnetic interactions on lattice distortion have been discussed.

iv

Atomic mobility in disordered γ and ordered γ ’ phases was modeled for the Ni-Al-Mo

ternary system, and a kinetic database was developed. The diffusion of Al in γ ’ was

simulated, and the formation energies of vacancy in different sublattices were calculated

by first-principles approach, both of which indicate the anti-site diffusion mechanism

being dominant for diffusion of Al.

The phase-field model for binary Ni-base superalloys was extended to ternary systems

and integrated with the corresponding thermodynamic, kinetic and lattice parameter

databases. The microstructure evolutions and coarsening kinetics of γ ’ precipitates in

Ni-Al-Mo alloys were studied by two-dimensional phase-field simulations. The effects

of volume fraction of precipitates and Mo concentration have been analyzed.

v

Table of Contents

Notation viii

List of Figures xi

List of Tables xvii

Acknowledgments xix

Chapter 1. Introduction 1

Chapter 2. Modeling of Lattice Parameter 8

2.1 Background 8

2.2 Models 9

2.2.1 Pure Element 9

2.2.2 Binary System 12

2.2.3 Ordered Phase 12

2.3 Application to Ni-Al System 13

2.3.1 Pure Al and Ni 13

2.3.2 Binary Ni-Al System 16

2.3.3 Lattice Misfit between γ and γ ’ Phases 23

2.4 Summary 24

Chapter 3. First-principles Study of Lattice Distortion in γ 39

3.1 Background 39

3.2 First-principle Calculations 40

3.3 Lattice Distortions 41

3.4 Lattice Parameter Change 45

vi

3.5 Summary 49

Chapter 4. First-principles Calculations and Phenomenological

Modeling of Lattice Misfit in Ni-base Superalloys 66

4.1 Background 66

4.2 Methodology 67

4.2.1 Lattice Parameter of Pure Metals and

Ordered Compounds 67

4.2.2 Effect of Chemical Ordering 68

4.2.3 Lattice Misfit 70

4.3 Results and Discussions 71

4.3.1 Ni-Al Binary System 71

4.3.2 Ni-Al-Mo Ternary System 71

4.3.3 Phase-field Simulation of γ ’ Precipitate Morphology 72

4.4 Summary 74

Chapter 5. Modeling of Atomic Mobility in γ and γ ’ of Ni-Al-Mo 79

5.1 Background 79

5.2 Model 80

5.3 Modeling of Atomic Mobility in γ 82

5.3.1 Ni-Al System 82

5.3.2 Ni-Mo and Al-Mo Systems 83

5.4 Modeling of Atomic Mobility in γ ’ 85

5.3.1 Ni-Al System 85

5.3.2 Ni-Al-Mo System 93

vii

5.5 Summary 94

Chapter 6. Coarsening Kinetics of γ ’ Precipitates in Ni-Al-Mo System 111

6.1 Background 111

6.2 Simulation Details 111

6.2.1 Model 112

6.2.2 Conditions and Parameters for Simulations 115

6.3 Results and Discussion 117

6.3.1 Microstructure Evolution 117

6.3.2 Coarsening Kinetics 118

6.4 Summary 122

Chapter 7. Conclusions and Future Directions 131

7.1 Conclusions 131

7.2 Future Directions 134

Appendix A. Thermodynamic Descriptions for γ and γ ’ in Ni-Al-Mo System 135

Appendix B. Diffusion Mobility and Atomic Mobility 139

Appendix C. Energies of Ni and Al in Ni3Al 142

References 153

viii

Notation

k Boltzmann’s constant

R Gas constant

klδ Kronecher-delta function

NS Standard deviation

T Temperature

t Time

c Composition variable

ix Mole fraction of element i

siy Site fraction of i in sublattice s

η Order parameter

G Gibbs energy

iµ Chemical potential of element i

Lα Thermal expansion coefficient

a Lattice parameter

δ Lattice misfit

w Inverse width of the Morse potential

D Inverse depth of the Morse potential

nr Nearest neighbor distance

ix

Dθ Debye temperature, and Tx DD /θ=

ik Linear regression coefficient for element i

N Number of sites in a supercell

fV , φ Volume fraction of precipitates

iM Atomic mobility of element i

0iM Frequency factor

iQ Activation energy

i∆Φ Generalized activation energy

*iD Tracer diffusion coefficient of an element i

nkjD Chemical diffusion coefficient of k in the gradient of j and with n as the

reference specie

VC Probability of the vacancy

ω Vacancy jump frequency

αf Correlation factor for phase α

NiAlP Anti-site factor

E Total energy

VaE∆ Vacancy formation energy

J Flux

r Space vector

imM Diffusion mobility of i with respect to the concentration gradient of element m

x

jnL Kinetic coefficient for the relaxation of the order parameter j with respect to the

gradient of the order parameter n

F Total free energy of a microstructure

f Free energy density

ele Elastic energy density

jβ Gradient energy coefficient of order parameter jη

( )jg η Double-well potential

0w Double-well potential height

( )jh η Separation function

u Space vector representing the local displacement field

ijσ Stress

klε Strain

ijklC Elastic constant

R Average particle radius

K Coarsening rate constant

pV Molar volume of the precipitate phase

xi

List of Figures

Figure 1.1 An integrated four-stage multi-scale approach for multi-component materials

modeling, simulation and design. 7

Figure 2.1 Calculated linear thermal expansion coefficient of fcc Al in comparison with

experimental data from the literature. 27

Figure 2.2 Comparison of lattice parameter data for Al and the model calculation. The

dotted and solid lines represent the results from Equation 2.8

( [ ])500(1001.1)500(1068.1exp0708.4 2285 −×+−×= −− TTa and

3103.1 −×=NS Å) and Equation 2.9

( [ ]1)500(1001.1)500(1068.10708.4 2285 +−×+−××= −− TTa and

3103.1 −×=NS Å), respectively. 28

Figure 2.3 Calculated linear thermal expansion coefficient of fcc Ni (solid line

T-85 1052.01022.1 ×+×= −α and 16 K103.1 −−×=NS for K300>T ) in

comparison with experimental data from the literature. 29

Figure 2.4 Comparison of lattice parameter data for Ni and the model calculation (solid

line [ ]1)900(1026.0)900(1022.15560.3 2285 +−×+−××= −− TTa and

3106.3 −×=NS Å). 30

Figure 2.5 Order parameter vs. temperature curve for the Ni-25at% Al alloy. 31

Figure 2.6 Average composition in the γ phase as a function of the holding time during

measurement at 953K. The initial and final compositions (dotted lines) refer to

xii

the equilibrium compositions at the previous annealing temperature (973K)

and the measurement temperature (953K), respectively. 32

Figure 2.7 Room-temperature lattice parameter of the γ phase in the Ni-Al system. The

solid line represents the results of the model calculation ( 3104.6 −×=NS Å).

33

Figure 2.8 Temperature dependence of the lattice parameter of the γ phase in various

Ni-Al alloys. The solid lines represent the results of the model calculation.

34

Figure 2.9 Room-temperature lattice parameter of the γ ’ phase. The solid line represents

the results of the model calculation ( 3104.7 −×=NS Å). 35

Figure 2.10 Temperature dependence of the γ ’ lattice parameter. The solid lines

represent the results of the model calculation. 36

Figure 2.11 The comparison of the experimental relative thermal expansion of the γ ’

phase (25 at% Al) and the model calculation (solid line) ( %06.0=NS ).

37

Figure 2.12 The calculated misfit between the γ and γ ’ phases. The solid curve shows

the misfit under the equilibrium condition, and the dashed lines represent

those under the frozen composition assumption ( 4101.4 −×=NS ). 38

Figure 3.1 Atomic radius difference between solute and solve (Ni) atoms vs. Linear

Regression Coefficient. 55

Figure 3.2 Electronic charge density (in units of e/Å3) of (a) Nb, (b) Mo and (c) Ru

solutes in the (001) plane of the fcc Ni lattice in period 5. 56

xiii

Figure 3.3 Electronic charge density (in units of e/Å3) of (a) Nb and (b) Ta solutes in the

(001) plane of the fcc Ni lattice in group 5. 57

Figure 3.4 Lattice parameter changes in Ni (γ ) solid solutions with additions of Al and

W ( 0059.0=NS Å for Al and 0015.0=NS Å for W). 58

Figure 3.5 Lattice parameter changes in Ni (γ ) solid solutions with additions of Co and

Hf ( 0011.0=NS Å for Co and 0017.0=NS Å for Hf). 59

Figure 3.6 Lattice parameter changes in Ni (γ ) solid solutions with additions of Nb

( 0027.0=NS Å for Nb). 60

Figure 3.7 Lattice parameter changes in Ni (γ ) solid solutions with additions of Mo

( 0041.0=NS Å for Mo). 61

Figure 3.8 Lattice parameter changes in Ni (γ ) solid solutions with additions of Re and

Ta ( 0013.0=NS Å for Re and 0026.0=NS Å for Ta). 62

Figure 3.9 Lattice parameter changes in Ni (γ ) solid solutions with additions of Ru

( 0055.0=NS Å for Ru). 63

Figure 3.10 Lattice parameter changes in Ni (γ ) solid solutions with additions of Ti

( 0011.0=NS Å for Ti). 64

Figure 3.11 Lattice parameter changes in Ni (γ ) solid solutions with additions of Cr

( 0088.0=NS Å for CrAvg). 65

xiv

Figure 4.1 Lattice misfit between γ and γ ’ in the Ni-Al binary system. Curves present

calculated results, and symbols are reported experimental values from

literature. 76

Figure 4.2 Lattice misfit γ and γ ’ in Ni-Al-Mo ternary system. Curves present

calculated results, and symbols are reported values from literature. 77

Figure 4.3 Comparison of precipitate morphologies, obtained by experiments. (a) and

2D phase-field simulations (b, c) in a Ni-12.5 at.% Al-2.0 at.% Mo alloy aged

at 1048K for 67h: (b) 0065.0=δ ; (c) 0035.0=δ . 78

Figure 5.1 Chemical diffusivity in the γ phase for Ni-Al as a function of Al

composition. The symbols are experimental data, and the solid line are

calculated from the mobility database developed by Engstrom and Agren.

98

Figure 5.2 Compositional dependence of the chemical diffusivity of the γ phase in Ni-

Mo. The solid line is calculated from our database, and the dashed one from

Campbell’s database. The symbols are experimental data. 99

Figure 5.3 Chemical diffusivities of the γ phase in Ni-15 at%Mo and Ni-3 at%Mo as a

function of inverse temperature. The solid lines are calculated from our

database, and the dashed ones from Campbell’s database. The symbols are

experimental data. 100

Figure 5.4 Diffusion coefficient of Mo in Al as a function of inverse temperature. The

solid line is calculated from our database, and the dashed one from

Campbell’s database. The symbols are experimental data. 101

xv

Figure 5.5 Tracer diffusivities of Ni in the stoichiometric Ni3Al. The solid line is

calculated from the assessment I, and the symbols present experimental data.

102

Figure 5.6 Chemical diffusivities in the stoichiometric Ni3Al. The solid line is calculated

from the assessment I, and the symbols present experimental data. 103

Figure 5.7 Calculated chemical diffusivity (assessment I) compared with the

experimental data from Fujiwara and Horita. 104

Figure 5.8 Calculated tracer diffusivities of Al in Ni3Al (L12) ordered phases at 1473K.

105

Figure 5.9 Concentration of the anti-site Al in Ni3Al (L12) ordered phases at 1473K.

106

Figure 5.10 Tracer diffusivities of Ni in the stoichiometric Ni3Al. The solid line is

calculated from the assessment II, and the symbols present experimental data.

107

Figure 5.11 Chemical diffusivities in the stoichiometric Ni3Al. The solid line is

calculated from the assessment II, and the symbols present experimental data.

108

Figure 5.12 Calculated chemical diffusivity (assessment II) compared with the

experimental data from Fujiwara and Horita. 109

Figure 5.13 Diffusion coefficient of Mo in Ni3Al. The symbols are experimental data.

110

xvi

Figure 6.1 Isothermal section of Ni-Al-Mo ternary phase diagram at 1048K. Symbols

show the compositions of selected samples and dotted lines present the tie-

lines for those compositions. 124

Figure 6.2 Lattice misfit between γ and γ ’ in Ni-Al-Mo ternary system at 1048K.

Symbols show the values for A1, A2 and A3 alloys. 125

Figure 6.3 Microstructure evolution of the γ ’ precipitates in Ni-Al-Mo alloys at 1048K.

Figures in the bottom row are from experiments, and others from 2D phase-

field simulations. 126

Figure 6.4 Plot of the cube of average particle size vs annealing time at 1048K. 127

Figure 6.5 Coarsening rate constant vs Mo concentration at 1048K. Open symbols are

from experiments and solid ones from 2D phase-field simulations. 128

Figure 6.6 Comparison of various ( )φf functions from different theories. 129

Figure 6.7 Coarsening rate constant vs volume fraction of γ ’ precipitates in Ni-Al-Mo

system at 1048K. 130

Figure C.1 Interactions between Al and Ni in the same sublattice. The open symbols are

calculated with constrained relaxations and the solid ones are from

unconstrained relaxations. 151

Figure C.2 Total energy of the L12 structured Ni-Al alloys at 0K. 152

xvii

List of Tables

Table 2.1 The optimized parameters for the γ and γ ’ phases in Ni-Al system (in Å).

26

Table 3.1 Total energies and lattice parameters for Ni107X1 fcc solutions from first-

principles calculations. 50

Table 3.2 Atomic radii and electron structure of solute atoms vs. lattice parameter

change in fcc Ni. 51

Table 3.3 Local lattice distortion in fcc Ni (in pm). 52

Table 3.4 Linear regression coefficients of solute atoms in fcc Ni (in Å/at.%). 53

Table 3.5 Comparison of calculated and experimental lattice parameter in Ni-base

superalloys. 54

Table 4.1 Lattice parameters of ordered and disordered phases. 75

Table 4.2 Linear coefficients of solute or anti-site elements in the γ and γ ’ phases (in

Å/at.%). 75

Table 5.1 Assessed mobility parameters of the disordered fcc phase in the Ni-Mo and Al-

Mo systems, (J/mole). 96

Table 5.2 Assessed mobility parameters of the ordered L12 phase in the Al-Ni system,

(J/mole). 96

Table 5.3 Formation energy of vacancy in the Ni3Al ordered phase. 97

Table 5.4 Model parameters for chemical ordering of L12 in Ni-Al-Mo, (J/mole). 97

Table 6.1 Some parameters for phase-field simulations ( KT 1048= ). 123

xviii

Table A.1 Thermodynamic properties for γ and γ ’ in Ni-Al-Mo system (in SI units)

137

Table C.1 Parameters for total energy description of the L12 structured Ni-Al alloys. (in

J/mole) 146

Table C.2 Structural descriptions of the SQS structure for L12 alloys. A and B are

randomly distributed species in one sublattice and C and D are atoms

occupying the other three sublattices. The atomic positions are given in direct

coordinates, and are for the ideal, unrelaxed structures. 147

xix

ACKNOWLEDGEMENTS

I wish to express my sincere gratitude to my advisors, Dr. Zi-Kui Liu and Dr. Long-Qing

Chen, for their constant advice, guidance and encouragement during my four-year Ph.D.

study at the Pennsylvania State University (PSU). I also wish to thank Dr. Jorge O. Sofo,

Dr. Padma Raghavan for their help and encouragement, and serving in my thesis

committee.

The assistance from Dr. Chris Wolverton of Ford Research Laboratory and Dr. Qiang Du

of the Department of Mathematics in my thesis research is gratefully acknowledged. I

also want to thank Dr. Michael G. Fahrmann of Special Metals Corp for providing the

original TEM micrographs.

I would also take the opportunity to thank all members in Dr. Zi-Kui Liu and Dr. Long-

Qing Chen’s research groups, as well as the faculty and staff of the Department of

Materials Science and Engineering, for their help and cooperation during my stay at PSU.

Finally, I would like to thank my parents, for their constant encouragement and

invaluable support without which this thesis could not have come true.

1

Chapter 1

Introduction

Since the microstructure (the size, shape, and spatial arrangement of the structural

features) plays a critical role in determining various properties of a material, e.g.

mechanical, electrical, magnetic and optical properties, the study of microstructures is a

very important portion in the field of materials science and engineering, and

microstructure controlling is vital process to obtain a material of desired properties. Used

in aircraft engines, Ni-base superalloys are demanded to be thermodynamically and

structurally stable at high temperatures and for a long period of time due to the high-

temperature operating environment of the aircraft engines. The microstructure of Ni-base

superalloys consists of γ ’ precipitates and a face-centered cubic (fcc) matrix γ . γ ’ has

an ordered fcc structure (L12) where one type of atoms prefer the face-centered sites and

the corner positions are occupied by another type of atoms. In Ni-base superalloys, the

γ ’ precipitate coherently embedded in the γ matrix is the primary strengthening phase,

and its volume fraction, morphology and size distribution strongly affects the mechanical

properties of the materials (e.g. strength, fatigue and creep). Thus the control of the

γ + γ ’ two-phase microstructure and its high-temperature stability becomes the key for

designing of Ni-base superalloys.

2

As a result of those fine γ ’ particles, a large surface area is presented, and thus the total

free energy can be decreased by smaller particles dissolving and mass transporting from

those smaller particles to larger ones. Such a process is called Ostwald ripening or

coarsening, which occurs at later stages of phase transformations. Due to the high service

temperature of Ni-base superalloys, coarsening is an important issue during service.

Thus a knowledge of the coarsening process of the γ ’ precipitates is essential for design

and application of Ni-base superalloys.

Since there are many factors, e.g. lattice misfit between the precipitate and matrix,

coherency of interface and diffusivities of various elements, affecting the microstructure

evolution and coarsening process, the theoretical predictions with the aid of computer is

the only practicable way for a systematic investigation on a complex system.

In current laboratory and industrial practice, the traditional trial-and-error method is still

a dominant technique to optimizing the alloy chemistry and processing conditions for

achieving desirable microstructures and mechanical properties of Ni-based superalloys.

This often highly expensive and empirical approach becomes more and more insufficient

to meet today’s increasingly demanding applications, and has been complemented by

various computational tools in last few decades. Due to the advances in computer and

information technology, computational materials science has been quickly developed,

and the required experiments can be dramatically reduced with the help of it. Some very

useful computational tools for materials scientists are atomic-scale first-principles

3

approach, CALPHAD (CALculation of PHAse Diagram) technique and phase-field

simulation.

Based on the density functional theory (DFT) [1], the first-principles approach can

calculate thermodynamic properties (e.g. enthalpy and entropy of formation), kinetic data

(e.g. diffusivity) and crystallography information (e.g. lattice parameter and interfacial

energy) using only the atomic numbers and crystal structure information as the inputs.

By considering the configurational and vibrational contributions, various properties at

finite temperatures can also be predicted. Although the first-principles approach has been

improved and extensively used in many fields in recent years, it is still unpractical to

determine the total energy for multi-component systems directly from the first-principles

approach, at least the accuracy is not comparable to the experimental measurements.

Another important tool in computational materials science, the CALPHAD approach [2]

is very powerful in predicting phase equilibria and phase transformations of multi-

component alloys. The CALPHAD approach allows one to determine the

thermodynamic descriptions for various phases in the multi-component system, and then

constructs the thermodynamic database. Both experiments and first-principles

calculations on simple low-order systems can provide data for deriving model parameters,

and various thermodynamic properties can be calculated from the thermodynamic

database. A similar strategy can be adopted for developing kinetic databases and

databases for lattice parameters, elastic constants and interfacial energies as a function of

composition and temperature [3].

4

The phase-field approach is one of the most powerful methods for modeling various

microstructure-related processes [4]. It describes a microstructure by a set of conserved

or non-conserved field variables, and those variables change smoothly from one

phase/domain to another across the interfacial region. Unlike other microstructure

models, the phase-field approach does not explicitly track the positions of interfaces, and

hence the temporal evolution of arbitrary microstructure can be predicted without any a

priori assumptions about their evolution path [3].

Recently, Liu and co-workers proposed a prototype to integrate various computational

tools for multi-component materials simulation and design [3]. The framework is shown

in Figure 1.1, and involves four major computational steps:

• Atomic-scale first-principles calculations for predicting thermodynamic properties,

kinetic data and crystallography information;

• CALPHAD (CALculation of PHAse Diagram) approach for developing

thermodynamic, kinetic and crystallographic databases;

• Phase-field modeling for simulating the evolution of microstructures;

• Finite element analysis for deriving mechanical properties from the simulated

microstructures.

As part of the framework, the present thesis is concerned with the first three steps, from

first-principles calculations to phase-field simulations. The main goal of this

investigation is to predict the microstructure evolution and coarsening kinetics of γ ’

precipitates in Ni-base superalloys by integrating these computational tools. As one of

5

the commonly used alloy elements in Ni-base superalloys, Mo can not only adjust the

lattice misfit between γ and γ ’ [5], but also control the coarsening rate of γ ’

precipitates [6]. Therefore, we focus our current study on Ni-Al-Mo superalloys.

Three objectives are planned for this work. The first objective is to develop a lattice

parameter database for the γ and γ ’ phases in the Ni-base superalloys, and then the

lattice misfit can be calculated from the database and employed to phase-field simulations.

In Chapter 2, we propose a phenomenological model to describe the lattice parameters of

substitutional solid solution as a function of temperature and composition. It is applied to

the γ and γ ’ phases in Ni-Al binary alloys and the model parameters are evaluated using

the CALPHAD approach from available experimental data in the literature. Since the

experimental results are usually very limited in multi-component systems, an integrated

computational approach is then developed in Chapter 3 for evaluating the lattice misfit

between γ and γ ’ in Ni-base superalloys by combining first-principles calculations,

existing experimental data and phenomenological modeling. In particular, the lattice

misfits in Ni-Al and Ni-Al-Mo alloys are studied by this approach. With the help of first-

principles calculations, we also study the effects of various alloy elements on the lattice

parameter and the local lattice distortion around the solute atom in binary fcc-Ni

solutions, which is shown in Chapter 4. The solute atoms considered include Al, Co, Cr,

Hf, Mo, Nb, Re, Ru, Ta, Ti and W. The contribution from the atomic size difference, the

electronic interactions, and the magnetic spin relations are discussed. Based on the

results from first-principles calculations, the linear composition coefficients of fcc Ni

6

lattice parameter for different solutes are determined, and the lattice parameters of multi-

component Ni-base superalloys as a function of solute composition are predicted.

The second objective is to construct an atomic mobility database for Ni-Al-Mo

superalloys. In particular, diffusion in disordered γ and ordered γ ’ phases is modeled in

Chapter 5 using phenomenological models provided by Andersson and Agren [7] and

Helander and Agren [8]. Diffusion data in various constituent binary systems are

collected from the literature and assessed to establish the kinetic database, and previous

modeling works for γ are reviewed and revised. The diffusion mechanisms in the

ordered γ ’ phase are discussed based on the results, and used to refine atomic mobility

modeling.

Finally, for the third objective, the mic rostructure evolutions and coarsening kinetics of

γ ’ precipitates in Ni-Al-Mo ternary alloys will be predicted by phase-field simulations in

Chapter 6. All simulations are linked with thermodynamic, kinetic and lattice parameter

databases to provide predictions in terms of experimental time and length scales. Both

the effects from volume fraction of precipitates and compositions are studied.

7

Figure 1.1 An integrated four-stage multi-scale approach for multi-component materials

modeling, simulation and design. Redrawn from [3]

First-principles calculations and experiments

Thermodynamic data of unary binary and ternary systems

Kinetic data of unary, binary and ternary systems

Interfacial energies, lattice parameters, elastic constants

CALPHAD approach to data optimization

Thermodynamic database for multi-component systems

Kinetic database for multi-component systems

Databases for lattice parameters, elastic constants, and interfacial energies

A multi-component phase-field model

Elastic constants and plasticity of phases

Simulated microstructures in 2 and 3 dimensions

OOF: Object-oriented finite element analysis of material microstructure

Mechanical response of simulated microstructures

8

Chapter 2

Modeling of Lattice Parameter

2.1 Background

The lattice parameter and thermal expansion are two important material properties that

are strongly correlated to many thermophysical properties. Because of their importance

in both theoretical study and practical applications, a large number of studies have been

carried out on this subject from many different points of view, using theoretical,

experimental, and empirical approaches. The composition dependency of lattice

parameters was modeled in various ways, such as elasticity theory, various potential

approaches, and first-principle calculation, but none of them is very successful [9],

neither simple nor accurate enough. The most widely used prediction of the lattice

parameters across a solid solution was the linear relationship proposed by Vegard [10].

However, the investigations on metallic systems always show some deviations from

Vegard’s law, because Vergard’s law is only valid when the electronic environment of

both atoms is undisturbed by the formation of the solid solution, but in reality electrons in

states just below the Fermi level can also participate in metallic bonding [11]. Due to the

limited and scattered experimental data, the temperature effect on the lattice parameter is

often overlooked, especially for multi-component alloys. In many cases, such an effect is

assumed to be small enough to be neglected or approximated by some arbitrary

polynomials (the linear relationship is often suggested). However, such an assumption is

9

seldom supported by the experimental results except in a very narrow temperature region.

In this chapter, a simple phenomenological model is developed to describe the lattice

parameters of solid solution phases as a function of composition and temperature. In

Section 2.2, the temperature effect on the linear expansion coefficient ( Lα ) of the pure

element is considered first, and the lattice parameters of pure elements are then calculated

from the thermal expansion. The contribution from substitutional solute is treated using

an approach similar to that used in the Gibbs energy modeling [12]. At the end of the

Section 2.2, the modeling of chemical ordering effect is discussed in relation to the

sublattice model. In Section 2.3, this model is applied to the Ni-Al system, the most

important constituent binary system in Ni-based superalloys, to calculate the difference

between the lattice parameters of precipitate γ ’ (L12) and matrix γ (fcc_A1), which

plays a very important role in the microstructure evolution and properties of Ni-based

superalloys [13-16]. We evaluate the model parameters describing the lattice parameters

of the γ ’ and γ phases, and then calculate the lattice misfit between these phases.

2.2 Models

2.2.1 Pure Element

Based on quantum physics, Ruffa [17] proposed the following equation to describe the

thermal expansion coefficients

10

)(2

33

DDn

L xgT

Dwrk

=

θα (2.1)

where T is the temperature, k is Boltzmann’s constant, w and D are the inverse width

and depth, respectively, of the Morse potential, nr the nearest neighbor distance, Dθ the

Debye temperature, and Tx DD /θ= . The integral )( Dxg is given by

∫ −=

Dx

x

x

D dxe

exxg

0 2

4

)1()( (2.2)

Since the Equation 2.1 only considers the first order in the frequency, it is usually

accurate only up to about Dθ7.0 , but gives the dominant contribution over the entire

temperature range. To extend this formula to higher temperatures, a correction term was

added [17].

+

= )(

2)(

23

1

3

DDDn

L xgD

kTxg

TDwr

α (2.3)

where

∫ −+

=Dx

x

xx

D dxe

eexxg

0 3

5

1)1(

)1()( (2.4)

For low temperatures ( DT θ<< ), one can obtain

34

52

DnL

TDwr

πα (2.5)

which implies Lα is proportional to the 3T for DT θ<< . When the temperature is high

( DT θ>> ),

11

+++

+≈TD

kT

Dk

Dk

Dwrk DDD

nL

110438

331

23 2θθθ

α (2.6)

Because the contribution of the T/1 term is not significant at high temperatures, Lα vary

almost linearly with temperature in the high temperature range, which will be adopted in

the present work, i.e., we can describe the thermal expansion coefficient by a linear

function of temperature as a first approximation.

BTAL +=α (2.7)

The parameters A and B can be defined by available information from the thermal

expansion experiments. Such a linear relationship was observed by several investigations

[18-20].

The lattice parameters a can be obtained from Equation 2.7 through integration of Lα

based on the definition dTda

aL

1=α , and the lattice parameter ( 0a ) at a given temperature

( 0T ) can be used to determine the integration constant

−+−= )(

2)(exp 2

02

00 TTB

TTAaa (2.8)

Lα is about 510− K-1 for metals, so the relative change in lattice parameter is very small

(about 1% with a change of 1000K in temperature), thus the lattice parameter can be

approximated by the following polynomial

+−+−= 1)(

2)( 2

02

00 TTB

TTAaa (2.9)

In the Section 2.3, we calculate the lattice parameters of Al by both Equations 2.8 and 2.9,

and the results are almost identical.

12

2.2.2 Binary System

Similar to the Gibbs energy modeling [12], we add an excess contribution to describe the

deviation of the lattice parameter from the Vegard’s law. Such a phenomenological model

can be written as

aaxa ex

iii += ∑ 0 (2.10)

where ix is the mole fraction of element i . ia0 denotes the lattice parameter of pure

element i defined by Equation 2.9. aex is the excess contribution expressed in the

Redlich-Kister polynomials [21]

∑∑ ∑> =

−=i ij

n

k

kjiji

kji

ex xxIxxa0

, )( (2.11)

The interaction parameter, jik I , , can be expressed as a function of temperature:

TBAI jik

jik

jik

,,, += (2.12)

2.2.3 Ordered Phase

The ordered phase and related disordered phase can be modeled by the sublattice model

[22, 23]. For a two-sublattice model qp BABA ),(),( , the lattice parameter a can be

expressed by the following equation:

∑∑∑∑∑∑ ∑

∑∑ ∑∑∑

> >>

>

++

+=

i ij k kllkji

IIl

IIk

Ij

Ii

i ij kjik

Ik

IIj

IIi

i ij kkji

IIk

Ij

Ii

i jji

IIj

Ii

IyyyyIyyy

Iyyyayya

,:,,:

:,:0

(2.13)

13

where Iiy and II

iy are the site fractions of i in the first and second sublattices. Similar to

Equation 2.12, the interaction parameters kjiI :, , jikI ,: and lkjiI ,:, can be expressed as

functions of temperature. jia :0 , the lattice parameter of the end member qp ji of the

sublattice model, can be written as:

≠+++

++

==

ijTDCaqp

qa

qpp

ijaa

jijiji

i

ji::

00

0

:0 (2.14)

where jiC : and jiD : are model parameters. The overall composition ix is connected with

the site fractions by IIi

Iii y

qpqy

qppx

++

+= . When II

iIii yyx == , the phase is

disordered, and Equation 2.13 is equivalent to Equation 2.10.

2.3 Application to Ni-Al System

In this section we will apply the above phenomenological model to describe the lattice

parameters of γ ’ and γ phases in the Ni-Al system.

2.3.1 Pure Al and Ni

Many reports on the measurement of thermal expansion coefficients and lattice

parameters for Al can be found in the literature. Touloukian et al. [24] referenced 71 sets

of data for Al in their review, and later, Wang and Reeber [25] cited 7 more in their report.

Their selection of experimental data are shown in Figure 2.1. Because the model is not

14

applicable to the low temperature range, only the experimental data measured above

300K are used to evaluate the parameters A and B in Equation 2.7. The results are

plotted as the solid line in Figure 2.1.

As shown in Figure 2.1, the coefficient of linear thermal expansion for Al shows a good

linear relationship with temperature from 300 to 800K (standard deviation

17 K107 −−×=NS , calculated as the square root of the sample variance of a set of values

[26]). The low temperature behavior deviates from the linearity because the model only

accurately describes the intermediate temperature behavior. When the temperature is

very high (close to the melting temperature mT ), the experimental results show a visible

deviation from the linear dependence, which is caused by the contribution from thermal

vacancies [25]. Based on the model parameters obtained from the above evaluation, the

lattice parameter of Al can be calculated by Equations 2.8 or 2.9. Figure 2.2 shows the

calculated lattice parameters compared with the experimental data, and it can be seen that

most experimental data can be well reproduced by the present model ( 3103.1 −×=NS Å).

As shown in Figure 2.2, when the temperature increases from the room temperature to the

melting temperature, the lattice parameter of Al only increases 2%. The dotted and solid

lines represent the results from Equations 2.8 and 2.9, respectively, and the differences

between them are less than 0.015%.

Many experimental studies on the thermal expansion of Ni were performed in a wide

temperature range, and more than 100 investigations before 1975 have been reviewed by

Touloukian et al. [24]. Using the dilatometry technique, Kollie [27] measured the

15

thermal expansions in temperature range from 300 to 1000K, while Mukherjee et al. [28]

investigated for low temperatures up to 300K. Yousuf et al. [29] studied the magnetic

effect on the lattice expansion of Ni by high-temperature X-ray diffractometry, and

reported the lattice parameters and the thermal expansion coefficients. The data [27-33]

are plotted in Figure 2.3. The experimental data show a clear peak around the Curie

temperature CT (633K) of Ni, which means the magnetic phase transition has significant

effect on the thermal expansion. In the low temperature range, the experimental results

are in good agreement with each other. However, at the high temperatures, the thermal

expansion coefficient by Yousuf et al. [29] is lower than those in previous reports [30, 31].

Since the purity of their samples [29] is higher than the others, the results from Yousuf et

al. [29] were used to determine the model parameters in Equation 2.7 in the present work.

The calculated linear thermal expansion coefficient of Ni is shown in Figure 2.3 as the

solid line. The low-temperature data ( K300<T ) were not used in the parameter

determination, and deviate from the solid line because Equation 2.7, as discussed in the

previous section, is only applicable in the high temperature range.

Lattice parameters for Ni have been measured by many investigators [29, 34-37]. But the

agreements among their results are quite poor, especially in the high temperature range.

One of the possible reasons is the effect of magnetism that was recently emphasized by

many investigators [28, 29, 38]. Figure 2.4 plots the selected data from the literature.

The poor agreements between the experimental data from different researchers can be

observed in the high temperature range. It therefore seems difficult to define a suitable

mathematic description just on the basis of the experimental lattice parameter data.

16

After determining the model parameters in Equation 2.7 from the above experimental

thermal expansion data, the lattice parameter of Ni can be calculated from Equation 2.9.

The calculated lattice parameter is shown in Figure 2.4 by the solid line. In the low

temperature range ( cTT < ), the experimental values are well reproduced by the

calculation ( 3105.1 −×=NS Å). At high temperatures, although the available

experimental data are relatively scattered, our modeling still shows a reasonable

description ( 3101.5 −×=NS Å). As shown in Figures 2.3 and 2.4, changes of slopes of

both thermal expansion and lattice parameter near the Curie temperature were observed

experimentally. To reproduce the phenomena requires a more accurate model that takes

into account the magnetic effect.

In the present work, we chose the 0T in Equations 2.8 and 2.9 as 2/mT , and the 0a is

evaluated from the available experimental data of the lattice parameter.

2.3.2 Binary Ni-Al System

2.3.2.1 Experimental Data

The lattice parameters of the γ solid solution in the Ni-Al system were measured by

several groups [39-45]. Most of those values were measured at room temperature on

samples quenched from high temperatures. The temperature dependence of the lattice

parameter of Ni-Al alloys was investigated by Kamara et al. [40] using high-temperature

17

X-ray diffractometry. Another in situ X-ray measurement was performed by Bottiger et

al. [39] on five different composition alloys, and the lattice parameters up to 553K were

reported.

The room-temperature lattice parameters of the 'γ phase were determined by many

investigators [40, 41, 44-59] on samples quenched mostly from 1173-1473K. The

temperature effect on the γ ’ phase were studied by Arbazov and Zelenkov in the

temperature range 293-974K [31] and Taylor and Floyd for 74-1273K [44] by the

dilatometry technique, and their results on the relative thermal expansion agree with each

other. Using high-temperature X-ray diffractometry, Kamara et al. [40], Rao et al. [60]

and Stoeckinger and Neumann [53] investigated the temperature dependencies of the γ ’

lattice parameter. A typical experimental method includes four steps: sample preparation,

heat treatment (homogenization or aging), quench and measurement (mostly at room

temperature). Kamara’s experiment procedure [40] can be described briefly as follows

1) Prepared the Ni-17.7 at% Al alloy by the arc melting;

2) Homogenized the samples at 1273K for 30 minutes and age them at 973K for 168

hours;

3) Quenched the annealed sample to room temperature;

4) Measured the lattice parameters at different temperatures (293, 563, 713, 843,

953K) for 0.6-2 hours.

The measured lattice parameter value is strongly affected by experimental details. On

one hand, the measuring temperature has direct effect on the lattice parameter because of

18

the thermal expansion; on the other hand, many experimental conditions can also impact

the lattice parameter by changing the composition and the order parameters of phases.

The order parameter η , the degree of ordering, can be calculated by the site-fraction of

various elements in the ordered phase. For γ ’ in the Ni-Al system, the order parameter

can be defined as

IIAl

IAl

IAl

IIAl

yyyy

+−

=3

η (2.15)

where IAly and II

Aly are the site fractions of Al in the first and second sublattices,

respectively. Using the thermodynamic descriptions by Dupin et al. [61, 62], the change

of the order parameter with temperature is shown in Figure 2.5 for the Ni-25 at% Al alloy.

Obviously, the order parameter changes very little in the low temperature range.

During measurement, the compositions of phases change with the measurement

temperature and the measurement time. This evolution can be simulated by Dictra

software [2]. The thermodynamic database from Dupin et al. [61, 62] and the mobility

database from Engstrom and Agren [63] were used in the simulation. Since the

diffusivity in the γ ’ ordered phase is much smaller than that in the γ disordered phase

(about one order of magnitude smaller in Ni-Al system [64]), only the diffusion in the γ

phase was considered in the present simulation. The average simulation size of the γ

phase (~1.5 mµ ) and the measurement temperatures (293, 563, 713, 843, 953K) were

obtained from Kamara’s experiment [40], and the initial composition (12.18 at% Al) is

taken as the equilibrium compositions at the aging temperature (973K). The change of

19

the average composition in the γ phase at 953K (the highest measurement temperature in

Kamara’s experiment [40]) is shown in Figure 2.6, and the two dashed lines refer to the

initial composition and the equilibrium composition at 953K, respectively. According to

this figure, the composition will not change significantly during the measurement

(usually less than 3 hours) if the temperature is lower than 953K.

We can thus assume the compositions of the samples measured below the aging

temperature are the same as those at the aging temperature (frozen composition

assumption).

2.3.2.2 Evaluation of Model Parameters

According to Equation 2.10, the lattice parameters of the γ disorder solution can be

described as

)( ,,00 TBAxxaxaxa NiAlNiAlNiAlNiNiAlAl +++= (2.16)

NiAlA , and NiAlB , are model parameters to be evaluated from the experimental data of the

γ phase.

As the γ and γ ’ phases are described with one single Gibbs energy by two sub-lattices

using the formula 13 ),(),( NiAlNiAl [62], for the γ ’ phase, Equation 2.14 is rewritten as

AlAlAl aa 0:

0 = (2.17)

NiNiNi aa 0:

0 = (2.18)

20

TDCaaa AlNiAlNiNiAlAlNi ::00

:0 75.025.0 +++= (2.19)

TDCaaa NiAlNiAlNiAlNiAl ::00

:0 25.075.0 +++= (2.20)

Due to the limited experimental data, the following assumptions were made to reduce the

number of independent parameters:

0,:, =NiAlAlNiI (2.21)

TBAI NiAlNiAlNiAl ,:*,:*,:* += (2.22)

)(33 ,:*,:*,:**:, TBAII NiAlNiAlNiAlAlNi +== (2.23)

where the asterisk refers to Al or Ni. Thus, Equation 2.13 can be simplified to

NiAlIINi

IIAlAlNi

INi

IAl

NiNiIINi

INiAlNi

IIAl

INiNiAl

IINi

IAlAlAl

IIAl

IAl

IyyIyy

ayyayyayyayya

,:*,:*

:0

:0

:0

:0

3 ++

+++= (2.24)

Since "' 25.075.0 iii yyx += , and the phase is disordered when "'iii yyx == , we obtain

=++=++

NiAlNiAlAlNiNiAl

NiAlNiAlAlNiNiAl

BDDBACCA

:::,:*

,::,:*

44

(2.25)

Using the experimental lattice parameter values of the γ ’ phase and the previously

obtained parameters NiAlA , and NiAlB , , the parameters of the ordered phase were

evaluated by the parrot module of Thermo-calc software [2]. The compositions and the

site fractions used in the present work were calculated from the thermodynamic

descriptions by Dupin et al. [61, 62]. All available experimental data were selected in the

present evaluation of model parameters, and the data obtained at room temperature were

given very low weight because of the large discrepancies.

21

2.3.2.3 Results and Discussion

All model parameters for the lattice parameters of the γ and γ ’ phases in the Ni-Al

system are listed in Table 2.1. It is shown that the lattice parameter of Al has a higher

temperature dependence than that of Ni. The Ni3Al has a rather weak temperature-

dependent lattice parameter, while the hypothetic Al3Ni phase has the highest temperature

dependence. The interaction parameter is negative and becomes more negative with

increasing temperature, which reduces the lattice misfit between γ and γ ’ as discussed

later.

The composition dependence of the lattice parameter of the γ solid solution at room

temperature is calculated and compared with experimental data in Figure 2.7, and the

calculated lattice parameters as a function of temperature for various compositions are

plotted with experimental values in Figure 2.8. Although the trends with temperature are

similar, results by Bottiger et al. [39] are significantly smaller than those reported by

Kamara et al. [40]. It is therefore impossible to reproduce both sets of data. The same

situation appears in the room-temperature data in Figure 2.7, where the values reported

by Bottiger et al. [39] are smaller than other available data [41-43, 65]. Since Bottiger’s

investigation was performed on thin-film samples from sputtering, their results are likely

to be influenced by several processing factors. For example, Bottiger et al. [39] found

that a higher sputtering pressure will reduce the lattice parameter. All experimental data

except for those from Bottiger et al. [39] were selected for the evaluation of model

22

parameters, and the calculation can represent most of those experimental data reasonably

well ( 3101.4 −×=NS Å).

The lattice parameters of the γ ’ phase measured at room temperature are plotted in

Figure 2.9, and they are quite scattered. Compared with measurements of the disordered

phase, the lattice parameter of the ordered phase is much more sensitive to experimental

procedures, e.g. heat treatment temperature and time. The treatment history will change

the composition and order parameter, and then cause large discrepancies on the measured

lattice parameters. The composition dependence of lattice parameter of the γ ’ phase at

room temperature was calculated by the present model as the solid line in Figure 2.9, and

the standard deviation NS is 3104.7 −× Å. The calculated curve lies among the

experimental data, and displays a similar slope to those reported by Noguchi et al. [51]

and Aoki and Izumi [52].

The calculated temperature dependence of the γ ’ phase lattice parameter is compared

with the experimental data in Figure 2.10. Both the data by Kamara et al. [40] and

Stoeckinger et al. [53] can be well reproduced by our model (Figures 2.10(a) and 2.10(b),

3105.2 −×=NS and 3107.2 −× , respectively). On the other hand, Rao et al. [60] did not

report the composition of their alloy. Their sample was prepared by arc melting and

homogenized at 1273K, and found to be in the single-phase region by metallographic

method. Thus the composition of their sample is probably between the Ni-rich part (23.0

at% Al) to Al-rich part (27.3 at% Al) of the γ ’ single phase region at 1273K. The

predicted values for these two compositions are plotted in Figure 2.10(c) as solid lines,

23

and Rao’s data lie between the two calculated curves and are closer to the Al-rich side of

the γ ’ phase.

The relative thermal expansion of the γ ’ phase (25 at% Al) shown in Figure 2.11 is given

by

'293

'293

'

'293

'

γ

γγ

γ

γ

aaa

aa −

=∆

(2.26)

where '293γa is the lattice parameter of the 'γ phase at 293K. The calculated results agree

reasonably well with the experimental data [31, 44] ( %06.0=NS ).

2.3.3 Lattice Misfit between γ and γ ’ Phases

The γ / γ ’ lattice misfit, δ , defined as the relative difference of the lattice parameters of

the matrix γ ( γa ) and the precipitate γ ’ ( 'γa )

γ

γγ

δa

aa −=

'

(2.27)

is considered to be an important microstructural quantity. Accurate lattice misfit data can

be used to analyze the microstructural evolution and have been a focus of many

investigations on commercial Ni-base alloys [66, 67].

The precipitation in a Ni-12.7 at% Al alloy aged at 973K was studied by Phillips [56]; the

misfit data was calculated from the lattice parameters obtained by X-ray measurements

on the quenched sample. Kamara et al. [40] aged a Ni-17.7 at% Al alloy at 973K for 168

24

hours, and measured the lattice parameters of the γ and γ ’ phases at different

temperatures (293, 563, 713, 843, 953K) by high temperature X-ray diffractometry.

From those data they calculated the corresponding misfits.

Using the results from the present work, the lattice misfit between γ and γ ’ phases in

Ni-Al binary system were predicted and plotted in Figure 2.12. The solid curve shows

the misfit between the two equilibrium phases. On the other hand, as pointed out earlier,

if the holding time is not long enough, the γ and γ ’ phases in the samples would

maintain their compositions at the aging temperature (973K), and their lattice parameters

should thus be calculated using the corresponding equilibrium compositions at the aging

temperature. The corresponding misfits thus calculated are shown by the dashed line in

Figure 2.12 for the aging temperature of 973K. Since an accurate determination of the

γ - γ ’ lattice misfit in the laboratory is often difficult because it is very sensitive to the

experimental conditions, such as sample preparation [68] and aging time [69], the data

reported by Phillips [56] and Kamara et al. [40] can be considered being well represented

by the present calculations ( 4101.4 −×=NS ). Furthermore, as shown in Figure 2.12, the

lattice misfit value decreases with increasing temperature and crosses zero at around

1252K. This phenomenon is also supported by several experimental investigations in

commercial Ni-based superalloys [13, 40, 70-72].

2.4 Summary

25

A phenomenological model is developed in this chapter to describe the lattice parameters

of substitutional solid solution. The lattice parameters of the pure elements are modeled

under the assumption of a linear temperature dependence of thermal expansion, and those

for solution phases are treated by an approach similar to that used in the Gibbs energy

modeling. This model has been applied to the Ni-Al system. Most available lattice

parameter data of the γ and γ ’ phases in Ni-Al system can be reproduced, and the γ - γ ’

lattice misfit can also be reasonably predicted by taking into account the slow diffusion

during the measurement.

26

Table 2.1 The optimized parameters for the γ and γ ’ phases in Ni-Al system (in Å).

28-5:

0 101237.4106.85724.0262 TTa AlAl−×+×+=

29-5:

0 102456.9104.32663.5098 TTa NiNi−×+×+=

Taaa AlAlNiNiAlNi62

:0

:0

:0 104818.8105743.825.075.0 −− ×−×−+=

Taaa AlAlNiNiNiAl41

:0

:0

:0 107764.1101893.175.025.0 −− ×+×−+=

TI NiAl41

*:, 102687.1104516.1 −− ×−×−=

TI NiAl52

,:* 102290.4108385.4 −− ×−×−=

27

Figure 2.1 Calculated linear thermal expansion coefficient of fcc Al (solid line

T-85 1003.21068.1 ×+×= −α and 16 K101.1 −−×=NS for K300>T ) in comparison with

experimental data from the literature (8[18], z [20], � [73], Ú [74], � [75], - [76]

and " [77]).

28

Figure 2.2 Comparison of lattice parameter data for Al and the model calculation (Ú[74],

- [76], � [78], � [79], z [80], 8 [81] and " [82]). The dotted and solid lines represent

the results from Equation 2.8

( [ ])500(1001.1)500(1068.1exp0708.4 2285 −×+−×= −− TTa and 3103.1 −×=NS Å) and

Equation 2.9 ( [ ]1)500(1001.1)500(1068.10708.4 2285 +−×+−××= −− TTa and

3103.1 −×=NS Å), respectively.

29

Figure 2.3 Calculated linear thermal expansion coefficient of fcc Ni (solid line

T-85 1052.01022.1 ×+×= −α and 16 K103.1 −−×=NS for K300>T ) in comparison with

experimental data from the literature (�[27] , � [28], Ú [29], z [30], - [31], 8 [32]

and " [33]).

30

Figure 2.4 Comparison of lattice parameter data for Ni (Ú[29], � [34], � [35], z [36]

and - [37]) and the model calculation (solid line

[ ]1)900(1026.0)900(1022.15560.3 2285 +−×+−××= −− TTa and 3106.3 −×=NS Å).

31

Figure 2.5 Order parameter vs. temperature curve for the Ni-25at% Al alloy.

32

Figure 2.6 Average composition in the γ phase as a function of the holding time during

measurement at 953K. The initial and final compositions (dotted lines) refer to the

equilibrium compositions at the previous annealing temperature (973K) and the

measurement temperature (953K), respectively.

33

Figure 2.7 Room-temperature lattice parameter of the γ phase in the Ni-Al system

(-[39], Ì [41], ! [42], Ú [43], K [44], L [56], z [59] and � [65]). The solid line

represents the results of the model calculation ( 3104.6 −×=NS Å).

34

Figure 2.8 Temperature dependence of the lattice parameter of the γ phase in various

Ni-Al alloys (-, 8, ", M, C [39] and � [40]). The solid lines represent the results of

the model calculation ( 3109.7 −×=NS Å for [39] and 3107.1 −×=NS Å for [40]).

35

Figure 2.9 Room-temperature lattice parameter of the γ ’ phase (�[40], Ì [41], K [44],

" [45], � [46], [47], � [48], � [49], , [50], 7 [51], 8 [52], B [53], C [54], M [55],

L [56], * [57], A [58], z [59], ! [83], - [84], and 5 [85]). The solid line represents

the results of the model calculation ( 3104.7 −×=NS Å).

36

Figure 2.10 Temperature dependence of the γ ’ lattice parameter (�[40], B [53], and 7

[60]). The solid lines represent the results of the model calculation ( 3105.2 −×=NS Å for

[40] and 3107.2 −×=NS Å for [53]).

37

Figure 2.11 The comparison of the experimental relative thermal expansion of the γ ’

phase (25 at% Al) (-[31] and K [44]) and the model calculation (solid line)

( %06.0=NS ).

38

Figure 2.12 The calculated misfit between the γ and γ ’ phases (�[40] and L [56]).

The solid curve shows the misfit under the equilibrium condition, and the dashed lines

represent those under the frozen composition assumption ( 4101.4 −×=NS ).

39

Chapter 3

First-principles Study of Lattice Distortion in γ

3.1 Background

In the previous chapter, we developed a phenomenological model to describe the lattice

parameter of γ and γ ’ as a function of temperature and composition, and the

contribution from alloying elements was treated using an approach similar to that used in

the Gibbs energy modeling [86]. It was then applied to Ni-Al binary alloys and a self-

consistent lattice parameter database for the γ and γ ’ phases was constructed. In that

particular case, the model parameters were evaluated using a large amount of

experimental data available in the literature. However, in general, the availability of

experimental data on lattice parameters of alloys is very limited. Very often there are

poor agreements among experimental data from different sources. As a result, extracting

modeling parameters based on experimental data alone can be difficult.

In the last decade, the quality of first-principles calculations of electronic and structural

properties has been improved considerably. For most cases, reliable formation energy of

alloys and compounds and band structures can be calculated at 0K. In this chapter, we

use the first-principles approach to understand the lattice distortion caused by alloying

elements. As a part of the development of lattice parameter database on Ni-base

superalloys, the current study is focused on the local and macroscopic lattice distortions

40

caused by various solute additions in γ (this chapter) and γ ’ (next chapter) of Ni-base

superalloys, and the goal is to predict the lattice parameter changes in Ni-base binary and

multi-component alloys as a function composition. Ten commonly used alloying

elements in Ni-base alloys are chosen for this study, namely, Al, Co, Cr, Hf, Mo, Nb, Re,

Ru, Ta, Ti and W, and the results are compared with available experimental

measurements.

3.2 First-principles Calculations

The first-principles calculations of the lattice parameter were performed using the Vienna

ab initio simulation package VASP (Version 4.6) [87], which allows one to minimize the

total energy with respect to the volume and shape of the cell and the positions of atoms

within the cell. In the present calculations, ultrasoft pseudopotentials and the generalized

gradient approximation (GGA) [88] are adopted. GGA partially corrects the overbinding

problem of the local density approximation (LDA) [89], and thus improves the

predictions for the equilibrium volumes [90, 91]. Supercells were employed to study the

lattice distortions caused by solute atoms. A convergence test on fcc Al by Sandberg et

al. [92] found that a supercell of 80 atoms was needed to achieve a convergency for the

formation energy of a single defect to be within 0.01 eV. We employ 108-atom

supercells with one solute atom in each supercell. To simulate the anti-ferromagnetic

property of Cr, at least two Cr atoms are required in the supercell, and thus a much larger

supercell and many configurations need to be considered. Instead of performing such

demanding calculations, two calculations were preformed, both with 107 Ni atoms and 1

41

Cr atom. In the first case, the Cr possesses the same spin direction as the surrounding Ni

atoms, denoted by “Cr+1”. In the second case, the spin direction of the Cr atom is

opposite to that of Ni atoms, denoted by “Cr-1”. The set of k points is adapted to the size

of the primitive cell, and a 444 ×× k -point mesh is selected for the supercell used in the

present calculations. The energy cutoff is determined by the choice of “high accuracy” in

the VASP. For a detailed description of the technical features and the computational

procedure of the VASP calculations we refer to the VASP’s manual [93].

3.3 Lattice Distortions

Introduction of solute atoms leads to redistribution of electron density and lattice

distortions. There are two kind of lattice distortions introduced. One is the macroscopic

lattice distortion represented by the overall lattice parameter change of an alloy. The

other is local lattice distortion. The overall lattice parameter change, a∆ , is defined as

puresol aaa −=∆ (3.1)

where purea is the lattice parameter of pure solvent, and sola is that of the solution

containing solution atoms. For dilute solutions, a∆ is approximated as a linear function

of compositions

∑=∆i

ii kxa (3.2)

where ix is the mole fraction of solute atom i , and ik is the linear regression coefficient.

In this work, we determine ik by first-principles calculations using N -site supercells.

42

The pure solvent is represented by decorating all sites with the solvent atom, and the

solution contains one solute atom. The composition of such a solution is

Nxi

1= (3.3)

And then ik can be calculated by the following equation

ii aNk ∆= (3.4)

The results from first-principles calculations are shown in Table 3.1, and the linear

regression coefficients for the ten solute atoms are listed in Table 3.2.

Empirically, for a known crystal structure, the lattice parameter is related to the atomic

radius, so the dependence of an alloy lattice parameter on the solute composition is

typically explained by the atomic radius of solute atoms. For example, the lattice

parameter of a solvent is expected to increase when solute atoms of larger atomic radii

are added. In Table 3.2, the atomic radii of the solute atoms are compared with their

effects on the lattice parameter of fcc Ni. All the ten solute elements have larger atomic

size than Ni, and they all increase the Ni lattice parameter as expected. In general, the

magnitude of the lattice parameter change increases with the size of the atomic radius of

the solute atom. However, this empirical relation is not always observed. For example,

the atomic radius of Al (1.43 Å) is larger than that of Re (1.37 Å) [94], while the lattice

parameter increase due to the addition of Al atoms ( 0.1587=Alk ) is considerably

smaller than that caused by Re ( 0.3903Re =k ). This may not be surprising because it is

commonly known that the radius of an atom depends on the environment. The atomic

radius is typically defined as one half of the internuclear distance between two adjacent

43

atoms at equilibrium. Such a definition is not scientifically rigorous and the values are at

best approximate. A more accurate prediction of atomic radius must take into account

the interactions between the solute and solvent atoms during alloying. One thus must

differentiate the classic atomic radii measured in pure elements and those in alloys.

All the solute considered except Al are transition elements, and their outermost electron

structures consist of s and d electrons. In Figure 3.1, the differences of atomic radii

between solute atoms and solvent atom (Ni) are plotted with the linear regression

coefficients of lattice parameters according to their periods (Figure 3.1(a)) and groups

(Figure 3.1(b)) in the periodic table. It is shown that for solutes within the same period,

the lattice parameter increases with the increase in the solute atom radius although the

relation is not linear. Figure 3.2 shows the electronic charge density of Nb, Mo and Ru

on the (001) planes of the Ni host lattice. All three elements belong to period 5.

Following observations can be made: i) the charge density of Nb shows a clear

interaction along the <110> directions with the nearest neighbor Ni atoms; ii) the Mo d

electrons are highly localized and less chemically inactive due to its half-filled d shell; iii)

the more outermost electrons of Ru has a higher and wider distribution of charge density

which represents a stronger interaction between Ru and neighboring Ni. Comparing with

Nb and Ru, Mo has a much weaker interaction with neighboring atoms, indicating an

easier compressing. As a result, the data point for Mo deviates from the line connecting

the Nb and Ru data points in Figure 3.2. In a given group, the outermost electron

structures are usually similar for all elements, so the interactions between those elements

(solute) and Ni (solvent) are expected to be similar. In this case, the lattice parameter

44

change is expected to have a close correlation with the atomic radius of the solute

element. Indeed, it is found that the effect of solute atoms in group 4 (Ti and Hf) and

group 6 (Cr, Mo amd W) on lattice parameters can be explained by their corresponding

radii. However, for group 5 (Nb and Ta), such a correlation is not observed. According

to Table 3.2, the atomic radius of Nb (1.46 Å) is slightly smaller than that of Ta (1.49 Å)

[94], but the lattice expansion caused by Nb solute ( 0.6123=Nbk ) is a little bit larger

than that caused by Ta ( 0.5859=Tak ). This anomaly may be explained by the difference

in the valance electronic structures between Nb (4d4 5s1) and Ta (5d3 6s2) [94] that causes

the interaction between the solvent atom and those two solute atoms are not similar any

more. The electronic charge densities of Nb and Ta solutes are shown in Figure 3.3. The

charge density of Nb exhibits a much stronger interaction with neighboring Ni atoms than

that between Ta and neighboring Ni atoms. Therefore, Nb atoms are much harder to be

compressed, leading to a larger lattice extension than Ta in Ni host lattice.

In addition to the electron density redistributions, the interactions between magnetic spins

also contribute to lattice distortions. Repulsion is expected between spins with the same

direction and attraction will occur with opposite directions. For example, for the case of

Cr substitution in fcc Ni, the atomic size difference between Cr and Ni is very small.

However, there is a significant effect of Cr addition on the fcc Ni lattice parameters as a

results of magnetic. When the spin direction of a Cr atom is the same spin as a

neighboring Ni atom (Cr+1), the linear regression coefficient is positive, i.e. lattice

parameter increases as a result of repulsive interactions between the magnetic spins.

Therefore, it is expected that if a Cr atom of opposite spin direction (Cr-1) is introduced,

45

the attractive force between magnetic spins decreases the lattice parameter, and the linear

regression coefficient becomes negative.

Near a solute atom, the local distortion is generally different from the macroscopic lattice

parameter change of a solid solution in magnitude and even in sign in some cases [95].

Experimentally, the local distortions are described by the shifts in the nearest-neighbor

distances around a solute atom. They can also be readily obtained using first-principles

calculations by relaxing both the supercell dimensions and internal atomic positions.

Using the X-ray absorption fine structure (XAFS) technique, Scheuer and Lehgeler [95]

systematically studied the lattice distortions around impurity atoms in dilute metal alloys

(the solute concentrations are between 1 and 2 at.%). In particular, they reported the

shifts in nearest-neighbor distances of Cr, Co, Mo, Nb, and Ti in fcc Ni. We compare the

experimentally measured values with those from our calculations in Table 3.3. It is found

that the calculated data agree with the experimental results within the experimental

uncertainties. Among them, Co exhibits the difference between the macroscopic lattice

parameter change and the local lattice distortion. Macroscopically, Co atoms expand the

fcc Ni (Table 3.2) while locally they decrease the nearest-neighbor distances (Table 3.3).

A comparison of the calculated results for Cr+1 and Cr-1 indeed show that the spin

direction has a strong effect on the local lattice distortions, i.e., Cr-1 atoms decrease the

nearest-neighbor distance and Cr+1 increases it.

3.4 Lattice Parameters Change

46

Using the data shown in Tables 3.1 and 3.2, the lattice parameters and the lattice

parameter changes in fcc-Ni solution can be predicted using Equations 3.1 and 3.2,

respectively. The predicted results can be compared with the experimentally measured

compositional dependence of lattice parameters in Ni-X binary alloys. The lattice

parameter measurements were often carried out by diffraction methods. The results are

usually sensitive to the experimental details, and leading to significant discrepancies

among data from different measurements. For example, both Taylor and Floyd [96] and

Pearson and Thompson [97] measured the lattice parameters of Ni-Cr alloys. The results

by Pearson and Thompson show 0.013Å lower than those by Taylor and Floyd, and the

magnitude of the discrepancy equals to that of the lattice parameter change by adding 10

at.% Cr. To reduce the systematic error within a particular measurement, we compare

the measured lattice parameter changes with our calculations. In extracting the lattice

parameter changes, the lattice parameter of pure Ni from the same investigation is taken

as the reference. In cases that the data for pure Ni were not available, the linearly

extrapolated value for pure Ni is used. The experimental data for binary Ni solid

solutions containing Al [39, 44, 59, 65], Co [36, 97-99], Cr [45, 96, 97, 100-102], Hf

[45], Mo [34, 45, 103-106], Nb [45, 107-111], Re [34, 112], Ru [113-115], Ta [45, 116],

Ti [45, 96, 117-120] and W [45, 121] are used for comparison.

The calculated lattice parameter changes in Ni-Al, Ni-Co, Ni-Hf, Ni-Mo, Ni-Nb, Ni-Re,

Ni-Ru, Ni-Ta, Ni-Ti and Ni-W, together with the related experimental data, are shown in

a series of plots in Figures 3.4-3.10. The differences are demonstrated by the standard

47

deviations NS , calculated as the square root of the sample variance of a set of values

[26]. Most of the experimental data are well-reproduced by the calculated results shown

in solid lines. The standard deviations for Ni-Al and Ni-Ru alloys are slightly higher than

others. For Ni-Al, the main reason is that the available experimental data are much more

scattered than other systems (Figure 3.4). In the case of Ni-Ru, in Figure 3.9, the

agreement between experimental data and our predictions is reasonable at low Ru

concentrations (< 10 at.%) while the large deviations are observed at high concentrations

where the linear approximation (Equation 3.2) may no longer be valid. The results of

calculations for “Cr+1” and “Cr-1” are shown in Table 3.4 and Figure 3.11. The linear

regression coefficient for Cr+1 is positive, while that for Cr-1 is negative, because the

magnetism has significant effect on lattice distortion as discussed in the previous section.

If the system is large enough and the interaction between the two kind of Cr atoms can be

ignored, the real case can then be taken as a weighted average of the above two cases,

which is supported by the comparison of thus evaluated and experimental lattice

parameter changes shown in Figure 3.11 where all experimental data lie between the Cr+1

and Cr-1 lines. The CAvg line indicates the half-half average of the Cr+1 and Cr-1 lines, and

is close to but a little lower than the experimental data, which means the Cr+1 case might

have a higher possibility to occur in the real system. Our energy calculations also show

that the total energy of the Cr+1 system is a little bit lower (200 J/mol) than the Cr-1

system.

There are some previous investigations [45, 71, 122, 123], found from the literature,

evaluating the linear regression coefficients of some solute atoms in fcc-Ni by fitting to

48

the experimental data. Mishima et al. [45] evaluated the linear regression coefficients

from experimentally determined lattice parameters in Ni-X binary systems, and all other

three investigations [71, 122, 123] are based on the data from multi-component nickel

alloys. Harada and Yamazaki [123] assumed the coefficients of the corresponding

alloying elements are the same for both of the γ and the γ ’ phases, while Watanabe and

Kuno [122] treated them to be different. Svetlov et al. [71] introduced some higher order

parameters in their model to describe the interaction of different solute elements. Their

linear regression coefficients are summarized in Table 3.4. As shown in Table 3.4, the

linear regression coefficients determined by the present work are very close to those

given by the three earlier investigations [45, 122, 123]. Svetlov’s results [71] are

somewhat different from others, even inconsistent to some experimental information.

For example, their strong negative linear regression coefficient for Ti indicates a

significant decrease in lattice parameter, but the experiments show that adding Ti atoms

will expand the fcc Ni lattice (Figure 3.10).

Lattice parameters of several Ni-Al-Cr-Co-Mo-Nb-Re-Ta-Ti-W alloys are calculated by

Equations 3.1 and 3.2, and two sets of linear regression coefficients (Mishima’s [45] and

ours) are used in the calculation. Since the linear regression coefficient for Re was not

determined by Mishima et al. [45], a value of 0.413 is given by fitting of the experimental

data shown in Figure 3.8. The calculated results are compared with the experimental data

[14, 124-127] in Table 3.5, and the standard deviations NS are also given. The standard

deviations for two different sets of coefficients are almost the same (~0.02 Å), which

confirm again the validity of our first-principles approach in analysis of lattice parameter

49

change caused by solute additions. The lattice parameter data reported by Li and Wahi

[124] and Volkl [125] cannot be well-reproduced by present calculations, and some

possible reasons are: i) the measurement uncertainty due to the experimental details; ii)

the invalidation of the linear relationship; iii) the interaction between different solute

atoms. The last two reasons are caused by the high concentrations of the solutes.

3.5 Summary

In this chapter, we present a first-principles approach to study the lattice distortion (both

macroscopic one and local one) caused by the solute atoms, and apply it to the fcc-Ni

lattice with Cr, Co, Mo, W, Ta, Re, Ru, Nb, Al, Ti and Hf as solutes. The effects of

atomic size, electronic interaction, and magnetic spin direction on lattice distortion have

been discussed through electronic charge density distributions. The calculated lattice

parameter changes and the local distortions in Ni-X alloys agree well with the

experimentally measured data in the literature, which demonstrated the validity of the

approach. Using the linear regression coefficients from the first-principles calculations,

the lattice parameters in multi-component Ni-base superalloys are predicted and

compared with available experimental observations, and good agreements are observed

when the concentrations of individual solutes are not too high.

50

Table 3.1 Total energies and lattice parameters for Ni107X1 fcc solutions from first-

principles calculations

Lattice Parameter (Å) X Total Energy

(eV/atom) a a∆

Ni -592.09001 10.596251 -

Al -591.85935 10.600658 0.004407

Co -593.69888 10.596616 0.000365

Cr+1 -595.91956 10.601294 0.005043

Cr-1 -595.69447 10.596021 -0.000230

Hf -597.83047 10.619488 0.023237

Mo -597.29946 10.607510 0.011259

Nb -597.40122 10.613259 0.017008

Re -598.55348 10.607091 0.010840

Ru -595.27323 10.607117 0.010866

Ta -599.39013 10.612527 0.016276

Ti -595.82672 10.605668 0.009417

W -599.50809 10.607606 0.011355

51

Table 3.2 Atomic radii and electron structure of solute atoms vs. lattice parameter

change in fcc Ni

Element Atomic radius, (Å)

[94]

Electron structure Linear regression

coefficients,

(Å/at.%)

Ni 1.24 3d84s2 -

Al 1.43 3s23p1 0.1587

Co 1.25 3d74s2 0.0132

Cr 1.30 3d54s1 0.181(Cr+1)

-0.008(Cr-1)

0.0865(CrAvg)

Hf 1.67 5d2 6s2 0.8365

Mo 1.39 4d5 5s1 0.4053

Nb 1.46 4d4 5s1 0.6123

Re 1.37 5d5 6s2 0.3903

Ru 1.34 4d7 5s1 0.3912

Ta 1.49 5d3 6s2 0.5859

Ti 1.45 3d2 4s2 0.3390

W 1.41 5d4 6s2 0.4088

52

Table 3.3 Local lattice distortion in fcc Ni (in pm)

Elements Reference [95] Present work

Al - 1.5

Co -0.4±0.6 -0.61

1.06 (Cr+1) Cr -1.1±0.7

-1.4 (Cr-1)

Hf - 7.37

Mo 2.4±0.3 1.92

Nb 5.4±0.7

4.0+0.9

4.5

Re - 1.53

Ru - 2.74

Ta - 4.14

Ti 2.2±0.4 2.48

W - 1.97

53

Table 3.4 Linear regression coefficients of solute atoms in fcc Ni (in Å/at.%)

Reference Solute

atom [122] [123] [45] [71]

Present work

Al 0.183 0.185 0.194 0.100 0.1587

Co 0.024 0.020 0.020 0.081 0.0132

Cr 0.130 0.120 0.112 0.133 0.181(Cr+1)

-0.008(Cr-1)

0.0865(CrAvg)

Hf - 0.700 0.990 - 0.8365

Mo 0.421 0.435 0.480 0.357 0.4053

Nb - 0.645 0.697 3.673 0.6123

Re - - - - 0.3903

Ru - - - - 0.3912

Ta - 0.630 0.697 0.593 0.5859

Ti 0.360 0.340 0.424 -1.222 0.3390

W 0.421 0.412 0.448 0.215 0.4088

54

Table 3.5 Comparison of calculated and experimental lattice parameter in Ni-base superalloys.

Composition, (at.%) Lattice parameter, (Å)

Al Cr Co Mo Nb Re Ta Ti W Ni

Ref.

Exp. [45] Present

work

9.00 17.83 10.80 1.55 0.00 2.05 0.43 0.64 1.58 56.12 3.579 3.5923 3.5806

9.41 17.23 10.72 1.54 0.00 2.31 0.50 0.63 1.81 55.85

[14]

3.584 3.5949 3.5830

4.62 28.58 11.09 1.79 0.58 0.00 0.54 1.24 0.77 50.79 3.556 3.5923 3.5788

6.55 24.90 9.27 1.79 0.82 0.00 0.98 1.99 0.46 53.24 3.553 3.5981 3.5838

7.36 24.83 9.40 1.67 0.56 0.00 0.40 1.94 0.58 53.26 3.553 3.5935 3.5799

7.79 24.25 9.28 1.79 0.82 0.00 0.90 2.25 0.51 52.41 3.554 3.6005 3.5858

7.37 23.76 9.49 1.67 0.77 0.00 0.65 2.04 0.74 53.51 3.554 3.5967 3.5827

6.80 24.85 9.08 1.74 0.67 0.00 0.58 1.96 0.87 53.45

[124]

3.558 3.5961 3.5822

5.21 25.78 21.26 0.55 0.00 1.29 0.06 0.24 1.02 44.58 [125] 3.600 3.5812 3.5700

2.58 8.05 9.35 0.56 0.00 8.72 0.36 0.08 2.60 67.70 [126] 3.583 3.5931 3.5856

3.50 19.80 18.40 0.70 0.00 2.70 0.30 0.30 2.70 51.60 [127] 3.588 3.5866 3.5763

NS 0.0221 0.0205

55

Figure 3.1 Atomic radius difference between solute and solve (Ni) atoms vs. Linear

Regression Coefficient.

56

(a) (b) (c)

Figure 3.2 Electronic charge density (in units of e/Å3) of (a) Nb, (b) Mo and (c) Ru solutes in the (001) plane of the fcc Ni lattice in

period 5.

57

(a) (b)

Figure 3.3 Electronic charge density (in units of e/Å3) of (a) Nb and (b) Ta solutes in the

(001) plane of the fcc Ni lattice in group 5.

58

Figure 3.4 Lattice parameter changes in Ni (γ ) solid solutions with additions of Al and

W ( 0059.0=NS Å for Al and 0015.0=NS Å for W). Experimental data: r [59], �

[44], £ [65], ¯ [39], s [45], Ï [121].

59

Figure 3.5 Lattice parameter changes in Ni (γ ) solid solutions with additions of Co and

Hf ( 0011.0=NS Å for Co and 0017.0=NS Å for Hf). Experimental data: r [97], s

[45], � [98], £ [99], ¯ [36].

60

Figure 3.6 Lattice parameter changes in Ni (γ ) solid solutions with additions of Nb

( 0027.0=NS Å for Nb). Experimental data: s [45], £ [107], r [108], ¯ [110], Ï

[111].

61

Figure 3.7 Lattice parameter changes in Ni (γ ) solid solutions with additions of Mo

( 0041.0=NS Å for Mo). Experimental data: s [45], Ï [103], r [104], £ [105], ̄

[106], È [34].

62

Figure 3.8 Lattice parameter changes in Ni (γ ) solid solutions with additions of Re and

Ta ( 0013.0=NS Å for Re and 0026.0=NS Å for Ta). Experimental data: s [45], È

[34], r [112], £ [116].

63

Figure 3.9 Lattice parameter changes in Ni (γ ) solid solutions with additions of Ru

( 0055.0=NS Å for Ru). Experimental data: � [113], r [114], £ [115].

64

Figure 3.10 Lattice parameter changes in Ni (γ ) solid solutions with additions of Ti

( 0011.0=NS Å for Ti). Experimental data: � [96], s [45], r [117], £ [118], ¯ [119],

Ï [120].

65

Figure 3.11 Lattice parameter changes in Ni (γ ) solid solutions with additions of Cr

( 0088.0=NS Å for CrAvg). Experimental data: � [96], r [97], £ [100], ̄ [101], Ï

[102], s [45].

66

Chapter 4

First-principles Calculations and Phenomenological Modeling of Lattice Misfit in Ni-base Superalloys

4.1 Background

It has been mentioned in previous chapter that one of the critical factors that control the

morphology of coherent γ ’ precipitates in the γ matrix is the magnitude and sign of the

stress-free lattice misfit between γ and γ ’. The lattice misfit is calculated from the

stress-free lattice parameters of the γ and γ ’ phases, which are typically measured by X-

ray diffraction method (XRD) [128] or convergent beam electron diffraction method

(CBED) [129]. The results are very sensitive to the details of alloy processing [86], and

the incoherent and equilibrium conditions must be satisfied for a measurement in a multi-

phase mixture. Consequently, the results of lattice misfit from different reports are

usually very scattered, especially for nickel-base superalloys where the lattice parameters

of γ and γ ’ are close to each other. The experimental data are even more scatted for

multi-component systems.

In this chapter, an integrated computational approach was developed to predict lattice

misfit by combining first-principles calculations and phenomenological modeling. In

particular, we applied this approach to obtaining the lattice misfits in both Ni-Al binary

and Ni-Al-Mo ternary systems. Agreement between calculated data and experimental

67

data is good for Ni-Al alloys, but not for some Ni-Al-Mo alloys. Such a discrepancy is

discussed in Section 4.3. Since the morphology of γ ’ precipitates will be significantly

changed by the lattice misfit value, comparisons of simulated morphologies of different

lattice misfits with the experimental observation is then used as a criteria of the reliability

of this approach. Phase-field simulations is carried out in Section 4.3.3 to obtain the

morphologies of different lattice misfits, and a detailed description of the phase-field

approach is given in Chapter 6.

4.2 Methodology

4.2.1 Lattice Parameters of Pure Metals and Ordered Compounds

In the last decade, first-principles calculations have been extensively used to obtain the

formation energies, band structures and lattice parameters of pure metals and compounds.

In the present work, the first-principle calculations of lattice parameters in the Ni-

superalloy system are performed using the Vienna ab initio simulation package VASP 4.6

[87]. The total energy of a system is minimized with respect to both the volume and

shape of a computational cell and the atom positions within the cell. In the present

calculations, the ultrasoft pseudopotentials and the generalized gradient approximation

(GGA) [88] were adopted. It has been generally known that GGA partially corrects the

overbinding problem of the local density approximation (LDA) [89] and thus improves

the predictions for the equilibrium volumes [90, 91]. The set of k points is chosen

according to the size of the computational cell. The energy cutoff is determined by the

68

choice of “high accuracy” in VASP. For a detailed description of the technical features

and the computational procedure of the VASP calculations we refer to the VASP’s

manual [93]. The calculated lattice parameters of pure Ni and Ni3Al compound ( 0a ) are

compared with the experimental data in Table 4.1.

To predict the lattice parameters at finite temperatures, thermal expansion information is

required. It can be determined experimentally (e.g., diffraction measurements) or

theoretically (e.g., first-principles linear-response theory) [130]. However, there is still

about a 10% uncertainty in the thermal expansion coefficients obtained from the

theoretical calculations due to various assumptions and approximations [130]. Such an

uncertainty can lead to an error of ~0.0035 in the misfit between γ and γ ’ in nickel-base

superalloys at 1000K. This error is significant since the measured lattice misfit in Ni-Al

binary alloys is only about 0.004 at 1000K [86]. Therefore, we still relied on

experimental data for the thermal expansion coefficients of γ and γ ’. In particular, we

used those values reported by Kamara et al. [40] who described the temperature effect

( Ta∆ ) by a quadratic function of temperature:

2cTbTaT +=∆ (4.1)

where b and c are constants (see Table 4.1).

4.2.2 Effect of Chemical Disordering

69

In the Ni-Al binary system, γ ’ has a L12 ordered fcc structure with two sublattices. One

sublattice is made up of face-centered sites occupied mostly by Ni atoms (Ni site), and

the other sublattice consists of fcc corner sites occupied mostly by Al atoms (Al site).

The degree of chemical order in γ ’ decreases with temperature increasing by mean of

anti-sites, i.e. Ni atoms go to Al site and Al atoms go to Ni site. The off-stoichiometry of

γ ’ is also realized by anti-site atoms. In multi-component systems, various solute

species are also expected to preferably distribute in one of the two sublattices. These

chemical disorders (due to the changes in composition and temperature) lead to the

changes in the lattice parameters. If these types of chemical disorders are relatively

small, we can approximate their effect on lattice parameter ( Ca∆ ) using a linear

combination:

∑∑=∆s i

si

siC yka (4.2)

where s indicates different sublattices, siy is the atomic fraction of element i in

sublattice s , and sik is the coefficient representing the effect of i in the s sublattice.

γ has a disordered fcc structure, where both sites are equivalent, and all atoms are in

random mixing. We can still use Equation 4.2 to describe the composition effect on the

lattice parameter change of γ , and the site fractions iy here are the same for all

sublattices and equal to ix , the atomic fractions in the γ phase.

70

In this work, we determine sik by using first-principles supercell calculations. Each

supercell contains one solute or antisite in a given sublattice. sik is then calculated using

the following equation:

)( 0aaNk si

si −= (4.3)

where 0a is the calculated lattice parameter for pure Ni or the completely ordered cell,

sia the calculated lattice parameter of the supercell containing one i atom in sublattice s ,

and N the total number of atoms in the supercell. In all calculations, the total number of

atoms is 108. The determined linear coefficients of solute or anti-site elements are

presented in Table 4.2.

4.2.3 Lattice Misfit

The dependences of lattice parameters of γ ( γa ) and γ ’ ( 'γa ) on temperature and

compositions are described by the following equation:

∑ ∑+++=∆+∆+=s i

si

siCT ykcTbTaaaaa 2

00',γγ (4.4)

For any given temperature T and composition ix , the site fractions in each phase can be

obtained from the thermodynamic databases by Dupin et al. [61] for Ni-Al and Zhou et

al. [131] for Ni-Al-Mo. The lattice misfit (δ ) between γ and γ ’ is defined as:

γ

γγδa

aa −= ' (4.5)

71

4.3 Results and Discussions

4.3.1 Ni-Al binary system

Using the approach described above, the lattice misfit between γ and γ ’ in Ni-Al binary

systems is plotted in Figure 4.1. The misfit shown by the solid curve was obtained by

assuming that the two phases are at thermodynamic equilibrium at each temperature. The

symbols represent the experimentally determined lattice misfits [40, 56] measured at

different temperatures on samples quenched from 973K. As mentioned in Chapter 2, the

phases in the samples were expected to maintain their equilibrium compositions at 973K

[86]. To test this hypothesis, we fixed the site fractions siy as the equilibrium values

corresponding to 973K, and then calculated the lattice misfits between the γ and γ ’

phases in Ni-Al binary system at different temperatures. The results are presented by the

dotted line in Figure 4.1. Indeed an excellent agreement is achieved by comparing the

dotted line and the experimental data points.

4.3.2 Ni-Al-Mo ternary system

Conley et al. [6] measured the room temperature lattice parameters of γ and γ ’ in three

Ni-Al-Mo alloys using the precision Debye-Scherrer powder X-ray method. The volume

fractions of γ ’, fV , are around 0.10. The sample were quenched from 1023K, and the

lattice misfits at 1023K were then evaluated assuming constant thermal expansion

72

coefficients for γ and γ ’. Fahrmann et al. [132] determined the lattice misfits of five

Ni-Al-Mo alloys at 1048K using high-temperature X-ray diffraction. The volume

fractions of γ ’ of alloys are between 0.10 and 0.20. Since the measurements were

performed on coherent microstructures, they increased the values for the lattice misfit by

a factor of 1.5 to approximate the coherency strain effect [132].

The compositional dependence of the predicted lattice misfit in Ni-Al-Mo is shown in

Figure 4.2. In general, with the increase in the Mo concentration, the lattice misfit

decreases and changes sign from positive to negative. This trend agrees with the above-

mentioned experimental measurements [6, 132]. As shown in Figure 4.2, the agreement

between experimental data (symbols) and our calculations (curves) are generally good

except for the alloy with the lowest Mo concentration (Ni-12.5 at.% Al-2.0 at.% Mo)

from Fahrmann et al. [132]. First, since the Mo concentration is very small in this alloy,

its lattice misfit should be close to that in Ni-Al binary alloys at the same temperature, i.e.

0.004 (see Figure 4.1). Second, Fahrmann et al. [132] reported that the microstructure of

this alloy was semicoherent instead of coherent, but they multiplied the same factor of 1.5

to estimate the corresponding stress-free misfit, which probably overestimated the stress-

free misfit.

4.3.3 Phase-field simulation of γ ’ precipitate morphology

73

Using the predicted lattice misfit, phase-field simulations were carried out to predict the

γ ’ precipitate morphology. A comparison between the predicted and experimentally

observed morphology provides another indirect evidence on the accuracies of the above

approach. In the phase-field method, a microstructure is described by a set of physical or

artificial fields, and its temporal and spatial evolution is governed by a set of

mathematical equations of the fields [133]. With reliable input data (properties of the

system, e.g., thermodynamic driving force, atomic mobility, lattice misfit and elastic

constant), the microstructure evolution and coarsening kinetics can be predicted

quantitatively. The simulation details can be found in Chapter 6.

We investigated the γ ’ precipitate morphologies in an alloy with composition Ni-12.5

at.% Al-2.0 at.% Mo using 2D phase-field simulations. Two sets of different misfit data

(0.0065 from Fahrmann et al. [132] and 0.0035 from the present work) were used. The

precipitate morphologies from experimental observation (a) and phase-field simulations

(b, c) are shown in Figure 4.3. The precipitate sizes from both simulations are somewhat

smaller than that from experiments. One of the reasons could be due to the 2D nature of

the simulations since the coarsening in 2D is slower than that in 3D because of the

reduced curvature. However, it is evident that the morphology using 0065.0=δ (Figure

4.3(b)) is quite different from that observed experimentally. The lenticular shape and

strong alignments are caused by large elastic stresses, indicating an overestimated misfit.

On the other hand, the particle shape in Figure 4.3(c) obtained using 0035.0=δ is very

74

similar to the experimental observation. According to those comparisons, we can

conclude that the predicted misfits from our work are more reliable.

4.4 Summary

In this chapter, we proposed an integrated computational approach for evaluating the

lattice misfit between γ and γ ’ in Ni-base alloys. It combines the first-principles

calculations of lattice parameters at 0K, experimental data on thermal expansion

coefficients, and phenomenological modeling. It was applied to the Ni-Al binary and the

Ni-Al-Mo ternary systems. A comparison between evaluated lattice mismatch and

experimental measurements shows good agreement in both its temperature and

composition dependences. Using the calculated lattice misfit, the precipitate morphology

of a Ni-Al-Mo alloy was predicted using a phase-field simulation and is shown to agree

well with experimental observation, which proves the reliability of this approach.

75

Table 4.1 Lattice parameters of ordered and disordered phases

0a (Å) 2cTbTaT +=∆ (Å)

Calculated

(0K)

Experimental

(298K) [86]

b (Å/K) [40] c (Å/K2) [40]

γ (Ni) 3.532 3.523 5.741×10-5 -1.010×10-9

γ ’ (Ni3Al) 3.573 3.552-3.589 6.162×10-5 -1.132×10-8

Table 4.2 Linear coefficients of solute or anti-site elements in the γ and γ ’ phases (in

Å/at.%)

γ γ ’ i

Iik I

ik IIik

Ni 0 0 -0.044

Al 0.159 1.077 0

Mo 0.405 0.819 0.042

76

Figure 4.1 Lattice misfit between γ and γ ’ in the Ni-Al binary system. (Curves present

calculated results, and symbols are reported experimental values from literature: ð[40],

◊[56])

77

Figure 4.2 Lattice misfit γ and γ ’ in Ni-Al-Mo ternary system. (Curves present

calculated results, and symbols are reported values from literature: Ο[132]; ∆[6])

78

(a) (b)

(c)

Figure 4.3 Comparison of precipitate morphologies, obtained by experiments [132] (a)

and 2D phase-field simulations (b, c) in a Ni-12.5 at.% Al-2.0 at.% Mo alloy aged at

1048K for 67h: (b) 0065.0=δ ; (c) 0035.0=δ .

79

Chapter 5

Modeling of Atomic Mobility in γ and γ ’ of Ni-Al-Mo

5.1 Background

During heat treatment and service, the average phase fraction, the composition profile

and the microstructure will change with time due to diffusion processes, affecting the

properties and lifetimes of the materials. With the increase of the computer power, the

computational experiment becomes more and more attractive, for example, many

diffusional phenomena can be computationally simulated by the DICTRA software [2].

A Ni-base kinetic database can aid us to study various processes theoretically, including

solidification, homogenization, γ ’ precipitation, boning, repairing, and protective

coatings [134].

Diffusion in the disordered γ phase has been studied widely, and a kinetic database for

γ in a 10-component Ni-base system was developed by Campbell and her coworkers

[134]. In recent years, many investigations have been devoted to the ordered

intermetallic compounds such as the L12-type and the B2-type due largely to their high-

temperature applications. For example, the intermetallic compounds based on the L12-

type Ni3Al ( γ ’ phase) and those based on the B2-type NiAl ( β phase) are major

components of the Ni-base superalloys, and their stabilities, microstructures and

distributions can significantly affect the mechanical properties of the materials.

80

Consequently, the diffusion behavior in such intermetallic phases attracts many scientific

interests from both the theory and the experiment.

In this chapter, we evaluate the diffusion mobilities for the γ and γ ’ phases in Ni-Al-Mo

system and develop the related kinetic database. At first, we introduce the model to

represent diffusion data, and then the available kinetic databases for the γ phase are

verified and revised with available experimental data from the literature. Using those

diffusion mobilities of the γ phase, the descriptions for chemical ordering of the γ ’

phase in Ni-Al binary and Ni-Al-Mo ternary systems are consequently extracted from the

available experimental diffusion data of the γ ’ phase.

5.2 Model

A model to represent diffusion data was provided by Andersson and Agren [7] and

modified by Jonsson [135], which suggests that the atomic mobility iM ( i for all

elements in the phase of interest) should be modeled, and can be described by a

frequency factor 0iM and an activation energy iQ

=RTQ

MRT

M iii exp

1 0 (5.1)

where R is the gas constant and T is the temperature. 0iM and iQ can be combined into

one parameter, i∆Φ [134]

81

∆Φ−

=RTRT

M ii exp

1 (5.2)

where 0ln iii MRTQ −=∆Φ .

With atomic mobility iM , all types of diffusivities can be calculated. For the mono-

vacancy diffusion mechanism in a substitutional sublattice, the relation between the tracer

diffusion coefficient *iD of an element i and its mobility iM is given as

ii RTMD =* (5.3)

And the chemical diffusion coefficient of k in the gradient of j and with n as the

reference specie, nkjD , can be expressed as

( )∑

∂−

∂∂

−=i n

j

j

iiikik

nkj

xxMxxD

µµδ (5.4)

where the Kronecker delta 1=ikδ when ki = and 0 otherwise. kx is the mole fraction

of the element k . The derivatives of the chemical potential iµ can be calculated from

the corresponding thermodynamic database.

For disordered solution phases, Agren and his co-workers [7, 135, 136] expressed the

composition and temperature dependence of the activation energy disi∆Φ by the Redlich-

Kister polynomial [21]

( )∑∑ ∑∑> =

−∆Φ+∆Φ=∆Φj jk

n

r

rkj

kji

rkj

l

lil

disi xxxxx

0

, (5.5)

82

where li∆Φ is the activation energy of element i in pure l surroundings, which is only

temperature dependent, and the interaction parameters kji

r ,∆Φ indicate the effect of the

kj − interaction on component i .

To describe the effect of chemical ordering, Helander and Agren [8] suggested the

following expression for the activation energy

ordi

disii ∆Φ+∆Φ=∆Φ (5.6)

where the contribution to the activation energy from chemical ordering ordi∆Φ is given by

∑∑≠

−∆Φ=∆Φj jk

kjkjordijk

ordi xxyy )( βα (5.7)

where αjy is the site fraction of component j in the sublattice α , and ord

ijk∆Φ the extra

energy for i due to the chemical ordering of the kj − atoms. This model had been

successfully applied to the B2 phase, an order bcc structure, in the binary Al-Ni and Al-

Fe systems [8] and the ternary Al-Fe-Ni system [137].

5.3 Modeling of Atomic Mobility in γ

5.3.1 Ni-Al System

The mobility database for the γ phase in the Ni-Al system was developed by Engstrom

and Agren [63] from experimental diffusivity data from Yamamoto et al. [138] with the

thermodynamic factors from the thermodynamic database from Ansara et al. [62]. In

83

recent years, the thermodynamic database for Ni-Al was updated [61], and new

experimental diffusion data were reported [139, 140] for γ in Ni-Al alloys. In Figure 5.1,

the chemical diffusion coefficients for the γ phase in the Ni-Al system are calculated

using the new thermodynamic database [61], and almost all of available experimental

data including those obtained in recent few years [139, 140] can still be well reproduced

by the mobility database from Engstrom and Agren [63].

5.3.2 Ni-Mo and Al-Mo Systems

The atomic mobilities in the γ phase of Ni-Mo and Al-Mo alloys were evaluated by

Campbell and her coworkers [134], and the results need to be revised in the current work

due to following reasons:

• Campbell’s work only considered limited experimental results, and many

important experimental data available in the literature weren’t taken into account;

• Campbell et al. [134] used NIST Ni-SuperAlloy thermodynamic database to

determine the thermodynamic factors, which is different from what we use in the

current work [131].

Campbell et al. [134] derived their Ni-Mo kinetic database only from Davin’s work [141]

and Swalin’s work [142], and both work measured the diffusivities in samples of very

low Mo concentrations. The chemical diffusivities with higher Mo concentrations were

reported by Heijwegen and Rieck ( 15.0=Mox ) [143] and Minamino et al. ( 03.0=Mox )

84

[144] at different temperatures. Using the Boltzmann-Matano method [145, 146], the

compositional dependence of the chemical diffusivity in γ was measured by Shueh et al.

[147] in Ni-Mo diffusion couples at 1323K. Taking into account all above experimental

data, the kinetic descriptions for Ni-Mo alloys were re-evaluated using the PARROT

module of the DICTRA software [2], and are listed in Table 5.1. The chemical

diffusivities in the γ phase with different Mo contents are calculated from Campbell’s

database [134] and ours, and compared with the experimental data in Figure 5.2. It was

observed that the chemical diffusivity decreases with the increase of Mo concentration,

and such a trend can be well reproduced by our database but not by Campbell’s database

[134]. As shown in Figure 5.3, the temperature dependence calculated from our database

agrees with the experimental data reported by Heijwegen and Rieck ( 15.0=Mox ) [143]

and Minamino et al. ( 03.0=Mox ) [144].

Since the solid solubility of Mo in Al is very small, diffusion measurements for the γ

phase can only be performed for very low Mo concentrations. The diffusivity of Mo in

Al was determined by Chi and Bergner [148] and Chang and Loretto [149]. Three

compositions were investigated (Al-0.05 at%Mo and Al-0.08 at%Mo by Chi and Bergner

[148], and Al-0.7 at% Mo by Chang and Loretto [149]), and the results are almost

identical, suggesting a very low concentration dependence of the Mo diffusion

coefficients in Al. The tracer diffusivity of Mo in Al was measued by Paul and Agarwala

[150] using 99Mo, and their data are obviously higher than those from Chi and Bergner

[148] and Chang and Loretto [149] (see Figure 5.4). Chang and Loretto doubted the

85

reliability of the data from Paul and Agarwala because those early works were usually

complicated by the presence of a very wider surface layer in which the concentration of

the transition metal did not conform to the expected diffusion profile [149]. Based on the

data from Chi and Bergner [148] and Chang and Loretto [149], the atomic mobility of

Mo in Al was estimated, and the calculated diffusion coefficients of Mo in Al are

compared with the experimental data in Figure 5.4.

5.4 Modeling of Atomic Mobility in γ ’

5.4.1 Ni-Al System

5.4.1.1 Experimental Data

Several tracer diffusion experiments [151-155] for Ni in the γ ’ phase have been carried

out, and most of them were focused on the stoichiometric compound Ni3Al. In the higher

temperature range (>1173K), the reported tracer diffusivities of Ni in the γ ’ phase from

different investigations [151-155] agree with each other within the experimental

uncertainty. The diffusivity data of a wide temperature range were measured by Frank et

al. [155] and Hoshino et al. [153]. Frank’s low-temperature data [155] keep the same

linear Arrhenius behavior as their high-temperature values, while Hoshino’s results [153]

reveal an enhanced diffusivity at low temperatures. Since Frank et al. [155] used the

single crystal while Hoshino et al. [153] used the polycrystals, the discrepancy is possibly

86

caused by the grain boundary diffusion. Because of the difficulty in the preparation of

the nuclide 26Al [156], no direct diffusion measurement of Al in the γ ’ phase was done

before, and then no reliable tracer diffusivity data for Al in the γ ’ phase are available.

Most of the chemical diffusivities data are determined experimentally by the diffusion

couple technique and the Boltzmann-Matano [145, 146] or similar methods. The inter-

diffusions of Ni-Al alloys in the temperature range from 1073K to 1473K have been

studied twice by Watanabe et al. [139, 157, 158], and the concentration profiles were

determined by the analytical electron microscope (AEM) and the electron probe

microanalysis (EPMA). In their first investigation [157], only two-phase diffusion

couples (γ / γ ’) were studied, and the chemical diffusivities were determined by the

Boltzmann-Matano method [145, 146]. In their later work [139, 158], they investigated

the diffusions in the single-phase (γ / γ , γ ’/γ ’ and β / β ), two-phase (γ / γ ’ and γ ’/ β )

and three-phase (γ / γ ’/ β ) couples, and applied the Sauer-Freise method [159] to the

diffusivity extraction in order to consider the variation in volume with the composition.

According to their report [139], at the lower temperatures, the chemical diffusivities in

the γ ’ phase is dependent on the types of the diffusion couples, and the values measured

from the three-phase couples were evidently greater (even one order higher at 1173K)

than those measured from the single-phase and two-phase couples. The more complex

the diffusion couple is, the greater uncertainty is expected, for example, the wavy γ / γ ’

and γ ’/ β interfaces were found in the low-temperature annealed three-phase couples

[158], which can bring additional measurement errors. Thus their data determined from

87

the three-phase couples are not used to evaluate the model parameters in the present

work. Fujiwara and Horita [160] have investigated the diffusion behavior of Al and Ni in

γ ’ phase and measured the chemical diffusivities between 1423K and 1523K by the

Sauer-Freise method [159]. They used the three-phase (γ / γ ’/ β ) couples, and the

concentration profiles were determined by EPMA. Their measurements were carried out

at quite higher temperatures, and Watanabe et al. [139] found the type of the diffusion

couple has no significant effect on the measured diffusivities at the higher temperatures,

so the data reported by Fujiwara and Horita [160] are also selected for the evaluation of

the model parameters. Janssen [161] and Shankar and Seigle [162] made inter-diffusion

experiments using the three-phase ( γ / γ ’/ β ) couples and measured the chemical

diffusivity in Ni3Al. In all of their experiments, the thicknesses of the Ni3Al layers are

not large enough for accurate measurements [163], and also their results are much higher

than others [139, 157, 158, 160, 163, 164], so their results are not used in the present

work.

Since most of the experiments used the polycrystal materials, no diffusivity data below

1000 K are used in the present assessment to avoid the effect of the grain boundary

diffusion.

5.4.1.2 Parameter Evaluation

88

To extract the ordering effect from the experimental data, both the kinetic description for

the corresponding disordered phase and the corresponding thermodynamic description

are needed. In this work, the thermodynamic database for Al-Ni system from Dupin et al.

[61, 62] and the mobility database for γ phase of Al-Ni alloys from Engstrom and Agren

[63] are used to evaluate the model parameters ordijk∆Φ . The assessed model parameters

are listed in Table 5.2 as assessment I. Comparison of the calculated and experimental

diffusion coefficients are shown in Figures 5.5-5.7, and most of the experimental data are

well reproduced.

Since there are only two types of experimental data, i.e. tracer diffusivities of Ni and

chemical diffusivities in a small composition range (22-29 at% Al), available, they may

not enough to distinguish four independent parameters ( ordAlNiAl∆Φ , ord

NiNiAl∆Φ , ordAlAlNi∆Φ and

ordNiAlNi∆Φ ) at the same time. A study on diffusion mechanism can help us to reduce the

number of independent parameters.

5.4.1.3 Diffusion Mechanism

The L12 structure is an ordered fcc, where different atoms have different preferences to

occupy the face centers or the corners. Taking the fully ordered Ni3Al as an example, the

Ni atoms occupy the face centers with 8 Ni and 4 Al atoms in the nearest neighboring

sites, and the corner positions are occupied by the Al atoms surrounded by 12 Ni atoms.

89

Unlike the diffusion in the disordered phase, the diffusion in the ordered phase is

restricted to retain the ordered structure, which complicates the diffusion mechanisms.

Ni diffusion in Ni3Al was studied by tracer diffusion experiments, and found to be

controlled by the intra-sublattice mechanism, i.e., the Ni atoms diffuse in their own

sublattice [64, 165]. Without direct measurements, the diffusion mechanism of Al in

Ni3Al is still uncertain, and several possible explanations have been proposed, namely the

six-jump cycle (both bent and straight) [166], anti-site (i.e., only Al atoms distributed

over Ni sublattice contribute to the long-range diffusion of Al) [64], anti-structural bridge

(i.e., Al atoms jump between Ni and Al sublattices) [167] and triple defect [168]

mechanisms.

Using the assessed mobility parameters, the calculated trace diffusivity of Al in Ni3Al

ordered phase is shown in Figure 5.8. The concentration of the anti-site Al in Ni3Al ( 'γ )

ordered phases calculated by the thermodynamic database from Dupin et al. [61, 62] is

shown in Figure 5.9. It was observed from Figure 5.8 and Figure 5.9 that the trace

diffusion coefficient of Al is proportional to the concentration of anti-site Al atom, which

supports the anti-site mechanism.

To further ascertain the diffusion mechanism of Al in Ni3Al, the vacancy formation

energies in different sublattice were calculated by first-principles approach. A supercell

90

of 108 atoms were used, and the ultrasoft pseudopotentials was adopted. The energies of

vacancy formation in Al and Ni sublattices, AlVaE∆ and

NiVaE∆ , can be calculated as

12122781

122681

LAl

LAlNi

LVaAlNiVa EEEE

Al+−=∆ (5.8)

12122781

122780

LNi

LAlNi

LVaAlNiVa EEEE

Ni+−=∆ (5.9)

where 122781

LAlNiE , 12

2681L

VaAlNiE and 122780

LVaAlNiE are the total energies of a 108-site L12 structure

with no defect, one Al-site vacancy and one Ni-site vacancy, respectively. 12LAlE ( 12L

NiE ) is

the total energy of a Al (Ni) atom on its own sublattice in the defect-free stoichiometric

alloy. The results from first-principles calculations are listed in Table 5.3. Obviously,

the formation energy of vacancy in Al-sublattice is significantly higher than that in Ni-

sublattice, which means the vacancy is difficult to form in the Al-sublattice. Thus the

anti-site diffusion mechanism is the most preferable diffusion mechanism, which will

dominate the diffusion of Al in the Ni3Al ordered phase.

5.4.1.4 Constraint on Model Parameters from Diffusion Mechanism

Assuming the vacancy mechanism, the self or trace diffusivities can be described by the

atom movement theory. The self-diffusivity of Ni in γ -Ni, NiNiD , can written as:

NiNiNiVNi

NiNi fCaD 0

2 ω= (5.10)

where a is the lattice parameter, VC is the probability of the vacancy, ω is the vacancy

jump frequency, and f is the correlation factor. The superscript indicates the phase, and

Ni here means all of those quantities are for γ -Ni. According to the five jump-frequency

91

model for impurity diffusion in dilute alloys [169], there are five types of vacancy jump

frequencies. 0ω is the jump frequency of a vacancy free from the effect of impurity, 1ω

is the frequency of the jump of a vacancy from a nearest-neighbor position of the

impurity to another nearest-neighbor position, 2ω is the exchange frequency of a

vacancy-impurity pair, and 3ω and 4ω are the frequencies of the dissociative and

associative jumps of a vacancy-solvent pair.

According to the five jump-frequency model [170], the trace diffusivity of Al in γ -Ni,

NiAlD , can expressed as:

NiNiNi

NiNi

VNiNiAl fCaD 2

3

42 ωωω

= (5.11)

Based on the intra-sublattice mechanism, the tracer diffusivity of Ni in the Ni3Al ordered

phase, AlNiNiD 3 , can be written as:

AlNiAlNiAlNiVAlNi

AlNiNi fCaD 33

032

33

32 ω= (5.12)

All quantities with a superscript Ni3Al are for the Ni3Al ordered phase.

For the anti-site mechanism, the tracer diffusivity of Al in the Ni3Al ordered phase,

AlNiAlD 3 , can be expressed as:

NiAl

AlNiAlNiAlNi

AlNiAlNi

VAlNiAlNi

Al PfCaD 3323

3

3432

33

32 ω

ωω

= (5.13)

92

where the anti-site factor NiAlP is defined as:

Al

NiAlNi

Al xy

P = (5.14)

which can be calculated from thermodynamic database.

The frequency product )/)(/( 0234 ωωωω , reflecting the impurity-vacancy and impurity-

matrix interactions, can be assumed to be equal for the diffusion of Al in Ni3Al and Ni

[64], thus from Equations 5.11-5.14

NiAlAlNi

Ni

NiNi

NiAl

AlNiAl P

DD

DD ≈3

3

(5.15)

which means the product of diffusion coefficients (left side) is decided by a

thermodynamic quantity (right side).

According to the Dictra modeling on the self and trace diffusivities (Equations 5.3, 5.5

and 5.6), the left side of Equation 5.15 can be re-written as

)'

exp(3

3

RTDD

DD ord

NiordAl

AlNiNi

NiNi

NiAl

AlNiAl ∆Φ−∆Φ−∆Φ

−= (5.16)

where disNi

NiAl

NiNi

disAl ∆Φ−∆Φ−∆Φ+∆Φ=∆Φ' , which can be calculated from the mobility

descriptions of the γ phase.

By considering the following reaction:

)( )()()1( 25.075.0325.075.023 fccAlNiAlNiLAlNi ↔

93

the anti-site factor NiAlP can be written as

)exp(RTG

fxy

Preact

fccAl

NiAlNi

Al

∆−≈≈= (5.17)

where fccf indicates the molar fraction of the γ phase, and reactG∆ is the energy for the

disordering of the 'γ phase, which can be calculated by the thermodynamic database.

Combining Equations 5.7 and 5.15-5.17, we get:

'')(*75.0*25.0

)(*)75.0*25.01(

GG reactordNiAlNi

ordAlAlNi

ordNiNiAl

ordAlNiAl

∆=∆Φ−∆=∆Φ−∆Φ

−∆Φ−∆Φ− (5.18)

Using the constraint from diffusion mechanisms (Equation 5.16), the description of the

mobility in Ni3Al ordered phase was refined. The re-assessed mobility parameters are

also listed in Table 5.2 and named assessment II. The agreement between the calculated

diffusivities and the experimental data shown in Figures 5.10-5.12 is good, which means

four independent parameters in assessment I are redundant.

5.4.2 Ni-Al-Mo System

The impurity diffusivities of Mo in γ ’ were determined by Minamino et al. [144] using a

(Ni-24.9 at%Al)/(Ni-23.0 at%Al-1.99 at%Al) diffusion couple. The concentration

profiles were measured by EPMA and the diffusion coefficients were obtained by the

Hall’s method [171].

94

According to Equation 5.7, it requires 18 model parameters to determine the effect of

chemical ordering in a ternary system. Besides the fours parameters in Ni-Al system

(Table 5.2), there are 14 extra parameters to be decided. Due to the lack of experimental

data, the number of independent parameters (freedoms) must be reduced before the

optimization process. Because the Mo atoms prefers to occupy Al-sites in the Ni3Al

ordered phase [144], we assume

• The effect of Al-Mo ordering can be ignored. ( 0** =∆Φ=∆Φ ordMoAl

ordAlMo )

• The Ni-Mo ordering and the Ni-Al ordering have similar effects.

( ordNiAl

ordNiMo ** ∆Φ=∆Φ and ord

AlNiordMoNi ** ∆Φ=∆Φ )

• The diffusion of Mo in the 'γ phase is similar to that of Al. (Equation 5.18)

Thus, only one independent parameter is left and can be derived from the experimental

data [144]. The thermodynamic factors were calculated from the thermodynamic

database developed by our group [131] and the mobility database constructed in Section

5.3 were used to separate the disordering part ( disi∆Φ ). The model parameters for

describing ordering effect in the Ni-Al-Mo system are listed in Table 5.4, and the

calculated diffusivity of Mo in Ni3Al are shown in Figure 5.13.

5.5 Summary

Atomic mobility in disordered γ and ordered γ ’ phases is modeled in this chapter for the

Ni-Al-Mo ternary system. For the γ phase, model parameters in the Ni-Mo and Al-Mo

95

binaries are evaluated from the experimental data in the literature, and the previous

modeling in the Ni-Al system is compared with recent experimental data. By combining

the above binary results, the kinetic description in γ is obtained for the Ni-Al-Mo system.

For the 'γ phase, the effect of chemical ordering on atomic mobility is described by a

phenomenological model [8]. The available experimental data for Ni3Al are used to

evaluated model parameters. The diffusion of Al in 'γ is simulated, indicating the anti-

site diffusion mechanism being dominant. The atomic mobility modeling of Al is then

refined based on the anti-site mechanism. By assuming the Mo atoms behave similarly

as the Al atom in 'γ , the atomic mobilities for 'γ in Ni-Al-Mo alloys are derived.

96

Table 5.1 Assessed mobility parameters of the γ phase in the Ni-Mo and Al-Mo systems,

(J/mole).

Ni-Mo

TMoMo ×−−=∆Φ 5.81254975 ** TNi

Mo ×−−=∆Φ 5.79267585 **

TMoNi ×−−=∆Φ 5.81286062 ** TNi

Ni ×−−=∆Φ 8.69287000 *

92958,0 −=∆Φ NiMoMo 138563,1 +=∆Φ NiMo

Mo

Al-Mo

TAlAl ×−−=∆Φ 1.72142000 * TMo

Al ×−−=∆Φ 8.59284000 **

TAlMo ×−−=∆Φ 0.66239668

*: from Ref. [63]

**: from Ref. [134]

Table 5.2 Assessed mobility parameters of the 'γ phase in the Al-Ni system, (J/mole).

ordAlNiAl∆Φ ord

NiNiAl∆Φ ordAlAlNi∆Φ ord

NiAlNi∆Φ

Assessment I -1.8804e+05 +9.3253e+04 -6.0418e+05 +5.8868e+05

Assessment II -9.5056e+04 +7.7234e+04 -2.4800e+05 +5.2878e+05

97

Table 5.3 Formation energy of vacancy in the Ni3Al ordered phase.

12LE (eV) VaE∆ (eV)

Ni81Al27 -590.530 -

Al -5.116* -

Ni -5.586* -

Ni81Al26Va -583.347 2.067

Ni80VaAl27 -583.491 1.454

*: see Appendix C.

Table 5.4 Model parameters for chemical ordering of 'γ in Ni-Al-Mo, (J/mole).

5104800.2 ×−=∆Φ=∆Φ ord

AlMoNiordAlAlNi

4105056.9 ×−=∆Φ=∆Φ ordAlNiMo

ordAlNiAl

5106236.6 ×−=∆Φ=∆Φ ordMoMoNi

ordMoAlNi

5103149.2 ×−=∆Φ=∆Φ ordMoNiMo

ordMoNiAl

5102878.5 ×+=∆Φ=∆Φ ordNiMoNi

ordNiAlNi

4107234.7 ×+=∆Φ=∆Φ ordNiNiMo

ordNiNiAl

0=∆Φ=∆Φ=∆Φ ordNiAlMo

ordMoAlMo

ordAlAlMo

0=∆Φ=∆Φ=∆Φ ordNiMoAl

ordMoMoAl

ordAlMoAl

98

Figure 5.1 Chemical diffusivity in the γ phase for Ni-Al as a function of Al composition.

The symbols are experimental data (MÚC�8�-[138], B7,![139], 4� [140]),

and the solid line are calculated from the mobility database developed by Engstrom and

Agren [63].

99

Figure 5.2 Compositional dependence of the chemical diffusivity of the γ phase in Ni-

Mo. The solid line is calculated from our database, and the dashed one from Campbell’s

database [134]. The symbols are experimental data (M[144], 8[147], "[143], -[141],

C[142]).

100

Figure 5.3 Chemical diffusivities of the γ phase in Ni-15 at%Mo and Ni-3 at%Mo as a

function of inverse temperature. The solid lines are calculated from our database, and the

dashed ones from Campbell’s database [134]. The symbols are experimental data

(M[144], "[143]).

101

Figure 5.4 Diffusion coefficient of Mo in Al as a function of inverse temperature. The

solid line is calculated from our database, and the dashed one from Campbell’s database

[134]. The symbols are experimental data (-[149], M"[148], �[150]).

102

Figure 5.5 Tracer diffusivities of Ni in the stoichiometric Ni3Al. The solid line is

calculated from the assessment I, and the symbols present experimental data (+[154];

×[153]; Ú[152]).

103

Figure 5.6 Chemical diffusivities in the stoichiometric Ni3Al. The solid line is calculated

from the assessment I, and the symbols present experimental data (-[160]; , [139]; M

[161]).

104

Figure 5.7 Calculated chemical diffusivity (assessment I) compared with the

experimental data from Fujiwara and Horita [160] (C: 1523K; 8: 1473K; -: 1423K)

and Watanabe et al. [139] (B: 1473K; 7: 1373K; ,: 1273K; !: 1173K).

105

Figure 5.8 Calculated tracer diffusivities of Al in Ni3Al ( 'γ ) ordered phases at 1473K.

106

Figure 5.9 Concentration of the anti-site Al in Ni3Al ( 'γ ) ordered phases at 1473K.

107

Figure 5.10 Tracer diffusivities of Ni in the stoichiometric Ni3Al. The solid line is

calculated from the assessment II, and the symbols present experimental data (+[154];

×[153]; Ú[152]).

.

108

Figure 5.11 Chemical diffusivities in the stoichiometric Ni3Al. The solid line is

calculated from the assessment II, and the symbols present experimental data (-[160];

, [139]; M [161]).

109

Figure 5.12 Calculated chemical diffusivity (assessment II) compared with the

experimental data from Fujiwara and Horita [160] (C: 1523K; 8: 1473K; -: 1423K)

and Watanabe et al. [139] (B: 1473K; 7: 1373K; ,: 1273K; !: 1173K).

110

Figure 5.13 Diffusion coefficient of Mo in Ni3Al. The symbols are experimental data

(M[144])

111

Chapter 6

Coarsening Kinetics of γ ’ Precipitates in Ni-Al-Mo System

6.1 Background

In the previous chapters, we determined the lattice parameters and mobilities in the γ

and γ ’ phases of Ni-Al-Mo alloys, and the corresponding property databases were

developed. In this chapter, we investigate the microstructural evolution and the

coarsening kinetics of γ ’ precipitates in the Ni-Al-Mo ternary system using two-

dimensional phase-field simulations. For the coarsening kinetics of γ ’ precipitates, there

have been a number of experimental measurements [5, 6, 41, 132] in Ni-Al-Mo alloys

and a few theoretical predictions in Ni-Al alloys [133, 172, 173]. In this work, we link

the phase-field simulations with thermodynamic, kinetic and lattice parameter databases

developed by our group. The simulations are performed in experimental time and length

scales, and the results can compare with experimental data quantitatively. Such

theoretical predictions provide a method to systematically study the effects of Mo

concentrations on morphological evolution and coarsening kinetics. The effect of the

volume fraction of the precipitates on coarsening kinetics is also a topic of discussion in

this chapter.

6.2 Simulation Details

112

6.2.1 Model

In the phase-field model, all phases or domains in a microstructure are characterized by a

set of field variables, e.g. compositions and order parameters, which change continuously

in the interface regions. To distinguish the disordered γ phase and the ordered γ ’ phase

with four types of ordered domains in Ni-Al-Mo ternary alloys, two composition

variables ( )tci ,r ( MoAli ,= ) and four artificial order parameters ( )tj ,rη ( 4,3,2,1=j )

are employed, which vary spatially (r ) and temporally (t ). The temporal evolution of

these field variables is described by the Cahn-Hilliard and Allen-Cahn (or Ginzburg-

Lauder) equations [4]

( )( )

∇∇=

∂∂

tcF

Mt

tc

mim

i

,,

rr

δδ

(6.1)

( )( )tF

Lt

t

njn

j

,

,

r

rδη

δη−=

∂∂

(6.2)

where F is the total free energy of the microstructure, imM is the diffusion mobility of i

with respect to the concentration gradient of element m and jnL is the kinetic coefficient

for the relaxation of the order parameter j with respect to the gradient of the order

parameter n .

The total free energy of an inhomogeneous microstructure can be described by the field

variables as [133]

113

( ) ( ) VcfFv

jj

jji d

2,

4

1

2∫ ∑

∇+=

=

ηβ

η (6.3)

where jβ is the gradient energy coefficient of the order parameter jη , and ( )jicf η,

denotes the local free energy density of the system expressed by [174]

[ ] elj

mi

mj

pi

pjji egwcfhcfhcf ++−+= )()()(1)()(),( 0 ηηηη (6.4)

Where ele is the elastic energy density, and 0w is the double-well potential height. The

chemical free energy densities of the precipitate and the matrix, pf and mf , are

obtained from the Ni-Al-Mo thermodynamic database [131]. The double-well potential

( )jg η and the separation function ( )jh η are selected as [133]

( ) ( )[ ]∑=

−=4

1

22 1j

jjjg ηηη (6.5)

( ) ( )[ ]∑=

+−=4

1

22 10156j

jjjjh ηηηη (6.6)

In this approach, the interface region is treated as a mixture of the γ matrix and the γ ’

precipitate with different compositions but equal chemical driving forces

( ) ( )

( ) ( )[ ]

−+=

∂∂

=∂

mij

piji

mi

mi

m

pi

pi

p

chchc

ccf

ccf

ηη 1

(6.7)

The elastic energy contribution arises from the lattice misfit between the γ and γ ’

phases. Assuming that the lattice parameter is a weighted average of the lattice

114

parameters of the γ and γ ’ phases with ( )jh η as the weighing factor, the local stress-

free strain 0klε is given by [175]

( )jklkl hηδεε 00 = (6.8)

where 0ε is the stress-free lattice misfit, and klδ the Kronecher-delta function. If there is

no macroscopic change in shape or volume, the elastic strain elklε can be written as

0klkl

elkl εδεε −= (6.9)

where )//(5.0 kllkkl ruru ∂∂+∂∂=δε is the local strain. The vector u represents the

local displacement field, which can be solved by using the Hooke’s law ( elklijklij C εσ = ,

where ijklC refers to the elastic constants), the local mechanical equilibrium condition

( 0/ =∂∂ jij rσ ), and Equations 6.8 and 6.9. After the displacement field is obtained, the

elastic strain and stress can be determined, and then the elastic energy density ele can be

evaluated by

elklkl

ele εσ21= (6.10)

The diffusion mobility imM in the Cahn-Hilliard equation (Equation 6.1) is expressed by

the following equation (see Appendix B)

∑=

−−=n

jjjijimmjim McccM

1

]][[ δδ (6.11)

where jM is the atomic mobility of element j , which are obtained from the atomic

mobility database for Ni-Al-Mo system [176]. The kinetic coefficient L in the Allen-

115

Cahn equation (Equation 6.2) is related to the interface mobility and is not well-

determined because the interface mobility is not available. Since an accurate value of L

is not necessary for a diffusion-controlled process [133], a constant value for L is

assigned. By comparing the simulated results of different L values, we used

sJmL /001.0= in this work because a larger L will not change the results any more,

which means sJm /001.0 is large enough and the simulated coarsening of the γ ’

precipitate is controlled by the diffusion process.

6.2.2 Conditions and Parameters for Simulations

To compare with experimental results [132], phase-field simulations at a temperature of

1048K were performed using a 512 × 512 grid with a unit grid size of 2nm. Three alloys

with different Mo concentrations were selected for simulations and their overall

compositions are shown in Figure 6.1 as solid circles, and the dashed lines are the tie-

lines for the three alloys calculated from the thermodynamic database [131]. More alloys

with different volume fractions of precipitates are selected along the tie-lines (open

symbols in Figure 6.1). The initial states were homogeneous solutions with small

composition fluctuations around the average compositions. West and Kirkwood [177]

observed that the maximum precipitate density (about 323 /10 m ) of γ ’ precipitates was

reached in a few seconds in Ni-Al alloys at 1063K. Consequently, approximate 1800

nuclei were introduced at an early stage of simulations ( st 10< ), and after that the

116

nucleation process was turned off. All nuclei were circles with an average radius of 6nm

and randomly distributed.

One of the critical factors that control the morphology of coherent γ ’ precipitates is the

magnitude and sign of the stress-free lattice misfit between γ and γ ’, very sensitive to

the details of experimental processing [86]. In Section 4, we proposed an integrated

computational approach for evaluating the lattice misfit between γ and γ ’ in Ni-base

superalloys by combining first-principles calculations, existing experimental data and

phenomenological modeling, and the values for the current simulations are shown in

Figure 6.2.

The interfacial energy is another essential parameter for phase-field simulations. Due to

the lack of data, a value of 13.5 mJ/m2 from binary Ni-Al alloys [178] was used in the

present work, in line with the value of 12 mJ/m2 for superalloy Nimonic 80a (Ni-Cr-Al-

Ti) reported by Zickel et al. [179]. By fitting to the interfacial energy, the gradient

energy coefficients jβ and the double-well potential height w can be determined for a

given interface width. Since the microstructure is described by artificial order parameters

jη in this model, an interface thickness wider than the actual physical one can be used

[133]. With the interface width being 5nm, jβ and 0w are mJ /100.9 11−× and

37 /105.3 mJ× , respectively.

117

The temperature and composition dependences of the elastic constants of γ and γ ’ in

Ni-Al binary alloys were studied by Prikhodko and his co-workers [180, 181], and only

few Ni-Al-Mo samples were measured in a very narrow temperature range [15, 182].

The elastic constants for current alloys were estimated from the above information and

listed in Table 6.1. Furthermore, the elastic homogeneity was assumed due to the lack of

data and the small difference between the γ and γ ’ elastic constants [133].

6.3 Results and Discussion

6.3.1 Microstructure Evolution

The 2D microstructures at different annealing times are shown in Figure 6.3, and a

comparison between the simulated and the experimental micrographs after 67h ageing is

also provided. The precipitate sizes from simulations are somewhat smaller than that

from experiments. One of the reasons could be due to the 2D nature of simulations since

the coarsening in 2D is slower than 3D because of the reduced curvature. As shown in

Figure 6.3, the simulated particle morphology is very similar to the experimental

observations for all three alloys. From A1 to A3, the magnitude of the lattice misfit

decreases with the increase of the Mo concentration (see Figure 6.2), resulting in more

circle-like morphology due to the low elastic energy. For A1 and A2 alloys of large

lattice misfit, the γ ’ particles gradually change their shapes from circle to rectangle with

the annealing time, while the particles in A3 of small misfit keep a circular morphology

118

for all the time. For all three alloys, the coalescence is observed between neighboring

domains of same order parameter, and such a phenomenon was also reported by the

previous phase-field simulations in Ni-Al alloys [133]. The particle alignment along

<10> direction of the γ matrix is also found in Figure 6.3, and the intensity depends on

the misfit strain. After 67h annealing, the γ ’ precipitates in A1 alloy ( 0036.00 =ε ) is

clearly aligned, and the degree of alignment in A2 alloy ( 0025.00 =ε ) is lower.

6.3.2 Coarsening Kinetics

The classical theory of coarsening process developed by Lifshitz and Slyozov [183] and

Wagner [184] (LSW) indicates that the average particle radius R obeys the following

temporal power law

( )ss ttKRR −=− 33 (6.12)

where K is the coarsening rate constant, and sR and st refer to the average particle size

and the time at the beginning of the steady state coarsening, respectively. To avoid the

ambiguity in determining the exact onset of the steady state coarsening, Equation 6.12

can be rewrite as

KtRR += 30

3 (6.13)

where ss KtRR −= 330 [133]. The coarsening rate constant K is described, under LSW

theory, as a function of the interfacial energy, σ , the diffusion coefficient in the matrix

119

phase, mD , and the second derivate of the free energy of the matrix phase with respect to

the concentration in the matrix 22 / mm cf ∂∂

( ) 2

22,,9

8

m

meqmeqp

mp

LSW

c

fcc

DVK

∂−

(6.14)

where pV is the molar volume of the precipitate phase, and eqmc , and eqpc , are the

equilibrium compositions of the matrix and the precipitate, respectively.

The relationship between the average particle size and annealing time at 1048K is plotted

in Figure 6.4. The average particle sizes at different annealing times were obtained by

following the same procedures used in the experiments [6, 132]. To compare the particle

coarsening rates for different alloys of different particle shapes, the radius of an area-

equivalent circle was assigned to each particle. In order to obtain accurate values for the

average particle size R , three independent run of simulations were performed for each

alloy. The linear relationships between 3R and t are observed for all three alloys, and

the coarsening constants K were obtained from a linear regression procedure for

Equation 6.13.

The coarsening rate constants from the simulations at 1048K are compared with the

experimental values [132] in Figure 6.5. Note that the experimental results from ref. [6]

are for 1023K. As it can be seen, both simulated and experimental coarsening rate

constants decrease with the increase of the Mo concentration. One reason for this

phenomenon could be due to the reduced diffusivity (see Table 6.1). The coarsening rate

120

constants from simulations are somewhat smaller than those from experiments. As

mentioned before, the coarsening in 2D is expected to be slower than that in 3D because

of the reduced curvature. The difference between simulations and experiments (about

300nm3/h for our cases, see Figure 6.5) is not very significant since a temperature

variance from 1048K to 1023K could bring a change of that magnitude in the coarsening

rate constants as shown in Figure 6.5.

Since the LSW theory was developed by solving the diffusion equation for a particle in

an infinite matrix, and did not consider the interactions between different particles, it can

only strictly applied to the case of zero volume fraction of the precipitate. It was reported

[185-192] that the volume fraction did not affect the cubic law of coarsening, but it will

change the value of coarsening rate constant and the shape of particle size distribution

curve. Thus the coarsening rate constant K is modified to be a function of the volume

fraction of particles, φ [193]

( )φfKK LSW= (6.15)

where ( )φf is an alloy-independent function, equal to 1 for the zero volume fraction in

line with the LSW theory. In Figure 6.6, ( )φf curves from different theories are

compared, and they are quite different.

Phase-field simulations were preformed for alloys with different volume fractions of

precipitates which are selected along the A1 and A3 tie-lines (open symbols in Figure

6.1), and the results are shown in Figure 6.7. As predicted by various theories, the

121

coarsening rate constant was observed to increase as the volume fraction of precipitates

increases due to the reduced diffusion distance. The anomalous dependence for small

volume fractions (i.e., the coarsening rate constant decreases instead of increases with

volume fraction) reported by Ardell and his co-workers in Ni-Al and Ni-Ti alloys [194] is

not found in present simulations. According to Figure 6.7, the growth of the coarsening

rate constant is sluggish in the low volume fraction range and was speeded up in the high

volume fraction range. Compared with the curves in Figure 6.6, such a trend is most

likely reproduced by theories provided by Asimov [185], Davies et al [188] and Voorhees

and Glicksman [190]. Figure 6.7 also indicates that the coarsening rate constants for A3

alloys have a stronger dependence on volume fraction than those for A1 alloys. For

volume fractions less than 0.3, coarsening for A3 alloys is slower than that for A1 alloys,

while A3 alloys have higher coarsening rate than A1 alloys when the volume fraction is

over 0.4. This observation may be explained by the particle shape. It is well known that

the coarsening process is driven by the variation in interfacial curvature and realized by

mass diffusion through the matrix. A1 alloys have high diffusivities and low curvature

effects (low driving forces) due to the particle shape of rectangle, while A3 alloys have

low diffusivities and high curvature effects (high driving forces) because of their circular

particles. When the volume fraction of precipitates is low, the diffusion distance is long

and the coarsening process is controlled by diffusion in the matrix, so coarsening in A1

alloys is faster than that in A3 alloys. However, for the cases of a high volume fraction,

the particles are very close to each other and the diffusion process becomes less

important, and the coarsening rate is controlled by the curvature. Thus the coarsening

rate constants for A3 alloys are large than those for A1 alloys.

122

6.4 Summary

The microstructure evolution and coarsening kinetics of γ ’ precipitates in Ni-Al-Mo

alloys were studied by phase-field simulations in terms of experimental length and time

scales. With increasing Mo concentration, the lattice misfit between γ and γ ’ decreases,

and the shape of the γ ’ particles changes from rectangle-like to circle-like. A linear

relationship between the cube of average particle size and annealing time is observed for

the coarsening stage and the coarsening rate constant increases with the volume fraction

of precipitates. With a low volume fraction, the increase of Mo concentration slows the

coarsening process due to the reduced diffusivity. If the volume fraction is high enough,

the coarsening rate can be speeded up by a higher Mo concentration that causes a larger

curvature effect by changing the particle shape.

123

Table 6.1 Some parameters for phase-field simulations ( KT 1048= ).

Alloy 0ε (%) 11C (GPa) 11C (GPa) 11C (GPa) mAlAlD (m2/s) m

MoMoD (m2/s)

A1 0.36 213 151 99 1710232.2 −× 1810446.2 −×

A2 0.25 218 153 100 1710655.1 −× 1810460.1 −×

A3 0.00 217 151 98 1710263.1 −× 1810826.0 −×

124

Figure 6.1 Isothermal section of Ni-Al-Mo ternary phase diagram at 1048K. Symbols

show the compositions of selected samples and dotted lines present the tie-lines for those

compositions.

125

Figure 6.2 Lattice misfit between γ and γ ’ in Ni-Al-Mo ternary system at 1048K.

Symbols show the values for A1, A2 and A3 alloys.

126

20h

40h

67h

67h

Ni-12.5 at%Al-2.0 at%Mo Ni-9.9 at%Al-5.0 at%Mo Ni-7.7 at%Al-7.9 at%Mo

Figure 6.3 Microstructure evolution of the γ ’ precipitates in Ni-Al-Mo alloys at 1048K.

Figures in the bottom row are from experiments [132], and others from 2D phase-field

simulations.

127

Figure 6.4 Plot of the cube of average particle size vs annealing time at 1048K. (Ο: Ni-

12.5 at%Al-2.0 at%Mo, hnmK /397 3= ; ∇: Ni-9.9 at%Al-5.0 at%Mo, hnmK /388 3= ;

ð: Ni-7.7 at%Al-7.9 at%Mo, hnmK /257 3= )

128

Figure 6.5 Coarsening rate constant vs Mo concentration at 1048K. Open symbols are

from experiments (Ο:[132], Ni-12.5 at%Al-2.0 at%Mo, Ni-9.9 at%Al-5.0 at%Mo and Ni-

7.7 at%Al-7.9 at%Mo; ∆: [6], Ni-10.2 at%Al-5.1 at%Mo, Ni-8.2 at%Al-7.9 at%Mo and

Ni-6.5 at%Al-9.8 at%Mo) and solid ones from 2D phase-field simulations.

129

Figure 6.6 Comparison of various ( )φf functions from different theories[185-188, 190].

130

Figure 6.7 Coarsening rate constant vs volume fraction of γ ’ precipitates in Ni-Al-Mo

system at 1048K.

131

Chapter 7

Conclusions and Future Directions

7.1 Conclusions

The main contributions of the present thesis are summarized as follows:

1. The lattice parameters of precipitate and matrix phases as a function of

temperature and composition were constructed using a phenomenological model.

Pure elements were modeled under the assumption that the thermal expansion

coefficients depend on temperature linearly. The lattice parameters of

substitutional solid solution phases are treated similar to the Gibbs energy

modeling in CALPHAD. Such a phenomenological approach was successfully

applied to Ni-Al binary system by evaluating the model parameters using

experimental data.

2. Due to a lack of experimental data for in multi-component systems, an integrated

computational approach was developed for evaluating the lattice misfit between

γ and γ ’ by combining first-principles calculations, existing experimental data

and phenomenological modeling. This approach was validated by comparing the

calculated lattice misfits with available experimental measurements as well as by

132

comparing the predicted γ ’ precipitate morphologies from phase-field

simulations with experimental observations for Ni-Al-Mo alloys.

3. The effects of various alloy elements (Al, Co, Cr, Hf, Mo, Nb, Re, Ru, Ta, Ti and

W) on the lattice parameter and the local lattice distortion around the solute atom

in the γ -Ni solution were studied using first-principles calculations. It is found

that the atomic size difference, the electronic interactions, and the magnetic spin

relations between the solute and solvent atoms all contribute to the lattice

distortions. Based on the results from first-principles calculations, the linear

composition coefficients of γ -Ni lattice parameter for different solutes are

determined, and the lattice parameters of multi-component Ni-base superalloys as

a function of solute composition are predicted.

4. Diffusion in disordered γ and ordered γ ’ phases was modeled for the Ni-Al-Mo

ternary system, and an atomic mobility database was developed. For the γ phase,

atomic mobilities in the Ni-Mo and Al-Mo binaries were evaluated based on the

experimental data in the literature, and the previous modeling in the Ni-Al system

was compared with recent experimental data. By combining the above binary

results, the kinetic description in γ was obtained for the Ni-Al-Mo system. For

the γ ’ phase, the effect of chemical ordering on atomic mobility was described by

an existing phenomenological model, and the model parameters were evaluated

by the available experimental data for Ni3Al. The diffusion of Al in 'γ was

133

simulated, and the formation energies of vacancy in different sublattices were

calculated by first-principles approach, both of which indicate the anti-site

diffusion mechanism being dominant for diffusion of Al. The atomic mobility

modeling of Al was then refined based on the anti-site mechanism. By assuming

the Mo atom behaves similarly as the Al atom in 'γ , the atomic mobilities for 'γ

in Ni-Al-Mo alloys were derived.

5. The phase-field model for binary Ni-base superalloys was extended to ternary

systems and integrated with the corresponding thermodynamic, kinetic and lattice

parameter databases. Two-dimensional phase-field simulations for Ni-Al-Mo

alloys were carried out. In particular, the effect of Mo concentration on

microstructure evolution and coarsening kinetics of γ ’ precipitates is studied.

For alloys of different compositions, the morphology and average size of

precipitates were predicted as a function of annealing time, and quantitative

comparisons between simulated results and experimental data showed good

agreements. It was observed that increasing Mo content decreases diffusivities

and the lattice misfit between γ and γ ’, and thus causes the changes from

cuboidal to spherical morphology and the variety in coarsening rate of precipitate

particles.

134

7.2 Future Directions

1. It will be straightforward to apply the present approach higher order Ni-base alloy

systems, e.g. Ni-Al-Mo-Ta. Also it will be interesting to test this approach in

other alloy systems.

2. Predicting properties from microstructures obtained in the present study will be

very useful as the behavior of materials is the most important information in the

field of materials science and engineering.

135

Appendix A

Thermodynamic Descriptions for γ and γ ’ in Ni-Al-Mo System

The disordered γ solution phase is described by a substitution model, and its Gibbs

energy is represented by the following equation in terms of one mole of atoms

GxxRTGxG xs

iii

iiim ++= ∑∑ ln0 γγ (A. 1)

where ix is the mole fraction of element i ( i = Ni, Al, Mo), and T the temperature. The

γiG0 denotes the Gibbs energy of pure element i with fcc structure taken from the

Scientific Group Thermodata Europe (SGTE) database [195]. Gxs is the excess Gibbs

energy with interaction parameters expressed in Redlich-Kister polynomials [21]

∑∑ ∑>

−=i ij m

mjiji

mji

xs xxLxxG )(,γ (A. 2)

where γji

m L , is the m th-order parameter for the ji − interaction.

The γ ’ phase is described by a two-sublattice model (Ni,Al,Mo)3(Ni,Al,Mo)1. As a

ordered phase of γ , its Gibbs energy is composed of two parts

( )IIi

Ii

ordmm yyGGG ,' ∆+= γγ (A. 3)

136

where the energy of disordered state is described by γmG , and Gord∆ is the contribution of

chemical ordering, which is a function of Iiy and II

iy (site fractions of element i in the

first and second sublattices, respectively) [62]

( ) ( )IIi

Ii

ordIIi

Ii

ordord xxGyyGG ,, −=∆ (A. 4)

( )

∑∑ ∑∑

∑ ∑∑∑ ∑∑

∑∑∑

> >

>>

+

++

++∆=

i ij k kllkji

IIl

IIk

Ij

Ii

j jkikj

Ik

Ij

IIi

j jkkji

IIk

IIj

Ii

i

IIi

IIi

Ii

Ii

i jji

IIj

Ii

ord

Lyyyy

LyyyLyyy

yyyyRTGyyG

,:,

i:,

i,:

',

ln25.0ln75.0γ

(A. 5)

where ',γ

jiG∆ is the formation energy of the L12 compound, ( i )3( j )1.

Thermodynamic descriptions for γ and γ ’ in the Ni-Al-Mo system are collected from

the literatures [61, 62, 131, 195] and listed in Table A.1.

137

Table A.1 Thermodynamic properties for γ and γ ’ in Ni-Al-Mo system (in SI units)

=γAlG0

70014.298 ≤≤ T

13723 74092107766481088466.1ln367224093137157976 -T+T.TTT.T.+. −− ×−×−−−

6.933700 ≤≤ T

13622 74092107642351085320.1ln584438048223211276 -T+T.TTT.T.. −− ×−×+−+−

30006.933 ≤≤T

928102341ln748231684188411278 −×−−+− T.TT.T..

=γMoG0

289614.298 ≤≤ T

1030927.1

658121066283510443403ln564123550132707453410

13723

T

TT.T.TT.T.. -

−−

×−

+×+×−−++

50002896 ≤≤T

933108493154ln638342190284415356 −×−−+− T.TT.T..

=γNiG0

172814.298 ≤≤ T

231084070.4ln096022854117165179 TTT.T.. −×−−+−

60001728 ≤≤ T

93110127541ln143135.279727840 −×+−+− T.TT.T.

TL MoAl 2092220,0 +−=γ

TL NiAl 213.16162407,0 +−=γ TL NiAl 914.3473418,

1 −+=γ TL NiAl 837.933471,2 −+=γ

TL NiAl 253.1030758,3 +−=γ

8.4007,0 −=γ

NiMoL TL NiMo 081.83151,1 +−=γ

272947,,0 +=γ

NiMoAlL 177424,,1 +=γ

NiMoAlL 136388,,2 +=γ

NiMoAlL

0', =∆ γAlAlG 0'

, =∆ γMoMoG 0'

, =∆ γNiNiG

0', =∆ γMoAlG 0'

, =∆ γAlMoG

NiAlNiAl UG ,

1', 3=∆ γ NiAl

AlNi UG ,1

', 3=∆ γ

NiMoNiMo UG ,

1'

, 3=∆ γ NiMoMoNi UG ,

1', 3=∆ γ

138

Table A.1 Thermodynamic properties for γ and γ ’ in Ni-Al-Mo system (in SI units)

(continued)

NiAlAlNiAl UL ,

1'

:,0 6=γ NiAl

AlNiAl UL ,4

':,

1 3=γ

NiAlNiMoAlMoNiAl UUL ,

1,,'

:,0 65.1 +=γ NiAlNiMoAl

MoNiAl UUL ,4

,,':,

1 35.1 +=γ

NiAlNiNiAl UL ,

1'

:,0 6=γ NiAl

NiNiAl UL ,4

':,

1 3=γ

NiAlNiAlAl UL ,

4'

,:1 =γ NiAl

NiAlMo UL ,4

',:

1 =γ NiAlNiAlNi UL ,

4'

,:1 =γ

NiMoNiMoAlAlNiMo UUL ,

1,,'

:,0 65.1 +−=γ NiMoNiMoAl

AlNiMo UUL ,4

,,':,

1 35.1 +−=γ

NiMoMoNiMo UL ,

1'

:,0 6=γ NiMo

MoNiMo UL ,4

':,

1 3=γ

NiMoNiNiMo UL ,

1'

:,0 6=γ NiMo

NiNiMo UL ,4

':,

1 3=γ

NiMoNiMoAl UL ,

4'

,:1 =γ NiMo

NiMoMo UL ,4

',:

1 =γ NiMoNiMoNi UL ,

4'

,:1 =γ

NiMoAlAlNiMoAl UL ,,'

:,,0 5.1−=γ NiMoAl

MoNiMoAl UL ,,':,,

0 5.1−=γ NiMoAlNiNiMoAl UL ,,'

:,,0 6=γ

TU NiAl 93047.26.14808,1 +−= TU NiAl 74273.360.7203,

4 −+=

TU NiMo 255.373.390,1 −−= TU NiMo 302.310.3748,

4 −−=

TU NiMoAl 033.1220375,, −+=

139

Appendix B

Diffusion Mobility and Atomic Mobility

In the kinetic database constructed by Dictra [2], the model relating the atomic mobility

is based on the generalize Onsager flux equations [7]. In a system with n components,

the flux of component k is expressed by

∑=

∇−=n

jjkjk LJ

1

µ (B. 1)

where kjL describes the linear dependence between the flux kJ and the chemical

potential gradient jµ∇ . jµ∇ constitute n driving force for diffusion, and only 1−n of

them are independent due to the Gibbs-Duhem equation,. Taking component n as the

reference, we can write the fluxes as function of the 1−n independent forces Φ∇ ,

∑−

=

Φ∇−=1

1

''~ n

iikik LJ (B. 2)

where

nn

iii V

V µµ ∇

−∇=Φ∇ (B. 3)

where iV is the molar volume of component i . The relationship between two sets of

linear dependence constants can be derived [7]

∑∑= =

−=

n

j

n

rjr

m

jkjk

m

riirki L

V

Vx

VV

xL1 1

'' δδ (B. 4)

140

where ix refers to the mole fraction of component i . irδ is the Kronecker delta, i.e.

equals to 1 when ri = and 0 otherwise.

Assuming that the volume fractions V are the same for all components and 0, =≠ jkkjL ,

( )( ) ( )( )

( )( )∑

∑∑

=

==

−−=

−−=−−=

n

jjjkjkiij

n

jjjkjkiij

n

jjjkjkiijki

Mxxxc

McxxLxxL

1

11

''

δδ

δδδδ

(B. 5)

where jM is the atomic mobility of component j , which can be calculated directly from

the kinetic database. The concentration variable jc is defined as moles of j per unit

volume, and c is total moles of atoms in a unit volume.

Inserting (B.5) in (B.2), we obtain

( ) ( )( )[ ]∑ ∑−

= =

−−∇−∇−=1

1 1

~ n

i

n

jjjkjkiijnik MxxxcJ δδµµ (B. 6)

Comparing (B.6) with the flux description from the phase-field theory,

( )∑−

=

∇−∇−=1

1

~ n

ini

dkik McJ µµ (B. 7)

we find that the diffusion mobility dkiM used in phase-field simulations can be described

by the atomic mobilities

( )( )∑=

−−=n

jjjkjkiij

dki MxxxM

1

δδ (B. 8)

141

For binary systems, the above equation can be re-written as

)( 21122111 MxMxxxM += (B. 9)

For ternary systems, the diffusion mobilities are:

332

1222

1112

111 )1( MxxMxxMxxM ++−= (B. 10)

33212212121112 )1()1( MxxxMxxxMxxxM +−−−−= (B. 11)

1233212212121121 )1()1( MMxxxMxxxMxxxM =+−−−−= (B. 12)

332

2222

2112

222 )1( MxxMxxMxxM +−+= (B. 13)

142

Appendix C

Energies of Ni and Al in Ni3Al

To study point defects in an alloy, the total energies of various atoms in the defect-free

stoichiometric alloy will be taken as reference states. For a one-element material, that

quantity can be measured with respect to the well-defined total energy per atom in the

perfect crystal. However, those quantities are not uniquely defined in a mixture of

different atoms. A convention would be to refer to the elements in their standard states.

Although it’s widely used in chemistry, it’s strictly speaking incorrect and less

convenient for atomistic calculation [196]. An alternative method is dividing those

quantities from the total energy of the related alloy. Taking the Ni3Al compound in a

perfect L12 structure as an example, we can write the total energy per atom, 123

LAlNiE , as

1212123 75.025.0 L

NiLAl

LAlNi EEE += (C.1)

where 12LAlE ( 12L

NiE ) is the total energy of a Al (Ni) atom on its own sublattice in the

defect-free stoichiometric alloy. However, 12LAlE and 12L

NiE always occur together and can

not be calculated separately, thus the division of the total energy between Al and Ni is not

unique based on the above equation, because an arbitrary energy could be added to 12LAlE

and subtracted from 12LNiE correspondingly. Since the choice of reference states is

ambiguous, the calculated energies based on such reference states are arbitrary [197].

143

From the point of view of thermodynamics, those total energies of various atoms can be

treated as chemical potentials. Once we know the total energy 12LE as a function of the

composition Alx , 12LAlE and 12L

NiE can then be calculated directly

∂−=

Al

L

AlL

AlNiLAl

xE

xEE12

123

12 (C.2)

∂−+=

Al

L

AlL

AlNiLNi

xE

xEE12

123

12 )1( (C.3)

According to the crystal structure, the L12 Ni3Al phase can be represented by a four

sublattice model, (Al,Ni)(Al,Ni)(Al,Ni)(Al,Ni), and the four sites are equilvalent. For the

completely ordered Ni3Al, one of those sublattice is occupied by Al atoms (Al-sublattice),

and the other three are taken by Ni atoms (Ni-sublattices). The total energy is then

expressed by

( )[ ]∑∑ ∑ ∑ ∑∑∑

∑∑∑∑∑∑

≠ ≠ ≠

=

−++

+=

s st tsu utsv j k l

sNi

sAllkjlkj

sNi

sAl

vl

uk

tj

s i

si

si

i j k l

Llkjilkji

L

yyLLyyyyy

yyRT

EyyyyE

, ,,::

1::

0

4

1

12:::

432112 ln4

(C.4)

where siy is the site fraction of atom i in sublattice s . 12

:::L

lkjiE is the total energy of a L12

structured compound whose four sublattices are occupied by kji ,, and l atoms,

respectively. Due to the symmetry of the structure, the terms 12:::

LlkjiE with the same

stoichiometry are equal. lkjL :: represents the interaction between species in one sublattice

144

and the other three sublattices are taken by lkj ,, atoms, respectively. Since the four

sublattices are equilvalent, jkljlkkjlljkkljlkj LLLLLL :::::::::::: ===== .

The total energies 12:::

LlkjiE are calculated by the first-principles approach using the Vienna

ab initio simulation package VASP (Version 4.6) [87]. The ultrasoft pseudopotentials

and the generalized gradient approximation (GGA) [88] are adopted for the current

calculations. The set of k points is adapted to the size of the primitive cell, and the

energy cutoff is determined by the choice of “high accuracy” in the VASP. The

calculated results are listed in Table C.1. The mixing of species in a given sublattice are

modeled using special quasi-random structures (SQS’s) that are specially designed small-

unit-cell periodic structures with only a few atoms per unit cell to closely mimic the most

relevant, near-neighbor pair and multi-site correlation functions of the random

substitutional alloys [198-200]. The 64-atom SQS structures for substitutionally random

31 )( CBA xx − and 211 )( DCBA xx − alloys at compositions 25.0=x and 5.0=x are

developed using Alloy-Theoretic Automated Toolkit (ATAT) [201], and the structural

descriptions of those SQS structures are shown in Table C.2. The total energy of

compounds with those SQS structures are calculated using VASP, and the interaction

parameters in Equation C.4 are then determined (see Figure C.1) and listed in Table C.1.

Two kinds of relaxations are introduced for the first-principle calculations, unconstrained

relaxation and constrained relaxation. The unconstrained relaxation allows the total

energy to be minimized with respect to the volume and shape of the cell and the positions

145

of atoms within the cell, while the shape of the cell and the positions of atoms are fixed in

the constrained relaxation.

At the absolute 0K, the Ni3Al alloy is completely ordered. The composition variance

only affects the site fraction in Al-sublattice when 25.0<Alx , and it only changes the site

fractions in Ni-sublattices when 25.0>Alx . Therefore, xy ∂∂ / is discontinuous at

25.0=Alx , and then ( ) 25.012 / =∂∂

AlxAlL xE can not be calculated directly at 0K. From Al-

rich side, ( ) +=∂∂ 25.012 /

AlxAlL xE equals to 107830J, and ( ) −=∂∂ 25.0

12 /AlxAl

L xE calculated

from Ni-rich side is –17435J. When the temperature increases, even a very small amount

of increase from 0K, xy ∂∂ / and AlL xE ∂∂ /12 become continuous due to the entropy

effects, and then

∂+

∂×=

−=+== 25.0

12

25.0

12

25.0

12

5.0AlAlAl xAl

L

xAl

L

xAl

L

xE

xE

xE (C.5)

Using the above equation, ( ) 25.012 / =∂∂

AlxAlL xE at 0K are assumed to be 45197J. And

12LAlE and 12L

NiE are then calculated by Eqs. C.2 and C.3 (See Figure C.2)

atomevmoleJELAl /116.5/49365312 −=−=

atomevmoleJELNi /586.5/53885012 −=−=

146

Table C.1 Parameters for total energy description of the L12 structured Ni-Al alloys. (in

J/mole)

35573612::: −=LAlAlAlAlE

42084612::: −=LNiAlAlAlE

49260612::: −=LNiNiAlAlE

52755112::: −=LNiNiNiAlE

52893412::: −=LNiNiNiAlE

1483::0 −=AlAlAlL

3772::0 +=NiAlAlL 2752::

1 +=NiAlAlL

4862::0 −=NiNiAlL 3125::

1 +=NiNiAlL

4734::0 +=NiNiNiL 1008::

1 +=NiNiNiL

147

Table C.2 Structural descriptions of the SQS structure for L12 alloys. A and B are

randomly distributed species in one sublattice and C and D are atoms occupying the other

three sublattices. The atomic positions are given in direct coordinates, and are for the

ideal, unrelaxed structures.

375.025.0 )( CBA

Lattice vectors

=1a (1.0 1.0 -2.0) =2a (1.0 -1.0 -2.0) =3a (-4.0 0.0 0.0) Atomic positions

A-(0.25000 0.25000 0.87500) A-(0.25000 0.25000 0.12500) A-(1.00000 1.00000 0.25000)

A-(1.00000 1.00000 1.00000) B-(0.75000 0.75000 0.87500) B-(0.75000 0.75000 0.62500)

B-(0.75000 0.75000 0.37500) B-(0.75000 0.75000 0.12500) B-(0.50000 0.50000 0.75000)

B-(0.50000 0.50000 0.50000) B-(0.50000 0.50000 0.25000) B-(0.50000 0.50000 1.00000)

B-(0.25000 0.25000 0.62500) B-(0.25000 0.25000 0.37500) B-(1.00000 1.00000 0.75000)

B-(1.00000 1.00000 0.50000) C-(1.00000 0.50000 0.75000) C-(0.62500 0.62500 0.68750)

C-(0.87500 0.37500 0.81250) C-(1.00000 0.50000 0.50000) C-(0.62500 0.62500 0.43750)

C-(0.87500 0.37500 0.56250) C-(1.00000 0.50000 0.25000) C-(0.62500 0.62500 0.18750)

C-(0.87500 0.37500 0.31250) C-(1.00000 0.50000 1.00000) C-(0.62500 0.62500 0.93750)

C-(0.87500 0.37500 0.06250) C-(0.75000 0.25000 0.62500) C-(0.37500 0.37500 0.56250)

C-(0.62500 0.12500 0.68750) C-(0.75000 0.25000 0.37500) C-(0.37500 0.37500 0.31250)

C-(0.62500 0.12500 0.43750) C-(0.75000 0.25000 0.12500) C-(0.37500 0.37500 0.06250)

C-(0.62500 0.12500 0.18750) C-(0.75000 0.25000 0.87500) C-(0.37500 0.37500 0.81250)

C-(0.62500 0.12500 0.93750) C-(0.50000 1.00000 0.75000) C-(0.12500 0.12500 0.68750)

C-(0.37500 0.87500 0.81250) C-(0.50000 1.00000 0.50000) C-(0.12500 0.12500 0.43750)

C-(0.37500 0.87500 0.56250) C-(0.50000 1.00000 0.25000) C-(0.12500 0.12500 0.18750)

C-(0.37500 0.87500 0.31250) C-(0.50000 1.00000 1.00000) C-(0.12500 0.12500 0.93750)

C-(0.37500 0.87500 0.06250) C-(0.25000 0.75000 0.62500) C-(0.87500 0.87500 0.56250)

C-(0.12500 0.62500 0.68750) C-(0.25000 0.75000 0.37500) C-(0.87500 0.87500 0.31250)

C-(0.12500 0.62500 0.43750) C-(0.25000 0.75000 0.12500) C-(0.87500 0.87500 0.06250)

C-(0.12500 0.62500 0.18750) C-(0.25000 0.75000 0.87500) C-(0.87500 0.87500 0.81250)

C-(0.12500 0.62500 0.93750)

148

35.05.0 )( CBA

Lattice vectors

=1a (1.0 2.0 1.0) =2a (1.0 0.0 -1.0) =3a (-3.0 2.0 -3.0) Atomic positions

A-(0.12500 0.50000 0.87500) A-(0.75000 0.50000 0.75000) A-(0.12500 1.00000 0.37500)

A-(0.87500 1.00000 0.62500) A-(1.00000 1.00000 1.00000) A-(0.50000 0.50000 1.00000)

A-(0.75000 1.00000 0.25000) A-(0.87500 0.50000 0.12500) B-(1.00000 0.50000 0.50000)

B-(0.25000 1.00000 0.75000) B-(0.37500 0.50000 0.62500) B-(0.62500 1.00000 0.87500)

B-(0.25000 0.50000 0.25000) B-(0.50000 1.00000 0.50000) B-(0.62500 0.50000 0.37500)

B-(0.37500 1.00000 0.12500) C-(0.37500 0.75000 0.87500) C-(0.25000 0.50000 0.75000)

C-(0.37500 0.25000 0.87500) C-(0.25000 0.75000 0.50000) C-(0.12500 0.50000 0.37500)

C-(0.25000 0.25000 0.50000) C-(0.50000 0.25000 0.75000) C-(0.37500 1.00000 0.62500)

C-(0.50000 0.75000 0.75000) C-(0.62500 0.75000 0.62500) C-(0.50000 0.50000 0.50000)

C-(0.62500 0.25000 0.62500) C-(0.87500 0.25000 0.87500) C-(0.75000 1.00000 0.75000)

C-(0.87500 0.75000 0.87500) C-(1.00000 0.75000 0.75000) C-(0.87500 0.50000 0.62500)

C-(1.00000 0.25000 0.75000) C-(0.37500 0.25000 0.37500) C-(0.25000 1.00000 0.25000)

C-(0.37500 0.75000 0.37500) C-(0.50000 0.75000 0.25000) C-(0.37500 0.50000 0.12500)

C-(0.50000 0.25000 0.25000) C-(0.75000 0.25000 0.50000) C-(0.62500 1.00000 0.37500)

C-(0.75000 0.75000 0.50000) C-(0.87500 0.75000 0.37500) C-(0.75000 0.50000 0.25000)

C-(0.87500 0.25000 0.37500) C-(0.12500 0.25000 0.62500) C-(1.00000 1.00000 0.50000)

C-(0.12500 0.75000 0.62500) C-(0.25000 0.25000 1.00000) C-(0.12500 1.00000 0.87500)

C-(0.25000 0.75000 1.00000) C-(0.62500 0.25000 0.12500) C-(0.50000 1.00000 1.00000)

C-(0.62500 0.75000 0.12500) C-(0.75000 0.75000 1.00000) C-(0.62500 0.50000 0.87500)

C-(0.75000 0.25000 1.00000) C-(1.00000 0.25000 0.25000) C-(0.87500 1.00000 0.12500)

C-(1.00000 0.75000 0.25000) C-(0.12500 0.75000 0.12500) C-(1.00000 0.50000 1.00000)

C-(0.12500 0.25000 0.12500)

2175.025.0 )( DCBA

Lattice vectors

=1a (1.0 2.0 0.0) =2a (1.0 0.0 -2.0) =3a (-3.0 1.0 -1.0) Atomic positions

A-(0.90625 0.65625 0.68750) A-(0.78125 0.53125 0.93750) A-(0.21875 0.46875 0.06250)

149

A-(0.28125 0.03125 0.93750) B-(0.34375 0.59375 0.81250) B-(0.46875 0.71875 0.56250)

B-(0.59375 0.84375 0.31250) B-(0.71875 0.96875 0.06250) B-(0.03125 0.78125 0.43750)

B-(0.15625 0.90625 0.18750) B-(0.40625 0.15625 0.68750) B-(0.53125 0.28125 0.43750)

B-(0.65625 0.40625 0.18750) B-(0.84375 0.09375 0.81250) B-(0.96875 0.21875 0.56250)

B-(0.09375 0.34375 0.31250) C-(0.06250 0.56250 0.87500) C-(0.18750 0.68750 0.62500)

C-(0.31250 0.81250 0.37500) C-(0.43750 0.93750 0.12500) C-(0.62500 0.62500 0.75000)

C-(0.75000 0.75000 0.50000) C-(0.87500 0.87500 0.25000) C-(0.12500 0.12500 0.75000)

C-(0.25000 0.25000 0.50000) C-(0.37500 0.37500 0.25000) C-(0.50000 0.50000 1.00000)

C-(0.56250 0.06250 0.87500) C-(0.68750 0.18750 0.62500) C-(0.81250 0.31250 0.37500)

C-(0.93750 0.43750 0.12500) C-(1.00000 1.00000 1.00000) D-(0.15625 0.40625 0.68750)

D-(0.31250 0.31250 0.87500) D-(0.28125 0.53125 0.43750) D-(0.43750 0.43750 0.62500)

D-(0.40625 0.65625 0.18750) D-(0.56250 0.56250 0.37500) D-(0.53125 0.78125 0.93750)

D-(0.68750 0.68750 0.12500) D-(0.71875 0.46875 0.56250) D-(0.87500 0.37500 0.75000)

D-(0.84375 0.59375 0.31250) D-(1.00000 0.50000 0.50000) D-(0.96875 0.71875 0.06250)

D-(0.12500 0.62500 0.25000) D-(0.21875 0.96875 0.56250) D-(0.37500 0.87500 0.75000)

D-(0.34375 0.09375 0.31250) D-(0.50000 1.00000 0.50000) D-(0.46875 0.21875 0.06250)

D-(0.62500 0.12500 0.25000) D-(0.59375 0.34375 0.81250) D-(0.75000 0.25000 1.00000)

D-(0.65625 0.90625 0.68750) D-(0.81250 0.81250 0.87500) D-(0.78125 0.03125 0.43750)

D-(0.93750 0.93750 0.62500) D-(0.90625 0.15625 0.18750) D-(0.06250 0.06250 0.37500)

D-(0.03125 0.28125 0.93750) D-(0.18750 0.18750 0.12500) D-(0.09375 0.84375 0.81250)

D-(0.25000 0.75000 1.00000)

215.05.0 )( DCBA

Lattice vectors

=1a (0.0 1.0 2.0) =2a (2.0 0.0 0.0) =3a (0.0 3.0 -2.0) Atomic positions

A-(0.12500 0.25000 0.12500) A-(0.37500 0.25000 0.37500) A-(0.62500 0.75000 0.62500)

A-(0.87500 0.75000 0.87500) A-(0.75000 0.25000 0.25000) A-(0.75000 0.75000 0.25000)

A-(1.00000 0.25000 0.50000) A-(1.00000 0.75000 0.50000) B-(0.25000 0.25000 0.75000)

B-(0.25000 0.75000 0.75000) B-(0.50000 0.25000 1.00000) B-(0.50000 0.75000 1.00000)

B-(0.12500 0.75000 0.12500) B-(0.37500 0.75000 0.37500) B-(0.62500 0.25000 0.62500)

B-(0.87500 0.25000 0.87500) C-(0.12500 1.00000 0.62500) C-(0.12500 0.50000 0.62500)

C-(0.37500 1.00000 0.87500) C-(0.37500 0.50000 0.87500) C-(1.00000 1.00000 1.00000)

150

C-(1.00000 0.50000 1.00000) C-(0.25000 1.00000 0.25000) C-(0.25000 0.50000 0.25000)

C-(0.50000 1.00000 0.50000) C-(0.50000 0.50000 0.50000) C-(0.75000 1.00000 0.75000)

C-(0.75000 0.50000 0.75000) C-(0.62500 1.00000 0.12500) C-(0.62500 0.50000 0.12500)

C-(0.87500 1.00000 0.37500) C-(0.87500 0.50000 0.37500) D-(0.31250 0.25000 0.56250)

D-(0.43750 1.00000 0.68750) D-(0.31250 0.75000 0.56250) D-(0.43750 0.50000 0.68750)

D-(0.56250 0.25000 0.81250) D-(0.68750 1.00000 0.93750) D-(0.56250 0.75000 0.81250)

D-(0.68750 0.50000 0.93750) D-(0.18750 0.25000 0.93750) D-(0.31250 1.00000 0.06250)

D-(0.18750 0.75000 0.93750) D-(0.31250 0.50000 0.06250) D-(0.43750 0.25000 0.18750)

D-(0.56250 1.00000 0.31250) D-(0.43750 0.75000 0.18750) D-(0.56250 0.50000 0.31250)

D-(0.68750 0.25000 0.43750) D-(0.81250 1.00000 0.56250) D-(0.68750 0.75000 0.43750)

D-(0.81250 0.50000 0.56250) D-(0.93750 0.25000 0.68750) D-(0.06250 1.00000 0.81250)

D-(0.93750 0.75000 0.68750) D-(0.06250 0.50000 0.81250) D-(0.81250 0.25000 0.06250)

D-(0.93750 1.00000 0.18750) D-(0.81250 0.75000 0.06250) D-(0.93750 0.50000 0.18750)

D-(0.06250 0.25000 0.31250) D-(0.18750 1.00000 0.43750) D-(0.06250 0.75000 0.31250)

D-(0.18750 0.50000 0.43750)

151

Figure C.1 Interactions between Al and Ni in the same sublattice. The open symbols are

calculated with constrained relaxations and the solid ones are from unconstrained

relaxations.

152

Figure C.2 Total energy of the L12 structured Ni-Al alloys at 0K.

153

REFERENCES

1. W. Kohn and L. J. Sham, Phys. Rev., 140 (1965) A1133.

2. J. O. Andersson, T. Helander, L. H. Hoglund, P. F. Shi and B. Sundman, Calphad,

26 (2002) 273-312.

3. Z. K. Liu, L. Q. Chen, R. Raghavan, Q. Du, J. O. Sofo, S. A. Langer and C.

Wolverton, J Comput-Aided Mater, 11 (2004) 183-199.

4. L. Q. Chen, Ann Rev Mater Res, 32 (2002) 113-140.

5. H. A. Calderon, G. Kostorz, Y. Y. Qu, H. J. Dorantes, J. J. Cruz and J. G.

CabanasMoreno, Mat Sci Eng a-Struct, 238 (1997) 13-22.

6. J. G. Conley, M. E. Fine and J. R. Weertman, Acta Metall, 37 (1989) 1251-1263.

7. J. O. Andersson and J. Agren, J Appl Phys, 72 (1992) 1350-1355.

8. T. Helander and J. Agren, Acta Mater, 47 (1999) 1141-1152.

9. M. L. Bhatia, A. K. Singh and T. K. Nandy, Intermetallics, 6 (1998) 141-146.

10. L. Vegard, Z. fur Physik, 5 (1921) 17-26.

11. H. W. King, J Mater Sci, 1 (1966) 79-90.

12. B. Sundman and J. Agren, J. Phys. Chem. Solids, 42 (1981) 297-301.

13. M. V. Nathal, R. A. Mackay and R. G. Garlick, Mater. Sci. Eng., 75 (1985) 195-

205.

14. D. F. Lahrman, R. D. Field, R. Darolia and H. L. Fraser, Acta Metall, 36 (1988)

1309-1320.

15. F. R. N. Nabarro, Metall Mater Trans A, 27 (1996) 513-530.

154

16. H. Mughrabi and U. Tetzlaff, Adv. Eng. Mater., 2 (2000) 319-326.

17. A. R. Ruffa, J Mater Sci, 15 (1980) 2268-2274.

18. F. L. Uffelmann, Phil. Mag., 10 (1930) 633-659.

19. L. Catz, Proc. Phys. Soc., 68 (1955) 957.

20. P. D. Pathak and N. G. Vasavada, J. Phys. C, 3 (1970) L44-L48.

21. O. Redlich and A. T. Kister, Ind. Eng. Chem., 40 (1948) 345-348.

22. J. O. Andersson, A. F. Guillermet, M. Hillert, B. Jansson and B. Sundman, Acta

Metall, 34 (1986) 437-445.

23. I. Ansara, B. Sundman and P. Willemin, Acta Metall, 36 (1988) 977-982.

24. Y. S. Touloukian, R. K. Kirby, R. E. Taylor and P. D. Desai, Thermophysical

Properties of Matter (IFI-Plenum, New York, 1975), vol. 12.

25. K. Wang and R. R. Reeber, Philos. Mag. A-Phys. Condens. Matter Struct. Defect

Mech. Prop., 80 (2000) 1629-1643.

26. J. F. Kenney and E. S. Keeping, Mathematics of statistics, Pt. 1 (Princeton, New

York, ed. 3rd, 1962).

27. T. G. Kollie, Phys Rev B, 16 (1977) 4872-4881.

28. G. D. Mukherjee, C. Bansal and A. Chatterjee, Int. J. Mod. Phys. B, 12 (1998)

449-470.

29. M. Yousuf, P. C. Sahu, H. K. Jajoo, S. Rajagopalan and K. G. Rajan, J. Phys. F,

Met. Phys., 16 (1986) 373-380.

30. Y. Tianji, J. Phys. Soc. Jpn., 31 (1971) 1366-1373.

31. M. P. Arbuzov and I. A. Zelenkov, Phys. Met. Metallogr. (Russia), 18 (1964)

149-150.

155

32. H. W. Altman, T. Rubin and H. L. Johnston, "Coefficient of Thermal Expansion

of Solids at Low Temperature. III. The Thermal Expansion of Pure Metals, with

the Data for Aluminum, Nickel, Titanium, and Zirconium" Cryogenic Laboratory

Rept. OSU-TR-264-27 (Ohio State University, 1954) pp. 10.

33. F. C. Nix and D. Macnair, Physical Review, 60 (1941) 597-605.

34. A. S. Pavlovic, V. S. Babu and M. S. Seehra, J Phys-Condens Mat, 8 (1996)

3139-3149.

35. V. V. Makeev and P. S. Popel, High Temp., 28 (1990) 525-529.

36. J. Bandyopadhyay and K. P. Gupta, Cryogenics, 17 (1977) 345-347.

37. R. Kohlhaas, P. Dunner and P. N. Schmitz, Z. Angew. Phys., 23 (1967) 245-249.

38. T. Faisst, Journal De Physique, 49 (1988) 65-66.

39. J. Bottiger, N. Karpe, J. P. Krog and A. V. Ruban, J. Mater. Res., 13 (1998) 1717-

1723.

40. A. B. Kamara, A. J. Ardell and C. N. J. Wagner, Metall Mater Trans A, 27 (1996)

2888-2896.

41. Y. Y. Qiu, J. Mater. Sci., 31 (1996) 4311-4319.

42. G. Ayub, F. A. Khwaja, A. U. Haq and Z. Ahmad, Acta Metall Mater, 43 (1995)

1457-1465.

43. T. Tarfa, B. Sitaud and O. Dimitrov, Acta Metall Mater, 41 (1993) 3191-3202.

44. A. Taylor and R. W. Floyd, J Inst Met, 81 (1952-53) 25-32.

45. Y. Mishima, S. Ochiai and T. Suzuki, Acta Metall., 33 (1985) 1161-1169.

46. P. V. M. Rao, K. S. Murthy, S. V. Suryanarayana and S. V. N. Naidu, Journal of

Materials Research (USA), 8 (1993) 741-744.

156

47. C. Bertrand, J. P. Dallas, J. Rzepski, N. Sidhom, M. F. Trichet and M. Cornet,

Mem. Etud. Sci. Rev. Met., 86 (1989) 334-346.

48. I. Baker, B. Huang and E. M. Schulson, Acta Metall., 36 (1988) 493-499.

49. Y. G. Kim, G. L. W. Yoon and N. S. Stoloff, J. Mater. Sci. Lett., 4 (1985) 1407-

1408.

50. S. C. Huang, A. I. Taub and K. M. Chang, Acta Metall., 32 (1984) 1703-1707.

51. O. Noguchi, Y. Oya and T. Suzuki, Metall. Trans. A, 12A (1981) 1647-1653.

52. K. Aoki and O. Izumi, J. Jpn. Inst. Met., Dec., 39 (1975) 1282-1289.

53. G. R. Stoeckinger and J. P. Neumann, Journal of Applied Crystallography, 1

(1970) 32-38.

54. C. L. Corey and B. Lisowsky, Aime Met. Soc. Trans., 239 (1967) 239-243.

55. E. Hornbogen and M. Roth, Z. Metallk., 58 (1967) 842-855.

56. V. A. Phillips, Acta Metall, 14 (1966) 1533-1547.

57. M. P. Arbuzov and I. A. Zelenkov, Fiz. Metal. Metalloved., 15 (1963) 725-728.

58. R. W. Guard and J. H. Westbrook, Transactions of the Metallurgical Society of

AIME, 215 (1959) 807-814.

59. A. J. Bradley and A. Taylor, Proc. Roy. Soc., A159 (1937)

60. P. V. M. Rao, S. V. Suryanarayana, K. S. Murthy and S. V. N. Naidu, J. Phys.,

Condens. Matter, 1 (1989) 5357-5361.

61. N. Dupin, I. Ansara and B. Sundman, Calphad, 25 (2001) 279-298.

62. I. Ansara, N. Dupin, H. L. Lukas and B. Sundman, J Alloy Compd, 247 (1997) 20-

30.

63. A. Engstrom and J. Agren, Z. Metallk., 87 (1996) 92-97.

157

64. H. Numakura, T. Ikeda, M. Koiwa and A. Almazouzi, Philos. Mag. A-Phys.

Condens. Matter Struct. Defect Mech. Prop., 77 (1998) 887-909.

65. H. A. Moreen, R. Taggart and D. H. Polonis, Metall. Trans., 2 (1971) 265-268.

66. H. Biermann, H. Mughrabi and M. Strehler, Metall. Trans. A, 27A (1996) 1003-

1014.

67. A. Royer, P. Bastie and M. Veron, Acta Mater, 46 (1998) 5357-5368.

68. R. F. Decker and J. R. Mihalisin, ASM. Trans. Quart., 62 (1969) 481-489.

69. D. Q. Wang, S. S. Babu, E. A. Payzant, P. G. Radaelli and A. C. Hannon, Metall.

Trans. A, 32A (2001) 1551-1552.

70. S. Yoshitake, T. Yokokawa, K. Ohno, H. Harada and M. Yamazaki, Temperature

dependence of gamma gamma ' lattice parameter misfit for nickel-base

superalloys, Materials for Advanced Power Engineering 1994. Part I, Liege,

Belgium (Kluwer Academic Publishers Group, P.O. Box 322, Dordrecht, 3300AH,

Netherlands, 1994), pp. 875-882.

71. I. L. Svetlov, I. V. Oldakovskii, N. V. Petrushin and I. A. Ignatova, Russ Metall+,

(1991) 139-145.

72. G. N. Maniar and E. J. Bridge, Metallography, 5 (1972) 91-93.

73. F. R. Kroeger and C. A. Swenson, J. Appl. Phys., 48 (1977) 853-864.

74. B. Guerard, H. Peisl and R. Zitzmann, Appl. Phys., 3 (1974) 37-43.

75. J. G. Collins and G. K. White, Journal of Low Temperature Physics, 10 (1973)

69-77.

76. A. J. C. Wilson, Proceedings of the Physical Society, 54 (1942) 487-491.

77. K. Honda and Y. Okubo, Sci. Rep. Tohoku Univ., 13 (1924) 101-107.

158

78. G. Langelaan and S. Saimoto, Rev Sci Instrum, 70 (1999) 3413-3417.

79. A. J. Cornish and J. Burke, J. Sci. Instrum., 42 (1965) 212-218.

80. H. M. Otte, W. G. Montague and D. O. Welch, J Appl Phys, 34 (1963) 3149-3150.

81. R. O. Simmons and R. W. Balluffi, Physical Review, 117 (1960) 52-61.

82. R. Feder and A. S. Nowick, Physical Review, 109 (1958) 1959-1963.

83. F. A. Mauer, R. G. Munro, G. J. Piermarini and S. Block, J Appl Phys, 58 (1985)

3727-3730.

84. I. M. Robertson and C. M. Wayman, Metallography, 17 (1984) 43-55.

85. R. S. Mints, G. F. Belyaeva and Y. S. Malkov, Russ. J. Inorg. Chem., 7 (1962)

1236-1239.

86. T. Wang, J. Z. Zhu, R. A. Mackay, L. Q. Chen and Z. K. Liu, Metall Mater Trans

A, 35A (2004) 2313-2321.

87. G. Kresse and J. Hafner, Phys Rev B, 48 (1993) 13115-13118.

88. J. P. Perdew, Electronic Structure of Solids. P. Ziesche, H. Eschrig, Eds.

(Akademie Verlag, Berlin, 1991).

89. R. O. Jones and O. Gunnarsson, Rev. Mod. Phys., 61 (1989) 689.

90. A. Lindbaum, J. Hafner, E. Gratz and S. Heathman, J Phys-Condens Mat, 10

(1998) 2933-2945.

91. A. Lindbaum, J. Hafner and E. Gratz, J Phys-Condens Mat, 11 (1999) 1177-1187.

92. N. Sandberg, B. Magyari-Kope and T. R. Mattsson, Phys Rev Lett, 89 (2002)

65901.

93. G. Kresse and J. Furthmuller, VASP the GUIDE,

http://cms.mpi.univie.ac.at/vasp/vasp/vasp.html.

159

94. C. Counterman, Periodic Table of the Elements,

http://web.mit.edu/3.091/www/pt/pert1.html.

95. U. Scheuer and B. Lengeler, Phys Rev B, 44 (1991) 9883-9894.

96. A. Taylor and R. W. Floyd, J. Inst. Met., 80 (1951-52) 577-587.

97. W. B. Pearson and L. T. Thompson, Can. J. Phys., 35 (1957) 349-357.

98. A. Taylor, J. Inst. Met., 77 (1950) 585.

99. M. Hnilicka and I. Karmazin, Scr. Metall., 8 (1974) 1029-1032.

100. B. M. Rovinshy, A. I. Samoilov and G. M. Rovensky, Fiz. Met. Metalloved., 7

(1959) 79-90.

101. I. I. Kornilov and A. Y. Snetkov, Issled. Zharoproch. Splavam, Akad. Nauk SSSR,

Inst., Met. Im. A. A. Baikova, 7 (1961) 106-111.

102. L. Karmazin and M. Svoboda, Kovove Mater., 17 (1979) 355-362.

103. W. Koster and W. Schmidt, Arch. Eisenhuttenwes., 8 (1934) 25-27.

104. G. Grube and H. Schlecht, Z. Elektrochem., 44 (1938) 413-422.

105. P. W. Guthrie and E. E. Stansbury, "X-ray and Metallographic Study of the Ni-

Rich Alloys of the Ni-Mo System II" ORNL-3078 (Oak Ridge Nat. Lab., Oak

Ridge, TN, 1961).

106. R. E. W. Casselton and W. Hume-Rothery, J. Less-Common Met., 7 (1964) 212-

221.

107. G. Grube, O. Kubaschewski and K. Zwiauer, Z. Elektrochem., 45 (1939) 881-884.

108. S. Arajs, H. Chessin and R. V. Colvin, Phys. Status Solidi, 3 (1963) 2337-2346.

109. I. J. Duerden and W. Hume-Rothery, J. Less-Common Met., 11 (1966) 381-387.

110. L. N. Guseva, R. S. Mints and E. S. Malkov, Russ. Metall., 5 (1969) 120-122.

160

111. D. S. Duvall and J. M. J. Donachie, J. Inst. Met., 98 (1970) 182-187.

112. E. M. Savitskii, M. A. Tylkina and E. P. Arskaya, Izv. V. U. Z. Tsvetn. Metall., 4

(1970) 114-116.

113. E. Raub and D. Menzel, Z. Metallkd., 52 (1961) 831-833.

114. W. L. Phillips, Trans. Metall. Soc. AIME, 230 (1964) 526-529.

115. I. I. Kornilov and K. P. Myasnikova, Izv. Akad. Nauk SSSR, Met. Govn. Delo,

(1964) 159-165.

116. H. Chessin, S. Arajs and R. V. Colvin, J. Appl. Phys., 35 (1964) 2419-2423.

117. T. H. Hazlett and E. R. Parker, Trans. ASM, 46 (1954) 701-715.

118. D. M. Poole and W. Hume-Rothery, J. Inst. Met., 83 (1954-55) 473-480.

119. K. Hashimoto, T. Tsujimoto and K. Saito, Trans., Jpn. Inst. Met., 22 (1981) 798-

806.

120. G. Chattopadhyay and H. Kleykamp, Z. Metallkd., 74 (1983) 182-197.

121. E. Epremian and D. Harker, Trans. AIME, 185 (1949) 267-273.

122. R. Watanabe and T. Kuno, Tetsu to Hagane, 61 (1975) 2274-2294.

123. H. Harada and M. Yamazaki, Tetsu to Hagane, 65 (1979) 1059-1068.

124. J. Li and R. P. Wahi, Acta Metall Mater, 43 (1995) 507-517.

125. R. Volkl, U. Glatzel and N. Feller-Kniepmeier, Scripta Mater, 38 (1998) 893-900.

126. C. Schulze and M. Feller-Kniepmeier, Mat Sci Eng a-Struct, 281 (2000) 204-212.

127. D. Sieborger, H. Brehm, F. Wunderlich, D. Moller and U. Glatzel, Z Metallkd, 92

(2001) 58-61.

128. U. Bruckner, A. Epishin, T. Link and K. Dressel, Mat Sci Eng a-Struct, 247 (1998)

23-31.

161

129. R. Volkl, U. Glatzel and M. Feller-Kniepmeier, Acta Mater, 46 (1998) 4395-4404.

130. Y. Wang, Z. K. Liu and L. Q. Chen, Acta Mater, 52 (2004) 2665-2671.

131. S. H. Zhou, Y. Wang, J. Z. Zhu, T. Wang, L. Q. Chen, R. A. MacKay and Z. K.

Liu, Superalloys, (2004) 969-975.

132. M. Fahrmann, P. Fratzl, O. Paris, E. Fahrmann and W. C. Johnson, Acta Metall

Mater, 43 (1995) 1007-1022.

133. J. Z. Zhu, T. Wang, A. J. Ardell, S. H. Zhou, Z. K. Liu and L. Q. Chen, Acta

Mater, 52 (2004) 2837-2845.

134. C. E. Campbell, W. J. Boettinger and U. R. Kattner, Acta Mater, 50 (2002) 775-

792.

135. B. Jonsson, Z Metallkd, 85 (1994) 502-509.

136. B. Jonsson, Z Metallkd, 86 (1995) 686-692.

137. T. Helander and J. Agren, Acta Mater, 47 (1999) 3291-3300.

138. T. Yamamoto, T. Takashima and K. Nishida, T Jpn I Met, 21 (1980) 601-608.

139. M. Watanabe, Z. Horita and M. Nemoto, Defect and Diffusion Forum, 143 (1997)

345-350.

140. D. F. Susan and A. R. Marder, Acta Mater, 49 (2001) 1153-1163.

141. A. Davin, V. Leroy, D. Coutsouradis and L. Habraken, Cobalt, 19 (1963) 51-56.

142. R. A. Swalin, A. Martin and R. Olson, T Am I Min Met Eng, 209 (1957) 936-939.

143. C. P. Heijwegen and G. D. Rieck, Acta Met., 22 (1974) 1269-1281.

144. Y. Minamino, H. Yoshida, S. B. Jung, K. Hirao and T. Yamane, Defect Diffus

for/Jr, 143 (1997) 257-262.

145. L. Boltzmann, Annu. Phys., 53 (1894) 960.

162

146. C. Matano, Jpn. J. Phys., 8 (1933) 109-113.

147. Y. S. Shueh, J. P. Hirth and R. A. Rapp, Metall Trans A, 22 (1991) 1501-1510.

148. N. Chi and D. Bergner, Trans Tech Publications, pp, (1983)

149. C. P. Chang and M. H. Loretto, Philos Mag A, 57 (1988) 593-603.

150. A. R. Paul and R. P. Agarwala, J Appl Phys, 38 (1967) 3790-&.

151. G. F. Hancock, Physica Status Solidi (a), 7 (1971) 535-540.

152. M. B. Bronfin, G. S. Bulatov and I. A. Drugova, Fiz. Metal. Metalloved., 40

(1975) 363-366.

153. K. Hoshino, S. J. Rothman and R. S. Averback, Acta Metall, 36 (1988) 1271-1279.

154. Y. Shi, G. Frohberg and H. Wever, Phys Status Solidi A, 152 (1995) 361-375.

155. S. Frank, U. Sodervall and C. Herzig, Phys. Status Solidi B-Basic Res., 191 (1995)

45-55.

156. H. Mehrer, Mater. Trans. JIM, 37 (1996) 1259-1280.

157. M. Watanabe, Z. Horita, D. J. Smith, M. R. Mccartney, T. Sano and M. Nemoto,

Acta Metall Mater, 42 (1994) 3381-3396.

158. M. Watanabe, Z. Horita and M. Nemoto, Defect and Diffusion Forum, 143 (1997)

637-642.

159. F. Sauer and V. Freise, Z. Electrochem., 66 (1962) 353-363.

160. K. Fujiwara and Z. Horita, Acta Mater, 50 (2002) 1571-1579.

161. M. M. P. Janssen, Metall. Trans., 4 (1973) 1623-1633.

162. S. Shankar and L. L. Seigle, Metall. Trans. A, 9 (1978) 1467-1476.

163. T. Ikeda, A. Almazouzi, H. Numakura, M. Koiwa, W. Sprengel and H. Nakajima,

Acta Mater, 46 (1998) 5369-5376.

163

164. T. Ikeda, A. Almazouzi, H. Numakura, M. Koiwa, W. Sprengel and H. Nakajima,

Defect and Diffusion Forum, 143 (1997) 275-278.

165. S. Frank, U. Sodervall and C. Herzig, Defect and Diffusion Forum, 143 (1997)

245-250.

166. I. V. Belova and G. E. Murch, Philos Mag A, 80 (2000) 1481-1493.

167. S. V. Divinski, S. Frank, U. Sodervall and C. Herzig, Acta Mater, 46 (1998)

4369-4380.

168. N. Kurita and M. Koiwa, Intermetallics, 10 (2002) 735-741.

169. A. B. Lidiard, Philos Mag, 46 (1955) 1218-1237.

170. A. D. Leclaire, J. Nucl. Mater., 69-7 (1978) 70-96.

171. L. D. Hall, J Chem Phys, 21 (1953) 87-89.

172. Y. Wang and A. G. Khachaturyan, Acta Metall Mater, 43 (1995) 1837-1857.

173. V. Vaithyanathan and L. Q. Chen, Acta Mater, 50 (2002) 4061-4073.

174. S. G. Kim, W. T. Kim and T. Suzuki, Phys Rev E, 60 (1999) 7186-7197.

175. A. G. Khachaturyan, Theory of Structural Transformation in Solids (Wiley, New

York, 1983).

176. T. Wang, L. Q. Chen and Z. K. Liu, Unpublished work, (2005)

177. A. W. West and D. H. Kirkwood, Scripta Metall Mater, 10 (1976) 681-686.

178. P. Staron and R. Kampmann, Acta Mater, 48 (2000) 701-712.

179. G. A. Zickler, B. Tian, C. Lind and O. Paris, Journal of Applied Crystallography,

36 (2003) 484-488.

180. S. V. Prikhodko, J. D. Carnes, D. G. Isaak and A. J. Ardell, Scripta Mater, 38

(1997) 67-72.

164

181. S. V. Prikhodko, J. D. Carnes, D. G. Isaak, H. Yang and A. J. Ardell, Metall

Mater Trans A, 30 (1999) 2403-2408.

182. M. Fahrmann, W. Hermann, E. Fahrmann, A. Boegli, T. M. Pollock and H. G.

Sockel, Determination of matrix and precipitate elastic constants in ( gamma -

gamma ') Ni-base model alloys, and their relevance to rafting, Materials Science

and Engineering A (Switzerland), vol. 260, no. 1-2, pp. 212-221, Great Oaks

Landing, Maryland, USA (Switzerland, 1999), pp.

183. I. M. Lifshitz and V. V. Slyozov, J. Phys. Chem. Solids, 19 (1961) 35-50.

184. C. Wagner, Z Elektrochem, 65 (1961) 581-591.

185. R. Asimow, Acta Metall, 11 (1963) 72-73.

186. A. J. Ardell, Acta Metall, 20 (1972) 61-&.

187. A. D. Brailsford and P. Wynblatt, Acta Metall, 27 (1979) 489-497.

188. C. K. L. Davies, P. Nash and R. N. Stevens, Acta Metall, 28 (1980) 179-189.

189. J. A. Marqusee and J. Ross, J Chem Phys, 80 (1984) 536-543.

190. P. W. Voorhees and M. E. Glicksman, Metall Trans A, 15 (1984) 1081-1088.

191. M. Tokuyama, K. Kawasaki and Y. Enomoto, Physica A, 134 (1986) 323-338.

192. J. H. Yao, K. R. Elder, H. Guo and M. Grant, Phys Rev B, 47 (1993) 14110-14125.

193. V. A. Snyder, N. Akaiwa, J. Alkemper and P. W. Voorhees, Metall Mater Trans

A, 30 (1999) 2341-2348.

194. D. M. Kim and A. J. Ardell, Scripta Mater, 43 (2000) 381-384.

195. A. T. Dinsdale, Calphad, 15 (1991) 317-425.

196. M. Hagen and M. W. Finnis, Philos Mag A, 77 (1998) 447-464.

197. M. de Koning, C. R. Miranda and A. Antonelli, Phys Rev B, 66 (2002) -.

165

198. C. Jiang, C. Wolverton, J. Sofo, L. Q. Chen and Z. K. Liu, Phys Rev B, 69 (2004)

-.

199. S. H. Wei, L. G. Ferreira, J. E. Bernard and A. Zunger, Phys Rev B, 42 (1990)

9622-9649.

200. A. Zunger, S. H. Wei, L. G. Ferreira and J. E. Bernard, Phys Rev Lett, 65 (1990)

353-356.

201. A. van de Walle, M. Asta and G. Ceder, Calphad, 26 (2002) 539-553.

VITA

Tao Wang was born on September 09, 1972. He graduated from Central South

University, Changsha, China in 1994 with a B.S. degree in Materials Science and

Engineering. He continued his graduate study at the same university and got his M.S.

degree in Materials Science and Engineering in 1997. In 2002, he joined the

Pennsylvania State University, University Park, PA, to pursue his Ph.D. degree in

Materials Science and Engineering. He is a member of the Minerals, Metals and

Materials Society (TMS). Listed below are his publications during his Ph.D. study:

1. Wang T, Zhu JZ, Mackay RA, Chen LQ and Liu ZK, “Modeling of lattice parameter in the Ni-Al system”, Metall. Mater. Trans., 35A (2004)

2. Wang T, Chen LQ and Liu ZK, “First-principles calculations and phenomenological modeling of lattice misfit in Ni-base superalloys”, Mater. Sci. Eng. A, (2006)

3. Wang T, Chen LQ and Liu ZK, “Lattice parameters and local lattice distortions in fcc-Ni solutions”, submitted to Metall. Mater. Trans. (2006)

4. Wang T, Liu ZK and Chen LQ, “Coarsening kinetics of 'γ precipitates in the Ni-Al-Mo system”, to be submitted, (2006)

5. Wang T, Chen LQ and Liu ZK, “Modeling of atomic mobility in Al-Ni L12 phase”, in preparation, (2006)


Recommended