thermodynamic modeling and mechanical properties …

224
The Pennsylvania State University The Graduate School THERMODYNAMIC MODELING AND MECHANICAL PROPERTIES MODELING OF LONG PERIODIC STACKING ORDERED (LPSO) PHASES A Dissertation in Materials Science and Engineering by Hongyeun Kim © 2019 Hongyeun Kim Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy August 2019

Upload: others

Post on 11-Feb-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

THERMODYNAMIC MODELING AND MECHANICAL PROPERTIES MODELING

OF LONG PERIODIC STACKING ORDERED (LPSO) PHASES

A Dissertation in

Materials Science and Engineering

by

Hongyeun Kim

© 2019 Hongyeun Kim

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

August 2019

ii

The dissertation of Hongyeun Kim was reviewed and approved* by the following:

Zi-Kui Liu

Distinguished Professor of Materials Science and Engineering

Dissertation Advisor

Chair of Committee

Ismaila Dabo

Assistant Professor of Materials Science and Engineering

Hojong Kim

Assistant Professor of Materials Science and Engineering

Adri van Duin

Professor of Department of Mechanical & Nuclear Engineering

Laszlo Kecskes

Special Member,

Adjunct Associate Research Scholar, Johns Hopkins University

Suzanne Mohney

Professor of Materials Science and Engineering

Administrative

*Signatures are on file in the Graduate School

iii

ABSTRACT

Recently, there has been an increasing interest in long periodic stacking ordered (LPSO)

phases in Mg alloys due to their lightweight, high elastic and mechanical properties. The Vickers

indentation hardness and Young’s modulus of LPSO phases have reached 470% and 140%,

respectively, of that of pure Mg. Although theoretical and experimental studies have revealed the

phase constitutions and crystal structures of LPSO phases including the formation of the noble

solute atom clusters, which is also known as L12-type clusters, their phase stabilities and the origin

of their enhanced mechanical properties are not yet solved. To further improve the properties and

design the alloys, a thorough understanding of the phase equilibria and the origin of the mechanical

properties of LPSO phases are therefore needed.

In this dissertation, the elastic properties of LPSO phases in the Mg-Al-Gd system were

studied using first-principles calculations. Since LPSO phases have been reported to enhance the

strength and ductility of Mg alloys due to their high elastic properties, the effects of atomic

arrangements in terms of Gd-Al L12-type clusters on LPSOs’ elastic properties in the Mg-Al-Gd

system were studied using first-principles calculations. Four types of LPSO phases (10H, 18R,

14H, and 24R) were investigated with and without an interstitial atom in the center of the L12-type

clusters. Furthermore, the calculated Poisson’s ratios of each LPSO phases from this study is also

used as an important parameter for obtaining thermodynamic properties.

Thermodynamic modelling of the four LPSO phases, i.e., 10H, 18R, 14H, and 24R, in the

Mg-Al-Gd system was performed using the CALPHAD (calculation of phase diagram) approach

with input from the present first-principles calculations and experimental data in the literature.

Sublattice models were developed to describe these LPSO phases. Especially, an L12-type clusters

in the FCC stacking layers of LPSO phases and the atomic occupancy in the center of L12 cluster

were considered based on experimental observations and energetics from first-principles

iv

calculations. The calculated phase equilibrium results are in good agreement with experiments

about the phase stability of 14H and 18R and the mole fraction of Gd and Al in these LPSO phases.

The present modeling provides a new approach to describe the thermodynamic properties of LPSO

phases that can be applied to other alloy systems.

Material hardness is a good indicator of mechanical properties. However, since there is no

hardness model that can be used for LPSO phases, a large portion of the effort in this dissertation

is devoted to developing a suite of hardness models, which can be divided into three categories:

hardness model for both brittle and ductile materials, temperature-dependent hardness model and

hardness model for layered structures. In turn, the hardness of the LPSO phases is

obtained/modeled, based on these hardness models that were developed.

Hardness, defined as the resistance of a material to deformation, is a quick and efficient

measure of mechanical performance of materials. However, to date no comprehensive predictive

models exist for both metals and ceramics. We present a physics-based model that is capable of

predicting Vickers indentation hardness of both brittle and ductile materials with model inputs from

either first-principles calculations or experiments. Particularly, we go beyond the elastic properties

of materials commonly used in the literature and introduce the plastic properties of materials in

terms of active slip systems, including the Peierls-Nabarro flow stress, Burgers vector and slip

plane spacing into the model. It is demonstrated that this model can predict hardness values from

below 0.1 GPa of pure aluminum to above 100 GPa of diamond. The predictive power of the new

model has the potential to significantly advance the computational discovery and design of new

materials with enhanced performance.

Furthermore, a new temperature dependent hardness model is also proposed based on the

thermally activated dislocation width in combination with our previous Vickers hardness model.

The thermally activated dislocation width, a basic building block for the temperature dependent

Peierls-Nabarro flow stress in the hardness model, captures dislocation-diffusion mechanisms

v

during the materials’ deformation. In the proposed model, the material hardness is determined by

(a) diffusion mechanisms, (b) slip systems, (c) diffusing species, and (d) phase transformations.

The model has been calibrated for and agrees well with experimental hot hardness results of 16

materials, which were available from the public domain, including metals and ceramics.

The hardness model for layered structures is also modeled in order to investigate the origin

of the Hall-Petch relation in structures with twinned, tilt and twist boundaries, especially, hardness

enhancement of these structures based on material’s active slip systems of the structure as well as

the elastic properties since the slip systems are crucial to understanding the deformation of

materials. The active slip systems in this model are modulated by the relaxation of atomic positions

near the boundaries. This proposed model explains the flow stress and the hardness changes as the

twin or grain size in the structure changes, that is previously considered as an outcome of the Hall-

Petch relation.

vi

TABLE OF CONTENTS

List of Figures .......................................................................................................................... viii

List of Tables ........................................................................................................................... xiii

Acknowledgements .................................................................................................................. xiv

Chapter 1 Introduction ............................................................................................................. 1

1.1 Motivation .................................................................................................................. 1 1.2 Overview .................................................................................................................... 2

Chapter 2 Elastic Properties of Long Periodic Stacking Ordered Phases in Mg-Al-Gd

Alloys: A First-Principles Study ...................................................................................... 4

2.1 Introduction ................................................................................................................. 4 2.2 Computational Methods .............................................................................................. 6 2.3 Results and Discussion ............................................................................................... 8

2.3.1 Structural Analysis of the LPSO Phases ......................................................... 8 2.3.2 Elastic Properties of the LPSO Phases ............................................................ 12 2.3.3 Electronic Properties of the LPSOs ................................................................. 21

Chapter 3 First-Principles Calculations and Thermodynamic Modelling of Long Periodic

Stacking Ordered (LPSO) Phases in Mg-Al-Gd .............................................................. 25

3.1 Introduction ................................................................................................................ 25 3.2 First-Principles Calculations ...................................................................................... 26 3.3 CALPHAD Modeling of Phase Equilibria ................................................................. 30 3.4 Results and Discussion ............................................................................................... 34

Chapter 4 Predictive Modeling of Hardness of Brittle and Ductile Materials ......................... 46

4.1 Introduction ................................................................................................................ 46 4.2 Presentation of the New Model .................................................................................. 48 4.3 Validation and Prediction ........................................................................................... 49 4.4 Discussion .................................................................................................................. 56 4.5 Full Derivation of the Hardness Model ...................................................................... 60

4.5.1 Derivation of the Hardness Equation .............................................................. 60 4.5.2 Evaluation of Model Parameters ..................................................................... 63 4.5.2.1 hT/hp Ratio ................................................................................................. 63 4.5.2.2 Parameter c ................................................................................................... 70

Chapter 5 Temperature Dependent Hardness Model: the Study of Thermally Activated

Dislocation Width ............................................................................................................ 85

5.1 Introduction ................................................................................................................ 85

vii

5.2 Results and Discussion ............................................................................................... 87 5.2.1 Change of Diffusion Mechanism .................................................................... 88 5.2.2 Change of the Active Slip System ................................................................... 91 5.2.3 Phase Transformation at Finite Temperature .................................................. 92 5.2.4 Change of the Diffusion Species ..................................................................... 95 5.2.5 Phase Transformations During Indentation ..................................................... 98

5.3 Modeling Procedure ................................................................................................... 98 5.3.1 Derivation ........................................................................................................ 98 5.3.2 Temperature-Dependent Elastic Properties ..................................................... 105

Chapter 6 Hardness Modeling for Layered Structures: The Origin of Hall-Petch Relation .... 110

6.1 Introduction ................................................................................................................ 110 6.2 Methodology .............................................................................................................. 113

6.2.1 Derivation of Peierls-Nabarro Flow Stress for Twinned Structures ............... 113 6.2.2 First-Principles Calculations ........................................................................... 119

6.3 Results and Discussion ............................................................................................... 119

Chapter 7 Hardness Modeling of LPSO Phases....................................................................... 127

7.1 Methodology .............................................................................................................. 127 7.2 Results and Discussion ............................................................................................... 128

Chapter 8 Conclusions and Future Work ................................................................................. 131

8.1 Conclusions ................................................................................................................ 131 8.2 Future Work ............................................................................................................... 133

Appendix A Complete Elastic Stiffness Matrixes of 10H, 18R and 24R LPSO Phases .......... 134

Appendix B Thermo-Calc Mg-Al-Gd Database ...................................................................... 136

Bibliography ............................................................................................................................ 166

viii

LIST OF FIGURES

Figure 2.1 The LPSO structures of 10H (a), 18R (b), 14H (c), and 24R (d) together with

the in-plane L12 cluster ordering (e) and the Gd8Al6 L12 cluster with an interstitial

(int.) atom Gd, Mg or Al (f). Blue box stands for the unit cell of each LPSO

structures and the red bracket with SB stands for structural block for each LPSO

structure. 𝑑𝑖𝑛𝑡𝑟𝑎𝑐𝑙𝑢𝑠𝑡𝑒𝑟 and 𝑑𝑖𝑛𝑡𝑒𝑟𝑐𝑙𝑠𝑡𝑒𝑟 stands for the 2NN RE-RE intracluster

and intercluster distances, 𝑤clusterand ℎcluster stands for the L12 cluster width and

height. ............................................................................................................................... 10

Figure 2.2 Calculated bulk moduli of the LPSO phases with respect to number of layers

in structural block; (a) bulk modulus from EOS fitting and (b) bulk modulus from

VRH approach. Red dash lines indicate the bulk and shear moduli of HCP Mg. ........... 15

Figure 2.3 (a) comparison of bulk moduli both from VRH and EOS fitting as a function

of formation energies of LPSOs, and (b) Young’s modulus along [0001] direction

trend as a function of volumetric formation energies (𝐸𝑓/𝑉) of LPSOs. ........................ 17

Figure 2.4 Changes in (a) C11, (b) C33, (c) C44 and (d) C66 elastic constants as a function

of the number of layers in structural blocks. .................................................................... 18

Figure 2.5 Comparison between C11 and the energy contribution of interstitial atom in

L12 cluster. ....................................................................................................................... 19

Figure 2.6 Relationship of L12 cluster width with (a) C11, and (b) C66 elastic constants. ........ 20

Figure 2.7 Crystallographic orientation dependence of the Young’s and Shear modulus of

10H, 18R, 14H and 24R LPSO phase at 0K, between [0001] and <1120> 𝜃 is the

angle from <1120>. The orientation dependencies of the Young’s modulus and

shear modulus of HCP Mg are shown for comparison. ................................................... 21

Figure 2.8 Differential charge density plots of the LPSOs with or without interstitial

atoms. Differential charge density plots of (a) 10H, (b) 18R, (c) 14H, and (d) 24R

LPSO. The reference states used in this study are (e) 2H and (f) 14H LPSO with Mg

only; (g), (h), and (i) are the 14H LPSO with Al-int., Gd-int., and Mg-int. Red

arrows indicate the charge density connections between the {0001] planes.

Isosurfaces are 0.0021 (e/Å 3) and the Mg atom sizes are exaggerated for better

visualization. .................................................................................................................... 23

Figure 3.1 Experimentally observed LPSO phase compositions of 14H (a) and 18R (b)

LPSO phases[8], [19], [23], [65], [91]–[101]. Dash lines refer to the composition

ranges of this model. ........................................................................................................ 31

Figure 3.2. (a) Phonon dispersion curves of HCP Mg with experimental data[105] (red

circles), and b) phonon DOS curves of HCP Mg and Mg LPSO phases of 10H, 18R,

14H, and 24R. .................................................................................................................. 35

ix

Figure 3.3 Comparison of (a) heat capacity of HCP Mg with experimental data from

SGTE[82], (b) heat capacities of Mg-only LPSO phases, and (c) Gibbs energy

differences of various pure Mg LPSO phases with respect to HCP Mg. ......................... 37

Figure 3.4 Formation energies of endmembers of the 10H, 18R, 14H, and 24R LPSO

phases at 0 K. The data sets of GdIIAlIIIMgIV-Saal (×) were taken from the

literature[14]. .................................................................................................................... 39

Figure 3.5 (a) Composition ranges of GdIIAlIII(Mg, Gd, Al, and Va)IV endmembers of the

10H(o), 18R(⟡), 14H(x), and 24R(∇) LPSOs, (b) formation energies (in kJ/mole-

atom) of the GdIIAlIII(Mg(x), Gd(∇), Al(⟡), and Va(o))IV endmembers at 0 K in

compositional space. ........................................................................................................ 40

Figure 3.6 Isothermal sections of the Mg-Al-Gd system at 673 K (a) and 798 K (b). All

experiment data (the thick lines and the symbols) at 673 K were measured by De

Negri et al.[110] (∇ : Al3Mg2 + FCC Al + Lav C36, ∆: GdMg + GdMg3, □: MgGd,

⟡: GdMg + AlGd2, ⧖:GdMg + Lav C15 + GdMg3), those at 798.15K were taken

from Kishida et al.[6], [109], including HCP Mg + Al2Gd (Laves C15) + 18R LPSO

(○[109]) and HCP Mg + Mg5Gd + 18R LPSO (∇[6] and ⟡[6]) phases. ........................ 41

Figure 3.7 Mg-corner of the isothermal sections of the Mg-Al-Gd system at (a) 838.15 K,

(b) 823.15 K, (c) 798.15 K, (d) 773.15 K, (e) 723.15 K, and (f) 673.15 K, with

experimental compositions from Lu et al.[111] at 838.15 K (∇) with HCP Mg +

Al2Gd (Laves C15) + 18R LPSO phases in equilibrium, at 823.15 K from Dai et

al.[112] (*) with 18R LPSO phase composition of Mg–7.9 at.% Al–10.9 at.%

(Gd+Y), at 798.15 K from Kishida et al. with HCP Mg + Al2Gd (Lav C15) + 18R

LPSO (○[109]) and HCP Mg + Mg5Gd + 18R LPSO (∇[6] and ⟡[6]) phases in

equilibrium, and at 773.15 K from Gu et al.[113] with 18R LPSO, respectively. The

small triangles represent the composition ranges of GdIIAlIII(Mg, Gd and Al)IV

endmembers. .................................................................................................................... 43

Figure 3.8 An enlarged view of the isothermal section of the Mg-Al-Gd system at 798.15

K, showing the composition homogeneity range of the 18R LPSO phase. Blue

triangle indicates the composition ranges of GdIIAlIII(Mg, Gd, Al, and Va)IV

endmembers as the same triangle as Figure 3.5. .............................................................. 44

Figure 3.9 Isopleth sections of the Mg-Al-Gd phase diagram with the molar ratio of

Al:Gd being 0.7 (a) and an enlarged view of the Mg-rich region (b), with

experimental compositions from Lu et al.[111] at 838.15 K (+) with HCP Mg, Lav

C15 and LPSO (18R) phases in equilibrium, and from Kishida et al.[17] at 673.15K

(*) with HCP Mg, Mg5Gd and LPSO (14H + 18R) phases, respectively. ....................... 45

Figure 4.1 Schematics of (a) Vickers tip geometry, (b) geometry changes during

indentation, and (c) a side view of indentation. ............................................................... 48

Figure 4.2 Hardness comparisons of (a) FCC, (b) BCC, and (c) HCP materials with

respect to experimental data. Solid and open symbols represent the predicted values

using elastic properties from experiments and first-principles calculations. Red

x

dashed lines indicate value equality, vertical dotted lines connect the hardness

between the slip systems. ................................................................................................. 53

Figure 4.3 Hardness comparisons of ceramic materials with respect to experimental data.

Solid and open symbols represent the predicted values using elastic properties from

experiments and first-principles calculations. Red dashed lines indicate value

equality, vertical dotted lines show the differences between glide edge and shuffle

screw slip systems. ........................................................................................................... 56

Figure 4.4 Hardness comparisons of all tested materials with respect to experimental data.

Solid and open symbols represent the predicted values using elastic properties from

experiments and first-principles calculations. The experiments for all materials both

hardness and elastic properties data from Table 4.6. Red dashed lines indicate value

equality. ............................................................................................................................ 57

Figure 4.5 Comparison of Peierls-Nabarro flow stress at 0 K with experimental yield

stress at low temperatures (4~77K) as a function of dislocation width. Data and

references are listed in Table 4.3. ..................................................................................... 58

Figure 4.6 Indentation ductility index as a function of dislocation width at room

temperature....................................................................................................................... 60

Figure 4.7 (a) Stress(𝜏)-strain(𝛾) curve during shear deformation. (b) typical Load(F)-

displacement(h) curve during indentation process. .......................................................... 62

Figure 4.8 Typical Load-displacement curves (a) F-h curve and (b) 𝐹-h curve. Red lines

are loading curves and blue lines are unloading curves, and green dot line represents

only the plastic contribution from Equation 4.16. ............................................................ 65

Figure 4.9 Peierls-Nabarro flow stress (𝜏𝑃𝑁) and ideal shear stress (𝜏𝑇) at 4.7 K and 7 K

in terms of dislocation width (𝑤0) from Refs.[206]–[209] shown in Table 4.3 and

Table 4.5........................................................................................................................... 68

Figure 4.10 Comparison of ℎ𝑇/ℎ𝑝 ratio between experiment from Table 4.4 and the

present model. .................................................................................................................. 70

Figure 4.11 Exponential relationship of the scaling factor c (from Equation 4.31) for FCC

metals with data from Table 4.4. ...................................................................................... 72

Figure 4.12 Plots of parameter c from experimental data (Equation 4.31) with data from

Table 4.4 (a) with respect to 𝑏/𝑠2 and (b) with respect to the model (Equation 4.36). ... 73

Figure 4.13 Comparison of experimental and calculated (VRH averaged) shear moduli

with elastic stiffness constant data from Shang et al.[42]. ............................................... 83

Figure 4.14 Comparison of experimental and calculated (VRH averaged) bulk moduli

with elastic stiffness constant data from Shang et al.[42]. ............................................... 84

xi

Figure 5.1 Activation energy for self-diffusion modeling. All the data and references are

in Table 5.3. ..................................................................................................................... 88

Figure 5.2 Predicted temperature dependent hardness of FCC metals. All the

experimental data is from Lozinskii[313]. ....................................................................... 90

Figure 5.3 Predicted temperature dependent hardness of FCC Rh (a) and Ir (b), and BCC

Mo (c) and W (d) metals. All the experimental data is from Lozinskii[313](■) and

Stephens et al.[316](▲). .................................................................................................. 92

Figure 5.4 Predicted temperature dependent hardness of HCP metals. All the

experimental data is from Lozinskii[313]. ....................................................................... 94

Figure 5.5 (a) Predicted temperature dependent flow stress of TiC comparison with

experiment results from Kurishita et al.[320] and (b) Predicted temperature

dependent hardness of TiC comparison with single crystal micro-Vickers hardness

(■, Expt.1) from Kumashiro et al.[321], single crystals of Vickers hardness (●,

Expt.2), equivalent x-cylinder hardness (▲, Expt.3), polycrystalline TiC equivalent

x-cylinder hardness (▼, Expt.4), and equivalent x-wedge hardness (◆, Expt.5),

experiment results from Atkins et al.[299], Vickers hardness of TiC0.94 (▶, Expt.6)

from Samsonov et al.[322] and Vickers hardness of TiC0.96 (★, Expt.7) from

Kohlstedt et al.[323] and predicted temperature dependent hardness of Si (c) and Ge

(d). The grey region in (c) is the phase transformation region from Domnich et

al.[159]. Experimental data of Si and Ge is from Atkins et al.[299]................................ 97

Figure 5.6 Validation of the hardness model from this work. (a) hT/hp and (b) hardness

between this model and experimental results. .................................................................. 104

Figure 5.7 Comparison of temperature-dependent hardness of BCC W between a) using

temperature-dependent elastic properties and b) using fixed elastic properties at 0 K.

the temperature-dependent elastic properties of BCC W is from Hu et al.[348]. ............ 105

Figure 6.1. Grain size dependent hardness of FCC Cu. Grain size (G) dependent hardness

(solid shapes) are from Chen et al.,[406] Sanders et al.,[407] Jiang et al.,[408]

Agnew et al.,[409] Gray et al.,[410] Valiev et al.,[411] Haouaoui et al.[412], and

Suryanarayanan et al.[413] twin size(T) dependent hardness are from You et

al.[403], Lu et al.[404] and Anderoglu et al.[405]. .......................................................... 112

Figure 6.2 The slip direction (plane) angles(𝜃𝑖) generated by twin boundaries in (a)

twinned FCC Cu and (b) twinned carbon diamond cubic. ............................................... 115

Figure 6.3 The slip direction (plane) angles(𝜃𝑝) of full and partial dislocations in FCC

and diamond cubic. .......................................................................................................... 116

Figure 6.4 Normalized 𝑏𝑖/𝑠𝑖 (with respect to that of each structures) changes of each

layers in (a) twinned carbon diamond cubic and (b) FCC Cu. ......................................... 117

xii

Figure 6.5 Schematics of the method of modeling of b/s in twinned structures. ..................... 118

Figure 6.6 Differential charge density plots of (a) carbon diamond cubic (reference state),

(b) FCC Cu (reference state), (c) twinned carbon diamond cubic and (d) twinned

FCC Cu structures. Red arrows indicate the close-up view of twin boundary area.

Isosurfaces are 0.0065 (e/Å 3) and the atom sizes are exaggerated for better

visualization. .................................................................................................................... 121

Figure 6.7 𝑏𝑠𝑑/𝑠𝑠𝑑 changes by the various twin layer distances in (a) twinned carbon

diamond cubic and (b) FCC Cu. ...................................................................................... 123

Figure 6.8 Hardness of diamond carbon as a function of twin bilayer distance. Expt.1 and

2 are from Huang et al.[426] and Irifune et al.[427], respectively. Open blue

triangles are obtained from relaxed structures calculated from first-principles

calculations....................................................................................................................... 124

Figure 6.9 Hardness of FCC (a) Cu and (b) Ag as a function of twin bilayer distance. For

(a) FCC Cu, Expt.1 from You et al.[403], Expt.2 from Lu et al.[404] and Expt.3

from Anderoglu et al.[405] are included. For (b) FCC Ag, Expt.1 from Bufford et

al.[428], Expt.2 from Bufford et al.[429] and Expt.3 from Furnish et al.[430] are

included. Red dash line is the hardness of their bulk state. .............................................. 124

Figure 6.10 Hall-Petch relationship in hardness of (a) carbon diamond, (b) FCC Cu and

(c) FCC Ag as a function of twin bilayer distance. References are from those in

Figure 6.8 and Figure 6.9. ★ in the plots are the hardness of bulk state, and these are

from Teter[117] for carbon diamond, from Samsonov[274] for FCC Cu and Ag. Red

dash lines are the slope for Hall-Petch relation. ............................................................... 126

Figure 7.1 Slip systems of (a) 18R and (b) 14H LPSOs. Thin solid lines are the pyramidal

slip, black thick lines are the slip direction within FCC layers, red thick lines are the

basal slip, and dash lines are the L12 cluster. .................................................................. 128

Figure 7.2 𝑏𝑖/𝑠𝑖 changes of (a) 18R and (b) 14H LPSOs. Pyramidal slip on {1108} for

18R and prismatic slip on {1100} for 14H are applied. .................................................. 129

Figure 7.3 Hardness prediction of 18R and 14H LPSO phases. Expt. 1 to Expt. 6 are from

[432] (Expt. 1), [433] (Expt. 2), [434] (Expt. 3), [435] (Expt. 4), [436] (Expt. 5),

[437] (Expt. 6), and the hardness of polycrystalline Mg as a reference[274] (Expt. 7),

respectively. ..................................................................................................................... 130

xiii

LIST OF TABLES

Table 2.1 Calculated lattice parameters (a, b, and c in Å , and , β , γ, in degrees, o ), area

per cluster (Acluster in Å 2/cluster) and equilibrium volumes (Veq in Å 3/atom) of the

LPSO phases in the Mg-Gd-Al alloys at 0 K. The formation energies (ΔEForm,

meV/atom) are calculated based on Equation 2.1. ........................................................... 9

Table 2.2 Calculated lattice features of LPSO structures. 𝑑𝑖𝑛𝑡𝑟𝑎𝑐𝑙𝑢𝑠𝑡𝑒𝑟(Å ) and

𝑑𝑖𝑛𝑡𝑒𝑟𝑐𝑙𝑢𝑠𝑡𝑒𝑟 (Å ) are the 2NN RE-RE intracluster and intercluster

distances, 𝑤cluster(Å ) and ℎcluster(Å ) are the L12 cluster width and height. ............... 11

Table 2.3 Calculated elastic properties of LPSO structures of the Mg-Gd-Al alloys at 0

K, including elastic stiffness constants (Cij's), Young's modulus (E), bulk modulus

(B) from both VRH approach and EOS fitting, and shear modulus (G) from the VRH

approach. The unit for each elastic property is GPa. ....................................................... 13

Table 3.1 Gibbs energies of LPSO end-members obtained from the Debye model, defined

as 𝐺𝐿𝑃𝑆𝑂 − 𝐻𝑆𝐸𝑅 = 𝑎 + 𝑏𝑇 + 𝑐𝑇ln𝑇 + 𝑑𝑇2 + 𝑒𝑇 − 1 + 𝑓𝑇3(J/mole-atom),

where HSER is reference state from SGTE[82]. ................................................................ 32

Table 3.2 Interaction parameters in individual sublattices (kJ/mol-atom). .............................. 34

Table 3.3 Calculated lattice parameters of pure elements in comparison with

computational and experimental data in the literature. .................................................... 35

Table 3.4 Formation energies (Eform in kJ/mole-atom) at 0 K and the scaling factors (s) of

the endmembers, see Equation 3.12, with the elastic constants from Kim et al.[81]. ...... 36

Table 4.1 Slip systems of different crystal structures. ............................................................. 50

Table 4.2 Slip systems at room temperature in BCC metals. ................................................... 53

Table 4.3 Comparison of PN flow stress at 0K with experimental yield stress (𝜏𝑌𝐺) at

low temperatures (4~77 K) as a function of dislocation width. ....................................... 58

Table 4.4 Experimental and calculated ℎ𝑇/ℎ𝑝 and parameter c for various materials. .......... 64

Table 4.5 Comparison of 𝜏𝑃𝑁𝐺 and 𝜏𝑇𝐺 ................................................................................ 68

Table 4.6 Hardness comparison between the present and previous models with slip

systems and elastic properties with S for screw dislocation. ........................................... 74

Table 5.1 The materials’ information used in Figure 5.2 to Figure 5.5. .................................. 95

Table 5.2 Crystal structures and their slip systems. ................................................................. 104

Table 5.3 Self-diffusion activation energy modeling. .............................................................. 106

Table 6.1 The angle (𝜃𝑝) of full and partial dislocations in structures. ................................... 115

xiv

ACKNOWLEDGEMENTS

There are many people that I would like to express my appreciation. More specifically, I

would like to thank four groups of people, without whom this dissertation would not have been

possible: my advisor, my thesis committee members, my lab mates, and my family.

I would like to express thanks to my advisor Dr. Zi-Kui Liu for mainly two aspects. First,

his deep knowledge in thermodynamics inspires me a lot, which now becomes an important part

of my knowledge. Second, he always mentions “critical thinking and communication with others”

that remind me all the times during my study.

In addition, I would like to thank the rest of my committee members, Dr. Adri van Duin,

Dr. Ismaila Dabo, Dr. Hojong Kim and Dr. Laszlo Kecskes for their time, encouragements and

suggestions during serving on my dissertation committee. Especially, I would like to thank to Dr.

Kecskes for the intuitive discussions.

I would like to thank many of my colleagues in the Phases Research Lab for their help

and discussions. Dr. Xuan Liu and Dr. Austin Ross taught me Thermo-Calc and has given many

useful advices since I joined the group. Dr. Shun-Li Shang and Dr. Yi Wang gave me suggestions

on calculation skills. The help and discussions from Dr. Yongjie Hu for dislocation study, Dr. Bi-

Cheng Zhou and Dr. Cassie Marker for thermodynamic modeling, Dr. Richard Otis for

thermodynamic intuition, Dr. Pinwen Guan for calculation details and skills, Brandon Bocklund

for python coding, Jorge Paz Soldan (dynamic duo) for thermodynamic discussions and Matthew

Feurer for DFT-TK help. Their help and discussions are priceless to me.

Lastly, I would like to express my deepest thanks to my lovely wife, Dr. Jungwan Yoon

for being with me always, and to my parents, Donghyune Kim and Kwangsook Ahn for their

supports.

1

Chapter 1

Introduction

1.1 Motivation

Magnesium (Mg) and its alloys are important structural materials in transportation,

aerospace, and consumer electronic industry applications[1]–[3] since they are lightweight.

However, due to their low ductility and low mechanical strength,[4], [5] which stem from a limited

number of slip systems for the hexagonal-close-packed (HCP) crystal structure, their applications

are limited, and many researchers have put great effort to improve the properties of Mg alloys. One

potential solution to overcoming these issues in Mg alloys is to introduce the face-centered-cubic

(FCC) stacking layers with an ABCABC (here, A, B, and C are different close-packed layers)

stacking sequence within the ABABAB HCP stacking layers, i.e., forming the long periodic

stacking ordered (LPSO) phases via TM (Transition Metal)-RE (Rare Earth) solute atoms

clusters[6]–[9]. It has been shown that the presence of LPSO phases improves the tensile strength,

hardness and ductility[10]–[13].

Among various candidates of Mg-TM-RE LPSO phases, Mg-Al-Gd LPSO phases have

obtained considerable attention for two major reasons. First, the alloying elements of Al are much

lighter than Zn or other TM elements. By alloying with Al, the LPSOs will be lighter than other

LPSO phases, that is the major concern for making lightweight structural metal. Second, among

various LPSO phase candidates in Mg-Al-RE ternary systems, only Mg-Al-Gd LPSOs are found

to be stable at finite temperature ranges[14].

In order to investigate the formation of LPSO phases in Mg-Al-Gd system, the phase

stability of LPSOs in Mg-Al-Gd system should be investigated first since the phase equilibria will

2

help to understand the conditions of alloy processing, i.e., temperature and composition ranges.

However, there has been no available thermochemical data for LPSO phases, and no research on a

thermodynamic model with solubilities of LPSO phases in the Mg-Al-Gd system. Furthermore, in

order to predict mechanical properties of LPSOs, i.e. hardness, a unified model which enables to

predict hardness not only for pure metals but also for complex compounds such as LPSO phases,

should be developed. So far, there is no hardness model for metals and alloys other than the

empirical expression of 𝐻𝑣 = 3𝜎𝑌 . Overall, although considerable efforts have been made to

understand and improve phase stability and mechanical properties of LPSO phases, there is still

lack of clear and systematic understanding of the relationship between the structure and the

resulting phase equilibria and mechanical properties, i.e., hardness, which will be the focus of the

present work.

1.2 Overview

The ultimate goal of this dissertation is to give a comprehensive description of the phase

equilibria of LPSO phases in Mg-Al-Gd system by a combined CALPHAD-DFT methodology,

and is to predict the mechanical properties of LPSOs by modeling hardness of polycrystalline

materials which implemented plastic deformations into the model. To achieve these goals, the

related methodologies are developed in the following chapters. Specifically, in Chapter 2, the

elastic properties of LPSO phases in Mg-Al-Gd system such as elastic stiffness constants were

calculated based on the first-principles calculations. In addition to this, orientation dependent shear

and Young’s moduli were discussed. In Chapter 3, the phase equilibria of the Mg-Al-Gd system

with LPSO phases were modeled by a combined CALPHAD-DFT methodology. Due to the lack

of sufficient experimental data, first-principles calculations played an important role in modeling

this system.

3

In Chapter 4, a Vickers hardness model for polycrystalline materials was developed since

there is no available model for predicting hardness of any metallic phases including LPSO phases.

The Vickers hardness model considers both elastic and plastic deformation of materials by

implementing Peierls-Nabarro flow stress in order to capture the plastic deformation of materials.

The developed hardness model agrees well with experimental results of metals and ceramics which

indicates the reliability of the model is from below 0.1 GPa to over 100 GPa. Especially the active

slip system as well as melting temperature and elastic properties (shear and bulk moduli and

Poisson’s ratio) played a significant role in determining materials’ hardness. In Chapter 5, a

temperature-dependent hardness model was developed by implementing diffusion mechanisms

such as dislocation (pipe), mono- and di-vacancy diffusions into the hardness model which was

developed in Chapter 4, since LPSOs are intermediate temperature phases. Especially, the modeling

of the activation energy for self(mono-vacancy)-diffusion, which is based on the earlier belief of

the Van Liempt rule, helped to simplify the model. This chapter was explained by the factors that

affect the hardness of materials as a function of temperature with some examples. The developed

temperature-dependent hardness model agrees well with 16 examples including FCC, BCC, HCP

and ceramic materials. In Chapter 6, the twin layer dependent hardness model was developed since

LPSO phases are layered structures. This chapter discussed how the active slip systems are changed

by the twin boundaries. This model has great agreement with experimental results of carbon

diamond cubic and FCC metals. In Chapter 7, the hardness of LPSOs in Mg-Al-Gd ternary systems

was predicted based on the hardness models developed in Chapters 4-6. In Chapter 8, the

conclusions were drawn and the future works were discussed.

4

Chapter 2

Elastic Properties of Long Periodic Stacking Ordered Phases in Mg-Al-Gd

Alloys: A First-Principles Study

2.1 Introduction

It has been shown that the presence of LPSO phases improves the tensile strength and

ductility[10]. For example, the Mg97Zn1Y2 (at.%) alloy, which includes the LPSO phase, reaches a

high yield strength of 480-610 MPa and an elongation of 5~16%, respectively[10]. While it is

known that such enhanced mechanical properties result from LPSO phases as well as grain

refinement,[11]–[13] the underlying mechanism of this phenomenon has not yet been fully

explained due to the plastic behavior of LPSOs. Since the mechanical properties can be estimated

from slip systems and elastic properties such as Poisson’s ratio and shear modulus[15], [16], the

elastic properties are one of the important factors in order to understand plastic deformations.

In order to understand the elastic properties of LPSOs, it is crucial to clarify the effects of

the crystal structures of LPSOs, especially the ordering of solute atoms in LPSOs, since the elastic

properties of LPSOs are largely affected by the nature of bonding, which is determined by the

crystal structures. Reported LPSO phases in the Mg-TM-RE ternary systems consist of 5-8 atomic

layers in the structural block (SB), a unit with the minimum number of stacking layers that includes

one set of stacking faults[6]–[8], [17]–[19]. For example, there are six layers in the 18R LPSO SB,

and seven layers in the 14H LPSO SB. These SBs are also referred to as the 10H, 18R, 14H, and

24R poly-types according to the Ramsdell notation[9], [20], [21], where the number represents total

layers in the repeating unit cell, and the letters H and R represent the hexagonal and rhombohedral

symmetries, respectively. Moreover, the solute atoms located in the 4-continued atomic layers and

5

these solute atoms form a specific in-plane ordering of the L12 cluster[6]–[9]. Kimizuka et al.[22]

verified the formation of the L12-type clusters in terms of Gd and Al in the Mg-Gd-Al system, using

the cluster expansion method. From the images, several investigators[17]–[19], [23] used scanning

transmission electron microscopy (STEM) to verify the L12 clusters of solute atoms in the stacking

fault regions of the LPSO phase and describe the two-dimensional (2D) close-packed in-plane

ordering of these L12 clusters. Furthermore, an interstitial atom (Mg, RE, or TM) at the center of

the L12 cluster has been observed in the Mg-Y-Zn system through STEM images[18], and

suggested by density functional theory (DFT) based first-principles calculations[24].

Furthermore, it is also crucial to clarify the effect of the L12 cluster interactions and the

contribution of the interstitial atom in the cluster since the L12 cluster is the key lattice feature of

the crystal structures of LPSOs. Recently, Kimizuka et al. described the cluster interaction of LPSO

phases with or without interstitial atom and the changes of RE-RE intracluster and intercluster

bonding distances[25]. The intracluster and intercluster distances represent the average 2nd nearest

neighbor (2NN) distances of RE atoms within and between the clusters, respectively. The smaller

cluster interaction undergoes the larger contraction of RE-RE intracluster bonding distance among

the L12 clusters with an interstitial atom. The intracluster bonding distance is related to the size of

cluster. Furthermore, Tane et al. reported there is a relationship between cluster interaction energy

or cluster density and elastic properties such as Young’s modulus and shear modulus[26]. The

findings of the previous literature imply that changes of the bonding environment around the L12

cluster should influence the elastic properties of LPSO phases via the changes of cluster interaction

or cluster density.

The present work aims to study the elastic properties of the Mg-Gd-Al LPSO phases (10H,

18R, 14H, and 24R) using first-principles calculations, where all of the possible L12-type clusters

with and without interstitial atoms are considered. The interstitial atoms Mg, Gd, and Al are

6

denoted as Mg-int., Gd-int. and Al-int., respectively. The predicted elastic properties of the LPSO

phases are interpreted by examining atomic bonding environments around the L12 clusters and

electronic structures.

2.2 Computational Methods

The crystal structure of 14H LPSO is P63/mcm, proposed by Egusa and Abe[23] based on

various theoretical and experimental results[17], [19], [23]. Space groups 18R and 24R LPSOs are

designated as C2/m[6], [18], [19], [23]. The crystal structures of 10H LPSO phase is designated as

Cmce, which was suggested by Kishida et al.[18] according to the stable LPSO phase in the Mg-

Y-Zn system. In order to describe the L12-type clusters in the DFT calculations, the number of

atoms in each LPSO phase are 240 (10H LPSO), 168 (14H LPSO), 144 (18R LPSO), and 192 (24R

LPSO), respectively, associated with the Gd and Al clustering in the stacking fault layers[18], [23].

First-principles calculations are conducted by using the Vienna Ab-initio Simulation

Package (VASP) [27], [28]. Electron-ion interactions are described by the projector augmented-

wave (PAW) method[29]. In order to describe the electron interactions including exchange and

correlation, the generalized gradient approximation (GGA) as implemented by Perdew, Burke, and

Ernzerhof (PBE)[30] is used. Plane wave cutoff energies of 350 eV are consistently used for all the

calculations, which are 1.3 times higher than the recommended ones by the VASP[31]. For HCP-

Mg with 2 atoms in the supercell, the 29 × 29 × 16 Γ-centered k-point grids are implemented. For

the crystal structures of LPSO phases, we use the 3 × 5 × 2 (10H LPSO with 240 atoms in the

supercell), 6 × 6 × 2 (14H LPSO with 168 atoms), 6 × 3 × 3 (18R LPSO with 144 atoms), and 3 ×

2 × 2 (24R LPSO with 192 atoms) Γ-centered k-point grids, respectively. The k-mesh guarantees

errors below 0.1 meV/atom (0.2 meV/atom for 24R LPSO due to the computational resource

limitations). The f-electrons of the Gd element are treated as core electrons, an approximation that

7

has shown to produce accurate thermodynamic properties for lanthanide-containing structures[32]–

[35]. After full relaxations, a final static calculation using the tetrahedral method with Blöch

corrections[36] is applied to ensure the accuracy of total energy. The energy convergence criterion

of the electronic self-consistency is set as 10-6 eV/atom for all of the calculations. The contour plots

of the differential charge density are generated using VESTA[37], [38].

The formation energies and the contribution of interstitial atom of LPSOs[14] are

calculated by Equation 2.1 and Equation 2.2, and listed in Table 2.1.

Equation 2.1 𝑬𝒇𝒐𝒓𝒎(𝑳𝑷𝑺𝑶) = 𝑬(𝑳𝑷𝑺𝑶) − 𝟏

𝑵∑ 𝑵𝒊𝑬𝒊𝒊

where Ei is the total energy of stable bulk state of species per atom of species i and Ni is the number

of atom of species i.

Equation 2.2 ∆𝑬𝒊𝒏𝒕𝒊 =

𝑬(𝑳𝑷𝑺𝑶+𝑵𝒊×𝒊𝒏𝒕)−𝑬(𝑳𝑷𝑺𝑶)− 𝑵𝒊𝑬𝒊

𝑵𝒊

In the present work, elastic stiffness constants are predicted at 0 K via DFT-based first-

principles calculations in terms of the stress–strain method [39]. To determine elastic constants for

a crystal from first-principles and Hooke’s law, a set of strains, expressed in Voigt notation with 𝜀

= (𝜀1, 𝜀2, 𝜀3, 𝜀4, 𝜀5, 𝜀6) (where 𝜀1, 𝜀2, 𝑎𝑛𝑑 𝜀3 are the normal strains and the others are the shear

strains), are placed on a crystal with lattice vectors R,

Equation 2.3 𝑹 = (

𝒂𝟏 𝒂𝟐 𝒂𝟑𝒃𝟏 𝒃𝟐 𝒃𝟑𝒄𝟏 𝒄𝟐 𝒄𝟑

)

After deformation, the resulting lattice vectors, R’, can be expressed as

Equation 2.4 𝑹′ = 𝑹(

𝟏 + 𝜺𝟏 𝝐𝟔/𝟐 𝝐𝟓/𝟐𝝐𝟔/𝟐 𝟏 + 𝜺𝟐 𝝐𝟒/𝟐𝝐𝟓/𝟐 𝝐𝟒/𝟐 𝟏 + 𝜺𝟑

)

Correspondingly, stresses 𝜎 = (𝜎1, 𝜎2, 𝜎3, 𝜎4, 𝜎5, 𝜎6) for each set of strains can be calculated

using first-principles to determine the 6 × 6 elastic stiffness constants matrix (C),

8

Equation 2.5

(

𝝈𝟏,𝟏𝝈𝟐,𝟏𝝈𝟑,𝟏𝝈𝟒,𝟏𝝈𝟓,𝟏𝝈𝟔,𝟏

𝝈𝟏,𝒏𝝈𝟐,𝒏𝝈𝟑,𝒏𝝈𝟒,𝒏𝝈𝟓,𝒏𝝈𝟔,𝒏)

=

(

𝑪𝟏𝟏𝑪𝟐𝟏𝑪𝟑𝟏𝑪𝟒𝟏𝑪𝟓𝟏𝑪𝟔𝟏

𝑪𝟏𝟐 𝑪𝟐𝟐𝑪𝟑𝟐𝑪𝟒𝟐𝑪𝟓𝟐𝑪𝟔𝟐

𝑪𝟏𝟑𝑪𝟐𝟑𝑪𝟑𝟑𝑪𝟒𝟑𝑪𝟓𝟑𝑪𝟔𝟑

𝑪𝟏𝟒𝑪𝟐𝟒𝑪𝟑𝟒𝑪𝟒𝟒𝑪𝟓𝟒𝑪𝟔𝟒

𝑪𝟏𝟓𝑪𝟐𝟓𝑪𝟑𝟓𝑪𝟒𝟓𝑪𝟓𝟓𝑪𝟔𝟓

𝑪𝟏𝟔𝑪𝟐𝟔𝑪𝟑𝟔𝑪𝟒𝟔𝑪𝟓𝟔𝑪𝟔𝟔)

(

𝜺𝟏,𝟏𝜺𝟐,𝟏𝜺𝟑,𝟏𝜺𝟒,𝟏𝜺𝟓,𝟏𝜺𝟔,𝟏

𝜺𝟏,𝒏𝜺𝟐,𝒏𝜺𝟑,𝒏𝜺𝟒,𝒏𝜺𝟓,𝒏𝜺𝟔,𝒏)

With n sets of strains 𝜀 (an n × 6 matrix, in general the linearly independent sets should be

6), the elastic stiffness constants matrix (C) are determined by C = 𝜀−1 𝜎. To obtain the elastic

stiffness components, linear strains of 𝜀 = 0.01 were applied to the cell in the reference

configuration with respect to the six independent components. Based on the single crystal elastic

stiffness constants, the aggregate properties associated with polycrystals, such as shear (G) moduli

are estimated by means of the Voigt-Reuss-Hill (VRH) approximation[40], [41], here, the average

results from VRH approach are reported, bulk moduli (B) are estimated by 4-parameter Birch-

Murnaghan (BM4) equation of states (EOS) fitting[42]. Correspondingly, Young's (E) and Bulk

moduli, and Poisson's ratio (𝜐) are also estimated by the following equations,

Equation 2.6 𝑬 = (𝟗 𝐁𝐆)

(𝟑 𝐁+𝐆)

Equation 2.7 𝝊 = (𝟑𝑩−𝟐𝑮)

(𝟔𝑩+𝟐𝑮)

2.3 Results and Discussion

2.3.1 Structural Analysis of the LPSO Phases

Table 2.1 summarizes the calculated lattice parameters, equilibrium volumes, formation

energies and in-plane areas per cluster (also see Figure 2.1e) of LPSO structures predicted by the

EOS fitting. It is found that the present prediction of lattice parameters of LPSO structures in Mg-

Gd-Al system agree well with previous results[9].

9

Table 2.1 Calculated lattice parameters (a, b, and c in Å , and , β , γ, in degrees, o ), area

per cluster (Acluster in Å 2/cluster) and equilibrium volumes (Veq in Å 3/atom) of the LPSO

phases in the Mg-Gd-Al alloys at 0 K. The formation energies (ΔEForm, meV/atom) are

calculated based on Equation 2.1.

Type Int.

atom a b c α β γ Acluster Veq Eform Eint Reference

10H

no 11.24 19.51 26.26 90 90 90 109.65 24.0 -83.7

Al 11.23 19.46 26.16 90 90 90 109.27 23.4 -116.6 -2.09

Mg 11.26 19.52 26.22 90 90 90 109.90 23.6 -117.8 -2.16

Gd 11.27 19.60 26.39 90 90 90 110.45 23.9 -118 -2.18

18R

no 11.23 19.46 16.13 90 76.4 90 109.27 23.8 -71

Al 11.22 19.43 16.09 90 76.5 90 109.00 23.4 -98.6 -2.08

Mg 11.24 19.47 16.12 90 76.5 90 109.42 23.5 -99.3 -2.14

Gd 11.27 19.50 16.21 90 76.5 90 109.88 23.7 -99.8 -2.17

11.2 19.4 16.2 90 76.7 90 Ref.[9]

14H

no 11.21 - 36.44 90 90 120 108.83 23.6 -61.9

Al 11.21 - 36.40 90 90 120 108.83 23.3 -84.7 -2.00

Mg 11.23 - 36.43 90 90 120 109.22 23.4 -85.5 -2.07

Gd 11.25 - 36.57 90 90 120 109.61 23.6 -86.4 -2.14

11.2 - 37.2 90 90 120 Ref.[9]

24R

no 11.20 19.41 21.12 90 79.8 90 108.70 23.4 -53.9

Al 11.20 19.37 21.11 90 79.8 90 108.47 23.2 -74.4 -2.04

Mg 11.22 19.41 21.13 90 79.8 90 108.89 23.3 -74.9 -2.09

Gd 11.24 19.43 21.22 90 79.8 90 109.20 23.5 -75.5 -2.15

10

Figure 2.1 The LPSO structures of 10H (a), 18R (b), 14H (c), and 24R (d) together with the

in-plane L12 cluster ordering (e) and the Gd8Al6 L12 cluster with an interstitial (int.) atom

Gd, Mg or Al (f). Blue box stands for the unit cell of each LPSO structures and the red bracket

with SB stands for structural block for each LPSO structure. 𝒅𝒊𝒏𝒕𝒓𝒂𝒄𝒍𝒖𝒔𝒕𝒆𝒓 and 𝒅𝒊𝒏𝒕𝒆𝒓𝒄𝒍𝒖𝒔𝒕𝒆𝒓 stands for the 2NN RE-RE intracluster and intercluster distances, 𝒘𝐜𝐥𝐮𝐬𝐭𝐞𝐫 and 𝒉𝐜𝐥𝐮𝐬𝐭𝐞𝐫 stands

for the L12 cluster width and height.

It can be seen that the lattice parameters of all LPSO supercell structures (larger than 11.20

Å ) are larger than those of HCP Mg, which corresponds to 11.07 Å (2√3 𝑎𝑀𝑔with aMg being 3.196

Å ). This represents that in the normal stacking layers (ABAB…), Mg atoms endure the tensile

stresses along the [1120] and [1010] directions due to the L12 clusters when it is compared with

HCP Mg structure. Among the LPSOs, the lattice parameter, a, is the largest (11.24 Å ) for the 10H

11

LPSO phase and decreases as the number of layers in SB increases to 11.20 Å for the 24R LPSO

phase. Since the distance between clusters is proportional to the lattice parameter, the number of

clusters in basal plane increases as the number of layers in SB increases. Furthermore, we also

examined the lattice relaxation of L12 clusters in the LPSO phases, since the cluster interaction can

be quantified by the changes of cluster dimensions. Kimizuka et al.[25] examined the 2NN RE-RE

bonding distances (intercluster and intracluster 2NN RE-RE distances as listed in Table 2.2) and

their effects on the intercluster interactions. Based on their work, it is also found that the types of

the interstitial atom induce changes of 2NN RE-RE bonding distances and intercluster interactions.

In this work, we examine the lattice relaxations of L12 cluster such as the in-plane area per cluster,

Acluster, the L12 cluster width (𝑤cluster: body diagonal distance between Gd atoms within stacking

fault region), and the L12 cluster height (ℎcluster: the distance between top and bottom Gd atoms

in the L12 cluster), listed in Table 2.2. It is found that the cluster with smaller interstitial atom, Al,

undergoes further inward contraction of the cluster.

Table 2.2 Calculated lattice features of LPSO structures. 𝒅𝒊𝒏𝒕𝒓𝒂𝒄𝒍𝒖𝒔𝒕𝒆𝒓 (Å ) and 𝒅𝒊𝒏𝒕𝒆𝒓𝒄𝒍𝒖𝒔𝒕𝒆𝒓 (Å ) are the 2NN RE-RE intracluster and intercluster

distances, 𝒘𝐜𝐥𝐮𝐬𝐭𝐞𝐫(Å ) and 𝒉𝐜𝐥𝐮𝐬𝐭𝐞𝐫(Å ) are the L12 cluster width and height.

Int.

type no no no no Al Al Al Al Mg Mg Mg Mg Gd Gd Gd Gd

SB 5 6 7 8 5 6 7 8 5 6 7 8 5 6 7 8

𝑑𝑖𝑛𝑡𝑟𝑎𝑐𝑙𝑢𝑠𝑡𝑒𝑟 4.1

0

4.1

2

4.1

4

4.1

3

4.0

8

4.1

0

4.1

1

4.1

0

4.1

0

4.1

2

4.1

2

4.1

1

4.1

3

4.1

6

4.1

6

4.1

4

𝑑𝑖𝑛𝑡𝑒𝑟𝑐𝑙𝑢𝑠𝑡𝑒𝑟 5.0

9

5.0

6

5.0

3

5.0

3

5.0

9

5.0

7

5.0

6

5.0

6

5.1

0

5.0

7

5.0

6

5.0

7

5.1

0

5.0

9

5.0

9

5.0

8

𝑤cluster 7.13

7.16

7.19

7.17

7.12

7.12

7.13

7.12

7.14

7.15

7.15

7.13

7.18

7.19

7.19

7.16

ℎcluster 7.2

7

7.2

6

7.2

5

7.2

9

7.1

7

7.1

6

7.1

5

7.1

7

7.2

2

7.2

1

7.1

8

7.2

1

7.2

7

7.2

9

7.2

4

7.2

9

12

2.3.2 Elastic Properties of the LPSO Phases

Calculated elastic properties Cij, B(EOS), B(VRH), G, E and 𝜈 (Poisson ratio) of the LPSO

phases are summarized in Table 2.3. For the comparison reason, the calculated elastic stiffness

matrix of 10H LPSO supercell and 18R and 24R LPSO supercell, orthorhombic and monoclinic

crystal structures, respectively, are converted based on hexagonal symmetry since 18R and 24R

LPSO supercells used in this study are based on Niggli reduced cell from hexagonal symmetry[43],

[44]. The 10H LPSO supercell used in this study shows lower formation energy than other 10H

LPSO supercells[18]. However, the complete elastic stiffness matrixes of 10H, 18R and 24R LPSO

phases are listed in Appendix A.

Since no existing elastic constants are available for the Mg-Gd-Al LPSO phases, first, we

compare the present elastic constants of HCP Mg from first-principles calculations with

experiments and other calculations[45]–[48]. The calculated elastic moduli of HCP Mg are in the

range of experiments or have small differences, less than 1.6% except for C12 which is 5.5%

different from experiments[46], also, bulk and Young’s moduli are in the range of

experiments[46]–[48]. Second, the elastic stiffness matrix of the Mg-Y-Zn 18R LPSO phases are

calculated and compared with experimental results[26], [47], [49] and other theoretical

calculations[50]. Experimental elastic properties include nanoindentation measurements using

resonant ultrasound spectroscopy combined with electromagnetic acoustic resonance (65.0±1.4

GPa along the [0001] and 54.0±0.6 GPa along the [1120] direction for Young’s modulus)[26], and

microindentation (66.7±4.9 GPa for Young’s modulus)[49]. Our calculations are in good

agreement with these experiments with ~3 % error, especially for the Mg-Y-Zn 18R LPSO phase

where the experimental data were collected at 5.5 K[26].

In order to investigate the prevailing lattice distortion induced by solute atoms in L12

cluster, Figure 2.2 plots the bulk moduli from elastic calculations (VRH) and EOS as a function of

13

the number of layers in the SB which are also reported in Table 2.3. For a reliable interpretation of

the bulk moduli results, we reported bulk moduli from both methods to see the trends. Both results

have similar trends, except the 10H LPSO. The discrepancy is due to the elastic calculation method

which uses smaller deformation ranges than that of EOS fitting and calculated from a fixed volume.

As shown in Figure 2.2a, with increasing the number of layers in the SB, bulk moduli from EOS

fitting and from VRH of the LPSO decrease slightly. In addition, for the same interstitial atom in

various LPSO phases, this trend is even more clearly shown. The bulk modulus increases from 40.4

GPa (24R) to 42.1 GPa (10H) for LPSO phase with Al-int. with decreasing the number of layers in

the SB. The previous studies indicated that bulk moduli are inversely correlated to equilibrium

volumes of pure elements[42] ( 𝐵 = 20422𝑉−1.868) and also in dilute Ni- and Mg-based

alloys[51], [52]. Since the cluster density, defined as the number of clusters in a unit volume (𝜌𝑉 =

𝑁𝑐𝑙/𝑉 = 2(𝑜𝑟 4 𝑓𝑜𝑟 10𝐻)/𝑠𝑢𝑝𝑒𝑟𝑐𝑒𝑙𝑙 𝑣𝑜𝑙𝑢𝑚𝑒), this trend can be rephrased as the denser the L12

cluster density is, the larger the bulk moduli will be.

Table 2.3 Calculated elastic properties of LPSO structures of the Mg-Gd-Al alloys at 0 K,

including elastic stiffness constants (Cij's), Young's modulus (E), bulk modulus (B) from

both VRH approach and EOS fitting, and shear modulus (G) from the VRH approach. The

unit for each elastic property is GPa.

Type System Int. C11 C33 C12 C13 C44 C66 BVRH BEOS G E 𝜈 Ref.

HCP

Mg

61.3 66.2 27.6 21.5 18.7 16.6 36.7 36.5 18.5 47.6 0.283 TW 59.3 61.4 25.9 21.6 16.3 - - - - [45]

59.5 61.6 25.9 21.8 16.4 - 35.6 17.3 44.6 [48]

63.5 66.5 25.9 21.7 18.4 18.7 36.9 19.4 49.5 [46]

- - - - - - - - 48

±4 [47]

10H

Mg-Gd-

Al no 75.6 87.4 27.7 17.5 24.3 22.4 40.8 37.8 25.4 63.1 0.239 TW

Mg-Gd-

Al Al 80.1 91.6 28.9 17.6 25.2 26.0 42.1 41.9 27.5 67.8 0.230 TW

Mg-Gd-

Al

M

g 78.9 90.4 28.8 17.5 23.8 24.6 41.9 41.5 26.3 65.2 0.237 TW

Mg-Gd-

Al Gd 72.0 90.3 29.5 19.1 23.9 19.7 41.7 41.2 22.3 56.7 0.268 TW

18R Mg-Gd-

Al no 73.3 84.3 27.7 15.3 25.2 22.0 38.8 37.7 25.5 62.8 0.226 TW

14

Mg-Gd-

Al Al 78.8 89.0 26.7 18.1 26.5 25.1 41.6 40.9 27.5 67.6 0.227 TW

Mg-Gd-

Al

M

g 77.6 88.9 27.2 17.8 26.9 23.8 41.4 40.6 27.1 66.8 0.228 TW

Mg-Gd-

Al Gd 75.8 87.9 28.4 18.0 24.4 22.0 41.3 40.4 25.3 62.9 0.242 TW

Mg-Y-

Zn no 70.4 85.3 30.1 19.4 22.9 20.0 40.5 23.2 58.5 0.256 TW

Mg-Y-

Zn Zn 69.8 84.6 32.4 19.5 21.8 19.4 40.6 22.5 56.9 0.263 TW

Mg-Y-

Zn

M

g 70.6 85.4 32.3 19.1 22.9 20.2 40.6 23.4 58.9 0.256 TW

Mg-Y-

Zn Y 70.7 84.3 30.4 19.7 22.9 18.4 41.0 22.5 57.1 0.263 TW

Mg-Y-

Zn NA

72.5

±0.

7

80.0

±1.

8

-

18.9

±1.

1

23.5

±0.

3

21.2

±0.

3

- -

73.0

±1.

9

58.4

±0.

3

[26]

Mg-Y-

Zn NA - - - - - - - -

66.7

±4.

9

[49]

Mg-Y-

Zn NA

67.7

±1.

0

72.9

±2.

0

28.3

±1.

1

19.5

±0.

8

21.5

±0.

3

19.7

±0.

3

38.0

±0.

7

65.0

±1.

4

54.0

±0.

6

[53]

Mg-Y-

Zn NA

68.1

±1.

0

67.2

±0.

9

21.6

±0.

7

24.0

±0.

8

20.6

±0.

2

23.2

±0.

2

-

21.8

±0.

1

54.9

±0.

4

[53]

Mg-Y-

Zn no 71.6 82.0 28.7 19.7 23.2 - [26] +

Mg-Y-

Zn

M

g 79.5 87.8 23.1 16.7 25 - 40 28.1 68.4 [50] +

Mg-Y-

Zn

M

g 77 82.3 18.2 15.8 26.6 - 37.3 28.9 69 [50] *

14H

Mg-Gd-

Al no 71.1 83.8 27.3 16.4 26.4 22.5 38.6 37.6 25.9 63.5 0.222 TW

Mg-Gd-

Al Al 75.1 87.4 28.2 17.9 26.0 23.5 40.6 40.2 26.4 65.2 0.230 TW

Mg-Gd-

Al

M

g 72.5 86.1 29.7 18.2 25.7 21.0 40.5 40.0 25.0 62.3 0.240 TW

Mg-Gd-

Al Gd 72.9 85.9 29.3 18.2 24.2 21.9 40.3 39.9 24.8 61.8 0.242 TW

24R

Mg-Gd-

Al no 73.1 82.9 24.5 15.7 24.6 21.3 38.5 37.5 24.9 61.5 0.229 TW

Mg-Gd-

Al Al 77.4 85.1 25.5 17.6 25.2 25.0 40.4 39.8 26.7 65.6 0.227 TW

Mg-Gd-

Al

M

g 75.4 85.7 27.4 17.1 26.5 24.0 40.0 39.7 26.8 65.8 0.224 TW

Mg-Gd-

Al Gd 72.9 86.6 28.8 17.1 28.1 20.4 40.2 39.6 25.9 63.9 0.230 TW

+ VASP and * SIESTA calculations, TW-this work, NA-did not mentioned

15

Figure 2.2 Calculated bulk moduli of the LPSO phases with respect to number of layers in

structural block; (a) bulk modulus from EOS fitting and (b) bulk modulus from VRH

approach. Red dash lines indicate the bulk and shear moduli of HCP Mg.

The introduction of an interstitial atom in the LPSO increases the bulk moduli. For

example, the bulk modulus (VRH) of 18R LPSO is 38.8 GPa while that of 18R LPSO with Al-int.

is 41.6 GPa. This could be explained by the change of bonding environment. Particularly, the

introduction of an interstitial atom in the L12 cluster creates new bonding within the L12 cluster.

16

This can be confirmed by the change of density of the L12 cluster due to the lattice relaxation and

the energy contribution by the interstitial atom in L12 cluster. The density of the cluster with an

interstitial atom (e.g., Gd8Al7, Gd9Al6 or Gd8Al6Mg) is higher than that of cluster without interstitial

atom (e.g. Gd8Al6) due to the atomic volume reduction by inserting an interstitial atom. Especially

the case of Al interstitial LPSO, by inserting an Al-int. into the cluster, the cluster width changes

from 7.16 Å to 7.12 Å and the cluster height changes from 7.26 Å to 7.16 Å for 18R and also, the

intracluster 2NN RE-RE distance (𝑑𝑖𝑛𝑡𝑟𝑎𝑐𝑙𝑢𝑠𝑡𝑒𝑟) also reduces from 4.12 Å to 4.10 Å (Table 2.2).

This results in a smaller equilibrium volume per atom and higher bulk moduli for the LPSO with

an interstitial atom. Thus, the slope of the bulk moduli of the interstitial LPSO is affected by the

equilibrium volume per atom which is originated from the local density of the L12 cluster.

The effects of the stacking sequence of the LPSO phases on the elastic properties were

examined in terms of the formation energy per unit volume. Figure 2.3 shows the bulk moduli

comparison between elastic calculations with VRH method and from EOS fitting, and Young’s

moduli along [0001] direction (E[0001]) in terms of formation energy per unit volume. In Figure

2.3a, both bulk moduli have the similar trends that they increased almost linearly with decreasing

formation energy per unit volume, although LPSO structures without an interstitial atom have a

different slope from those with an interstitial atom. The slope difference originates from the

changes of the bonding nature around L12 clusters. The bulk modulus discrepancy between elastic

calculations with VRH and EOS fitting is from the types of applied pressure, although the bulk

moduli from both methods should be the same in principle. For example, the bulk modulus from

EOS fitting is calculated from the second derivative of energy over isotropic volume changes, while

that from VRH is an average value from energy with applied anisotropic pressure. Since we have

used 9 volumes for EOS fitting, the bulk modulus from EOS fitting may be more accurate.

Furthermore, E[0001] of LPSO structures increased almost linearly with decreasing formation energy

per unit volume, shown in Figure 2.3b. It is observed that the formation energy per unit volume

17

decreases with the increasing number of layers in SB, i.e. the addition of Mg layers between clusters

along [0001] direction, resulting in lower cluster density along [0001] direction and smaller E[0001].

Therefore, the Young’s modulus along [0001] direction, E[0001], related to the formation energy of

LPSO due to the atomic bonding changes between the stacking layers, especially the cluster density

changes along [0001] direction.

Figure 2.3 (a) comparison of bulk moduli both from VRH and EOS fitting as a function of

formation energies of LPSOs, and (b) Young’s modulus along [0001] direction trend as a

function of volumetric formation energies (𝑬𝒇/𝑽) of LPSOs.

Based on the present calculations shown in Figure 2.4, elastic constants such as C11, C33,

C44, and C66 of the LPSOs are larger than those of HCP Mg due to the introduction of solute atoms

and the formation of L12 cluster. The elastic stiffness component of C33 shows a linear trend with

the number of layers in the structural block (SB) as depicted in Figure 2.4b. The cluster density

decreases with the increasing number of layers in SB. This is related to the bonds in the [0001]

direction, which lead to the decrease of C33 shown in Figure 2.4b. For example, C33 of LPSO

without an interstitial atom decreased from 87.4 to 82.9 GPa. Furthermore, C11 (Figure 2.4a) is

mainly related to the atomic bonds within the basal plane. As Kimizuka et al.[54] described, 2NN

RE-RE intercluster and intracluster bonding distances are related to cluster interactions, which

indicates that the C11 may be determined by the competition between intercluster and intracluster

18

bonding distances. We analyzed the intercluster and intracluster bonding distances as well as the

cluster heights and widths. Since the clusters are closed packed in {0001} planes, the bonding

distances mainly affect the elastic property along {0001} plane, C11. The intracluster bonding

distances and cluster widths increase with the increasing number of layers in SB. This tendency

then changes when the number of layers in SB approaches 7 with the intracluster bonding distances

and cluster widths starting to decrease. This trend is very similar to that of C11.

Figure 2.4 Changes in (a) C11, (b) C33, (c) C44 and (d) C66 elastic constants as a function of the

number of layers in structural blocks.

Figure 2.5 shows the linear relationship between C11 of LPSOs and the energy contribution

resulting from the insertion of an interstitial atom (∆𝐸𝑖𝑛𝑡𝑖 ). The more negative ∆𝐸𝑖𝑛𝑡

𝑖 is, the higher

19

C11 becomes. This trend was valid for LPSOs with the interstitial atom, except the 10H LPSO with

the interstitial Gd atom, probably due to the high out-of-plane interaction between L12 clusters.

This linear relationship between the ∆𝐸𝑖𝑛𝑡𝑖 and C11 mainly originates from the changes of the

bonding distances around the L12 cluster (listed in Table 2.2) because the ∆𝐸𝑖𝑛𝑡𝑖 stems from the

bonding environment change around the L12 cluster due to the interstitial atom. For example, the

∆𝐸𝑖𝑛𝑡𝐴𝑙 increased from -2.09 meV/int (10H) to -2.00 meV/int (14H), as the intracluster distance

increased from 4.08Å (10H) to 4.11Å (14H). This trend of ∆𝐸𝑖𝑛𝑡𝑖 as a function of the number of

layers in the SB is very similar to that of C11.

Figure 2.5 Comparison between C11 and the energy contribution of interstitial atom in L12

cluster.

Moreover, it is found that C66 shows a similar trend with respect to that of the C11. C66 is

determined by the shear force along [1010] or [2110] direction while C11 is determined by the

tensile or compressive force along [10 1 0] or [21 1 0] direction. Both the elastic stiffness

components, C11 and C66, should be related to all the bonds along that direction. Since elastic

20

properties are related to cluster interaction or cluster density, C11 and C66 are related to the

intercluster and intracluster bonding distances. Figure 2.6 shows that the relationship between the

elastic stiffness components, the C11 and C66, and the L12 cluster width. This represents that the

smaller cluster width indicates the larger C11 and C66.

Figure 2.6 Relationship of L12 cluster width with (a) C11, and (b) C66 elastic constants.

Figure 2.7 shows the first principles calculated orientation dependent Young’s modulus

and shear modulus with the angle from 0 to 90 degrees from the cij components. Directional

Young’s and shear modulus are calculated from elastic compliance matrix (Sij) of hexagonal

system[55] and orientation dependent Young’s and shear modulus are calculated from equations

by Tromans et al.[56]. The LPSO phases contain clusters of L12, L12(Al), L12(Mg), L12(Gd) as a

function of the number of layers in the SB. Figure 2.7a shows the Young’s modulus of the LPSO

phases between [0001] and [1120]. As C11 and C33 discussed in the previous paragraph, it is clearly

shown that the Young’s moduli of the LPSO phases are more orientation dependent than that of

HCP Mg. Also, both [0001] and [1120], have the same trend as that of C33 and C11, respectively.

However, the shear moduli of LPSOs are not quite orientation dependent compared to that of HCP

Mg as shown in Figure 2.7b. Interestingly, the smaller L12 cluster size LPSOs such as L12(Al), in

terms of wcluster and hcluster, have different trend from other LPSOs and HCP Mg. This indicates that

21

C44 for L12(Al) clustered LPSOs are smaller than or similar to that of C66. This is due to the large

shrinkage of the cluster, and the RE-RE intercluster bonding distances are much larger than that of

intracluster bonding distances, which results in the localized cluster in Mg matrix. The localized

cluster does not seem to interfere shear force.

Figure 2.7 Crystallographic orientation dependence of the Young’s and Shear modulus of

10H, 18R, 14H and 24R LPSO phase at 0K, between [0001] and <11��0> 𝜽 is the angle from

<11��0>. The orientation dependencies of the Young’s modulus and shear modulus of HCP

Mg are shown for comparison.

2.3.3 Electronic Properties of the LPSOs

Based on DFT theory, the charge density can provide the information of the bonding

strength and the anisotropy of the bonding (elasticity)[57]. To study the bonding strength and the

anisotropy of bonding (elasticity), differential charge densities[57]–[60] are computed as follows

Equation 2.8 ∆𝝆 = 𝝆𝒊𝒏𝒕𝒆𝒓 − 𝝆𝒏𝒐𝒏−𝒊𝒏𝒕𝒆𝒓

where 𝜌𝑖𝑛𝑡𝑒𝑟 is the charge density after electronic relaxations, and 𝜌𝑛𝑜𝑛−𝑖𝑛𝑡𝑒𝑟 the reference

(or non-interacting) charge density calculated from one electronic step. The contour values of ∆𝜌

22

are in in 𝐞/Å𝟑. This is applied to HCP Mg, pure Mg 14H LPSO (LPSO structure with Mg atoms

only), and the LPSO phases with L12 clusters.

Figure 2.8 plots the isosurface of ∆𝜌 = 0021 𝐞/Å𝟑. It can be seen that in the HCP stacking

blocks, the isosurface shape is an prism (rectangular in 2 Dimension), and in the FCC stacking

region, the isosurface shape is tetragonal (triangle) [60]. The isosurface shape in pure Mg 14H

LPSO (Figure 2.8f) changes to tetragonal. It is worth noting that there are no connections between

the {0001} planes with ∆𝜌 in HCP Mg and pure Mg 14H LPSO[57]. However, the formation of

L12 cluster by solute atoms not only increases the charge density at the FCC stacking faults region,

but also connects the charge density in the HCP stacking blocks (between L12 clusters along [0001]

direction) (red arrows in Figure 2.8a, b, c, d). Such connections between the {0001] planes are

likely to result from the solute atom rich stacking faults regions. Since the denser charge density

imply the stronger bonding between atoms[57] and also Young’s modulus is proportional to

∆𝜌[60], the origin of the enhanced Young’s modulus of LPSOs comes from not only the formation

of the L12 cluster but also the connections of the {0001} planes in the HCP stacking layers.

23

Figure 2.8 Differential charge density plots of the LPSOs with or without interstitial atoms.

Differential charge density plots of (a) 10H, (b) 18R, (c) 14H, and (d) 24R LPSO. The

reference states used in this study are (e) 2H and (f) 14H LPSO with Mg only; (g), (h), and (i)

are the 14H LPSO with Al-int., Gd-int., and Mg-int. Red arrows indicate the charge density

connections between the {0001] planes. Isosurfaces are 0.0021 (e/Å 3) and the Mg atom sizes

are exaggerated for better visualization.

24

It is interesting to know how the contributions of elastic properties of the HCP layers in a

LPSO phase are changing. According to Miedema et al.[61] and Wu et al.[62], for pure alkali

metals and non-transition metals, √𝐵/𝑉𝑚 is linearly proportional to charge density at the boundary

of the Wigner-Seitz cell (𝑛𝑊𝑆), where B is the bulk modulus and 𝑉𝑚 the molar volume of the

element. This relation can be applied to the HCP stacking layers in the LPSO phases in order to

explain the partial charge to these stacking regions. Since the HCP stacking layers in the LPSO

phases are compressed along (0001) direction compared to HCP Mg structure (c.f. in Figure 2.8)

and the nearest neighbor interatomic distance of those region (3.162 Å ) is smaller than that of HCP

Mg structure (3.178 Å ), the volume of those regions is smaller than the HCP Mg structure.

Moreover, the bulk modulus of those regions should be larger than that of the HCP Mg structure

because the correlation between the bulk modulus and the nearest neighbor interatomic distance,

bulk modulus decreases owing to lattice expansion, described by Ganeshan et al.[51]. Thus, the

√𝐵/𝑉𝑚 values of the HCP stacking region in the LPSO phases are larger than the corresponding

values of HCP Mg. This means that the charge densities of the HCP stacking regions in the LPSO

are larger than that of the HCP Mg.

25

Chapter 3

First-Principles Calculations and Thermodynamic Modelling of Long

Periodic Stacking Ordered (LPSO) Phases in Mg-Al-Gd

3.1 Introduction

Among a variety of Mg alloys, the Mg-Al based alloys, such as AZ-91D and AM-50A,

have been widely used because of their excellent mechanical strength, corrosion resistance, and die

castability[63]. To further increase their strength and usage at higher temperatures, such as in

automotive powertrains above 125 oC[63], Mg alloys containing long periodic stacking ordered

(LPSO) phases[64] have received considerable attention due to their improved creep resistance and

strength[10], [12], [13], [65]. For example, it has been reported that the Mg97Zn1Y2 alloys with

various LPSO phases show outstanding creep resistance[64], [66] as well as excellent tensile yield

strength above 600 MPa and an elongation of 5% at room temperature[10], [67].

The LPSO phases observed in Mg-TM (Transition Metal)-RE (Rare Earth) ternary systems

[6], [8], [17], [19] consist of periodic ordered FCC stacking layers in the structural block (SB);

[17]–[19] see Figure 2.1. Among various LPSOs, 14H and 18R are frequently observed[6], [8],

[17]–[19]. By analyzing the crystal structures of LPSOs in the Mg-Al-Gd system, Egusa et al.[23],

Kishida et al.[6], [17], [18], and Yokobayashi et al.[8], [19] discovered the L12 type clusters in SBs

of the 14H and 18R LPSO phases in terms of the in-plane ordering of Gd and Al atoms in FCC

stacking layer regions, and Kimizuka et al.[25], [68] and Kishida et al.[17], [69] reported the

periodic formation of Gd8Al6 with a L12 type atomic arrangement is long range order.

26

Furthermore, Kishida et al.[18], [69] suggested that there might exist interstitial atoms at

the center of the L12 clusters in both Mg-Zn-Y and Mg-Al-RE systems based on scanning

transmission electron microscopy (STEM) observations. First-principle calculations based on

density functional theory (DFT) by Saal et al.[14] and Kishida et al.[18], [69] support this

suggestion. Thermodynamic stability of many LPSO phases at 0 K was investigated by Saal et

al.[14] using DFT-based first-principles calculations with the interstitial atoms at the center of the

L12 clusters considered. However, their thermodynamic stability at finite temperatures has not been

studied except in the Mg-Y-Zn[70], [71] and Mg-Gd-Zn[72] systems, where the 14H and 18R

LPSO phases were treated as stoichiometric compounds of Mg12ZnRE and Mg10ZnRE, respectively.

In the present work, the thermodynamic properties of the 10H, 14H, 18R and 24R LPSO

phases in the Mg-Al-Gd ternary system are modeled by means of the CALPHAD (calculation of

phase diagram) method[73]. The L12 clusters and the existence of interstitial atoms within their

center are considered in terms of the compound energy formalism (CEF)[74]. DFT-based first-

principles calculations are performed to provide thermodynamic properties at finite-temperatures

[75] for CALPHAD modeling.

3.2 First-Principles Calculations

The space groups of 10H and 14H LPSO phases used in the present work are Cmce and

P63/mcm, respectively, while the space group of 18R and 24R LPSO phases are C2/m [6], [18].

Crystal structures of 10H, 18R, 14H, and 24R LPSO phases are shown in Figure 2.1a-d,

respectively, and the L12 cluster with an interstitial atom is depicted in Figure 2.1f. The lattice of

the LPSO phases can be divided into four sublattices based on their Wyckoff positions as follows

[17]–[19]

Equation 3.1 (𝑴𝒈)𝟔𝟖+𝟐𝟒𝒙(𝑴𝒈,𝑮𝒅, 𝑨𝒍)𝟏𝟔(𝑴𝒈,𝑮𝒅,𝑨𝒍)𝟏𝟐(𝑴𝒈,𝑮𝒅, 𝑨𝒍, 𝑽𝒂)𝟐

27

where x=1, 2, 3, and 4 for 10H, 18R, 14H, and 24R LPSO phases, respectively. The 1st sublattice

(sublattice Ⅰ) represents the layers outside of the L12 cluster which consists of mostly Mg atoms

located in the HCP and FCC lattices, the 2nd sublattice (sublattice Ⅱ) the corner positions of the L12

cluster, the 3rd sublattice (sublattice Ⅲ) the face-centered positions of the L12 cluster, and the 4th

sublattice (sublattice Ⅳ) the interstitial octahedral site of the L12 cluster. The supercells of the

10H, 14H, 18R and 24R LPSO phases contain 244 or 240, 170 or 168, 146 or 144, and 194 or 192

atoms with or without the interstitial atoms in the L12 cluster, respectively[18], [23]. As it can be

seen in the sublattice model above, the mixture of all elements is considered in the sublattices Ⅱ,

Ⅲ and Ⅳ.

Each endmember in CEF is defined by the sublattice model shown in Eq. 1 with only one

element in each sublattice. For mixing in sublattices Ⅱ and Ⅲ, the dilute solutions are considered

by substituting one atom in the sublattice of the supercells. The enthalpies of mixing in dilute

solutions in the sublattice Ⅳ with a supercell twice the size of the LPSO supercells are calculated

to be very small about 10 J/mole-atom. This is due to the large distance between interstitial atoms,

about six-times than that between Mg atoms, indicating ideal mixing in the sublattice.

First-principles calculations are performed using the Vienna Ab-initio Simulation Package

(VASP)[27], [28], [31]. Electron-ion interactions are described by the projector augmented-wave

(PAW) method[29], and the exchange-correlation energy functional is depicted by the generalized

gradient approximation (GGA) as implemented by Perdew, Burke, and Ernzerhof (PBE)[30]. The

f-electrons of Gd are treated as core electrons (so-called “frozen” potential), an approximation that

has shown to produce accurate thermodynamic properties for lanthanide compounds[32]–[35].

Plane wave cutoff energy of 350 eV is used for all calculations, which is at least 1.3 times higher

than the recommended values by VASP[31]. For HCP Mg with 2 atoms in the unit cell, the 29 ×

29 × 16 k-point grids are employed. The k-point meshes of 3 × 5 × 2 (10H LPSO), 6 × 6 × 2 (14H

LPSO), 6 × 3 × 3 (18R LPSO), and 3 × 2 × 2 (24R LPSO) are used for the LPSO phases. These k-

28

point meshes guarantee errors below 0.1 meV/atom (0.2 meV/atom for 24R LPSO). These

structures are fully relaxed by the Methfessel-Paxton method.[76] After relaxations, a final static

calculation using the tetrahedral method with Blöch corrections[36] is applied to predict accurate

total energy.

The Helmholtz energy, 𝐹(𝑉, 𝑇), of the present structures of interest is evaluated in terms

of the quasi-harmonic approach as a function of volume (V) and temperature (T) [75], [77]:

Equation 3.2 𝑭(𝑽, 𝑻) = 𝑬𝟎𝑲(𝑽) + 𝑭𝒗𝒊𝒃(𝑽, 𝑻) + 𝑭𝒆𝒍(𝑽, 𝑻)

where E0K(V) is the static contribution at 0 K without the zero-point vibrational energy. Fvib(V, T)

and Fel(V, T) represent the vibrational and thermal-electronic contributions to the Helmholtz

energy, respectively. To estimate E0K(V), a four-parameter Birch–Murnaghan (BM4) equation of

state (EOS),[75], [78] is used to fit the energy versus volume (E-V) data points from first-principles

calculations,

Equation 3.3 𝑬(𝑽) = 𝑨𝟏 + 𝑨𝟐𝑽−𝟐

𝟑 + 𝑨𝟑𝑽−𝟒

𝟑 +𝑨𝟒𝑽−𝟐

where A1, A2, A3 and A4 are fitting parameters. The E-V data points used in the EOS fitting are

relaxed with respect to ionic positions and cell shape at the given volumes, and nine E-V data points

are usually used. Note that the Helmholtz energy is equated to the Gibbs energy due to the zero (or

ambient) external pressure used in the present work. The thermal electronic contribution to the

Helmholtz energy is estimated based on the electronic density of states (DOS) in terms of the

Fermi–Dirac statistics for metallic systems[75].

The quasi-harmonic vibrational contributions can be obtained through phonon or the Debye

model. In calculations of phonon, the vibrational contribution to the Helmholtz energy can be

expressed as

Equation 3.4 𝑭𝒗𝒊𝒃(𝑽, 𝑻) = 𝒌𝑩𝑻∫ 𝐥𝐧 [𝟐 𝐬𝐢𝐧𝐡ℏ𝝎

𝟐𝒌𝑩𝑻]

𝟎𝒈(𝝎,𝑽)𝒅𝝎

29

where ℏ is the reduced Planck constant, ω the phonon frequency, and g(ω, V) the phonon DOS as

a function of frequency ω and volume V. In the present work, phonon calculations are carried out

for Mg, Al, and Gd, and the pure Mg LPSO phases by the supercell approach as implemented in

the YPHON code[79]. The primitive cells of pure Mg LPSO contains 10 atoms for 10H and 18

atoms for 18R. The 3 × 3 × 1 supercells are used in calculations of phonon with their force

constants calculated by VASP in terms of k-point mesh of 5 × 5 × 1 and the finite displacement

method (the step size is 0.015 Å ).

Vibrational contribution to the Helmholtz energy via Debye model is as follows,[75]

Equation 3.5 𝑭𝒗𝒊𝒃(𝑽, 𝑻) =𝟗

𝟖𝒌𝑩𝚯𝑫(𝑽) − 𝒌𝑩𝑻 {𝑫(

𝚯𝑫(𝑽)

𝑻) + 𝟑𝒍𝒏(𝟏 − 𝒆−

𝚯𝑫(𝑽)

𝑻 )}

where kB is the Boltzmann constant, T the temperature, and D the Debye function, ΘD the Debye

temperature given by

Equation 3.6 𝚯𝑫 = 𝒔𝑨𝑽𝟎𝟏/𝟔(𝑩𝟎

𝑴)𝟏/𝟐(𝑽𝟎

𝑽)𝜸

where A is a constant equal to (6π2)1/3ℏ/kB, s a scaling factor to adjust Debye temperature, ℏ the

reduced Planck constant, V0 the equilibrium volume, B0 the bulk modulus, M the atomic mass, and

γ the Debye-Gruneisen parameter. The scaling factor s of each LPSO phase is calculated based on

the following equation[80],

Equation 3.7 𝒔(𝝊) = 𝟑𝟓/𝟔 [𝟒√𝟐 (𝟏+𝝊

𝟏−𝟐𝝊)𝟑/𝟐+ (

𝟏+𝝊

𝟏−𝝊)𝟑/𝟐]−𝟏/𝟑

where 𝜐 is the Poisson’s ratio to be predicted from elastic stiffness constants[42] and the Voigt-

Reuss-Hill (VRH) approximation[40], [41]. The details of elastic properties of the LPSO

endmembers from first-principles calculations are based on Kim et al.[81].

30

3.3 CALPHAD Modeling of Phase Equilibria

The Gibbs energies of pure Mg, Gd, and Al are taken from the Scientific Group

Thermodata Europe (SGTE) pure element database [82]. For the three binary systems, the Mg-Al

system was modelled by Liang et al.[83] and Zhong et al. [84]. In the present work, the modeling

work by Zhong et al. is used since it incorporates the energetics of compounds from first-principles

calculations and latest experiments by Czeppe et al.,[85] which changes the upper temperature limit

of ε-Al30Mg23. The Mg-Gd system was modelled by Cacciamani et al.[86] and Guo et al.[87]. The

modeling work by Guo et al. is used in the present work since the cooling and heating differential

thermal analysis (DTA) results from Manfrinetti et al.[88] were included in their modeling,

resulting in a better description of phase boundary between Gd and B2-GdMg and a finite solubility

in B2-GdMg. The Al-Gd system modelled by Cacciamani et al.[86] is adopted in the present work

since their model matches well with the experimental data by Gschneidner et al.[89] and Saccone

et al.[90].

The sublattice model of the LPSO phases is expressed in Equation 3.1. The LPSO

compositions in the Mg-RE-TM (TM=Al, Zn, Cu, Ni and RE=Gd, Y, Er) systems observed by

energy dispersive X-ray spectroscopy (EDS)[8], [19], [23], [65], [91]–[101] in the literature are

shown in Figure 3.1. It can be seen that the composition range of the present sublattice model could

cover the observed LPSO composition data.

31

Figure 3.1 Experimentally observed LPSO phase compositions of 14H (a) and 18R (b)

LPSO phases[8], [19], [23], [65], [91]–[101]. Dash lines refer to the composition ranges of this

model.

The Gibbs energy for the four-sublattice model of the LPSO phases is given by

Equation 3.8 𝑮𝒎 = ∑ ∑ ∑ 𝒚𝑴𝒈Ⅰ 𝒚𝒋

Ⅱ𝒚𝒌Ⅲ𝒚𝒍

Ⅳ𝑮𝑴𝒈:𝒋:𝒌:𝒍𝒍𝒌𝒋 − 𝑻𝑺𝒎 + 𝑮𝒎𝒙𝒔

where 𝐺𝑀𝑔:𝑗:𝑘:𝑙 denotes the Gibbs energy of endmembers with the species j, k, and l occupying the

2nd, 3rd, and 4th sublattices, respectively, and 𝑆𝑚 and 𝐺𝑚𝑥𝑠 are the ideal entropy and the excess

Gibbs energy of mixing. The Gibbs energy of endmembers of the LPSO phases (𝐺𝑀𝑔𝑝𝐺𝑑𝑞𝐴𝑙𝑟) is

described as follows:

Equation 3.9 𝑮𝑴𝒈𝒑𝑮𝒅𝒒𝑨𝒍𝒓 − ∑ 𝒏𝒊𝑯𝒊𝑺𝑬𝑹

𝒊 = 𝒂 + 𝒃𝑻 + 𝒄𝑻 𝐥𝐧(𝑻) + 𝒅𝑻𝟐 + 𝒆𝑻−𝟏 + 𝒇𝑻𝟑

where a, b, c, d, e and f are the model parameters determined from thermodynamic properties at

finite temperatures obtained from the DFT-based first-principles calculations, see Equation 3.2;

𝑛𝑖 is the mole of species 𝑖; 𝐻𝑖𝑆𝐸𝑅 refers to the SGTE enthalpies of species 𝑖 at 298.15 K, 1 bar, and

its stable structure, referred as the stable element reference (SER),[82] such as HCP Mg, HCP Gd,

and FCC Al. For DFT calculations, the same SER reference states are used to estimate the Gibbs

energy in Equation 3.9. The a-f model parameters of all the endmembers are listed in Table 3.1.

32

Table 3.1 Gibbs energies of LPSO end-members obtained from the Debye model, defined as

𝑮𝑳𝑷𝑺𝑶 −𝑯𝑺𝑬𝑹 = 𝒂 + 𝒃𝑻 + 𝒄𝑻 𝐥𝐧(𝑻) + 𝒅𝑻𝟐 + 𝒆𝑻−𝟏 + 𝒇𝑻𝟑(J/mole-atom), where HSER is

reference state from SGTE[82].

End-Members a b c d e f

10H

(Mg)92(Gd)16(Al)12(Al)2 -19299.44 139.14 -25.17 -1.96×10-3 7.54×104 -2.55×10-7

(Mg)92(Gd)16(Al)12(Gd)2 -19393.17 135.51 -25.05 -2.19×10-3 5.95×104 -2.07×10-7

(Mg)92(Gd)16(Al)12(Mg)2 -19346.74 138.40 -25.13 -2.00×10-3 7.19×104 -2.31×10-7

(Mg)92(Al)16(Al)12(Va)2 -6522.63 145.47 -25.51 -1.90×10-3 1.08×105 -5.04×10-7

(Mg)92(Gd)16(Al)12(Va)2 -16298.83 143.74 -25.22 -2.14×10-3 1.04×105 -3.26×10-7

(Mg)92(Mg)16(Al)12(Va)2 -7313.50 152.19 -25.53 -1.93×10-3 5.63×104 -4.11×10-7

(Mg)92(Al)16(Gd)12(Va)2 -9956.37 130.28 -25.26 -2.05×10-3 3.98×104 -2.74×10-7

(Mg)92(Gd)16(Gd)12(Va)2 -9063.78 128.65 -25.21 -2.45×10-3 3.57×104 -3.09×10-7

(Mg)92(Mg)16(Gd)12(Va)2 -7591.30 149.94 -25.18 -2.41×10-3 2.83×104 -2.51×10-7

(Mg)92(Al)16(Mg)12(Va)2 -7691.75 141.49 -25.23 -2.06×10-3 8.81×104 -3.45×10-7

(Mg)92(Gd)16(Mg)12(Va)2 -11146.29 132.49 -25.02 -2.39×10-3 4.77×104 -1.86×10-7

(Mg)92(Mg)16(Mg)12(Va)2 -8896.54 162.25 -29.30 5.10×10-3 1.26×105 -2.07×10-6

18R

(Mg)116(Gd)16(Al)12(Al)2 -17719.08 136.10 -25.16 -1.94×10-3 7.89×104 -2.82×10-7

(Mg)116(Gd)16(Al)12(Gd)2 -17863.77 135.43 -25.33 -1.89×10-3 7.39×104 -3.02×10-7

(Mg)116(Gd)16(Al)12(Mg)2 -17777.03 135.22 -25.04 -1.63×10-3 7.70×104 -1.68×10-7

(Mg)116(Al)16(Al)12(Va)2 -6777.82 144.40 -25.45 -1.95×10-3 1.14×105 -4.75×10-7

(Mg)116(Gd)16(Al)12(Va)2 -15128.37 135.86 -25.29 -1.97×10-3 7.60×104 -3.03×10-7

(Mg)116(Mg)16(Al)12(Va)2 -7411.63 142.00 -25.43 -1.95×10-3 9.94×104 -4.05×10-7

(Mg)116(Al)16(Gd)12(Va)2 -9850.91 135.48 -25.22 -2.03×10-3 7.23×104 -2.85×10-7

(Mg)116(Gd)16(Gd)12(Va)2 -8496.11 127.81 -25.05 -2.43×10-3 5.03×104 -2.06×10-7

(Mg)116(Mg)16(Gd)12(Va)2 -7987.33 133.92 -25.26 -2.18×10-3 6.58×104 -2.84×10-7

(Mg)116(Al)16(Mg)12(Va)2 -7705.89 142.30 -25.23 -2.07×10-3 1.07×105 -3.53×10-7

(Mg)116(Gd)16(Mg)12(Va)2 -10509.56 133.67 -25.19 -2.25×10-3 6.77×104 -2.10×10-7

(Mg)116(Mg)16(Mg)12(Va)2 -7928.70 141.33 -25.86 3.62×10-3 7.45×104 -7.77×10-7

14H

(Mg)140(Gd)16(Al)12(Al)2 -16389.11 136.51 -25.15 -1.96×10-3 8.05×104 -2.91×10-7

(Mg)140(Gd)16(Mg)12(Al)2 -10549.06 133.66 -25.11 -2.25×10-3 6.77×104 -2.38×10-7

(Mg)140(Gd)16(Al)12(Gd)2 -16550.29 135.17 -25.16 -1.98×10-3 7.42×104 -2.77×10-7

(Mg)140(Gd)16(Mg)12(Al)2 -9590.46 132.75 -25.14 -2.22×10-3 6.37×104 -2.35×10-7

(Mg)140(Gd)16(Al)12(Mg)2 -16470.41 136.02 -25.16 -1.94×10-3 7.76×104 -2.84×10-7

(Mg)140(Gd)16(Mg)12(Al)2 -10241.18 133.54 -25.14 -2.23×10-3 6.65×104 -2.38×10-7

(Mg)140(Al)16(Al)12(Va)2 -6879.13 143.78 -25.47 -1.94×10-3 1.09×105 -4.58×10-7

(Mg)140(Gd)16(Al)12(Va)2 -14406.22 136.43 -25.21 -2.06×10-3 7.96×104 -3.21×10-7

(Mg)140(Mg)16(Al)12(Va)2 -7514.17 139.61 -25.45 -1.97×10-3 8.44×104 -3.95×10-7

33

(Mg)140(Al)16(Gd)12(Va)2 -9623.40 134.45 -25.31 -1.98×10-3 6.58×104 -2.98×10-7

(Mg)140(Gd)16(Gd)12(Va)2 -8395.93 127.54 -25.01 -2.42×10-3 4.80×104 -2.18×10-7

(Mg)140(Mg)16(Gd)12(Va)2 -7979.11 131.84 -25.26 -2.16×10-3 5.63×104 -2.95×10-7

(Mg)140(Al)16(Mg)12(Va)2 -7785.07 141.21 -25.23 -2.09×10-3 9.91×104 -3.54×10-7

(Mg)140(Gd)16(Mg)12(Va)2 -10189.43 133.36 -25.18 -2.20×10-3 6.51×104 -2.44×10-7

(Mg)140(Mg)16(Mg)12(Va)2 -8781.04 157.77 -28.55 3.99×10-3 1.15×105 -1.81×10-6

24R

(Mg)164(Gd)16(Al)12(Al)2 -15402.29 140.66 -25.16 -2.01×10-3 8.40×104 -2.96×10-7

(Mg)164(Gd)16(Al)12(Gd)2 -15513.55 140.05 -25.15 -2.02×10-3 8.09×104 -2.53×10-7

(Mg)164(Gd)16(Al)12(Mg)2 -15603.34 143.75 -25.62 -1.22×10-3 9.37×104 -4.26×10-7

(Mg)164(Al)16(Al)12(Va)2 -7094.23 145.47 -25.61 -1.82×10-3 1.05×105 -4.82×10-7

(Mg)164(Gd)16(Al)12(Va)2 -13517.54 141.48 -25.42 -1.86×10-3 8.40×104 -3.80×10-7

(Mg)164(Mg)16(Al)12(Va)2 -7643.64 142.57 -25.32 -2.08×10-3 9.31×104 -3.81×10-7

(Mg)164(Al)16(Gd)12(Va)2 -9477.02 138.97 -25.26 -2.01×10-3 7.29×104 -3.11×10-7

(Mg)164(Gd)16(Gd)12(Va)2 -8379.11 134.18 -25.09 -2.33×10-3 5.40×104 -2.40×10-7

(Mg)164(Mg)16(Gd)12(Va)2 -8063.77 137.85 -25.23 -2.17×10-3 6.80×104 -3.02×10-7

(Mg)164(Al)16(Mg)12(Va)2 -7856.82 143.25 -25.29 -2.03×10-3 9.82×104 -3.64×10-7

(Mg)164(Gd)16(Mg)12(Va)2 -9979.53 138.29 -25.28 -2.08×10-3 6.94×104 -2.73×10-7

(Mg)164(Mg)16(Mg)12(Va)2 -9126.87 164.94 -29.73 5.69×10-3 1.33×105 -2.22×10-6

The ideal entropy and the excess Gibbs energy of mixing in per mole of formula are

represented by:

Equation 3.10 𝑺𝒎 = −𝑹{(𝟔𝟖 + 𝟐𝟒𝒙)𝒚𝑴𝒈

Ⅰ 𝒍𝒏(𝒚𝑴𝒈Ⅰ ) + 𝟏𝟔∑ 𝒚𝒋

Ⅱ 𝒍𝒏(𝒚𝒋Ⅱ)𝒋

+𝟏𝟐∑ 𝒚𝒌Ⅲ 𝒍𝒏(𝒚𝒌

Ⅲ)𝒌 + 𝟐∑ 𝒚𝒍Ⅳ 𝒍𝒏(𝒚𝒍

Ⅳ)𝒍

}

Equation 3.11

𝑮𝒎 𝒙𝒔 =∑∑∑∑𝒚𝑴𝒈

Ⅰ 𝒚𝒋Ⅱ𝒚𝒌

Ⅲ𝒚𝒍Ⅳ𝒚𝒎

Ⅱ𝑳𝑴𝒈:𝒋,𝒎:𝒌:𝒍𝒎>𝒋𝒍𝒌𝒋

+∑∑∑∑ 𝒚𝑴𝒈Ⅰ 𝒚𝒋

Ⅱ𝒚𝒌Ⅲ𝒚𝒍

Ⅳ𝒚𝒎Ⅲ𝑳𝑴𝒈:𝒋:𝒌,𝒎:𝒍

𝒎>𝒌𝒍𝒌𝒋

+∑∑∑∑𝒚𝑴𝒈Ⅰ 𝒚𝒋

Ⅱ𝒚𝒌Ⅲ𝒚𝒍

Ⅳ𝒚𝒎Ⅳ𝑳𝑴𝒈:𝒋:𝒌:𝒍,𝒎

𝒎>𝒍𝒍𝒌𝒋

+⋯

+∑∑∑∑∑𝒚𝑴𝒈Ⅰ 𝒚𝒋

Ⅱ𝒚𝒌Ⅲ𝒚𝒍

Ⅳ𝒚𝒎Ⅱ𝒚𝒏

Ⅲ𝑳𝑴𝒈:𝒋,𝒎:𝒌,𝒏:𝒍𝒏>𝒌𝒎>𝒍𝒍𝒌𝒋

+⋯

where 𝐿𝑀𝑔:𝑗,𝑚:𝑘:𝑙𝐿𝑃𝑆𝑂𝑣 is the vth interaction parameter between species j and m in the second sublattice,

and the same for other interaction parameters. They are evaluated from the enthalpy of formation

34

from DFT-based first-principles calculations. 𝐿𝑀𝑔:𝑗,𝑚:𝑘:𝑙𝐿𝑃𝑆𝑂𝑣 of all the endmembers are listed in

Table 3.2.

Table 3.2 Interaction parameters in individual sublattices (kJ/mol-atom).

LPSO 10H 18R 14H 24R

i,j Al-

Gd

Al-

Mg

Gd-

Mg

Al-

Gd

Al-

Mg

Gd-

Mg

Al-

Gd

Al-

Mg

Gd-

Mg

Al-

Gd

Al-

Mg

Gd-

Mg

𝐿Mg:i,j:Al:Va𝐿𝑃𝑆𝑂0 -1.139 -0.018 1.287 -1.127 0.217 -1.757 0.733 1.259 -2.037 5.767 -0.096 -5.577

𝐿Mg:i,j:Al:Va𝐿𝑃𝑆𝑂1 5.594 0.928 1.598 5.214 0.081 4.240 4.491 -0.423 3.473 -1.642 -6.829 -3.572

𝐿Mg:i,j:Gd:Va𝐿𝑃𝑆𝑂0 -0.409 0.471 1.435 -2.160 -0.585 0.961 0.162 0.248 -0.078 0.845 0.363 -0.617

𝐿Mg:i,j:Gd:Va𝐿𝑃𝑆𝑂1 2.162 0.102 0.679 3.073 -0.147 1.094 1.194 -0.272 0.337 -6.257 -7.507 -6.422

𝐿Mg:i,j:Mg:Va𝐿𝑃𝑆𝑂0 -0.574 -0.253 -0.842 -1.709 0.210 0.811 -0.992 -0.081 0.189 3.060 -0.825 -3.837

𝐿Mg:i,j:Mg:Va𝐿𝑃𝑆𝑂1 5.454 0.345 2.032 3.504 0.354 1.142 2.819 0.118 0.848 -1.948 -7.166 -3.960

𝐿Mg:Al:i,j:Va𝐿𝑃𝑆𝑂0 -0.202 -0.024 0.297 0.061 0.098 0.194 0.717 0.597 0.204 -3.863 -3.009 1.738

𝐿Mg:Al:i,j:Va𝐿𝑃𝑆𝑂1 -7.607 0.057 -3.640 -5.992 -0.126 -3.025 -5.938 -0.691 -2.680 -7.611 -2.409 -8.004

𝐿Mg:Gd:i,j:Va𝐿𝑃𝑆𝑂0 9.128 7.159 -1.364 5.889 4.652 -1.627 4.977 3.702 -1.393 1.436 1.102 -0.593

𝐿Mg:Gd:i,j:Va𝐿𝑃𝑆𝑂1 -5.591 -6.042 0.697 -3.321 -3.987 0.066 -3.032 -3.218 -0.037 -3.028 -2.808 -2.552

𝐿Mg:Mg:i,j:Va𝐿𝑃𝑆𝑂0 0.655 -0.119 -0.629 0.484 -0.043 0.021 0.557 -0.034 -0.135 0.187 -0.435 -0.312

𝐿Mg:Mg:i,j:Va𝐿𝑃𝑆𝑂1 -5.223 -0.389 -1.509 -4.607 -0.204 -1.210 -4.042 -0.392 -0.977 -6.875 -5.013 -5.545

3.4 Results and Discussion

To benchmark the reliability of first-principles calculations for the Mg-Al-Gd system,

lattice parameters of pure elements are calculated and compared to experimental and calculated

data available in the literature in Table 3.3 [102]–[104]. As can be seen in the table, the lattice

parameters of Mg, Al, and Gd are in good agreement with the corresponding experimental results.

The relative errors between calculated and experimental lattice parameters of these pure elements

are less than 1%. Figure 3.2 shows the phonon results of HCP Mg and pure Mg LPSO endmembers,

(Mg)68+24x(Mg)16(Mg)12(Va)2, and the phonon dispersion curves of HCP Mg (Figure 3.2a) are

calculated at its equilibrium volume, comparing favorably with experiments[105]. The phonon

DOS curves between HCP Mg and the Mg LPSO endmembers are plotted in Figure 3.2b. All the

35

phonon DOS curves have similar trends especially the slopes at low frequencies (< 3.5 THz),

indicating the similar thermodynamic properties between HCP Mg and pure Mg LPSO

endmembers since thermodynamic properties are dominated by phonon at low frequencies[106].

Table 3.3 Calculated lattice parameters of pure elements in comparison with computational

and experimental data in the literature.

Element Lattice Parameter

a(Å) Error (%) c(Å) Error (%) Reference Mg 3.195 -0.56 5.176 -0.71 This work

3.213 5.213 Expt.[102] 3.189 5.099 Calc.[107]

Gd 3.643 0.19 5.728 -0.01 This work 3.636 5.783 Expt.[103] 3.624 5.715 Calc.[42]

Al 4.040 -0.22 This work 4.049 Expt.[104] 4.046 Calc.[42]

𝐸𝑟𝑟𝑜𝑟 (%) = (𝐶𝑎𝑙𝑐. −𝐸𝑥𝑝𝑡. )/𝐸𝑥𝑝𝑡.× 100(%)

Figure 3.2. (a) Phonon dispersion curves of HCP Mg with experimental data[105] (red

circles), and b) phonon DOS curves of HCP Mg and Mg LPSO phases of 10H, 18R, 14H, and

24R.

The predicted heat capacities of HCP Mg, based on both the quasi-harmonic phonon

method and the quasi-harmonic Debye model are shown in Figure 3.3a. At low temperatures, the

predicted heat capacities, both by phonon and the Debye model (with a scaling factor of 0.789; see

36

Equation 3.7), are in good agreement with data from SGTE[82]. Therefore, scaling factors from

the predicted elastic constants[81] are used to calculate thermodynamic properties in terms of the

Debye model. The calculated scaling factors of all endmembers are listed in Table 3.4 with input

from Kim et al.[81]. In Figure 3.3b, the quasi-harmonic calculations of phonon show that the heat

capacities of pure Mg LPSO endmembers are slightly larger than that of HCP Mg (such as < 0.02

J/mole-atom at 300 K). Figure 3.3c shows that HCP Mg is more stable than the pure Mg LPSO

endmembers in the whole temperature range (up to melting temperature of 923 K for HCP Mg).

This is in contradiction to the conclusion by Iikubo et al.[108], in which it was concluded that 14H

and 18R Mg LPSO endmembers are stable over HCP Mg phase at high temperatures ( > 600 K and

> 400 K, respectively). Our conclusion is more reasonable because imaginary frequencies were not

observed in the phonon calculations which cause an error on the force constant, and the small

entropy contribution (slopes at low frequency region in Figure 3.2b) is not enough to overcome

the enthalpy contribution. Therefore, Mg LPSO endmembers including FCC stacking layers should

be less stable than HCP Mg without FCC stacking layers as shown in Figure 3.3c.

Table 3.4 Formation energies (Eform in kJ/mole-atom) at 0 K and the scaling factors (s) of the

endmembers, see Equation 3.12, with the elastic constants from Kim et al.[81].

End-Members 10H (x=1) 18R (x=2) 14H (x=3) 24R (x=4)

𝐸𝑓𝑜𝑟𝑚 s 𝐸𝑓𝑜𝑟𝑚 s 𝐸𝑓𝑜𝑟𝑚 s 𝐸𝑓𝑜𝑟𝑚 s

(Mg)68+24x(Al)16(Al)12(Va)2 1.853 0.837 1.576 0.877 1.486 0.858 1.314 0.833

(Mg)68+24x(Al)16(Gd,)12(Va)2 -1.679 0.602 -1.583 0.832 -1.315 0.762 -1.195 0.808

(Mg)68+24x(Al)16(Mg)12(Va)2 0.608 0.770 0.580 0.864 0.509 0.830 0.452 0.826

(Mg)68+24x(Gd)16(Al)12(Va)2 -8.076 0.882 -6.856 0.910 -6.154 0.914 -5.194 0.901

(Mg)68+24x(Gd)16(Gd)12(Va)2 -0.859 0.713 -0.318 0.854 -0.218 0.807 -0.162 0.820

(Mg)68+24x(Gd)16(Mg)12(Va)2 -2.952 0.749 -2.265 0.860 -1.943 0.822 -1.697 0.824

(Mg)68+24x(Mg)16(Al)12(Va)2 1.100 0.571 0.945 0.826 0.856 0.749 0.684 0.805

(Mg)68+24x(Mg)16(Gd)12(Va)2 0.666 0.519 0.294 0.816 0.309 0.728 0.217 0.800

(Mg)68+24x(Mg)16(Mg)12(Va)2 0.347 0.600 0.196 0.832 0.221 0.761 3.162 0.808

(Mg)68+24x(Gd)16(Al)12(Al)2 -11.250 0.900 -9.492 0.906 -8.157 0.898 -7.167 0.905

(Mg)68+24x(Gd)16(Al)12(Mg)2 -11.366 0.885 -9.577 0.902 -8.232 0.879 -7.216 0.911

37

(Mg)68+24x(Gd)16(Al)12(Gd)2 -11.385 0.849 -9.582 0.842 -8.316 0.876 -7.275 0.900

Figure 3.3 Comparison of (a) heat capacity of HCP Mg with experimental data from

SGTE[82], (b) heat capacities of Mg-only LPSO phases, and (c) Gibbs energy differences of

various pure Mg LPSO phases with respect to HCP Mg.

38

Calculated formation energies of all LPSO endmembers at 0 K with pure element reference

states are obtained from Equation 3.12 and summarized in Table 3.4 and Figure 3.4.

Equation 3.12 𝑬𝒇𝒐𝒓𝒎(𝑳𝑷𝑺𝑶) = 𝑬(𝑳𝑷𝑺𝑶) − 𝟏

𝑵∑ 𝑵𝒊𝑬𝒊𝒊

where Ei is the total energy of stable bulk state of species per atom of species i, and Ni the moles

of species i. The calculated formation energies at 0 K are in good agreement with the results by

Saal et al.[14]. To illustrate the energetics for atomic occupancy at the interstitial site in the L12

cluster, the formation energies of endmembers with and without interstitial elements are plotted in

Figure 3.5. It can be seen in Figure 3.5b that the formation energy of GdIIAlIIIVaIV, i.e., with the

interstitial site being vacant, is substantially higher than the surfaces of the formation energy

bounded by GdIIAlIII(Al or Gd or Mg)IV, i.e. with the interstitial site occupied. This signifies that

the interstitial sites are energetically favored with atoms and confirms the conclusions by Kishida

et al.[18], [69]. The parameters in Gibbs energy functions of all endmembers, see Equation 3.9,

and sublattice interaction parameters, see Equation 3.11, are evaluated and listed in Table 3.1 and

Table 3.2, also in Appendix B. It can be seen that the interaction parameters in the interstitial

sublattice, i.e., sublattice IV, is rather small, indicating the near ideal mixing in the sublattice due

probably to their very small compositional variation.

39

5 6 7 8

-12

-10

-8

-6

-4

-2

0

2

24R14H18R

GdIIAl

IIIVa

IV

GdIIAl

IIIAl

IV

GdIIAl

IIIMg

IV

GdIIAl

IIIGd

IV

AlIIAl

IIIVa

IV

AlIIGd

IIIVa

IV

AlIIMg

IIIVa

IV

GdIIGd

IIIVa

IV

GdIIMg

IIIVa

IV

MgIIAl

IIIVa

IV

MgIIGd

IIIVa

IV

MgIIMg

IIIVa

IV

GdIIAl

IIIMg

IV-Saal

Form

ation

Energ

y (

kJ/m

ol-

ato

m)

Number of layers in Structural Block

10H

Figure 3.4 Formation energies of endmembers of the 10H, 18R, 14H, and 24R LPSO phases

at 0 K. The data sets of GdIIAlIIIMgIV-Saal (×) were taken from the literature[14].

40

Figure 3.5 (a) Composition ranges of GdIIAlIII(Mg, Gd, Al, and Va)IV endmembers of the

10H(o), 18R(⟡), 14H(x), and 24R(∇) LPSOs, (b) formation energies (in kJ/mole-atom) of the

GdIIAlIII(Mg(x), Gd(∇), Al(⟡), and Va(o))IV endmembers at 0 K in compositional space.

41

Figure 3.6 shows the isothermal sections of the Mg-Al-Gd system at 673.15 K and 798.15

K, respectively, together with the experimental results in the literature[6], [109], [110]. It is noted

that at 673.15 K (Figure 3.6a), the observed phases in the experimental work by De Negri et

al.[110] is well reproduced. At 798.15 K (Figure 3.6b) the observed three-phase regions, HCP

Mg+Al2Gd (Laves C15)+18R LPSO[109] and HCP Mg+Mg5Gd+18R LPSO[6] are reproduced.

Figure 3.6 Isothermal sections of the Mg-Al-Gd system at 673 K (a) and 798 K (b). All

experiment data (the thick lines and the symbols) at 673 K were measured by De Negri et

al.[110] (𝛁 : Al3Mg2 + FCC Al + Lav C36, ∆: GdMg + GdMg3, □: MgGd, ⟡: GdMg +

AlGd2, ⧖:GdMg + Lav C15 + GdMg3), those at 798.15K were taken from Kishida et al.[6],

[109], including HCP Mg + Al2Gd (Laves C15) + 18R LPSO (○[109]) and HCP Mg + Mg5Gd

+ 18R LPSO (𝛁[6] and ⟡[6]) phases.

To further illustrate the stability of the LPSO phases, the isothermal sections from 838.15

K to 673.15 K at the Mg-corner are calculated and plotted in Figure 3.7. It can be seen that the 18R

LPSO phase is stable between 838.15 K to 723.15 K (see Figure 3.7a-e), which is in good

agreement with the measured three-phase regions at 838.15 K by Lu et al.[111], at 823.15 K by Dai

et al.[112], at 798.15 K by Kishida et al.[6], [109], and at 773.15 K by Gu et al.[113]. It should be

noted that the experimental data point at 823.15 K from Dai et al.[112] is treated as Mg–7.9 at.%

42

Al–10.9 at.% Gd(+Y) based on their assumption that Gd appears to be partially substituted by Y,

although the exact composition is Mg–7.9 at.% Al–7.8 at.% Gd–3.1 at.% Y. The small

compositional triangles in Figure 3.7d-f represent the composition ranges of three types of

interstitial atoms in the L12 cluster. It can be seen that the equilibrium composition of the 18R

LPSO phase is with Gd at the interstitial site, while the 14H LPSO phase has Mg at the interstitial

site.

43

Figure 3.7 Mg-corner of the isothermal sections of the Mg-Al-Gd system at (a) 838.15 K, (b)

823.15 K, (c) 798.15 K, (d) 773.15 K, (e) 723.15 K, and (f) 673.15 K, with experimental

compositions from Lu et al.[111] at 838.15 K (𝛁) with HCP Mg + Al2Gd (Laves C15) + 18R

LPSO phases in equilibrium, at 823.15 K from Dai et al.[112] (*) with 18R LPSO phase

composition of Mg–7.9 at.% Al–10.9 at.% (Gd+Y), at 798.15 K from Kishida et al. with HCP

Mg + Al2Gd (Lav C15) + 18R LPSO (○[109]) and HCP Mg + Mg5Gd + 18R LPSO (𝛁[6] and

⟡[6]) phases in equilibrium, and at 773.15 K from Gu et al.[113] with 18R LPSO, respectively.

The small triangles represent the composition ranges of GdIIAlIII(Mg, Gd and Al)IV

endmembers.

44

Figure 3.8 shows an enlarged view of the composition of the 18R LPSO phase in the

isothermal section at 798.15 K, slightly away from the endmember with Gd in the interstitial site

and the stable phase composition of 18R LPSO is on the line between GdIIAlIIIGdIV and

GdIIAlIIIVaIV. The calculated vacancy concentration at the interstitial site is less than 1% which

means that most of the interstitial sites are filled by Gd. Kishida et al.[18] recently indeed observed

the coexistence of vacancy and atoms at the interstitial site in the Mg-Y-Zn LPSO phases by STEM.

They also illustrated the atomic type of the occupancy at the interstitial site is to be Y with high

probability by annular bright-field (ABF)-STEM image, which agrees well with the present model.

Based on their average phase composition and STEM images of 18R LPSO, the concentration of

vacancies at the interstitial site is higher than those from the present model. More experimental data

are needed to refine the model.

Figure 3.8 An enlarged view of the isothermal section of the Mg-Al-Gd system at 798.15 K,

showing the composition homogeneity range of the 18R LPSO phase. Blue triangle indicates

the composition ranges of GdIIAlIII(Mg, Gd, Al, and Va)IV endmembers as the same triangle

as Figure 3.5.

Figure 3.9 shows the calculated isopleths of the Mg-Al-Gd system with the molar ratio of

Al : Gd = 0.7. As shown in the isothermal section at 838.15 K in Figure 3.7a, the three-phase

region of Laves C15 + HCP Mg + 18R LPSO is reproduced in the alloy of Mg91.5Al3.5Gd5 (at.%)

45

observed by Lu et al.[111]. Kishida et al.[17] and Yokobayashi et al.[19] heat-treated the sample

of Mg91.5Al3.5Gd5 (at.%) at 823 K for 2 hours and at 673 K for 10 hours. They observed the Mg5Gd

+ HCP Mg + LPSO (18R+14H) phases, while the calculated phase equilibrium includes Mg5Gd +

HCP Mg + 14H LPSO phases. Since a four-phase equilibrium in a ternary system is an invariant

equilibrium based on Gibbs phase rule, this discrepancy indicates that the sample may have not

reached full equilibrium. Nevertheless, the 18R phase and the 14H + 18R phases observed at 798

K and 673 K, respectively, indicate that there may be a phase boundary between these two

temperatures, which is predicted by the present thermodynamic modeling of the LPSO phases to

be at 760 K.

Figure 3.9 Isopleth sections of the Mg-Al-Gd phase diagram with the molar ratio of Al:Gd

being 0.7 (a) and an enlarged view of the Mg-rich region (b), with experimental compositions

from Lu et al.[111] at 838.15 K (+) with HCP Mg, Lav C15 and LPSO (18R) phases in

equilibrium, and from Kishida et al.[17] at 673.15K (*) with HCP Mg, Mg5Gd and LPSO

(14H + 18R) phases, respectively.

46

Chapter 4

Predictive Modeling of Hardness of Brittle and Ductile Materials

4.1 Introduction

The scientific community expanded great efforts to design ultrahard materials for cutting

and polishing applications[114], [115] by interpreting hardness as “the extent to which a given solid

resists both elastic and plastic deformation”[116]. Based on this interpretation, previous ceramic

material models primarily focused on elastic properties[117]–[119] as most ceramics behave

elastically with no plastic deformation up to fracture at room temperature[120]. However, the

differentiation between metals and ceramics is somewhat arbitrary as it is simply based on the

extent of the observed macro-scale plastic deformation or the apparent lack of it, leading to the

separate treatment of ductile and brittle material classes in previous research. More importantly,

this limited interpretation of hardness does not consider the ratios between reversible elastic and

irreversible plastic deformations, another important concept in modeling as it provides information

on the extent to which a material is ductile.

Without such considerations, researchers have developed semi-empirical theoretical

hardness models, either focusing on elastic properties (e.g., shear (G)[117], bulk (B)[118] moduli,

and Pugh ratio (G/B)[119]) or based on chemical bond properties (e.g., length[121], charge

density[121], iconicity[121], strength[116], and electronegativity[122]). Despite the contribution

by previous models, they do not fully capture the complexity of hardness, only limited to certain

materials. For example, Chen et al.[119] correlated the hardness to elasticity based on the Pugh

ratio, predicting hardness for ceramics, mostly zincblende, rocksalt and diamond structures.

47

Nevertheless, this model is not predictive for metals, since plastic deformation is not considered

and the Pugh ratio, as a measure of ductility, is restrictive and only valid within similar crystal

structures and melting temperatures[123].

In order to improve the predictive power of a unified hardness model, the material’s

fundamental deformation behavior must be elucidated, capturing both reversible elastic and

irreversible plastic deformation characteristics[124] as well as the ratio of total to plastic

indentation depth. Especially the material’s plastic deformation must be modeled correctly since

first, the plasticity characteristics affect the resultant deformation during the hardness measurement,

and second without this information the ratio cannot be calculated. Plastic deformation is mainly

due to the creation and motion of dislocations[124] and macroscopically affects the flow stress or

the critical shear stress for dislocation motion,[125] via edge or screw dislocations. Further, flow

stress is significantly affected by the operating slip systems that govern the propagation of

dislocations[125]. Therefore, the key to modeling plastic deformation is considering the materials'

slip systems.

Herein, we present a physics-based model for indentation hardness capable of predicting

the response of both ceramics and metals regardless of their bonding types. It uses experimentally

measured hardness and the ratio of total to plastic indentation depth obtained from load-

displacement curves. The flow stress is estimated from the Peierls-Nabarro (PN) stress and the

material slip systems are considered through the dependence of the PN stress on slip systems. Two

model parameters are determined based on input data from either first-principles calculations or

experiments, accelerating materials design for optimized ultrahard performance. Ultimately, our

research contributes to a greater understanding of the deformation behavior during hardness

measurements.

48

4.2 Presentation of the New Model

Due to the simplicity of its usage, the Vickers’ hardness measurement is the ubiquitous

method for evaluating the mechanical behavior of materials. It uses a pyramidal shaped diamond

tip. The working principle of the Vickers hardness (𝐻𝑣) is the division of the maximum applied

force (𝐹) by the plastically deformed area (𝐴𝑝) obtained from the two diagonals of the indented

surface after unloading the tip,[126], [127] as shown in Figure 4.1.

Figure 4.1 Schematics of (a) Vickers tip geometry, (b) geometry changes during indentation,

and (c) a side view of indentation.

Equation 4.1 𝑯𝒗 =𝑭

𝑨𝒑

During loading, the total deformation consists of both elastic and plastic deformation; while

during unloading the elastic deformation is recovered, the resulting indentation represents the

plastic deformation only. Based on this concept, we obtain the modified hardness model equation

below. The detailed derivation is in 4.5 Full Derivation of the Hardness Model.

49

Equation 4.2 𝑯𝒗 = 𝒄 𝑮 (𝒉𝑻

𝒉𝒑)𝟐

𝐭𝐚𝐧𝟑/𝟐𝜶 𝐜𝐨𝐬 𝜶

where ℎ𝑇 and ℎ𝑝 are the indentation depths at the maximum loading and after unloading,

respectively; 𝛼 is the angle of the indenter tip (𝛼 = 22° for a Vickers diamond tip), 𝐺 is the shear

modulus of the material, and 𝑐 is a scaling factor.

In this model, we closely look into ℎ𝑇/ℎ𝑝 ratio values in the load-displacement curves of

Vickers and Berkovich hardness measurements[127]–[130]. For purely plastic materials, ℎ𝑇/ℎ𝑝 =

1, and for purely elastic materials, ℎ𝑇/ℎ𝑝 = ∞. Different from prior approaches is evident in the

fact that the ℎ𝑇/ℎ𝑝 ratio is equivalent to the dissipated energy ratio between (elastic+plastic) and

plastic deformations. Most critical to the new model, and what follows is the estimation of the

plastic energy, which originates from dislocation energy, and plays a significant role in this model.

For capturing the dislocation energy, the concept of flow stress is introduced. After modeling the

ℎ𝑇/ℎ𝑝 ratio and scaling factor c, Equation 4.3 was obtained.

Equation 4.3 𝑯𝒗 = 𝟎. 𝟏𝟔𝟏𝟓 𝑮 (𝟏 +𝟖𝟎𝟎𝟎

𝟑

𝒘

𝒃(𝒃

𝒔)𝟐(𝝉𝑷𝑵

𝑮)𝟐)𝟐

𝒆−𝟐.𝟐𝒌 (𝒃

𝒔)𝟒𝐭𝐚𝐧𝟑/𝟐 𝒂𝐜𝐨𝐬𝒂

where 𝜏𝑃𝑁 and 𝑤 are the Peierls-Nabarro stress and dislocation width, respectively, 𝑏/𝑠 is the

Burgers vector and slip plane spacing, and 𝑘 is the ratio of shear and bulk moduli. A detailed

derivation is shown in 4.5 Full Derivation of the Hardness Model.

4.3 Validation and Prediction

In the present hardness model, the active slip systems, 𝑏/𝑠, and dislocation width, 𝑤, of

materials as well as their elastic properties and melting temperatures constitute the input data.

Typically, in a crystalline material there are many slip systems that coexist, but specific slip systems

are activated at a given condition. The active slip systems of materials can be determined not only

50

by the crystal structures but also by materials’ bonding nature such as the electronic structure of

materials[131]. Note, for the purposes of this model, we will only consider the material hardness at

room temperature. Furthermore, the active slip system in a material usually is accompanied by

other equivalent slip systems (typically more than one). For example, typical active slip systems of

the diamond cubic structure are shuffle and glide, and those of body centered cubic (BCC) structure

are ½ [111]{110}, ½ [111]{211} and ½ [111]{321} slip systems. Therefore, the active slip systems

can be determined by portions of the energetic descriptions of each slip system. The present model

has considered the major active slip systems of each structure shown in Table 4.1.

Table 4.1 Slip systems of different crystal structures.

Structur

e

Dislocatio

n type

Burgers

vector, �� �� length, b

Slip

plane Interspacing, s

example

FCC edge 1

2[110] 𝑎0/√2 {111} √(

√3𝑎03)

2

+ (√3𝑎03)

2

Ni, Al,

Au, Ir, Rh

FCC screw 1

2[110] 𝑎0/√2 {110}

1

2√𝑎02 + (

√2𝑎02)

2

Ir, Rh

BCC screw 1

2[111] 𝑎0√3/2 {110}

√2𝑎0

√3

Mo, V,

Fe, W

BCC screw 1

2[111] 𝑎0√3/2 {211}

𝑎0

√2 Nd, W

HCP-

basal edge

1

3[1210] 𝑎0 {0001}

𝑐

2

Mg, Zn,

Cd

HCP-

prism edge

1

3[1210] 𝑎0 {1010}

√3𝑎02

Ti, Zr

HCP-

pyramidal edge

1

3[1210] 𝑎0 {1011}

1

√43(1𝑎0)2

+ (1𝑐)2

Co

HCP-

twin edge

1

2[1011]

1

2√𝑎0

2 + 𝑐2 {1012}

1

√43(1𝑎0)2

+ (2𝑐)2

Be

Diamon

d cubic edge

1

2[110] 𝑎0/√2 {111} √(

√3𝑎06)

2

+ (√3𝑎06)

2

C(diamon

d), Si, Ge

51

Diamon

d cubic screw

1

2[110] 𝑎0/√2 {111} √(

√3𝑎06)

2

+ (√3𝑎06)

2

Si, Ge

Zinc

blende edge

1

2[110] 𝑎0/√2 {111} √(

√3𝑎06)

2

+ (√3𝑎06)

2

SiC,

ZnSe,

ZnS

Zinc

blende edge

1

2[110] 𝑎0/√2 {110} √(

√2𝑎04)

2

+ (√2𝑎04)

2

ZnSe,

ZnS

Rocksal

t edge

1

2[110] 𝑎0/√2 {111} √(

√3𝑎06)

2

+ (√3𝑎06)

2

TiC, TiN,

ZrC

Rocksal

t edge

1

2[110] 𝑎0/√2 {110} √(

√2𝑎04)

2

+ (√2𝑎04)

2

MgO

Spinel edge 1

2[110] 𝑎0/√2 {111} √(

√3𝑎06)

2

+ (√3𝑎06)

2

MgAl2O4

We will review our results for each crystallographic system. Face-centered cubic (FCC)

metals are ductile with low flow stresses, resulting in their ℎ/ℎ𝑝 ratios being close to 1. The active

slip system of FCC metals at room temperature is mainly ½ [110]{111} since the most densely

packed planes of FCC metals are {111}[132]. The calculated hardness values from the present

hardness model are in good agreement with experimental data except for Ir, Rh, and Th[132], [133]

as shown in Figure 4.2a. Although the crystal structures of Ir and Rh are FCC, their slip systems

are complex since both planar and non-planar core structures can coexist[134], [135]. Through

atomistic simulations using a bond-order potential Cawkwell et al.[131] found that the non-planar

core structure is due to the ½ [110] screw dislocation originating from unsaturated d-bonds. This

aspect of interatomic bonding distinguishes Ir, Rh, and Th from most other FCC metals. The present

model can predict the hardness with different slip systems as shown in Figure 4.2a for Ir, Rh, and

Th, and the actual deformation process is likely a result of a mixture of various slip systems.

52

53

Figure 4.2 Hardness comparisons of (a) FCC, (b) BCC, and (c) HCP materials with respect

to experimental data. Solid and open symbols represent the predicted values using elastic

properties from experiments and first-principles calculations. Red dashed lines indicate value

equality, vertical dotted lines connect the hardness between the slip systems.

BCC metals are usually brittle due to the higher flow stress of screw dislocations, resulting

in higher hardness values than other metals as shown in Figure 4.2b. The active slip systems of

BCC structures at room temperature are either or combinations of ½ [111]{110}, ½ [111]{211} and

½ [111]{321} screw dislocations[136]. The active slip system of Mo[137], Fe[138], V[139], and

Cr[140] at room temperature is the ½ [111]{110} screw dislocation, while those of Nb[136], [141]

and W[136], [142] at room temperature consist of both ½ [111]{110} and ½ [111]{211} screw

dislocations. Based on the calculated hardness from the present model, the active slip systems of

W are predicted to be the mixture of ½ [111]{110} and ½ [111]{211} screw dislocations. The active

slip systems of BCC metals predicted by the present model agree well with those observed

experimentally as shown in Table 4.2. It should be noted that in general the active slip systems

change with the operating temperature. This is especially true for Nb and W; the active slip systems

of Nb and W at room temperature are different from those at 77 K[136]. The prediction of

temperature dependence of hardness will be discussed in subsequent chapter.

Table 4.2 Slip systems at room temperature in BCC metals.

Model W Mo Nb Fe V Cr

This model

prediction 110, 211 110 110, 211 110 110 110

Weinberger et

al.[136] 110, 211

110,

some 211 110 110 110, 211 -

Fritz et al.[140] - - - - - 110

Bressers et

al.[139] - - - -

110,

Rarely 211 -

Schadler et

al.[143] 110, 211 - - - - -

54

Butt model[144] 110[144],

211[145] 110[144] 211[144] 211[144] 211[144] 211[145]

Finnis/Sinclair

potentials[146] 110 110 211 - TW 211

Deformation modes in hexagonal close-packed (HCP) metals are mainly slip and

deformation twinning[147]–[149]. In this study, slip and the formation of deformation twinning are

considered since the formation of twins can be considered as multiple slips on the slip plane which

corresponds to the twin boundary. However, the interactions between slip and twinning are not

considered in the present study due to their complexity, which will be discussed in a separate study.

The slip systems in HCP metals are similar to those in FCC metals due to the close-packed nature

of both crystal structures. In HCP metals the dominant slip direction is the close-packed direction

1

3[1120], which is also the Burgers vector. The active slip planes are mainly the basal {0001},

prismatic {1010}, and pyramidal {1012} planes. The basal plane, like the {111} planes in FCC

metals, is the favored slip plane in Cd, Zn, and Mg, but in Ti and Zr, the prismatic plane is more

strongly favored due to unsaturated d-electrons[132]. The predictions from the present model agree

well with experimental data as shown in Figure 4.2c. It should be noted that the deformation modes

of HCP Be, Y, and Hf include not only basal and prismatic slip, but also {1012} deformation

twinning along the [1011] direction.[147]–[149] Based on the hardness comparison in Figure 4.2c,

noticeable portions of deformation twinning as well as basal and prismatic slips contributes to the

hardness of Be and Y.

Active slip systems of ceramic materials are even more complicated than those of metals.

Ionicity and covalency of the materials affect the slip systems even in the same crystal structure.

For example, within the same zincblende (ZB) structure, carbides and nitrides such as SiC and AlN

have the {111} slip systems with a Burgers vector of 1

2[110], while ionic bonded ZB structures

such as ZnSe, ZnS, and ZnTe have the {110} and {111} slip planes. Therefore, ZB structures are

55

divided into ionic and covalent ZB structures. Furthermore, the active slip systems of NaCl

structures are the same as those of ZB structures, i.e., the {111} slip system for carbide and nitride

NaCl structures such as TiN and Ti,C, and the {110} and {001}slip systems[150]–[152] for ionic

bonded NaCl structures such as MgO and NaCl, respectively.

Similarly, understanding of slip systems of semiconductor materials is crucial for

determining dislocation behavior and plastic deformation during epitaxial growth and device

processing in order to improve their optical and electronical properties and homogeneity[153].

Zincblende and diamond cubic crystal structured semiconductor materials have two major slip

systems, which are glide edge and shuffle screw dislocations[154]–[156]. The present model is

capable of predicting the energetic contribution of each active slip system on hardness. It is known

that the active slip systems in semiconductor materials such as Si and GaAs are combinations of

glide edge and shuffle screw dislocations,[154]–[156] where the zig-zag shaped 1

6[121] partial

dislocations of the glide-set in ZB and diamond cubic crystal structures can be treated as a 1

2[110]

full dislocation in the present model. As shown in Figure 4.3, the experimental hardness values of

the semiconductor materials such as GaAs, GaP, and AlSb are in between the predicted values from

these glide edge and shuffle screw slip systems. This indicates that the glide-set and shuffle-set slip

systems coexist during the indentation process, but the hardness contribution from shuffle screw

dislocations is much smaller than those from glide edge dislocations. The experimental hardness

values (12 and 8.8 GPa) of Si and Ge seem to be located in between the contributions of glide-set

(8.6 and 5.3 GPa) and shuffle-set (28.5 and 13.8 GPa) slips, respectively. If only contributions from

these two sets of slip types, the glide-set of slip is dominant, representing 83% and 59% of the total

hardness for Si and Ge, respectively, though the possible phase transformation to a tetragonal β-tin

structure during indentation process between 0 and 300 °C[157]–[159] with pop-out or elbow in

the load-displacement curve could potentially complicate the situation. Furthermore, the

56

quantitative contributions from the individual slip systems would also depend on the migration

barrier of the dislocation in terms of possible kink formation and migration[160].

0 5 10 150

10

20

3040

45

Si

ZnS

ZnSe

ZnTe InSb

GaSb

AlSb

InAs

AlAs

InP

GaAs

Ge

AlP

GaP

Expt. G&B Glide

Calc. G&B Glide

Expt. G&B Shuffle

Calc. G&B Shuffle

Mo

de

lle

d H

v (

GP

a)

Experimental Hv (GPa)

Figure 4.3 Hardness comparisons of ceramic materials with respect to experimental data.

Solid and open symbols represent the predicted values using elastic properties from

experiments and first-principles calculations. Red dashed lines indicate value equality,

vertical dotted lines show the differences between glide edge and shuffle screw slip systems.

4.4 Discussion

It is known that the indentation hardness represents the materials’ ability to resist plastic

deformation. The lack of materials’ properties on the imparted plastic deformation in existing

hardness models in the literature has limited the predictive capability of these models. It is noted

that the Vickers and Berkovich hardness tests use a slow and fixed loading rate to avoid the effects

of impact and strain rate,[161] thus enabling the use of the static attributes of plastic deformation

in the present hardness model, including the Peierls-Nabarro flow stress, dislocation width, Burgers

vector, and slip plane spacing in addition to the elastic properties conventionally used in existing

hardness models. Based on the well-established concept of active slip systems during plastic

57

deformation, the present model captures the competition among possible slip systems among pure

elements and compounds originating from their unsaturated d-bonds[131].

With model parameters evaluated from available experimental data in the literature, the

present model is able to satisfactorily cover the hardness ranges from 0.1 GPa of pure metals to 100

GPa of ceramic materials as shown in Figure 4.4. For materials with more than one slip system

being activated, the apparent hardness values seem to fall between the hardness values of the

activated slip systems. Figure 4.5 and Table 4.3 further shows that the PN flow stress decreases

exponentially with increasing dislocation width, and the PN flow stress agrees well with

experimentally determined yield stress at low temperature.

0 20 40 60 80 100 120

0

20

40

60

80

100

120

0 2 4 6 8 100

2

4

6

8

10

From Expt. G&B

From Calc. G&B

Mo

de

lled

Hv (

GP

a)

Experimental Hv (GPa)

Figure 4.4 Hardness comparisons of all tested materials with respect to experimental data.

Solid and open symbols represent the predicted values using elastic properties from

experiments and first-principles calculations. The experiments for all materials both

hardness and elastic properties data from Table 4.6. Red dashed lines indicate value equality.

58

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4

10-6

10-5

10-4

10-3

10-2

10-1

100

Dislocation width, w0K

Flo

w o

r Y

ield

str

ess, P

N/G

or Y

/G

Expt. Y/G at low T

Diamond C and Si

BCC metals

FCC metals

HCP metals

Calc. PN/G at 0K

Carbides or Nitrides

Semiconductors

Other ceramics

BCC metals

FCC metals

HCP metals

Figure 4.5 Comparison of Peierls-Nabarro flow stress at 0 K with experimental yield stress

at low temperatures (4~77K) as a function of dislocation width. Data and references are listed

in Table 4.3.

Table 4.3 Comparison of PN flow stress at 0K with experimental yield stress (𝝉𝒀

𝑮) at low

temperatures (4~77 K) as a function of dislocation width.

Material Structure 𝝉𝑷𝑵𝑮

𝝉𝒀𝑮

References

C Diamond cubic 2.34 × 10−2 5.71 × 10−2 [162]

Si Diamond cubic

1.21 × 10−2 ~2.66× 10−2

1.00 × 10−1 [163]

Fe BCC 2.67 × 10−3 5.50 × 10−3 [164]

Mo BCC 2.67 × 10−3 5.11 × 10−4 [165], [166]

Nb BCC 2.67 × 10−3 8.80 × 10−4 [165]

Al FCC 2.23 × 10−5 1.09 × 10−5 [165]

Al FCC 4.00 × 10−5 [167]

Ni FCC 4.28 × 10−5 6.58 × 10−5 [165]

Cu FCC 2.33 × 10−5 5.31 × 10−6 [165]

Cu FCC 1.04 × 10−5 [168]

Ag FCC 1.60 × 10−5 2.00 × 10−5 [169]

Ag FCC 2.60 × 10−5 [170]

Be HCP (basal

slip) 6.26 × 10−3 7.16 × 10−4 [171]

59

Be HCP (prism

slip) 3.67 × 10−3 1.35 × 10−3 [165]

Mg HCP 1.07 × 10−3 1.00 × 10−4 [172]

Mg HCP (prism

slip) 6.65 × 10−4 2.00 × 10−3 [165]

Cd HCP (basal

slip) 1.83 × 10−4 2.10 × 10−5 [173]

Zn HCP 5.00 × 10−4 2.00 × 10−5 [174]

Ti HCP 4.60 × 10−4 1.16 × 10−3 [175]

Ti HCP 1.74 × 10−3 [175]–[177] 𝜏𝑃𝑁

𝐺 was calculated from Equation 4.27 or Equation 4.28 with the information in Table 4.6.

As discussed in detail in 4.5 Full Derivation of the Hardness Model, the ℎ𝑇/ℎ𝑝 ratio

equals the ratio of the total energy and plastic energy dissipated during the indentation, i.e., 𝐸𝑇/𝐸𝑝.

This energy ratio can be correlated with the plasticity index[178] or the indentation ductility

index[179]. In the present work, we define the indentation ductility index as follows with the details

presented in 4.5 Full Derivation of the Hardness Model.

Equation 4.4 𝑫 =𝑬𝒑

𝑬𝒆=𝒉𝒑

𝒉𝒆=

𝟏

𝟖𝟎𝟎𝟎

𝟑

𝒘

𝒃(𝒃

𝒔)𝟐(𝝉𝑷𝑵𝑮)𝟐

where 𝐸𝑒 and ℎ𝑒 are the elastic energy and elastic deformation depth during indentation,

respectively. Figure 4.6 plots 𝐷 with respect to the dislocation width at room temperature. For

highly plastic materials, i.e., ductile materials, 𝐷 approaches ∞, while for highly elastic materials

𝐷 approaches zero.

60

0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.410

-1

101

103

105

107

109

Carbides and Nitrides

Semiconductors

other cubic ceramics

BCC metals

FCC metals

HCP metalsIn

denta

tion

du

ctilit

y inde

x, D

Dislocation width, w

Figure 4.6 Indentation ductility index as a function of dislocation width at room temperature.

Finally, it is important to recognize that the power of the present hardness model is in its

ease of expandability to a more comprehensive hardness model which includes the effects of

temperature, grain size or twin boundary thickness, and solute atom size. For example, the

temperature dependent hardness, i.e., hot hardness, can be calculated based on the temperature

dependence of the PN flow stress and dislocation width; therefore, these factors can be possibly

included in the extended model. The details will be discussed in future publications.

4.5 Full Derivation of the Hardness Model

4.5.1 Derivation of the Hardness Equation

The Vickers hardness is defined as follows:

61

Equation 4.5 𝑯𝒗 =𝑭𝒎𝒂𝒙

𝑨𝒑

where 𝐹𝑚𝑎𝑥 is the maximum applied force, and 𝐴𝑝 the plastically deformed area after the indenter

tip is unloaded (see Figure 4.1a). 𝐴𝑝 can be obtained from the resulting indentation as follows[180]

Equation 4.6 𝑨𝒑 =𝒅𝒑𝟐

𝟐 𝐜𝐨𝐬 𝒂′

where 𝑑𝑝 is the diagonal length of the resulting indentation on the sample surface, and 𝛼′ the angle

between the sample surface and the indented surface after unloading the tip. (△ABO’ in Figure

4.1b).

To estimate the applied maximum force, 𝐹𝑚𝑎𝑥, we assume that the pyramid indentation can

be divided into four triangular pyramid indenters and each part is a pure shear process[181] with

two forces acting on xz(𝜏𝑦𝑥) and xy(𝜏𝑧𝑥) planes, respectively, as shown in Figure 4.1b. As the

indentation process contains both elastic and plastic deformations, 𝐹𝑚𝑎𝑥 can be written as

follows[182]

Equation 4.7 𝑭𝒎𝒂𝒙 = 𝑨𝑻𝝉𝑻

where 𝐴𝑇 is the shear area, parallel to the applied force during loading, and 𝜏𝑇 the total shear stress,

both at maximum loading[182]. The shear area is the parallelogram marked by AO’PQ in Figure

4.1b. It is a complex quantity due to the difficulty in determining the elastic deformation depth

(𝑂′𝑃 in Figure 4.1b), which includes the small atomic displacements by a single dislocation

nucleation and reaches further down to the depth of the orange colored area in Figure 4.1c. The

shear area is conventionally assumed to be proportional to the deformed area (△AOO’ in Figure

4.1b) of the indentation during loading[181] with a material dependent coefficient, i.e.,

Equation 4.8 𝑨𝑻 =𝟏

𝟐𝒄𝟏𝒅𝑻

𝟐 𝐭𝐚𝐧𝜶

where 𝑑𝑇 is the diagonal length of the indentation at maximum loading, 𝛼 the angle between the

sample surface and indented surface during indentation, being 22o for a Vickers diamond tip, and

62

𝑐1 the material dependent coefficient representing the complexity in determining the shear area to

be discussed later.

In various indentation techniques including spherical and Vickers indentations,[183]–[191]

typical shear stress-strain curves follow approximately the Taylor relation[192]–[194] as follows

Equation 4.9 𝝉𝑻 = 𝒄𝟐𝑮√𝜸𝑻

as shown in Figure 4.7 with 𝑐2 being a geometrical factor, 𝐺 the elastic shear modulus, and 𝛾𝑇 the

total shear strain dictated by the geometry of the indentation tip as shown below

Equation 4.10 𝜸𝑻 =𝒉𝑻

𝒅𝑻= 𝐭𝐚𝐧𝜶

where ℎ𝑇 is the indentation depth at maximum loading.

Figure 4.7 (a) Stress(𝝉 )-strain(𝜸 ) curve during shear deformation. (b) typical Load(F)-

displacement(h) curve during indentation process.

By inserting Equation 4.8, Equation 4.9 and Equation 4.10 into Equation 4.7, the

applied force is obtained

Equation 4.11 𝑭𝒎𝒂𝒙 =𝟏

𝟐𝒄 𝑮 𝒅𝑻

𝟐𝒕𝒂𝒏𝟑/𝟐𝜶

where 𝑐 = 𝑐1𝑐2 to be discussed in section 4.5.2.2 Parameter c.

Based on Equation 4.6 and Equation 4.11, Equation 4.5 can be re-organized as follows

Equation 4.12 𝑯𝒗 = 𝒄 𝑮 (𝒅𝑻

𝒅𝒑)𝟐

𝒕𝒂𝒏𝟑/𝟐𝜶 𝒄𝒐𝒔 𝜶′

63

where 𝑑𝑝 is the diagonal length of the indentation due to the plastic deformation after unloading

the indentation tip (see Figure 4.1c).

The term 𝑑𝑇/𝑑𝑝 can be re-expressed as the indentation depth, ℎ𝑇 tan𝛼

ℎ𝑝 tan𝛼, from the geometry

of the tip (Equation 4.10) and indented surface (ℎ𝑝 = 𝑑𝑝 tan𝛼′) where ℎ𝑝 is the depth of the

indentation due to the plastic deformation after unloading the indentation tip. It can be seen if 𝛼′

equals to 𝛼 , 𝑑𝑇/𝑑𝑝 equals to ℎ𝑇/ℎ𝑝 , i.e., the ratio between the total deformation and plastic

deformation is the same on the sample surface and along the indentation depth. This is an

approximation commonly used in the literature[195]–[197] and also adopted in the present work,

i.e.,

Equation 4.13 𝜶 ≈ 𝜶′

It should be noted that the error introduced by the above approximation is partially canceled when

𝑐 in Equation 4.11, is evaluated, see section 4.5.2.2 Parameter c. Therefore, Equation 4.12 can

be re-written as follows

Equation 4.14 𝑯𝒗 = 𝒄 𝑮 (𝒉𝑻

𝒉𝒑)𝟐

𝒕𝒂𝒏𝟑/𝟐𝜶 𝒄𝒐𝒔 𝜶

4.5.2 Evaluation of Model Parameters

4.5.2.1 𝒉𝑻/𝒉𝒑 Ratio

For modelling of ℎ𝑇

ℎ𝑝 in Equation 4.14, we collected experimentally measured hardness

(𝐻𝑣) and load-displacement (F-h) curves in the literature as listed in Table 4.4 including pure

metals and ceramic materials, with ℎ𝑇 and ℎ𝑝 denoted by the intercepts on the x-axis at maximum

load and after unloading, respectively, as shown in Figure 4.8a. Both Vickers and Berkovich

hardness data are considered because the hardness values from Vickers hardness (four face

64

pyramid) and Berkovich hardness (three face pyramid with an area to depth function which is the

same as that of a Vickers indenter) are equivalent to each other[195], [198]–[200]. All the

experimental elastic properties and hardness data in Table 4.4 are at room temperature.

Furthermore, due to its better agreement on polycrystalline materials[201]–[203], the elastic

properties used in this study are based on Voigt-Reuss-Hill approximation (VRH) method[40],

[41]. When the data from the references are not obtained based on the VRH method or the

references do not provide elastic stiffness matrix (cij), the data from the references are directly used

as mentioned in Table 4.6. They are usually based on the Voigt method[204] which provides the

upper limit of the true polycrystalline constants[201].

Table 4.4 Experimental and calculated 𝒉𝑻/𝒉𝒑 and parameter c for various materials.

structure

indented plane

Ghkl (GPa)

B (GPa)

k v TMelting

(K) slip

system

ℎ𝑇/ℎ𝑝

(exp.)

ℎ𝑇/ℎ𝑝

(calc.)

c

(exp.

)

c

(calc.

)

Al[205],B FCC single

crystal 26.0 77.3 0.34 0.35 933.45 [110]{111}

1.01[1

95] 1.00 0.042 0.046

Cu[205],B FCC (100) 50.9 148.1 0.34 0.35 1358.15 [110]{111} 1.05[2

06] 1.00 0.027 0.032

Ag[205],

B FCC (111) 19.3 102.0 0.19 0.41 1234.95 [110]{111}

1.06[2

07] 1.00 0.076 0.069

Ni[205],B FCC (100) 76.0 168.3 0.45 0.30 1728.15 [110]{111} 1.13[2

08] 1.00 0.027 0.027

Au[205] FCC not

listed 27.5 171.7 0.16 0.42 1337.15 [110]{111}

1.01[2

09] 1.00 0.080 0.075

W[205],B BCC single

crystal 160.0 309.7 0.52 0.28 3695.15

[111]{211}

S

1.15[1

95] 1.01 0.107 0.093

W[205],B BCC single crystal

160.0 309.7 0.52 0.28 3695.15 [111]{110}

S

1.15[1

95] 1.09 0.107 0.168

Fe[205],V BCC not

listed 81.5 166.7 0.49 0.29 1811.15

[111]{110}

S

1.02[2

10] 1.01 0.088 0.090

ZnSe[205],B

ZB (100) 28.0 56.6 0.49 0.29 1798.15 [110]{110} 1.14[1

28] 1.04 0.160 0.202

ZnSe[20

5],B ZB (100) 28.0 56.6 0.49 0.29 1798.15 [110]{111}

1.14[1

28] 1.43 0.160 0.454

GaAs[205],B

ZB (100) 49.5 79.9 0.62 0.24 1511.15 [110]{111} 1.49[1

28] 1.60 0.421 0.310

GaP[205],B

ZB (100) 59.3 93.7 0.63 0.24 1750.15 [110]{111} 1.60[1

28] 1.67 0.395 0.313

SiC[211],

B ZB

not listed

192.0 225.0 0.85 0.17 3003.15 [110]{111} 2.04[2

12] 1.95 0.281 0.307

SiC[211],

B ZB

polycry

stalline 192.0 225.0 0.85 0.17 3003.15 [110]{111}

1.79[2

13] 1.95 0.281 0.307

SiC[211],

V ZB

polycrystalline

192.0 225.0 0.85 0.17 3003.15 [110]{111} 1.79[2

13] 1.95 0.365 0.307

SiC[211],

V ZB

polycry

stalline 192.0 225.0 0.85 0.17 3003.15 [110]{111}

1.72[2

10] 1.95 0.365 0.307

Si[205],B DC (100) 69.2 101.1 0.68 0.22 1687.15 [110]{111} 2.26[1

28] 1.48 0.313 0.404

65

Si[205],B DC (100) 69.2 101.1 0.68 0.22 1687.15 [110]{111}

S

2.26[1

28] 3.47 0.313 0.444

c-

BN[214],

B ZB (111) 405.4 399.7 1.01 0.12 3246.15 [110]{111}

2.50[2

12] 2.42 0.162 0.200

BC2N[21

5],B ZB

not

listed 445.0 408.0 1.09 0.10 3273.15 [110]{111}

2.85[2

12] 2.52 0.139 0.184

MgO[205],V

NaCl polycrystalline

130.0 160.0 0.81 0.18 3125.15 [110]{100} 1.16[2

10] 1.09 0.225 0.150

MgAl2O4

[210],V

spine

l

polycry

stalline 96.0 180.0 0.53 0.27 2408.15 [110]{111}

1.47[2

10] 1.35 0.427 0.602

*B and V in column 1 for Berkovich and Vickers indentation methods, respectively.

Figure 4.8 Typical Load-displacement curves (a) F-h curve and (b) √𝑭-h curve. Red lines are

loading curves and blue lines are unloading curves, and green dot line represents only the

plastic contribution from Equation 4.16.

As shown by Sakai et al.,[210], [213] there is a linear relationship between the square root

of the load and displacement, i.e., √𝐹 ∝ ℎ, as shown in Figure 4.8b. Moreover, Oliver et al.[130],

[195] also observed this relationship for the pyramid shaped indenter. Using the Maxwell

combination of the elastic and plastic elements in a viscoelasticity model, the following equations

can be obtained[216]

Equation 4.15 𝒉𝑻 = 𝒉𝒑 + 𝒉𝒆

Equation 4.16 𝑭𝒎𝒂𝒙 = 𝒌𝑻𝒉𝑻𝟐 = 𝒌𝒆𝒉𝒆

𝟐 = 𝒌𝒑𝒉𝒑𝟐

66

where 𝑘𝑇 , 𝑘𝑒 , and 𝑘𝑝 are the coefficients for total, elastic, and plastic deformation, respectively,

and ℎ𝑒 is the indentation depth due to elastic deformation (see Figure 4.8a). The total, elastic, and

plastic deformation energies, 𝐸𝑇, 𝐸𝑒 , and 𝐸𝑝, correspond to the areas in a F-h curve[217]–[221] as

shown in Figure 4.8a and are related by the following equation.

Equation 4.17 𝑬𝑻 = ∫ 𝑭𝒅𝒉𝒉𝑻𝟎

=𝒌𝑻𝒉𝑻

𝟑

𝟑= 𝑬𝒆 + 𝑬𝒑 =

𝒌𝒆𝒉𝒆𝟑

𝟑+𝒌𝒑𝒉𝒑

𝟑

𝟑=𝑭𝒎𝒂𝒙

𝟑(𝒉𝒆 + 𝒉𝒑)

The ℎ𝑇

ℎ𝑝 ratio can thus be expressed in terms of the energy ratio as follows[222]–[224]

Equation 4.18 𝒉𝑻

𝒉𝒑= 𝟏 +

𝒉𝒆

𝒉𝒑= 𝟏 +

𝑬𝒆

𝑬𝒑

From the Frenkel’s classical elastic theory,[225], [226] the elastic shear strain of a defect-

free material is written as 𝛾𝑒 =𝑏

2𝜋𝑠, and the corresponding elastic energy can be calculated by the

following equation

Equation 4.19 𝑬𝒆 =𝟏

𝟐𝑮𝑽𝒆𝜸𝒆

𝟐 =𝟏

𝟐𝑮𝑽𝒆 (

𝒃

𝟐𝝅𝒔)𝟐

where 𝑉𝑒 is the volume of elastic deformation during indentation, b the Burgers vector, and s the

interplanar spacing of the slip plane. The obtained b and s from crystal structures are listed in Table

4.1.

Since plastic deformation is mainly due to dislocations, the plastic deformation energy in the

present work is calculated from the average dislocation line energy of 𝐺𝑏2 as follows[227]

Equation 4.20 𝑬𝒑 ≈ 𝝆𝑻𝑽𝒑𝑮𝒃𝟐

where 𝜌𝑇 and 𝑉𝑝 are the total dislocation density generated by the indentation process and the

plastic deformation volume during indentation, respectively. 𝜌𝑇 contains both statistically stored

dislocation density (𝜌𝑆𝑆𝐷) which is due to the multiplication of dislocations by the deformation and

geometrically necessary dislocation density (𝜌𝐺𝑁𝐷) which is due to the required dislocations to

match plastic strain gradient.[194], [228] 𝜌𝑇𝑉𝑝 is the number of dislocations to fill up the plastic

67

deformation volume. Inserting Equation 4.19 and Equation 4.20 into Equation 4.18, we obtain

the ℎ

ℎ𝑝 ratio as follows

Equation 4.21 𝒉𝑻

𝒉𝒑= 𝟏 +

𝟏

𝟖𝝅𝟐𝝆𝑻𝒔𝟐

𝑽𝒆

𝑽𝒑

In the following, the evaluations of quantities in Equation 4.18 are discussed. Based on

the Taylor relation[192]–[194] the relation between stress and dislocation density can be written

as follows

Equation 4.22 𝝉 − 𝝉𝑷𝑵 = 𝒒𝑮𝒃√∆𝝆

where 𝜏𝑃𝑁 is the PN flow stress for dislocation gliding to start, 𝑞 a constant and ∆𝜌 is the

dislocation density difference by indentation process (∆𝜌 = 𝜌 − 𝜌0), where 𝜌0 initial dislocation

density before indentation. 𝜌0 is negligible compare to 𝜌𝑇 when the sample is not work-hardened.

To obtain the saturated hardness without the effect of indentation size due to the strain gradient

(𝜌𝑆𝑆𝐷 >> 𝜌𝐺𝑁𝐷), the applied maximum shear stress (𝜏𝑇) should be equal to or larger than the ideal

shear strength[229] which is the first maximum in the shear stress–strain curve[230].

Let us consider the initiation of ∆𝜌 with ∆𝜌 = 1/𝑚2 when the stress is slightly above the

PN stress, i.e., 𝜏 = (𝑧 + 1)𝜏𝑃𝑁 with z being a very small number. One obtains from Equation 4.22

Equation 4.23 𝒛𝝉𝑷𝑵 = 𝒒𝑮𝒃√𝟏/𝒎𝟐

Since 𝜏𝑇 ≫ 𝜏𝑃𝑁 as shown in Table 4.5 and Figure 4.9, Equation 4.22 can be approximated as

follows

Equation 4.24 𝝉𝑻 ≈ 𝒒𝑮𝒃√𝝆𝑻

which has the similar form of Equation 4.9. Dividing Equation 4.23 by Equation 4.24 and re-

arranging gives the following equation

Equation 4.25 𝝆𝑮𝑵𝑫 ≈𝟏(𝒎−𝟐)𝒃𝟐

𝒛𝟐𝟏

𝒃𝟐(𝝉𝑻

𝑮)𝟐(𝝉𝑷𝑵

𝑮)−𝟐

68

0.5 1.0 1.5 2.010

-6

10-5

10-4

10-3

10-2

10-1

Cu

PN

/G

Carbides & Nitrides

Semiconductors

other cubic ceramics

BCC

FCC

HCP

T/G

Diamond

cBN

BCC

FCC

PN

/G o

r T

/G a

t 0

K

Dislocation width at 0K, w0K

AlNb

Mo

cBNDiamond

Figure 4.9 Peierls-Nabarro flow stress (𝝉𝑷𝑵) and ideal shear stress (𝝉𝑻) at 4.7 K and 7 K in

terms of dislocation width (𝒘𝟎) from Refs.[231]–[234] shown in Table 4.3 and Table 4.5.

Table 4.5 Comparison of 𝝉𝑷𝑵

𝑮 and

𝝉𝑻

𝑮

Materials 𝜏𝑃𝑁𝐺

𝜏𝑇𝐺

C (diamond) 1.97× 10−2 0.17 (96.3GPa/548GPa) [231], [232]

c-BN 1.62 × 10−2 0.17 (70.5GPa/405.4GPa) [231]

Mo (BCC) 2.00 × 10−3 0.13 (15.8GPa/124.2GPa) [233]

Nb (BCC) 1.87 × 10−3 0.17 (6.4GPa/37.5GPa) [233]

Al (FCC) 6.80 × 10−6 0.15 (3.8GPa/25.4GPa) [234]

Cu (FCC) 1.03 × 10−5 0.09 (3.61GPa/40.9GPa) [234] 𝜏𝑃𝑁

𝐺 are calculated from Equation 4.27 or Equation 4.28 with the information from Table 4.3

and Table 4.6.

It can be seen in Table 4.5 and Figure 4.9 that 𝜏𝑇

𝐺 does not change much from diamond to FCC

elements,[231]–[234] while 𝜏𝑃𝑁

𝐺 varies exponentially from 0.02 for diamond to ~10−6 for FCC

elements. Then, Equation 4.25 can be approximated as

69

Equation 4.26 𝝆𝑮𝑵𝑫(𝒎−𝟐) =

𝒆

𝒃𝟐(𝝉𝑷𝑵

𝑮)−𝟐

where 𝑒 (=1(𝑚−2)𝑏2

𝑧2(𝜏𝑇

𝐺)2) is treated as a constant to be evaluated.

The PN flow stresses for edge and screw dislocations are expressed below[15], [235]

Equation 4.27 𝝉𝑷𝑵

𝑮(𝒆𝒅𝒈𝒆) =

𝟏

(𝟏−𝒗)𝐞𝐱𝐩 (

−𝟐𝝅𝒘

𝒃)

Equation 4.28 𝝉𝑷𝑵

𝑮(𝒔𝒄𝒓𝒆𝒘) = 𝐞𝐱𝐩 (

−𝟐𝝅𝒘

𝒃)

where 𝑤 is the dislocation width. Its temperature dependency is approximated in the literature as

follows[236]–[238]

Equation 4.29 𝒘 = 𝒘𝟎 (𝟏 +𝑻

𝟑𝑻𝑴)

where 𝑤0 is the dislocation width at 0 K, equal to 𝑠 for screw dislocations, and 𝑠/(1 − 𝑣) for edge

dislocations with 𝑣 being the Poisson’s ratio at low temperatures,[15] and 𝑇𝑀 the melting

temperature.

The atomic displacement by a single dislocation occurs anisotropically over a few lattice

layers along the Burgers vector direction on the slip plane,[239]–[241] and this deformation is

regarded as the elastic volume by a single dislocation and proportional to the dislocation

width[238], [242], [243]. Consequently, we propose that (𝑉𝑒

𝑉𝑝) ∝

𝑤

𝑏, and by inserting this into

Equation 4.21, we obtained as follows

Equation 4.30 𝒉𝑻

𝒉𝒑= 𝟏 +

𝟖𝟎𝟎𝟎

𝟑

𝒘

𝒃(𝒃

𝒔)𝟐(𝝉𝑷𝑵

𝑮)𝟐

where the constant, 8000

3(≈

1

8𝜋2𝑒

𝑉𝑒

𝑉𝑝

𝑏

𝑤) is evaluated from the measured data of

ℎ𝑇

ℎ𝑝 and the materials

property data in the equation. The calculated value of ℎ𝑇

ℎ𝑝 from Equation 4.30 are compared with

the experimental data in Table 4.4 and Figure 4.10, showing good agreement.

70

1.0 1.5 2.0 2.5 3.0 3.5 4.0

1.0

1.5

2.0

2.5

3.0

3.5

4.0

ZnSe{110}

ZnSe{111}

b-SiCc-BN

BC2N

Si(100)-glide

Calc

. h

T/h

p

Expt. hT/hp

Si(100)-shuffle

Figure 4.10 Comparison of 𝒉𝑻/𝒉𝒑 ratio between experiment from Table 4.4 and the present

model.

4.5.2.2 Parameter c

In the recent hardness modeling by Chen et al.[181], the model parameter 𝑐 in Equation

4.11 was assumed to be unity. Let us examine this assumption from experimental measurements

by re-organizing Equation 4.14 as follows

Equation 4.31 𝒄 =𝑯𝒗

𝑮(𝒉𝑻𝒉𝒑)𝟐

𝒕𝒂𝒏𝟑/𝟐𝜶 𝒄𝒐𝒔 𝜶

Table 4.4 lists the available experimental data and the calculated values of 𝑐 from the

above equation. It can be seen that 𝑐 varies more than an order of magnitude and thus needs to be

better modeled. In Table 4.4, the ratio of shear and bulk moduli, i.e., 𝑘 = 𝐺/𝐵 first used by

Pugh[123] and later used in the hardness modelling by Chen et al.[181], is also included along with

the slip systems and elastic properties.

71

Recently, Cheng et al.[244]–[247] derived an equation for the applied force during

indentation process using a scaling approach and revealed the approximate relationships between

hardness and elastic and plastic properties such as yield stress, elastic modulus, work hardening

exponent and Poisson’s ratio as follows

Equation 4.32 𝑭 = 𝑬𝒉𝟐∏(𝒀

𝑬, 𝝂, 𝒏, 𝜽)

with Y, E, 𝜈 , 𝑛, and 𝜃 being yield stress (flow stress in the present work), Young’s modulus,

Poisson’s ratio, work-hardening exponent, and indentation angle (𝑎 in the present study). It is

understood that 𝑌

𝐸 and 𝑛 are related to materials’ slip systems in terms of the Peierls-Nabarro flow

stress,[15], [235] 𝜈 is related to 𝑘 (=3(1−2𝑣)

2(1+𝑣)), and 𝑎 is fixed to 22° from the indentation tip angle.

Therefore, 𝑐 can be considered as a function of 𝑏

𝑠 and 𝑘.

Starting from the FCC pure metals with the same slip system ½ [110](111), 𝑏

𝑠= √3/2,

and ℎ𝑇

ℎ𝑝= 1, the following correlation is observed from Figure 4.11.

Equation 4.33 𝒄 ∝ 𝒆−𝟐.𝟐𝒌

The exponential relationship in Equation 4.33 is based on the relationship between Poisson’s ratio

and compressive stress found by Walsh[248] since 𝑘 is a function of Poisson’s ratio, and the

compressive stress is assumed to have the similar trend to the shear stress.

72

0.00 0.02 0.04 0.06 0.08 0.100.0

0.2

0.4

0.6

0.8

1.0

Ni

Cu

exp

(-2.2

k)

Expt. c

Al

Ag

Au

Figure 4.11 Exponential relationship of the scaling factor c (from Equation 4.31) for FCC

metals with data from Table 4.4.

To find out the relationship between 𝑏/𝑠 and 𝑐, the elastic shear energies from Equation

4.19 and from Equation 4.17 and Equation 4.11 are compared since the shear area A is mostly

from elastic deformation.

Equation 4.34 𝑬𝒆 =𝟏

𝟐𝑮𝑽𝒆 (

𝒃

𝟐𝝅𝒔)𝟐=𝒉𝒆

𝟑𝑭𝒎𝒂𝒙 =

𝒉𝒆

𝟑

𝟏

𝟐𝒄 𝑮 𝒅𝑻

𝟐𝒕𝒂𝒏𝟑/𝟐𝜶

with 𝑐 derived as

Equation 4.35 𝒄 =𝟑

𝟒𝝅𝟐𝑽𝒆

𝒅𝑻𝟐𝒉𝒆𝒕𝒂𝒏

𝟑/𝟐𝜶(𝒃

𝒔)𝟐

In Figure 4.12a, the values of 𝑐 evaluated from experimental data shown in Table 4.4 are plotted

with respect to (𝑏

𝑠)2, and a parabolic correlation is observed instead. Consequently, the following

relation is obtained, and the comparison between the calculated and experimental values of

parameter 𝑐 is plotted in Figure 4.12.

Equation 4.36 𝒄 = 𝟎. 𝟏𝟔𝟏𝟓 𝒆−𝟐.𝟐𝒌 (𝒃

𝒔)𝟒

73

Inserting Equation 4.30 and Equation 4.36 into Equation 4.14, we obtain the hardness

equation as follows

Equation 4.37 𝑯𝒗 = 𝟎. 𝟏𝟔𝟏𝟓 𝑮 (𝟏 +𝟖𝟎𝟎𝟎

𝟑

𝒘

𝒃(𝒃

𝒔)𝟐(𝝉𝑷𝑵

𝑮)𝟐)𝟐

𝒆−𝟐.𝟐𝒌 (𝒃

𝒔)𝟒𝐭𝐚𝐧𝟑/𝟐 𝒂𝐜𝐨𝐬𝒂

Figure 4.12 Plots of parameter c from experimental data (Equation 4.31) with data from

Table 4.4 (a) with respect to (𝒃/𝒔)𝟐 and (b) with respect to the model (Equation 4.36).

74

Table 4.6 Hardness comparison between the present and previous models with slip systems

and elastic properties with S for screw dislocation.

Compounds struc

ture G B k v

TMelting

(K) Slip system

Hv

calc. Hv exp.

Hv

Chen

[181]

H

Gao

[121

]

H

��im��nek[1

16]

C exp, [249] DC 578 443 1.30 0.05 3823.15 [110]{111} 100.

0 96±5[2

15] 109.7 93.6 95.4

C cal, [249] DC 548.3 465.5 1.18 0.08 3823.15 [110]{111} 100.

5 96±5[2

15] 93.9 93.6 95.4

C exp, [250],a DC 535.5 442.3 1.21 0.07 3823.15 [110]{111} 96.9 96±5[2

15] 95.7 93.6 95.4

C exp, [215] DC 535 443 1.21 0.07 3823.15 [110]{111} 96.9 96±5[2

15] 95.4 93.6 95.4

BC2N cal,

[251], b ORT 446 403 1.11 0.10 3273.15 [110]{111} 78.5 76[252] 76.9 78 71.9

BC2N exp,

[215] ORT 445 408 1.09 0.10 3273.15 [110]{111} 78.7 76[252] 75.4 78 71.9

c-BN exp,

[214],a ZB 405.4 400 1.01 0.12 3246.15 [110]{111} 72.8

66[253]

,

62[252]

,

63±5[2

15]

65.1 64.5 63.2

c-BN cal,

[249] ZB 403 404 1.00 0.13 3246.15 [110]{111} 72.5

66[253]

,

62[252]

,

63±5[2

15]

63.7 64.5 63.2

c-BN cal,

[249] ZB 382 376 1.02 0.12 3246.15 [110]{111} 68.5

66[253]

,

62[252]

,

63±5[2

15]

63.0 64.5 63.2

c-BN cal,

[254] ZB 405 384 1.05 0.11 3246.15 [110]{111} 72.0

66[253]

,

62[252]

,

63±5[2

15]

68.4 64.5 63.2

c-BN exp,

[215] ZB

409

±6

400

±3 1.02 0.12 3246.15 [110]{111} 73.3

66[253]

,

62[252]

,

63±5[2

15]

66.2 64.5 63.2

b-SiC exp,

[211] ZB 192 225 0.85 0.17 3003.15 [110]{111} 34.0

34[253]

,

28±3[2

15]

33.0 30.3 31.1

b-SiC cal,

[249] ZB 196.5 224.9 0.87 0.16 3003.15 [110]{111} 34.7

34[253]

,

28±3[2

15]

34.5 30.3 31.1

b-SiC cal,

[255] ZB 219 223 0.98 0.13 3003.15 [110]{111} 38.2

34[253]

,

28±3[2

15]

42.8 30.3 31.1

b-SiC exp,

[256] ZB 186.5 220.3 0.85 0.17 3003.15 [110]{111} 33.0

34[253]

,

28±3[2

15]

32.1 30.3 31.1

b-SiC exp,

[215] ZB

196

±13

226

±9 0.87 0.16 3003.15 [110]{111} 34.6

34[253]

,

28±3[2

15]

34.1 30.3 31.1

SiO2 exp,

[215]

stish

ovite 220 305 0.72 0.21 1983.15 [110]{111} 32.1

33±2[2

15] 29.0 30.4

75

SiO2 exp,

[215] stishovite

239 340 0.70 0.22 1983.15 [110]{111} 34.9 33±2[2

15] 29.6 30.4

VC cal, [181] NaCl 209.1 305.5 0.68 0.22 3083.15 [110]{111} 37.6 29[116] 26.2 27.2

ZrC exp,

[257],a NaCl 169.7 223.1 0.76 0.20 3805.15 [110]{111} 32.9

27±2[2

15] 26.3

ZrC cal, [258] NaCl 185 228 0.81 0.18 3805.15 [110]{111} 36.0 27±2[2

15] 30.2

ZrC cal, [258] NaCl 185 225 0.82 0.18 3805.15 [110]{111} 36.0 27±2[2

15] 30.7

ZrC exp,

[259],a NaCl 169.6 223.3 0.76 0.20 3805.15 [110]{111} 32.9

27±2[2

15] 26.2

ZrC exp,

[215] NaCl

166

±2 223 0.74 0.20 3805.15 [110]{111} 32.2

27±2[2

15] 25.2

TiC exp,

[257],a NaCl 182.2 242.0 0.75 0.20 3433.15 [110]{111} 34.1

29±3[2

15] 27.1 18.8

TiC cal, [260] NaCl 177 250 0.71 0.21 3433.15 [110]{111} 33.1 29±3[2

15] 24.6 18.8

TiC exp,

[124] NaCl 198.3 286 0.69 0.22 3433.15 [110]{111} 37.1

29±3[2

15] 25.8 18.8

TiC exp,

[261] NaCl 186 200 0.93 0.15 3433.15 [110]{111} 34.6

29±3[2

15] 36.1 18.8

TiC exp,

[215] NaCl

188

±6

241

±1 0.78 0.19 3433.15 [110]{111} 35.2

29±3[2

15] 29.0 18.8

TiN cal, [262]

NaCl 187.1 282.0 0.66 0.23 3203.15 [110]{111} 34.1 23[263] 23.4 18.7

TiN exp,

[264],a NaCl 187.1 318.3 0.59 0.25 3203.15 [110]{111} 34.2 23[263] 19.9 18.7

TiN cal, [265]

NaCl 212.7 294.6 0.72 0.21 3203.15 [110]{111} 38.8 23[263] 28.4 18.7

TiN cal,

[266],a NaCl 207.9 326.3 0.64 0.24 3203.15 [110]{111} 38.0 23[263] 23.8 18.7

RuO2 cal,

[267] Fluor

ite 226.0 343.7 0.66 0.23 1473.15 [110]{111} 28.0

20[124]

, [268],K 26.2 20.6

RuO2 cal,

[269],a

Fluor

ite 130.6 296.3 0.44 0.31 1473.15 [110]{111} 18.4

20[124]

, [268],K 10.3 20.6

RuO2 cal,

[267] Fluor

ite 198.2 267.9 0.74 0.20 1473.15 [110]{111} 23.8

20[124]

, [268],K 28.0 20.6

RuO2 exp,

[268]

Fluor

ite 144* 399 0.36 0.34 1473.15 [110]{111} 21.9

20[124]

, [268],K 8.1 20.6

NbC cal, [270],a

NaCl 171.0 333.3 0.51 0.28 3763.15 [110]{111} 33.1 23±3[2

15] 15.5 18.3

NbC exp,

[259],a NaCl 171.7 340.0 0.51 0.28 3763.15 [110]{111} 33.3

23±3[2

15] 15.2 18.3

AlN cal, [271],a

ZB 138.2 208.7 0.66 0.23 2473.15 [110]{111} 22.7

18[272]

,

12±1[2

15]

19.1 21.7 17.6

AlN cal, [273],a

ZB 130.0 208.0 0.63 0.24 2473.15 [110]{111} 21.5

18[272]

,

12±1[2

15]

16.9 21.7 17.6

AlN exp,

[215]

Not

listed

128

±2

203

±5 0.63 0.24 2473.15 [110]{111} 21.1

18[272]

,

12±1[2

15]

16.9 21.7 17.6

NbN exp,

[274] NaCl 165 292 0.57 0.26 2846.15 [110]{111} 29.1

25[275]

,

20[274]

,

14±1[2

15]

17.3 19.5

NbN exp,

[215] NaCl 156

315

±28 0.50 0.29 2846.15 [110]{111} 27.9

25[275]

,

20[274]

,

14±1[2

15]

13.9 19.5

NbN

cal,[276],MP NaCl 130 305 0.43 0.31 2846.15 [110]{111} 23.8

25[275]

, 9.7 19.5

76

20[274]

,

14±1[2

15]

HfN cal, [249]

NaCl 164.8 278.7 0.59 0.25 3578.15 [110]{111} 31.2

19.5[27

4],

17±2[2

15]

18.4

HfN exp, [274], +

NaCl 202 306 0.66 0.23 3578.15 [110]{111} 38.3

19.5[27

4],

17±2[2

15]

24.5

ZrO2 cal,

[254]

MN

C 88 187 0.47 0.30 2988.15 [110]{111} 16.1 13[253] 8.4 10.8

ZrO2 exp,

[277] MNC

93.6 187 0.50 0.29 2988.15 [110]{111} 17.0 13[253] 9.7 10.8

Si exp, [257],a DC 66.6 97.9 0.68 0.22 1687.15 [110]{111} 8.9 12[253] 11.9 13.6 11.3

Si cal, [278],a DC 65.4 93.4 0.70 0.22 1687.15 [110]{111} 8.7 12[253] 12.2 13.6 11.3

Si cal, [278],a DC 62.5 92.9 0.67 0.23 1687.15 [110]{111} 8.4 12[253] 11.1 13.6 11.3

Si cal, [279],a DC 61.7 97.0 0.64 0.24 1687.15 [110]{111} 8.4 12[253] 10.1 13.6 11.3

Si cal, [279],a DC 61.7 96.3 0.64 0.24 1687.15 [110]{111} 8.4 12[253] 10.2 13.6 11.3

Si exp, [205],a DC 66.3 97.0 0.68 0.22 1687.15 [110]{111} 8.9 12[253] 11.9 13.6 11.3

Si exp, [257],a DC 66.6 97.9 0.68 0.22 1687.15 [110]{111}S 29.1 12[253] 11.9 13.6 11.3

Si cal, [278],a DC 65.4 93.4 0.70 0.22 1687.15 [110]{111}S 27.6 12[253] 12.2 13.6 11.3

Si cal, [278],a DC 62.5 92.9 0.67 0.23 1687.15 [110]{111}S 27.7 12[253] 11.1 13.6 11.3

Si cal, [279],a DC 61.7 97.0 0.64 0.24 1687.15 [110]{111}S 29.1 12[253] 10.1 13.6 11.3

Si cal, [279],a DC 61.7 96.3 0.64 0.24 1687.15 [110]{111}S 28.9 12[253] 10.2 13.6 11.3

GaP exp, [259],a

ZB 55.7 88.2 0.63 0.24 1750.15 [110]{111} 7.7

7.73[28

0],

8.8[281

],

9.5[282

],K

9.3 8.9 8.7

GaP exp,

[257],a ZB 55.8 88.8 0.63 0.24 1750.15 [110]{111} 7.8

7.73[28

0],

8.8[281

],

9.5[282

],K

9.2 8.9 8.7

GaP exp, [205],a

ZB 56.1 88.6 0.63 0.24 1750.15 [110]{111} 7.8

7.73[28

0],

8.8[281

],

9.5[282

],K

9.4 8.9 8.7

GaP cal,

[283],a ZB 61.9 89.7 0.69 0.22 1750.15 [110]{111} 8.4

7.73[28

0],

8.8[281

],

9.5[282

],K

11.5 8.9 8.7

GaP

cal,[276],MP ZB 52 76 0.68 0.22 1750.15 [110]{111} 7.1

7.73[28

0],

8.8[281

],

9.5[282

],K

9.9 8.9 8.7

AlP exp,

[284],a ZB 48.8 86 0.57 0.26 2803.15 [110]{111} 8.5

9.4[282

],K 7.1 9.6 7.9

AlP cal, [283],a

ZB 55.2 90 0.61 0.25 2803.15 [110]{111} 9.6 9.4[282

],K 8.8 9.6 7.9

AlP cal,

[285],a ZB 47.6 89.7 0.53 0.27 2803.15 [110]{111} 8.4

9.4[282

],K 7.0 9.6 7.9

Ge cal, [278] DC 48.4 72.3 0.67 0.23 1211.35 [110]{111} 5.2 8.8[282

],K 9.1 11.7 9.7

77

Ge

cal,[276],MP DC 45 59 0.76 0.20 1211.35 [110]{111} 4.6

8.8[282

],K 10.5 11.7 9.7

Ge exp,

[257],a DC 54.3 74.9 0.72 0.21 1211.35 [110]{111} 5.7

8.8[282

],K 11.2 11.7 9.7

Ge exp, [227],a

DC 54.7 75.2 0.73 0.21 1211.35 [110]{111} 5.7 8.8[282

],K 11.3 11.7 9.7

GaAs exp,

[205],a ZB 46.5 75 0.62 0.24 1511.15 [110]{111} 5.9

6.8[280

],

7[281],

7.5[282

],K

7.8 8 7.4

GaAs exp,

[286],a ZB 46.7 75.5 0.62 0.24 1511.15 [110]{111} 6.0

6.8[280

],

7[281],

7.5[282

],K

7.8 8 7.4

GaAs cal,

[276],MP ZB 41 61 0.67 0.23 1511.15 [110]{111} 5.1

6.8[280

],

7[281],

7.5[282

],K

8.0 8 7.4

GaAs exp,

[257],a ZB 46.7 75.4 0.62 0.24 1511.15 [110]{111} 6.0

6.8[280

],

7[281],

7.5[282

],K

7.8 8 7.4

Y2O3 cal,

[249] CSC 72.4 166.0 0.44 0.31 2698.15 [110]{111} 13.0

7.5[287

],K 6.3 7.7

Y2O3 cal,

[249] CSC 66.6 155.0 0.43 0.31 2698.15 [110]{111} 12.0

7.5[287

],K 5.7 7.7

Y2O3 exp,

[288] CSC

66.3

±0.8

149.5

±1.0 0.44 0.31 2698.15 [110]{111} 11.9

7.5[287

],K 6.0 7.7

InP exp,

[286],a ZB 34.3 71.1 0.48 0.29 1335.15 [110]{111} 4.4

5.5[281

],

5.4[282

],K

3.7 6 5.1

InP exp,

[205],a ZB 34.4 72.5 0.47 0.30 1335.15 [110]{111} 4.5

5.5[281

],

5.4[282

],K

3.6 6 5.1

InP

cal,[276],MP ZB 31 59 0.53 0.28 1335.15 [110]{111} 3.9

5.5[281

],

5.4[282

],K

4.0 6 5.1

AlAs exp,

[289],a ZB 44.8 77.9 0.58 0.26 2013.15 [110]{111} 6.8

5.2[281

],

5[282],K

6.7 8.5 6.8

AlAs exp, [286],a

ZB 44.6 78.3 0.57 0.26 2013.15 [110]{111} 6.8 5.2[281

],

5[282],K

6.5 8.5 6.8

AlAs

cal,[276],MP ZB 39 70 0.56 0.27 2013.15 [110]{111} 6.0

5.2[281

],

5[282],K

5.6 8.5 6.8

GaSb exp,

[286],a ZB 34.2 56.3 0.61 0.25 985.15 [110]{111} 3.3

4.5[281

],

4.4[282

],K

5.8 6 5.6

GaSb exp, [257],a

ZB 34.1 56.4 0.60 0.25 985.15 [110]{111} 3.3

4.5[281

],

4.4[282

],K

5.8 6 5.6

GaSb exp,

[205],a ZB 34.2 56.3 0.61 0.25 985.15 [110]{111} 3.3

4.5[281

],

4.4[282

],K

5.8 6 5.6

GaSb

cal,[276],MP ZB 30 45 0.67 0.23 985.15 [110]{111} 2.7

4.5[281

],

4.4[282

],K

6.1 6 5.6

AlSb cal,

[290],a ZB 31.5 56.1 0.56 0.26 1333.15 [110]{111} 3.8

4.2[281

],

4[282],K

4.7 4.9 4.9

78

AlSb exp, [286],a

ZB 31.9 58.2 0.55 0.27 1333.15 [110]{111} 3.9 4.2[281

],

4[282],K

4.5 4.9 4.9

AlSb exp,

[257],a ZB 32.5 59.3 0.55 0.27 1333.15 [110]{111} 4.0

4.2[281

],

4[282],K

4.6 4.9 4.9

AlSb exp, [205],a

ZB 31.9 58.2 0.55 0.27 1333.15 [110]{111} 3.9 4.2[281

],

4[282],K

4.5 4.9 4.9

AlSb

cal,[276],MP ZB 30 49 0.61 0.25 1333.15 [110]{111} 3.6

4.2[281

],

4[282],K

5.2 4.9 4.9

InAs exp,

[286],a ZB 29.5 57.9 0.51 0.28 1215.15 [110]{111} 3.5

4[281],

3.8[282

],K

3.6 5.7 4.5

InAs exp, [257],a

ZB 29.5 57.9 0.51 0.28 1215.15 [110]{111} 3.5 4[281],

3.8[282

],K

3.6 5.7 4.5

InAs

cal,[276],MP ZB 25 49 0.51 0.28 1215.15 [110]{111} 3.0

4[281],

3.8[282

],K

3.0 5.7 4.5

InSb exp,

[205],a ZB 22.9 46.0 0.50 0.29 800.15 [110]{111} 2.1

3.0[281

],

2.2[282

],K

2.5 4.3 3.6

InSb exp, [286],a

ZB 22.9 46.5 0.49 0.29 800.15 [110]{111} 2.1

3.0[281

],

2.2[282

],K

2.5 4.3 3.6

InSb exp,

[257],a ZB 23.0 46.9 0.49 0.29 800.15 [110]{111} 2.1

3.0[281

],

2.2[282

],K

2.4 4.3 3.6

InSb

cal,[276],MP ZB 19 35 0.54 0.27 800.15 [110]{111} 1.6

3.0[281

],

2.2[282

],K

2.5 4.3 3.6

ZnS exp,

[257],a ZB 32.7 78.4 0.42 0.32 1458.15 [110]{110} 1.6

1.7[282

],K 2.5 2.7

ZnS exp,

[205],a ZB 31.5 77.1 0.41 0.32 1458.15 [110]{110} 1.5

1.7[282

],K 2.3 2.7

ZnS

cal,[276],MP ZB 33 68 0.49 0.29 1458.15 [110]{110} 1.4

1.7[282

],K 3.6 2.7

ZnSe exp,

[257],a ZB 29.4 59.5 0.49 0.29 1798.15 [110]{110} 1.3

1.1[280

],

1.3[282

],K

3.3 2.6

ZnSe exp,

[205],a ZB 28.8 63.1 0.46 0.30 1798.15 [110]{110} 1.3

1.1[280

],

1.3[282

],K

2.7 2.6

ZnSe

cal,[276],MP ZB 28 58 0.48 0.29 1798.15 [110]{110} 1.2

1.1[280

],

1.3[282

],K

3.0 2.6

ZnTe exp,

[205],a ZB 23.4 51.0 0.46 0.30 1511.15 [110]{110} 1.0

0.9[282

],K 2.1 2.3

ZnTe exp,

[257],a ZB 23.4 51.0 0.46 0.30 1511.15 [110]{110} 1.0

0.9[282

],K 2.1 2.3

ZnTe

cal,[276],MP ZB 22 46 0.48 0.29 1511.15 [110]{110} 0.9

0.9[282

],K 2.1 2.3

MgAl2O4 spine

l 96 180 0.53 0.27 2408.15 [110]{111} 16.0

13.4[21

0]

MgO

pero

vskit

e

119 151 0.79 0.19 3125.15 [110]{110} 3.4 5.95[21

0]

V exp, [205],a BCC 47.5 156.7 0.30 0.36 2183.15 [111]{110}S 0.92 0.63[29

1] 1.7

V cal,[42],a BCC 30.4 182.9 0.17 0.42 2183.15 [111]{110}S 0.77 0.63[29

1] -1.2

Fe exp,

[205],a BCC 81.5 166.7 0.49 0.29 1811.15 [111]{110}S 1.02

1.13[21

0] 8.4

79

Fe cal,[42],a BCC 80.7 189.3 0.43 0.31 1811.15 [111]{110}S 1.15 1.13[21

0] 6.6

Cr exp, [205],a

BCC 114.6 160.7 0.71 0.21 2180.15 [111]{110}S 0.95

1.06

[291],

1.42

[291],M

18.6

Cr cal,[42],a BCC 131.7 190.1 0.69 0.22 2180.15 [111]{110}S 1.13

1.06

[291],

1.42

[291],M

19.6

Mo exp,

[205],a BCC 124.2 263.7 0.47 0.30 2896.15 [111]{110}S 1.81

1.53[29

1],

1.66-

2.02

[291],M

10.9

Mo cal,[42],a BCC 117.5 260.4 0.45 0.30 2896.15 [111]{110}S 1.79

1.53[29

1],

1.66-

2.02

[291],M

9.8

Nb exp,

[205],a BCC 37.5 169.7 0.22 0.40 2742.15 [111]{211}S 0.91

1.32

[291] -0.2

Nb cal,[42],a BCC 25.0 172.3 0.15 0.43 2742.15 [111]{211}S 0.70 1.32

[291] -1.6

Nb exp,

[205],a BCC 37.5 169.7 0.22 0.40 2742.15 [111]{211}S 1.77

1.32

[291] -0.2

Nb cal,[42],a BCC 25.0 172.3 0.15 0.43 2742.15 [111]{211}S 1.37 1.32

[291] -1.6

Eu exp,[257] BCC 7.9 8.3 0.95 0.14 1099.15 [111]{211}S 0.06 0.17

[291] 3.3

Eu cal,[42],a BCC 9.5 13.1 0.73 0.21 1099.15 [111]{211}S 0.12 0.17

[291] 2.1

W exp,[205],a BCC 160 309.7 0.52 0.28 3695.15 [111]{110}S 2.21 3.43[29

1] 15.0

W exp,[205],a BCC 160 309.7 0.52 0.28 3695.15 [111]{211}S 4.44 3.43[29

1] 15.0

W cal,[42],a BCC 146.6 302.2 0.49 0.29 3695.15 [111]{110}S 2.17 3.43[29

1] 12.9

W cal,[42],a BCC 146.6 302.2 0.49 0.29 3695.15 [111]{211}S 4.34 3.43[29

1] 12.9

Al exp,[205],a FCC 26.0 77.3 0.34 0.35 933.45 [110]{111} 0.18 0.17

[291] 0.8

Al cal,[42],a FCC 23.1 74.3 0.31 0.36 933.45 [110]{111} 0.17 0.17

[291] 0.2

Ni exp, [205],a FCC 83.2 184.3 0.45 0.30 1728.15 [110]{111} 0.53 0.64

[291] 7.5

Ni cal, [42],a FCC 92.3 195.6 0.47 0.30 1728.15 [110]{111} 0.57 0.64

[291] 8.7

Cu exp,[205],a FCC 47.3 137.7 0.34 0.35 1358.15 [110]{111} 0.36 0.37

[291] 2.5

Cu cal, [42],a FCC 49.6 137.5 0.36 0.34 1358.15 [110]{111} 0.37 0.37

[291] 3.0

Pd exp,[205],a FCC 46.7 187.7 0.25 0.39 1828.15 [110]{111} 0.47 0.46

[291] 0.7

Pd cal, [42],a FCC 43.3 163.7 0.26 0.38 1828.15 [110]{111} 0.42 0.46

[291] 0.8

Ag

exp,[205],a FCC 29.2 102 0.29 0.37 1234.95 [110]{111} 0.25

0.25

[291] 0.3

Ag cal, [42],a FCC 28.1 91.3 0.31 0.36 1234.95 [110]{111} 0.23 0.25

[291] 0.5

Pt exp,[205],a FCC 63.5 283 0.22 0.40 2041.15 [110]{111} 0.69 0.55

[291] 0.9

Pt cal, [42],a FCC 43.9 243.4 0.18 0.41 2041.15 [110]{111} 0.53 0.55

[291] -0.5

Au

exp,[205],a FCC 27.5 171.7 0.16 0.42 1337.15 [110]{111} 0.31

0.22

[291] -1.4

Au cal, [42],a FCC 19.2 137.6 0.14 0.43 1337.15 [110]{111} 0.23 0.22

[291] -1.9

Pb exp,[205],a FCC 8.5 43.9 0.19 0.41 600.65 [110]{111} 0.07 0.03

[291],M -2.0

Pb cal, [42],a FCC 15.7 40.6 0.39 0.33 600.65 [110]{111} 0.08 0.03

[291],M 0.3

Ir exp,[205],a FCC 224.3 373.3 0.60 0.25 2720.15 [110]{111} 1.12 1.76

[291] 23.1

80

Ir exp,[205],a FCC 224.3 373.3 0.60 0.25 2720.15 [110]{110}S 3.60 1.76

[291] 23.1

Ir cal, [42],a FCC 216.3 342.8 0.63 0.24 2720.15 [110]{111} 1.01 1.76

[291] 24.1

Ir cal, [42],a FCC 216.3 342.8 0.63 0.24 2720.15 [110]{110}S 3.26 1.76

[291] 24.1

Rh exp,[205],a FCC 149.4 267 0.56 0.26 2236.15 [110]{111} 0.79 1.25

[291] 16.0

Rh exp,[205],a FCC 149.4 267 0.56 0.26 2236.15 [110]{110}S 2.49 1.25

[291] 16.0

Rh cal, [42],a FCC 146.0 253.4 0.58 0.26 2236.15 [110]{111} 0.75 1.25

[291] 16.4

Rh cal, [42],a FCC 146.0 253.4 0.58 0.26 2236.15 [110]{110}S 2.36 1.25

[291] 16.4

Th exp,[205],a FCC 28.6 57.7 0.50 0.29 2028.15 [110]{111} 0.17 0.29

[291] 3.3

Th exp,[205],a FCC 28.6 57.7 0.50 0.29 2028.15 [110]{110}S 0.53 0.29

[291] 3.3

Th cal, [42],a FCC 39.0 56.6 0.69 0.22 2028.15 [110]{111} 0.15 0.29

[291] 8.0

Th cal, [42],a FCC 39.0 56.6 0.69 0.22 2028.15 [110]{110}S 0.48 0.29

[291] 8.0

Be exp, [292] HCP 150.3 111.9 1.34 0.04 1560.15 [1120] {0001} 0.71

1.67

[291]

~0.95[1

47]

0.74[14

7]

50.0

Be exp, [292] HCP 150.3 111.9 1.34 0.04 1560.15 [1120]{1010} 0.43

1.67

[291]

~0.95[1

47]

0.74[14

7]

49.8

Be exp, [292] HCP 150.3 111.9 1.34 0.04 1560.15 [1011]{1012}

T 1.74

1.67

[291]

~0.95[1

47]

0.74[14

7]

50.0

Be cal, [42] HCP 158.1 121.1 1.31 0.05 1560.15 [1120] {0001} 0.80

1.67

[291]

~0.95[1

47]

0.74[14

7]

49.8

Be cal, [42] HCP 158.1 121.1 1.31 0.05 1560.15 [1120]{1010} 0.50

1.67

[291]

~0.95[1

47]

0.74[14

7]

50.0

Be cal, [42] HCP 158.1 121.1 1.31 0.05 1560.15 [1011]{1012}

T 1.94

1.67

[291]

~0.95[1

47]

0.74[14

7]

49.8

Mg exp, [292] HCP 17.3 35.3 0.49 0.29 923.15 [1120] {0001} 0.35 0.31

[291],M 4.3

Mg exp, [292] HCP 17.3 35.3 0.49 0.29 923.15 [1120]{1010} 0.27 0.31

[291],M 4.3

Mg exp, [292] HCP 17.3 35.3 0.49 0.29 923.15 [1120]{1011} 0.20 0.31

[291],M 4.3

Mg cal, [42] HCP 15.9 35.7 0.45 0.31 923.15 [1120] {0001} 0.35 0.31

[291],M 3.7

Mg cal, [42] HCP 15.9 35.7 0.45 0.31 923.15 [1120]{1010} 0.27 0.31

[291],M 3.7

Mg cal, [42] HCP 15.9 35.7 0.45 0.31 923.15 [1120]{1011} 0.21 0.31

[291],M 3.7

Cd exp, [292] HCP 21.7 62.3 0.35 0.34 594.25 [1120] {0001} 0.26 0.29

[291],M 3.3

Cd cal, [42] HCP 16.5 35.8 0.46 0.30 594.25 [1120] {0001} 0.16 0.29

[291],M 3.9

Zn exp, [292] HCP 36.1 63.9 0.56 0.26 692.65 [1120] {0001} 0.32 0.20[29

1],M 7.9

81

Zn cal, [42] HCP 30 51.8 0.58 0.26 692.65 [1120] {0001} 0.25 0.20[29

1],M 7.3

Ti exp, [292] HCP 42.7 108.9 0.39 0.33 1941.15 [1120]{1010} 1.01 0.97

[291] 5.7

Ti cal, [42] HCP 44.3 112.8 0.39 0.33 1941.15 [1120]{1010} 1.05 0.97

[291] 5.8

Zr exp, [292] HCP 36.2 98.4 0.37 0.34 2128.15 [1120]{1010} 0.92 0.90

[291] 4.8

Zr cal, [42] HCP 33.1 95.3 0.35 0.34 2128.15 [1120]{1010} 0.88 0.90

[291] 4.2

Y exp, [292] HCP 25.3 41.2 0.61 0.25 1799.15 [1120] {0001} 0.54 1.08[29

1],M 7.1

Y exp, [292] HCP 25.3 41.2 0.61 0.25 1799.15 [1120]{1010} 0.36 1.08[29

1],M 7.1

Y exp, [292] HCP 25.3 41.2 0.61 0.25 1799.15 [1120]{1011} 0.28 1.08[29

1],M 7.1

Y cal, [42] HCP 27.1 40.8 0.66 0.23 1799.15 [1120] {0001} 0.52 1.08[29

1],M 8.1

Y cal, [42] HCP 27.1 40.8 0.66 0.23 1799.15 [1120]{1010} 0.35 1.08[29

1],M 8.1

Y cal, [42] HCP 27.1 40.8 0.66 0.23 1799.15 [1120]{1011} 0.27 1.08[29

1],M 8.1

Sc exp, [292] HCP 30.4 56.2 0.54 0.27 1814.15 [1120] {0001} 0.72 1.02

[291],M 6.8

Sc exp, [292] HCP 30.4 56.2 0.54 0.27 1814.15 [1120]{1010} 0.51 1.02

[291],M 6.8

Sc cal, [42] HCP 33.8 54.9 0.62 0.24 1814.15 [1120] {0001} 0.68 1.02

[291],M 8.4

Sc cal, [42] HCP 33.8 54.9 0.62 0.24 1814.15 [1120]{1010} 0.48 1.02

[291],M 8.4

Gd exp, [292] HCP 21.6 37.9 0.57 0.26 1585.15 [1120] {0001} 0.47 0.56

[291] 5.9

Gd exp, [292] HCP 21.6 37.9 0.57 0.26 1585.15 [1120]{1010} 0.33 0.56

[291] 5.9

Gd cal, [42] HCP 22 38.5 0.57 0.26 1585.15 [1120] {0001} 0.48 0.56

[291] 6.0

Gd cal, [42] HCP 22 38.5 0.57 0.26 1585.15 [1120]{1010} 0.34 0.56

[291] 6.0

Tb exp, [292] HCP 22.4 38.9 0.58 0.26 1629.15 [1120] {0001} 0.50 0.45

[291] 6.1

Tb exp, [292] HCP 22.4 38.9 0.58 0.26 1629.15 [1120]{1010} 0.34 0.45

[291] 6.1

Tb cal, [42] HCP 23.8 39.4 0.60 0.25 1629.15 [1120] {0001} 0.50 0.45

[291] 6.7

Tb cal, [42] HCP 23.8 39.4 0.60 0.25 1629.15 [1120]{1010} 0.34 0.45

[291] 6.7

Dy exp, [292] HCP 25 40.6 0.62 0.24 1685.15 [1120] {0001} 0.52 0.41

[291] 7.0

Dy exp, [292] HCP 25 40.6 0.62 0.24 1685.15 [1120]{1010} 0.35 0.41

[291] 7.0

Dy cal, [42] HCP 24.6 40.7 0.60 0.25 1685.15 [1120] {0001} 0.53 0.41

[291] 6.8

Dy cal, [42] HCP 24.6 40.7 0.60 0.25 1685.15 [1120]{1010} 0.35 0.41

[291] 6.8

Ho exp, [292] HCP 26.4 40.9 0.65 0.23 1747.15 [1120] {0001} 0.53 0.41

[291] 7.7

Ho exp, [292] HCP 26.4 40.9 0.65 0.23 1747.15 [1120]{1010} 0.35 0.41

[291] 7.7

Ho cal, [42] HCP 27.7 42.5 0.65 0.23 1747.15 [1120] {0001} 0.55 0.41

[291] 8.0

Ho cal, [42] HCP 27.7 42.5 0.65 0.23 1747.15 [1120]{1010} 0.36 0.41

[291] 8.0

Er exp, [292] HCP 28.2 44.8 0.63 0.24 1802.15 [1120] {0001} 0.66 0.43

[291] 7.7

Er exp, [292] HCP 28.2 44.8 0.63 0.24 1802.15 [1120]{1010} 0.39 0.43

[291] 7.7

Er cal, [42] HCP 30.3 44.4 0.68 0.22 1802.15 [1120] {0001} 0.56 0.43

[291] 8.9

Er cal, [42] HCP 30.3 44.4 0.68 0.22 1802.15 [1120]{1010} 0.37 0.43

[291] 8.9

Hf exp, [292] HCP 55.6 110 0.51 0.28 2504.15 [1120] {0001} 1.55 1.76

[291] 8.9

Hf exp, [292] HCP 55.6 110 0.51 0.28 2504.15 [1120]{1010} 1.07 1.76

[291] 8.9

82

Hf exp, [292] HCP 55.6 110 0.51 0.28 2504.15 [1120]{1011} 0.83 1.76

[291] 8.9

Hf cal, [42] HCP 55 109.1 0.50 0.28 2504.15 [1120] {0001} 1.55 1.76

[291] 8.8

Hf cal, [42] HCP 55 109.1 0.50 0.28 2504.15 [1120]{1010} 1.06 1.76

[291] 8.8

Hf cal, [42] HCP 55 109.1 0.50 0.28 2504.15 [1120]{1011} 0.83 1.76

[291] 8.8

Co exp, [292] HCP 75.4 193.7 0.39 0.33 1768.15 [1120] {0001} 2.29 1.40

[291] 7.8

Co exp, [292] HCP 75.4 193.7 0.39 0.33 1768.15 [1120]{1010} 1.76 1.40

[291] 7.8

Co exp, [292] HCP 75.4 193.7 0.39 0.33 1768.15 [1120]{1011} 1.34 1.40

[291] 7.8

Co cal, [42] HCP 103.5 212.5 0.49 0.29 1768.15 [1120] {0001} 2.54 1.40

[291] 12.3

Co cal, [42] HCP 103.5 212.5 0.49 0.29 1768.15 [1120]{1010} 1.95 1.40

[291] 12.3

Co cal, [42] HCP 103.5 212.5 0.49 0.29 1768.15 [1120]{1011} 1.48 1.40

[291] 12.3

Re exp, [292] HCP 177.1 368.6 0.48 0.29 3455.15 [1120] {0001} 5.00 2.45

[291] 16.5

Re exp, [292] HCP 177.1 368.6 0.48 0.29 3455.15 [1120]{1010} 3.77 2.45

[291] 16.5

Re exp, [292] HCP 177.1 368.6 0.48 0.29 3455.15 [1120]{1011} 2.88 2.45

[291] 16.5

Re cal, [42] HCP 166.3 366.8 0.45 0.30 3455.15 [1120] {0001} 5.00 2.45

[291] 14.9

Re cal, [42] HCP 166.3 366.8 0.45 0.30 3455.15 [1120]{1010} 3.75 2.45

[291] 14.9

Re cal, [42] HCP 166.3 366.8 0.45 0.30 3455.15 [1120]{1011} 2.87 2.45

[291] 14.9

S and T denote screw dislocation and deformation twin a shear modulus acquired from VRH method of Cij results at 25 C b shear modulus acquired from Voigt method since they used the same method for bulk modulus

[205] experimental values from Ref. [205] is carefully treated after comparing with 2 other

experimental elastic constants sources[227], [257]. MP the data from materials project

* used C44 instead of shear modulus + the elastic constants determined from neutron data have limited accuracy (10–15%). K, M Knoop and microhardness data is used since Vickers hardness is not available.

83

0 100 200 300 400 500 600

0

100

200

300

400

500

600

Ca

lcula

ted

Sh

ea

r M

od

ulu

s, G

ca

l (G

Pa

)

Experimental Shear Modulus, Gexp

(GPa)

Figure 4.13 Comparison of experimental and calculated (VRH averaged) shear moduli with

elastic stiffness constant data from Shang et al.[42].

84

0 100 200 300 400 500

0

100

200

300

400

500

Experimental Bulk Modulus, Bexp

(GPa)

C

alc

ula

ted

Bu

lk M

od

ulu

s, B

ca

l (G

Pa

)

Figure 4.14 Comparison of experimental and calculated (VRH averaged) bulk moduli with

elastic stiffness constant data from Shang et al.[42].

85

Chapter 5

Temperature Dependent Hardness Model:

the Study of Thermally Activated Dislocation Width

5.1 Introduction

Indentation hardness as a fingerprint of materials’ deformation behavior has been widely

used in designing materials for abrasives and wear resistant coatings[293] due to its simplicity and

low cost. As a variable which influences significantly on the hardness, temperature plays a

significant role in determining strength properties of materials, especially for high-temperature

applications[294] since materials become soft at high temperature that was observed as a reduction

in the material’s ability to resist indentation hardness due to the increase in irreversible plastic

deformation. The irreversible plastic deformation is originated from temperature-dependent flow

stress, or fundamentally dislocation motion as a function of temperature[159]. It is supported by

the observation that the trend of hardness with temperature is similar to that of a number of

mechanical properties such as flow stress[295]–[297]. Thus, understanding of the temperature

dependent flow stress is crucial for understanding the hardness behavior as a function of

temperature.

The thermally activated dislocation width in the temperature dependent Peierls-Nabarro

(PN) flow stress changes with the temperature[236]–[238] and this dislocation width term should

capture the diffusion mechanisms at each temperature region since the indentation hardness test at

elevated temperature performs with already thermally activated materials. To be specific, the major

deformation mechanism of hardness at high temperature is the self-diffusion by the dislocation

climb since Sherby and Armstrong predicted the self-diffusion activation energy for creep

deformation and self-diffusion from hot hardness test[298]. While that at low temperature is the

86

dislocation (pipe) diffusion by the dislocation glides which is proved by a number of

researchers[298]–[302].

Without such considerations, the previous temperature dependent flow stress and hardness

models limit to the specific temperature ranges and does not predict the transition temperatures.

For example, the temperature dependent flow stress models is only valid at low temperature

ranges[236]–[238] since they did not include diffusion mechanisms into their models. Furthermore,

the temperature dependent hardness model, first proposed by Ito[303] and Shishokin[304], and has

been explored subsequently by a number of investigators[298], [299], [305], [306]. Although this

model provides fundamental understandings of the relationship that is Arrhenius-type diffusion

mechanisms, 𝑯(𝑻) = 𝑨𝒊 𝐞𝐱𝐩 (𝑩𝒊

𝑻), the pre-exponential (𝑨𝒊) and softening (𝑩𝒊) coefficients does

not provide the physical meaning of thermally activated dislocation motion since 𝑩𝒊 coefficient is

dependent on the power law exponent which is another parameter to be modelled, it does not predict

the transition temperature where the slope changes due to the change of mechanism, and it does not

capture other deformation mechanisms such as slip system change as a function of

temperature[136], [137], [307], [308].

In this work, the temperature-dependent dislocation width term in Peierls-Nabarro (PN)

flow stress which includes not only dislocation diffusion and self-diffusion but also slip system

changes and the changes of diffusion species, is modelled based on our finding which is related to

materials’ deformation diffusion behavior. This modeling also includes the relationship between

melting temperature and slip systems, and the activation energies. Then, the temperature dependent

hardness model is proposed from the modelled temperature dependent PN flow stress and is

developed from the previous Vickers hardness model from Chapter 4.

87

5.2 Results and Discussion

In order to model the thermally activated dislocation width, the mono-vacancy (self-

)diffusion activation energy was modeled with its melting temperature and the dislocation climb

based structural factor. This activation energy model extends the previous belief of the linear

relationship between self-diffusion activation energy and melting temperature, also known as Van

Liempt rule[309] (𝑄 = 17𝑅𝑇𝑀), since the rule is valid only within the same crystal structure[310]–

[314]. The dislocation climb based structural factor describes the atomic jumps to the nearest

neighbor in diffusion process and it can be considered as the slip system for dislocation climb.

Modeled mono-vacancy diffusion activation energy in this study agrees well with mono-vacancy

diffusion of 62 materials’ experimental data and that of 39 materials’ first-principles calculation

results as shown in Figure 5.1. This model also predicts the activation energies of both the diffusion

species in binary compounds, such as Ti and C vacancy diffusion in TiC. The detailed derivation

and a table for all the data are in 5.3 Modeling Procedure.

0 1 2 3 4 5 6 7 80

1

2

3

4

5

6

7

8

M

od

ell

ed

Ac

tiv

ati

on

En

erg

y (

eV

)

Expt. Activation Energy (eV)

Expt. BCC

Expt. FCC

Expt. HCP

Expt. Me in Carbides

Expt. C in Carbides

Expt. Alkali Halides

Expt. Semiconductors

Calc. BCC

Calc. FCC

Calc. HCP

88

Figure 5.1 Activation energy for self-diffusion modeling. All the data and references are in

Table 5.3.

Our temperature-dependent hardness model is derived based on the thermally activated

dislocation width and self-diffusion activation energy model as an input since the hardness is

dependent on the plastic deformation as well as elastic deformation. This thermally activated

dislocation width includes various deformation mechanisms that affect the dislocation width, such

as the change of diffusion mechanisms, diffusion species and the active slip system, and also phase

transformations at high temperature. We show here that the prediction of the temperature dependent

hardness quantitatively and they agree well with the following experimental data of 16 materials

including the temperature dependent hardness for FCC, BCC, HCP metals and rocksalt and

zincblende ceramics as shown in Figure 5.2 to Figure 5.5. Each deformation mechanism will be

discussed the details with the examples of the temperature dependent hardness and full derivations

are in 5.3 Modeling Procedure.

5.2.1 Change of Diffusion Mechanism

First, diffusion mechanisms directly affect the temperature dependent dislocation width,

and this influences the hardness of materials. As shown in Figure 5.2a-f, this temperature

dependent hardness model of FCC metals are in good agreement with experimental hardness results

from Lozinskii[315], especially the slope change at the critical temperature around 0.5Tm. The

critical temperature is the temperature where the dislocation diffusion interaction that are dominant

at low temperature is equal to the di-vacancy (2V) diffusion which is dominant at high temperature

based on this model. Furthermore, it is also in agreement with previous research[316] that the other

critical temperature where the mono-vacancy diffusion is equal to the di-vacancy diffusion is

89

around 0.4Tm as shown in Figure 5.2a-f. Since this study assumed 𝑄𝑑 = 0.65𝑄1𝑉 and the entropy

contribution of di-vacancy diffusion is fixed as other metals, the hardness of FCC Pd is slightly off

from the experimental result, and this can be addressed that 𝑄𝑑 is probably higher than 0.65𝑄1𝑉

and the entropy contribution is smaller than that of other FCC metals. For the accurate modelling

of temperature dependent hardness of FCC Pd, the experimental or calculated 𝑄𝑑 and entropy

contribution of di-vacancy should be required. During the indentation deformation of FCC metals,

therefore, the major deformation diffusion mechanisms are the dislocation diffusion at low

temperature and di-vacancy (2V) diffusion at higher temperature, and the mono-vacancy diffusion

never become a major mechanism for FCC metals.

90

Figure 5.2 Predicted temperature dependent hardness of FCC metals. All the experimental

data is from Lozinskii[315].

91

5.2.2 Change of the Active Slip System

Second, the temperature dependent dislocation width is also affected by the active slip

systems of materials. As it is well described in the previous studies[134], [135] and the previous

hardness model from Chapter 4, FCC Rh and Ir undergo the active slip system mixture between

½ [110]{111} edge and ½ [110]{110} screw dislocations during deformation at room temperature

due to the unsaturated d-bonds[131]. As shown in Figure 5.3a-b, the portions of screw dislocations

become larger at higher temperature. Over 0.5Tm, the hardness trend of FCC Rh tends to follow the

di-vacancy diffusion model trend for ½ [110]{111} edge dislocations, although it is difficult to

judge due to the lack of experiment data.

Furthermore, BCC metals are typical examples of slip change as a function of temperature.

Figure 5.3c-d shows temperature dependent hardness of BCC Mo and BCC W. This hardness

model well described the positive slope change of BCC Mo and BCC Was a function of temperature

due to the slip system changes. The slip system of BCC Mo is ½ [111]{110} screw dislocations at

room temperature, while the slip system become the mixture between ½ [111]{110} and

½ [111]{211} screw dislocations at 77 K[136], [137], [307], [308]. The slip system of BCC W at

room temperature is the mixture between ½ [111]{110} and ½ [111]{211} screw dislocations[136],

[142], [317]. As the temperature increases, the slip systems of BCC Mo and BCC W become

½ [111]{110} screw dislocations and the hardness change with temperature becomes moderate.

92

Figure 5.3 Predicted temperature dependent hardness of FCC Rh (a) and Ir (b), and BCC

Mo (c) and W (d) metals. All the experimental data is from Lozinskii[315](■) and Stephens

et al.[318](▲).

5.2.3 Phase Transformation at Finite Temperature

Third, the temperature dependent dislocation width is affected by phase transformations.

Figure 5.4 shows temperature dependent hardness of HCP Co, Ti and Zr. HCP Co, Ti and Zr

undergo phase transformations to FCC or BCC structures at high temperature. When a phase

transforms, it influences the hardness as shown in Figure 5.4. This hardness model well describes

93

the hardness changes due to transformation to other structures as a function of temperature. To be

specific, the temperature dependent hardness of HCP Co follows the pyramidal slip path, and then

HCP Co transforms to FCC phase at 0.42 Tm. The hardness of FCC Co between 0.42 and 0.6 Tm in

Figure 5.4a is due to mono-vacancy diffusion, while above 0.6 Tm, di-vacancy diffusion dominates

the hardness as other FCC metals does. The hardness trends of HCP Ti and Zr follow the prismatic

slip path with mono-vacancy and di-vacancy diffusion at low to intermediate temperature ranges.

After phase transformation to BCC structure, the hardness trends follow that of BCC metals.

Interestingly, this model also capture the abnormally fast diffusion behavior of BCC Ti above 0.54

Tm, although the diffusion mechanism of BCC Ti is not yet fully understood[319], [320]. This

model predicts that the activation energy for this abnormal diffusion behavior of BCC Ti is likely

that of the di-vacancy diffusion which is not usual in BCC metals[316]. The low shear modulus of

BCC Ti[321] as listed in Table 5.1, also play a role in the hardness drop. Furthermore, this model

can predict the hardness of metastable phases such as FCC Co, BCC Ti and BCC Zr as shown in

Figure 5.4. For example, BCC Ti phase after quenching to room temperature, the hardness of the

metastable BCC Ti at room temperature (1.12 GPa) is slightly higher than that of HCP Ti (0.90

GPa). It is worth to mention that the hardness trends of HCP metal are rather continuously change

the slope not like that of FCC metals due to the c/a ratio change as a function of temperature which

affects the thermally activated dislocation width. Since it is not considered in Figure 5.4, the

predictions do not fit well.

94

Figure 5.4 Predicted temperature dependent hardness of HCP metals. All the experimental

data is from Lozinskii[315].

95

Table 5.1 The materials’ information used in Figure 5.2 to Figure 5.5.

Elements Structure G(GPa) B(GPa) k 𝜈 𝑇𝑀(K)

[322] Considered Slip system

Au [42],a FCC 19.2 137.6 0.14 0.43 1337.15 [110]{111}

Ag [42],a FCC 28.1 91.3 0.31 0.36 1234.95 [110]{111}

Al[42],a FCC 23.1 74.3 0.31 0.36 933.45 [110]{111}

Ni [42],a FCC 92.3 195.6 0.47 0.30 1728.15 [110]{111}

Pd [42],a FCC 43.3 163.7 0.26 0.38 1828.15 [110]{111}

Pt [42],a FCC 43.9 243.4 0.18 0.41 2041.15 [110]{111}

Rh [42],a FCC 146 253.4 0.58 0.26 2236.15 [110]{111}

[110]{110}S

Ir [42],a FCC 216.3 342.8 0.63 0.24 2720.15 [110]{111}

[110]{110}S

Mo[42],a BCC 117.5 260.4 0.45 0.30 2896.15 [111]{110}S

[111]{211}S

W[42],a BCC 146.6 302.2 0.49 0.29 3695.15 [111]{110}S

[111]{211}S

Co [42] HCP 103.5 212.5 0.49 0.29 1768.15

Basal

Prismatic

Pyramidal

Co FCC 102.0 212.0 0.48 0.29 1768.15 [110]{111}

Ti [42] HCP 44.3 112.8 0.39 0.33 1941.15

Basal

Prismatic

Pyramidal

Ti BCC 20.0 87.7 0.23 0.39 1941.15 [111]{110}S

[111]{211}S

Zr [42] HCP 33.1 95.3 0.35 0.34 2128.15

Basal

Prismatic

Pyramidal

Zr BCC 6.0 89 0.07 0.47 2128.15 [111]{110}S

[111]{211}S

TiC [260] Rocksalt 176.9 250.3 0.71 0.19 3433.15 [110]{111}

Si[278],a Diamond

Cubic 62.5 92.9 0.67 0.23 1687.15

[110]{111}

[110]{111}S

Ge[278] Diamond

Cubic 48.4 72.3 0.67 0.23 1211.35

[110]{111}

[110]{111}S

SiC[249] Zinc

blende 196.5 224.9 0.87 0.16 3003.15

[110]{111}

[110]{111}S a shear modulus acquired from VRH method of Cij results

5.2.4 Change of the Diffusion Species

Fourth, the temperature dependent dislocation width is affected by the diffusion

mechanisms of different species in binary compounds. For titanium carbide (TiC), the flow stress

96

as a function of temperature is first checked in Figure 5.5a in order to make sure the reliability of

this model. As shown in Figure 5.5a, the slope of the flow stress both from experiments by

Kurishita et al.[323] and the carbon vacancy diffusion model (above 0.33Tm) shows matches each

other. The difference of y-intercept is due to the difference of strain rate since the indentation

hardness such as Vickers hardness uses very slow strain rate. For the temperature dependent

hardness of TiC as shown in Figure 5.5b, this hardness model agrees well with previous

experiments[301], [305], [324]–[328] especially the trend from Kumashiro et al.[324] (Expt.1) and

Kohlstedt et al.[326] (Expt.7) which capture the slope change at the critical temperature (0.33 Tm).

Interestingly, it is mentioned previous studies[301], [324] that the hardness of TiC falls rapidly

between 0.2 and 0.4 Tm. The critical temperature where the slope changes is predicted to be at 0.33

Tm from this model. This critical temperature is due to the change of diffusion mechanism and

diffusion species. The dislocation diffusion by Ti atom is the major deformation mechanism at low

temperature up to 0.33Tm, while above 0.33Tm, the carbon mono-vacancy diffusion is the major

mechanism[323], [324]. To be specific, the major deformation mechanism at low temperature is

the dislocation diffusion by dislocation glide of Ti atom on {111} plane[329]. Although carbon

atoms are easy to diffuse, Ti atom should diffuse for the TiC structure to physically deform. At

high temperature, the dislocation movement by Ti atom can climb to the carbon vacancy site[298]

which is corresponds to mono-vacancy diffusion of carbon atom in TiC[323], [324] since the

carbon vacancies are located next to Ti atoms and the number of carbon vacancy sites are more

than that of Ti vacancy sites. This mechanism at high temperature is confirmed that the activation

energy for indentation creep was closely equivalent to that obtained for the self-diffusion of carbon

in TiC from the temperature dependency of indentation behavior above the critical temperature

(0.34-0.43 Tm) by Kumashiro et al.[324]. Kurishita et al.[323] also point out the deformation

mechanism at low temperature is due to the PN stress while that at high temperature is due to the

self-diffusion of carbon in TiC.

97

Figure 5.5 (a) Predicted temperature dependent flow stress of TiC comparison with

experiment results from Kurishita et al.[323] and (b) Predicted temperature dependent

hardness of TiC comparison with single crystal micro-Vickers hardness (■, Expt.1) from

Kumashiro et al.[324], single crystals of Vickers hardness (●, Expt.2), equivalent x-cylinder

hardness (▲, Expt.3), polycrystalline TiC equivalent x-cylinder hardness (▼, Expt.4), and

equivalent x-wedge hardness (◆, Expt.5), experiment results from Atkins et al.[301], Vickers

hardness of TiC0.94 (▶, Expt.6) from Samsonov et al.[325] and Vickers hardness of TiC0.96 (★,

Expt.7) from Kohlstedt et al.[326] and predicted temperature dependent hardness of Si (c)

and Ge (d). The grey region in (c) is the phase transformation region from Domnich et al.[159].

Experimental data of Si and Ge is from Atkins et al.[301].

98

5.2.5 Phase Transformations During Indentation

The modelled temperature dependent hardness of Diamond structure Si and Ge agree

well with that of experiments[301], [330]–[332] as shown in Figure 5.5c-d. Some

semiconductors such as Si and Ge are more complicated to model since they undergo phase

transformations to β-tin phase, indentation-induced metallization of silicon[157], [159], during

indentation process between 0 and 300 oC (0.16-0.35 Tm) which is marked as a grey area in

Figure 5.5c. Above 0.35 Tm, the mono-vacancy diffusion by dislocation climb dominates the

temperature dependent hardness up to 0.6 Tm. Above 0.6 Tm, since di-vacancy diffusion model is

the only model except dislocation and mono-vacancy diffusion in this study, other possible

diffusion mechanisms in Si should be considered to determine the major deformation mechanism

at that temperature. However, the activation energy of the major mechanism above 0.6 Tm should

be equivalent to that of di-vacancy diffusion based on this model and the entropy contribution

probably lower than that of di-vacancy diffusion modelled in this model.

5.3 Modeling Procedure

5.3.1 Derivation

The temperature dependent hardness modeling procedure starts from the flow stress as a

function of temperature since flow stress is a major factor to determine hardness. The governing

equation brought from Kocks et al.[333] and Laasraoui et al.[334] as follows

Equation 5.1 𝒍𝒏𝝉𝑷𝑵(𝑻)

𝝉𝑷𝑵,𝟎𝑲= −

𝒏𝒊𝒌𝑩𝑻

𝑸𝒍𝒏 (

��𝟎𝑲

��(𝑻))

99

Where, 𝜏𝑃𝑁,0𝐾 and 𝜏𝑃𝑁(𝑇) is the flow stress at 0K and finite temperature, ��0𝐾 and ��(𝑇) is the

strain rate at 0K and finite temperature. Deformation diffusion mechanism coefficient 𝑛𝑖 is added

in order to validate the strain rate term for the same strain rate and is to be modelled later, 𝑘𝐵 is the

Boltzmann constant, 𝑇 is the temperature in Kelvin (K) and 𝑄 is the activation energy for materials’

deformation diffusion. The types of the activation energy depend on the materials’ deformation

mechanism. For example, dislocation diffusion by dislocation glide is major mechanism at low

temperature range while, self-diffusion by dislocation climb is the major mechanism at high

temperature range.

Since the indentation process usually keeps the strain rate constant[294], the strain rate

terms (��0𝐾 and ��(𝑇)) should be the same and can be removed, and the Equation 5.1 can be re-

express as follows.

Equation 5.2 𝝉𝑷𝑵(𝑻) = 𝝉𝑷𝑵,𝟎𝑲 𝒆𝒙𝒑(−𝒏𝒊𝒌𝑩𝑻

𝑸)

Then, Peierls-Nabarro flow stress[15], [335], [336] is applied into Equation 5.2.

Equation 5.3 𝝉𝑷𝑵

𝑮(𝑻) =

𝟏

(𝟏−𝒗)𝒆𝒙𝒑(−

𝟐𝝅𝒘𝟎𝑲

𝒃) 𝒆𝒙𝒑 (−

𝒏𝒊𝒌𝑩𝑻

𝑸)

Where, 𝑣 is the Poisson ratio, 𝑏 is the magnitude of Burgers vector, and 𝒘𝟎𝑲 is the dislocation

width at 0 K. The dislocation width at 0 K for edge dislocations is expressed as follows.

Equation 5.4 𝒘𝟎𝑲 =𝒔

(𝟏−𝒗)

Where, 𝑠 is the interspacing distance. For the screw dislocations[15], 𝑤0𝐾 = 𝑠.

Then, the temperature dependent flow stress (Equation 5.3) is compared with the previous

temperature dependent flow stress equation suggested by Nabarro[238] and Dietze[236] as follows.

Equation 5.5 𝝉𝑷𝑵

𝑮(𝑻) =

𝟏

(𝟏−𝒗)𝒆𝒙𝒑(−

𝟐𝝅𝒘𝟎𝑲

𝒃(𝟏 +

𝒂𝑻

𝑻𝑴))

Where, 𝑇𝑀 is a melting temperature, and 𝑎 is a modelling parameter and is known to be 1/3 or 1/10

at low temperature[236]–[238], [337] and 𝑎 = 𝑎𝑆𝐷 at high temperature.

100

The flow stress equation can be re-express as Equation 5.6 which contains the concept of

dislocation width as a function of temperature.

Equation 5.6 𝝉𝑷𝑵

𝑮(𝑻) =

𝟏

(𝟏−𝒗)𝒆𝒙𝒑(−

𝟐𝝅𝒘(𝑻)

𝒃)

The temperature dependent dislocation width term at high temperature, 𝑤(𝑇), from both

Equation 5.3 and Equation 5.5 are re-organized as follows.

Equation 5.7 𝒘(𝑻) = 𝒘𝟎𝑲 (𝟏 +𝒂𝑺𝑫𝑻

𝑻𝑴) = 𝒘𝟎𝑲 (𝟏 +

𝒏𝑺𝑫(𝟏−𝒗)𝒃𝒌𝑩𝑻

𝟐𝝅𝒔𝑸𝑺𝑫)

Where, 𝑛𝑆𝐷 is a self-diffusion mechanism coefficient that will be discussed later. The two-

temperature dependent dislocation width terms in Equation 5.7 should be the same. Since self-

diffusion, which is due to dislocation climb, is the major deformation mechanism at high

temperature, we have collected the mono-vacancy diffusion activation energies of all the materials

in order to make a correlation since mono-vacancy diffusion is the same as self-diffusion for most

metal at low temperature region.

First, the self-diffusion activation energy is correlated with materials’ melting

temperature[309], [338], [339] since Van Liempt shows the linear relationship between self-

diffusion activation energy and melting temperature[309] (as known as Van Liempt rule, 𝑄 =

17𝑅𝑇𝑀) and many research have also shown the linear relationship between the self-diffusion

activation energies and their melting temperatures within the same crystal structures such as FCC,

BCC, alkali halide and carbide rocksalt crystal structures[310]–[314]. However, the relationship

between the activation energy and melting temperature is not valid for different crystal

structures[310]–[313] as Gibbs[313] also pointed out the discrepancy by the crystal structures such

as 27.9𝑇𝑀 𝑐𝑎𝑙/𝑚𝑜𝑙𝑒 for FCC and 22.6𝑇𝑀 𝑐𝑎𝑙/𝑚𝑜𝑙𝑒 for BCC.

Second, in order to combine this relationship with various crystal structures, structural

factors should be considered. As the structural factor, materials’ slip systems are carefully

considered since the atomic jumps to the nearest neighbor in diffusion process can be described as

101

the slip systems in flow stress and the dislocation climb does not occurs on the same slip plane

where dislocation glide moves. To be specific, the dislocation climb direction can become another

½ [110]{111} slip system for cubic materials due to the identical slip systems, while HCP materials

should be considered based on their major slip systems since the dislocation climb direction is not

identical to the dislocation glide direction. For example, the atomic jump to nearest neighbor in

FCC structure happens to the [110] directions on (111) planes, which is the same as the slip system

of FCC materials, while, for basal slip dominant HCP materials, such as Mg and Zn, dislocation

climb may occur the direction on prismatic or pyramidal plane not on the basal plane. It is worth to

mention that for BCC crystal structure, the relevant slip system for the atomic jump, which is [111]

directions on (110) planes, are treated as the relevant slip system of [110] since the [211] and [321]

slip systems does not contain the nearest neighbor. For zinc blende and diamond cubic structured

semiconductor materials, the partial dislocation, 𝑏 =1

6[121], is considered since the partial

dislocation is the nearest neighboring atomic jumps. For the mono-vacancy diffusion activation

energy in alkali halide ionic structures, Schottky and Frenkel defects are considered dependent on

their major diffusion mechanisms and grab the mono-vacancy diffusion activation energies.

Furthermore, the local lattice expansion during the atomic jumps is captured by the Poisson’s ratio,

(1 − 𝑣).

Therefore, the activation energy for mono-vacancy diffusion, 𝑄1𝑉, is expressed as follows,

based on the melting temperature and the slip systems.

Equation 5.8 𝑸𝟏𝑽 = 𝟖𝝅(𝟏 − 𝒗) (𝒔

𝒃)𝟏𝑽𝒌𝑩𝑻𝑴

Where, (𝑠

𝑏)1𝑉

describes the relevant slip system for mono-vacancy diffusion atomic jumps, or the

dislocation climb. (𝑠

𝑏)1𝑉

is not always the same as 𝑠

𝑏. As shown in Figure 5.1 in the main document

and listed in Table 5.3, the modelled activation energies of materials for mono-vacancy diffusion

are well agreed with the collected experimental results. Based on these derivations, the temperature

102

dependent dislocation width at high temperature due to mono-vacancy diffusion is obtained as

follows.

Equation 5.9 𝒘(𝑻) = 𝒘𝟎𝑲 (𝟏 +𝒏𝟏𝑽

(𝟒𝝅)𝟐(𝒃

𝒔) (

𝒃

𝒔)𝟏𝑽

𝑻

𝑻𝑴)

For modeling temperature dependent flow stress, materials’ diffusion mechanisms are

needed to be fully understood first. The total diffusion coefficient during materials’ deformation is

the sum of the coefficient of each diffusion mechanism. The major deformation mechanism at low

temperature region is the dislocation (pipe) diffusion which corresponds to the dislocation glide,

while that at high temperature region such as Harper-Dorn creep deformation is majorly determined

by the self-diffusion, which is related to dislocation climb. The self-diffusion includes mono-

vacancy (1V), interstitial, or di-vacancy (2V) diffusion mechanism depends on their major

diffusion mechanisms at target temperature.

The dislocation diffusion 𝐷𝑑 and di-vacancy diffusion 𝐷2𝑉 follows the Arrhenius

relation[316], [340], [341] as shown in Equation 5.10 and Equation 5.11.

Equation 5.10 𝑫𝒅 = 𝑫𝒅𝟎 𝐞𝐱𝐩 (−

𝑸𝒅

𝒌𝑩𝑻)

Equation 5.11 𝑫𝟐𝑽 = 𝑫𝟐𝑽𝟎 𝐞𝐱𝐩 (−

𝑸𝟐𝑽

𝒌𝑩𝑻)

Where, 𝐷𝑑0 and 𝐷2𝑉

0 are the pre-exponential factors for dislocation and di-vacancy diffusion, and

𝑄𝑑 and 𝑄2𝑉 are the activation energies for dislocation and di-vacancy diffusion, respectively.

Due to the limited number of experimental data for 𝑄𝑑 , 𝑄2𝑉 , 𝐷𝑑0 and 𝐷2𝑉

0 , we made

assumptions based on the previous experimental[341]–[344] and computational research[340],

[345]–[347]. Since the typical activation energy of dislocation diffusion, 𝑄𝑑, is approximately 0.6–

0.7 of the activation energy of self-diffusion 𝑄𝑆𝐷 (or 𝑄1𝑉)[340]–[344], it is assumed that 𝑄𝑑 =

0.65𝑄1𝑉. Since the dislocation diffusion 𝐷𝑑 is close to the typical values for self-diffusion 𝐷1𝑉 at

infinite high temperature (𝐷0,𝑑 = 𝐷0,1𝑉 at 𝑇 = 𝑇𝑀)[340], [342], it is assumed that (𝜏𝑝

𝐺)𝑑= (

𝜏𝑝

𝐺)1𝑉

103

at melting temperature. Moreover, for the di-vacancy diffusion for all the materials in this study,

we assumed 𝑄2𝑉 = 1.4 𝑄1𝑉 based on previous studies[316], [341], and (𝜏𝑝

𝐺)2𝑉= 𝑒𝑥𝑝(4.5) (

𝜏𝑝

𝐺)1𝑉

from the fitting of hardness experiment data of FCC Au.

The diffusion mechanism coefficient 𝑛𝑖 is determined based on the ratio of activation

energy with respect to that of mono-vacancy diffusion, i.e., 𝑄𝑖

𝑄1𝑉, and the number of dislocation

climb jump at once (𝑗 ), i.e. 2 jumps for di-vacancy diffusion in terms of dislocation climb.

Therefore, we have obtained the diffusion mechanism coefficient, 𝑛𝑖 = 𝑗𝑛𝑄𝑖

𝑄𝑆𝐷, where 𝑗 is the

number of diffusion species, 𝑛 is a constant value, (10

3𝜋)2 , i.e. 𝑛𝑑 = (

10

3𝜋)2

𝑄𝑑

𝑄1𝑉 and 𝑛2𝑉 =

2(10

3𝜋)2

𝑄2𝑉

𝑄1𝑉 in Equation 5.13. The thermally activated dislocation width for metals at low

temperature, 𝒘𝟎𝑲 (𝟏 +𝒏𝒅(𝟒𝝅)𝟐

(𝒃

𝒔) (

𝒃

𝒔)𝟏𝑽

𝑻

𝑻𝑴), is the same as the dislocation width from previous

research[236]–[238], 𝒘𝟎𝑲 (𝟏 +𝟏

𝟑

𝑻

𝑻𝑴).

Equation 5.12 𝑯𝒗 = 𝟎. 𝟏𝟔𝟏𝟓 𝑮 (𝟏 + 𝟐𝟖𝟓𝟕. 𝟏𝒘

𝒃(𝒃

𝒔)𝟐(𝝉𝑷𝑵

𝑮)𝟐)𝟐

𝒆−𝟐.𝟐𝒌 (𝒃

𝒔)𝟒𝐭𝐚𝐧𝟑/𝟐 𝒂𝐜𝐨𝐬𝒂

Equation 5.13 𝝉𝑷𝑵

𝑮(𝑻)(𝒆𝒅𝒈𝒆) =

𝟏

(𝟏−𝒗)𝒆𝒙𝒑(−

𝟐𝝅𝒘𝟎𝑲

𝒃(𝟏 +

𝒏𝒊

(𝟒𝝅)𝟐(𝒃

𝒔) (

𝒃

𝒔)𝟏𝑽

𝑻

𝑻𝑴))

The term, 1

8𝜋2𝑒

𝑉𝑒

𝑉𝑝 from previous hardness model from Chapter 4, is modified to 2857.1

𝑤

𝑏

from 8000

3

𝑤

𝑏, since

𝑉𝑒

𝑉𝑝 was obtained from the previous dislocation width assumption, 𝑤(𝑇) =

𝑤0𝐾 (1 +1

3

𝑇

𝑇𝑀). Therefore, the

𝑉𝑒

𝑉𝑝 is not only a function of temperature but also a function of crystal

structures similarly to the activation volume during plastic deformation[348]–[350]. The present

model has considered the major active slip systems of each structure shown in Table 5.2. Figure

5.6 shows the comparisons of hT/hp and hardness (𝑯𝒗) between the modified hardness model in

this study (Equation 5.12) and experimental results used in Chapter 4.

104

Figure 5.6 Validation of the hardness model from this work. (a) hT/hp and (b) hardness

between this model and experimental results.

Table 5.2 Crystal structures and their slip systems.

Structure Dislocation

type Slip system (

𝑏

𝑠)𝑃𝑁

(𝑏

𝑠)𝑆𝐷

Examples

FCC edge [110]{111} √3/2 √3/2 Ni, Al, Au, Ir,

Rh

FCC screw [110]{110} 3√2/4 3√2/4 Ir, Rh

BCC screw [111]{110} 3√2/4 3√2/4 Mo, V, Fe, W

BCC screw [111]{211} √6/2 3√2/4 Nd, W

HCP Basal-edge [1210]{0001} 2𝑎0/𝑐 2√3/3 Be, Mg, Zn,

Cd

HCP Prism-edge [1210]{1010} 2√3/3 2𝑎0/𝑐 Ti, Zr

HCP Pyramidal-

edge [1210]{1011} 2𝑎0/√(

𝑐

2)2

+ (5𝑎03)2

2𝑎0/𝑐 Co

Diamond

cubic edge [110]{111} √3 1/√3

C(diamond),

Si, Ge

Diamond

cubic screw [110]{111} √3 1/√3 Si, Ge

Zinc

blende edge [110]{111} √3 1/√3

SiC, ZnSe,

ZnS

Zinc

blende screw [110]{111} √3 1/√3 ZnSe, ZnS

Rocksalt edge [110]{111} √3 1 (for Me)

√3 (for C) TiC, TiN, ZrC

105

Rocksalt edge [110]{110} √2 1 (for Me)

√2 (for O) MgO

Rocksalt edge [110]{100} 1 1 NaCl, MgO

5.3.2 Temperature-Dependent Elastic Properties

The elastic properties such as G, B, k, and 𝒗 do change as a function of temperature. For

example, the elastic stiffness matrix (i.e. 𝑪𝟏𝟏, 𝑪𝟏𝟐 𝐚𝐧𝐝 𝑪𝟒𝟒 ) of pure BCC W decrease as the

temperature increases[351]. However, their contribution to the temperature-dependent hardness is

negligible compared to that of thermally activated dislocation width as shown in Figure 5.7.

Therefore, in this study, we used the elastic properties at 0 K from first-principles calculations.

Figure 5.7 Comparison of temperature-dependent hardness of BCC W between a) using

temperature-dependent elastic properties and b) using fixed elastic properties at 0 K. the

temperature-dependent elastic properties of BCC W is from Hu et al.[351].

106

Table 5.3 Self-diffusion activation energy modeling.

Element structure expt. 𝜈 (𝑏

𝑠)𝑆𝐷

* 𝑇𝑀(𝐾)[322] Modeled

Q (eV) Expt. Q (eV)

Calc.

Q[352]

(eV)

Li BCC 0.362 1.061 453.69 0.591

0.518[353]

0.556[353]

0.548[354]

0.546[354]

0.58[354]

0.584

Be HCP 0.036 1.275 1560.15 2.553 1.71(//c)[353][354]

1.63(⊥c)[353][354]

1.722(//c)

1.908(⊥c)

Na BCC 0.310 1.061 370.87 0.523

0.365[353]

0.43[354]

0.453[355]

0.453

Mg HCP 0.289 1.232 923.15 1.153 1.40-1.44(//c)[354]

1.41-1.43(⊥c)[354]

1.193(//c)

1.215(⊥c)

Al FCC 0.349 0.866 933.47 1.520

1.26[353]

1.48[354]

1.50[354]

1.31[354]

1.307

K BCC 0.350 1.061 336.53 0.447

0.386[353]

0.41[354]

0.423[355]

0.374

Sc HCP 0.271 1.255 1814.15 2.283 2.610(//c)

2.612(⊥c)

Ti HCP 0.327 1.260 1941.15 2.247 1.75[354]

2.0(⊥c)[355]

2.629(//c)

2.729(⊥c)

V BCC 0.362 1.061 2183.15 2.843

3.47[354]

4.24[354]

3.194[355]

3.09[355]

3.086

Cr BCC 0.212 1.061 2180.15 3.508

4.58[353][354]

4.2[355]

3.1F

3.846

Fe BCC 0.290 1.061 1811.15 2.626

2.99[353]

3.00[353]

2.92[354]

2.60[354]

2.615[355]

3.130[355]

3.136

Co HCP 0.328 1.232 1768.15 2.089

2.99[354]

3.15[354]

2.91[355]

3.113(//c)

3.102(⊥c)

Ni FCC 0.304 0.866 1728.15 3.009

2.88[353]

2.96[355]

2.91[355]

2.774

Cu FCC 0.346 0.866 1357.77 2.221 2.06[353]

2.19[355] 1.948

Zn HCP 0.262 1.077 692.68 1.027

0.95(//c)[354]

1.00(⊥c)[354]

0.945(//c)[355]

0.780(//c)

0.748(⊥c)

107

1.054(⊥c) [355]

Y HCP 0.245 1.274 1799.15 2.309 2.62(//c)[354]

2.91(⊥c)[354]

2.495(//c)

2.521(⊥c)

Zr HCP 0.336 1.255 2128.15 2.437

3.2[353]

1.16-3.64[353]

3.17(//c)[355]

2.605(//c)

2.779(⊥c)

Nb BCC 0.397 1.061 2750.15 3.386 3.62[353]

4.163[355] 3.365

Mo BCC 0.296 1.061 2896.15 4.161 4.53[353]

4.99[355] 4.180

Rh FCC 0.264 0.866 2237.15 4.117 4.05[353] 3.958

Pd FCC 0.385 0.866 1828.05 2.811

2.76[353]

2.76[354]

2.59[355]

2.430

Ag FCC 0.369 0.866 1234.93 1.948

1.76[353]

1.96[355]

1.912[355]

1.76[355]

1.597

Cd HCP 0.344 1.061 594.22 0.796

0.81(//c)[353][354]

0.85(⊥c)[353][354]

0.789(//c)[355]

0.828(⊥c)[355]

0.475(//c)

0.408(⊥c)

Gd HCP 0.261 1.260 1586.15 2.017 1.42[354] 2.161(//c)

2.282 (⊥c)

Tb HCP 0.258 1.265 1629.15 2.068 2.242(//c)

2.334(⊥c)

Dy HCP 0.245 1.271 1685.15 2.169 2.344(//c)

2.414(⊥c)

Ho HCP 0.234 1.274 1747.15 2.274 2.397(//c)

2.452(⊥c)

Er HCP 0.240 1.274 1770.15 2.288 3.13(//c)[354]

3.14(⊥c)[354]

2.498(//c)

2.539(⊥c)

Hf HCP 0.284 1.266 2506.15 3.072

3.84(//c)[354]

3.61(⊥c)[354]

3.345[355]

3.191(//c)

3.364(⊥c)

Ta BCC 0.342 1.061 3290.15 4.421

3.8[353]

4.39[353]

4.13[353]

4.07[353]

4.28[354]

4.39[354]

3.849

W BCC 0.280 1.061 3695.15 5.435

5.45[353]

6.084[355]

6.487[355]

5.210

Re HCP 0.293 1.239 3459.15 4.275 5.3[353] 5.662(//c)

4.915 (⊥c)

Ir FCC 0.250 0.866 2739.15 5.140 4.54[354] 4.639

Ir FCC-s 0.250 0.943 2739.15 4.196 4.54[354] 4.639

Pt FCC 0.396 0.866 2041.45 3.086

2.96[354]

2.89[354]

2.67[355]

2.208

108

Au FCC 0.424 0.866 1337.33 1.926

1.806[354]

1.826[354]

1.810[355]

1.73[355]

1.247

Pb FCC 0.409 0.866 600.61 0.888

1.11[354]

1.05[355]

1.075[355]

1.031

Th FCC 0.287 0.866 2023.15 3.606

3.1-3.32[353]

3.11[354]

3.22[355]

3.469

Th FCC-s 0.287 0.943 2023.15 3.312

3.1-3.32[353]

3.11[354]

3.22[355]

3.469

TiC

[Ti] NaCl 0.218 1.000 3430 5.806 7.649[356]

TiC

[C] NaCl 0.218 1.633 3430 3.556 3.404[314]

ZrC

[Zr] NaCl 0.181 1.000 3805.15 6.752 7.45[357]

ZrC

[C] NaCl 0.181 1.633 3805.15 4.135 3.148[314],a

NbC[N

b] NaCl 0.284 1.000 3763.15 5.836 5.51[357], [358]

NbC

[C] NaCl 0.284 1.633 3763.15 3.574 3.113[314],a

TaC

[C] NaCl 0.230 1.633 4215.15 4.305 3.487[314],a

MgO NaCl 0.188 1.000 3125.15[359

] 5.496

4.258

4.77[360]

AgCl NaCl 0.3590.

420 1.000 728.15

1.011

0.915

1.567[311]

0.736-0.829(𝑄𝐼𝑆𝐷)

[361]

1.02(𝑄𝑉𝑆𝐷) [361]

AgBr NaCl 0.345

0.440 1.000 707.15

1.000

0.858

1.276[311]

0.622-

0.777(𝑄𝐼𝑆𝐷)[361]

0.881(𝑄𝑉𝑆𝐷) [361]

CsCl CsCl 0.222

0.280 1.000 918.15

1.548

1.432

1.561[311]

1.555[361]

KCl NaCl 0.101 1.000 1049.15 2.043 2.096[311]

2.021[361]

KI NaCl 0.061 1.000 959 1.950 1.767[311]

KBr NaCl 0.118 1.000 1003 1.916 2.131[311]

KF NaCl 0.151 1.000 1130 2.078 2.196[311]

LiBr NaCl 0.200 1.000 823.15 1.426 1.347[361]

LiCl NaCl 0.272

0.230 1.000 883.15

1.399

1.473

1.647[311]

1.503[361]

LiF NaCl 0.254

0.210 1.000 1121.15

1.801

1.918

2.196[311]

1.917[361]

LiI NaCl 0.240 1.000 722.15 1.189 0.985[361]

NaCl NaCl 0.164 1.000 1073.15 1.945 2.136[311]

109

0.240 1.766 1.969[361]

NaBr NaCl 0.170 1.000 1028 1.848 2.016[311]

NaF NaCl 0.174 1.000 1261 2.256 2.690[311]

RbCl NaCl 0.131 1.000 988 1.859

1.986 (𝑄𝑉𝑆𝐷) [311]

1.606-2.125(𝑄𝐼𝑆𝐷)

[361]

Si Diamond

Cubic 0.220 0.612 1687.15 4.654

5.141[311]

4.65 (𝑄𝑉𝑆𝐷) [362]

4.82 (𝑄𝐼𝑆𝐷) [362]

4.644[363]

4.77[364]

Ge Diamond

Cubic 0.207 0.612 1211.35 3.396

3.36[311]

3.14[365]

3.05[366]

3.09[367]

GaAs ZnS 0.243 0.612 1511.15 4.045

4.24(𝑄𝐺𝑎𝑆𝐷)[368]

4[369]

4.0[370]

5.6(𝑄𝐺𝑎𝑆𝐷)[371]

InP ZnS 0.276 0.612 1335.15 3.417 3.85(𝑄𝐼𝑛

𝑆𝐷)[372]

5.65(𝑄𝑃𝑆𝐷)[372]

GaSb ZnS 0.227 0.612 985.15 2.692

3.148(𝑄𝐺𝑎𝑆𝐷) [373]

3.443(𝑄𝑆𝑏𝑆𝐷) [373]

1.59(𝑄𝑆𝑏𝑆𝐷)[374]

3.24(𝑄𝐺𝑎𝑆𝐷)[374]

InSb ZnS 0.287 0.612 800.15 2.019

1.813(𝑄𝐼𝑛𝑆𝐷) [373]

1.934(𝑄𝑆𝑏𝑆𝐷) [373]

1.45 (𝑄𝐼𝑛𝑆𝐷) [343]

1.91(𝑄𝑆𝑏𝑆𝐷) [343]

ZnS ZnS 0.291 0.707 1458.15 3.166 3.15[375]

ZnSe ZnS 0.300 0.707 1798.15 3.855 3.0[376]

3.45[377]

ZnTe ZnS 0.301 0.707 1511.15 3.235 2.69[378]

*(𝑏

𝑠)𝑆𝐷 is obtained from Table 5.2.

[ ] diffusion species

110

Chapter 6

Hardness Modeling for Layered Structures: The Origin of Hall-Petch

Relation

6.1 Introduction

The yield (or flow) stress and hardness of polycrystalline materials usually increase as the

grain size (d) decreases from millimeters to tens of nanometers according to the Hall-Petch

relationship[379], [380]. This behavior is mainly due to the dislocations being hindered (pinned)

by grain boundaries[381]. Such dislocation hindering originates from a higher energy barrier at the

grain boundary for dislocations to pass through the boundary than for dislocations to move within

grain. That is to say, as the grains become smaller, the number of piled up dislocations at grain

boundaries increase due to the increase in the density of higher energy barriers; in turn, the effect

of dislocation blocking increases, strengthening materials[382]. The increased yield stress or

hardness is inversely proportional to the square root of grain size as expressed in Equation 6.1 and

Equation 6.2.

Equation 6.1 𝝈 = 𝝈𝟎 + 𝒌𝒅−𝟏

𝟐

Equation 6.2 𝑯𝑽 = 𝑯𝑽,𝟎 + 𝒌𝒅−𝟏

𝟐

Where 𝜎0 and 𝐻𝑉,0 are yield and hardness of the bulk state, k is a Hall-Petch coefficient, and d is

the grain size. Due to the simplicity and well-fitting nature of the equation, Equation 6.1 and

Equation 6.2 are widely used to demonstrate the effect of grain boundary strengthening in pure

metals and alloys[383]–[396]. The coefficient k has been estimated for pure metals and alloys[389].

111

Many attempts have been made to interpret the physical meaning behind the Hall-Petch

relationship by establishing a universal relation for all materials[397]–[405]. Some researchers

modified the original Hall-Petch relation as part of a power law expansion in order to capture the

yield (flow) stress and hardness of different crystal structured pure metals and alloys. Despite their

great efforts, the physical meaning behind this empirical relation has not yet been fully resolved.

To be specific, due to the simplicity of Equation 6.1 and Equation 6.2, these relations do not

consider crystal structures, slip systems, grain misorientations, other slip system activation at grain

boundaries, and other defect structures beyond grain boundaries, etc.

The grain misorientation or grain boundary energy, one of these key factors, plays an

important role in increasing 𝜎 (or 𝐻𝑉) for the same grain size. In the grain size (or twin size)

dependent hardness plot of FCC Cu as shown in Figure 6.1, the hardness values for the same grain

size deviate more than 1 GPa, although the average hardness seems to follow the Hall-Petch relation

trend. In part, this deviation is due to variations in grain misorientation and/or grain boundary

energy since the experimental hardness data of all the twin boundary cases[406]–[408], which has

a lower grain boundary energy, are lower than that of grain size effect[409]–[416]. Thus, the Hall-

Petch relation breaks down, that is, the twin-size dependent hardness, since the Hall-Petch line

should be parabolic to meet the bulk state hardness, 0.362 GPa at 0 (grain size)-1/2. It should be

noted that the grain size dependent hardness experimental data[409]–[416] are mostly from the

ECAP process which results in high angle grain misorientations[417]. Furthermore, Li et al.[418]

showed that hardness increases by increasing the grain misorientation from 2 to 20 degrees in their

dislocation-grain boundary penetration model. Li et al. also pointed out a changing of the Hall-

Petch coefficient (k) due to grain misorientation. Therefore, the grain misorientation or grain

boundary energy plays an important role in determining hardness.

112

0.0 0.1 0.2 0.3 0.4 0.5

1

2

3

4

H

ard

ness (

GP

a)

Grain size-1/2

(nm-1/2

)

Chen et al.(G)

Sanders et al.(G)

Jiang et al.(G)

Agnew et al.(G)

Gray et al.(G)

Valiev et al.(G)

Haouaoui et al.(G)

Suryanarayanan et al.(G)

You et al.(T)

Schwaiger et al.(T)

Anderoglu et al.(T)

0.37

Figure 6.1. Grain size dependent hardness of FCC Cu. Grain size (G) dependent hardness

(solid shapes) are from Chen et al.,[409] Sanders et al.,[410] Jiang et al.,[411] Agnew et

al.,[412] Gray et al.,[413] Valiev et al.,[414] Haouaoui et al.[415], and Suryanarayanan et

al.[416] twin size(T) dependent hardness are from You et al.[406], Lu et al.[407] and

Anderoglu et al.[408].

In addition, other slip systems at the boundary due to the presence of the twin (or grain)

boundaries possibly can be activated depending on the crystal structure. These activated slip

systems near the boundary will increase the energy barrier for dislocations to move, if so, then this

results in dislocation pile-ups near the boundary. For example, it is reported that the 1

2< 110 >

screw dislocations, which are originated from the 1

6[112] partial dislocations near twin boundary

regions in twinned FCC structures, are observed from both experiments and calculations[386],

[419]–[424], although the major slip system for FCC metals are edge dislocations of ½

<110>{111}.

In the present work, in order to systematically investigate the origin of the Hall-Petch

relation from the microscale, twinned FCC Cu, Ag, and carbon diamond structures with various

113

bilayer distances are used since twin boundaries can be considered to be coherent grain boundaries.

For modeling twin-size dependent hardness, the Peierls-Nabarro (PN) flow stress are used since

the PN flow stress equation estimates well the various crystals’ experimental flow stress[15], [425].

To effectively model the slip system angle term, one of the assumptions in classical PN flow stress

is that a dislocation moves through the straight-forward slip direction; this enables to calculate the

flow stress of curved slip direction. Further, additional slip system activation due to the existence

of twin boundaries in FCC metals is also considered. As such, this model will give an idea of how

a curved slip system increases the flow stress and the indentation hardness.

6.2 Methodology

6.2.1 Derivation of Peierls-Nabarro Flow Stress for Twinned Structures

The Peierls-Nabarro (PN) flow stress is the minimum shear stress to overcome the energy

barrier (PN potential) for a dislocation and to move the dislocation. The PN flow stress at 0 K for

edge and screw dislocations are expressed below[15], [235]

Equation 6.3 𝝉𝑷𝑵

𝑮(𝒆𝒅𝒈𝒆) =

𝟏

(𝟏−𝝂)𝒆𝒙𝒑(

−𝟐𝝅𝒘

𝒃)

Equation 6.4 𝝉𝑷𝑵

𝑮(𝒔𝒄𝒓𝒆𝒘) = 𝒆𝒙𝒑(

−𝟐𝝅𝒘

𝒃)

where 𝐺 is shear stress, 𝜈 the Poisson’s ratio, b the Burgers vector and w the dislocation width. The

dislocation width, w, is equal to 𝑠

(1−𝑣) for edge dislocations at 0 K and 𝑠 for screw dislocations[15].

Here, the Burgers vector and the dislocation width are based on full dislocations, i.e., <110> slip

direction on {111} planes for FCC crystals and their derivatives (including diamond cubic)

structures since the PN flow stress equation is for a straight-forward dislocation movement[426],

[427].

114

However, the PN flow stress of twinned structures is more complex since the full

dislocation and the slip planes in twinned structures are not in a straight line and not on a flat slip

plane, but instead they are distorted with certain angles near the twin boundaries as shown in Figure

6.2. Thus, in order to model the angle term in the twinned structures, the angle relationship between

two partial dislocations in FCC and diamond cubic structures are investigated, i.e., 120 for two

1/6<211> Shockley partial dislocations in FCC and 109.47 for two ¼ <111> glide partial

dislocations. Since the PN potential, estimated from these partial dislocations, should be equal to

that estimated from full dislocations, the angle between two partial dislocations plays a crucial role

in determining the PN flow stress. Therefore, the distorted angle should be captured by b/s in

Equation 6.3 and Equation 6.4 as expressed as follows.

Equation 6.5 𝒃𝒇𝒖𝒍𝒍

𝒔𝒇𝒖𝒍𝒍=𝒃𝒑𝒂𝒓𝒕𝒊𝒂𝒍

𝒔𝒑𝒂𝒓𝒕𝒊𝒂𝒍𝒇(𝜽𝒑)

Where 𝜃𝑝 is the angle between two partial dislocation vectors, i.e., 60 for FCC and 70.53 for

diamond cubic structures. The b/s of partial and full dislocations and the angles for FCC and

diamond cubic structures are listed in Table 6.1 and Figure 6.3.

115

Figure 6.2 The slip direction (plane) angles(𝜽𝒊) generated by twin boundaries in (a) twinned

FCC Cu and (b) twinned carbon diamond cubic.

Table 6.1 The angle (𝜽𝒑) of full and partial dislocations in structures.

Angle

(𝜽𝒑)

Structure

(dislocation type) b s 𝒇(𝜽𝒑)

0

FCC and

HCP(c/a=1.633)

(full)

𝑎0√2

2 √(

√3𝑎03)

2

+ (√3𝑎03)

2

1.00

0

Diamond cubic

and Zinc blende

(full)

𝑎0√2

2 √(

√3𝑎06)

2

+ (√3𝑎06)

2

1.00

116

60

FCC and

HCP(c/a=1.633)

(partial)

𝑎0√6

6

𝑎0√2

2 2.12

70.53

Diamond cubic

and Zinc blende

(partial)

𝑎0√3

4

𝑎0√6

6 2.31

Figure 6.3 The slip direction (plane) angles(𝜽𝒑) of full and partial dislocations in FCC and

diamond cubic.

Based on the information from Table 6.1 and Figure 6.3, the function of angle term, 𝑓(𝜃𝑝),

is obtained as follows.

Equation 6.6 𝒃𝒇𝒖𝒍𝒍

𝒔𝒇𝒖𝒍𝒍=𝒃𝒑𝒂𝒓𝒕𝒊𝒂𝒍

𝒔𝒑𝒂𝒓𝒕𝒊𝒂𝒍(𝟏 + 𝟎. 𝟎𝟏𝟖𝟔𝟏 × 𝜽𝒑(𝒊𝒏 𝒅𝒆𝒈𝒓𝒆𝒆𝒔))

𝑜𝑟 𝒃𝒇𝒖𝒍𝒍

𝒔𝒇𝒖𝒍𝒍=𝒃𝒑𝒂𝒓𝒕𝒊𝒂𝒍

𝒔𝒑𝒂𝒓𝒕𝒊𝒂𝒍(𝟏 + 𝟏. 𝟎𝟔𝟔𝟐 × 𝜽𝒑(𝒊𝒏 𝒓𝒂𝒅𝒊𝒂𝒏𝒔))

Then, we have considered the twinned super-structures as a unit cell structure since the slip

direction of the structures is shifted by the presence of the twin boundaries. After relaxation of the

twinned FCC and diamond cubic structures from the first principles calculations, of which details

are included in Chapter 6.2.2, the angles and calculated 𝑏𝑖/𝑠𝑖 of each layer in FCC and diamond

117

cubic structures for the three possible slip directions, which move across the twin boundary, are

obtained as shown in Figure 6.4 based on Equation 6.6. Then, the 𝑏𝑖/𝑠𝑖 of each layer is averaged

by a root mean square (RMS) averaging scheme as shown in Figure 6.5 since the slip system

modulation are similar to that of the electric current. The three possible slip directions in twinned

diamond cubic and FCC structures towards twin boundaries are averaged by arithmetic mean of

three possible slip directions. Therefore, the twin boundary bilayer based bsd

ssd term is obtained as

follows.

Equation 6.7 𝐛𝐬𝐝

𝐬𝐬𝐝= √

𝟏

𝐧∑ (

𝐛𝐢

𝐬𝐢𝐟(𝛉𝐢))

𝟐

𝐧𝐢

Where n is the number of layers between twin boundaries and 𝑖 is the 𝑖th layer, bsd and ssd are

Burgers vector and slip plane spacing for the superdislocation, respectively.

Figure 6.4 Normalized 𝒃𝒊/𝒔𝒊 (with respect to that of each structures) changes of each layers

in (a) twinned carbon diamond cubic and (b) FCC Cu.

118

Figure 6.5 Schematics of the method of modeling of b/s in twinned structures.

By putting Equation 6.7 into Equation 6.3, the flow stress of twinned structure was obtained

as follows.

Equation 6.8 𝝉𝑷𝑵

𝑮(𝒆𝒅𝒈𝒆) =

𝟏

(𝟏−𝝂)𝐞𝐱𝐩 (

−𝟐𝝅

(𝟏−𝝂)

𝐬𝐬𝐝

𝐛𝐬𝐝)

Furthermore, the modeled hardness was calculated by integrating Equation 6.8 into the

hardness model proposed in Chapter 4 and Chapter 5 as follows. The modeled hardness is room

temperature based, see Chapter 5 for detailed information.

Equation 6.9 𝑯𝒗 = 𝟎. 𝟏𝟔𝟏𝟓 𝑮 (𝟏 + 𝟐𝟖𝟓𝟕. 𝟏𝒘𝐬𝐝

𝐛𝐬𝐝(𝐛𝐬𝐝

𝐬𝐬𝐝)𝟐(𝝉𝑷𝑵

𝑮)𝟐)𝟐

𝒆−𝟐.𝟐𝒌 (𝐛𝐬𝐝

𝐬𝐬𝐝)𝟒𝐭𝐚𝐧𝟑/𝟐 𝒂 𝐜𝐨𝐬𝒂

Where k is the Pugh ratio which is shear modulus over bulk modulus (G/B) and 𝑎 is the indenter

angle which is 22 .

119

6.2.2 First-Principles Calculations

First-principles calculations are performed using the Vienna Ab-initio Simulation Package

(VASP)[27], [28], [31]. Electron-ion interactions are described by the projector augmented-wave

(PAW) method[29], and the X-C functionals depicted by the generalized gradient approximation

(GGA), as implemented by Perdew, Burke, and Ernzerhof (PBE)[30], are used. The plane wave

cutoff energy of 520 (C), 390 (Cu) and 360 (Ag) eV is used for all calculations; this is 1.3 times of

the recommended values by VASP[31]. For primitive diamonds and face centered cubic structures

and their twinned structures, more than 8000 k-points are implemented. The k-mesh guarantees

errors below 0.1meV/atom. The structures are relaxed by the Methfessel-Paxton method[76]. After

relaxations, a final static calculation using the tetrahedral method with Blöch corrections[36] is

applied to ensure the accuracy of total energy. The crystal structures of diamond carbons, Cu and

Ag are considered to be diamond cubic (Fd3m) and face centered cubic (Fm3m) structures, and the

twinned structures of both diamond cubic and FCC used in this study are sigma 3 boundaries on

(111) plane.

In the present work, elastic stiffness constants are predicted at 0 K via the stress–strain

method from first-principles calculations[39]. The detailed descriptions are the same as the

calculation of elastic stiffness constants of LPSO in Chapter 2.2

6.3 Results and Discussion

In the present twin bilayer dependent hardness model, the local active slip system 𝑏𝑖/𝑠𝑖

changes due to the presence of twin boundaries in FCC and diamond cubic structures are

investigated. The active slip system, 𝑏 = ½ < 110 >, of the twinned structure is not straight-

forward but curved near the twin boundaries as shown in Figure 6.2, Figure 6.4 and Figure 6.5,

120

while it is are straight-forward in the bulk state of FCC and diamond cubic structures. This 𝑏𝑖/𝑠𝑖

change can be interpreted as that a dislocation requires a certain minimum energy in order to move

to another position which corresponds to PN barrier. Due to the slip angle change at twin

boundaries, the required barrier increases at the twin boundaries by as much as indicated in

Equation 6.6 and Figure 6.4. Due to this increased barrier, dislocations are piled up at the

boundaries. Therefore, it can be re-interpreted that the alternating stress fields (PN barrier) in a

modulated structure by systematic defects (i.e., twin boundaries) inhibit dislocation motion[428].

Furthermore, due to the presence of the twin boundaries, other slip systems can be

sometimes activated near the twin boundary regions. For example, it is reported that the 1

2< 110 >

screw dislocations, which originate from 1

6[112] partial dislocations near twin boundary regions in

the twinned FCC structure, are observed in both experiments and calculations[386], [419]–[424].

The activation of other slip systems by twin boundaries is due to the change of electronic structures

near the twin boundary regions[419]. In order to check the change of electronic structures near the

boundary region, the differential charge density plots of carbon diamond cubic and FCC Cu

structures, the same method used in Chapter 2.3.3 and Equation 2.8, are generated as shown in

Figure 6.6. In the differential charge density of the carbon diamond cubic structure, we find that

there is no change in the isosurface shapes at the twin boundary. Therefore, it can be assumed that

the dislocation mechanism near the twin boundary in the carbon diamond cubic structure is as

follows

Equation 6.10 𝒃 (𝟏

𝟐[𝟏𝟏𝟎]) → 𝒃(

𝟏

𝟐[𝟏𝟏𝟎]𝑻) 𝒇𝒐𝒓 𝑪𝒂𝒓𝒃𝒐𝒏 𝒅𝒊𝒂𝒎𝒐𝒏𝒅

121

Figure 6.6 Differential charge density plots of (a) carbon diamond cubic (reference state),

(b) FCC Cu (reference state), (c) twinned carbon diamond cubic and (d) twinned FCC Cu

structures. Red arrows indicate the close-up view of twin boundary area. Isosurfaces are

0.0065 (e/Å 3) and the atom sizes are exaggerated for better visualization.

However, we found the change of the isosurface shape in twinned FCC Cu from a

tetragonal (triangle in 2 Dimension) to a prism (rectangular in 2D) near the twin boundaries. This

indicates that the active slip system near the regions is probably not the same as that of the bulk

state, which is expected to be 𝟏

𝟐[𝟏𝟏𝟎] . Based on previous research on possible dislocation

mechanisms near twin boundaries for FCC metals,[386], [419]–[424] the Shockley partial

122

dislocations (𝟏

𝟔[𝟏𝟏��]) are activated near twin boundaries,[419] and these partial dislocation pairs

mostly have combined screw characteristics instead of edge characteristic [386] as expressed in

Equation 6.11. The active slip system is originated from the electronic structure, i.e., the

unsaturated d-bonds[131]. Therefore, it is assumed that the region where the partial or screw

dislocations are activated is limited to two layers away from the twin boundary as shown in Figure

6.6.

Equation 6.11 𝒃 (𝟏

𝟐[𝟏𝟏𝟎]) → 𝒃𝟏 (

𝟏

𝟐[𝟏𝟏𝟎]𝑻) + 𝒃𝟐 (

𝟏

𝟔[𝟏𝟏��] ) 𝒇𝒐𝒓 𝑭𝑪𝑪 𝒎𝒆𝒕𝒂𝒍𝒔

Based on Equation 6.10 for carbon diamond cubic and Equation 6.11 for FCC Cu, we have

calculated b/s as shown in Figure 6.4. Interestingly, 𝐛𝐢

𝐬𝐢 for carbon diamond cubic at the twin

boundary increases up to 2.3 times of that of the bulk state, while that for FCC Cu at the twin

boundary increases up to 11 times of that of the bulk state due to the activation of screw

dislocations. After averaging 𝐛𝐢

𝐬𝐢 based on the methodology as shown in Figure 6.5 and expressed

in Equation 6.7, 𝐛𝐬𝐝

𝐬𝐬𝐝 of the relaxed twinned structures with different twin bilayer distances are

calculated as shown in Figure 6.7. It should be noted that the relative change of 𝐛𝐬𝐝

𝐬𝐬𝐝 in the carbon

diamond cubic structure for a 10 nm twin bilayer increases 2.7% with respect to that of the bulk

state, while that in FCC Cu for a 10 nm twin bilayer increases 84.2% from that of the bulk state

due to the activation of partial dislocations near the twin boundaries.

123

Figure 6.7 𝒃𝒔𝒅/𝒔𝒔𝒅 changes by the various twin layer distances in (a) twinned carbon

diamond cubic and (b) FCC Cu.

The obtained twin bilayer dependent 𝑏𝑠𝑑/𝑠𝑠𝑑 was put into Equation 6.9 in order to

calculate the twin bilayer dependent hardness. Figure 6.8 shows the twin bilayer dependent

hardness model for the twinned carbon diamond cubic structure as well as the experimental values

from Huang et al.[429] and Irifune et al.[430]. This model agrees well with the previous

experimental results[429], [430]. Figure 6.9 also shows the agreement between the hardness from

the current hardness model and various experimental results on FCC Cu and Ag. It should be noted

that we have collected twin bilayer dependent hardness experimental data[406]–[408], [431]–[433]

with large grain size experiments (> 0.5~1m𝜇m) and columnar grained nanotwin data (excluding

equiaxed grained nanotwin data) in order to separate out the grain size effect on hardness.

Furthermore, the twin bilayer distances are approximately determined by considering a 100% twin

fraction. For example, if the average twin bilayer distance is 20 nm and the fraction of twin

boundaries is 50%, then the average twin bilayer is considered to be 40 nm since the twin

boundaries do not interact with each other as shown in Figure 6.4. For Expt.1[406] of twinned

FCC Cu, the hardness is assumed to follow the Tabor’s relation (i.e., 𝐻𝑣 = 3𝜎𝑌).

124

Figure 6.8 Hardness of diamond carbon as a function of twin bilayer distance. Expt.1 and 2

are from Huang et al.[429] and Irifune et al.[430], respectively. Open blue triangles are

obtained from relaxed structures calculated from first-principles calculations.

Figure 6.9 Hardness of FCC (a) Cu and (b) Ag as a function of twin bilayer distance. For (a)

FCC Cu, Expt.1 from You et al.[406], Expt.2 from Lu et al.[407] and Expt.3 from

Anderoglu et al.[408] are included. For (b) FCC Ag, Expt.1 from Bufford et al.[431], Expt.2

from Bufford et al.[432] and Expt.3 from Furnish et al.[433] are included. Red dash line is

the hardness of their bulk state.

In order to compare the hardness model with the Hall-Petch relation, Figure 6.10 shows

the hardness as a function of (twin bilayer)-1/2 of twinned carbon diamond, FCC Cu and FCC Ag.

This also shows that the classical Hall-Petch relation breaks down especially for carbon diamonds.

The experimental hardness of a twinned carbon diamond in Figure 6.10 clearly shows that the

125

experimental data do not follow the linear Hall-Petch relation from the point of (0,100) which is

the hardness of the bulk state, but the hardness trend of the experiment data follows this model. For

FCC metals, although the trend is not stronger than that of carbon diamond since it is very difficult

to separate out twin bilayer effect on hardness, the hardness trend does not follow the Hall-Petch

relation (i.e., the linear line in Figure 6.10). However, the experimental hardness trend is captured

by this model.

126

Figure 6.10 Hall-Petch relationship in hardness of (a) carbon diamond, (b) FCC Cu and (c)

FCC Ag as a function of twin bilayer distance. References are from those in Figure 6.8 and

Figure 6.9. ★ in the plots are the hardness of bulk state, and these are from Teter[215] for

carbon diamond, from Samsonov[291] for FCC Cu and Ag. Red dash lines are the slope for

Hall-Petch relation.

127

Chapter 7

Hardness Modeling of LPSO Phases

In this chapter, the hardness models developed in Chapter 4, Chapter 5 and Chapter 6 are

applied to LPSO phases in order to predict the hardness of LPSO phases in the Mg-Al-Gd system.

This chapter mainly focuses on the prediction of the hardness of LPSO phases in the Mg-Al-Gd

system by the determination of the factors that affect the hardness, also discussed in previous

chapters.

7.1 Methodology

Based on the hardness models presented in Chapter 4, Chapter 5, and Chapter 6, the

chemical composition contribution (i.e., diluted phases such as Cu-2at.%Ni) on hardness are

captured by the elastic properties (shear and bulk moduli) and the change of the active slip systems

(the mechanical paths how the dislocations move in the crystal lattice). The slip systems are

distorted near the defects, i.e., vacancies, impurity atoms and twin boundaries, and the slip systems

are no longer the same as that without defects.

In order to predict the hardness of any material, the following information is needed: (1)

active slip systems, (2) melting temperature and (3) elastic properties (described in Chapter 4), (4)

diffusion mechanisms (described in Chapter 5), and (5) slip system angle term and additional slip

system activation if the slip systems are not straight-forward (described in Chapter 6).

For the prediction of the hardness of LPSO phases, the melting temperature is obtained

from Figure 3.9 in Chapter 3, and elastic properties of LPSO phases used in this study are from

Table 2.3. The diffusion mechanisms of LPSO phases are assumed to be the same as that of HCP

Mg since slip occurs along the weakest bonds, which are between mg layers on basal plane or move

128

between L12 clusters surrounded by Mg atoms. In order to determine active slip systems of LPSO

phases, we first determine the LPSO structures, (Mg)116(Gd)16(Al)12(Gd)2 for 18R and

(Mg)140(Gd)16(Al)12(Mg)2 for 14H, based on their phase stability discussed in Chapter 3. For the

active slip systems of LPSO phases, basal slip on {0001} for 14H and 18R LPSOs, pyramidal slip

on {1108} for 18R and prismatic slip on {1100}for 14H are considered based on Hagihara et

al.[434] as shown in Figure 7.1. The slip system angle terms for LPSO phases are calculated using

Equation 6.6 as discussed in Chapter 6.

Figure 7.1 Slip systems of (a) 18R and (b) 14H LPSOs. Thin solid lines are the pyramidal

slip, black thick lines are the slip direction within FCC layers, red thick lines are the basal

slip, and dash lines are the L12 cluster.

7.2 Results and Discussion

Since the values of indentation hardness and flow stress change exponentially as the change

of b/s (slip system term), the determinations of the 𝑏𝑖/𝑠𝑖 of 14H and 18R LPSO structures are

crucial to predict hardness. Although the basal slip systems of both 14H and 18R LPSOs are simply

129

obtained since they are straight-forward, the pyramidal slips are calculated based on the 𝑏𝑖/𝑠𝑖 of

14H and 18R LPSO structures as shown in Figure 7.2.

Figure 7.2 𝒃𝒊/𝒔𝒊 changes of (a) 18R and (b) 14H LPSOs. Pyramidal slip on {1��08} for 18R

and prismatic slip on {1��00} for 14H are applied.

As there are no hardness data for LPSO phases in the Mg-Al-Gd systems, we instead used

the hardness data for LPSO phases in the Mg-Zn-Y systems[435]–[440] since the elastic properties

of both Mg-Al-Gd and Mg-Zn-Y are very similar (less than 8% difference in shear modulus, see

Table 2.3) and the atomic displacements in Mg layers are negligible.

Based on the hardness models and the above information, the predicted hardness of 18R

and 14H LPSO phases in the Mg-Al-Gd system is 1.81 GPa (18R) and 1.38 GPa (14H) for

pyramidal slip and 0.87 GPa (18R) 0.86 GPa (14H) for basal slip. Since the proposed hardness

model considers individual slip systems due to the lack of the slip system ratio information, and

also the indentation hardness measurement is orientation dependent property that affects the ratio

of the active slip systems, the predicted LPSOs’ hardness would rather give the ranges of the

hardness based on individual slip systems such as basal and pyramidal slips. For example, the

model with basal slip is the minimum hardness and that with pyramidal slip is the maximum

hardness. This trend agrees well with the experimental values from literatures[435]–[440] as shown

130

in Figure 7.3. Furthermore, the hardness of LPSO phases is 2.5 (basal) to 6 (pyramidal) times larger

for 18R and 2.5(basal) to 4.5 (pyramidal) times larger for 14H than that of HCP Mg[291].

Figure 7.3 Hardness prediction of 18R and 14H LPSO phases. Expt. 1 to Expt. 6 are from

[435] (Expt. 1), [49] (Expt. 2), [437] (Expt. 3), [438] (Expt. 4), [439] (Expt. 5), [440] (Expt. 6),

and the hardness of polycrystalline Mg as a reference[291] (Expt. 7), respectively.

131

Chapter 8

Conclusions and Future Work

8.1 Conclusions

Chapter 2 investigates the effects of L12 clusters and interstitial atoms (Mg, Gd, and Al)

on the structural and elastic properties of the Mg-Gd-Al LPSO phases via first-principles

calculations. Key conclusions can be summarized as follows:

(i) The number of layers in the SB affects the cluster densities along [0001] direction,

and this results in the changes of C33 and E[0001].

(ii) The size of L12 clusters is a key lattice feature to determine the C11 and C66.

(iii) For a SB with the same number of fault layers, an L12 cluster with an interstitial

atom in the LPSO increases the bulk modulus, Young’s modulus, and shear

modulus.

(iv) Effects of the number of fault layers in SB on elastic properties, such as bulk

modulus, are traceable to the redistribution of differential charge densities caused

by alloying elements.

In Chapter 3, thermodynamic properties of the 10H, 14H, 18R, and 24R LPSO phases in

the Mg-Al-Gd ternary system are predicted by first-principles calculations and modeled via the

CALPHAD method. The LPSO phases are modelled by four sublattices to capture the L12 cluster

embedded in the FCC stacking layers, including the atomic occupancy in the center interstitial site

within the cluster. Thermodynamic properties of the LPSO endmembers are obtained through the

quasiharmonic phonon and Debye models from first-principles calculations. It is observed that the

pure Mg endmembers of the LPSO phases are less stable with respect to HCP Mg; the occupancy

132

of the interstitial site by atoms are energetically favorable; and the mixing in the interstitial site is

nearly ideal. The presently thermodynamic description of the LPSO phases reproduces well the

phase equilibria reported in the literature, for example, the 18R is stable at high temperatures, the

14H is stable at low temperatures, and the 10H and 24R phases are not stable in the Mg-Ga-Al

system.

In Chapter 4, a new hardness model applicable to both ductile/brittle materials is developed

through the consideration of both elastic and plastic deformations of materials for the first time. It

incorporates the Peierls-Nabarro flow stress, dislocation width, Burgers vector, and slip plane

spacing in addition to the shear and bulk moduli. The model is based on the fundamental

understanding of elastic and plastic deformations during the indentation experiment, and the two

model parameters which are evaluated from the ratio of total and plastic indentation depths and the

experimental hardness values reported in the literature. In addition to hardness, the present model

can give insights on the possible dominant active slip systems by comparing the predicted hardness

values of various slip systems. The present hardness model provides a long-missing capability in

quantitatively predicting the mechanical properties of materials and future development of

ultrahard materials.

In Chapter 5, the unified temperature dependent hardness model for Vickers hardness is

proposed based on the dislocation width as a function of temperature. The dislocation width is truly

affected by the materials’ deformation diffusion mechanisms which include dislocation diffusion

and self-diffusion. This temperature dependent hardness model will help to predict materials

hardness as a function of temperature as well as other mechanical properties such as flow stress.

This model will accelerate materials design for mechanical properties as a function of temperature.

In Chapter 6, the twin bilayer dependent Vickers hardness model is proposed based on the

dislocation width as a function of twin bilayer. The dislocation width is affected by the twin

boundaries. Importantly the presence of the twin boundaries will accelerate the activation of the

133

screw dislocations in FCC metals which is generated from the partial dislocation of 1

6[112]. This

twin bilayer dependent hardness model will help to predict materials hardness as a function of twin

bilayer as well as other mechanical properties such as flow stress. This model will accelerate

materials design for mechanical properties.

In Chapter 7, the Vickers hardness models developed by Chapter 4, Chapter 5 and Chapter

6 are applied to 18R and 14H LPSO phases in Mg-Al-Gd system. Basal slip and pyramidal slip are

applied to this hardness model for LPSO phases. This model agrees well with various experimental

results.

8.2 Future Work

(1) Determination of the active slip systems in various crystal structures from first-

principles calculations.

(2) Modeling hardness as a function of the grain size by taking considerations of grain

misorientation angles and other slip system mechanisms.

(3) Modeling of LPSO hardness as a function of temperature by calculations of

diffusion activation energies.

134

Appendix A

Complete Elastic Stiffness Matrixes of 10H, 18R and 24R LPSO Phases

Type System Int. C11 C22 C33 C12 C13 C23 C44 C55 C66 C15 C25 C35 C46 B G E 𝜈 Ref.

10H

Mg-Gd-Al no 74.9 76.4 87.4 27.7 17.6 17.4 24.8 23.7 22.4 0 0 0 0 40.8 25.4 63.1 0.239 This work

Mg-Gd-Al Al 79.9 80.3 91.6 28.9 17.5 17.6 25.2 25.1 26.0 0 0 0 0 42.1 27.5 67.8 0.230 This work

Mg-Gd-Al Mg 78.8 79.0 90.4 28.8 17.5 17.5 23.7 23.8 24.6 0 0 0 0 41.9 26.3 65.2 0.237 This work

Mg-Gd-Al Gd 75.6 75.3 86.8 29.5 19.4 18.8 24.1 23.7 19.7 0 0 0 0 42.2 23.9 60.4 0.256 This work

18R

Mg-Gd-Al no 72.8 73.8 84.3 27.6 16.9 13.7 23.8 26.7 22.0 3.8 -4.1 0.6 -1.2 38.8 25.5 62.8 0.226 This work

Mg-Gd-Al Al 78.0 79.5 89.0 26.7 18.9 17.2 26.0 27.0 25.1 1.9 -2.6 0.8 -2.2 41.6 27.5 67.6 0.227 This work

Mg-Gd-Al Mg 77.1 78.1 88.9 27.2 18.5 17.1 26.3 27.4 23.8 3.3 -4.4 1.6 -2.1 41.4 27.1 66.8 0.228 This work

Mg-Gd-Al Gd 76.0 75.7 87.9 25.4 18.3 25.7 23.7 25.1 22.0 3.0 -5.1 1.1 -1.3 41.3 25.3 62.9 0.242 This work

Mg-Y-Zn no 70.4 70.4 85.3 30.1 19.8 19.0 21.7 24.1 20.0 -3.3 2.1 0.9 1.5 40.5 23.2 58.5 0.256 This work

Mg-Y-Zn Zn 70.5 69.1 84.6 32.4 18.5 20.4 20.3 23.3 19.4 -4.0 2.7 1.1 4.1 40.6 22.5 56.9 0.263 This work

Mg-Y-Zn Mg 71.6 69.6 85.4 32.3 17.9 20.3 21.7 24.2 20.2 -3.8 3.2 1.2 4.1 40.6 23.4 58.9 0.256 This work

Mg-Y-Zn Y 71.6 69.7 84.3 30.4 19.5 19.9 22.6 23.1 18.4 -2.2 -0.4 2.0 3.9 41.0 22.5 57.1 0.263 This work

Mg-Y-Zn 72.5

±0.7

80.0

±1.8 -

18.9

±1.1

23.5

±0.3

21.2

±0.3 - -

73.0 ±1.9 58.4 ±0.3

Exp. [26]

no 71.6 70.6 82.0 28.7 19.7 19.6 23.2 22.9 21.1 -2.5 2.3 0.5 1.7 Calc. [26]

135

Zn 72.3 73.4 84.2 28.8 18.8 18.8 24.2 24.2 21.5 -3.4 3.2 -0.1 2.6 Calc. [26]

Mg 71.9 72.7 85.2 29.1 18.8 18.7 24.8 24.7 21.8 -3.8 3.5 0.1 2.6 Calc. [26]

Y 72.7 73.2 83.8 28.1 19.4 18.9 24.6 24.3 22.1 -3.4 3.3 0.3 2.0 Calc. [26]

Mg-Y-Zn - - - - - - - - 66.7 ±4.9

Exp. [49]

Mg-Y-Zn 67.7 ±1.0

72.9 ±2.0

28.3 ±1.1

19.5 ±0.8

21.5 ±0.3

19.7 ±0.3

38.0 ±0.7

65.0 ±1.4 54.0 ±0.6

Exp. RT

[53]

Mg-Y-Zn 68.1 ±1.0

67.2 ±0.9

21.6 ±0.7

24.0 ±0.8

20.6 ±0.2

23.2 ±0.2

- 21.8 ±0.1

54.9 ±0.4

Exp. RT

[53] Mg-Y-Zn no 71.6 70.6 82.0 28.7 19.7 19.6 23.2 22.9 21.1 -2.5 2.3 0.5 1.7 Calc. [53]

24R

Mg-Gd-Al no 73.1 73.1 82.9 24.5 16.1 15.3 24.9 24.3 21.3 2.8 -2.5 0.2 -1.3 38.5 24.9 61.5 0.229 This work

Mg-Gd-Al Al 77.7 77.0 85.1 25.5 17.1 18.1 25.1 25.3 25.0 1.9 -1.5 -0.2 -1.1 40.4 26.7 65.6 0.227 This work

Mg-Gd-Al Mg 76.2 74.6 85.7 27.4 15.5 18.8 26.0 27.1 24.0 3.1 -2.8 -0.1 -0.5 40.0 26.8 65.8 0.224 This work

Mg-Gd-Al Gd 74.1 71.7 86.6 28.8 15.4 18.9 27.8 28.5 20.4 2.8 -1.2 -1.7 -0.5 40.2 25.9 63.9 0.230 This work

136

Appendix B

Thermo-Calc Mg-Al-Gd Database

$ *********************

$ File name: MgGdAlwithLPSOphases.TDB

$ *********************

$-----------------------------------------------------------------------------

$ Thermodynamic database of the Mg-Gd-Al system, modeled by Hongyeun Kim

$ + Phases Research Lab

$ Department of Materials Science and Engineering

$ The Pennsylvania State University

$ Last update: May 20, 2018

$-----------------------------------------------------------------------------

$ the model parameters of LPSO phases are mole-formula based

$ while those in the manuscript are mole-atom based

$

ELEMENT /- ELECTRON_GAS 0.0000E+00 0.0000E+00 0.0000E+00!

ELEMENT VA VACUUM 0.0000E+00 0.0000E+00 0.0000E+00!

ELEMENT AL FCC_A1 2.6982E+01 4.5773E+03 2.8322E+01!

ELEMENT GD HCP_A3 1.5725E+02 0.0000E+00 0.0000E+00!

137

ELEMENT MG HCP_A3 2.4305E+01 4.9980E+03 3.2671E+01!

FUNCTION GHSERAL 298.15 -7976.15+137.093038*T-24.3671976*T*LN(T)

-.001884662*T**2-8.77664E-07*T**3+74092*T**(-1); 700 Y

-11276.24+223.048446*T-38.5844296*T*LN(T)+.018531982*T**2

-5.764227E-06*T**3+74092*T**(-1); 933.47 Y

-11278.378+188.684153*T-31.748192*T*LN(T)-1.230524E+28*T**(-9); 2900 N

!

FUNCTION GHSERGD 200 -6834.5855+97.13101*T-24.7214131*T*LN(T)

-.00285240521*T**2-3.14674076E-07*T**3-8665.73348*T**(-1); 1000 Y

-6483.25362+95.6919924*T-24.6598297*T*LN(T)-.00185225011*T**2

-6.61211607E-07*T**3; 1508.15 Y

-123124.992+699.125537*T-101.800197*T*LN(T)+.0150644246*T**2

-6.39165948E-07*T**3+29356890.3*T**(-1); 3600 N !

FUNCTION GHSERMG 298.15 -8367.34+143.677875*T-26.1849782*T*LN(T)

+4.858E-04*T**2-1.393669E-06*T**3+78950*T**(-1); 923 Y

-14130.185+204.718543*T-34.3088*T*LN(T)+1.038192E+28*T**(-9); 3000 N !

FUNCTION GA12GDMG 298.15 -50000+50*T; 3000 N !

FUNCTION GA12MGAL 298.15 +97875-101.5*T; 3000 N !

FUNCTION GA12MG 298.15 +133469-87.319*T; 3000 N !

FUNCTION GA12ALMG 298.15 -52780-50.75*T; 3000 N !

FUNCTION GALBCC 298.15 +10083-4.813*T+GHSERAL#; 6000 N !

FUNCTION GGDFCC 298.15 +1800-.1*R#*T+GHSERGD#; 3600 N !

FUNCTION GALLAV 298.15 +15000+3*GHSERAL#; 3000 N !

138

FUNCTION GGDLAV 298.15 +15000+3*GHSERGD#; 3000 N !

FUNCTION GC15ALGD 298.15 -160000+32.64*T; 3000 N !

FUNCTION GC15GDMG 298.15 -45000+9.972*T; 3000 N !

FUNCTION GMGLAV 298.15 +15000+3*GHSERMG#; 3000 N !

FUNCTION GC36ALGD 298.15 +GC15ALGD#+13000; 3000 N !

FUNCTION GC36GDMG 298.15 +GC15GDMG#+15000; 3000 N !

FUNCTION GALLIQ 298.15 +3028.879+125.251171*T-24.3671976*T*LN(T)

-.001884662*T**2-8.77664E-07*T**3+74092*T**(-1)+7.9337E-20*T**7; 700 Y

-271.21+211.206579*T-38.5844296*T*LN(T)+.018531982*T**2

-5.764227E-06*T**3+74092*T**(-1)+7.9337E-20*T**7; 933.47 Y

-795.996+177.430178*T-31.748192*T*LN(T); 2900 N !

FUNCTION GGDLIQ 100 +6225.4407+88.8092103*T-24.7214131*T*LN(T)

-.00285240521*T**2-3.14674076E-07*T**3-8665.73348*T**(-1); 1000 Y

+146262.037-1208.70685*T+159.352082*T*LN(T)-.108247135*T**2

+1.06945505E-05*T**3-19678357*T**(-1); 1508.15 Y

-5397.314+192.336215*T-38.5075*T*LN(T); 3600 N !

FUNCTION GMGLIQ 298.15 +8202.24-8.83693*T-8.01759E-20*T**7+GHSERMG#; 923

Y

+8690.32-9.39216*T-1.03819E+28*T**(-9)+GHSERMG#; 6000 N !

FUNCTION UN_ASS 298.15 +0.0; 300 N !

TYPE_DEFINITION % SEQ *!

DEFINE_SYSTEM_DEFAULT ELEMENT 2 !

DEFAULT_COMMAND DEF_SYS_ELEMENT VA /- !

139

PHASE AL2GD3 % 2 2 3 !

CONSTITUENT AL2GD3 :AL : GD : !

PARAMETER G(AL2GD3,AL:GD;0) 298.15 -220000+55.89*T+2*GHSERAL#

+3*GHSERGD#; 3000 N REF0 !

PHASE AL3GD % 2 3 1 !

CONSTITUENT AL3GD :AL : GD : !

PARAMETER G(AL3GD,AL:GD;0) 298.15 -165000+29.61*T+3*GHSERAL#

+GHSERGD#; 3000 N REF0 !

PHASE ALGD % 2 1 1 !

CONSTITUENT ALGD :AL : GD : !

PARAMETER G(ALGD,AL:GD;0) 298.15 -98000+23*T+GHSERAL#

+GHSERGD#; 3000 N REF0 !

PHASE ALGD2 % 2 1 2 !

CONSTITUENT ALGD2 :AL : GD : !

140

PARAMETER G(ALGD2,AL:GD;0) 298.15 -115500+30.4*T+GHSERAL#

+2*GHSERGD#; 3000 N REF0 !

PHASE ALMG_BETA % 2 140 89 !

CONSTITUENT ALMG_BETA :AL : MG : !

PARAMETER G(ALMG_BETA,AL:MG;0) 298.15 -803385+105.238*T

+140*GHSERAL#+89*GHSERMG#; 6000 N REF0 !

PHASE ALMG_EPSILON % 2 30 23 !

CONSTITUENT ALMG_EPSILON :AL : MG : !

PARAMETER G(ALMG_EPSILON,AL:MG;0) 298.15 -170832-8.047*T+30*GHSERAL#

+23*GHSERMG#; 6000 N REF0 !

PHASE ALMG_GAMMA % 3 5 12 12 !

CONSTITUENT ALMG_GAMMA :GD,MG : AL,MG : AL,MG : !

PARAMETER G(ALMG_GAMMA,GD:AL:AL;0) 298.15 +GA12GDMG#+GA12MGAL#

-GA12MG#; 3000 N REF0 !

PARAMETER G(ALMG_GAMMA,MG:AL:AL;0) 298.15 +8360+20.338857*T

+5*GHSERMG#+24*GHSERAL#; 6000 N REF0 !

141

PARAMETER G(ALMG_GAMMA,GD:MG:AL;0) 298.15 +GA12GDMG#-GA12MG#

+GA12ALMG#; 3000 N REF0 !

PARAMETER G(ALMG_GAMMA,MG:MG:AL;0) 298.15 -104308.83+23.495281*T

+17*GHSERMG#+12*GHSERAL#; 6000 N REF0 !

PARAMETER G(ALMG_GAMMA,GD:AL:MG;0) 298.15 +GA12GDMG#+GA12MGAL#

-GA12ALMG#; 3000 N REF0 !

PARAMETER G(ALMG_GAMMA,MG:AL:MG;0) 298.15 +180556-138.069*T

+17*GHSERMG#+12*GHSERAL#; 6000 N REF0 !

PARAMETER G(ALMG_GAMMA,GD:MG:MG;0) 298.15 +GA12GDMG#; 3000 N

REF0 !

PARAMETER G(ALMG_GAMMA,MG:MG:MG;0) 298.15 +139371-87.319*T

+29*GHSERMG#; 6000 N REF0 !

PARAMETER G(ALMG_GAMMA,MG:AL:AL,MG;0) 298.15 +113100-14.5*T; 6000 N

REF0 !

PARAMETER G(ALMG_GAMMA,MG:MG:AL,MG;0) 298.15 +113100-14.5*T; 6000 N

REF0 !

TYPE_DEFINITION & GES A_P_D BCC_A2 MAGNETIC -1.0 4.00000E-01 !

PHASE BCC_A2 %& 2 1 3 !

CONSTITUENT BCC_A2 :AL,GD,MG : VA : !

PARAMETER G(BCC_A2,AL:VA;0) 298.15 +GALBCC#; 3000 N REF0 !

PARAMETER G(BCC_A2,GD:VA;0) 100 -3600.77684+95.0191641*T

-24.7214131*T*LN(T)-.00285240521*T**2-3.14674076E-07*T**3

142

-8665.73348*T**(-1); 1000 Y

+152792.743-1349.58873*T+180.097094*T*LN(T)-.119550229*T**2

+1.17915728E-05*T**3-22038836*T**(-1); 1508.15 Y

-15783.7618+202.222057*T-38.960425*T*LN(T); 1586.15 Y

-19850.5562+224.817909*T-41.904333*T*LN(T)+8.58222759E-04*T**2

-3.77570269E-08*T**3+995428.573*T**(-1); 3600 N REF0 !

PARAMETER G(BCC_A2,MG:VA;0) 298.14 -5267.34+141.575547*T

-26.1849782*T*LN(T)+4.858E-04*T**2-1.393669E-06*T**3+78950*T**(-1); 923 Y

-11030.185+202.616215*T-34.3088*T*LN(T)+1.038192E+28*T**(-9); 3000 N REF0 !

PARAMETER G(BCC_A2,AL,GD:VA;0) 298.15 -80000+30.64*T; 6000 N REF0 !

PARAMETER G(BCC_A2,GD,MG:VA;0) 298.15 -45347.5+25.5692*T; 6000 N

REF0 !

PARAMETER G(BCC_A2,GD,MG:VA;1) 298.15 +10195+1.3355*T; 6000 N

REF0 !

PARAMETER G(BCC_A2,GD,MG:VA;2) 298.15 +3267.1-3.5551*T; 6000 N

REF0 !

PHASE FCC_A1 % 2 1 1 !

CONSTITUENT FCC_A1 :AL,GD,MG : VA : !

PARAMETER G(FCC_A1,AL:VA;0) 298.15 +GHSERAL#; 3000 N REF0 !

PARAMETER G(FCC_A1,GD:VA;0) 298.15 +GGDFCC#; 3000 N REF0 !

PARAMETER G(FCC_A1,MG:VA;0) 298.15 +2600-.9*T+GHSERMG#; 3000 N

REF0 !

143

PARAMETER G(FCC_A1,AL,MG:VA;0) 298.15 +1593+2.149*T; 6000 N REF0 !

PARAMETER G(FCC_A1,AL,MG:VA;1) 298.15 +1014-.66*T; 6000 N REF0 !

PARAMETER G(FCC_A1,AL,MG:VA;2) 298.15 -673; 6000 N REF0 !

PHASE GDMG3 % 2 1 3 !

CONSTITUENT GDMG3 :GD : MG : !

PARAMETER G(GDMG3,GD:MG;0) 298.15 -57000+13.0162*T+3*GHSERMG#

+GHSERGD#; 3000 N REF0 !

PHASE GDMG5 % 2 1 5 !

CONSTITUENT GDMG5 :GD : MG : !

PARAMETER G(GDMG5,GD:MG;0) 298.15 -60521.6+11.2668*T

+5*GHSERMG#+GHSERGD#; 3000 N REF0 !

TYPE_DEFINITION ' GES A_P_D HCP_A3 MAGNETIC -3.0 2.80000E-01 !

PHASE HCP_A3 %' 2 1 .5 !

CONSTITUENT HCP_A3 :AL,GD,MG : VA : !

PARAMETER G(HCP_A3,AL:VA;0) 298.15 +5481-1.8*T+GHSERAL#; 3000

N REF0 !

144

PARAMETER TC(HCP_A3,GD:VA;0) 200 +293.4; 3600 N REF0 !

PARAMETER BMAGN(HCP_A3,GD:VA;0) 200 +3; 3600 N REF0 !

PARAMETER G(HCP_A3,GD:VA;0) 298.15 +GHSERGD#; 3000 N REF0 !

PARAMETER G(HCP_A3,MG:VA;0) 298.15 +GHSERMG#; 3000 N REF0 !

PARAMETER G(HCP_A3,AL,MG:VA;0) 298.15 +4336-2.863*T; 6000 N REF0 !

PARAMETER G(HCP_A3,AL,MG:VA;1) 298.15 -449-.135*T; 6000 N REF0 !

PARAMETER G(HCP_A3,AL,MG:VA;2) 298.15 -1963; 6000 N REF0 !

PARAMETER G(HCP_A3,GD,MG:VA;0) 298.15 -33346.6+19.3451*T; 6000 N

REF0 !

PARAMETER G(HCP_A3,GD,MG:VA;1) 298.15 +13854; 6000 N REF0 !

PHASE LAV_C14 % 2 1 2 !

CONSTITUENT LAV_C14 :AL,GD : AL,GD : !

PARAMETER G(LAV_C14,AL:AL;0) 298.15 +GALLAV#; 3000 N REF0 !

PARA G(LAV_C14,GD:AL;0) 298.15 +0; 6000 N!

PARA G(LAV_C14,AL:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LAV_C14,GD:GD;0) 298.15 +GGDLAV#; 3000 N REF0 !

PHASE LAV_C15 % 2 1 2 !

CONSTITUENT LAV_C15 :AL,GD,MG : AL,GD,MG : !

PARAMETER G(LAV_C15,AL:AL;0) 298.15 +GALLAV#; 3000 N REF0 !

145

PARAMETER G(LAV_C15,GD:AL;0) 298.15 +GC15ALGD#+2*GHSERAL#

+GHSERGD#; 3000 N REF0 !

PARAMETER G(LAV_C15,MG:AL;0) 298.15 +15000+2*GHSERAL#+GHSERMG#;

3000 N REF0 !

PARAMETER G(LAV_C15,AL:GD;0) 298.15 -GC15ALGD#+30000+GHSERAL#

+2*GHSERGD#; 3000 N REF0 !

PARAMETER G(LAV_C15,GD:GD;0) 298.15 +GGDLAV#; 3000 N REF0 !

PARAMETER G(LAV_C15,MG:GD;0) 298.15 -GC15GDMG#+32000+2*GHSERGD#

+GHSERMG#; 3000 N REF0 !

PARAMETER G(LAV_C15,AL:MG;0) 298.15 +15000+GHSERAL#+2*GHSERMG#;

3000 N REF0 !

PARAMETER G(LAV_C15,GD:MG;0) 298.15 -50777+17.149*T+2*GHSERMG#

+GHSERGD#; 3000 N REF0 !

PARAMETER G(LAV_C15,MG:MG;0) 298.15 +GMGLAV#; 3000 N REF0 !

PARAMETER G(LAV_C15,GD:AL,MG;0) 298.15 +29500; 3000 N REF0 !

PARAMETER G(LAV_C15,GD:AL,MG;1) 298.15 -10000; 3000 N REF0 !

PHASE LAV_C36 % 2 1 2 !

CONSTITUENT LAV_C36 :AL,GD,MG : AL,GD,MG : !

PARAMETER G(LAV_C36,AL:AL;0) 298.15 +GALLAV#; 3000 N REF0 !

PARAMETER G(LAV_C36,GD:AL;0) 298.15 +GC36ALGD#+2*GHSERAL#

+GHSERGD#; 3000 N REF0 !

PARAMETER G(LAV_C36,MG:AL;0) 298.15 -500+2*GHSERAL#+GHSERMG#;

146

3000 N REF0 !

PARAMETER G(LAV_C36,AL:GD;0) 298.15 -GC36ALGD#+30000+GHSERAL#

+2*GHSERGD#; 3000 N REF0 !

PARAMETER G(LAV_C36,GD:GD;0) 298.15 +GGDLAV#; 3000 N REF0 !

PARAMETER G(LAV_C36,MG:GD;0) 298.15 -GC36GDMG#+30000+2*GHSERGD#

+GHSERMG#; 3000 N REF0 !

PARAMETER G(LAV_C36,AL:MG;0) 298.15 +30500+GHSERAL#+2*GHSERMG#;

3000 N REF0 !

PARAMETER G(LAV_C36,GD:MG;0) 298.15 +GC36GDMG#+GHSERGD#

+2*GHSERMG#; 3000 N REF0 !

PARAMETER G(LAV_C36,MG:MG;0) 298.15 +GMGLAV#; 3000 N REF0 !

PARAMETER G(LAV_C36,GD,MG:AL;0) 298.15 -45000+9*T; 3000 N REF0 !

PHASE LIQUID % 1 1.0 !

CONSTITUENT LIQUID :AL,GD,MG : !

PARAMETER G(LIQUID,AL;0) 298.15 +GALLIQ#; 3000 N REF0 !

PARAMETER G(LIQUID,GD;0) 298.15 +GGDLIQ#; 3000 N REF0 !

PARAMETER G(LIQUID,MG;0) 298.15 +GMGLIQ#; 3000 N REF0 !

PARAMETER G(LIQUID,AL,GD;0) 298.15 -166500+52.36*T; 6000 N

REF0 !

PARAMETER G(LIQUID,AL,GD;1) 298.15 -23790; 6000 N REF0 !

PARAMETER G(LIQUID,AL,GD;2) 298.15 +18520; 6000 N REF0 !

PARAMETER G(LIQUID,AL,GD,MG;0) 298.15 -20000; 6000 N REF0 !

147

PARAMETER G(LIQUID,AL,MG;0) 298.15 -9019+4.794*T; 6000 N REF0 !

PARAMETER G(LIQUID,AL,MG;1) 298.15 -1093+1.412*T; 6000 N REF0 !

PARAMETER G(LIQUID,AL,MG;2) 298.15 +494; 6000 N REF0 !

PARAMETER G(LIQUID,GD,MG;0) 298.15 -36681.3+16.2484*T; 6000 N

REF0 !

PARAMETER G(LIQUID,GD,MG;1) 298.15 +34233.8-10.7783*T; 6000 N

REF0 !

PARAMETER G(LIQUID,GD,MG;2) 298.15 -7352.9; 6000 N REF0 !

PHASE LPSO10H % 4 92 16 12 2 !

CONSTITUENT LPSO10H :MG : AL,GD,MG : AL,GD,MG : AL,GD,MG,VA : !

PARA G(LPSO10H,MG:AL:AL:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO10H,MG:GD:AL:AL;0) 298.15 -2354532.08+16774.99*T

-3070.98*T*LN(T)-.2397*T**2+9194000*T**(-1)-3.106E-05*T**3;

3000 N REF0 !

PARA G(LPSO10H,MG:MG:AL:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:AL:GD;0) 298.15 +0; 6000 N!

148

PARAMETER G(LPSO10H,MG:GD:AL:GD;0) 298.15 -2365966.81+16331.7*T

-3055.52*T*LN(T)-.2668*T**2+7265000*T**(-1)-2.528E-05*T**3;

3000 N REF0 !

PARA G(LPSO10H,MG:MG:AL:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:AL:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO10H,MG:GD:AL:MG;0) 298.15 -2360302.63+16684.48*T

-3065.66*T*LN(T)-.2439*T**2+8772000*T**(-1)-2.815E-05*T**3;

3000 N REF0 !

PARA G(LPSO10H,MG:MG:AL:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:AL:MG:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:GD:MG:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO10H,MG:MG:MG:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO10H,MG:AL:AL:VA;0) 298.15 -782715.93+17456.09*T

-3061.63*T*LN(T)-.2279*T**2+12980000*T**(-1)-6.045E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:GD:AL:VA;0) 298.15 -1955859.12+17048.53*T

149

-3026.71*T*LN(T)-.2572*T**2+12450000*T**(-1)-3.906E-05*T**3;

3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:AL:VA;0) 298.15 -877620.27+18263.01*T

-3063.8*T*LN(T)-.2318*T**2+6751000*T**(-1)-4.928E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:AL:GD:VA;0) 298.15 -1194764.44+15633.3*T

-3031.19*T*LN(T)-.2458*T**2+4779000*T**(-1)-3.293E-05*T**3;

3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:GD:VA;0) 298.15 -1087653.56+15437.85*T

-3025.65*T*LN(T)-.2935*T**2+4279000*T**(-1)-3.703E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:MG:GD:VA;0) 298.15 -910956.4+17993.29*T

-3021.13*T*LN(T)-.2888*T**2+3392000*T**(-1)-3.015E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:AL:MG:VA;0) 298.15 -923009.74+16979.14*T

-3027.52*T*LN(T)-.2475*T**2+10570000*T**(-1)-4.14E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:GD:MG:VA;0) 298.15 -1337554.87+15898.4*T

-3002.15*T*LN(T)-.2862*T**2+5720000*T**(-1)-2.227E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO10H,MG:MG:MG:VA;0) 298.15 -1067584.41+19470.57*T

-3516.43*T*LN(T)+.612039*T**2+15152000*T**(-1)-2.483E-04*T**3; 3000

N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:AL:VA;0) 298.15 -136647; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:AL:VA;1) 298.15 +671235; 3000 N REF0 !

150

PARAMETER G(LPSO10H,MG:AL,MG:AL:VA;0) 298.15 -2177; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,MG:AL:VA;1) 298.15 +111307; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:AL,GD:VA;0) 298.15 -24208; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:AL,GD:VA;1) 298.15 -912856; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:AL,MG:VA;0) 298.15 -2825; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:AL,MG:VA;1) 298.15 +6782; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:AL:VA;0) 298.15 +154409; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:AL:VA;1) 298.15 +191805; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:AL,GD:VA;0) 298.15 +1095340; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:AL,GD:VA;1) 298.15 -670940; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:AL,MG:VA;0) 298.15 +859085; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:AL,MG:VA;1) 298.15 -725095; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:AL,GD:VA;0) 298.15 +78596; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:AL,GD:VA;1) 298.15 -626763; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:AL,MG:VA;0) 298.15 -14320; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:AL,MG:VA;1) 298.15 -46645; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:GD:VA;0) 298.15 -49024; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:GD:VA;1) 298.15 +259459; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,MG:GD:VA;0) 298.15 +56565; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,MG:GD:VA;1) 298.15 +12235; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:GD,MG:VA;0) 298.15 +35675; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL:GD,MG:VA;1) 298.15 -436853; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:GD:VA;0) 298.15 +172246; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:GD:VA;1) 298.15 +81486; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD:GD,MG:VA;0) 298.15 -163698; 3000 N REF0 !

151

PARAMETER G(LPSO10H,MG:GD:GD,MG:VA;1) 298.15 +83590; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:GD,MG:VA;0) 298.15 -75509; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:MG:GD,MG:VA;1) 298.15 -181093; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:MG:VA;0) 298.15 -68913; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,GD:MG:VA;1) 298.15 +654423; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,MG:MG:VA;0) 298.15 -30388; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:AL,MG:MG:VA;1) 298.15 +41340; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:MG:VA;0) 298.15 -101085; 3000 N REF0 !

PARAMETER G(LPSO10H,MG:GD,MG:MG:VA;1) 298.15 +243856; 3000 N REF0 !

PHASE LPSO14H % 4 140 16 12 2 !

CONSTITUENT LPSO14H :MG : AL,GD,MG : AL,GD,MG : AL,GD,MG,VA : !

PARA G(LPSO14H,MG:AL:AL:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:AL:AL;0) 298.15 -2786149.21+23028.16*T

-4275.39*T*LN(T)-.3339*T**2+13680000*T**(-1)-4.94E-05*T**3;

3000 N REF0 !

PARA G(LPSO14H,MG:MG:AL:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:GD:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:MG:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:MG:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:MG:AL;0) 298.15 -1793340.98+22721.72*T

-4268.55*T*LN(T)-.3833*T**2+11510000*T**(-1)-4.049E-05*T**3; 6000 N REF0 !

152

PARA G(LPSO14H,MG:MG:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:AL:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:AL:GD;0) 298.15 -2813548.65+22801.14*T

-4276.6*T*LN(T)-.3373*T**2+12620000*T**(-1)-4.709E-05*T**3;

3000 N REF0 !

PARA G(LPSO14H,MG:MG:AL:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:GD:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:MG:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:MG:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:MG:GD;0) 298.15 -1630377.48+22568*T

-4273.96*T*LN(T)-.3773*T**2+10830000*T**(-1)-3.994E-05*T**3; 6000 N REF0 !

PARA G(LPSO14H,MG:MG:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:AL:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:AL:MG;0) 298.15 -2799969.07+22945.45*T

-4277.98*T*LN(T)-.3293*T**2+13190000*T**(-1)-4.828E-05*T**3;

3000 N REF0 !

PARA G(LPSO14H,MG:MG:AL:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:GD:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:MG:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO14H,MG:AL:MG:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO14H,MG:GD:MG:MG;0) 298.15 -1741000.25+22702.32*T

-4274.51*T*LN(T)-.379*T**2+11310000*T**(-1)-4.046E-05*T**3; 6000 N REF0 !

PARA G(LPSO14H,MG:MG:MG:MG;0) 298.15 +0; 6000 N!

153

PARAMETER G(LPSO14H,MG:AL:AL:VA;0) 298.15 -1155694.45+24155.39*T

-4279.5*T*LN(T)-.3259*T**2+18370000*T**(-1)-7.698E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:GD:AL:VA;0) 298.15 -2420244.24+22742.55*T

-4234.51*T*LN(T)-.3454*T**2+13370000*T**(-1)-5.397E-05*T**3;

3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:AL:VA;0) 298.15 -1262380.29+23455.24*T

-4275.57*T*LN(T)-.3317*T**2+14180000*T**(-1)-6.633E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:AL:GD:VA;0) 298.15 -1616730.58+22587.99*T

-4252.14*T*LN(T)-.3323*T**2+11050000*T**(-1)-5.01E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:GD:GD:VA;0) 298.15 -1410515.73+21426.21*T

-4201.07*T*LN(T)-.4069*T**2+8064000*T**(-1)-3.667E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:MG:GD:VA;0) 298.15 -1340489.73+22148.91*T

-4244.11*T*LN(T)-.3631*T**2+9456000*T**(-1)-4.961E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:AL:MG:VA;0) 298.15 -1307892.46+23724.09*T

-4239.32*T*LN(T)-.3504*T**2+16650000*T**(-1)-5.953E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:GD:MG:VA;0) 298.15 -1711823.78+22404.3*T

-4230.22*T*LN(T)-.3693*T**2+10930000*T**(-1)-4.105E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO14H,MG:MG:MG:VA;0) 298.15 -1475214.62+26504.93*T

154

-4796.04*T*LN(T)+.670536*T**2+19341800*T**(-1)-3.04561E-04*T**3;

3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:AL:VA;0) 298.15 +123091; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:AL:VA;1) 298.15 +754417; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:AL:VA;0) 298.15 +211429; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:AL:VA;1) 298.15 -71145; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL:AL,GD:VA;0) 298.15 +120491; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL:AL,GD:VA;1) 298.15 -997608; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL:AL,MG:VA;0) 298.15 +100358; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL:AL,MG:VA;1) 298.15 -116095; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:AL:VA;0) 298.15 -342208; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:AL:VA;1) 298.15 +583484; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:AL,GD:VA;0) 298.15 +836065; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:AL,GD:VA;1) 298.15 -509455; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:AL,MG:VA;0) 298.15 +621910; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:AL,MG:VA;1) 298.15 -540640; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:AL,GD:VA;0) 298.15 +93659; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:AL,GD:VA;1) 298.15 -678979; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:AL,MG:VA;0) 298.15 -5754; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:AL,MG:VA;1) 298.15 -65837; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:GD:VA;0) 298.15 +27178; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:GD:VA;1) 298.15 +200656; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:GD:VA;0) 298.15 +41584; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:GD:VA;1) 298.15 -45712; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL:GD,MG:VA;0) 298.15 +34281; 3000 N REF0 !

155

PARAMETER G(LPSO14H,MG:AL:GD,MG:VA;1) 298.15 -450286; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:GD:VA;0) 298.15 -13145; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:GD:VA;1) 298.15 +56621; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:GD,MG:VA;0) 298.15 -234097; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD:GD,MG:VA;1) 298.15 -6258; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:GD,MG:VA;0) 298.15 -22621; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:MG:GD,MG:VA;1) 298.15 -164212; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:MG:VA;0) 298.15 -166652; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,GD:MG:VA;1) 298.15 +473551; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:MG:VA;0) 298.15 -13587; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:AL,MG:MG:VA;1) 298.15 +19826; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:MG:VA;0) 298.15 +31690; 3000 N REF0 !

PARAMETER G(LPSO14H,MG:GD,MG:MG:VA;1) 298.15 +142429; 3000 N REF0 !

PHASE LPSO18R % 4 116 16 12 2 !

CONSTITUENT LPSO18R :MG : AL,GD,MG : AL,GD,MG : AL,GD,MG,VA : !

PARA G(LPSO18R,MG:AL:AL:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:AL:AL;0) 298.15 -2586985.99+19691.21*T

-3673.73*T*LN(T)-.2831*T**2+11520000*T**(-1)-4.11E-05*T**3;

3000 N REF0 !

PARA G(LPSO18R,MG:MG:AL:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:GD:GD:AL;0) 298.15 +0; 6000 N!

156

PARA G(LPSO18R,MG:MG:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:MG:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:MG:AL;0) 298.15 +0; 6000 N REF0 !

PARA G(LPSO18R,MG:MG:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:AL:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:AL:GD;0) 298.15 -2608111.07+19592.82*T

-3698.42253*T*LN(T)-.2755*T**2+10790000*T**(-1)-4.403E-05*T**3

; 3000 N REF0 !

PARA G(LPSO18R,MG:MG:AL:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:GD:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:MG:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:MG:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:MG:GD;0) 298.15 +0; 6000 N REF0 !

PARA G(LPSO18R,MG:MG:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:AL:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:AL:MG;0) 298.15 -2595446.79+19562.46*T

-3655.39*T*LN(T)-.2375*T**2+11240000*T**(-1)-2.451E-05*T**3;

3000 N REF0 !

PARA G(LPSO18R,MG:MG:AL:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:GD:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:MG:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO18R,MG:AL:MG:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:GD:MG:MG;0) 298.15 +0; 6000 N REF0 !

157

PARA G(LPSO18R,MG:MG:MG:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO18R,MG:AL:AL:VA;0) 298.15 -976005.86+20793.54*T

-3664.53*T*LN(T)-.2801*T**2+16450000*T**(-1)-6.836E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:GD:AL:VA;0) 298.15 -2178485.73+19383.18*T

-3641.7*T*LN(T)-.2841*T**2+10940000*T**(-1)-4.37E-05*T**3;

3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:AL:VA;0) 298.15 -1067274.12+20448.04*T

-3662.62*T*LN(T)-.2804*T**2+14320000*T**(-1)-5.826E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:AL:GD:VA;0) 298.15 -1418530.48+19508.42*T

-3632.06*T*LN(T)-.2916*T**2+10410000*T**(-1)-4.107E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:GD:GD:VA;0) 298.15 -1223440.54+18404.78*T

-3607.33*T*LN(T)-.3505*T**2+7236000*T**(-1)-2.967E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:MG:GD:VA;0) 298.15 -1150175.91+19284*T

-3637.79*T*LN(T)-.3133*T**2+9482000*T**(-1)-4.093E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:AL:MG:VA;0) 298.15 -1109648.79+20491.61*T

-3632.81*T*LN(T)-.2985*T**2+15380000*T**(-1)-5.076E-05*T**3; 3000 N

REF0 !

PARAMETER G(LPSO18R,MG:GD:MG:VA;0) 298.15 -1513376.25+19248.82*T

-3627.28*T*LN(T)-.3246*T**2+9746000*T**(-1)-3.023E-05*T**3; 3000 N

REF0 !

158

PARAMETER G(LPSO18R,MG:MG:MG:VA;0) 298.15 -1141733.28+20351.3*T

-3723.22*T*LN(T)+.00521359*T**2+10727300*T**(-1)-1.11855E-04*T**3;

3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:AL:VA;0) 298.15 -162225; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:AL:VA;1) 298.15 +750803; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:AL:VA;0) 298.15 +31241; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:AL:VA;1) 298.15 +11691; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL:AL,GD:VA;0) 298.15 +8821; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL:AL,GD:VA;1) 298.15 -862883; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL:AL,MG:VA;0) 298.15 +14109; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL:AL,MG:VA;1) 298.15 -18144; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:AL:VA;0) 298.15 -253010; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:AL:VA;1) 298.15 +610629; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:AL,GD:VA;0) 298.15 +848010; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:AL,GD:VA;1) 298.15 -478210; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:AL,MG:VA;0) 298.15 +669930; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:AL,MG:VA;1) 298.15 -574230; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:AL,GD:VA;0) 298.15 +69656; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:AL,GD:VA;1) 298.15 -663374; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:AL,MG:VA;0) 298.15 -6153; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:AL,MG:VA;1) 298.15 -29366; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:GD:VA;0) 298.15 -311046; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:GD:VA;1) 298.15 +442561; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:GD:VA;0) 298.15 -84259; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:GD:VA;1) 298.15 -21399; 3000 N REF0 !

159

PARAMETER G(LPSO18R,MG:AL:GD,MG:VA;0) 298.15 +27872; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL:GD,MG:VA;1) 298.15 -435540; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:GD:VA;0) 298.15 +138374; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:GD:VA;1) 298.15 +157546; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:GD,MG:VA;0) 298.15 -234342; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD:GD,MG:VA;1) 298.15 +9435; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:GD,MG:VA;0) 298.15 +2960; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:MG:GD,MG:VA;1) 298.15 -174247; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:MG:VA;0) 298.15 -246037; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,GD:MG:VA;1) 298.15 +504558; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:MG:VA;0) 298.15 +30252; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:AL,MG:MG:VA;1) 298.15 +50907; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:MG:VA;0) 298.15 +116778; 3000 N REF0 !

PARAMETER G(LPSO18R,MG:GD,MG:MG:VA;1) 298.15 +164380; 3000 N REF0 !

PHASE LPSO24R % 4 164 16 12 2 !

CONSTITUENT LPSO24R :MG : AL,GD,MG : AL,GD,MG : AL,GD,MG,VA : !

PARA G(LPSO24R,MG:AL:AL:AL;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO24R,MG:GD:AL:AL;0) 298.15 -2988043.96+27088.766*T

-4880.482*T*LN(T)-.3896*T**2+16290000*T**(-1)-5.75E-05*T**3;

3000 N REF0 !

PARA G(LPSO24R,MG:MG:AL:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:GD:AL;0) 298.15 +0; 6000 N!

160

PARA G(LPSO24R,MG:GD:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:GD:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:GD:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:MG:AL;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:AL:GD;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO24R,MG:GD:AL:GD;0) 298.15 -3009628.46+26970.485*T

-4879.113*T*LN(T)-.3915*T**2+15690000*T**(-1)-4.904E-05*T**3;

3000 N REF0 !

PARA G(LPSO24R,MG:MG:AL:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:GD:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:GD:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:GD:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:MG:GD;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:AL:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO24R,MG:GD:AL:MG;0) 298.15 -3027048.32+27687.271*T

-4971.07*T*LN(T)-.2373*T**2+18170000*T**(-1)-8.271E-05*T**3;

3000 N REF0 !

PARA G(LPSO24R,MG:MG:AL:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:GD:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:GD:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:AL:MG:MG;0) 298.15 +0; 6000 N!

161

PARA G(LPSO24R,MG:GD:MG:MG;0) 298.15 +0; 6000 N!

PARA G(LPSO24R,MG:MG:MG:MG;0) 298.15 +0; 6000 N!

PARAMETER G(LPSO24R,MG:AL:AL:VA;0) 298.15 -1362092.82+27931.036*T

-4916.712*T*LN(T)-.3487*T**2+20220000*T**(-1)-9.255E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:GD:AL:VA;0) 298.15 -2595366.92+26963.741*T

-4881.001*T*LN(T)-.3574*T**2+16120000*T**(-1)-7.289E-05*T**3;

3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:AL:VA;0) 298.15 -1467578.63+27372.835*T

-4861.816*T*LN(T)-.3986*T**2+17880000*T**(-1)-7.313E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:AL:GD:VA;0) 298.15 -1819588.26+26682.013*T

-4848.967*T*LN(T)-.3853*T**2+14000000*T**(-1)-5.962E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:GD:GD:VA;0) 298.15 -1608789.53+25762.18*T

-4817.989*T*LN(T)-.4472*T**2+10370000*T**(-1)-4.612E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:MG:GD:VA;0) 298.15 -1548243.75+26466.342*T

-4844.667*T*LN(T)-.4165*T**2+13060000*T**(-1)-5.796E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:AL:MG:VA;0) 298.15 -1508509.73+27504.883*T

-4855.804*T*LN(T)-.3905*T**2+18860000*T**(-1)-6.985E-05*T**3; 3000

N REF0 !

PARAMETER G(LPSO24R,MG:GD:MG:VA;0) 298.15 -1916069.55+26551.645*T

-4853.711*T*LN(T)-.3992*T**2+13320000*T**(-1)-5.232E-05*T**3; 3000

162

N REF0 !

PARAMETER G(LPSO24R,MG:MG:MG:VA;0) 298.15 -1752358.44+31668.18*T

-5707.35*T*LN(T)+1.09164*T**2+25601000*T**(-1)-4.26788E-04*T**3; 3000 N

REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:AL:VA;0) 298.15 +1107315; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:AL:VA;1) 298.15 -315275; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,MG:AL:VA;0) 298.15 -18471; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,MG:AL:VA;1) 298.15 -1311081; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:AL,GD:VA;0) 298.15 -741605; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:AL,GD:VA;1) 298.15 -1461285; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:AL,MG:VA;0) 298.15 -577654; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:AL,MG:VA;1) 298.15 -462475; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:AL:VA;0) 298.15 -1070760; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:AL:VA;1) 298.15 -685841; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:AL,GD:VA;0) 298.15 +275680; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:AL,GD:VA;1) 298.15 -581375; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:AL,MG:VA;0) 298.15 +211582; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:AL,MG:VA;1) 298.15 -539226; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:AL,GD:VA;0) 298.15 +35991; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:AL,GD:VA;1) 298.15 -1319965; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:AL,MG:VA;0) 298.15 -83468; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:AL,MG:VA;1) 298.15 -962515; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:GD:VA;0) 298.15 +162278; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:GD:VA;1) 298.15 -1201362; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,MG:GD:VA;0) 298.15 +69761; 3000 N REF0 !

163

PARAMETER G(LPSO24R,MG:AL,MG:GD:VA;1) 298.15 -1441289; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:GD,MG:VA;0) 298.15 +333709; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL:GD,MG:VA;1) 298.15 -1536678; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:GD:VA;0) 298.15 -118402; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:GD:VA;1) 298.15 -1232962; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:GD,MG:VA;0) 298.15 -113859; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD:GD,MG:VA;1) 298.15 -489985; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:GD,MG:VA;0) 298.15 -59976; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:MG:GD,MG:VA;1) 298.15 -1064666; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:MG:VA;0) 298.15 +587530; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,GD:MG:VA;1) 298.15 -373951; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,MG:MG:VA;0) 298.15 -158384; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:AL,MG:MG:VA;1) 298.15 -1375874; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:MG:VA;0) 298.15 -736695; 3000 N REF0 !

PARAMETER G(LPSO24R,MG:GD,MG:MG:VA;1) 298.15 -760313; 3000 N REF0 !

PHASE MG2GD % 2 2 1 !

CONSTITUENT MG2GD :MG : GD : !

PARAMETER G(MG2GD,MG:GD;0) 298.15 +GHSERGD#+2*GHSERMG#-50777

+17.149*T; 6000 N REF0 !

PHASE MG41M5 % 2 41 5 !

164

CONSTITUENT MG41M5 :MG : GD : !

PARAMETER G(MG41M5,MG:GD;0) 298.15 +41*GHSERMG#+5*GHSERGD#

-302608+56.334*T+11485.6-9.988*T; 6000 N REF0 !

$ THIS PHASE HAS A DISORDERED CONTRIBUTION FROM BCC_A2

TYPE_DEFINITION ( GES AMEND_PHASE_DESCRIPTION MGM DIS_PART BCC_A2,,,!

PHASE MGM %( 3 .5 .5 1 !

CONSTITUENT MGM :AL,GD,MG : AL,GD,MG : VA : !

PARA G(MGM,AL:AL:VA;0) 298.15 +0; 6000 N!

PARAMETER G(MGM,GD:AL:VA;0) 298.15 -64000+12*T; 6000 N REF0 !

PARA G(MGM,MG:AL:VA;0) 298.15 +0; 6000 N!

PARAMETER G(MGM,AL:GD:VA;0) 298.15 -64000+12*T; 6000 N REF0 !

PARA G(MGM,GD:GD:VA;0) 298.15 +0; 6000 N!

PARAMETER G(MGM,MG:GD:VA;0) 298.15 -28695.4+6.1567*T; 6000 N

REF0 !

PARA G(MGM,AL:MG:VA;0) 298.15 +0; 6000 N!

PARAMETER G(MGM,GD:MG:VA;0) 298.15 -28695.4+6.1567*T; 6000 N

REF0 !

PARA G(MGM,MG:MG:VA;0) 298.15 +0; 6000 N!

PARAMETER G(MGM,GD:AL,MG:VA;0) 298.15 -50000+36*T; 6000 N REF0 !

PARAMETER G(MGM,AL,MG:GD:VA;0) 298.15 -50000+36*T; 6000 N REF0 !

PARAMETER G(MGM,GD,MG:GD:VA;0) 298.15 -15446.6; 6000 N REF0 !

165

PARAMETER G(MGM,GD,MG:GD:VA;1) 298.15 -16247.3; 6000 N REF0 !

PARAMETER G(MGM,GD:GD,MG:VA;0) 298.15 -15446.6; 6000 N REF0 !

PARAMETER G(MGM,GD:GD,MG:VA;1) 298.15 -16247.3; 6000 N REF0 !

LIST_OF_REFERENCES

NUMBER SOURCE

!

166

Bibliography

[1] G. S. Frankel, “Magnesium alloys: Ready for the road,” Nat Mater, vol. advance on, no.

12, pp. 1189–1190, Oct. 2015.

[2] M. K. Kulekci, “Magnesium and its alloys applications in automotive industry,” Int. J.

Adv. Manuf. Technol., vol. 39, no. 9–10, pp. 851–865, Nov. 2008.

[3] T. M. Pollock, “Weight Loss with Magnesium Alloys,” Science (80-. )., vol. 328, no.

5981, pp. 986–987, May 2010.

[4] Z. Wu and W. A. Curtin, “The origins of high hardening and low ductility in magnesium,”

Nature, vol. 526, no. 7571, pp. 62–67, Sep. 2015.

[5] W. Xu et al., “A high-specific-strength and corrosion-resistant magnesium alloy,” Nat.

Mater., no. October, pp. 1–8, 2015.

[6] K. Kishida, H. Yokobayashi, and H. Inui, “The most stable crystal structure and the

formation processes of an order-disorder (OD) intermetallic phase in the Mg–Al–Gd

ternary system,” Philos. Mag., vol. 93, no. 21, pp. 2826–2846, Jul. 2013.

[7] H. Yokobayashi, K. Kishida, H. Inui, M. Yamasaki, and Y. Kawamura, “Enrichment of

Gd and Al atoms in the quadruple close packed planes and their in-plane long-range

ordering in the long period stacking-ordered phase in the Mg-Al-Gd system,” Acta Mater.,

vol. 59, no. 19, pp. 7287–7299, Nov. 2011.

[8] H. Yokobayashi, K. Kishida, H. Inui, M. Yamasaki, and Y. Kawamura, “Structure

analysis of a long period stacking ordered phase in Mg-Al-Gd alloys,” in Materials

Research Society Symposium Proceedings, 2011, vol. 1295, no. 5, pp. 267–272.

[9] K. Kishida, H. Yokobayashi, H. Inui, M. Yamasaki, and Y. Kawamura, “The crystal

structure of the LPSO phase of the 14H-type in the Mg-Al-Gd alloy system,”

167

Intermetallics, vol. 31, pp. 55–64, 2012.

[10] Y. Kawamura, K. Hayashi, A. Inoue, and T. Masumoto, “Platform Science and

Technology for Advanced Magnesium Alloys. Rapidly Solidified Powder Metallurgy

Mg97Zn1Y2Alloys with Excellent Tensile Yield Strength above 600 MPa.,” Mater.

Trans., vol. 42, no. 7, pp. 1172–1176, Sep. 2001.

[11] K. Hagihara, N. Yokotani, and Y. Umakoshi, “Plastic deformation behavior of Mg 12

YZn with 18R long-period stacking ordered structure,” Intermetallics, vol. 18, pp. 267–

276, 2009.

[12] M. Yamasaki, K. Hagihara, S.-I. Inoue, J. P. Hadorn, and Y. Kawamura,

“Crystallographic classification of kink bands in an extruded Mg-Zn-Y alloy using

intragranular misorientation axis analysis,” Acta Mater., vol. 61, pp. 2065–2076, 2013.

[13] K. Hagihara et al., “Effect of long-period stacking ordered phase on mechanical properties

of Mg97Zn1Y2 extruded alloy,” Acta Mater., vol. 58, no. 19, pp. 6282–6293, 2010.

[14] J. E. Saal and C. Wolverton, “Thermodynamic stability of Mg-based ternary long-period

stacking ordered structures,” Acta Mater., vol. 68, pp. 325–338, Apr. 2014.

[15] J. N. Wang, “Prediction of Peierls stresses for different crystals,” Mater. Sci. Eng. A, vol.

206, no. 2, pp. 259–269, 1996.

[16] E. Orowan, “Problems of plastic gliding,” Proc. Phys. Soc., vol. 52, no. 1, p. 8, 1940.

[17] K. Kishida, H. Yokobayashi, H. Inui, M. Yamasaki, and Y. Kawamura, “The crystal

structure of the LPSO phase of the 14H-type in the Mg–Al–Gd alloy system,”

Intermetallics, vol. 31, pp. 55–64, Dec. 2012.

[18] K. Kishida, K. Nagai, A. Matsumoto, A. Yasuhara, and H. Inui, “Crystal structures of

highly-ordered long-period stacking-ordered phases with 18R, 14H and 10H-type stacking

sequences in the Mg–Zn–Y system,” Acta Mater., vol. 99, pp. 228–239, Oct. 2015.

[19] H. Yokobayashi, K. Kishida, H. Inui, M. Yamasaki, and Y. Kawamura, “Enrichment of

168

Gd and Al atoms in the quadruple close packed planes and their in-plane long-range

ordering in the long period stacking-ordered phase in the Mg–Al–Gd system,” Acta

Mater., vol. 59, no. 19, pp. 7287–7299, Nov. 2011.

[20] L. S. Ramsdell, “S’TUDIES ON SILICON CARBIDE,” Am. Mineral., vol. 32, no. 1–2, p.

64, 1947.

[21] Parthe E and K. YVON, “On the Crystal Chemistry of the Close Packed Transition Metal

Carbides, II. A Proposal for the Notation of the Different Crystal Structures,” Acta

Crystallogr. Sect. B, vol. B26, pp. 153–163, Jul. 1970.

[22] H. Kimizuka and S. Ogata, “Predicting Atomic Arrangement of Solute Clusters in Dilute

Mg Alloys,” Mater. Res. Lett., vol. 1, no. 4, pp. 213–219, 2013.

[23] D. Egusa and E. Abe, “The structure of long period stacking/order Mg–Zn–RE phases

with extended non-stoichiometry ranges,” Acta Mater., vol. 60, no. 1, pp. 166–178, Jan.

2012.

[24] S. Y. Ma, L. M. Liu, and S. Q. Wang, “The predominant role of Zn6Y9 cluster in the long

period stacking order structures of Mg-Zn-Y alloys: a first-principles study,” J. Mater.

Sci., vol. 48, no. 4, pp. 1407–1412, 2013.

[25] H. Kimizuka, M. Fronzi, and S. Ogata, “Effect of alloying elements on in-plane ordering

and disordering of solute clusters in Mg-based long-period stacking ordered structures: A

first-principles analysis,” Scr. Mater., vol. 69, no. 8, pp. 594–597, 2013.

[26] M. Tane et al., “Effects of stacking sequence and short-range ordering of solute atoms on

elastic properties of Mg–Zn–Y alloys with long-period stacking ordered structures,” Acta

Mater., vol. 96, pp. 170–188, Sep. 2015.

[27] G. Kresse, “Efficient iterative schemes for ab initio total-energy calculations using a

plane-wave basis set,” Phys. Rev. B, vol. 54, no. 16, pp. 11169–11186, Oct. 1996.

[28] G. Kresse and J. Furthmüller, “Efficiency of ab-initio total energy calculations for metals

169

and semiconductors using a plane-wave basis set,” Comput. Mater. Sci., vol. 6, no. 1, pp.

15–50, Jul. 1996.

[29] G. Kresse, “From ultrasoft pseudopotentials to the projector augmented-wave method,”

Phys. Rev. B, vol. 59, no. 3, pp. 1758–1775, Jan. 1999.

[30] J. P. Perdew, K. Burke, and M. Ernzerhof, “Generalized Gradient Approximation Made

Simple,” Phys. Rev. Lett., vol. 77, no. 18, pp. 3865–3868, Oct. 1996.

[31] “VASP: The Guide, Wein, Austria, 2013.” [Online]. Available:

http://cms.mpi.univie.ac.at/vasp/vasp.pdf. [Accessed: 24-Apr-2016].

[32] M. C. Gao, A. D. Rollett, and M. Widom, “Lattice stability of aluminum-rare earth binary

systems: A first-principles approach,” Phys. Rev. B, vol. 75, no. 17, p. 174120, May 2007.

[33] Z. Mao, D. N. Seidman, and C. Wolverton, “First-principles phase stability, magnetic

properties and solubility in aluminum–rare-earth (Al–RE) alloys and compounds,” Acta

Materialia, May-2011. [Online]. Available: http://ac.els-cdn.com/S1359645411001340/1-

s2.0-S1359645411001340-main.pdf?_tid=a6b15ffe-cd86-11e4-8935-

00000aacb35e&acdnat=1426694073_e24f5f851261a8213df5060416b406d9. [Accessed:

20-Mar-2015].

[34] J. E. Saal and C. Wolverton, “Solute–vacancy binding of the rare earths in magnesium

from first principles,” Acta Mater., vol. 60, no. 13–14, pp. 5151–5159, Aug. 2012.

[35] A. Issa, J. E. Saal, and C. Wolverton, “Physical factors controlling the observed high-

strength precipitate morphology in Mg–rare earth alloys,” Acta Mater., vol. 65, pp. 240–

250, Feb. 2014.

[36] P. E. Blöchl, O. Jepsen, and O. K. Andersen, “Improved tetrahedron method for Brillouin-

zone integrations,” Phys. Rev. B, vol. 49, no. 23, pp. 16223–16233, Jun. 1994.

[37] K. Momma, F. Izumi, and IUCr, “VESTA : a three-dimensional visualization system for

electronic and structural analysis,” J. Appl. Crystallogr., vol. 41, no. 3, pp. 653–658, Jun.

170

2008.

[38] K. Momma, F. Izumi, and IUCr, “VESTA 3 for three-dimensional visualization of crystal,

volumetric and morphology data,” J. Appl. Crystallogr., vol. 44, no. 6, pp. 1272–1276,

Dec. 2011.

[39] S. Shang, Y. Wang, and Z.-K. Liu, “First-principles elastic constants of α- and θ-Al[sub

2]O[sub 3],” Appl. Phys. Lett., vol. 90, no. 10, p. 101909, Mar. 2007.

[40] R. Hill, “The Elastic Behaviour of a Crystalline Aggregate,” Proc. Phys. Soc. Sect. A, vol.

65, no. 5, pp. 349–354, May 1952.

[41] Z. K. Liu, H. Zhang, S. Ganeshan, Y. Wang, and S. N. Mathaudhu, “Computational

modeling of effects of alloying elements on elastic coefficients,” Scr. Mater., vol. 63, no.

7, pp. 686–691, 2010.

[42] S. L. Shang et al., “First-principles calculations of pure elements: Equations of state and

elastic stiffness constants,” Comput. Mater. Sci., vol. 48, no. 4, pp. 813–826, Jun. 2010.

[43] P. Niggli, Krystallographische und Strukturtheoretische Grundbegriffe, Handbuch der

Experimentalphysik. Akademische Verlagsgesellschaft, 1928.

[44] A. Santoro, A. D. Mighell, and IUCr, “Determination of reduced cells,” Acta Crystallogr.

Sect. A, vol. 26, no. 1, pp. 124–127, Jan. 1970.

[45] S. Eros and C. . Smith, “Low-temperature elastic constants of magnesium alloys,” Acta

Metall., vol. 9, no. 1, pp. 14–22, Jan. 1961.

[46] L. J. Slutsky and C. W. Garland, “Elastic Constants of Magnesium from 4.2°K to 300°K,”

Phys. Rev., vol. 107, no. 4, pp. 972–976, Aug. 1957.

[47] E. Oñorbe et al., “The evolution of internal strain in Mg–Y–Zn alloys with a long period

stacking ordered structure,” Scr. Mater., vol. 65, no. 8, pp. 719–722, Oct. 2011.

[48] A. R. Wazzan and L. B. Robinson, “Elastic constants of magnesium-lithium alloys,” Phys.

Rev., vol. 155, no. 3, pp. 586–594, 1967.

171

[49] Y. Chino, M. Mabuchi, S. Hagiwara, H. Iwasaki, A. Yamamoto, and H. Tsubakino,

“Novel equilibrium two phase Mg alloy with the long-period ordered structure,” Scr.

Mater., vol. 51, no. 7, pp. 711–714, Oct. 2004.

[50] S.-Y. Ma, L.-M. Liu, and S.-Q. Wang, “The microstructure, stability, and elastic

properties of 14H long-period stacking-ordered phase in Mg–Zn–Y alloys: a first-

principles study,” J. Mater. Sci., vol. 49, no. 2, pp. 737–748, 2013.

[51] S. Ganeshan, S. L. Shang, Y. Wang, and Z.-K. Liu, “Effect of alloying elements on the

elastic properties of Mg from first-principles calculations,” Acta Mater., vol. 57, no. 13,

pp. 3876–3884, 2009.

[52] S. L. Shang, D. E. Kim, C. L. Zacherl, Y. Wang, Y. Du, and Z. K. Liu, “Effects of

alloying elements and temperature on the elastic properties of dilute Ni-base superalloys

from first-principles calculations,” J. Appl. Phys., vol. 112, no. 5, p. 053515, 2012.

[53] M. Tane, Y. Nagai, H. Kimizuka, K. Hagihara, and Y. Kawamura, “Elastic properties of

an Mg–Zn–Y alloy single crystal with a long-period stacking-ordered structure,” Acta

Mater., vol. 61, no. 17, pp. 6338–6351, Oct. 2013.

[54] H. Kimizuka, M. Fronzi, and S. Ogata, “Effect of alloying elements on in-plane ordering

and disordering of solute clusters in Mg-based long-period stacking ordered structures: A

first-principles analysis,” Scr. Mater., vol. 69, no. 8, pp. 594–597, Oct. 2013.

[55] C. F. Cline, H. L. Dunegan, and G. W. Henderson, “Elastic constants of hexagonal BeO,

ZnS, and CdSe,” J. Appl. Phys., vol. 38, no. 4, pp. 1944–1948, 1967.

[56] D. Tromans, “Elastic Anisotropy of HCP Metal Crystals and Polycrystals,” Ijrras, vol. 6,

no. March, pp. 462–483, 2011.

[57] S. L. Shang et al., “Generalized stacking fault energy, ideal strength and twinnability of

dilute Mg-based alloys: A first-principles study of shear deformation,” Acta Mater., vol.

67, pp. 168–180, 2014.

172

[58] V. R. Manga et al., “Anomalous phonon stiffening associated with the (1 1 1) antiphase

boundary in L12 Ni3Al,” Acta Mater., vol. 82, pp. 287–294, 2015.

[59] K. Li et al., “Discovering a First-Order Phase Transition in the Li–CeO 2 System,” Nano

Lett., vol. 17, no. 2, pp. 1282–1288, 2017.

[60] P. N. H. Nakashima, A. E. Smith, J. Etheridge, and B. C. Muddle, “The bonding electron

density in aluminum.,” Science (80-. )., vol. 331, no. 6024, pp. 1583–6, Mar. 2011.

[61] A. R. Miedema, F. R. de Boer, and P. F. de Chatel, “Empirical description of the role of

electronegativity in alloy formation,” J. Phys. F Met. Phys., vol. 3, no. 8, pp. 1558–1576,

Aug. 1973.

[62] P. Wu, “Correlation of Bulk Modulus and the Constituent Element Properties of Binary

Intermetallic Compounds,” Chemistry of Materials, Dec-2001. [Online]. Available:

http://dx.doi.org/10.1021/cm0104203. [Accessed: 16-Apr-2016].

[63] A. A. A. Luo, “Recent magnesium alloy development for elevated temperature

applications,” Int. Mater. Rev., vol. 49, no. 1, pp. 13–30, 2004.

[64] D. Xu, E. H. Han, and Y. Xu, “Effect of long-period stacking ordered phase on

microstructure, mechanical property and corrosion resistance of Mg alloys: A review,”

Progress in Natural Science: Materials International, vol. 26, no. 2. pp. 117–128, 2016.

[65] K. Hagihara, N. Yokotani, and Y. Umakoshi, “Plastic deformation behavior of Mg12YZn

with 18R long-period stacking ordered structure,” Intermetallics, vol. 18, no. 2, pp. 267–

276, 2010.

[66] W. Hu, Z. Yang, J. Liu, and H. Ye, “Creep of A Mg-Zn-Y Alloy At Elevated

Temperatures,” in Magnesium Technology 2016, 2016, pp. 169–174.

[67] A. Inoue, Y. Kawamura, M. Matsushita, K. Hayashi, and J. Koike, “Novel hexagonal

structure and ultrahigh strength of magnesium solid solution in the Mg–Zn–Y system,” J.

Mater. Res., vol. 16, no. 07, pp. 1894–1900, Jan. 2001.

173

[68] H. Kimizuka, S. Kurokawa, A. Yamaguchi, A. Sakai, and S. Ogata, “Two-dimensional

ordering of solute nanoclusters at a close-packed stacking fault: modeling and

experimental analysis.,” Sci. Rep., vol. 4, p. 7318, Jan. 2014.

[69] K. Kishida, H. Yokobayashi, and H. Inui, “A formation criterion for Order-Disorder (OD)

phases of the Long-Period Stacking Order (LPSO)-type in Mg-Al-RE (Rare Earth)

Ternary Systems,” Sci. Rep., vol. 7, no. 1, p. 12294, Dec. 2017.

[70] G. Shao, V. Varsani, and Z. Fan, “Thermodynamic modelling of the Y-Zn and Mg-Zn-Y

systems,” Calphad-Computer Coupling Phase Diagrams Thermochem., vol. 30, no. 3, pp.

286–295, 2006.

[71] J. Grobner et al., “Phase equilibria and transformations in ternary Mg-rich Mg-Y-Zn

alloys,” Acta Mater., vol. 60, no. 17, pp. 5948–5962, 2012.

[72] H. Y. Qi, G. X. Huang, H. Bo, G. L. Xu, L. B. Liu, and Z. P. Jin, “Experimental

investigation and thermodynamic assessment of the Mg-Zn-Gd system focused on Mg-

rich corner,” J. Mater. Sci., vol. 47, no. 3, pp. 1319–1330, 2012.

[73] Z. K. Liu, “First-Principles Calculations and CALPHAD Modeling of Thermodynamics,”

J. Phase Equilibria Diffus., vol. 30, no. 5, pp. 517–534, 2009.

[74] M. Hillert and L. I. Staffansson, “Regular-solution model for stoichiometric phases and

ionic melts,” Acta Chem. Scand., vol. 24, no. 10, pp. 3618–3626, 1970.

[75] S.-L. Shang, Y. Wang, D. Kim, and Z.-K. Liu, “First-principles thermodynamics from

phonon and Debye model: Application to Ni and Ni3Al,” Comput. Mater. Sci., vol. 47, no.

4, pp. 1040–1048, Feb. 2010.

[76] M. Methfessel and A. T. Paxton, “High-precision sampling for Brillouin-zone integration

in metals,” Phys. Rev. B, vol. 40, no. 6, pp. 3616–3621, Aug. 1989.

[77] Y. Wang, Z.-K. K. Liu, and L.-Q. Q. Chen, “Thermodynamic properties of Al, Ni, NiAl,

and Ni3Al from first-principles calculations,” Acta Mater., vol. 52, no. 9, pp. 2665–2671,

174

May 2004.

[78] F. Birch, “Finite Elastic Strain of Cubic Crystals,” Phys. Rev., vol. 71, no. 11, pp. 809–

824, Jun. 1947.

[79] Y. Wang, L. Q. Chen, and Z. K. Liu, “YPHON: A package for calculating phonons of

polar materials,” Comput. Phys. Commun., vol. 185, no. 11, pp. 2950–2968, Nov. 2014.

[80] X. L. Liu, B. K. Vanleeuwen, S. L. Shang, Y. Du, and Z. K. Liu, “On the scaling factor in

Debye-Gruneisen model: A case study of the Mg-Zn binary system,” Comput. Mater. Sci.,

vol. 98, pp. 34–41, Feb. 2015.

[81] H. Kim, W. Y. Wang, S.-L. Shang, L. J. Kecskes, K. A. Darling, and Z.-K. Liu, “Elastic

properties of long periodic stacking ordered phases in Mg-Gd-Al alloys: A first-principles

study,” Intermetallics, vol. 98, pp. 18–27, 2018.

[82] A. T. Dinsdale, “SGTE DATA FOR PURE ELEMENTS,” Calphad, vol. 15, no. 4, pp.

317–425, 1991.

[83] P. LIANG et al., “Experimental investigation and thermodynamic calculation of the

central part of the Mg-Al phase diagram,” Zeitschrift f{ü}r Met., vol. 89, no. 8, pp. 536–

540, 1998.

[84] Y. Zhong, M. Yang, and Z. K. Liu, “Contribution of first-principles energetics to Al-Mg

thermodynamic modeling,” Calphad Comput. Coupling Phase Diagrams Thermochem.,

vol. 29, no. 4, pp. 303–311, 2005.

[85] T. Czeppe, W. Zakulski, E. Bielańska, and E. Bielanska, “Study of the thermal stability of

phases in the Mg-Al system,” vol. 24, no. 3, pp. 249–254, May 2003.

[86] G. Cacciamani, S. De Negri, A. Saccone, and R. Ferro, “The Al-R-Mg (R = Gd, Dy, Ho)

systems. Part II: Thermodynamic modelling of the binary and ternary systems,”

Intermetallics, vol. 11, no. 11–12, pp. 1135–1151, 2003.

[87] C. Guo, Z. Du, and C. Li, “A thermodynamic description of the Gd-Mg-Y system,”

175

Calphad Comput. Coupling Phase Diagrams Thermochem., vol. 31, no. 1, pp. 75–88,

2007.

[88] P. Manfrinetti and K. A. Gschneidner, “Phase equilibrium in the La- Mg (0–65 at.% Mg)

and Gd-Mg systems,” J. Less Common Met., vol. 123, pp. 267–275, 1986.

[89] K. A. Gschneidner and F. W. Calderwood, “Gd-Al and Dy-Al systems,” Bull. Alloy Phase

Diagrams, vol. 9, no. 6, pp. 680–683, 1988.

[90] A. Saccone, A. M. Cardinale, S. Delfino, and R. Ferro, “Gd-Al and Dy-Al systems: phase

equilibria in the 0 to 66.7 at.% Al composition range,” Zeitschrift f{ü}r Met., vol. 91, no.

1, pp. 17–23, 2000.

[91] S. Zhang, G. Y. Yuan, C. Lu, and W. J. Ding, “The relationship between (Mg,Zn)3RE

phase and 14H-LPSO phase in Mg–Gd–Y–Zn–Zr alloys solidified at different cooling

rates,” J. Alloys Compd., vol. 509, no. 8, pp. 3515–3521, 2011.

[92] X. H. Shao, H. J. Yang, J. T. M. De Hosson, and X. L. Ma, “Microstructural

Characterization of Long-Period Stacking Ordered Phases in Mg97Zn1Y2 (at.{%})

Alloy,” Microsc. Microanal., vol. 19, no. 6, pp. 1575–1580, Dec. 2013.

[93] M. Jiang, S. Zhang, Y. Bi, H. Li, Y. Ren, and G. Qin, “Phase equilibria of the long-period

stacking ordered phase in the Mg-Ni-Y system,” Intermetallics, vol. 57, pp. 127–132,

2015.

[94] M. Jiang, X. Su, H. Li, Y. Ren, and G. Qin, “The phase equilibria and thermal stability of

the long-period stacking ordered phase in the Mg-Cu-Y system,” J. Alloys Compd., vol.

593, pp. 141–147, 2014.

[95] T. Horiuchi et al., “Liquid-Solid Equilibrium and Intermediate Phase Formation during

Solidification in Mg-1.3 at%Zn-1.7 at%Y Alloy,” Mater. Trans., vol. 49, no. 10, pp.

2247–2253, 2008.

[96] Y. J. Wu, D. L. Lin, X. Q. Zeng, L. M. Peng, and W. J. Ding, “Formation of a lamellar

176

14H-type long period stacking ordered structure in an as-cast Mg–Gd–Zn–Zr alloy,” J.

Mater. Sci., vol. 44, no. 6, pp. 1607–1612, Mar. 2009.

[97] M. Yamasaki, T. Anan, S. Yoshimoto, and Y. Kawamura, “Mechanical properties of

warm-extruded Mg–Zn–Gd alloy with coherent 14H long periodic stacking ordered

structure precipitate,” Scr. Mater., vol. 53, no. 7, pp. 799–803, 2005.

[98] M. Yamasaki, M. Sasaki, M. Nishijima, K. Hiraga, and Y. Kawamura, “Formation of 14H

long period stacking ordered structure and profuse stacking faults in Mg-Zn-Gd alloys

during isothermal aging at high temperature,” Acta Mater., vol. 55, no. 20, pp. 6798–6805,

Dec. 2007.

[99] W. J. Ding, Y. J. Wu, L. M. Peng, X. Q. Zeng, G. Y. Yuan, and D. L. Lin, “Formation of

14H-type long period stacking ordered structure in the as-cast and solid solution treated

Mg-Gd-Zn-Zr alloys,” J. Mater. Res., vol. 24, no. 5, pp. 1842–1854, May 2009.

[100] T. Itoi, T. Seimiya, Y. Kawamura, and M. Hirohashi, “Long period stacking structures

observed in Mg97Zn1Y2 alloy,” Scr. Mater., vol. 51, no. 2, pp. 107–111, 2004.

[101] Y. M. Zhu, A. J. Morton, and J. F. Nie, “The 18R and 14H long-period stacking ordered

structures in Mg–Y–Zn alloys,” Acta Mater., vol. 58, no. 8, pp. 2936–2947, 2010.

[102] P. Karen, A. Kjekshus, Q. Huang, and V. L. Karen, “The crystal structure of magnesium

dicarbide,” J. Alloys Compd., vol. 282, no. 1–2, pp. 72–75, Jan. 1999.

[103] D. M. Sweger, R. Segnan, and J. J. Rhyne, “Temperature dependence of hyperfine

interactions in Dy-Gd alloys,” Phys. Rev. B, vol. 9, no. 9, pp. 3864–3870, May 1974.

[104] S. Popovic, B. Grzeta, V. Ilakovac, R. Kroggel, G. Wendrock, and H. Loffler, “Lattice

constant of the F.C.C. Al-rich α-Phase of Al-Zn alloys in equilibrium with GP zones and

the β(Zn)-Phase,” Phys. status solidi, vol. 130, no. 2, pp. 273–292, Apr. 1992.

[105] H. R. Schober and P. H. Dederichs, “Mg,” in Phonon States of Elements. Electron States

and Fermi Surfaces of Alloys, K.-H. Hellwege and J. L. Olsen, Eds. Berlin, Heidelberg:

177

Springer Berlin Heidelberg, 1981, pp. 82–86.

[106] Charles Kittel, Introduction to Solid State Physics, 8th ed. Hoboken, NJ: John Wiley &

Sons, Inc., 2005.

[107] Y. Wang et al., “Ab initio lattice stability in comparison with CALPHAD lattice stability,”

Calphad Comput. Coupling Phase Diagrams Thermochem., vol. 28, no. 1, pp. 79–90,

2004.

[108] S. Iikubo, K. Matsuda, and H. Ohtani, “Phase stability of long-period stacking structures

in Mg-Y-Zn: A first-principles study,” Phys. Rev. B - Condens. Matter Mater. Phys., vol.

86, no. 5, p. 54105, 2012.

[109] K. Kishida, A. Inoue, H. Yokobayashi, and H. Inui, “Deformation twinning in a Mg-Al-

Gd ternary alloy containing precipitates with a long-period stacking-ordered (LPSO)

structure,” Scr. Mater., 2014.

[110] S. De Negri, A. Saccone, G. Cacciamani, and R. Ferro, “The Al–R–Mg (R=Gd, Dy, Ho)

systems. Part I: experimental investigation,” Intermetallics, vol. 11, no. 11–12, pp. 1125–

1134, 2003.

[111] F. Lu et al., “Significantly improved corrosion resistance of heat-treated Mg-Al-Gd alloy

containing profuse needle-like precipitates within grains,” Corros. Sci., vol. 94, pp. 171–

178, 2015.

[112] J. Dai, S. Zhu, M. A. Easton, W. Xu, G. Wu, and W. Ding, “Precipitation process in a Mg-

Gd-Y alloy grain-refined by Al addition,” Mater. Charact., vol. 88, pp. 7–14, 2014.

[113] X. F. Gu, T. Furuhara, L. Chen, and P. Yang, “Study on the planar segregation of solute

atoms in Mg-Al-Gd alloy,” Scr. Mater., vol. 150, pp. 45–49, 2018.

[114] Q. Li, H. Wang, and Y. M. Ma, “Predicting new superhard phases,” J. Superhard Mater.,

vol. 32, no. 3, pp. 192–204, 2010.

[115] H. Y. Chung et al., “Synthesis of ultra-incompressible superhard rhenium diboride at

178

ambient pressure,” Science (80-. )., vol. 316, no. 5823, pp. 436–439, Apr. 2007.

[116] A. Šimůnek and J. Vackář, “Hardness of covalent and ionic crystals: First-principle

calculations,” Phys. Rev. Lett., vol. 96, no. 8, pp. 5–8, 2006.

[117] D. M. Teter, “Computational Alchemy: The Search for New Superhard Materials,” MRS

Bull., vol. 23, no. 01, pp. 22–27, 1998.

[118] A. Y. Liu and M. L. Cohen, “Prediction of new low compressibility solids,” Science (80-.

)., vol. 245, no. 4920, pp. 841–842, Aug. 1989.

[119] X.-Q. Q. Chen, H. Niu, D. Li, and Y. Li, “Modeling hardness of polycrystalline materials

and bulk metallic glasses,” Intermetallics, vol. 19, no. 9, pp. 1275–1281, 2011.

[120] D. W. Richerson, W. E. Lee, and D. Richerson, Modern Ceramic Engineering: Properties,

Processing, and Use in Design, Third Edition. Taylor & Francis, 1992.

[121] F. Gao et al., “Hardness of covalent crystals.,” Phys. Rev. Lett., vol. 91, no. 1, p. 015502,

Jul. 2003.

[122] K. Li, X. Wang, F. Zhang, and D. Xue, “Electronegativity identification of novel

superhard materials,” Phys. Rev. Lett., vol. 100, no. 23, pp. 1–4, 2008.

[123] S. F. Pugh, “Relations between the elastic moduli and the plastic properties of

polycrystalline pure metals,” Philos. Mag. Ser. 7, vol. 45, no. 367, pp. 823–843, 1954.

[124] J. Haines, J. M. L, and G. Bocquillon, “Synthesis and Design of Superhard Materials,”

Annu. Rev. Mater. Res., vol. 1955, no. 1, pp. 1–23, 2001.

[125] R. Peierls, “The size of a dislocation,” Proc. Phys. Soc., vol. 52, no. 1, pp. 34–37, 1940.

[126] American Society for Testing and Materials, “Standard Test Method for Knoop and

Vickers Hardness of Materials,” ASTM Stand. E384-11, pp. 1–43, 2012.

[127] W. C. Oliver and G. M. Pharr, “An Improved Technique for Determining Hardness and

Elastic-Modulus Using Load and Displacement Sensing Indentation Experiments,” J.

Mater. Res., vol. 7, no. 6, pp. 1564–1583, 1992.

179

[128] S. E. Grillo et al., “Nanoindentation of Si, GaP, GaAs and ZnSe single crystals,” J. Phys.

D. Appl. Phys., vol. 36, pp. L5–L9, Jan. 2003.

[129] G. M. Pharr, W. C. Oliver, and F. R. Brotzen, “On the generality of the relationship

among cantact stiffness,contact area, and elastic-modulus during indentation,” J. Mater.

Res., vol. 7, no. 3, pp. 613–617, 1992.

[130] W. C. Oliver and G. M. Pharr, “Measurement of hardness and elastic modulus by

instrumented indentation: Advances in understanding and refinements to methodology,” J.

Mater. Res., vol. 19, no. 01, pp. 3–20, 2004.

[131] M. J. Cawkwell, D. Nguyen-Manh, D. G. Pettifor, and V. Vitek, “Construction,

assessment, and application of a bond-order potential for iridium,” Phys. Rev. B -

Condens. Matter Mater. Phys., vol. 73, no. 6, p. 064104, 2006.

[132] V. Vitek and V. Paidar, “Chapter 87 Non-planar Dislocation Cores: A Ubiquitous

Phenomenon Affecting Mechanical Properties of Crystalline Materials,” in Dislocations in

Solids, Elsevier Masson SAS, 2008, pp. 439–514.

[133] S. S. Hecker, D. L. Rohr, and D. F. Stein, “Brittle fracture in iridium,” Metall. Trans. A,

vol. 9, no. April, pp. 481–488, 1978.

[134] M. J. Cawkwell, D. Nguyen-Manh, C. Woodward, D. G. Pettifor, and V. Vitek, “Origin of

brittle cleavage in iridium,” Science (80-. )., vol. 309, no. 5737, pp. 1059–1062, 2005.

[135] M. J. Cawkwell, C. Woodward, D. Nguyen-Manh, D. G. Pettifor, and V. Vitek,

“Atomistic study of athermal cross-slip and its impact on the mechanical properties of

iridium,” Acta Mater., vol. 55, no. 1, pp. 161–169, 2007.

[136] C. R. Weinberger, B. L. Boyce, and C. C. Battaile, “Slip planes in bcc transition metals,”

Int. Mater. Rev., vol. 58, no. 5, pp. 296–314, 2013.

[137] L. L. Hsiung, “On the mechanism of anomalous slip in bcc metals,” Mater. Sci. Eng. A,

vol. 528, pp. 329–337, 2010.

180

[138] D. Caillard, “Kinetics of dislocations in pure Fe. Part I. In situ straining experiments at

room temperature,” Acta Mater., vol. 58, pp. 3493–3503, 2010.

[139] J. Bressers and P. De Meester, “Slip plane choice in vanadium at deformation

temperatures T ≤ 0.15Tm,” J. Less-Common Met., vol. 84, no. C, pp. 11–23, 1982.

[140] R. Fritz, V. Maier-Kiener, D. Lutz, and D. Kiener, “Interplay between sample size and

grain size: Single crystalline vs. ultrafine-grained chromium micropillars,” Mater. Sci.

Eng. A, vol. 674, pp. 626–633, 2016.

[141] R. A. Foxall, M. S. Duesbery, and P. B. Hirsch, “The deformation of niobium single

crystals,” Can. J. Phys., vol. 45, no. 2, pp. 607–629, Feb. 1967.

[142] P. Beardmore and D. Hull, “Deformation and fracture of tungsten single crystals,” J. Less-

Common Met., vol. 9, no. 3, pp. 168–180, 1965.

[143] H. W. Schadler, “Deformation behavior of zone-melted tungsten single crystals,” Trans.

Met. Soc. AIME, vol. 218, pp. 649–655, 1960.

[144] D. Ali, N. Mushtaq, and M. Z. Butt, “Investigation of active slip-systems in some body-

centered cubic metals,” J. Mater. Sci., vol. 46, no. 11, pp. 3812–3821, 2011.

[145] M. Z. Butt, “Kinetics of flow stress in crystals with high intrinsic lattice friction,” Philos.

Mag., vol. 87, no. 24, pp. 3595–3614, 2007.

[146] M. W. Finnis and J. E. Sinclair, “A simple empirical N-body potential for transition

metals,” Philos. Mag. A Phys. Condens. Matter, Struct. Defects Mech. Prop., vol. 50, no.

1, pp. 45–55, Jul. 1984.

[147] N. A. Hill and J. W. S. Jones, “The crystallographic dependence of low load indentation

hardness in beryllium,” J. Nucl. Mater., vol. 3, no. 2, pp. 138–155, Feb. 1961.

[148] T. G. Carnahan and T. E. Scott, “Deformation modes of hcp Yttrium at 77, 298, and 497

K,” Metall. Trans., vol. 4, no. 1, pp. 27–32, 1973.

[149] C. a. Yablinsky, E. K. Cerreta, G. T. Gray, D. W. Brown, and S. C. Vogel, “The effect of

181

twinning on the work-hardening behavior and microstructural evolution of hafnium,”

Metall. Mater. Trans. A, vol. 37, no. June, pp. 1907–1915, 2006.

[150] A. P. Gerk, “The effect of work-hardening upon the hardness of solids: minimum

hardness,” J. Mater. Sci., vol. 12, no. 4, pp. 735–738, 1977.

[151] L. Liu, X. Z. Wu, R. Wang, H. F. Feng, and S. Wu, “On the generalized stacking energy,

core structure and Peierls stress of the 110{110} dislocations in alkali halide,” Eur. Phys.

J. B, vol. 85, p. 58, 2012.

[152] F. Granzer, G. Wagner, and J. EisenblAtter, “Atomistische Bereechnung der Core-

Struktur, Core-Energie und Peierls-Spannung einer Stufenversetzung in NaCl,” Phys.

status solidi, vol. 30, pp. 587–600, 1968.

[153] I. Yonenaga, “Hardness, Yield Strength, and Dislocation Velocity in Elemental and

Compound Semiconductors,” Mater. Trans., vol. 46, no. 9, pp. 1979–1985, 2005.

[154] M. S. Duesbery, “Dislocation motion in silicon: The shuffle-glide controversy,” Philos.

Mag. Lett., vol. 74, no. 4, pp. 253–258, 1996.

[155] J. F. Justo, M. de Koning, W. Cai, and V. V Bulatov, “Vacancy interaction with

dislocations in silicon: The shuffle-glide competition,” Phys. Rev. Lett., vol. 84, no. 10,

pp. 2172–2175, 2000.

[156] C. Z. Wang, J. Li, K. M. Ho, and S. Yip, “Undissociated screw dislocation in Si: Glide or

shuffle set?,” Appl. Phys. Lett., vol. 89, no. 5, p. 051910, Jul. 2006.

[157] V. Domnich and Y. Gogotsi, “Phase Transformations in Silicon Under Contact Loading,”

Rev. Adv. Mater. Sci., vol. 3, pp. 1–36, 2002.

[158] I. V. Gridneva, Y. V. Milman, and V. I. Trefilov, “Phase transition in diamond-structure

crystals during hardness measurements,” Phys. Status Solidi, vol. 14, no. 1, pp. 177–182,

Nov. 1972.

[159] V. Domnich, Y. Aratyn, W. M. Kriven, and Y. Gogotsi, “Temperature dependence of

182

silicon hardness: Experimental evidence of phase transformations,” Rev. Adv. Mater. Sci.,

vol. 17, no. 1–2, pp. 33–41, 2008.

[160] Y.-J. Hu et al., “Solute-induced solid-solution softening and hardening in bcc tungsten,”

Acta Mater., vol. 141, pp. 304–316, Dec. 2017.

[161] G. D. P. J. P. Quinn, “Effect of Loading Rate Upon Conventional Ceramic

Microindentation Hardness,” J. Res. Natl. Inst. Stand. Technol., vol. 107, no. 3, pp. 299–

306, 2002.

[162] D. J. Weidner, Y. Wang, and M. T. Vaughan, “Strength of diamond,” Science, vol. 266,

no. 5184, pp. 419–422, Oct. 1994.

[163] T. Suzuki and H. Kojima, “Dislocation motion in silicon crystals as measured by the lang

X-ray technique,” Acta Metall., vol. 14, no. 8, pp. 913–924, 1966.

[164] S. Takeuchi, T. Hashimoto, and K. Maeda, “Plastic Deformation of bcc Metal Single

Crystals at Very Low Temperatures,” Trans. Japan Inst. Met., vol. 23, no. 2, pp. 60–69,

1982.

[165] W. J. M. G. Tegart, Elements of Mechanical Metallurgy. Mac Millan, 1966.

[166] D. F. Stein, “The effect of orientation and impurities on the mechanical properties of

molybdenum single crystals,” Can. J. Phys., vol. 45, no. 2, pp. 1063–1074, 1967.

[167] S. Howe, B. Liebmann, and K. Lücke, “High temperature deformation of aluminum single

crystals,” vol. 9, no. 7, pp. 625–631, 1961.

[168] K. Marukawa, “Dislocation motion in copper single crystals,” J. Phys. Soc. Japan, vol. 22,

no. 2, pp. 499–510, Feb. 1967.

[169] H. Suga, Y. Sueki, M. Higuchi, and T. Imura, “The Anomaly in Temperature Dependence

of the Yield Stress of Silver Base Solid Solutions,” Japanese J. Appl. Physics, Part 1

Regul. Pap. Short Notes Rev. Pap., vol. 15, no. 2, pp. 379–380, Feb. 1976.

[170] G. Thomas, J. Washburn, and L. R. L. I. M. R. Division, Electron Microscopy and

183

Strength of Crystals: Proceedings of the First Berkeley International Materials

Conference: the Impact of Transmission Electron Microscopy on Theories of the Strength

of Crystals. Interscience Publishers, 1963.

[171] P. I. Treharne and A. Moore, “Tensile deformation of beryllium single crystals in various

orientations between 25°C and 600°C,” J. Less-Common Met., vol. 4, no. 3, pp. 275–285,

1962.

[172] E. Schmid, “Beiträge zur Physik und Metallographie des Magnesiums,” Zeitschrift für

Elektrochemie, vol. 37, no. 8‐9, pp. 447–459, Aug. 1931.

[173] W. Boas and E. Schmid, “Über die Temperaturabhängigkeit der kritischen Schubspannung

von Cadmiumkristallen,” Zeitschrift für Phys., vol. 57, no. 9–10, pp. 575–581, Sep. 1929.

[174] W. Fahrenhorst and E. Schmid, “Über die Temperaturabhängigkeit der Kristallplastizität.

II,” Zeitschrift für Phys., vol. 64, no. 11–12, pp. 845–855, Sep. 1930.

[175] M. P. Biget and G. Saada, “Low-temperature plasticity of high-purity α-titanium single

crystals,” Philos. Mag. A, vol. 59, no. 4, pp. 747–757, Apr. 1989.

[176] E. D. Levine, “Deformation mechanisms in titanium at low temperatures,” IME Trans.,

vol. 236, pp. 1558–1564, 1966.

[177] A. Akhtar and E. Teghtsoonian, “Prismatic slip in α-titanium single crystals,” Metall.

Mater. Trans. A, vol. 6, no. 12, pp. 2201–2208, Dec. 1975.

[178] Y. T. Cheng and C. M. Cheng, “Relationships between hardness, elastic modulus, and the

work of indentation,” Appl. Phys. Lett., vol. 73, no. 5, pp. 614–616, 1998.

[179] M. Sakai, “Meyer hardness: A measure for plasticity?,” J. Mater. Res., vol. 14, no. 9, pp.

3630–3639, 1999.

[180] D. Chicot, D. Mercier, F. Roudet, K. Silva, M. H. Staia, and J. Lesage, “Comparison of

instrumented Knoop and Vickers hardness measurements on various soft materials and

hard ceramics,” J. Eur. Ceram. Soc., vol. 27, no. 4, pp. 1905–1911, 2007.

184

[181] X.-Q. Chen, H. Niu, D. Li, and Y. Li, “Modeling hardness of polycrystalline materials and

bulk metallic glasses,” Intermetallics, vol. 19, no. 9, pp. 1275–1281, 2011.

[182] J. M. Gere, mechanics of materials, 6th ed. Belmont, CA: Thomson Learning, Inc, 2004.

[183] E. G. Herbert, G. M. Pharr, W. C. Oliver, B. N. Lucas, and J. L. Hay, “On the

measurement of stress – strain curves by spherical indentation,” Thin Solid Films, vol.

399, pp. 331–335, 2001.

[184] S. Pathak and S. R. Kalidindi, “Spherical nanoindentation stress-strain curves,” Mater.

Sci. Eng. R Reports, vol. 91, pp. 1–36, 2015.

[185] S. Basu, A. Moseson, and M. W. Barsoum, “On the determination of spherical

nanoindentation stress-strain curves,” J. Mater. Res., vol. 21, no. 10, pp. 2628–2637,

2006.

[186] S. Jayaraman, G. T. T. Hahn, W. C. C. Oliver, C. A. A. Rubin, and P. C. C. Bastias,

“Determination of monotonic stress-strain curve of hard materials from ultra-low-load

indentation tests,” Int. J. Solids Struct., vol. 35, no. 5–6, pp. 365–381, Feb. 1998.

[187] M. Beghini, L. Bertini, and V. Fontanari, “Evaluation of the stress-strain curve of metallic

materials by spherical indentation,” Int. J. Solids Struct., vol. 43, no. 7–8, pp. 2441–2459,

2006.

[188] J. H. Ahn and D. Kwon, “Derivation of plastic stress-strain relationship from ball

indentations: Examination of strain definition and pileup effect,” J. Mater. Res., vol. 16,

no. 11, pp. 3170–3178, 2001.

[189] K. Matsuda, “Prediction of stress-strain curves of elastic-plastic materials based on the

Vickers indentation,” Philos. Mag. A Phys. Condens. Matter, Struct. Defects Mech. Prop.,

vol. 82, no. 10, pp. 1941–1951, 2002.

[190] K. D. Bouzakis, N. Michailidis, S. Hadjiyiannis, G. Skordaris, and G. Erkens, “The effect

of specimen roughness and indenter tip geometry on the determination accuracy of thin

185

hard coatings stress-strain laws by nanoindentation,” Mater. Charact., vol. 49, no. 2, pp.

149–156, 2002.

[191] K. D. Bouzakis, N. Michailidis, and G. Erkens, “Thin hard coatings stress-strain curve

determination through a FEM supported evaluation of nanoindentation test results,” Surf.

Coatings Technol., vol. 142–144, pp. 102–109, 2001.

[192] G. I. Taylor, “The Mechanism of Plastic Deformation of Crystals. Part II. Comparison

with Observations,” Proc. R. Soc. A Math. Phys. Eng. Sci., vol. 145, no. 855, pp. 388–404,

Jul. 1934.

[193] G. I. Taylor, “The Mechanism of Plastic Deformation of Crystals. Part I. Theoretical,”

Proc. R. Soc. A Math. Phys. Eng. Sci., vol. 145, no. 855, pp. 362–387, Jul. 1934.

[194] W. D. Nix and H. Gao, “Indentation size effects in crystalline materials: A law for strain

gradient plasticity,” J. Mech. Phys. Solids, vol. 46, no. 3, pp. 411–425, 1998.

[195] W. C. Oliver and G. M. Pharr, “An improved technique for determining hardness and

elastic modulus using load and displacement sensing indentation experiments,” J. Mater.

Res., vol. 7, no. 06, pp. 1564–1583, 1992.

[196] “Nano vs. Micro Indentation Hardness Testing.” [Online]. Available:

http://nanomechanicsinc.com/indentation-hardness/. [Accessed: 02-Oct-2017].

[197] “Hysitron service document.” [Online]. Available:

https://www.hysitron.com/media/1683/t-003-v30-probe-calibration.pdf.

[198] B. Rother, A. Steiner, D. A. Dietrich, H. A. Jehn, J. Haupt, and W. Gissler, “Depth-

sensing indentation measurements with Vickers and Berkovich indenters,” J. Mater. Res.,

vol. 13, no. 8, pp. 2071–2076, 1998.

[199] L. Min, C. Wei-min, L. Nai-gang, and W. Ling-Dong, “A numerical study of indentation

using indenters of different geometry,” J. Mater. Res., vol. 19, no. 01, pp. 73–78, 2004.

[200] N. A. Sakharova, J. V Fernandes, J. M. Antunes, and M. C. Oliveira, “Comparison

186

between Berkovich, Vickers and conical indentation tests: A three-dimensional numerical

simulation study,” Int. J. Solids Struct., vol. 46, no. 5, pp. 1095–1104, 2009.

[201] O. L. Anderson, “A simplified method for calculating the debye temperature from elastic

constants,” J. Phys. Chem. Solids, vol. 24, no. 7, pp. 909–917, 1963.

[202] F. Birch, “The velocity of compressional waves in rocks to 10 kilobars: 2.,” J. Geophys.

Res., vol. 66, no. 7, pp. 2199–2224, Jul. 1961.

[203] D. H. Chung and W. R. Buessem, “The Voigt-Reuss-Hill approximation and elastic

moduli of polycrystalline Mgo, CaF2, β-ZnS, ZnSe, and CdTe,” J. Appl. Phys., vol. 38,

no. 6, pp. 2535–2540, 1967.

[204] S. Hirsekorn, “Elastic Properties of Polycrystals: A Review,” Textures Microstruct., vol.

12, no. 1–3, pp. 1–14, 1990.

[205] A. G. Every and A. K. McCurdy, “Second and Higher Order Elastic Constants,” in Low

Frequency Properties of Dielectric Crystals’ of Landolt-Börnstein - Group III Condensed

Matter, vol. 29a, D. F. Nelson, Ed. Berlin/Heidelberg: Springer-Verlag, 1992, pp. 259–

419.

[206] S. V Hainsworth, H. W. Chandler, and T. F. Page, “Analysis of nanoindentation load-

displacement loading curves,” J. Mater. Res., vol. 11, no. 8, pp. 1987–1995, 1996.

[207] G. M. Pharr and W. C. Oliver, “Nanoindentation of silver-relations between hardness and

dislocation structure,” J. Mater. Res., vol. 4, no. 01, pp. 94–101, 1989.

[208] K. Durst, O. Franke, A. Böhner, and M. Göken, “Indentation size effect in Ni-Fe solid

solutions,” Acta Mater., vol. 55, no. 20, pp. 6825–6833, 2007.

[209] Y. V Milman, B. A. Galanov, and S. I. Chugunova, “Plasticity characteristic obtained

through hardness measurement,” Acta Metall. Mater., vol. 41, no. 9, pp. 2523–2532, 1993.

[210] M. Sakai and Y. Nakano, “Elastoplastic load-depth hysteresis in pyramidal indentation,”

J. Mater. Res., vol. 17, no. 8, pp. 2161–2173, 2002.

187

[211] R. D. Carnahan, “Elastic Properties of Silicon Carbide,” J. Am. Ceram. Soc., vol. 51, no.

4, pp. 223–224, Apr. 1968.

[212] V. L. Solozhenko, S. N. Dub, and N. V Novikov, “Mechanical properties of cubic BC2N,

a new superhard phase,” Diam. Relat. Mater., vol. 10, no. 12, pp. 2228–2231, 2001.

[213] M. Sakai, S. Shimizu, and T. Ishikawa, “The Indentation Load-depth Curve of Ceramics,”

J. Mater. Res., vol. 14, no. 04, pp. 1471–1484, 1999.

[214] M. Grimsditch, E. S. Zouboulis, and A. Polian, “Elastic constants of boron nitride,” J.

Appl. Phys., vol. 76, pp. 832–834, 1994.

[215] D. M. Teter, “Computational Alchemy: The Search for New Superhard Materials,” MRS

Bull., vol. 23, no. 01, pp. 22–27, 1998.

[216] M. Sakai, “Energy principle of the indentation-induced inelastic surface deformation and

hardness of brittle materials,” Acta Metall. Mater., vol. 41, no. 6, pp. 1751–1758, 1993.

[217] D. Chicot, M. Y. N’Jock, F. Roudet, X. Decoopman, M. H. Staia, and E. S. Puchi-

Cabrera, “Some improvements for determining the hardness of homogeneous materials

from the work-of-indentation,” Int. J. Mech. Sci., vol. 105, pp. 279–290, Jan. 2016.

[218] B. J. Briscoe, L. Fiori, and E. Pelillo, “Nano-indentation of polymeric surfaces,” J. Phys.

D. Appl. Phys., vol. 31, no. 19, pp. 2395–2405, Oct. 1998.

[219] M. Lei, B. Xu, Y. Pei, H. Lu, and Y. Q. Fu, “Micro-mechanics of nanostructured

carbon/shape memory polymer hybrid thin film,” Soft Matter, vol. 12, no. 1, pp. 106–114,

2015.

[220] F. Pöhl, C. Hardes, and W. Theisen, “Scratch behavior of soft metallic materials,” AIMS

Mater. Sci., vol. 3, no. 2, pp. 390–403, 2016.

[221] M. Yetna Njock, F. Roudet, M. Idriss, O. Bartier, and D. Chicot, “Work-of-indentation

coupled to contact stiffness for calculating elastic modulus by instrumented indentation,”

Mech. Mater., vol. 94, pp. 170–179, 2016.

188

[222] J. Malzbender and G. De With, “Indentation load–displacement curve, plastic

deformation, and energy,” J. Mater. Res., vol. 17, no. 2, pp. 502–511.

[223] J. Malzbender, “Comment on hardness definitions,” J. Eur. Ceram. Soc., vol. 23, no. 9,

pp. 1355–1359, 2003.

[224] T. A. Venkatesh, K. J. Van Vliet, A. E. Giannakopoulos, and S. Suresh, “Determination of

elasto-plastic properties by instrumented sharp indentation: guidelines for property

extraction,” Scr. Mater., vol. 42, no. 9, pp. 833–839, Apr. 2000.

[225] U. Messerschmidt, Dislocation Dynamics During Plastic Deformation. Springer-Verlag

Berlin Heidelberg, 2010.

[226] J. Frenkel, “Zur Theorie der Elastizitatsgrenze und der Festigkeit kristallinischer Korper,”

Zeitschrift fur Phys., vol. 37, no. 7–8, pp. 572–609, Jul. 1926.

[227] J. P. Hirth and J. Lothe, Theory of dislocations, 2nd ed. New York: Wiley, 1982.

[228] D. Kuhlmann-Wilsdorf, “Theory of plastic deformation: Properties of low energy

dislocation structures,” Mater. Sci. Engng. A, vol. 113, pp. 1–41, 1989.

[229] C. R. Krenn, D. Roundy, M. L. Cohen, D. C. Chrzan, and J. W. Morris, “Connecting

atomistic and experimental estimates of ideal strength,” Phys. Rev. B, vol. 65, p. 134111,

2002.

[230] P.-Y. Tang, G.-H. Huang, Q.-L. Xie, and J.-Y. Li, “Ideal shear strength and deformation

behaviours of L10 TiAl from first-principles calculations,” Bull. Mater. Sci., vol. 39, no.

6, pp. 1411–1418, 2016.

[231] Y. Zhang, Sun Hong, and C. Chen, “Superhard cubic BC2N compared to diamond,” Phys.

Rev. Lett., vol. 93, no. 19, p. 195504, 2004.

[232] D. Roundy and M. L. Cohen, “Ideal strength of diamond, Si, and Ge,” Phys. Rev. B -

Condens. Matter Mater. Phys., vol. 64, p. 212103, 2001.

[233] W. Luo, D. Roundy, M. L. Cohen, and J. W. Morris, “Ideal strength of bcc molybdenum

189

and niobium,” Phys. Rev. B - Condens. Matter Mater. Phys., vol. 66, no. 9, pp. 1–7, 2002.

[234] S. Ogata, J. Li, and S. Yip, “Ideal pure shear strength of aluminum and copper.,” Science,

vol. 298, no. 5594, pp. 807–11, Oct. 2002.

[235] F. R. N. Nabarro, “Fifty-year study of the Peierls-Nabarro stress,” Mater. Sci. Eng. A, vol.

234–236, pp. 67–76, 1997.

[236] H. D. Dietze, “Die Temperaturabhängigkeit der Versetzungsstruktur,” Zeitschrift für

Phys., vol. 132, no. 1, pp. 107–110, 1952.

[237] Z. Wu, H. Bei, G. M. Pharr, and E. P. George, “Temperature dependence of the

mechanical properties of equiatomic solid solution alloys with face-centered cubic crystal

structures,” Acta Mater., vol. 81, pp. 428–441, 2014.

[238] F. R. N. Nabarro, “The Peierls stress for a wide dislocation,” Mater. Sci. Eng. A, vol. 113,

no. C, pp. 315–326, 1989.

[239] S. Rao, C. Hernandezt, J. P. Slmmons, T. A. Parthasarathy, and C. Woodward, “Green’s

function boundary conditions in two-dimensional and three-dimensional atomistic

simulations of dislocations,” Philos. Mag. A, vol. 77, no. 1, pp. 231–256, 1998.

[240] C. Woodward, D. R. Trinkle, L. G. Hector, and D. L. Olmsted, “Prediction of dislocation

cores in aluminum from density functional theory,” Phys. Rev. Lett., vol. 100, p. 045507,

2008.

[241] J. A. Yasi, T. Nogaret, D. R. Trinkle, Y. Qi, L. G. Hector, and W. A. Curtin, “Basal and

prism dislocation cores in magnesium: Comparison of first-principles and embedded-

atom-potential methods predictions,” Model. Simul. Mater. Sci. Eng., vol. 17, no. 5, p.

055012, Jul. 2009.

[242] Y. Zhang, J. Liu, H. Chu, and J. Wang, “Elastic fields of a core-spreading dislocation in

anisotropic bimaterials,” Int. J. Plast., vol. 81, pp. 231–248, 2016.

[243] G. Lu, “The Peierls—Nabarro Model of Dislocations: A Venerable Theory and its Current

190

Development,” in Handbook of Materials Modeling, Dordrecht: Springer Netherlands,

2005, pp. 793–811.

[244] Y.-T. T. Cheng and C.-M. M. Cheng, “Relationships between hardness, elastic modulus,

and the work of indentation,” Appl. Phys. Lett., vol. 73, no. 10, pp. 614–616, 1998.

[245] Y. T. Cheng and C. M. Cheng, “Can stress-strain relationships be obtained from

indentation curves using conical and pyramidal indenters?,” J. Mater. Res., vol. 14, no. 9,

pp. 3493–3496, 1999.

[246] C.-M. Cheng and Y.-T. Cheng, “On the initial unloading slope in indentation of elastic-

plastic solids by an indenter with an axisymmetric smooth profile,” Appl. Phys. Lett., vol.

71, pp. 2623–2625, 1997.

[247] Y. T. Cheng, Z. Li, and C. M. Cheng, “Scaling relationships for indentation

measurements,” Philos. Mag. A Phys. Condens. Matter, Struct. Defects Mech. Prop., vol.

82, no. 10, pp. 1821–1829, 2002.

[248] J. B. Walsh, “The effect of cracks in rocks on Poisson’s ratio,” J. Geophys. Res., vol. 70,

no. 20, pp. 5249–5257, Oct. 1965.

[249] H. Yao, L. Ouyang, and W. Y. Ching, “Ab initio calculation of elastic constants of

ceramic crystals,” J. Am. Ceram. Soc., vol. 90, no. 10, pp. 3194–3204, Oct. 2007.

[250] H. J. McSkimin and P. Andreatch, “Elastic moduli of diamond as a function of pressure

and temperature,” J. Appl. Phys., vol. 43, no. 7, pp. 2944–2948, Jul. 1972.

[251] J. Chang, X. R. Chen, D. Q. Wei, and X. L. Yuan, “Elastic constants and anisotropy of β-

BC2N under pressure,” Phys. B Condens. Matter, vol. 405, no. 17, pp. 3751–3755, 2010.

[252] V. L. Solozhenko, D. Andrault, G. Fiquet, M. Mezouar, and D. C. Rubie, “Synthesis of

superhard cubic BC2N,” Appl. Phys. Lett., vol. 78, no. 10, pp. 1385–1387, Mar. 2001.

[253] J. H. Westbrook and H. Conrad, The Science of Hardness Testing and Its Research

Applications. American Society for Metals, 1973.

191

[254] M. Iuga, G. Steinle-Neumann, and J. Meinhardt, “Ab-initio simulation of elastic constants

for some ceramic materials,” Eur. Phys. J. B, vol. 58, no. 2, pp. 127–133, 2007.

[255] W. R. L. Lambrecht, B. Segall, M. Methfessel, and M. Van Schilfgaarde, “Calculated

elastic constants and deformation potentials of cubic SiC,” Phys. Rev. B, vol. 44, no. 8, pp.

3685–3694, 1991.

[256] R. Hearmon, Numerical data and functional relationships in science and technology, vol.

11. Springer-Verlag, 1984.

[257] D. R. Lide, CRC Handbook of Chemistry and Physics, 84th Edition, 2003-2004, 84th ed.,

vol. 53. CRC Press, 2003.

[258] H. Fu, W. Peng, and T. Gao, “Structural and elastic properties of ZrC under high

pressure,” Mater. Chem. Phys., vol. 115, no. 2–3, pp. 789–794, 2009.

[259] W. Weber, “Lattice Dynamics of Transition-Metal Carbides,” Phys. Rev. B, vol. 8, no. 11,

p. 5082, 1973.

[260] K. Chen and L. Zhao, “Elastic properties, thermal expansion coefficients and electronic

structures of Ti0.75X0.25C carbides,” J. Phys. Chem. Solids, vol. 68, no. 9, pp. 1805–

1811, 2007.

[261] D. G. Clerc and H. M. Ledbetter, “Mechanical hardness: A semiempirical theory based on

screened electrostatics and elastic shear,” J. Phys. Chem. Solids, vol. 59, no. 6–7, pp.

1071–1095, 1998.

[262] V. Podgursky, “Ab initio calculations of elastic properties of isotropic and oriented Ti 1−

x Al x N hard coatings,” J. Phys. D. Appl. Phys., vol. 40, no. 13, pp. 4021–4026, 2007.

[263] P. Ettmayer and W. Lengauer, Nitrides: Transition Metal Solid-State Chemistry.

Chichester, UK: John Wiley & Sons, Ltd, 1994.

[264] J. O. Kim, J. D. Achenbach, P. B. Mirkarimi, M. Shinn, and S. A. Barnett, “Elastic

constants of single-crystal transition-metal nitride films measured by line-focus acoustic

192

microscopy,” J. Appl. Phys., vol. 72, no. 5, pp. 1805–1811, Sep. 1992.

[265] K. Chen, L. R. Zhao, J. Rodgers, and J. S. Tse, “Alloying effects on elastic properties of

TiN-based nitrides,” J. Phys. D. Appl. Phys., vol. 36, no. 21, pp. 2725–2729, Nov. 2003.

[266] W. Wolf, R. Podloucky, T. Antretter, and F. D. Fischer, “First-principles study of elastic

and thermal properties of refractory carbides and nitrides,” Philos. Mag. Part B, vol. 79,

no. 6, pp. 839–858, 1999.

[267] Z. J. Yang et al., “Elastic and electronic properties of fluorite RuO2: From first principle,”

Condens. Matter Phys., vol. 15, no. 1, pp. 1–9, 2012.

[268] J. M. Léger, P. Djemia, F. Ganot, J. Haines, A. S. Pereira, and J. A. H. Da Jornada,

“Hardness and elasticity in cubic ruthenium dioxide,” Appl. Phys. Lett., vol. 79, no. 14,

pp. 2169–2171, 2001.

[269] J. Tse, D. Klug, K. Uehara, Z. Li, J. Haines, and J. Léger, “Elastic properties of potential

superhard phases of RuO2,” Phys. Rev. B - Condens. Matter Mater. Phys., vol. 61, no. 15,

pp. 10029–10034, 2000.

[270] J. Chen, L. L. Boyer, H. Krakauer, and M. J. Mehl, “Elastic constants of NbC and MoN:

Instability of B1-MoN,” Phys. Rev. B, vol. 37, no. 7, pp. 3295–3298, 1988.

[271] M. B. Kanoun, A. E. Merad, G. Merad, J. Cibert, and H. Aourag, “Prediction study of

elastic properties under pressure effect for zincblende BN, AlN, GaN and InN,” Solid.

State. Electron., vol. 48, no. 9, pp. 1601–1606, 2004.

[272] I. Yonenaga, T. Shima, and M. H. F. Sluiter, “Nano-indentation hardness and elastic

moduli of bulk single-crystal AlN,” Japanese J. Appl. Physics, Part 1 Regul. Pap. Short

Notes Rev. Pap., vol. 41, no. 7 A, pp. 4620–4621, 2002.

[273] A. F. Wright, “Elastic properties of zinc-blende and wurtzite AlN, GaN, and InN,” J.

Appl. Phys., vol. 82, no. 6, pp. 2833–2839, Aug. 1997.

[274] X.-J. Chen et al., “Hard superconducting nitrides,” Proc. Natl. Acad. Sci., vol. 102, no. 9,

193

pp. 3198–3201, 2005.

[275] M. Benkahoul, E. Martinez, A. Karimi, R. Sanjinés, and F. Lévy, “Structural and

mechanical properties of sputtered cubic and hexagonal NbNx thin films,” Surf. Coatings

Technol., vol. 180–181, pp. 178–183, 2004.

[276] M. De Jong et al., “Charting the complete elastic properties of inorganic crystalline

compounds,” Sci. Data, vol. 2, p. 150009, Mar. 2015.

[277] S. ‐K Chan et al., “Temperature Dependence of the Elastic Moduli of Monoclinic

Zirconia,” J. Am. Ceram. Soc., vol. 74, no. 7, pp. 1742–1744, Jul. 1991.

[278] K. Kang and W. Cai, “Brittle and ductile fracture of semiconductor nanowires - molecular

dynamics simulations,” Philos. Mag., vol. 87, no. 14–15, pp. 2169–2189, 2007.

[279] B. Minisini, J. Roetting, and F. Tsobnang, “Elastic and thermodynamic properties of OsSi,

OsSi2 and Os2Si3,” Comput. Mater. Sci., vol. 43, no. 4, pp. 812–817, 2008.

[280] I. Yonenaga, “Thermo-mechanical stability of wide-bandgap semiconductors: High

temperature hardness of SiC, AlN, GaN, ZnO and ZnSe,” Phys. B Condens. Matter, vol.

308–310, pp. 1150–1152, 2001.

[281] J. J. Gilman, “Insulator-metal transitions at microindentations,” J. Mater. Res., vol. 7, no.

3, pp. 535–538, 1992.

[282] C. M. Sung and M. Sung, “Carbon nitride and other speculative materials,” Mater.

Chemisry Phys., vol. 43, no. 95, pp. 1–18, 1996.

[283] P. Rodriguez-Hernandez and A. Munoz, “Ab initio calculations of electronic structure and

elastic constants in AlP,” Semicond. Sci. Technol., vol. 7, no. 12, pp. 1437–1440, 1992.

[284] Landolt-Börstein, Semicondutors: Physics of Group IV Elements and III–V compounds

vol. III/17a. Berlin: Springer-Verlag, 1992.

[285] S. Aouadi, P. Rodriguez-Hernandez, K. Kassali, and A. Muñoz, “Lattice dynamics

properties of zinc-blende and Nickel arsenide phases of AlP,” Phys. Lett. Sect. A Gen. At.

194

Solid State Phys., vol. 372, no. 32, pp. 5340–5345, 2008.

[286] T. Azuhata, T. Sota, and K. Suzuki, “Elastic constants of III -V compound

semiconductors: modification of Keyes’ relation Elastic constants of III–V compound

semiconductors: modification of Keyes’ relation,” J. Phys. Condens. Matter, vol. 8, no. 8,

pp. 3111–3111, 1996.

[287] M. J. Weber, CRC Handbook of Laser Science and Technology. Florida: CRC Press, 1987.

[288] J. W. Palko, W. M. Kriven, S. V Sinogeikin, J. D. Bass, and A. Sayir, “Elastic constants of

yttria (Y2O3) monocrystals to high temperatures,” J. Appl. Phys., vol. 89, no. 12, pp.

7791–7796, 2001.

[289] S. Gehrsitz, H. Sigg, N. Herres, K. Bachem, K. Ko hler, and F. K. Reinhart,

“Compositional dependence of the elastic constants and the lattice parameter of AlxGa1-

xAs,” Phys. Rev. B - Condens. Matter Mater. Phys., vol. 60, no. 16, pp. 11601–11610,

1999.

[290] S. Q. Wang and H. Q. Ye, “First-principles study on elastic properties and phase stability

of III - V compounds,” Phys. Status Solidi Basic Res., vol. 240, no. 1, pp. 45–54, Nov.

2003.

[291] G. Samsonov, “Handbook of the Physicochemical Properties of the Elements,” in

Handbook of the Physicochemical Properties of the Elements, Springer, 1968, pp. 387–

446.

[292] D. Tromans, “Elastic Anisotropy of HCP Metal Crystals and Polycrystals,” Ijrras, vol. 6,

no. March, pp. 462–483, 2011.

[293] X. Li and B. Bhushan, “Micro/nanomechanical and tribological studies of bulk and thin-

film materials used in magnetic recording heads,” in Thin Solid Films, 2001, vol. 398–

399, pp. 313–319.

[294] B. D. Beake and J. F. Smith, “High-temperature nanoindentation testing of fused silica

195

and other materials,” Philos. Mag. A Phys. Condens. Matter, Struct. Defects Mech. Prop.,

vol. 82, no. 10, pp. 2179–2186, 2002.

[295] T. K. Harris, E. J. Brookes, and C. J. Taylor, “The effect of temperature on the hardness of

polycrystalline cubic boron nitride cutting tool materials,” Int. J. Refract. Met. Hard

Mater., vol. 22, no. 2–3, pp. 105–110, Jan. 2004.

[296] S. G. Roberts, P. D. Warren, and P. B. Hirsch, “Hardness anisotropies: A new approach,”

Mater. Sci. Eng., vol. 105–106, no. PART 1, pp. 19–28, 1988.

[297] D. R. Trinkle and C. Woodward, “The Chemistry of Deformation: How Solutes Soften

Pure Metals,” Science (80-. )., vol. 310, no. 5754, pp. 1665–1667, 2005.

[298] O. D. Sherby and P. E. Armstrong, “Prediction of activation energies for creep and self-

diffusion from hot hardness data,” Metall. Mater. Trans. B, vol. 2, no. 12, pp. 3479–3484,

Dec. 1971.

[299] H. D. Merchant, G. S. Murty, S. N. Bahadur, L. T. Dwivedi, and Y. Mehrotra, “Hardness-

temperature relationships in metals,” J. Mater. Sci., vol. 8, no. 3, pp. 437–442, 1973.

[300] T. Hirai and K. Niihara, Hot hardness of SiC single crystal, vol. 14, no. 9. Kluwer

Academic Publishers, 1979, pp. 2253–2255.

[301] A. G. Atkins and D. Tabor, “Hardness and Deformation Properties of Solids at Very High

Temperatures,” Proc. R. Soc. A Math. Phys. Eng. Sci., vol. 292, no. 1431, pp. 441–459,

Jun. 1966.

[302] Y. D. Han, H. Y. Jing, S. M. L. Nai, L. Y. Xu, C. M. Tan, and J. Wei, “Temperature

dependence of creep and hardness of Sn-Ag-Cu lead-free solder,” J. Electron. Mater., vol.

39, no. 2, pp. 223–229, 2010.

[303] K. ITO, “Scientific Papers, Sendai,” Tohoku Univ. Ser. 1, vol. 12, p. 137, 1923.

[304] W. Schischokin, “Die Harte und der FlieRdruck der Metalle bei verschiedenen

Temperaturen,” Zeitschrift für Anorg. und Allg. Chemie, vol. 189, no. 1, pp. 263–282,

196

Apr. 1930.

[305] G. E. Hollox, “Microstructure and mechanical behavior of carbides,” Mater. Sci. Eng.,

vol. 3, no. 3, pp. 121–137, Sep. 1968.

[306] K. A. Padmanabhan, “On the nature of the stress function for thermally activated flow,”

Scr. Metall., vol. 7, no. 2, pp. 137–144, 1973.

[307] J. Richter, “The Influence of Temperature on Slip Behaviour of Molybdenum Single

Crystals Deformed in Tension in the Range from 293 to 573 °K .II. Slip Geometry and

Structure of Slip Bands,” Phys. Status Solidi, vol. 46, no. 1, pp. 203–215, Jul. 1971.

[308] L. Kaun, A. Luft, J. Richter, and D. Schulze, “Slip Line Pattern and Active Slip Systems

of Tungsten and Molybdenum Single Crystals Weakly Deformed in Tension at Room

Temperature,” Phys. Status Solidi, vol. 26, no. 2, pp. 485–499, Jan. 1968.

[309] J. A. M. Van Liempt, “Die Berechnung der Auflockerungswarme der Metalle aus

Rekristallisationsdaten.,” Zeit. Phys., vol. 96, no. 7–8, pp. 534–541, 1935.

[310] N. Boden, “Self-diffusion and melting in cubic solids,” Chem. Phys. Lett., vol. 46, no. 1,

pp. 141–145, 1977.

[311] A. M. Brown and M. F. Ashby, “Correlations for diffusion constants,” Acta Metall., vol.

28, pp. 1085–1101, 1980.

[312] J. M. Chezeau and J. H. Strange, “Diffusion in molecular crystals,” Physics Reports, vol.

53, no. 1. pp. 1–92, 1979.

[313] G. B. Gibbs, “Self-diffusion and the melting parameters of cubic metals,” Acta

Metallurgica, vol. 12, no. 5. pp. 673–675, 1964.

[314] F. J. J. van Loo and G. F. Bastin, “On the diffusion of carbon in titanium carbide,” Metall.

Trans. A, vol. 20, no. 3, pp. 403–411, 1989.

[315] M. G. Lozinskii, “CHAPTER III – Methods for measuring the modulus of elasticity,

internal friction and hardness, and for investigating the deformation of metals and alloys at

197

high temperatures in vacuo,” in High Temperature Metallography, 1961, pp. 319–433.

[316] Gerhard Neumann and Cornelis Tuijn, “Introduction,” in SELF-DIFFUSION AND

IMPURITY DIFFUSION IN PURE METALS: HANDBOOK OF EXPERIMENTAL DATA,

vol. 14, Pergamon Materials Series, 2008, pp. 1–35.

[317] R. G. Garlick and H. B. Probst, “Investigation of room-temperature slip in zone-melted

tungsten single crystals,” Trans. Metall. Soc. AIME, vol. 230, pp. 1120–1125, 1964.

[318] J. R. Stephens and W. R. Witzke, “The role of electron concentration in softening and

hardening of ternary molybdenum alloys,” J. Less-Common Met., vol. 41, no. 2, pp. 265–

282, Jul. 1975.

[319] G. Vogl, W. Petry, T. Flottmann, and A. Heiming, “Direct determination of the self-

diffusion mechanism in bcc Beta- titanium,” Phys. Rev. B, vol. 39, no. 8, pp. 5025–5034,

1989.

[320] J. Askill and G. B. Gibbs, “Tracer Diffusion in β‐Titanium,” Phys. status solidi, vol. 11,

no. 2, pp. 557–565, 1965.

[321] H. Ledbetter, H. Ogi, S. Kai, S. Kim, and M. Hirao, “Elastic constants of body-centered-

cubic titanium monocrystals,” J. Appl. Phys., vol. 95, p. 4642, 2004.

[322] C. Kittel, Introduction to solid state physics. Wiley, 2005.

[323] H. Kurishita, K. Nakajima, and H. Yoshinaga, “The High Temperature Deformation

Mechanism in Titanium Carbide Single Crystals,” Mater. Sci. Eng., vol. 54, pp. 177–190,

1982.

[324] Y. Kumashiro, A. Itoh, T. Kinoshita, and M. Sobajima, “The micro-Vickers hardness of

TiC single crystals up to 1500° C,” J. Mater. Sci., vol. 12, no. 3, pp. 595–601, 1977.

[325] G. V. Samsonov et al., “Temperature dependence of hardness of titanium carbide in the

homogeneity range,” Phys. status solidi, vol. 1, no. 2, pp. 327–331, Feb. 1970.

[326] D. L. Kohlstedt, “The temperature dependence of micro- hardness of the transition-metal

198

carbides,” J. Mater. Sci., vol. 8, pp. 777–786, 1973.

[327] G. V Samsonov, “Imperfection of carbon sublattice: Effect on the properties of refractory

carbides of transition metals,” Powder Metall. Met. Ceram., vol. 47, no. 1–2, pp. 13–20,

2008.

[328] M. S. Koval’chenko, V. V. Dzhemelinskii, and V. A. Borisenko, “Temperature

dependence of the hardness of titanium, zirconium, and hafnium carbides,” Strength

Mater., vol. 1, no. 5, pp. 515–518, Nov. 1969.

[329] W. B. Li, J. L. Henshall, R. M. Hooper, and K. E. Easterling, “The mechanisms of

indentation creep,” Acta Metall. Mater., vol. 39, no. 12, pp. 3099–3110, 1991.

[330] I. V. Gridneva, Y. V. Milman, and V. I. Trefilov, “Phase transition in diamond‐structure

crystals during hardness measurements,” Phys. status solidi, vol. 14, no. 1, pp. 177–182,

Nov. 1972.

[331] J. J. Gilman, “Flow of covalent solids at low temperatures,” vol. 46, no. 10, 1975.

[332] J. J. Gilman, “Physical chemistry of intrinsic hardness,” Mater. Sci. Eng. A, vol. 209, no.

1–2, pp. 74–81, 1996.

[333] U. F. Kocks, “Laws for Work-Hardening and Low-Temperature Creep,” J. Eng. Mater.

Technol., vol. 98, no. 1, p. 76, Jan. 1976.

[334] A. Laasraoui and J. J. Jonas, “Prediction of steel flow stresses at high temperatures and

strain rates,” Metall. Trans. A, vol. 22, no. 7, pp. 1545–1558, 1991.

[335] F. R. N. Nabarro, “Philosophical Magazine A Theoretical and experimental estimates of

the Peierls stress,” Philos. Mag. A, vol. 753, no. 3, pp. 703–711, 1997.

[336] J. N. Wang, “A new modification of the formulation of Peierls stress,” Acta Mater., vol.

44, no. 4, pp. 1541–1546, 1996.

[337] D. Kuhlmann-Wilsdorf, “Frictional stress acting on a moving dislocation in an otherwise

perfect crystal,” Phys. Rev., vol. 120, no. 3, pp. 773–781, 1960.

199

[338] G. P. Tiwari, R. S. Mehtrota, and Y. Iijima, “Solid-State Diffusion and Bulk Properties,”

in Diffusion processes in advanced technological materials, Devendra Gupta, Ed.

Norwich, New York: Springer, 2005, p. 69.

[339] J. Philibert, “Some Thoughts and/or Questions about Activation Energy and Pre-

Exponential Factor,” Defect Diffus. Forum, vol. 249, pp. 61–72, 2006.

[340] G. P. Purja Pun and Y. Mishin, “A molecular dynamics study of self-diffusion in the cores

of screw and edge dislocations in aluminum,” Acta Mater., vol. 57, no. 18, pp. 5531–5542,

2009.

[341] G. Neumann and V. Tolle, “Monovacancy and divacancy contributions to self-diffusion in

face-centered cubic metals,” Philos. Mag. A, vol. 54, no. 5, pp. 619–629, 1986.

[342] R. W. Balluffi, “On measurements of self‐diffusion rates along dislocations in F.C.C.

Metals,” physica status solidi (b), vol. 42, no. 1. WILEY‐VCH Verlag, pp. 11–34, 01-Jan-

1970.

[343] A. Rastogi and K. V Reddy, “Self-diffusion in polycrystalline InSb films,” Semicond. Sci.

Technol, vol. 9, pp. 2067–2072, 1994.

[344] J. Mimkes, “Pipe diffusion along isolated dislocations,” Thin Solid Films, vol. 25, no. 1,

pp. 221–230, 1975.

[345] S. Soltani, N. Abdolrahim, and P. Sepehrband, “Molecular dynamics study of self-

diffusion in the core of a screw dislocation in face centered cubic crystals,” Scr. Mater.,

vol. 133, pp. 101–104, 2017.

[346] J. Huang, M. Meyer, and V. Pontikis, “Is pipe diffusion in metals vacancy controlled? a

molecular dynamics study of an edge dislocation in copper,” Phys. Rev. Lett., vol. 63, no.

6, pp. 628–631, 1989.

[347] J. Huang, M. Meyer, and V. Pontikis, “Migration of point defects along a dissociated edge

dislocation in copper: A molecular dynamics study of pipe diffusion,” Philos. Mag. A

200

Phys. Condens. Matter, Struct. Defects Mech. Prop., vol. 63, no. 6, pp. 1149–1165, Jun.

1991.

[348] A. S. Krausz, “On the Analysis of the Activation Volume,” Z. Naturforsch, vol. 31, pp.

728–730, 1976.

[349] K. K. Ray and A. K. Mallik, “An investigation on the activation volume for plastic flow in

nickel determined by micro and macro tests,” Phys. status solidi, vol. 75, no. 2, pp. 451–

458, 1983.

[350] O. D. Sherby, J. L. Robbins, and A. Goldberg, “Calculation of Activation Volumes for

Self-Diffusion and Creep at High Temperature,” J. Appl. Phys., vol. 41, pp. 3961–3968,

1970.

[351] Y.-J. Hu et al., “Effects of alloying elements and temperature on the elastic properties of

W-based alloys by first-principles calculations,” J. Alloys Compd., vol. 671, no. 5, pp.

267–275, 2016.

[352] S. L. Shang et al., “A comprehensive first-principles study of pure elements: Vacancy

formation and migration energies and self-diffusion coefficients,” Acta Mater., vol. 109,

pp. 128–141, 2016.

[353] P. Ehrhart, P. Jung, H. Schultz, and H. Ullmaier, “Atomic Defects in Metals, in: Landolt-

Börnstein, New Series, Group III,” in Atomic Defects in Metals, vol. 1, H. Ullmaier, Ed.

Berlin/Heidelberg: Springer-Verlag, 1991, pp. 385–386.

[354] E. A. Brandes and G. B. Brook, Smithells metals reference book, 7th ed. Oxford: Elsevier

Butterworth-Heinemann, 1992.

[355] G. (Gerhard) Neumann and C. Tuijn, Self-diffusion and impurity diffusion in pure metals :

handbook of experimental data. Amsterdam: Elsevier, 2009.

[356] S. Sarian, “Diffusion of 44Ti in TiCx,” J. Appl. Phys., vol. 40, no. 9, pp. 3515–3520, Aug.

1969.

201

[357] R. Andrievskii, Y. F. Khromov, and I. S. Alekseeva, “Self-Diffusion of Carbon and Metal

Atoms in Zirconium and Niobium Carbides,” Fiz. Met. Met., vol. 32, pp. 5–22, 1971.

[358] B. B. Yu and R. F. Davis, “Self-diffusion of 95Nb in single crystals of NbCx,” J. Phys.

Chem. Solids, vol. 42, no. 2, pp. 83–87, 1981.

[359] L. E. Linda Jones and R. E. Tressler, “The High Temperature Creep Behavior of Oxides

and Oxide Fibers,” 1991.

[360] J. Narayan and J. Washburn, “Self diffusion in magnesium oxide,” Acta Metall., vol. 21,

no. 5, pp. 533–538, 1973.

[361] J. Maier, “Defect Chemistry: Composition, Transport, and Reactions in the Solid State;

Part I: Thermodynamics,” Angewandte Chemie International Edition in English, vol. 32,

no. 3. Hüthig & Wepf Verlag, pp. 313–335, 01-Mar-1993.

[362] T. Südkamp and H. Bracht, “Self-diffusion in crystalline silicon: A single diffusion

activation enthalpy down to 755 C,” Phys. Rev. B, vol. 94, no. 12, p. 125208, 2016.

[363] L. Kalinowski and R. Seguin, “Self-diffusion in intrinsic silicon,” Appl. Phys. Lett., vol.

35, no. 3, pp. 211–213, 1979.

[364] R. F. Peart, “Self Diffusion in Intrinsic Silicon,” Phys. status solidi, vol. 15, no. 2, pp.

K119–K122, 1966.

[365] G. Vogel, G. Hettich, and H. Mehrer, “Self-diffusion in intrinsic germanium and effects of

doping on self-diffusion in germanium,” J. Phys. C Solid State Phys., vol. 16, no. 32, pp.

6197–6204, 1983.

[366] H. D. Fuchs et al., “Germanium Ge70/74Ge isotope heterostructures: An approach to self-

diffusion studies,” Phys. Rev. B, vol. 51, no. 23, pp. 16817–16821, 1995.

[367] M. Werner, H. Mehrer, and H. D. Hochheimer, “Effect of hydrostatic pressure,

temperature, and doping on self-diffusion in germanium,” Phys. Rev. B, vol. 32, no. 6, pp.

3930–3937, 1985.

202

[368] L. Wang et al., “Ga self-diffusion in GaAs isotope heterostructures,” Phys. Rev. Lett., vol.

76, no. 13, pp. 2342–2345, 1996.

[369] T. Y. Tan et al., “Disordering in 69GaAs/71GaAs isotope superlattice structures,” J. Appl.

Phys., vol. 72, no. 10, pp. 5206–5212, 1992.

[370] J.-C. Lee, T. E. Schlesinger, and T. F. Kuech, “Interdiffusion of Al and Ga in

(Al,Ga)As/GaAs superlattices,” J. Vac. Sci. Technol. B Microelectron. Process. Phenom.,

vol. 5, no. 4, pp. 1187–5815, 1987.

[371] B. Goldstein, “Diffusion in compound semiconductors,” Phys. Rev., vol. 121, no. 5, pp.

1305–1311, 1961.

[372] B. Goldstein, “Diffusion in compound semiconductors,” Phys. Rev., vol. 121, no. 5, pp.

1305–1311, 1961.

[373] F. H. Eisen and C. E. Birchenall, “Self-diffusion in indium antimonide and gallium

antimonide,” Acta Metall., vol. 5, no. 5, pp. 265–274, 1957.

[374] H. Bracht, S. P. Nicols, W. Walukiewicz, J. P. Silveira, F. Briones, and E. E. Haller,

“Large disparity between gallium and antimony self-diffusion in gallium antimonide,”

Nature, vol. 408, no. 6808, pp. 69–72, 2000.

[375] V. A. Williams, “Diffusion of some impurities in zinc sulphide single crystals,” J. Mater.

Sci., vol. 7, pp. 807–812, 1972.

[376] M. M. Hexkicberg and A. Stevexsok, “Self‐diffusion of Zn and Se in ZnSe,” Phys. status

solidi, vol. 48, no. 1, pp. 255–269, 1971.

[377] E. A. Secco and C.-H. Su, “Gas-solid exchange reactions : Zinc vapor and polycrystalline

zinc selenide,” Can. J. Chem., vol. 46, no. 10, pp. 1621–1624, 1968.

[378] R. A. Reynolds and D. A. Stevenson, “Self-diffusion of zinc and tellurium in zinc

telluride,” J. Phys. Chem. Solids, vol. 30, no. 1, pp. 139–147, 1969.

[379] E. O. Hall, “The Deformation and Ageing of Mild Steel: III Discussion of Results,” Proc.

203

Phys. Soc. Sect. B, vol. 64, no. 9, pp. 747–753, Sep. 1951.

[380] PETCH and N. J., “The Cleavage Strength of Polycrystals,” J. Iron Steel Inst., vol. 174,

pp. 25–28, 1953.

[381] J. Schiotz and K. W. Jacobsen, “A Maximum in the Strength of Nanocrystalline Copper,”

Science (80-. )., vol. 301, no. 5638, pp. 1357–1359, 2003.

[382] S. Yip, “Nanocrystals: The strongest size,” Nature, vol. 391, no. 6667, pp. 532–533, Feb.

1998.

[383] M. Hakamada et al., “Relationship between hardness and grain size in electrodeposited

copper films,” Mater. Sci. Eng. A, vol. 457, no. 1–2, pp. 120–126, 2007.

[384] H. W. Song, S. R. Guo, and Z. Q. Hu, “Coherent polycrystal model for the inverse Hall-

Petch relation in nanocrystalline materials,” Nanostructured Mater., vol. 11, no. 2, pp.

203–210, 1999.

[385] N. Hansen, “Hall-petch relation and boundary strengthening,” Scr. Mater., vol. 51, no. 8

SPEC. ISS., pp. 801–806, 2004.

[386] C. S. Pande, B. B. Rath, and M. A. Imam, “Effect of annealing twins on Hall-Petch

relation in polycrystalline materials,” Mater. Sci. Eng. A, vol. 367, no. 1–2, pp. 171–175,

Feb. 2004.

[387] L. H. Friedman and D. C. Chrzan, “Scaling Theory of the Hall-Petch Relation for

Multilayers,” Phys. Rev. Lett., vol. 81, no. 13, pp. 2715–2718, 1998.

[388] D. Wu, J. Zhang, J. C. Huang, H. Bei, and T. G. Nieh, “Grain-boundary strengthening in

nanocrystalline chromium and the Hall-Petch coefficient of body-centered cubic metals,”

Scr. Mater., vol. 68, no. 2, pp. 118–121, 2013.

[389] Z. C. Cordero, B. E. Knight, and C. A. Schuh, “Six decades of the Hall–Petch effect – a

survey of grain-size strengthening studies on pure metals,” Int. Mater. Rev., vol. 61, no. 8,

pp. 495–512, Nov. 2016.

204

[390] J. A. Knapp and D. M. Follstaedt, “Hall–Petch relationship in pulsed-laser deposited

nickel films,” J. Mater. Res., vol. 19, no. 01, pp. 218–227, 2004.

[391] R. W. Armstrong, “60 Years of Hall-Petch: Past to Present Nano-Scale Connections,”

Mater. Trans., vol. 55, no. 1, pp. 2–12.

[392] L. L. Shaw, A. L. Ortiz, and J. C. Villegas, “Hall-Petch relationship in a nanotwinned

nickel alloy,” Scr. Mater., vol. 58, no. 11, pp. 951–954, 2008.

[393] M. Kato, “Hall-Petch Relationship and Dislocation Model for Deformation of Ultrafine-

Grained and Nanocrystalline Metals,” Mater. Trans., vol. 55, no. 1, pp. 19–24, 2013.

[394] Y. Li, A. J. Bushby, and D. J. Dunstan, “The Hall-Petch effect as a manifestation of the

general size effect,” Proc. R. Soc. A Math. Phys. Eng. Sci., vol. 472, no. 2190, 2016.

[395] R. W. Armstrong, “Engineering science aspects of the Hall-Petch relation,” in Acta

Mechanica, 2014, vol. 225, no. 4–5, pp. 1013–1028.

[396] C. S. Pande and K. P. Cooper, “Nanomechanics of Hall-Petch relationship in

nanocrystalline materials,” Progress in Materials Science, vol. 54, no. 6. pp. 689–706,

2009.

[397] R. Armstrong, I. Codd, R. M. Douthwaite, and N. J. Petch, “The plastic deformation of

polycrystalline aggregates,” Philos. Mag. A J. Theor. Exp. Appl. Phys., vol. 7, no. 73, pp.

45–58, Jan. 1962.

[398] V. Bata and E. V Pereloma, “An alternative physical explanation of the Hall–Petch

relation,” Acta Mater., vol. 52, no. 3, pp. 657–665, Feb. 2004.

[399] L. H. Friedman and D. C. Chrzan, “Continuum analysis of dislocation pile-ups: Influence

of sources,” Philos. Mag. A, vol. 77, no. 5, pp. 1185–1204, May 1998.

[400] A. Navarro and E. ~R. de Los Rios, “An alternative model of the blocking of dislocations

at grain boundaries,” Philos. Mag. Part A, vol. 57, pp. 37–42, 1988.

[401] A. A. Nazarov, “On the pile-up model of the grain size-yield stress relation for

205

nanocrystals,” Scr. Mater., vol. 34, no. 5, pp. 697–701, Mar. 1996.

[402] A. H. Cottrell, “THEORY OF BRITTLE FRACTURE IN STEEL AND SIMILAR

METALS,” Trans. Met. Soc. AIME, vol. 212, p. 192, 1958.

[403] M. A. Meyersm and E. Ashworth, “A model for the effect of grain size on the yield stress

of metals,” Philos. Mag. A, vol. 46, no. 5, pp. 737–759, Nov. 1982.

[404] A. W. Thompson, M. I. Baskes, and W. F. Flanagan, “The dependence of polycrystal

work hardening on grain size,” Acta Metall., vol. 21, no. 7, pp. 1017–1028, Jul. 1973.

[405] M. F. Ashby, “The deformation of plastically non-homogeneous materials,” Philos. Mag.

A J. Theor. Exp. Appl. Phys., vol. 21, no. 170, pp. 399–424, Feb. 1970.

[406] Z. S. You, L. Lu, and K. Lu, “Tensile behavior of columnar grained Cu with preferentially

oriented nanoscale twins,” Acta Mater., vol. 59, no. 18, pp. 6927–6937, Oct. 2011.

[407] L. Lu, R. Schwaiger, Z. W. Shan, M. Dao, K. Lu, and S. Suresh, “Nano-sized twins induce

high rate sensitivity of flow stress in pure copper,” Acta Mater., vol. 53, no. 7, pp. 2169–

2179, 2005.

[408] O. Anderoglu, A. Misra, H. Wang, and X. Zhang, “Thermal stability of sputtered Cu films

with nanoscale growth twins,” J. Appl. Phys., vol. 103, no. 9, p. 094322, May 2008.

[409] Zachary C. Cordero, Braden E. Knight, and Christopher A. Schuh, “Six Decades of the

Hall-Petch Effect – A Survey of Grain-Size Strengthening Studies on Pure Metals,” Int.

Mater. Rev., vol. 61, no. 8, pp. 495–512, 2016.

[410] P. G. Sanders, J. A. Eastman, and J. R. Weertman, “Elastic and tensile behavior of

nanocrystalline copper and palladium,” Acta Mater., vol. 45, no. 10, pp. 4019–4025, Oct.

1997.

[411] H. Jiang, Y. T. Zhu, D. P. Butt, I. V Alexandrov, and T. C. Lowe, “Microstructural

evolution, microhardness and thermal stability of HPT-processed Cu,” Mater. Sci. Eng. A,

vol. 290, no. 1, pp. 128–138, Oct. 2000.

206

[412] S. . Agnew, B. . Elliott, C. . Youngdahl, K. . Hemker, and J. . Weertman, “Microstructure

and mechanical behavior of nanocrystalline metals,” Mater. Sci. Eng. A, vol. 285, no. 1–2,

pp. 391–396, Jun. 2000.

[413] G. T. Gray, T. C. Lowe, C. M. Cady, R. Z. Valiev, and I. V. Aleksandrov, “Influence of

strain rate & temperature on the mechanical response of ultrafine-grained Cu, Ni, and Al-

4Cu-0.5Zr,” Nanostructured Mater., vol. 9, no. 1–8, pp. 477–480, Jan. 1997.

[414] R. Z. Valiev, E. V. Kozlov, Y. F. Ivanov, J. Lian, A. A. Nazarov, and B. Baudelet,

“Deformation behaviour of ultra-fine-grained copper,” Acta Metall. Mater., vol. 42, no. 7,

pp. 2467–2475, Jul. 1994.

[415] M. Haouaoui, I. Karaman, H. J. Maier, and K. T. Hartwig, “Microstructure evolution and

mechanical behavior of bulk copper obtained by consolidation of micro- and nanopowders

using equal-channel angular extrusion,” Metall. Mater. Trans. A Phys. Metall. Mater. Sci.,

vol. 35 A, no. 9, pp. 2935–2949, 2004.

[416] R. Suryanarayanan Iyer, C. A. Frey, S. M. . Sastry, B. . Waller, and W. . Buhro, “Plastic

deformation of nanocrystalline Cu and Cu–0.2 wt.% B,” Mater. Sci. Eng. A, vol. 264, no.

1–2, pp. 210–214, 2002.

[417] M. Furukawa, Z. Horita, M. Nemoto, and T. G. Langdon, “Processing of metals by equal-

channel,” J. Mater. Sci., vol. 36, pp. 2835–2843, 2001.

[418] Z. Li, C. Hou, M. Huang, and C. Ouyang, “Strengthening mechanism in micro-

polycrystals with penetrable grain boundaries by discrete dislocation dynamics simulation

and Hall-Petch effect,” Comput. Mater. Sci., vol. 46, no. 4, pp. 1124–1134, Oct. 2009.

[419] L. Lu, R. Schwaiger, Z. W. Shan, M. Dao, K. Lu, and S. Suresh, “Nano-sized twins induce

high rate sensitivity of flow stress in pure copper,” Acta Mater., vol. 53, no. 7, pp. 2169–

2179, Apr. 2005.

[420] Q. Fang and F. Sansoz, “Influence of intrinsic kink-like defects on screw dislocation e

207

coherent twin boundary interactions in copper,” 2017.

[421] Z. H. Jin et al., “Interactions between non-screw lattice dislocations and coherent twin

boundaries in face-centered cubic metals,” Acta Mater., vol. 56, no. 5, pp. 1126–1135,

2008.

[422] J. Zhang, H. Zhang, H. Ye, and Y. Zheng, “Twin Boundaries merely as Intrinsically

Kinematic Barriers for Screw Dislocation Motion in FCC Metals,” Sci. Rep., vol. 6, p.

22893, Sep. 2016.

[423] T. Ezaz, M. D. Sangid, and H. Sehitoglu, “Philosophical Magazine Energy barriers

associated with slip-twin interactions Energy barriers associated with slip-twin

interactions,” Philos. Mag., vol. 91, no. 10, pp. 1464–1488, 2011.

[424] Y. T. Zhu, J. Narayan, J. P. Hirth, S. Mahajan, X. L. Wu, and X. Z. Liao, “Formation of

single and multiple deformation twins in nanocrystalline fcc metals,” Acta Mater., vol. 57,

pp. 3763–3770, 2009.

[425] Y. Kamimura, K. Edagawa, and S. Takeuchi, “Experimental evaluation of the Peierls

stresses in a variety of crystals and their relation to the crystal structure,” Acta Mater., vol.

61, no. 1, pp. 294–309, 2013.

[426] G. Schoeck, “The Peierls model: Progress and limitations,” Mater. Sci. Eng. A, vol.

400401, pp. 7–17, 2005.

[427] Y. Xiang, H. Wei, P. Ming, and W. E, “A generalized Peierls-Nabarro model for curved

dislocations and core structures of dislocation loops in Al and Cu,” Acta Mater., vol. 56,

no. 7, pp. 1447–1460, 2008.

[428] M. Shinn, L. Hultman, and S. A. A. Barnett, “Growth, structure, and microhardness of

epitaxial TiN/NbN superlattices,” J. Mater. Res., vol. 7, no. 4, pp. 901–911, Apr. 1992.

[429] Q. Huang et al., “Nanotwinned diamond with unprecedented hardness and stability,”

Nature, vol. 510, no. 7504, pp. 250–253, 2014.

208

[430] T. Irifune, A. Kurio, S. Sakamoto, T. Inoue, and H. Sumiya, “Materials: Ultrahard

polycrystalline diamond from graphite,” Nature, vol. 421, no. 6923, pp. 599–600, Feb.

2003.

[431] D. Bufford, H. Wang, and X. Zhang, “High strength, epitaxial nanotwinned Ag films,”

Acta Mater., vol. 59, no. 1, pp. 93–101, Jan. 2011.

[432] D. Bufford, H. Wang, and X. Zhang, “Thermal stability of twins and strengthening

mechanisms in differently oriented epitaxial nanotwinned Ag films,” J. Mater. Res., vol.

28, no. 13, pp. 1729–1739, 2013.

[433] T. A. Furnish and A. M. Hodge, “On the mechanical performance and deformation of

nanotwinned Ag,” APL Mater., vol. 2, no. 4, p. 46112, 2014.

[434] Y. Kawamura, Y. Fukusumi, K. Hagihara, M. Yamasaki, and T. Nakano, “Non-Basal Slip

Systems Operative in Mg12ZnY Long-Period Stacking Ordered (LPSO) Phase with 18R

and 14H Structures,” Mater. Trans., vol. 54, no. 5, pp. 693–697, 2013.

[435] J. Chen, Y. Du, X. Zeng, L. Peng, W. Ding, and Y. Wu, “Formation of lamellar phase

with 18R-type LPSO structure in an as-cast Mg96Gd3Zn1(at%) alloy,” Mater. Lett., vol.

169, pp. 168–171, Apr. 2015.

[436] H. Iwasaki, H. Tsubakino, A. Yamamoto, S. Hagiwara, Y. Chino, and M. Mabuchi,

“Novel equilibrium two phase Mg alloy with the long-period ordered structure,” Scr.

Mater., vol. 51, no. 7, pp. 711–714, Oct. 2004.

[437] J.-S. Zhang, X.-F. Niu, S.-Z. Wu, C.-X. Xu, Z.-Y. You, and K.-B. Nie, “Microstructure

and mechanical properties of a compound reinforced Mg95Y2.5Zn2.5 alloy with long

period stacking ordered phase and W phase,” China Foundry, vol. 14, no. 1, pp. 34–38,

2017.

[438] X. H. Shao, Z. Q. Yang, and X. L. Ma, “Strengthening and toughening mechanisms in

Mg-Zn-Y alloy with a long period stacking ordered structure,” Acta Mater., vol. 58, no.

209

14, pp. 4760–4771, Aug. 2010.

[439] D. Li, J. Zhang, Z. Que, C. Xu, and X. Niu, “Effects of Mn on the microstructure and

mechanical properties of long period stacking ordered Mg95Zn2.5Y2.5 alloy,” Mater.

Lett., vol. 109, pp. 46–50, Oct. 2013.

[440] L. CAO, Y. WU, L. PENG, Q. WANG, and W. DING, “Microstructure and tribological

behavior of Mg–Gd–Zn–Zr alloy with LPSO structure,” Trans. Nonferrous Met. Soc.

China, vol. 24, no. 12, pp. 3785–3791, Dec. 2014.

210

VITA

Hongyeun Kim was born on September 06th, 1982 in the city of Suwon in South Korea.

He received a B.S. and M.S. in Materials Science and Engineering in 2008 and in 2010 from Yonsei

University in Seoul, South Korea. Also he received second M.S. in Materials Science and

Engineering in 2014 from University of Florida. In 2014, he joined the Phases Research Laboratory

at the Pennsylvania State University towards to obtain his Ph.D. degree. During his study at Penn

State, he has received two Larry-Kaufmann scholarships and five the department Travel Award

grants.

Listed below are his publications during his Ph.D. study:

[1] W. Y. Wang, S. L. Shang, Y. Wang, H. Y. Kim, K. A. Darling, L. J. Kecskes, S. N.

Mathaudhu, X. Hui and Z. K. Liu, “Solid solution hardening Mg-Gd-TM (TM = Ag, Zn and Zr)

alloys: An integrated density functional theory and electron work function study”, JOM, 67

(2015) 2433-2441. http://dx.doi.org/10.1007/s11837-015-1555-9

[2] W. Y. Wang, Y. Wang, S. L. Shang, K. A. Darling, H. Kim, B. Tang, H. C. Kou, S. N.

Mathaudhu, X. Hui, J. S. Li, L. Kecskes, and Z. K. Liu, Strengthening Mg alloys by Self-

dispersed Nano-lamellar Faults. Materials Research Letters, 5 (2017) 415-425.

http://dx.doi.org/10.1080/21663831.2017.1308973

[3] H. Kim, W. Y. Wang, S.L. Shang, L. Kecskes, K. Darling, Z. K. Liu, “Elastic Properties of

Long Periodic Stacking Ordered Phases in Mg-Gd-Al Alloys: A First-Principles Study”,

INTERMETALLICS, 98 (2018) 18-27. https://doi.org/10.1016/j.intermet.2018.04.009

[4] H. Kim, A.J. Ross, S.L. Shang, Y. Wang, L.J. Kecskes and Z.K. Liu, “First-principles

calculations and thermodynamic modelling of long periodic stacking ordered (LPSO) phases in

Mg-Al-Gd”, Materiala, 4 (2018) 192-202. https://doi.org/10.1016/j.mtla.2018.09.013

[5] H. Kim, S. L. Shang, L. J. Kecskes, Z. K. Liu, “Predictive Modeling of Hardness of Brittle

and Ductile Materials”, submitted to Acta Materialia, under review

[6] H. Kim, L. J. Kecskes, Z. K. Liu, “Temperature Dependent Hardness Model: the Study of

Thermally Activated Dislocation Width”, to be submitted

[7] H. Kim, Z. K. Liu, “Hardness Modeling for Layered Structures: The Origin of Hall-Petch

Relation”, to be submitted