Bond Order Potentials for Group IV Semiconductors

183
1 Bond Order Potentials for Group IV Semiconductors A Dissertation Presented to The Faculty of the School of Engineering and Applied Science University of Virginia In Partial Fulfillment of The Requirements for the Degree Doctor of Engineering Physics by Brian Andrew Gillespie January, 2009

Transcript of Bond Order Potentials for Group IV Semiconductors

Page 1: Bond Order Potentials for Group IV Semiconductors

1

Bond Order Potentials for Group IV Semiconductors

A Dissertation Presented to

The Faculty of the School of Engineering and Applied Science

University of Virginia

In Partial Fulfillment of

The Requirements for the Degree

Doctor of Engineering Physics

by

Brian Andrew Gillespie

January, 2009

Page 2: Bond Order Potentials for Group IV Semiconductors

2

Approval Sheet

This dissertation is submitted in partial fulfillment of the requirements for the degree

Doctor of Engineering Physics

Brian Andrew Gillespie

This Dissertation has been read and approved by the Examining Committee:

Leonid Zhigilei, Materials Science, Committee Chair

Vittorio Celli, Physics

James Groves, Material Science

Robert E. Johnson, Engineering Physics

Haydn N.G. Wadley, Materials Science, Advisor

Accepted for the School of Engineering and Applied Science

James H. Aylor, Dean

Page 3: Bond Order Potentials for Group IV Semiconductors

3

Abstract

The atomic scale assembly mechanisms utilized to make semiconductor devices

are difficult to experimentally characterize. Large scale, computer simulations are

beginning to be used to identify these atomistic mechanisms. Computational modeling

of semiconductor systems is complicated by the intricacies and subtleties of the method

by which the materials form atomic bonds. The semiconductors silicon and germanium

are group IV elements which form hybrid sp3 atomic orbital’s to create a diamond cubic

crystalline lattice at standard temperature and pressure. While density functional

methods are well suited for the analysis of the atomic scale structure of these covalent

systems, computational resource limitations prohibit their application to situations

where time dependent reassembly phenomena occur. To model these phenomena,

molecular dynamics methods are used, and these must use an interatomic potential

containing terms to account for the open local atomic environment. Many empirical

interatomic potentials have been proposed over the years; most notable the Stillinger-

Weber and Tersoff potentials. The silicon and germanium parameterizations of these

potentials have been evaluated for their predictive ability for small clusters, melting

temperature, bulk properties for a wide range of crystal structures (dc, sc, fcc, bcc, -Sn,

hcp, and bc8), and the energy of low index surface reconstructions. These potentials are

Page 4: Bond Order Potentials for Group IV Semiconductors

4

shown to give reliable estimates of the bulk properties of the dc phase, but are also

shown to be inadequate for the study of many other structures encountered during the

atomic assembly of both silicon and germanium.

A tight binding description of covalent bonding is used here to propose bond

order potentials (BOP) for the group IV semiconductors silicon and germanium. The

potential addresses both the and bonding of these sp-valent elements. A promotion

energy term associated with the formation of the hybrid orbitals is included in the

formulism. The potential is parameterized using ab-initio and experimental data. The

BOP potential’s predictions for the cohesive energy, atomic volume, and bulk modulus

of the dc, fcc, bcc, bc8, hcp, and -Sn phases of silicon and germanium compare

favorably with estimates obtained using Density Functional Theory (DFT) using local

density approximation (LDA). The BOP also gives point defect formation energies that

are in good agreement with ab initio estimates. The structure of small atomic clusters,

the melting transition temperature and the atomic structure of low index surfaces are

also used to assess the validity of the BOP potential. The functional form of the BOPs for

silicon and germanium are the same, facilitating their eventual use for the study of the

binary Si-Ge alloy system. These improved potentials have then been used for large

scale molecular dynamics simulations of a solid phase epitaxial re-growth of an

amorphous thin film (a prototypical device fabrication process) and the rate limiting

atomic scale mechanism identified.

Page 5: Bond Order Potentials for Group IV Semiconductors

5

The low temperature vapor deposition of silicon thin films and the ion

implantation of silicon often results in the formation of amorphous silicon layers on a

crystalline silicon substrate. These amorphous layers can be crystallized by a thermally

activated solid-phase epitaxial (SPE) growth process. The transformations are rapid and

initiate at the buried amorphous to crystalline interface within the film. The initial stages

of the transformation are investigated here using a molecular dynamics simulation

approach based upon the bond order potential for silicon. The method is used first to

predict an amorphous structure for a rapidly cooled silicon melt. The radial distribution

function of this structure is shown to be similar to that observed experimentally.

Molecular dynamics simulations of its subsequent crystallization indicate that the early

stage, rate limiting mechanism appears to be removal of tetrahedrally coordinated

interstitial defects in the nominally crystalline region just behind the advancing

amorphous to crystalline transition front. The activation barriers for interstitial

migration within the crystal lattice are calculated and found to be comparable to the

activation energy of the overall solid-phase epitaxial growth process simulated here.

Page 6: Bond Order Potentials for Group IV Semiconductors

6

Table of Contents

List of Figures . . . . . . . . . 9

List of Tables . . . . . . . . . 12

List of Terms and Abbreviations . . . . . . 14

1. Introduction . . . . . . . . 16

2. Molecular Dynamics Methods . . . . . . 28

2.1. Stillinger-Weber Potential . . . . . . 32

2.2. Tersoff Potential . . . . . . . 37

2.3. Bond Order Potential . . . . . . 42

3. SW and Tersoff Silicon Assessment . . . . . 52

3.1. Bulk Properties . . . . . . . 53

3.2. Small Clusters . . . . . . . 61

3.3. Point Defects . . . . . . . 69

3.4. Melting Temperature . . . . . . 72

3.5. Surface Reconstructions . . . . . . 74

4. Silicon BOP Assessment . . . . . . . 78

4.1. Bulk Properties . . . . . . . 78

4.2. Small Clusters . . . . . . . 82

Page 7: Bond Order Potentials for Group IV Semiconductors

7

4.3. Melting Temperature . . . . . . 85

4.4. Point Defects . . . . . . . 86

4.5. Surface Reconstructions . . . . . . 87

5. Solid Phase Epitaxy of Silicon . . . . . . 89

5.1. Introduction . . . . . . . . 89

5.2. Simulation Details . . . . . . . 93

5.3. Amorphous Characterization . . . . . 95

5.4. Epitaxial Crystallization . . . . . . 97

5.5. Discussion . . . . . . . . 108

6. SW and Tersoff Germanium Assessment . . . . 110

6.1. Bulk Properties . . . . . . . 111

6.2. Small Clusters . . . . . . . 118

6.3. Melting Temperature . . . . . . 122

6.4. Point Defects . . . . . . . 123

6.5. Surface Reconstructions . . . . . . 125

7. Germanium BOP Assessment . . . . . . 127

7.1. Bulk Properties . . . . . . . 127

7.2. Small Clusters . . . . . . . 132

7.3. Melting Temperature . . . . . . 135

7.4. Point Defects . . . . . . . 136

7.5. Surface Reconstructions . . . . . . 138

8. Discussion . . . . . . . . 140

Page 8: Bond Order Potentials for Group IV Semiconductors

8

8.1. Stillinger-Weber Overview . . . . . . 141

8.2. Tersoff Overview . . . . . . . 142

8.3. Improved Fidelity of the BOP Approach . . . . 144

9. Future Works: SiGe Alloy Parameterization . . . . 150

10. Conclusion . . . . . . . . 153

Appendix A: Mathematica Fitting . . . . . . 156

Appendix B. DFT Calculations . . . . . . 171

Appendix C. Qualitative Analysis Determinants . . . . 175

References . . . . . . . . . 176

Page 9: Bond Order Potentials for Group IV Semiconductors

9

List of Figures

1. Section 1

1.1. SiGe band energy diagram . . . . . . 18

1.2. Periodic Table entries for Si and Ge . . . . . 19

1.3. sp3 orbital hybridization . . . . . . 20

1.4. SiGe film growth . . . . . . . 22

1.5. SiGe critical film thickness . . . . . . 22

1.6. Ion implantation schematic . . . . . . 24

2. Section 2

2.1. Molecular Dynamics flowchart . . . . . 30

2.2. Bond Order Potential electronic hopping paths . . . 43

3. Section 3

3.1. Crystal lattices . . . . . . . 54

3.2. SW and Tersoff Si binding energy curves . . . . 55

3.3. SW and Tersoff vibrational spectrum . . . . 62

3.4. Silicon small clusters . . . . . . . 64

3.5. Point defect structures . . . . . . 72

3.6. Si (100) 2x1 surface . . . . . . . 75

Page 10: Bond Order Potentials for Group IV Semiconductors

10

3.7. Si (113) 3x2 surface . . . . . . . 76

3.8. Si (111) 7x7 surface . . . . . . . 77

4. Section 4

4.1. BOP Si binding energy curves . . . . . 80

4.2. BOP Si vibrational spectrum . . . . . . 82

5. Section 5

5.1. Radial distribution function for a-Si . . . . . 95

5.2. Activation barrier to SPE . . . . . . 97

5.3. SPE snapshots at 700 K . . . . . . 98

5.4. SPE snapshots at 900 K . . . . . . 99

5.5. Number of crystal atoms over time . . . . . 100

5.6. Bond angle distribution in transition region . . . . 101

5.7. Vacancy to crystalline site defect migration . . . . 103

5.8. Tetrahedral to [110]-split defect migration . . . . 104

5.9. Tetrahedral to hexagonal defect migration . . . . 106

5.10. Tetrahedral to crystalline site defect migration . . . 107

6. Section 6

6.1. SW and Tersoff Ge binding energy curves . . . . 112

6.2. SW and Tersoff Ge vibrational spectrum . . . . 117

6.3. Ge (111) c2x8 surface . . . . . . 126

7. Section 7

7.1. BOP Ge binding energy curves . . . . . 130

Page 11: Bond Order Potentials for Group IV Semiconductors

11

7.2. BOP Ge vibrational spectrum . . . . . 131

8. Section 8

9. Section 9

9.1. SiGe phase diagram . . . . . . . 151

10. Section 10

A. Appendix A

A.1 CG-NNL.nb description . . . . . . 157

A.2 fitGSP.nb description . . . . . . 160

A.3 Variation in , and

A.4 Variation in A and

A.5 Variation in r1 and rcut . . . . . . 167

A.6 Variation in r0, rc, rc and rc

Page 12: Bond Order Potentials for Group IV Semiconductors

12

List of Tables

2.1 Stillinger-Weber Parameters . . . . . . 35

2.2 Tersoff Parameters . . . . . . . 40

2.3 Bond Order Potential Parameters . . . . . 50

3.1 Atomic Volume as a function of lattice constants . . . 57

3.2 SW and Tersoff silicon atomic volumes . . . . 57

3.3 SW and Tersoff silicon cohesive energies . . . . 58

3.4 SW and Tersoff silicon bulk moduli . . . . . 60

3.5 SW silicon small clusters . . . . . . 65

3.6 Tersoff silicon small clusters . . . . . . 66

3.7 SW and Tersoff silicon point defects . . . . . 71

3.8 SW and Tersoff silicon surface reconstructions . . . 76

4.1 BOP silicon atomic volumes . . . . . . 81

4.2 BOP silicon cohesive energies . . . . . 81

4.3 BOP silicon bulk moduli . . . . . . 81

4.4 BOP silicon small clusters . . . . . . 84

4.5 BOP silicon point defects . . . . . . 87

4.6 BOP silicon surface reconstructions . . . . . 88

6.1 SW and Tersoff germanium atomic volumes . . . 114

6.2 SW and Tersoff germanium cohesive energies . . . 115

Page 13: Bond Order Potentials for Group IV Semiconductors

13

6.3 SW and Tersoff germanium bulk moduli . . . . 116

6.4 SW germanium small clusters . . . . . 119

6.5 Tersoff germanium small clusters . . . . . 120

6.6 SW and Tersoff germanium small clusters . . . . 124

6.7 SW and Tersoff germanium surface reconstructions . . 126

7.1 BOP germanium atomic volumes . . . . . 128

7.2 BOP germanium cohesive energies . . . . . 128

7.3 BOP germanium bulk moduli . . . . . 129

7.4 BOP germanium small clusters . . . . . 133

7.5 BOP germanium point defects and surface reconstructions . 137

8.1 Qualitative evaluation of all three potentials . . . 146

Page 14: Bond Order Potentials for Group IV Semiconductors

14

List of Terms and Abbreviations

In order of appearance.

dc . . diamond cubic crystal structure

sc . . simple cubic crystal structure

fcc . . face-centered cubic crystal structure

bcc . . body-centered cubic crystal structure

b-Sn . . beta-Tin crystal structure

hcp . . hexagonal close packed crystal structure

bc8 . . bc8 crystal structure

BOP . . Bond Order Potential

DFT . . Density Functional Theory

LDA . . Local Density Approximation

SPE . . Solid Phase Epitaxy

SW . . Stillinger-Weber interatomic potential

LED . . Light Emitting Diode

p-MOSFET . p-type Metal-Oxide-Semiconductor Field Effect Transistor

a-Si . . amorphous silicon

c-Si . . crystalline silicon

MD . . Molecular Dynamics

EAM . . Embedded Atom Method

Page 15: Bond Order Potentials for Group IV Semiconductors

15

TB . . Tight Binding

IP . . Interatomic Potential

GSP . . Goodwin Skinner Pettifor pair term

HF . . Hartree Fock

T, H and X . Tetrahedral, Hexagonal and (110)-Split Interstitials

DAS . . Dimer Adatom Stacking fault for (111) 7x7 surface

EDIP . . Environment Dependent Interatomic Potential

RDF . . Radial Distribution Function

VASP . . Vienna Ab-initio Simulation Package

IBIEC . . Ion Beam Induced Epitaxial Crystallization

ZB . . Zinc Blende crystal structure

Page 16: Bond Order Potentials for Group IV Semiconductors

16

I. Introduction

Silicon is an abundant element composing ~25% of the Earth’s crust by mass in

the form of silica and various silicates [1]. It is one of the most intensely studied

materials in the Periodic table because of its many engineering applications [2]. Silicon,

in the form of silica and silicates, is vital to the construction industry as a principle

constituent of natural stone, glass, concrete and cement [3]. Silicon is also commonly

alloyed with aluminum to produce easily cast metallic alloy parts for industry [4]. Silicon

even sees service in the children’s toy Silly Putty, which contains significant amounts of

elemental silicon (silicon binds to the silicone and allows the material to bounce 20%

higher) [5]. However, silicon’s greatest technological impact derives from its use in

microelectronic applications. Elemental silicon is the principle component of most

semiconductor devices, most importantly integrated circuits. Ultra-pure silicon can be

doped with other elements to adjust its electrical response by controlling the number

and polarity of its charged current carriers [2]. Such control is necessary for solid state

transistors, solar cells, integrated circuits, microprocessors, and other semiconductor

devices which are used in microelectronics.

Page 17: Bond Order Potentials for Group IV Semiconductors

17

Germanium, the group IV element located directly below silicon in the periodic

table, is also a material of great scientific and engineering interest. The development of

the germanium transistor ushered in the era of modern microelectronics [6]. The

germanium transistor is considered by many to be the greatest invention of the

twentieth century [7]. Despite silicon’s superior electrical properties, germanium was

the original semiconductor of choice because silicon required much higher purity;

purities that could not be achieved commercially at the time [2]. Germanium continues

to be of interest to modern microelectronics because of its unique combinations of

electrical and other physical properties. For instance, germanium is an indirect band

gap material with a small band gap. It therefore has low light absorbance at infrared

wavelengths, and because it is easily cut and polished, it can be used for infrared lenses

and windows in the 8-14 micron wavelength range [8-9]. The small bandgap of Ge and

Ge-Si alloys also make it a useful material in solar cell applications where its high

absorbance enable the use of thinner layers of active material [10-11]. Light emitting

diodes (LEDs) have also been fabricated to take advantage of germanium’s unique

properties [12-13]. These LEDs are based on Ge-Si alloy self-assembled quantum dots,

and have exhibited a broad emission peaked at a wavelength centered upon 1.45 m

[12-13].

The performance of traditional silicon-based semiconductor technologies is also

being extended through the use of SixGe1-x alloys. These performance increases can be

tied to the differences in physical and electronic properties. For example, while

elemental silicon has a band gap of 1.12 eV [14], the inclusion of germanium reduces

Page 18: Bond Order Potentials for Group IV Semiconductors

18

this to ~0.93 eV for Si0.5Ge0.5. A

graph showing the effect of

germanium additions upon the

band gap of SixGe1-x is given in

Figure 1.1 (the red line). In Figure

1.1 (the green line) the intrinsic

carrier resistivities are shown for

SixGe1-x [14]. This greatly increases

the number of intrinsic charge carriers with the semiconductor. For example, at 300 K,

silicon has an intrinsic carrier concentration of 1.0x1010 cm-3, germanium has 2.4x1013

cm-3, and Si0.5Ge0.5 has 1.2x1013 cm-3 [14]. Another key advantage of SiGe alloys are their

high electron hole mobility. At room temperature (300 K) silicon has a hole mobility of

450 cm2/V*s and electron mobility 3,900 cm2/V*s, while Si0.5Ge0.5 has a hole mobility of

1,175 cm2/V*s and an electron mobility of 7,700 cm2/v*s. A recent SiGe p-MOSFET

device designed by Masashi Shima displayed an increase in hole mobility of around 30%

compared with a similar device built using silicon technology alone [15]. Interest in SiGe

microelectronics has also been aided by a realization that the massive investment in

existing silicon foundry fabrication tools can be used for most SiGe alloy process steps

[16].

Both silicon and germanium are group IV metalloids with a valence state of s2p2;

the Periodic Table entries for silicon and germanium are shown in Figure 1.2 [1]. When

Page 19: Bond Order Potentials for Group IV Semiconductors

19

two or more atoms of the same or similar electronegativities react, a complete transfer

of electrons does not occur [17]. Instead the atoms achieve noble gas valence shell

electronic configurations by sharing pairs of electrons. This form of interatomic bonding

is called covalent bonding [17]. Silicon and germanium form covalent bonds between

themselves and with each other and with the common group III and V dopants [17].

Covalent bonds are characterized by their strength and directional dependence.

In order to fill the s and p valence electron shell, silicon and germanium need to borrow

four electrons. However, in the s2p2 electronic ground state, silicon and germanium are

only capable of forming 2 covalent bonds. The s2p2 electronic orbitals hybridize into the

sp3 configuration to create four electron states capable of forming four equivalent

covalent bonds [17]. Figure 1.3 shows the structure of the sp3 configuration. The

tetrahedral arrangement of these covalent bonds leads to the equilibrium, ambient

temperature and pressure condensed phase of silicon and germanium to be the

diamond cubic structure [1].

This dissertation has been motivated by interest in the growth processes of

silicon, germanium and their alloys. The growth of silicon crystals from the

Page 20: Bond Order Potentials for Group IV Semiconductors

20

condensation of a vapor has been thoroughly studied [18-21]. Temperature plays an

important role in vapor phase growth processes. The structure of a semiconductor film

grown at low temperature is often amorphous (i.e no long range order in the solid

phase). As the substrate temperature is increased, the mobility of atoms during the

growth process increases, and the structure becomes more crystalline; and in the case

of silicon, assumes the diamond cubic structure. An experimental study performed by

Zalm et al. at the Philips Research Laboratories in the Netherlands found that a

minimum yield of 40% crystalline 60% amorphous was obtained when depositing Si at a

rate of 0.1 nm/s at a substrate temperature of ~650 K [18]. Further experiments at

higher temperature (~745 K and 835 K) showed a significant increase in crystallinity.

These results are consistent with other works [19-21]. The growth of crystalline silicon

generally proceeds by a planar (step flow) mode as opposed to an island growth scheme

[22]. It is possible, however, to grow silicon in a columnar structure by altering the

angle of incidence of adatom flux as shown experimentally by Xie et al. [22]. They found

that columnar structures grew in the same direction as the adatom incidence angle.

They also found that the column width was dependent on the deposition temperature

Page 21: Bond Order Potentials for Group IV Semiconductors

21

(~40 at 300 K and ~130 at 425 K, note that these temperatures are a relative

measurement as the exact surface temperature was not experimentally measured) [22].

When germanium condenses from the vapor phase the structure that is formed

depends on the temperature of the condensate [23-24]. The structure of germanium

deposited at low temperature is amorphous. As temperature is increased, the volume

fraction of the grown film that is crystalline increases [23-24]. The growth mechanism

of germanium on silicon is very similar to the homoepitaxy of Si for the first 3

monolayers [23-24]. Germanium grows in a 2-D planar method for the first three

monolayers, after which the growth proceeds via the formation island formation on the

substrate [23-24]. This is due to the 4% lattice mismatch between Si and Ge and the

resulting strain in the lattice [25-26]. A deposition temperature of at least 525 K is

needed to ensure a 95% crystalline growth region [27], and typically a ~900 K anneal is

performed to smooth out the surface.

The growth of SiGe films proceeds via a 2-D (planar) mechanism or a 3-D

(islanding) mechanism [28]. Figure 1.4 details the temperature and composition for the

two types of growth. Strain effects prevent unlimited planar growth [28]. It has been

found that at a given temperature and germanium fraction, SiGe can be grown on Si up

to a critical thickness before islanding begins [28]. For example, Si0.2Ge0.8 can be grown

in a planar fashion up to ~2500 Å, whereas Si0.5Ge0.5 can be grown up to ~100 Å before

the onset of misfit dislocation formation and islanding [29]. Figure 1.5 displays

experimental and theoretical results correlating germanium fraction to critical thickness.

Page 22: Bond Order Potentials for Group IV Semiconductors

22

Substrate temperature for crystalline deposition varies from the temperature for

homoepitaxy of Si. The inclusion of Ge degrades the crystallinity of deposition at low

temperatures (~650 K) [18].

The crystalline quality of the silicon and germanium grown films is of great

importance because the crystalline structure generally has more desirable electronic

characteristics [2]. Fortunately the amorphous phases of both silicon and germanium

are less stable than the crystalline structures and as a result there exists driving force, in

the form of Gibbs free energy, for the rearrangement of the amorphous material to the

diamond cubic structure [30]. The spontaneous rearrangement of amorphous material

to crystalline material, also known as solid phase epitaxy when it occurs on crystal

substrates, is temperature dependent and well described by an Arrhenius equation [30].

In the case of high purity silicon films for example, the interface between the two solid

Figure 1.4 Plot of Si1-xGex film morphology

vs. growth temperature and germanium

fraction.

Figure 1.5 Plot of critical thickness of

Si1-xGex films vs. germanium fraction.

Experimental data taken from

reference 28.

Page 23: Bond Order Potentials for Group IV Semiconductors

23

phases has been observed to move at a velocity of several m/s at 725 °C [30]. The

interface velocity is directly dependent on the temperature; at higher temperatures the

interface moves more rapidly, and at lower temperatures the interface moves more

slowly [30].

While silicon, germanium and their alloys are intrinsic semiconductors, the true

strength of their use in the semiconductor industry is derived by the inclusion of

controlled defects within the lattice [2]. These defects are usually introduced in the

form of dopant elements. These dopants greatly increase the number of charge carriers

(electrons or holes) within the semiconductor and are a necessary component for most

modern electronic devices [2]. The most commonly employed method for the

introduction of dopants into the semiconductor is through ion implantation processes

[31]. A typical ion implantation setup is diagrammed in Figure 1.6. The ion implantation

process typically involves a high purity ion source, where ions of the desired element are

produced, an accelerator, where the ions are eletrostatically accelerated to high energy,

and a target chamber containing the silicon or germanium substrate. An energetic ion

entering a solid interacts with the atoms of the solid in two principle ways. Atoms are

displaced from their lattice sites and set in motion through momentum transferring

collisions. In addition, electrons are excited by Coulomb interactions with the moving

atoms. These two types of energy losses are often referred to as nuclear and electronic

energy loss respectively. Atoms set in motion by nuclear collisions of the primary ion

may collide with other atoms and set them in motion and these, in turn, can do the

same to others [30]. This sequence is typically called a collision cascade.

Page 24: Bond Order Potentials for Group IV Semiconductors

24

The common feature of all ion

implantation processes in a solid is the

production of lattice disorder (even for self-

implantation) [30]. The amount of disorder

generated is controlled by the ion energy,

mass, dose (the total number of ions

implanted), dose rate (the rate at which

ions are implanted) and the substrate

temperature [30]. The disorder can be so

severe that the implanted region becomes amorphous. Recovery of this damage is

required to restore the crystalline structure and to electrically activate the dopants [2].

In particular, the transition from amorphous silicon (a-Si) to crystalline silicon (c-Si) is a

process of great technological importance. In solid phase epitaxy (SPE), the

amorphous/crystalline interface propagates by a thermally induced epitaxial crystal

growth into the metastable amorphous region [30]. The SPE process has been the

subject of many research efforts in recent years [30-40]. The great challenge to

experimentalists is that the SPE process occurs within the bulk of the semiconductor

material which remains inaccessible to surface experimental probes [35]. This difficulty

has limited the usefulness of experimental methods to investigate the SPE process to

measurements of macroscopic quantities such as the activation energy for growth.

While the growth from the vapor phase of silicon, germanium and their alloys is

well studied; a better understanding of the atomic scale assembly mechanisms that

Page 25: Bond Order Potentials for Group IV Semiconductors

25

occur during vapor deposition and the atomic mechanisms responsible for the

rearrangement of atoms at the amorphous/crystalline interface in solid phase epitaxy

would be helpful to development of new techniques for the growth of semiconductor

thin films. A fundamental study of the ion impact process and atomic reassembly

induced by it would also be of benefit.

Molecular dynamics (MD) techniques have demonstrated the ability to identify

and capture the atomistic mechanisms responsible for experimental phenomenon [41-

45]. Molecular dynamics simulations make use of interatomic force calculations

obtained using interatomic potentials. These potentials coarse grain the quantum

mechanical behavior of the many electrons into a force law that relates the interatomic

force to the local atom configuration. As a result, any fundamental insights that can be

derived from the examination of the electronic ground state are sacrificed. In return,

interatomic potentials allow for the simulation of a sufficient number of atoms to

realistically approximate atomic assembly processes requiring thousands of atoms over

a nanosecond time scale. The investigation of these atomistic assembly mechanisms

inherently involves considerations of interatomic forces, therefore kinetic Monte Carlo

simulation methods which can accommodate many more atoms are untenable.

Molecular dynamics has been used successfully in studies of the surfaces of

metals during their assembly from the vapor phase [41-45]. The approach is most

precisely implemented with fcc metals where the bond forces are dependent only on

the nearest neighbor distance (r) and are not dependent on the angle () formed by an

Page 26: Bond Order Potentials for Group IV Semiconductors

26

atom with its nearest neighbors. The simulated growth of close packed metallic

multilayers has utilized spherically symmetric embedded atom method (EAM) potentials

to examine hot atom and ion assisted deposition and surfactant mediated growth [43-

44]. In the case of ion assisted growth it was shown that inert gas ion beams could be

used to aid in the growth of Cu/Ni/Cu multilayers by smoothing out the interfaces [44].

It was discovered that even low energy (~3 eV) inert gas impacts could cause atomic

exchange at the Cu/Ni interface [44]. It was shown that by modulating the ion beam

assistance it is possible to significantly reduce the deleterious intermixing, while at the

same time enabling ion assisted flattening [44]. These observations were subsequently

corroborated experimentally [45]. Despite these successes, the EAM potentials are not

suitable for modeling the complex covalent bonding observed in silicon and germanium

systems because it treats only omnidirectional bonding.

Numerous groups have attempted to incorporate the physical concepts

underlying covalent bonding in many-body interatomic potentials [46-54]. This has led

to semi-empirical sets of equations that attempt to approximate the phenomenological

nature of the bond. An assessment for the GaAs system has shown that this approach

results in mixed success [55]. In an alternative approach, Pettifor et al. have shown that

it is possible to derive an analytic, many-body interatomic potential by coarse graining

the electronic structure within the orthogonal two-center tight-binding (TB)

representation of covalent bonding [56-59]. These analytic bond order potentials

(BOPs) explicitly link the bond order (and therefore the bond energy) to the positions of

Page 27: Bond Order Potentials for Group IV Semiconductors

27

atomic neighbors. Applications of the approach to the GaAs system and to

hydrocarbons systems have given encouraging results [60].

This dissertation explores the application of the BOP formalism developed by

Pettifor and co-workers to elemental silicon and germanium. The resulting interatomic

potentials are assessed by comparison of their predictions with those of the widely used

Stillinger-Weber (SW) and Tersoff potentials as well as experimental and ab initio data.

Each potential is assessed for its ability to predict cohesive energies, elastic moduli,

atomic volumes, small clusters, defect formations energies, melting temperature and

surface reconstructions. The potentials for both Si and Ge have been formulated in a

manner that will simplify the future development of SiGe potentials. The utility of the

silicon potential is investigated by examining the atomistic mechanisms responsible for

the rearrangement of atoms at the amorphous/crystalline interface during the early

stages of solid phase epitaxy.

Page 28: Bond Order Potentials for Group IV Semiconductors

28

II. Molecular Dynamics Methods

If the interatomic forces between the atoms of an ensemble can be written as a

function of atom separation, the solution to Newton’s equations of motion for the

ensemble provide a detailed description of the ensembles atomic coordinates, the

vibrational states (and the excursions of these that result in atom migration), and the

time varying forces acting upon all the atoms in the system. This approach is known as

molecular dynamics (MD) and it was first introduced by Alder and Wainwright in the

1950’s [61, 62]. System properties such as pressure [63], the internal energy [63], the

viscosity [64], diffusion rates [64], and the specific heat [65] of atomic ensembles can be

well predicted using this method. The approach numerically integrates Newton’s

equations of motion to track the particle positions as a function of time. The emergence

of powerful modern computers enables large and/or complicated systems to be

simulated. One such example is the simulation of supersonic crack propagation in two-

dimensional crystal lattices by Farid Abraham et al. at IBM [66]. Another is the work of

V. V. Zhakhovskii et al. [67] in which extreme overheating is applied to a planar structure

in vacuum, consisting of between 105 and 106 atoms, causing it to expand.

Page 29: Bond Order Potentials for Group IV Semiconductors

29

The theory behind MD can be traced back to Gibbs in 1902 [68]. In his landmark

work “Elementary Principles in Statistical Mechanics”, Gibb’s introduces the idea of a

canonical ensemble. A canonical ensemble refers to an infinite number of systems of a

constant volume V, at a constant temperature T, each containing N atoms. In such an

ensemble, the properties are allowed to change over time, and can vary from one

system to another system. Despite the statistical variation it is possible to calculate

observable properties via two different methods. The first is to calculate the property at

a fixed time for all the systems. The average value would be the observable property of

the ensemble. The second method takes the average of the property over a long period

of time for a single system in the ensemble. Given an evolution in time that sampled all

potential structures the second method gives the same value as the first. MD is an ideal

approach for studying short time dynamics. It can also be used (using the methods

described above) to calculate the average (observable) properties of a system.

The general method involved in molecular dynamics is displayed in a flowchart in

Figure 2.1. The first step is to create the system under consideration. This involves

setting the initial position and velocities for all the atoms in the system. Next the forces

between the atoms are calculated using an interatomic potential describing the forces

between atoms in terms of their local configuration. In the third step the atoms are

moved by integrating the equations of motion over the time step chosen for the system.

A periodic boundary condition is then applied to move atoms back into the central

Page 30: Bond Order Potentials for Group IV Semiconductors

30

simulation cell. A desired property is then calculated from the new atom positions. The

program iterates this process until the total number of time steps is equal to the time

limit set.

In principle, any analytic or numerical method for calculating the forces between

atoms allowing the equations of motion to be solved can be employed in MD

simulations. Many such methods have been proposed, including density functional

theory [69], tight binding [70], and interatomic potentials [71]. The computational

expense of each method versus the level of approximation is often the deciding factor in

the selection of a modeling regime. Because of the need to simulate a moderate

number of atoms (in the thousands) in a dynamic fashion, this dissertation has

employed interatomic potentials.

Figure 2.1 Flowchart displaying the basic logical pathways in the MD process.

Setup initial

configuration

Calculate force

on each atom

Simulation time

exceeded?

Average time-dependent property values

to obtain accurate property values YES

NO

Calculate the desired

properties

Integrate

equations

of motion

Page 31: Bond Order Potentials for Group IV Semiconductors

31

Interatomic potentials abandon the time consuming calculation of the electronic

ground state in favor of empirically generated equations to capture the interaction of

atoms [71]. These potentials consist of a set of empirical functions dependent upon

atomic positions and “fitting” parameters. These functions and parameters can give a

complete description of system energy and forces acting on each particle (excluding

energetic interactions within individual atoms).

Recent work on non-empirical, physically motivated bond order potentials (BOP)

by David Pettifor and his collaborators has expanded the nature of the classical

interatomic potential [56-60]. The functional form of the BOP has been derived from

orthogonal TB and draws heavily on the molecular orbital bonding there. The BOP

format includes extrinsic terms to account for both s and p contributions to covalent

bonding as well as a term to account for the orbital hybridization. This format holds

great promise for atomistic modeling due to its ability to operate on a timescale close to

that of empirical IPs yet still retaining an element of the electronic interactions which

lend predictive validity to TB methods. It must be noted that the predictive validity of

these interatomic potentials is inextricably tied to the quality of their parameterization.

The parameterization process, or potential fitting, is by no means an easy process. In

fact the bulk of the time spent during the course of this research has been dedicated to

the fitting of the silicon and germanium BOP potentials. Great care must be employed

to ensure that modeling predictions are reasonable. This requires rigorous evaluations

of an IP’s transferability across a wide range of properties.

Page 32: Bond Order Potentials for Group IV Semiconductors

32

2.1 Stillinger-Weber Potential

The Stillinger-Weber (SW) many-body interatomic potential was proposed in

1985 in response to the growing interest in the utilization of computer simulation

modeling techniques. Previously developed pair potentials, such as the Lennard-Jones

potential [72], are incapable of stabilizing the equilibrium diamond cubic structure of

silicon under ambient conditions [73]. These pair potentials invariably prefer close

packed crystalline phases [73]. The silicon system presented interesting challenges,

most notably, unlike most other elemental systems, silicon is observed to increase in

density when it melts [1]. Diffraction experiments have shown that the melting process

causes the crystal structure to collapse, substantially increasing the coordination from 4

to an average value in excess of 6 [74-76]. This process is also accompanied by a shift in

the electronic properties of the element from semiconductor to metal [77]. The SW

potential (or family of potentials) was introduced to model these condensed phases of

silicon with the desire to predict details of the change in local order for tetrahedral

semiconductors as they melt [46].

The SW potential remains one of the more widely used interatomic potential

formats for the simulation of open structure materials. This is due to its combination of

ease of implementation and reasonable predictive validity. The potential employs a

common approximation approach for the interaction energy of a system of N identical

particles in which the energy is expressed in terms of sums of all possible one-body (i),

Page 33: Bond Order Potentials for Group IV Semiconductors

33

two-body (i, j), three-body (i, j, k), etc. interactions terms. Thus, the potential energy is

represented by:

2.1

This representation depends heavily on the rapid convergence of the energy with

increasing interaction order. The SW potential truncates the form of Eq. 2.1 to include

only the two and three-body terms [46]. Some modifications/extensions of the SW

format include the four-body term; however these potentials will not be discussed here

[78].

The two-body term, f2, is required to be a function only of radial distance, and is

written:

2.2

Where the multiplicative combination of parameters a and set the range of the

potential. B is an adjusted value that determines the location of the potential minimum,

and A is set such that the value of f2 at the potential minimum is -1. The parameter is

a scaling parameter. The values chosen for simulation of silicon are A = 7.04956, B =

0.60222, = 2.0951 Å, = 2.1702 eV, a = 1.8, p = 4, and q = 0 [46] (values with no unit

are unitless). The potential naturally approaches a value of 0 at r = a·, thus avoiding

Nji

N

jii

NjifjififU

),(),()( 21

0

exp2)(2

ar

rrB

A

rf

qp

ar

ar

Page 34: Bond Order Potentials for Group IV Semiconductors

34

the need for a separate cutoff function. The smooth approach to 0 aids the potential in

avoiding the generation of artifacts in the simulation [46].

The three-body term, f3, is required to have rotational and translational

symmetry [46]. It is expressed as a symmetrized sum:

2.3

Where h(ijk) is dependent on the atomic separation distances rij and rjk, and the angle

ijk. Given both radii are less than a·, h has the following form:

2.4

If either rij or rjk are greater than a·, h = 0. The angular term is strongly preferential in

favor of the tetrahedral configuration yet does not preclude the possibility of alternative

geometries [46]. The SW potential assigns values of 21 and 1.2 to the parameters and

respectively [46] (both are unitless).

The SW potential for Si was parameterized using a least mean squares fitting

routine with an emphasis on several criteria [46]. First and foremost was the

requirement that the diamond cubic structure be the most energetically stable at zero

temperature and pressure [46]. This is of course the nature of real silicon [1]. Secondly

the potential was required to predict in “reasonable accord” the melting point and

liquid structure [46]. In these two criteria the SW potential is quite successful in

reproducing the desired experimental properties of silicon [46]. The potential, however,

)()()()(3 jkihjikhijkhijkf

23

1cosexp),,(

ijk

jkij

ijkjkijarar

rrh

Page 35: Bond Order Potentials for Group IV Semiconductors

35

does not predict a -Sn transition at high

pressure [63,79]. This raises concerns

regarding the transferability quality of the

structural and mechanical energy calculations.

The SW potential has been further employed

to examine many features of the elemental Si

system, and a few of these are discussed here.

The SW parameters for silicon and germanium

can be found in Table 2.1.

Feuston, Kalia, and Vashishta performed a detailed molecular dynamics study of

the silicon microclusters using the SW potential [80]. Their results indicated the

presence of unusually stable “magic number” clusters of size 4, 6, and 10. This result is

mirrored in photofragmentation experiments performed by Bloomfield et al. [81].

While the energetic stability of the magic number clusters matches that found in

experiment, the geometric structure of small silicon clusters of size n = 3-6 are

inaccurate [82,83]. The geometric structures preferred by the SW potential for clusters

of size 3, 4, 5, and 6 are the equilateral triangle, square, pentagon, and two equilateral

triangles directly on top of each other respectively [80]. Experiment shows the proper

geometric structures to be a bent chain, rhombus, trigonal bipyramid and distorted

octahedron [82,83].

Table 2.1 Parameters for silicon and

germanium for the Stillinger-Weber

potential

Si Ge

A 7.04956 7.04956

B 0.60222 0.60222

Å 2.0951 2.181

eV 2.1702 1.93

a 1.8 1.8

p 4 4

q 0 0

21 31

1.2 1.2

Page 36: Bond Order Potentials for Group IV Semiconductors

36

Khor and Das Sarma employed the SW potential to examine low index surfaces

of silicon [84,85]. Of particular interest was the investigation of the (100) surface [84]

and the dispute (at the time) between buckled and symmetric dimers [86-89]. Also of

interest was the Pandey defect reconstruction model which had been observed using

scanning tunneling microscopy [84]. The SW potential was found to provide an accurate

description of the symmetric dimers on the (100) surface and that the Pandey-type

defect reconstruction may be favorable in energy compared to the dimer reconstruction

[84]. The SW potential was unable to predict the presence of buckled dimers even as

higher energy metastable structures [84].

Ding and Andersen extended the SW potential format to germanium and

employed it to examine amorphous germanium [54]. They attempted to parameterize

the SW format to accommodate three separate phases of germanium: the liquid, the

diamond lattice crystalline solid, and the amorphous solid. However, they found no

parameter set that simultaneously gave a good description for all three phases of

interest [54]. As a result they settled on a good description of the amorphous and

crystalline solid phases. The resulting parameter set for Ge is: A = 7.04956, B = 0.60222,

p = 4, q = 0, a = 1.8, = 31, = 1.2, = 1.93, and = 2.181 [54]. This potential is shown

to accurately reproduce radial distributions in the amorphous and crystalline phase of

germanium as well as a reasonable prediction (within 10%) of the phonon dispersion

curves [54].

Page 37: Bond Order Potentials for Group IV Semiconductors

37

2.2 Tersoff Potential

The Tersoff interatomic potential was proposed in 1986 with the intent to

develop a potential with which it was feasible to calculate the structure and energetics

of complex covalently bonded systems [47-49]. Tersoff recognized that previous

interatomic potentials, SW included, did not attempt to describe accurately the

properties of nontetrahedral forms of silicon [47]. The potential format that Tersoff

proposed was motivated by intuitive ideas about the dependence of bond order upon

the local environment. The Tersoff potential represents the first such potential to

incorporate the bond order in the functional format, however empirically [47]. The

Tersoff potential is modeled off a Morse-type pair potential, which allows a physical

interpretation of the potential parameters [47-49].

As one of the earliest interatomic potentials to become readily adopted, the

Tersoff potential had a fairly simple functional format. Over the course of a few years,

Tersoff continued to make minor adjustments to the functional format [47-49]; the

format employed in this research and presented here is that published in Ref. 49. If one

were to consider a system of atoms, the total potential energy of the system, U, could

be captured as the sum of the energy of the individual bonds in the system. The Tersoff

potential [49] represents this empirically with:

ji

ijAijijRijc rfbrfrfU )()()(2

1

2.5

Page 38: Bond Order Potentials for Group IV Semiconductors

38

The function fc is a cutoff function designed to restrict the range of the potential. The

first term in square brackets in Eq. 2.5 is interpreted to represent the repulsive

electrostatic force that two atoms encounter when brought within close proximity to

each other. The second term is interpreted to represent the bonding energy between

two atoms. This term is adjusted by bij which includes considerations of the bond order

and the local environment.

The functions fR and fA represent repulsive and attractive pair potentials. The

choice of exponential functional forms for these, as in a Morse potential, was based on

the “universal” bonding behavior discussed by Ferrante, Smith, and Rose [90, 91]. They

had shown that a large number of calculated binding energy curves could be mapped

onto a single dimensionless curve which could be well described by the assumption of a

Morse-type pair potential [90, 91]. This lead to fR and fA to have the form:

ijijijijR rArf exp)(, 2.6

ijijijijA rBrf exp)( , 2.7

where Aij and Bij are the geometric average of the fitted parameters A and B for atoms i

and j. The parameters ij and ij are the average of the fitted parameters and for

atoms i and j.

The function bij represents the only deviation from a potential that is otherwise

wholly pair-wise [47-49]. It represents a measure of the interatomic bond order and has

been written:

Page 39: Bond Order Potentials for Group IV Semiconductors

39

iii nn

ij

n

iijijb2

1

)1(

, 2.8

)()(,

ijkikik

jik

cij grf

, 2.9

22

2

2

2

))cos((1)(

ijkii

i

i

i

ijkhd

c

d

cg

. 2.10

where ijkis the bond angle between bonds ij and ik. The parameter ij has a value of

unity if between two atoms of the same type, for a bond between silicon and

germanium this term has a value of 1.00061. This small adjustment was included to

properly simulate the heat of formation of the zinc-blende structure [49]. Similarly the

parameter ij also has a value of unity, and was included in the potential for possible

future flexibility. The parameters , n, c, d, and h are all dependent only on atom i.

Values for these parameters for the silicon and germanium systems are listed in Table

2.2 [49].

Unlike the Stillinger-Weber potential, the Tersoff functional format includes an

explicit cutoff term [47-49]. This step function is designed to create a smooth transition

to between the separation ranges of R and S.

ijij

ijijij

ijij

ijij

ijij

ijC

rS

SrRRS

Rr

Rr

rf

0

)cos(2

1

2

1

1

)( .

Page 40: Bond Order Potentials for Group IV Semiconductors

40

The Tersoff parameters for Si-Si,

Ge-Ge, and Si-Ge interactions are listed

in Table 2.2. The 11 unknowns per

material (parameters) are fitted to 7

different materials properties: cohesive

energies of real and hypothetical bulk

crystals (exact structures are not

specified in the publication [49]), bulk

modulus, diamond bond length, and

elastic constants c11, c12 and c44

constrained to within 20%. Tersoff justified the use of a small database during the

fitting by subsequently comparing the resulting potential against a much larger database

to verify its suitability [47-49].

The Tersoff potential has been widely used for molecular dynamics research due

to the ease of its implementation and its physical motivation (albeit empirically

formulated). For example, Dyson and Smith utilized the Tersoff potential in the

examination of the (001) and (111) 2x1 surface reconstructions of diamond [92]. They

compared the Tersoff predictions to other interatomic potentials; the Stillinger-Weber

and two proposed modifications to the Tersoff potential by Brenner [53, 92]. They

concluded that while the Tersoff potential gave better results than the SW potential, it

was still not in qualitative agreement with other work. The Tersoff-Brenner potentials

Table 2.2 Parameters for silicon and

germanium for the Tersoff potential.

Si Ge

A (eV) 1830.8 1769

B (eV) 471.18 419.23

Å) 2.4799 2.4451

Å) 1.7322 1.7047

1.1 x 10-6 9.0166 x 10-7

n 0.78734 0.75627

c 1.0039 x 105 1.0643 x 105

d 16.217 15.652

h -0.59825 -0.43884

R Å) 2.7 2.8

S Å) 3.0 3.1

Page 41: Bond Order Potentials for Group IV Semiconductors

41

[53] performed significantly better in surface property prediction, and they cautioned

the use of empirical potentials fitted to bulk properties when examining surfaces [92].

Other research, such as that performed by Mura et al., has shown more

promising results [93]. Mura et al. employed the Tersoff potential to examine the

structural trends in SiC amorphous alloys [93]. They were able to characterize the local

coordination found in amorphous Si1-xCx as a function of the alloy composition. They

concluded that for high carbon content the disorder is limited by the distortion of the

silicon sublattice [93]. Tarus and Nordlund used the Tersoff potential in the simulation

of germanium surface segregation during deposition of Si on Ge/Si (001) surfaces [94].

Their molecular dynamics research uncovered a surface segregation mechanism

involving the movement of two atoms at the surface. This mechanism, they observed, is

thermally activated with a fairly short time scale for the process (~1 ns) [94].

Of particular interest to this research is the research of Motooka of Kyushu

University in Japan [31, 95]. Motooka investigated amorphization and crystallization

processes in ion implanted silicon [31, 95]]. Using the Tersoff potential and cross-

sectional transmission electron microscopy he determined that amorphization in a

crystalline lattice can be induced when the concentration of the di-vacancy and di-

interstitial pairs exceeds a certain threshold value [31, 95]. He also observed that

epitaxial growth mechanisms at the amorphous/crystalline interface are quite similar to

crystal growth from the melt [31, 95]. The solid phase epitaxial growth of silicon will be

discussed later in section V along with calculations of the process using the BOP.

Page 42: Bond Order Potentials for Group IV Semiconductors

42

2.3 Bond Order Potential

If an interatomic potential for a covalently bonded material is to be used to

simulate assembly of the condensed state, it must reliably model the radial and angular

dependence of the dynamic interatomic forces present during the formation and

breaking of atomic bonds. Germanium has 2 s and 2 p electrons and its covalent

bonding within the condensed phases can be described using sp-hybridized atomic

orbitals whose overlap results in bonding and anti-bonding molecular orbitals with or

symmetry [17]. The strength of a covalent bond (its bond order) can be characterized

as one-half the difference between the number of electrons in bonding and anti-

bonding molecular orbitals within the bond [96]. Bond orders, when multiplied by bond

(hopping) integrals, then yield bonding energies [97]. The bond integrals are related to

the probability that an electron will hop between atoms, and so depends upon the

interatomic separation and angle and the overlapping orbital type [98]. Figure 2.2

illustrates the specific hopping paths that are directly accounted for within the BOP.

A general analytic expression for the reduced (separate and bond-order

contributions) BOP derived from two-center orthogonal tight binding theory has been

detailed by Pettifor and his collaborators [56-59]. While this analytic bond order

potential might at first appear complicated, it is composed of several terms with straight

forward physical interpretations. The bond order potential seeks to describe the real

space potential energy function E(r,) (where r is the interatomic separation and an

angular dependence) for a collections of atoms. If the atoms are bound by sp hybridized

Page 43: Bond Order Potentials for Group IV Semiconductors

43

electron states, this potential can be written as the sum of three terms; an electrostatic

repulsive term (Urep), an attractive bonding term (Ubond), and a promotion energy term

(Uprom) associated with creation of the sp3 hybridized orbitals:

prombondrep UUUE 2.12

The atoms in the condensed phases of group IV elements such as germanium can

form bonds with their neighbors that have either (primary covalent) or (secondary

covalent) bonding characteristics. The bonding term in Eq. 2.12 (Ubond) is therefore a

sum of the separate and contributions to the bonding energy:

i ijj

ijijijijbondU,

,,,, 222

1 , 2.13

where and are the and bond (hopping) integrals between atoms i and j. The

coefficients and are the bond orders of the and bonds between atoms i and j.

Figure 2.2 Hopping integration paths of length 2 (a and d), and length 4 (b, c and e)

that contribute to the potential energy of the bond between atoms I and j. The

position of atoms k and k’ determine the local environment of the bond.

Page 44: Bond Order Potentials for Group IV Semiconductors

44

The bond integrals are radially dependant functions of the atomic separation between

atoms i and j [56]. The bond orders are dependent on the local environment of the ij

bond [56]. This methodology connects well with the single, double, and triple bond

terminology from chemistry [17]. It also results in maximum allowable values for the s

and p orbital contributions to the bond order. These are 1 for bonds, and 2 for

bonds.

The repulsive potential energy component of Eq. 1, Urep, approximates the

electrostatic repulsion between the atomic cores using a simple pair potential :

i ijj

ijrepU,2

1 , 2.14

The and bond integrals, and , found in Eq. 2.13, and the repulsive pair potential

found in Eq. 2.14 can be writing in a similar fashion:

n

rf0, 2.15

n

rf0, , 2.16

mrf0 2.17

where f(r) is a radially dependent Goodwin-Skinner-Pettifor (GSP) function [99], 0,

0,and 0 are fitted parameters that control the magnitude of the bond energy, and

nand n are fitted parameters that control the rate at which and respectively

approach zero as r increases. The value m/n is a measure of the hardness of the

Page 45: Bond Order Potentials for Group IV Semiconductors

45

repulsive potential. It is fully soft in the limit m/n = 1 and totally hard for m/n = ∞

[100]. For group IV sp valent materials such as silicon and germanium, a m/n value of

~2 is expected [100]. The germanium potential described here has a fitted m/n ratio of

1.88.

The GSP function is a pairwise function derived from the TB expression for the

binding energy [99] of the form:

cc n

c

n

c r

r

r

r

r

rrf 00 exp , 2.18

where r is the interatomic spacing between bonded atoms, r0 is the spacing at which f(r)

= 1, rc is a characteristic radius, and nc is a characteristic exponent. We have assumed

that the parameters r0, rc, and nc have identical values for the three pair functions in Eqs.

2.15-2.17. This simple constraint significantly eased the subsequent parameterization

process used to fit the potential framework to the condensed states of Ge and did not

decrease the potential’s performance.

The bond order, in Eq. 2.13, has been derived for half-full valence shell

systems such as germanium and later generalized to include systems of generalized

valence band occupancy [56-59]. The half-full valence shell bond-order has been

employed here and has the derived form:

2/1

2

4

42242, ~

1

~2

~~2

1

jiij

ij

R 2.19

Page 46: Bond Order Potentials for Group IV Semiconductors

46

where

, 2.20

. 2.21

The 2 term in Eq. 2.19 describes the self-returning, second moment two-hop

contributions to a bond between atoms i and j, and is illustrated in Fig. 2.2a. This term

has the form:

jik ij

ik

jikij g,

2

,

,2

,,2

2.22

where g is a function introducing angular dependant contributions to the bond order

resulting from orbital overlap. It has the form:

1cos1, jikjik pg 2.23

where ijk is the angle, centered on atom i, between the atom pairs ij and ik. The term

g has a single free parameter, p. This parameter has physical significance. For s

orbital bonding, because of radial symmetry, there is no angular dependence (p = 0);

for p orbitals, because of the orthogonalality, the energy minimum is found at 90°

angles and therefore the function g has a cos form (p = 1) [100]. This parameter

ji

224

4

4

~

ji

ji

ji

224

22

22

~~

Page 47: Bond Order Potentials for Group IV Semiconductors

47

allows for the potential to control the hybridized atomic orbital overlap dependency

[100]. The angular term can be optimized for the sp3 diamond cubic structure by

choosing p = 0.75

The remaining two terms in Eq. 2.19 represent hopping paths of length four. The

4term describes the self-returning hopping paths of length four linking atoms i and j

while the R4 term represents a 4-atom ring-type interference path linking atoms i and j.

These terms are illustrated in Fig. 2.2b-c respectively and are written in the form:

2.24

,

jikk ij

jk

ij

kk

ij

ik

kijjkkkikjikij ggggR,, ,

,

,

,

,

,

,,,,4,

2.25

The analytic expression of the bond order for a half full valence shell with hops

of length two and four can be written as:

ij, 2.26

where

jik ij

ki

ij

ik

ijkkkijik

jik ij

ik

jik gggg,

2

,

,

2

,

,

,,,

,

4

,

,2

,4

jikk ij

kk

ij

ik

kikjik gg,,

2

,

,

2

,

,2

,,

Page 48: Bond Order Potentials for Group IV Semiconductors

48

21

,4

,2,2

21

ij

jiij

2.27

The 2 terms are defined as:

jikk ij

ik

jikikij

,,

2

,

,2

,,2 2sinˆ

2.28

where the angle jik is described in Fig. 2.2d and

2

,

,

2

,

,

ij

ik

ij

ik

ik p

2.29

The fitting parameter p is species dependant only on the central atom i (for

multicomponent systems). The f4 term includes four-body dihedral angles, fkk’, that

are influential in p type bonding.

kjijik

jikk

kiikij

22

,,

2

,

2

,,4 sinsinˆˆ

kijjikkjik 222

,

2

, sinsinˆˆ 2.30

kjiijkjkki 222

,

2

, sinsinˆˆ

4

2cossinsinˆˆ 222

,

2

,kk

kijijkkjjk

where the angles are defined in Fig. 2.2e and the dihedral angle is defined as

Page 49: Bond Order Potentials for Group IV Semiconductors

49

1sinsin

coscoscos22cos

22

2

kjijik

jikkjikki

kk

2.31

The promotion energy term in Eq. 2.12 represents the energy penalty incurred by

elevating a s level electron into a p level atomic orbital and reconfiguring the atomic

orbitals into the hybrid sp3 configuration found in the diamond structure. Following

reference 56, the promotion energy can be approximated by:

212

,11

ij

ij

prom AU

2.32

where and A are fitting parameters.

To summarize, the format of the BOP employed herein uses a potential energy

function (Eq. 2.12) that is expressed in terms of bonding (Eq. 2.13), repulsive (Eq. 2.14)

and promotion energy (Eq. 2.32) components. The repulsive component is purely

pairwise in nature whereas the bonding component has been further subdivided into

and terms. The and bonding elements are the products of bond integrals (Eq. 2.15

and 2.16) and bond orders (Eq. 2.19 and 2.26) respectively. The bond order (Eq. 2.19)

contains hopping paths of length 2 and 4 including a 4-atom ring-type interference path

(Eq. 2.25). The bond order (Eq. 2.26) contains hopping paths of length 4 and includes

dihedral angle effects.

There are a total of 15 free parameters that need to be fitted for the BOP

defined above. The best estimates of these 15 parameters are given in Table 2.3. There

Page 50: Bond Order Potentials for Group IV Semiconductors

50

are 6 GSP parameters (r0, rc, m, n, n and nc), 3 prefactors (repulsive 0, bonding ,0

and ,0), 2 angular terms (p and p), 2 promotion energy parameters ( and A) and 2

cutoff parameters (r1 and rcut). These have been determined for the BOP using a

method adapted from the work of Albe et al. [101].

A systematic three-step process for free parameter fitting first determined the

pair function (the bond energy of cubic structures and dimers can be expressed solely as

a function of bond length) and then in a second step incorporated the angular function

parameters. Because the promotion energy term as described in previous publications

Table 2.3 Parameters for silicon and germanium for the Bond Order Potential. The

silicon potential uses term sensitive values for the rc and nc parameters.

Symbol Quantity Value

Ge Si

r0 GSP reference radius (Å) 2.32 2.349

rc GSP characteristic radius (Å) 3.634 : 14.932 : 11.675 : 2.771

m GSP repulsive exponent 6.187 8.742

n GSP attractive exponent 3.405 2.944

n GSP attractive exponent 1.012 4.040

nc GSP decay exponent 10.0 : 2.523 : 6.533 : 29.482

r1 Spline start radius (Å) 3.2 3.1

rcut Spline cutoff radius (Å) 3.7 3.6

0 Repulsive energy prefactor (eV) 2.415 1.012

,0 bond integral prefactor (eV) 2.171 2.224

,0 bond integral prefactor (eV) 0.299 0.196

p bond 3-body angular term 0.75 0.732

p p bond 3-body angular term 0.1 0.1

Promotion energy term (eV) 6.605 10.179

A Promotion energy term 0.798 1.852

Page 51: Bond Order Potentials for Group IV Semiconductors

51

[56-59] cannot be expressed in the form of a pair function, it could not be employed

with the first two steps of the fitting. The third step therefore introduced the

promotion energy term. The methodology for the first two steps of this approach is

discussed in detail in Appendix A of reference 102. The promotion energy contributions

to the various structures were determined using a least mean squares fitting routine.

This enabled incorporation of a physically meaningful promotion energy term into the

potential following an iterative process of parameter optimization. A detailed look at

the methodology of this fitting process is presented in Appendix A.

There exist a great many different methods of employing DFT and each method

often predicts slightly different energies for similar systems. Because of this, it was

necessary to construct a consistent database of bulk material properties (including

cohesive energy (Ec in eV), atomic volume (Va in Å3), lattice constants (a and c in Å), and

bulk moduli (B in GPa)) of various equilibrium and metastable silicon and germanium

phases. These calculations were performed with the local density approximation (LDA)

of DFT using the VASP DFT package. A description of the details of these VASP

calculations can be found in Appendix B. In the fitting of silicon and germanium higher

priority (weights) were given to the atomic volume and cohesive energy of the diamond

cubic and -Sn phases since these are well established silicon and germanium phases.

Accurate structural and energetic predictions of surface energies and defect formation

energies were considered important. An accurate simulation of theoretical bulk phases

(such as fcc and bcc) and the configuration (but not the energy) of clusters were given a

lower weight, and as a result are less well predicted.

Page 52: Bond Order Potentials for Group IV Semiconductors

52

III. SW and Tersoff Silicon Assessment

It would generally be considered unwise to employ an interatomic potential to

study the atomistic mechanisms of a material system if it is known that the potential’s

predictions of those properties are poor. It is therefore necessary to perform a robust

assessment of an interatomic potential to determine which properties a potential is well

suited to model. This research is motivated by an interest in atomic assembly; therefore

the material properties assessed are all ones which can be tied in some way to atomic

assembly processes. The following properties have been examined for a wide range of

structures with varying atomic coordination (sc, fcc, bcc, dc, -Sn, hcp and bc8): atomic

volume, cohesive energy and bulk modulus. The following bulk properties of the

diamond cubic phase have also been calculated: the elastic constants (c11, c12 and c44),

the phonon vibration spectrum, the Cauchy pressure and the melting temperature (Tm).

The structure and energetics of small silicon clusters (Sin with n=2-6), defect formation

energies and the energy of low index surface reconstructions have also been examined.

This chapter details the simulation methods employed to calculate the

properties listed above using the Stillinger-Weber and Tersoff potentials for the silicon

Page 53: Bond Order Potentials for Group IV Semiconductors

53

material system. Subsequent chapters will forgo explanations of calculation methods

and merely present data and comments in a similar format. Examination of the SW and

Tersoff Ge predictions can be found in chapter V, while the BOP predictions of these

properties can be found in chapters IV and VI for Si and Ge respectively.

3.1 Bulk Properties

An essential requirement of an interatomic potential is a proper prediction of

the equilibrium phase under ambient conditions. Failure to meet this critical

requirement will jeopardize the confidence in predictive validity that is an essential

component of modeling research. Furthermore, the ability of a potential to model the

local bonding environment can be inferred by examination of the structure of various

other phases where the atomic coordination ranges from 4 to 12. The structures that

have been selected are as follows (abbreviations and atomic coordination are listed in

parenthesis): simple cubic (sc, 6), face centered cubic (fcc, 12), body centered cubic (bcc,

8), diamond cubic (dc, 4), -Sn (4), hexagonal close packed (hcp, 12) and bc8 (4). In the

case of silicon, the lowest energy equilibrium phase at room temperature and

atmospheric pressure is the dc structure. The unit cell for each of these structures and

the DFT predicted lattice constants can be found in Figure 3.1.

The bulk properties (atoms volume, cohesive energy and bulk modulus) were all

calculated concurrently. Crystal files (labeled “r” files) were generated using a

Page 54: Bond Order Potentials for Group IV Semiconductors

54

Mathematica program developed by Dewey Murdick called CG-NNL (this program will

be discussed in more detail in a chapter IX) for each of the seven examined crystal

structures. These crystals were then minimized in energy by inducing strains on the

crystal to find the lattice constant a (a and c in the case of -Sn and hcp). In this way we

constructed the binding energy curves, useful graphical tools for determining the fidelity

of a potential. From these curves the atomic volume, cohesive energy and bulk modulus

g)-Sn

Page 55: Bond Order Potentials for Group IV Semiconductors

55

Figure 3.2 Binding energy vs. relative volume (with respect to equilibrium dc

volume) for a range of Si bulk structures. a) Stillinger-Weber b) Tersoff c) DFT.

dc

-Sn

bc8

sc

bcc

fcc

hcp

Page 56: Bond Order Potentials for Group IV Semiconductors

56

can be extracted. The binding energy curves for silicon as predicted by SW, Tersoff and

DFT can be found in Figure 3.2.

The binding energy curves are a quick way of examining a potential’s viability

over a wide range of structures. As such, they are the first property that has been

examined for each potential. The binding energy curves give a wide spectrum visual

overview of a potential’s predictions of energy, volume and moduli of varying phases

simultaneously. As can be observed from Figure 3.2 the SW potential for silicon predicts

with reasonable accuracy the relative energy differences between the bulk lattice

phases (the minimums of the curves lie approximately where DFT calculations indicate

they should). The SW potential, however, does not predict proper volume relations for

the phases. The relative volumes of the curves differ significantly from the DFT

calculations. The binding energy curves also reveal that the SW potential incorrectly

places -Sn and bc8 phases in reversed order [79]. As pressure increases (relative

volume V/V0 decreases) the bc8 phase curve should cross the dc phase curve first

followed by the -Sn phase curve [63]. The SW potential predicts the opposite. The

Tersoff potential slightly overestimates the energy differences between structures. The

most notable failure of the Tersoff potential binding energy curves is that the bc8 phase

curve does not at any point become the lowest energy. Instead the sc phase finds itself

located between the dc and -Sn phase curves. This suggests that the Tersoff potential

is unlikely to reproduce the -Sn to bc8 phase transition properly.

Page 57: Bond Order Potentials for Group IV Semiconductors

57

The volume of an atom in a given crystal

structure can be written as a function of the lattice

constants of that structure. Table 3.1 lists these

functions for each of the crystal structures examined.

The SW and Tersoff predictions of the atomic volumes

are listed in Table 3.2. Also included are the

percentage differences from the DFT calculations. The

average percent difference in atomic volume predicted

by the SW potential is ~17%. Only the dc and bc8

phases are predicted within 10% of the DFT

calculations. The Tersoff predictions are notably

superior, with all but the hcp phase predicted within

10%. These results are not encouraging for the

potential’s transferability to those phases.

The cohesive energy of an atom is defined as the energy required to remove an

atom from the bulk and move it beyond interaction range (i.e. infinity) [98]. This can be

Table 3.2 Atomic Volume Predictions (Å3)

Structure dc sc fcc bcc bc8 -Sn hcp

DFT 19.79 0.789 0.704 0.717 0.869 0.749

SW 20.013 0.889 0.892 0.854 0.902 0.861 0.936

% diff 1.13 13.91 28.04 20.42 4.92 16.10

Tersoff 20.013 0.818 0.739 0.735 0.918 0.764 0.937

% diff 1.13 4.78 6.09 3.61 6.80 3.09

Table 3.1 Atomic Volume as a

function of lattice constants

structure Va

sc 3a

fcc 4

3a

bcc 2

3a

dc 8

3a

-Sn 4

2ca

hcp ca 2

4

3

bc8 16

3a

Page 58: Bond Order Potentials for Group IV Semiconductors

58

found by taking the total sum of the energy of a system of simulated bulk atoms and

dividing by the total number of atoms in the system. Table 3.3 contains the cohesive

energies for the DFT, SW and Tersoff predictions, as well as the percentage differences

between the DFT and interatomic predictions. DFT calculations indicate an energy-

structure trend from minimum to maximum of dc, bc8, -Sn, sc, hcp, bcc and fcc. The

SW potential captures the energy trends well; the order of structures from minimum to

maximum energy is the same as the DFT predictions with the exception of the bcc phase

and hcp phase switching places. The Tersoff potential switches the bcc and hcp phases

as well, and more importantly also switches the sc and b-Sn phases. Of particular note

is the SW prediction of the dc phase cohesive energy. The SW potential predicts a

cohesive energy of -4.34 eV for the dc phase of silicon. Experimental measurements

have found that the energy is -4.63 eV [1]. Stillinger and Weber address this issue in

their original publication and recognize that a rescaling of certain parameters (A and B)

would allow for a proper energy prediction [46]. They instead chose to prioritize the

solid-liquid transition temperature. They also indicate that augments to the potential in

Table 3.3 Cohesive Energy Predictions (eV)

Structure Dc sc fcc bcc bc8 -Sn hcp

DFT -4.63 0.292 0.468 0.451 0.122 0.213

SW -4.34 0.274 0.396 0.281 0.191 0.20 0.305

% diff 6.26 6.27 5.24 2.87 7.96 6.27

Tersoff -4.63 0.319 0.761 0.432 0.247 0.329 0.522

% diff 0.0 0.62 7.04 0.44 2.77 2.63

Page 59: Bond Order Potentials for Group IV Semiconductors

59

the form of single particle, position independent terms could be used to correct the

cohesive energy [46].

The bulk modulus can be calculated by considering the sound derivative of the

total energy with respect to strain near the equilibrium. The bulk modulus can be

defined [98] as

2

2

9

1

tot

c

E

VB 3.1

where Vc is the unit cell volume and a strain parameter. A similar method is used to

obtain the elastic constants c11 and c44 where

2

2

11

1

tot

c

E

Vc and 3.2

2

2

444

1

tot

c

E

Vc . 3.3

The elastic constant c12 can be calculated from B and c11 (c12 = (3B-c11)/2) [98]. A more

in-depth description of the method used can be found in a book by M.W. Finnis entitled

“Interatomic Forces in Condensed Matter” [98].

The bulk moduli predictions can be found in Table 3.4. Of the bulk properties

examined, the bulk moduli are the poorest predicted by the SW and Tersoff potentials.

Both the SW and Tersoff potential’s predictions of the dc bulk modulus are in

reasonable accord with the DFT calculation. The other structures, however, do not fare

Page 60: Bond Order Potentials for Group IV Semiconductors

60

as well. The average percent difference of the prediction from the DFT is 240% and

182% for the SW and Tersoff respectively. It should be noted however that the Tersoff

average is skewed heavily by an aberrantly large prediction of the fcc bulk modulus. In

general, the Tersoff predictions are in good agreement with DFT.

The three independent elastic constants (c11, c12 and c44) have also been

calculated for the dc structure. Silicon has values of 165.78, 63.94 and 79.62 GPa for c11,

c12 and c44 (relaxed) respectively [8]. The SW potential predicts corresponding values of

90.32, 107.69 and 39.55 GPa. The Tersoff potential predicts values of 139.77, 77.26 and

29.01 GPa. These predictions lie within 50% of the elastic moduli, a result which is

within the acceptable range limitations found within Tight Binding models. The Cauchy

pressure (c12 – c44) of the strong covalently bonded material silicon is negative (-16 GPa)

[8]. This is because the high hardness and stiffness of silicon arise not from a high bulk

modulus, but from a high shear modulus due to strong resistance to interatomic

bending [98]. The SW and Tersoff predictions are 68.14 and 48.25 GPa respectively.

The prediction of a positive Cauchy pressure is a common feature among many

empirical interatomic potentials [98].

Table 3.4 Bulk Modulus Predictions (GPa)

Structure dc sc fcc bcc bc8 -Sn hcp

DFT 105.4 105.6 93.54 111.3 128.1 155.9

SW 101.9 190.6 714.5 742.7 74.8 410.2 322.4

% diff 3.32 80.49 663.8 567.3 41.6 163.1

Tersoff 98.1 132.6 1180.9 153.3 103.9 143.4 139.9

% diff 6.92 25.57 1162.45 37.74 18.89 8.02

Page 61: Bond Order Potentials for Group IV Semiconductors

61

The phonon spectrum of the diamond cubic phase was also calculated. This

probes the performance of the potential at near equilibrium bond spacing conditions. A

diamond cubic crystal of 512 atoms was annealed at a simulated temperature of 300 K

and the velocities of a randomly selected 50 atom sample were tracked and used to

calculate the velocity-velocity autocorrelation function. The vibrational spectrum for

the system was then calculated by taking the Fourier transform of this correlation

function. The resulting vibrational spectrum is shown in Figure 3.3. Silicon exhibits a

single strong peak at 520 cm-1 [103]. The SW and Tersoff potentials predictions both

exhibit strong peaks in the 500-540 cm-1 range, however the Tersoff potential shows a

primary peak at 110 cm-1.

3.2 Small Clusters

The calculation of the structure and energetics of small silicon clusters is

straightforward. The xyz position data for many high symmetry configurations of each

cluster size was input into the MD code and a molecular statics minimization was

performed. The resulting structure and bond lengths are recorded and the total energy

of the system is the binding energy of the small cluster. The investigation of silicon

clusters helps characterize the nature of atomic bonding as predicted by the

corresponding potential because the structure and bond energies of small clusters are

heavily reliant on both the angular and radial component of the interatomic potential.

Thus examination of cluster properties is a useful gauge of a potential’s transferability.

Page 62: Bond Order Potentials for Group IV Semiconductors

62

Examined herein are the structure of various silicon small clusters containing up to six

atoms by means of the SW and Tersoff potentials.

A wide range of high symmetry small clusters have been examined for silicon.

These are: the dimer, the trimer: chain (D∞h), bent chain (C2v), equilateral triangle

(D3h), tetramer: chain (D∞h), square (D4h), rhombus (C2v), flagged triangle (c2v),

tetrahedron (Td), pentamer: pentagon (d5h), pyramid (c4v), trigonal bipyramid (d3h),

hexamer: octahedron (Oh), edge-capped trigonal bipyramid (c2v), face-capped trigonal

Page 63: Bond Order Potentials for Group IV Semiconductors

63

bipyramid (c2v), and bent-chair hexagon (c2v). The structures examined are the same

structures examined in references 104 and 105. These structures are pictured in Figure

3.4 along with the Hartree-Fock predicted bond lengths [104, 105]. The Hartree-Fock

method (HF) is an approximate method for the determination of the ground state wave

function and energy of a many-body system. It is often used as the starting point in

many ab initio studies of molecules. The SW and Tersoff potential predictions for the

structure and energies of these small clusters are listed in Tables 3.5 and 3.6 for the SW

results and Tersoff results respectively.

The dimer is the simplest of the silicon clusters examined (and in fact the

simplest silicon cluster possible). It consists solely, by definition, of 2 silicon atoms

bound together. The angular dependence terms found within an interatomic potential

have no influence on the determination of the dimer properties. Experimental studies

have shown the bond length of the silicon dimer to be 2.246 Å and the binding energy to

be -2.41 eV. HF calculations predict a bond length of 2.265 Å and binding energy of -

3.06 eV [104, 105]. The SW potential predicts a bond length of 2.35 Å and a binding

energy of -2.17 eV. The Tersoff potential predicts a bond length of 2.313 Å and a

binding energy of -2.62 eV. The SW potential overestimates the dimer bond length by

4.7% and underestimates its energy by 10% (although much of the energy difference is

due to the scaling of the energy parameters to better match the melting temperature as

previously mentioned). The Tersoff potential overestimates the dimer bond length by

2.9% and overestimates the binding energy by 8.7%. These results are representative of

Page 64: Bond Order Potentials for Group IV Semiconductors

64

Page 65: Bond Order Potentials for Group IV Semiconductors

65

Table 3.5 SW Si Small Clusters.

Structure

Point

Group Bond

Bond

Length (Å)

Binding Energy

(eV)

SW HF

dimer dimer D∞h 1-2 2.352 -2.17 -3.06

trimers

linear chain D∞h 1-2 2.423 -3.81 ~

bent chain C2v 1-2 2.352 -4.34 -3.04

triangle D3h 1-2 2.563 -4.44 ~

tetramers

linear chain D∞h 1-2

2-3

2.413

2.512 -5.49 -7.29

square D4h 1-2 2.388 -8.15 -8.84

rhombus D2h ~ ~ ~ -10.71

tetrahedron Td 1-2 2.714 -6.68 -8.09

flagged triangle C2v

1-2

1-3

2-4

2.618

2.553

2.419

-6.11 -8.86

pentamers

pentagon D5h 1-2 2.352 -10.85 -10.08

trigonal bipyramid D3h

1-2

1-5

2-3

2.497

3.289

3.254

-10.32 -13.92

pyramid C4v 1-2

2-3

2.853

2.456 -10.74 -13.47

hexamers

edge capped

trigonal bipyramid C2v

1-2

1-3

2-3

3-4

3-5

2.435

2.563

3.333

3.273

2.435

-14.01 -18.26

face capped trigonal

bipyramid C2v

1-2

1-3

1-5

3-4

3-5

2.374

3.604

2.371

3.305

2.426

-14.15 ~

octahedron Oh 1-2 2.733 -12.86 ~

bent chair hexagon D3d 1-2

1-3 2.352 -13.02 -13.32

Page 66: Bond Order Potentials for Group IV Semiconductors

66

Table 3.6 Tersoff Si Small Clusters

Structure Point

Group Bond

Bond

Length (Å)

Binding Energy

(eV)

Tersoff HF

dimer dimer D∞h 1-2 2.3 -2.66 -3.06

trimers

linear chain D∞h 1-2 2.313 -5.25 ~

bent chain C2v 1-2 2.313 -5.25 -3.04

triangle D3h 1-2 2.313 -7.87 ~

tetramers

linear chain D∞h 1-2

2-3 2.319 -7.98 -7.29

square D4h 1-2 2.38 -8.64 -8.84

rhombus D2h ~ ~ ~ -10.71

tetrahedron Td 1-2 2.623 -7.09 -8.09

flagged triangle C2v

1-2

1-3

2-4

2.305

2.305

2.181

-6.65 -8.86

pentamers

pentagon D5h 1-2 2.323 -12.43 -10.08

trigonal bipyramid D3h

1-2

1-5

2-3

2.451

2.997

3.361

-10.95 -13.92

pyramid C4v 1-2

2-3

2.654

2.53 -10.34 -13.47

hexamers

edge capped

trigonal bipyramid C2v

1-2

1-3

2-3

3-4

3-5

2.389

3.002

2.53

2.88

2.483

-11.37 -18.26

face capped trigonal

bipyramid C2v

1-2

1-3

1-5

3-4

3-5

2.599

2.701

2.604

3.083

2.552

-13.23 ~

octahedron Oh 1-2

2-3

2.563

2.945 -12.38 ~

bent chair hexagon D3d 1-2

1-3

2.30

3.984 -15.79 -13.32

Page 67: Bond Order Potentials for Group IV Semiconductors

67

a reasonable level of potential predictive ability given that the silicon dimer (or any

other cluster property) was not considered during the parameterization of these

potentials.

Of the three different trimer configurations examined (the linear chain, the bent

chain and the equilateral triangle) ab initio calculations consistently determine that the

lowest energy configuration is the bent chain [104, 105]. The HF predictions have

determined the bond lengths of the bent chain to be 2.165 Å with an apex angle of 77.8°

[104, 105]. Neither the SW nor the Tersoff potential are able to stabilize a structure

with that apex angle, instead preferring to minimize into the equilateral triangle. Both

the SW and Tersoff potentials predict the equilateral triangle as the lowest energy

trimer structure. The SW potential predicts the bent chain structure to be only 0.1 eV

higher in energy than the equilateral triangle. The bent chain structure is very

competitive in energy because it is optimized at an angle of 109°. This angle

corresponds to the tetrahedral angle found in the diamond cubic lattice and is the

optimum angle for the SW three-body term [46]. The Tersoff potential predicts a clear

energetic advantage for the equilateral triangle. The linear chain and bent chain

structures are predicted by the Tersoff potential to have identical bond lengths and

energies; however, the apex bond angle of the bent chain structure is 132°.

We have examined five geometric arrangements as possible ground state

configurations of the silicon tetramer. These are: the linear chain, the square, the

rhombus, the tetrahedron and the flagged triangle. Ab initio studies indicate that the

Page 68: Bond Order Potentials for Group IV Semiconductors

68

lowest energy silicon tetramer can be found in the rhombus configuration [104, 105].

The SW and Tersoff potentials are unable to stabilize the rhombus structure. Instead

the higher symmetry square structure is preferred. The highest symmetry structure, the

tetrahedron, is not energetically competitive despite having a greater number of first

nearest neighbor bonds. This can be attributed to the large amount of bond bending

found in the structure.

Three pentamer geometries have been investigated here: the pentagon, the

trigonal bipyramid and the square pyramid. Two initial configurations for the trigonal

bipyramid and the square pyramid were considered. For the trigonal bipyramid the first

configurations had a large distance between atoms 1 and 5 (atom numbers are defined

in Figure 3.4) and a compact central triangle between atoms 2, 3 and 4. The alternative

configuration was distorted such that atoms 1 and 5 were closer together and the

central triangle was large. The second configuration has been predicted by ab initio

calculations to have a small energy advantage over the first (0.9 eV) and is the HF

predicted lowest energy pentamer configuration [104, 105]. The two square pyramid

configurations consist of a tight square with long bond lengths to atom 1 and a large

square with short bond lengths to atom 1. The SW potential predicts that all three

pentamer geometries are close in energy (only a 0.53 eV total energy range between

the structures) with the planar pentagon as the lowest in energy. The 108° angle found

in the planar pentagon is very close to the 109° optimal angle found in the SW potential,

therefore little bond bending is occurring. As a result, despite having fewer bonds than

the other configurations (5 compared to 8 in the square pyramid and 9 in the trigonal

Page 69: Bond Order Potentials for Group IV Semiconductors

69

bipyramid) the planar pentagon is predicted the lowest in energy. The Tersoff potential

predicts a clear energetic advantage for the planar pentagon structure for similar

reasons.

Four of the possible configurations for the silicon hexamer have been

investigated. These are: the edge-capped trigonal bipyramid (Ecap), the face-capped

trigonal bipyramid (Fcap), the octahedron (tetragonal bipyramid), and the chair-bent

hexagon. Ab initio studies have shown that the Ecap structure is the lowest energy

[104, 105]. The SW potential predicts the two capped trigonal bipyramid structures to

be very close in energy, with a small energy advantage for the Fcap structure (0.14 eV

lower in energy). The Tersoff potential predicts the chair-bent hexagon as the lowest

energy structure. This result is interesting in that it is the only cluster size for which a

diamond cubic lattice fragment is predicted by the Tersoff to be lowest in energy.

3.3 Point Defects

Point defects enhance diffusion rates in materials and are therefore important to

control during the synthesis of semiconductor devices [106]. For example, ion

implantation of dopants into a semiconductor substrate results in a super-saturation of

point defects [107]. After annealing at high temperature the diffusion rate for these

defects is anomalously high; a transient effect dependent on intrinsic carrier

concentration at the annealing temperature [108]. Some progress has been made in

examining the contributions of point defects to self-diffusion through the use of ab initio

Page 70: Bond Order Potentials for Group IV Semiconductors

70

calculations, such as local density approximation [109, 110], and, to some extent,

empirical descriptions for the energy [111, 112]. However the defect migration

pathways and diffusivities are still not well established. In order to address these issues

a potential that gives a reasonable approximation for defect formation energies is

required.

The defect formation energy, Ef, can be defined using the approach proposed

by Finnis [98] as

)0,(),(lim NEN

NxNEE d

Nf 3.4

where E(N,0) is the total energy of a perfect crystal with N occupied lattice sites, E(N,a)

is the total energy of a crystal with N lattice sites and x defects, and Nd is the number of

atoms in the defected crystal. Four point defects were considered: the vacancy (V), the

tetrahedral interstitial (T), the hexagonal interstitial (H) and the (110)-split interstitial

(X). Computational crystals were assembled with 1536 lattice sites (N) and the Finnis

method used to estimate defect energies. We note that this number of atoms was

sufficient to have converged to a good estimate of the energy.

The defect formation energies predicted by the SW and Tersoff potentials are of

the same order as those calculated by DFT methods [113-119]. The SW and Tersoff

predictions of defect formation energies are listed in Table 3.7. The lowest energy

interstitial configuration has been determined by DFT to be the (110)-split interstitial.

This interstitial, as well as the tetrahedral and hexagonal are illustrated in Figure 3.5.

The SW potential correctly predicts the X interstitial as the lowest energy and correctly

Page 71: Bond Order Potentials for Group IV Semiconductors

71

determines that the tetrahedral and

hexagonal are the second and third

in energy respectively. The SW,

however, fails in predicting the

magnitude its estimates. The SW

overestimates the formation energy of the X interstitial by 1.2 eV, the T by 0.8 eV and

the H by 2.0 eV. The SW potential is also unable to stabilize the H interstitial, instead

upon molecular statics minimization rearranges to a T interstitial (the transition

pathway from H-to-T is kinetically easier than the H-to-X, this will be discussed in section

V). The Tersoff potential, in contrast to the SW, is able to successfully stabilize the H

interstitial. The Tersoff potential, however, predicts the T interstitial as the lowest

energy configuration. Despite this, the Tersoff potential predictions are very close to

the DFT estimates.

The silicon vacancy has also been considered. This defect plays a key role in

diffusion mechanics [4]. Ab initio measurements indicate that the formation energy of

the silicon vacancy lies in the range of 3.3-4.3 eV [118]. These estimates have also

reported that vacancy formation is accompanied by a relaxation of the surrounding

lattice inwards towards the vacancy thereby decreasing the vacancy volume [119]. The

SW potential fails to capture this decrease in volume upon relaxation of the lattice. The

SW potential does not exhibit any lattice relaxation inwards or outwards; the atoms

surrounding the vacancy remain in their bulk lattice positions. As a result of this, the

defect formation energy of the silicon vacancy as predicted by the SW potential is

Table 3.7 Point Defect Formation Energies (eV)

Defect SW Tersoff DFT

V 4.34 3.73 3.3 - 4.3

X 4.46 4.45 3.3

T 4.98 3.52 3.7 – 4.8

H 6.57 4.67 4.3 – 5.0

Page 72: Bond Order Potentials for Group IV Semiconductors

72

identical to the cohesive energy (-4.34 eV). In contrast to ab initio estimates [119], the

Tersoff potential predicts a non-physical increase in volume surrounding the vacancy.

The magnitude of the volume increase estimated by the Tersoff potential is 24%.

3.4 Melting Temperature

Prediction of a materials melting temperature (Tm) is a good test of an

interatomic potential. Near the melting temperature, the interatomic spacing is

significantly larger than the separation at room temperature to which the potential was

fitted. The melting temperature therefore provides insight about the strength of the

interatomic bond and the shape of the interatomic potential at large interatomic

Figure 3.5 The diamond cubic lattice (a) and the three low energy

interstitial configurations: b) tetrahedral, c) hexagonal, and d) [110]-split.

Page 73: Bond Order Potentials for Group IV Semiconductors

73

separation (where the spline smoothing function may affect the interactions). Good

estimates of the melting temperature are often correlated with the accurate prediction

of surface structures, surface evaporation, and the rate of surface diffusion.

Experimental testing has shown that silicon undergoes a phase transition from the solid

to liquid at a temperature of 1691 K at ambient pressure.

The melting temperature was estimated for the potential following an approach

of Morris et al. in which a half-liquid/half-solid supercell is allowed to achieve an

equilibrium temperature under constant pressure [120]. A large supercell (2160 atoms,

60 plans of 36 atoms each) was used and two temperature control regions were applied,

one well above Tm and the other well below. The system was allowed to equilibrate for

20 picoseconds at which point the supercell was a half melted and half crystalline. The

temperature control regions were then removed and the system allowed to achieve

equilibrium (this was assumed to occur within 500 ps). If the resulting atomic system

retained a combination of crystalline and liquid material the uniform temperature of the

system was the melting temperature. The calculation was repeated several times and

predicted uncertainty was ±50 K.

The SW potential was parameterized with a strong focus of obtaining a model for

the simulation of liquid silicon properties [46]. As such, an accurate approximation of

the melting temperature was a weighty consideration. The SW potential predicts a

melting temperature 1750 K. The predictive uncertainty of ±50 K places this value very

close to the experimentally observed value of 1691 K [1]. The Tersoff potential does not

Page 74: Bond Order Potentials for Group IV Semiconductors

74

predict the melting temperature with accuracy. The Tersoff potential estimates the

melting temperature of silicon at 2800 K, significantly above the experimental value.

Little attempt is made in the original publications to justify this value accept to say that

given the short length of time of simulation, overheating may play a significant role [47-

49].

3.5 Surface Reconstructions

A robust description of surface morphology is an important part of the

simulation of vapor phase deposition. The most common surfaces that are used for

crystal growth in silicon are the (100) and (111) surfaces [2]. The widely observed

surface reconstructions for those surfaces are the (2×1) dimer row [121, 122] and the

(7×7) dimer adatom stacking fault (DAS) [123] respectively. At low temperatures the

(100) surface is often seen to have a c(2×4) buckled dimer configuration [124]. It is

thought that at higher temperatures the buckled dimers oscillate at a high frequency

such that they appear to be symmetrical [122, 125, 126]. Here we examine the (100)

(2×1) [121, 122], the (113) (3×2) [127-130], and the (111) (7×7) DAS [123]. These

surfaces are illustrated in Figures 3.6 to 3.8. The energies of surfaces were calculated

from the surface area, number of atoms, bulk cohesive energy, and total energy of the

computational supercell of the reconstructed surface. Each supercell had between

Page 75: Bond Order Potentials for Group IV Semiconductors

75

1000-2200 atoms with reconstructed top and bottom surfaces. The calculated surface

energies relative to the unreconstructed surfaces are compared with the ab initio/TB

data in Table 3.8.

The SW predicted values of the (100) 2x1 symmetric dimer rows is quite

accurate. The SW, however, fails to accurately capture the energy advantage of the

Page 76: Bond Order Potentials for Group IV Semiconductors

76

(111) 7x7 reconstructed surface, predicting an energy underestimation by a factor of 2.

The (113) 3x2 surface reconstruction is

overestimated by nearly a factor of 2 as

well. DFT calculations, however, have

not obtained a clear minimum energy

reconstruction for the (113) surface

Table 3.8 Surface Reconstruction Energies

(eV/Å2)

Surface SW Tersoff DFT

(100) 2x1 -0.052 -0.061 -0.054

(111) 7x7 -0.231 0.586 -0.403

(113) 3x2 -0.067 -0.075 -0.036

Page 77: Bond Order Potentials for Group IV Semiconductors

77

rendering an evaluation of a potential’s

predictions difficult [128-130]. The Tersoff potential also successfully captures the

simple dimer rows of the (100) 2x1 reconstruction. The Tersoff potential notably

predicts a positive surface energy for the (111) 7x7 reconstruction. This would imply

that the surface reconstruction is not stable. As a result, any simulation that employs

the Tersoff potential to examine this surface is called into question. The Tersoff

potential also overestimates the surface energy of the (113) 3x2 reconstruction;

however for reasons already stated this potential prediction cannot be adequately

critiqued.

Page 78: Bond Order Potentials for Group IV Semiconductors

78

IV. Silicon BOP Assessment

The previous chapter presented a detailed assessment of the Stillinger-Weber

[46] and Tersoff [47-49] silicon potentials. In this chapter, the recently published silicon

bond order potential [131] is presented and assessed in the same style as the SW and

Tersoff potentials in chapter III. Namely, the bulk properties for a range of experimental

and theoretical crystal structures, the energy and structure of small clusters, the melting

temperature, the defect formation energies, and surface reconstructions are all

examined. This potential has been developed because of a perceived weakness on the

part of other currently available interatomic potentials to explicitly account for the

physical concepts involved in covalent bonding in an analytic, non-empirical fashion.

The BOP format, which has been detailed in chapter II, incorporates additional physical

concepts such as -bonding and promotion energy. The inclusion of these bonding

concepts allows the BOP to capture additional features of silicon.

4.1 Bulk Properties

At standard temperature and pressure (273 K and 1 atm) the lowest energy

crystalline phase of Si has the diamond cubic structure [8]. The dc structure has an

Page 79: Bond Order Potentials for Group IV Semiconductors

79

atomic volume of ~20 Å3/atom and a cohesive energy of -4.63 eV/atom [8]. If sufficient

pressure (~12.5 GPa) is applied to Si, it will distort from the four-fold coordinated dc

structure to a six-fold coordinated Sn phase [79]. This transition is also accompanied

by an electronic shift from the semiconducting to metallic state [79]. When pressure is

removed from Sn phase silicon, it does not return to the dc phase. Instead it follows

a lower energy kinetic path to the bc8 phase at a pressure of ~8 GPa [79]. The structural

changes can be qualitatively observed in the DFT calculated binding energy curves for

silicon shown in Figure 4.1. The bulk energetics of the Si BOP have been explored by

producing binding energy vs. atomic volume curves for the structures; these are also

graphed in Figure 4.1. It is apparent that at a very high pressure the transition from dc

to Sn is properly modeled by the BOP. The BOP also correctly places the bc8 phase as

a stable phase at an intermediate pressure. Comparison between the DFT calculations

and the BOP predictions indicate that in general the BOP reproduces well the volume

dependent relative energies of different silicon phases.

The atomic volumes, cohesive energies, and bulk moduli of crystalline structures

predicted by the Si BOP are in reasonable agreement with DFT calculations. The atomic

volumes predicted by the BOP shown in Table 4.1 are within ±3% of the VASP DFT

estimates, with the exception of the bcc and bc8 structures (both within ±8%). The

cohesive energies summarized in Table 4.2, are within ±6% of DFT results. The bulk

moduli, summarized in Table 4.3, are within 50% of DFT results. This level of bulk

modulus prediction is within the acceptable predictive range limitations [60] found

Page 80: Bond Order Potentials for Group IV Semiconductors

80

within the tight binding model from which the BOP is derived and represents an

improvement over the SW and Tersoff potentials.

The elastic properties (c11, c12 and c44) are a noteworthy success of the Si BOP.

The BOP predicts c11, c12 and c44 to have values of 134.89 (∆19%), 70.98 (∆11%) and

84.03 (∆5%) GPa respectively. These predictions are very close to the experimentally

determined values of 165.78, 63.94 and 79.62 GPa respectively [8]. Of particular note is

the Cauchy pressure prediction. The Cauchy pressure (c12 – c44relaxed) was calculated to

Page 81: Bond Order Potentials for Group IV Semiconductors

81

be -13.05 GPa, in good agreement with the experimental value of -16.0 GPa [8]. The

realistic modeling of the Cauchy pressure makes the silicon BOP unusual amongst other

available interatomic potentials and indicates that the silicon BOP is uniquely capable of

reliably modeling the elastic behavior of silicon. The successful prediction of a negative

Cauchy pressure has been attributed to the inclusion of an environment dependent

repulsive promotion energy term [131].

Table 4.1 Atomic Volume Predictions (Å3). Comparable to Table 3.2.

Structure dc sc fcc bcc bc8 -Sn

DFT 19.79 0.789 0.704 0.717 0.869 0.749

BOP 20.01 0.804 0.696 0.647 0.924 0.719

% diff 1.1 3.1 0.04 8.7 7.6 2.8

Table 4.2 Cohesive Energy Predictions (eV). Comparable to Table 3.3.

Structure dc sc fcc bcc bc8 -Sn

DFT -4.63 0.292 0.468 0.451 0.122 0.213

BOP -4.639 0.434 0.536 0.539 0.134 0.224

% diff 0.2 3.1 1.3 1.9 0.07 0.04

Table 4.3 Bulk Modulus Predictions (GPa). Comparable to Table 3.4.

Structure dc sc fcc bcc bc8 -Sn

DFT 105.4 105.6 93.54 111.3 128.1 155.9

BOP 100.76 224.36 86.86 119.02 163.23 90.31

% diff 4.4 112.5 7.1 6.9 27.4 42.1

Page 82: Bond Order Potentials for Group IV Semiconductors

82

Examining the bulk phonon spectrum of the diamond cubic phase is also a useful

method of evaluating the performance of the potential near equilibrium. This

calculation is performed by annealing a sample crystal of sufficient size (512 atoms were

employed) and calculating the velocity-velocity autocorrelation function of the system.

The vibrational spectrum for the system can then be calculated by taking the Fourier

transform of the correlation function. The resulting vibrational spectrum is illustrated in

Figure 4.2. The highest peak calculated is at 550 cm-1 which is within ~5% of the

experimentally observed highest intensity peak at 520 cm-1 [103].

4.2 Small Clusters

Silicon dimer properties (energy and separation) were employed as target values

during the parameterization of the silicon BOP. It is therefore not possible to tout the

Page 83: Bond Order Potentials for Group IV Semiconductors

83

accuracy of their predictions as outputs of the potential. The experimental values for

the silicon dimer’s binding energy and separation distance are -2.41 eV and 2.34 Å [104,

105]. The BOP predictions are -2.61 eV and 2.336 Å. These data, as well as the BOP

predictions for all other cluster structures and energies are found in Table 4.4.

Three possible configurations for the Si trimer, the linear chain, the bent chain,

and the equilateral triangle, were examined. The BOP Si potential predicted that the

equilateral triangle had the lowest energy. Data for the linear chain and equilateral

triangle were not available for HF calculations [104, 105]. HF calculations predicted that

the bent chain has the minimum energy at an apex angle of ~80°. The apex angle that

minimizes the energy of the bent chain was found to ~111.6° using the BOP, matching

the minimum of the angular term.

Four Si tetramer structures were examined including the linear chain, the square,

the trigonal pyramid, and the corner-capped triangle. The BOP Si potential predicted

that the square had the lowest energy. The HF calculations indicated the square and the

capped triangle had the lowest energy (around -8.85 eV) [104, 105]. Both the bond

length and energy of the square predictions of the BOP agree well with the HF

calculations. However, there are some deviations for the other higher energy

structures. For example the Si4 linear chain structure has 2 separate bond lengths, as

shown in Figure 3.4. The BOP predicted that the 2-3 bond is significantly (~0.4 Å) longer

than the 1-2 bond, whereas the HF method predicted that the 2-3 bond is shorter (~0.1

Page 84: Bond Order Potentials for Group IV Semiconductors

84

Å) than the 1-2 bond [104, 105]. This may indicate an overestimation of the -bonding

nature in the 1-2 bond by the BOP.

The BOP predictions for the Si pentamer and hexamer structures are comparable

to the HF results [104, 105]. With BOP, the lowest energy pentamer is the planar

pentagon, however this was not the most stable structure with HF [104, 105]. The

Table 4.4 BOP Si Small Clusters

Structure

Point

Group Bond

Bond

Length (Å)

Binding Energy

(eV)

BOP HF

dimer dimer D∞h 1-2 2.336 -2.61 -3.06

trimers

linear chain D∞h 1-2 2.37 -4.84 ~

bent chain C2v 1-2 2.39 -4.64 -3.04

triangle D3h 1-2 2.47 -4.91 ~

tetramers

linear chain D∞h 1-2

2-3

2.433

2.815 -8.19 -7.29

square D4h 1-2 2.43 -8.87 -8.84

rhombus D2h ~ ~ ~ -10.71

tetrahedron Td 1-2 2.581 -7.26 -8.09

flagged triangle C2v

1-2

1-3

2-4

2.498

2.487

2.342

-7.18 -8.86

pentamers

pentagon D5h 1-2 2.384 -11.61 -10.08

trigonal bipyramid D3h

1-2

1-5

2-3

2.515

3.124

3.415

-11.38 -13.92

pyramid C4v 1-2

2-3

2.617

2.50 -11.29 -13.47

hexamers

edge capped

trigonal bipyramid C2v

1-2

1-3

2-3

3-4

3-5

2.466

2.506

3.544

3.062

2.466

-15.11 -18.26

octahedron Oh 1-2 2.554 -15.01 ~

Page 85: Bond Order Potentials for Group IV Semiconductors

85

planar pentagonal structure favored by the BOP does not provide a significant energetic

advantage over the competing high symmetry cluster structure’s predicted energies in

BOP predictions. The planar pentagon structure is preferred by BOP due to the minimal

amount of bond bending required to obtain the structure (bond angles of 108°). The

edge-capped trigonal bipyramid was found by the BOP to be the minimum energy

hexamer structure, in agreement with the HF predictions [104, 105].

In general the BOP Si potential prefers larger atomic spacing in clusters than the

HF calculations [104, 105]. The source for this discrepancy arises from the selection of

the target dimer separation in the BOP parameterization (2.336 Å). The HF

underestimates the dimer separation at 2.227 Å (~0.11 Å shorter than the experiment)

[8]. Even though the BOP does not predict the same low energy structures for Si

clusters as those predicted by the HF calculation (with the exception of the Si hexamer),

the improvement of the predictions of the overall binding energies for different clusters

compared to other potentials is anticipated to improve the simulations of complex

atomic arrangement processes.

4.3 Melting temperature

The predicted melting temperature (Tm) can affect surface structures,

evaporation rates, and surface diffusion. Near the melting temperature the interatomic

spacing is significantly larger than the separation at room temperature to which the

potential has been fitted. Therefore the melting temperature can be a combinational

Page 86: Bond Order Potentials for Group IV Semiconductors

86

measure of the strength of the interatomic bond (depth of the interatomic potential

well) and the shape of the interatomic potential at large separation. During deposition

large interatomic separation distances are frequently encountered. It is therefore

important to examine the melting temperature.

The Tm estimate was determined following the approach by Morris et al. in which

a half-liquid/half-solid supercell is allowed to achieve an equilibrium temperature under

constant pressure [120]. A large supercell (2160 atoms, 60 plans of 36 atoms each) was

used and two temperature control regions were applied, one well above Tm and one

well below. After 20 picoseconds, the supercell was ½ melted and ½ crystalline. The

temperature control regions were then removed and the system allowed to reach

thermal equilibrium (assumed to occur within 500 ps). At equilibrium the boundary

between the liquid and solid phases will have stopped moving, and the temperature of

this equilibrated region is taken as the melting temperature. The Tm obtained in this

manner has an uncertainty of ±50 K (obtained from variations in repeated runs).

The BOP silicon potential predicts a melting temperature of 1650 ±50 K. This

temperature range includes the experimentally observed value of Tm = 1687 K [1]. It

indicates that the BOP potential models well the interaction of silicon at large

separation distances.

4.4 Point defects

Precise defect formation energies were not expected because defects were not

incorporated in the fitting process. As a result, the relative order of the defect energies

Page 87: Bond Order Potentials for Group IV Semiconductors

87

predicted by the BOP and ab initio calculations is

not exactly the same. These results can be found

in Table 4.5. For example, the BOP predicts that

the T interstitial has the lowest energy, whereas

the DFT predicts that the X-split interstitial has the

lowest energy [113-119]. Nonetheless the BOP predicted interstitial energies are

generally close to previous DFT calculations. The vacancy formation energy of 2.759 eV

predicted by the BOP also well matches the past DFT value of 3.17 eV [118]. The

vacancy volume was seen to shrink during energy minimization. This again matches the

observations from ab initio calculations [119].

4.5 Surface reconstructions

A robust description of surface morphology is an important part of the

simulation of vapor phase deposition. The most common surfaces that are used for

crystal growth in silicon are the (100) and (111) surfaces. The widely observed surface

reconstructions for those surfaces are the 2×1 dimer row [121, 122] and the 7×7 dimer

adatom stacking fault (DAS) [123] respectively. At low temperatures the (100) surface is

often seen to have a c(2×4) buckled dimer configuration [124]. It is thought that at

higher temperatures the buckled dimers oscillate at a high frequency such that they

appear to be symmetrical [122, 125, 126]. The BOP potential was fitted approximately

to the (100) 2×1 surface energy. Here we examine the (100) 2×1 [121, 122], the (113)

Table 4.5 Point Defect Formation Energies (eV)

Defect BOP DFT

V 2.76 3.3 - 4.3

X 3.37 3.3

T 2.63 3.7 – 4.8

H 3.85 4.3 – 5.0

Page 88: Bond Order Potentials for Group IV Semiconductors

88

3×2 [128-130], and the (111) 7×7 DAS [123].

These surfaces have been illustrated in Figures

3.6-3.8. The energies of surfaces were calculated

from the surface area, number of atoms, bulk

cohesive energy, and total energy of the

computational supercell of the reconstructed surface. Each supercell had between

1000-2200 atoms with reconstructed top and bottom surfaces. The calculated surface

energies relative to the unreconstructed surfaces are compared with the ab initio/TB

data in Table 4.6.

It can be seen from Table 4.6 that the surface free energy relative to the

unreconstructed surface is within 0.06 eV/Å2 for each surface except the (113) 3×2

surface. However, this is not an issue because there is no clear minimum energy

reconstruction for the (113) surface in the literature [128-130]. Most importantly, the

highly complex (111) 7×7 surface reconstruction is found to be stable with nearly the

same relative free energy as tight binding calculation [132]. The BOP shows a marked

improvement in the calculation of surface energies over other available silicon

interatomic potentials [63].

Table 4.6 Surface Reconstruction Energies (eV/Å2)

Surface BOP DFT

(100) 2x1 -0.046 -0.054

(111) 7x7 -0.379 -0.403

(113) 3x2 -0.139 -0.036

Page 89: Bond Order Potentials for Group IV Semiconductors

89

V. Solid Phase Epitaxy of Silicon

5.1 Introduction

The synthesis of epitaxial silicon thin films by the low temperature condensation

of vapor [133], followed by the subsequent solid state transformation of these

amorphous layers [30] is widely utilized during the manufacture of microelectronic

devices. Amorphous layers can also be formed during ion implantation and are

recrystallized by thermal treatments [31]. The solid phase epitaxial growth (SPEG)

process that occurs during the annealing of these amorphous layers proceeds by the

thermally induced epitaxial growth of a crystal seed into the metastable amorphous

region [30]. The need to control point defect populations, dislocation types and

densities and stacking fault concentrations in films grown by these processes has

stimulated significant interest in the mechanisms of atomic reassembly at the epitaxial

interface.

Experimental thermal annealing studies indicate the growth rate of the

crystalline phase into an amorphous silicon system is well characterized by an Arrhenius

relation with a single activation energy of 2.7 eV which is thought to consist of a defect

formation energy of 2.4 eV and a defect migration energy of 0.4eV [30]. These

Page 90: Bond Order Potentials for Group IV Semiconductors

90

experimental studies also indicate that the transformations rates can be rapid. For

example, a 2500 Å thick a-Si film transforms fully to a crystalline structure in 2.5 seconds

at a temperature of 725 °C (a growth rate of 0.1 m/s) [30]. The SPEG rate can be

effected by self ion bombardment of the films [30]. The ion bombardment of

amorphous films results in a lowering of the activation energy for the SPEG process to

0.18-0.4 eV [32, 33]. Bernstein, Aziz, and Kaxiras argue that this low activation energy

reduction results from the ion impact assisted formation of the rate limiting defects. The

transformation from the amorphous to crystalline state of the ion irradiated structure is

then only controlled by the migration of these defects with an activation energy in the

0.4eV range [34].

A detailed understanding of the atomistic mechanisms involved in SPEG has

been impeded by the difficulties of high resolution imaging of the moving (sometimes

very rapidly) buried interface [35]. Computational modeling has therefore been used to

investigate the SPEG process [34-40]. The use of computational modeling techniques is

restricted by the relatively large number of atoms (of order 103) that must be used to

characterize each phase [34]. Additionally these systems must be simulated for an

extended time (>1ns) to observe even initial movement of the a-c interface [34]. The

use of high fidelity, unbiased parameter-free quantum mechanical calculations such as

density functional theory (DFT), is therefore prohibited [35]. A molecular dynamics

(MD) approach appears the most promising approach provided it employs an

interatomic potential that adequately captures the radial and angular dependences of

the interatomic interactions.

Page 91: Bond Order Potentials for Group IV Semiconductors

91

A computational study by Motooka et al. employing on the order of 1000 atoms

over a nanosecond timescale and utilizing the Tersoff potential [48] found two

temperature driven growth regimes in contrast to the single experimentally observed

temperature dependency [35]. Their MD results indicated that at low temperature

SPEG proceeded via a 2-D planar growth mechanism with an activation energy of 2.6 eV.

At higher temperatures however, {111} facets were formed at the interface and the

activation energy for growth decreased to 1.2 eV. Bernstein et al. also identified two

temperature driven growth regimes in a study employing an environmentally

dependent interatomic potential (EDIP) [34, 134]. However, they report an activation

energy of only 0.4 ± 0.2 eV at low temperature and an energy of 2.0 ± 0.5 eV in the high

temperature regime. These activation energies are clearly in conflict with the Tersoff

potential predictions. While both studies argue that removal of lattice defects at the a-c

interface is rate limiting, the defect whose migration controls the transformation rate

remained unclear.

The most recent computational studies of SPEG performed by Garter and Weber

[36-40] employ both the Tersoff and the Stillinger Weber potential [36-40, 46, 135].

They examined the morphology of the a-c interface and observed that the interface is

not sharp, but rather extends over a 6-8 atomic monolayer thick region. They argue

that the rearrangement of atomic defects in the transition region is the limiting atomic

mechanism in SPEG. They also show that the concentration of the defects predicted by

the Tersoff potential is about double that predicted by the Stillinger Weber approach

[135].

Page 92: Bond Order Potentials for Group IV Semiconductors

92

Here, we use the recently developed BOP for silicon and a molecular dynamics

simulation method to examine the initial stages of the amorphous to crystalline

transformation of silicon. Computational resource limitations constrain our simulations

using the BOP approach to systems of 1000 atoms and for short time scales (~1 ns). As a

result, the simulation is capable of only resolving the initial atomic reassembly

processes. We note at the outset that the initial growth rates observed below and in

previous studies of the simulations of the SPEG process in silicon are several orders of

magnitude faster than that observed experimentally [34-40]. We are confident that this

result does not indicate flaws in our work because previous studies [34-40] have

obtained similar results when examining the initial growth period. We suspect there

exists a SPE mechanism that is activated after the initial stage examined here is

complete. The discovery of this mechanism awaits much larger timescale MD

simulations.

The atomic scale structure of the amorphous film created using the BOP has

been examined, and is found to contain a high concentration of both 3 and 5

coordinated atoms. This indicates that the BOP predicts a highly defected amorphous

film similar to that observed in ion implanted films. These defects result in fast diffusion

pathways and a correlation is drawn between removal of interstitial defects at the a-c

interface and rapid epitaxial growth. The BOP based simulations are then employed to

investigate the activation energy barriers to the migration of interstitial defects within

the bulk. These energy barriers are observed to be similar to the calculated overall

activation energy for solid phase epitaxial growth process.

Page 93: Bond Order Potentials for Group IV Semiconductors

93

5.2 Simulation Details

There are numerous ways to synthesize amorphous silicon including quenching

from the melt [136], low temperature vapor deposition [137-139], and ion

bombardment [140, 141]. The a-Si films generated by each method have different

atomic scale structures [30]. Amorphous silicon generated by rapidly cooling the liquid

phase results in the formation of a network of tetrahedrally coordinated atoms with no

long range order (an “ideal” amorphous film) [136]. Low temperature vapor deposition

grown films are amorphous but also sometimes contain low density regions, or voids

[30]. These are a consequence of self-shadowing during the deposition process

combined with low atom mobility on the film surface [30]. When the deposited atoms

are unable to significantly migrate across the surface, the atoms assemble into a

random network with no long range periodicity. Ion implantation causes atoms to be

displaced from their lattice sites by primary and secondary collision processes [30].

Continuous ion bombardment results in the overlap of the damage zone of individual

impacts eventually leading to a continuous amorphous structure. Amorphous silicon

films created by ion bombardment contain a large concentration of three and five-fold

bonding defects [142].

To prepare a computational amorphous/crystalline (a-c) interface a 23.0375 Å

[10 1 ] by 43.44 Å [010] by 23.0375 Å [101] volume single crystal was created. This crystal

was made up of 1152 atoms distributed in 32 (010) layers. Periodic boundary conditions

were employed in the [10 1 ] and [101] directions. The top 24 monolayers were then

Page 94: Bond Order Potentials for Group IV Semiconductors

94

melted elevating their temperature to 2000 K for 50 ps (for 25000 time steps with a

time step, t = 2 fs), while the bottom 8 monolayers were thermally constrained to 500

K with the lowest 2 layers rigidly fixed in space. This resulted in a thin crystalline

substrate with a layer of liquid silicon on top. This system was then quenched to the

desired temperature over a period of 50 ps to create a computational sample containing

a region of crystalline and amorphous material separated by an amorphous to

crystalline interface. This system was then thermally annealed for 5 ns and the resulting

epitaxial growth rate and rate limiting defect migration energy barriers investigated.

The epitaxial growth rate was measured by tracking the rate of change of the

number of crystalline atoms present in the system by identifying the bonding

environment of an atom and that of its neighbors. In order for an atom to be classified

as part of a crystalline region it was required to maintain, within a tolerance of 10°, the

109o tetrahedral arrangement of its bonds with its four nearest neighbors. The increase

in number of atoms that satisfy this condition gives the epitaxial growth rate in atoms

per unit time interval.

The interstitials present in the crystallized region were also identified and their

diffusion pathways were also examined using molecular statics minimization techniques

[143]. Starting from a given interstitial configuration, the interstitial atom was

incrementally moved in the direction of minimum energy along a path to another

interstitial position. After each incremental movement, the entire system was allowed

to relax in energy around the constrained interstitial atom. This then allowed the

Page 95: Bond Order Potentials for Group IV Semiconductors

95

energy along the migration path to be computed and the energy barrier to interstitial

migration within the crystal to be determined.

5.3 Amorphous Characterization

The radial distribution function (RDF) for the a-Si film produced by the rapid

solidification simulation is shown in Figure 5.1. The simplest view of amorphous silicon

is that it consists of a continuous random network of tetrahedrally coordinated atoms

[30]. Therefore one would expect that the RDF would display a large slender peak

centered near the bulk silicon equilibrium nearest neighbor distance, followed by

broadening secondary and tertiary peaks. The RDF obtained by analyzing the quenched

Figure 5.1 Radial distribution function g(r) for a-Si compared to experimental a-Si

data. Inset: the atomic coordination distribution of silicon atoms in the a-Si region

graphed as a percentage of frequency as predicted by simulation.

Page 96: Bond Order Potentials for Group IV Semiconductors

96

structure generally agrees with this interpretation. The data obtained from the

simulation also agree well with experiment [144]. While this is an encouraging

affirmation of accurate modeling of a-Si by the BOP approach, it is not conclusive [145].

Many of the other interatomic potentials poorly predict the random tetrahedral

network of a-Si, predicting a large number of 3 and 5 coordinated atoms [145]. The

inset in Figure 5.1 shows the distribution of atomic coordinations in the amorphous

region predicted by the BOP analysis. The BOP predicts a broad distribution of atomic

coordination with a maximum at 4 and an average coordination of 4.16. Generally

amorphous silicon films are considered to be a uniform random network of coordination

4 atoms, however the various experimental methods for generating amorphous thin

films result in different amorphous states [30]. For example, the amorphous film

generated by ion implantation techniques contains many atoms that are not four-fold

coordinated [30]. Spectroscopic studies of self-implanted a-Si, have shown a large

concentration of dangling bonds associated with three-fold coordinated atoms and

floating bonds resulting from five-fold coordinated atoms [142]. In simulation, the many

atoms that do not have four-fold coordination suggest the amorphous region generated

by the fast quenching technique is highly defected. The high concentration of defects in

the very rapidly cooled BOP a-Si structure suggests that the simulated structure is

similar to ion beam irradiated structures where defects are introduced to the system by

a non-thermal ion collision process.

Page 97: Bond Order Potentials for Group IV Semiconductors

97

5.4 Epitaxial Crystallization

The quenched computational system was annealed at numerous temperatures

between 700 and 950K and the growth rate of the a-c interface during the SPEG process

was determined. This is plotted as a function of inverse absolute temperature in Figure

5.2. The growth rate data is reasonably well fitted by an Arrhenius relation with an

activation energy barrier of 0.87 eV. This activation energy is about twice that

experimentally reported for ion beam irradiated silicon [31]. Earlier simulation studies

have reported a wide range of activation energies [34-40]. Using their environmentally

dependent modeling approach, Bernstein et al [34] have shown that the low

temperature activation barrier (0.2 eV) observed in their simulations corresponds to the

migration of defects while the larger

value (2.0 eV) seen at higher

temperatures corresponds to a more

complicated process involving defect

formation and diffusion. The a-Si film

formed by quenching the Si-BOP structure

contained a large concentration of

“frozen in” defects and we suspect this is

responsible for our observation of an

activation barrier lying between the

Bernstein et al limits.

Page 98: Bond Order Potentials for Group IV Semiconductors

98

To investigate the atomic scale details of the a-c transformation, time resolved

atomic structures near the a-c interface during simulations of the SPEG process at

annealing temperatures of 700 and 900 K are shown in Figures 5.3 and 5.4. They show

the advancing amorphous to crystalline front moving through the amorphous region. It

can be seen that the most recently crystallized region in each time resolved consists of a

Page 99: Bond Order Potentials for Group IV Semiconductors

99

predominantly crystalline lattice containing a high concentration of interstitial defects.

Detailed examination of the transition region indicates that overlap of the lattice strains

of these interstitial defects eventually gives rise to the continuous random network of

the a-Si layer as one moves upward through the simulated region.

To investigate the influence of these remnant defects upon the advance of the a-

c interface we have plotted the number of crystallized atoms in the simulated structure

Page 100: Bond Order Potentials for Group IV Semiconductors

100

against transformation time for a transformation at 800K, Figure 5.5. It can be seen that

the growth of the crystalline phase is unsteady with sudden jumps in transformation

rate interspersed by periods of significantly slower transformation. The overall growth

rate of the crystalline phase is therefore limited by the atomistic rearrangements

occurring during these periods of low crystalline growth. An examination of the atomic

structure of the system before and just after a shift from a slow to fast growth mode

reveals that the velocity jump occurs upon elimination of an interstitial atom in the

crystalline phase just behind the transition region. This has been to occur at all of the

simulated temperatures. An example of such a rapid crystallization after the elimination

of a near interface defect can be clearly seen in the time resolved atomic structures of

Figures 5.4(c) to 5.4(d). This result is in qualitative agreement with the conclusions of Lu

et al. who argued that removal of defects residing at the a-c interface was the rate

controlling mechanism for SPEG [142].

Page 101: Bond Order Potentials for Group IV Semiconductors

101

To characterize the nature of the defects within the partially crystallized system,

we have calculated bond angle distributions functions, g(), for each region, Figure 5.6.

The angular distribution function for the transition region, Figure 5.6(b), shows a broad

peak centered on the tetrahedral bond angle (109o) with a shoulder extending towards

~75°. A small secondary peak is also evident at a bond angle of ~50°. We note that a

[110]-split type interstitial defect has a bond angle of ~50°, while a tetrahedral defect

has bonds with an angle of ~70°. The average coordination of these interstitials was

obtained by visually identifying 100 interstitial atoms in the various simulations and

determining their coordinations. This resulted in an average coordination of 4.05 for

Page 102: Bond Order Potentials for Group IV Semiconductors

102

the transition region interstitials. The large angular distribution concentrations at 50°

and 75° within the transition region indicate the presence of a high concentration of

defects with configurations that include those bonding angles, such as the [110]-split

and tetrahedral interstitial types. Previous calculations, presented in section 4, have

determined the formation energy of these point defects to be 3.37 eV and 2.63 eV

respectively.

The diffusion pathways associated with these two interstitials have been

investigated using a molecular statics approach. We have employed notational

shorthand for the point defects as follows: silicon vacancy (V), tetrahedral interstitial (T),

hexagonal interstitial (H), and the [110]-split interstitial (X). The specific defect

migration pathways that have been examined are presented in Figures 5.7-5.10. In no

particular order these are the X-to-C, T-to-C, V-to-C, T-to-X, and T-to-H. It is important

to remember that these are intended to be approximations of the motions encountered

at the a-c interface. These combinations were chosen because each involves a single

defect migrating to either another defect location or to a crystal lattice site. The energy

barriers to motion range from 0.49 – 1.88 eV.

Of the examined atomistic motions, migration of a vacancy into an adjacent

lattice site has the lowest activation energy. Figure 5.7 details atomistic motions that

occur as the vacancy is moved. The formation energy of the silicon vacancy has been

previously predicted by the BOP to be 2.76 eV. The Si-BOP predicts a volume decrease

Page 103: Bond Order Potentials for Group IV Semiconductors

103

of ~40% for the silicon vacancy as the adjacent atoms relax inwards. The migration

pathway for the silicon vacancy is simple; an adjacent atom switches places with the

vacancy. The energy barrier to this motion is 0.49 eV. In more detail, the adjacent

atom, labeled 1 in Figure 5.7, moves in a [111] direction towards the vacancy site. At

Page 104: Bond Order Potentials for Group IV Semiconductors

104

the midpoint of this motion, the third configuration in Figure 5.7(c), bonds are formed

with 3 new atoms. Atom 1 after this point has 6 atoms to which it is bonded. As the

atom moves closer to the vacancy site, the original three bonds are subjected to

significant bonding strains (and subsequently the atoms bonded to those atoms as well)

and shift inwards toward atom 1. The original three heavily strained bonds finally break

when atom 1 occupies the former vacancy site. The non-symmetric nature of the

energy curve in Figure 5.7(f) reflects this atomic mechanism. The observed drop in

energy when those bonds finally break corresponds to the relaxation of the stored strain

energy in the original bonds.

Page 105: Bond Order Potentials for Group IV Semiconductors

105

The second lowest energy migration pathway involves the switch between a

tetrahedral interstitial and a [110]-split interstitial. This calculation has been performed

in two ways. The first, the T-to-X has atoms T and C, as labeled in Figure 5.8, fixed in

space and moved incrementally to the minimum energy configuration of the X

interstitial. The second method, the X-to-C, has atom X1, as labeled in Figure 5.8, fixed

in space and incrementally moved into its associated lattice site; atom X2 is allowed to

move freely and as a result minimizes into a tetrahedral site. In both of these methods

all other atoms are allowed to reach their minimum energy configuration at each

increment. Both of these methods simulate the motion between a T and X interstitial

configuration. The energy barrier to this migration is 0.75 – 0.97 eV. Figure 5.8 shows

the high energy configuration for this atomic motion. Significant lattice distortions

occur in the neighboring atoms. Because of the very small energy barrier to migration in

the X-to-T direction (0.14 eV) it is unlikely that X type interstitials will have a long

lifespan in simulation.

The third defect migration of interest is the migration of a tetrahedral site to

another tetrahedral site shown in Figure 5.9. It should be noted that the midpoint

between any two tetrahedral sites is always occupied by a hexagonal site. Therefore, by

symmetry, the energetics of the migration can be fully considered by examining the T-

to-H pathway. The hexagonal interstitial is predicted by the Si-BOP to be metastable

with an energy of formation of 3.85 eV. Any small thermally induced distortion from

perfect symmetry will result in the hexagonal defect transitioning into a tetrahedral site.

Page 106: Bond Order Potentials for Group IV Semiconductors

106

As a result the energy barrier to migration from one tetrahedral site to another is

approximately the same as the difference between their formation energies, 1.10 eV.

The final and highest energy defect migration examined is for a tetrahedral atom

to directly displace a neighboring lattice site atom. This atomic motion is labeled T-to-C.

The tetrahedral interstitial atom is incrementally moved towards a neighboring lattice

atom site. This pushes the crystalline atom out of its lattice site towards a nearby low

energy tetrahedral site (not the same tetrahedral site that is already occupied). The

high energy configuration of this motion is shown in Figure 5.10. Considerable lattice

distortion is encountered in this motion due to the motion requiring the breaking and

reforming of 6 strong atomic bonds. This atomistic mechanism encounTersoff an energy

barrier of 1.88 eV, considerably higher than any of the other motions examined. Due to

Page 107: Bond Order Potentials for Group IV Semiconductors

107

the considerably higher energy involved, it is unlikely that this motion is encountered

with any frequency in simulation.

If we examine the possible energy pathways for one tetrahedral interstitial to

migrate to a nearby tetrahedral interstitial site there are a number of ways of going

about it. The tetrahedral interstitial atom could directly displace an adjacent crystal

atom, pushing that atom into a nearby tetrahedral site. This motion, the T-to-C

mechanism, has been shown to have a very large energy of activation, 1.88 eV. Another

path the tetrahedral interstitial could take would be to migrate through a hexagonal

interstitial site. This pathway has been found to have an activation barrier of 1.1 eV,

much preferable to the T-to-C mechanism but still higher than the T-to-X pathway. The

Page 108: Bond Order Potentials for Group IV Semiconductors

108

lowest energy pathway for tetrahedral interstitial migration is to first overcome a 0.75-

0.97 EV energy barrier to form a X interstitial. Secondly, the X interstitial then must

overcome a 0.14 eV barrier to transform back into a T interstitial in a new tetrahedral

site. Because this last motion has the lowest energy, it is the mechanism that is most

likely to occur.

5.5 Discussion

We have studied the solid phase epitaxial growth of silicon using the recently

developed bond order potential. The solid phase epitaxial process involves the

spontaneous, thermally activated rearrangement of atomic bonds at the

amorphous/crystal interface [30]. For an ion-implanted amorphous surface,

experimental studies have shown that this process results in the motion of a sharp a-c

interface towards the free surface [30]. The growth rate of the crystalline region (or the

velocity of the a-c interface) can be well modeled by an Arrhenius relation with

activation energy of 2.7 eV [30]. Ion bombardment introduces defects into the

amorphous film which enhance the growth rate and reduce the activation energy to

0.18-0.4 eV [34]. The rapid quenching from the liquid phase method used to obtain the

amorphous film used in the present simulations results in a highly defected amorphous

film. As a result the a-Si found in the present simulations is similar to an unrelaxed

amorphous film under ion bombardment. The BOP predicts activation energy of 0.87 eV

for the SPE process.

Page 109: Bond Order Potentials for Group IV Semiconductors

109

Previous atomistic modeling research has shown that defects play a key role in

SPEG [34-40]. The exact nature of that role has differed depending on which

interatomic potential was employed. The BOP also indicated that defects are an integral

piece of the atomistic mechanism that limits the SPE process. The rapidly advancing

crystalline front is slowed by the need to remove trapped interstitials at the interface.

These interstitials distort the surrounding crystal lattice and prevent further

crystallization until they are annealed out. These interstitials have been found to be

four-fold coordinated, indicating that they are predominantly of the (110)-split (X) and

tetrahedral (T) type. The bonding environment of the transition region in which these

interstitials are found is predominantly crystalline. We therefore believe that the

activation energy for the movement of these interstitials through a crystalline lattice will

be comparable to the activation energy of their movement in the transition region. The

X-to-T transition was found to have an energy barrier between 0.75 and 0.97 eV. This

energy is comparable to the 0.87 eV activation energy for SPE as predicted by the BOP.

This result suggests that the annihilation of interstitial defects at the a-c interface is the

rate limiting mechanism for solid phase epitaxial growth.

Page 110: Bond Order Potentials for Group IV Semiconductors

110

VI. SW and Tersoff Germanium

Assessment

Germanium is a covalently bonded material of great scientific and engineering

interest because of its unique combinations of electrical and other physical properties.

For instance, germanium is an indirect band gap material with a small band gap. It

therefore has low light absorbance at infrared wavelengths, and because it easily cut

and polished, it can be used for infrared lenses and windows in the 8-14 micron

wavelength range [8, 9]. The small bandgap of Ge and Ge-Si alloys also make it a useful

material in solar cell applications where its high absorbance enable the use of thinner

layers of active material [10, 11]. Light emitting diodes (LEDs) have also been fabricated

to take advantage of germanium’s unique properties [12, 13]. These LEDs are based on

Ge-Si alloy self-assembled quantum dots and have exhibited a broad emission peaked at

a wavelength centered upon 1.45 m [12, 13].

The electronic structure of germanium is very similar to silicon due to the fact

that both are group IV elements [1]. As mentioned in the introduction, the ground state

equilibrium structure of germanium is the diamond cubic phase [8]. The nearest

neighbor bond length in this phase is 2.446 Å which is approximately 4% larger than for

Page 111: Bond Order Potentials for Group IV Semiconductors

111

silicon [8]. Germanium also forms weaker bonds than silicon; the cohesive energy of the

diamond cubic phase of germanium is -3.82 eV compared to the -4.63 eV for silicon [8].

This arises because germanium has a weaker electronegativity than silicon, and

subsequently forms a shallower energy well [17]. These differences necessitate a

reparameterization of the existing potentials to accommodate germanium. In this

chapter we present the SW and Tersoff predictions for germanium properties using the

published parameter sets [49, 54]. In chapter VII, we shall present the predictions of the

recently developed Ge BOP.

6.1 Bulk Properties

Binding energy curves were constructed for germanium using the VASP DFT

software package [146-149] and the Stillinger-Weber [46, 54] and Tersoff [48-50]

potentials. These results can be seen in Figure 6.1. The SW binding energy curve

prediction for germanium shows a significant weakness in the potential to predict the

energetics of alternative bulk structures. The Ding and Andersen fit of Ge [54]

significantly overestimates the energy differences between the phases; the sole

exception being the bc8 phase. It is not surprising that the bc8 phase is modeled

reasonably given that the atoms in the bc8 structure possess a similar local bonding

environment to those found in the dc structure. The close packed phases (sc, fcc and

bcc) as well as -Sn are significantly more open (i.e. higher relative volume) than what is

expected from the DFT estimates; these structures are seen to approach a relative

Page 112: Bond Order Potentials for Group IV Semiconductors

112

Figure 6.1 Binding energy vs. relative volume (with respect to equilibrium dc

volume) for a range of Si bulk structures. a) Stillinger-Weber b) Tersoff c) DFT.

dc

-Sn

bc8

sc

bcc

fcc

hcp

Page 113: Bond Order Potentials for Group IV Semiconductors

113

volume of unity with respect to the diamond cubic phase. Experimental studies of

germanium phases observe an atomic rearrangement to the -Sn structure at high

pressure [150]. In contrast to this, the SW predicted binding energy curves suggest that

the dc phase remains the equilibrium structure even under high pressure.

The Tersoff binding energy curves indicate that the Tersoff potentials provides a

reasonably approximation of the energies of many of the other examined structures.

Two of the phases (fcc and bcc) have their energy minimums at significantly lower

relative volumes than expects however. This results in the aberrant prediction of the

bcc phase becoming the minimum energy phase at high pressure. Also absent from the

Tersoff predictions is a region of slightly elevated pressure for which the minimum

energy phase is bc8 (in DFT this occurs at ~0.81 V/V0). Another odd presentation of the

Tersoff binding energy curves is the curvature of the fcc phase. Normally it is difficult to

estimate a value for the bulk modulus visually, the fcc curve appears abnormally sharp

(calculations show that the predicted modulus for the fcc phase is 914.3 GPa quite in

excess of the DFT estimate of 63.5 GPa).

DFT calculations have predicted an atomic volume trend for germanium of (in

decreasing order): dc, bc8, sc, -Sn, fcc, bcc and hcp. This trend is similar to the DFT

estimates obtained for silicon. In those calculations the first four structures were in the

same order. The DFT, SW and Tersoff atomic volume predictions are presented in Table

6.1. Stillinger-Weber predictions for atomic volume are poor. The SW predicts the

structure with the greatest atomic volume to be the hcp structure. This result is

Page 114: Bond Order Potentials for Group IV Semiconductors

114

contradictory to the DFT results. Additionally, the second highest atomic volume

structure as predicted by DFT, the bc8 phase, is found by the SW to be ranked 5th in

terms of atomic volume. The Ge SW potential systematically overestimates the atomic

volume of the close packed structures. The Tersoff potential handles the atomic volume

predictions much better. The only failure observed by the Tersoff potential is the

overestimation of the hcp phase atomic volume. The Tersoff also switches the order of

the bcc and fcc phases, however, the two phases are very close in atomic volume and

this is a minor issue.

Cohesive energy trends have been calculated for germanium and this data is

found in Table 6.2. The DFT calculations have determined an energy trend of (in

decreasing order): dc, bc8, -Sn, sc, fcc, hcp and bcc. The SW predicts the dc and bc8

phase energies with an accuracy of 2%, however the remaining phases are poorly

predicted. The average percentage difference for the remaining phases (-Sn, sc, fcc,

Table 6.1 Atomic Volumes as estimated by DFT and by the SW and Tersoff

potentials. The atomic volume of the dc structure is given in cubic angstroms while

all other values are given relative to the dc volume. For example the fcc Tersoff

atomic volume can be obtained by multiplying the dc value by the number listed;

22.54 * 0.72 = 16.23.

Structure dc sc fcc bcc bc8 -Sn hcp

DFT 22.54 0.83 0.81 0.80 0.92 0.81 0.80

SW 22.54 0.99 0.97 0.92 0.93 0.92 1.01

% diff 0.0 18.96 19.56 15.44 0.84 13.16 26.07

Tersoff 22.54 0.83 0.72 0.73 0.93 0.78 0.82

% diff 0.0 0.46 10.68 9.08 0.84 4.06 3.15

Page 115: Bond Order Potentials for Group IV Semiconductors

115

bcc and hcp) is approximately 16%. It should be noted the SW does predict the order of

the four lowest energy phases properly despite the considerable energy difference. The

Tersoff potential predicts the energies of the germanium phases much more accurately

than the SW. The Tersoff potential does not predict the b-Sn phase to be the third

lowest in energy however. Rather unusually, the Tersoff predicts the sc, bcc, -Sn and

hcp phases to have nearly identical energies (all within ±0.015 eV of each other).

Similar to the SW and Tersoff predictions for silicon, the predictions of the bulk

moduli for the examined phases of germanium are poor. These predictions are

presented in Table 6.3. Both the SW and Tersoff are in reasonable agreement for the dc

phase bulk moduli (32% and 19% differences respectively), however the remaining

phases are not as well modeled. Particularly noteworthy are the extreme predictions of

the fcc bulk modulus which the SW and Tersoff overestimate by 856% and 1340%

respectively. These results strongly argue against the use of these potentials in

simulations where these structures might be present.

Table 6.2 Cohesive energy estimates by DFT and the SW and Tersoff potentials. The

dc values are listed in terms of eV and all other values are given relative to the dc

value. i.e. the DFT estimation of the sc cohesive energy can be found by taking the

dc value and adding the 0.24 to obtain -3.58 eV.

Structure dc sc fcc bcc bc8 -Sn hcp

DFT -3.82 0.24 0.34 0.35 0.13 0.24 0.34

SW -3.86 0.81 1.12 0.99 0.25 0.73 0.84

% diff 1.05 14.72 21.18 17.38 2.06 12.62 13.07

Tersoff -3.85 0.32 0.53 0.35 0.26 0.33 0.32

% diff 0.78 6.89 4.71 0.75 2.79 1.7 1.35

Page 116: Bond Order Potentials for Group IV Semiconductors

116

The three independent elastic constants (c11, c12 and c44) have also been

calculated for the dc structure. Experimental measurements of germanium have

determined that these values are 131.5, 49.5 and 68.4 GPa for c11, c12 and c44

respectively [8]. A related property, the Cauchy pressure (c12 – c44) for germanium is -

19 GPa. The SW potential predicts values of 90.3, 75.3 and 39.6 GPa for c11, c12 and

c44 respectively and a Cauchy pressure of 35.7 GPa. The Tersoff potential predicts

values of 138.4, 78.2 and 23.3 GPa for c11, c12 and c44 respectively and a Cauchy

pressure of 54.9 GPa. A negative Cauchy pressure is indicative of a material in which

high hardness and stiffness arise not from a high bulk modulus, but from a high shear

modulus [98]. The prediction of a positive Cauchy pressure indicates that the atomic

bonds are more likely to undergo bending rather than stretching [98]. This non-physical

prediction of a fundamental germanium property is distressing.

The final examination of the dc bulk properties is the phonon spectrum.

Germanium exhibits a single strong vibration spectrum peak at 390 ± 10 cm-1 [103]. The

Table 6.3 Bulk modulus estimates by DFT and the SW and Tersoff potentials. All

values are listed in GPa.

Structure dc sc fcc bcc bc8 -Sn hcp

DFT 60.72 67.38 63.49 63.59 64.45 68.69 70.45

SW 80.31 155.71 607.37 554.01 74.79 300.34 239.83

% diff 32.27 131.09 856.64 771.22 16.04 337.24 240.42

Tersoff 72.21 98.98 914.29 122.88 78.41 104.19 196.03

% diff 18.92 46.89 1340.05 93.24 21.66 51.68 178.25

Page 117: Bond Order Potentials for Group IV Semiconductors

117

SW and Tersoff predictions are illustrated in Figure 6.2. The SW potential predicts a

single strong peak at 315 ± 10 cm-1. The Tersoff potential predicts a single strong peak

at 300 ± 10 cm-1. The vibration spectrum is a measure of how fast the lattice is

vibrating, a lower primary peak indicates that the oscillation of atoms in their lattice

sites is occurring at an increased rate. This may adversely influence other potential

predictions such as thermodynamic properties (i.e. melting temperature).

Page 118: Bond Order Potentials for Group IV Semiconductors

118

6.2 Germanium Small Clusters

Extensive ab initio data was not readily available for germanium small clusters,

therefore the construction of a germanium equivalent of Figure 3.4 was not possible.

Despite this, similar configurations were employed as the initial atomic configurations in

the MS minimizations (albeit with bond lengths increased by 4%). Published data on

germanium small clusters reveals the minimum energy configuration and the binding

energy of those configurations; this data is listed in Tables 6.4 and 6.5 together with the

SW and Tersoff cluster predictions. The absence of ab initio data for germanium

clusters precludes the inclusion of such data in Tables 6.4 and 6.5.

Experimental studies have determined that the germanium dimer has a

separation distance of ~2.3 Å with a binding energy of ~2.64 eV [151]. Both the SW and

Tersoff potentials overestimate the dimer separation distance (2.45 Å and 2.43 Å for SW

and Tersoff respectively). These two potentials obtain a minimum energy separation

closer to the nearest neighbor distance of the diamond cubic structure (2.446 Å). Both

potentials also significantly underestimate the binding energy of the Ge dimer.

Ab initio calculations have determined that a bent 3 member chain is the lowest

energy structure for the Ge trimer [151, 152]. In contrast to their predictions for silicon,

both the SW and Tersoff germanium potentials are able to correctly predict this. The

SW potential predicts an energy preference for the bent chain structure over the

equilateral triangle of 0.45 eV. Not surprisingly, the apex angle of the ent chain is

predicted to be 109° by the SW potential. This is the result of the potential’s preference

Page 119: Bond Order Potentials for Group IV Semiconductors

119

Table 6.4 Stillinger-Weber germanium cluster predictions

Structure Point

Group Bond

Bond

Length (Å)

Binding

Energy (eV)

dimer dimer D∞h 1-2 2.448 -1.93

trimers

linear chain D∞h 1-2 2.557 -3.19

bent chain C2v 1-2 2.448 -3.86

triangle D3h 1-2 2.752 -3.41

tetramers

linear chain D∞h 1-2

2-3

2.529

2.716 -4.55

square D4h 1-2 2.515 -6.99

rhombus D2h ~

tetrahedron Td 1-2 2.924 -5.65

tetrahedral fragment C3v 1-2 2.462 -4.85

pentamers

pentagon D5h 1-2 2.448 -9.64

trigonal bipyramid D3h

1-2

1-5

2-3

2.656

3.461

3.490

-8.65

pyramid C4v 1-2

2-3

3.125

2.561 -8.11

hexamers

edge capped

trigonal bipyramid C2v

1-2

1-3

2-3

3-4

3-5

2.552

2.766

3.616

3.453

2.552

-11.34

face capped trigonal

bipyramid C2v

1-2

1-3

1-5

3-4

3-5

2.482

3.771

2.478

3.514

2.566

-12.12

square bipyramid C4v 1-2

2-3

2.846

3.162 -9.39

bent chair hexagon D3d 1-2

1-3

2.448

3.998 -11.58

Page 120: Bond Order Potentials for Group IV Semiconductors

120

Table 6.5 Tersoff germanium cluster predictions

Structure Point

Group Bond

Bond

Length (Å)

Binding

Energy (eV)

dimer dimer D∞h 1-2 2.432 -2.01

trimers

linear chain D∞h 1-2 2.482 -3.60

bent chain C2v 1-2 2.452 -4.01

triangle D3h 1-2 2.622 -3.79

tetramers

linear chain D∞h 1-2

2-3

2.429

2.761 -5.07

square D4h 1-2 2.497 -6.86

rhombus D2h ~ ~ ~

tetrahedron Td 1-2 2.733 -5.78

flagged triangle C2v

1-2

1-3

2-4

2.622

2.622

2.432

-5.48

pentamers

pentagon D5h 1-2 2.442 -9.79

trigonal bipyramid D3h

1-2

1-5

2-3

2.561

3.098

3.521

-8.86

pyramid C4v 1-2

2-3

2.758

2.638 -8.48

hexamers

edge capped

trigonal bipyramid C2v

1-2

1-3

2-3

3-4

3-5

2.652

2.708

2.813

3.098

2.559

-10.79

face capped trigonal

bipyramid C2v

1-2

1-3

1-5

3-4

3-5

2.694

2.801

2.706

3.308

2.657

-10.59

octahedron Oh 1-2

2-3 2.754 -10.98

bent chair hexagon D3d 1-2

1-3

2.431

4.125 -12.06

Page 121: Bond Order Potentials for Group IV Semiconductors

121

for the tetrahedral bonding angle. The Tersoff potential predicts an apex angle of 115°.

Both interatomic potentials underestimate the bonding energy of the trimer by slightly

less than 3 eV.

Similar to their predictions of the silicon tetramer, HF based calculations have

determined that the germanium tetramer has its lowest energy in a rhombus

configuration [151, 152]. Neither the SW nor Tersoff potential are able to stabilize a

rhombus structure. When molecular statics minimization is performed both the SW and

Tersoff potential’s predict a shift to a square configuration. The square configuration is

predicted by both the SW and Tersoff potentials to be the lowest in energy.

Interestingly, the SW potential was also unable to stabilize the flagged triangle

structure. Instead when a MS minimization was applied the resulting structure was a

tetrahedral fragment (the 1-3 bond was broken, and atom 2 was raised out of the plane

created by the other 3 atoms).

Ab initio data on the structure and energetics of germanium clusters of sizes 5

and 6 that are consistent with the values for clusters of size 2-4 are not readily available.

As a result, no direct comparison between the SW and Tersoff potentials and highly

accurate DFT calculations can be performed. The supposition that germanium clusters

of size 5 and 6 will behave similarly to silicon clusters of size 5 and 6 is not unreasonable,

however it must be recognized as an assumption. Both silicon and germanium display

similar characteristics for clusters of size 3 and 4 as well as having very similar electron

valence. Regardless of which high symmetry structure has the lowest binding energy, it

Page 122: Bond Order Potentials for Group IV Semiconductors

122

is fairly reasonable to assume that it isn’t the planar pentagon structure (which is the

least stable of the pentamers examined by silicon HF calculations). The SW and Tersoff

potentials both predict the planar pentagon as the lowest in energy.

The SW and Tersoff potentials predict differing structures as the lowest energy.

The SW potential prefers the face-capped trigonal bipyramid and the Tersoff potential

prefers the chair-bent hexagon. The SW potential predicts considerable distortions

from the initial f-cap structure, however. This includes an increased in the 1-3 bond (as

defined in figure 3.4) from 2.79 Å to 3.77 Å. Also the 3-4 bond increases from 2.97 to

3.51 Å. These lengths are approaching, but not exceeding, the SW cutoff distance for

germanium of 3.92 Å. As a result the final structure is similar to a five member ring with

a single atom replaced by a dimer pair. The chair-bent hexagon structure preferred by

the Tersoff potential is similar to a tetrahedral fragment. The 1-2 bond length is very

close to the nearest neighbor distance encountered in the bulk diamond cubic lattice.

6.3 Melting Temperature

Prediction of a materials melting temperature (Tm) is a good test of an

interatomic potential. Near the melting temperature, the interatomic spacing is

significantly larger than the separation at room temperature to which the potential was

fitted. The melting temperature therefore provides insight about the strength of the

interatomic bond and the shape of the interatomic potential at large interatomic

separation (where the spline smoothing function may affect the interactions). Good

Page 123: Bond Order Potentials for Group IV Semiconductors

123

estimates of the melting temperature are often correlated with the accurate prediction

of surface structures, surface evaporation, and the rate of surface diffusion. The

melting temperature was estimated for germanium in the same fashion used to

calculate the melting temperature of silicon.

The SW germanium potential predicts a melting temperature of 1350±50 K, a

temperature range above the experimentally observed Tm of 1211 K [1]. This result is a

significant variance from the experimental melting temperature, however it does reflect

the priorities assigned in the parameterization of the Ge SW potential. Ding and

Andersen [54] were unable to find a set of parameters for the SW potential that gave a

good description of all three phases of germanium: diamond crystal, amorphous solid,

and liquid. As their primary interest was in the development of a germanium potential

to model the amorphous structure, and because they recognized the importance of

properly capturing the structure of the diamond crystal, they chose to sacrifice liquid

property predictive accuracy [54]. The Tersoff germanium potential, similar to its silicon

melting temperature predictions, is a significant overestimation at 2400±50 K. It is

possible that the short cutoff distance of the Tersoff potential (3.1 Å) hampers its ability

to accurately model the interaction of germanium atoms at large separation distances,

such as encountered during melting simulation.

6.4 Point Defects

The formation energy of point defects has been examined for germanium in the

same fashion as for silicon. Table 6.6 lists the SW and Tersoff defect formation energy

Page 124: Bond Order Potentials for Group IV Semiconductors

124

predictions. Similar to its prediction

for the silicon vacancy, the SW

germanium potential also predicts

the vacancy formation energy to be

the same magnitude as the cohesive

energy. This is accompanied by the lack of any lattice relaxation around the vacancy.

The Tersoff potential significantly overestimates the formation energy of the vacancy,

placing it at 3.6 eV (nearly double the DFT estimation). The DFT calculations of the

germanium vacancy have shown a distinct relaxation of the surrounding lattice atoms

inwards towards the vacancy, resulting in a 25-40% volume decrease for the vacancy

[153]. The Tersoff potential predicts an increase in vacancy volume of 4.7%.

The SW germanium potential has very poor interstitial formation energy

predictions. The SW overestimates the formation energies by a factor of ~3. The source

of this prediction may be traced back to the change in the parameter. The germanium

and silicon SW parameterizations are identical with the exception of the energy and

length scale and the parameter (which was increased from 21 to 31 for germanium).

This parameter represents the reduced strength of the three-body interaction. The

increase in this term increases the scale of the energy penalty for deviations from the

tetrahedral angle. As a result, the decidedly non-tetrahedral angles found in the local

environment of the interstitials are found to be erroneously high in energy. The Tersoff

potential presents a reasonable set of predictions for interstitial formation energies,

Table 6.6 Point Defect Formation Energies (eV)

Defect SW Tersoff DFT

V 3.86 3.60 1.9

X 9.16 3.59 3.19, 3.84

T 7.52 4.44 2.29, 3.55

H 9.83 4.86 2.94, 3.99

Page 125: Bond Order Potentials for Group IV Semiconductors

125

although does display a preference for the T interstitial over the X interstitial in contrast

to some DFT predictions [153, 154].

6.5 Surface Reconstructions

The local bonding environment found at the surface differs significantly from the

bonding structure of the bulk lattice phases employed in the parameterization.

Therefore the examination of surface energies is a stringent test of a potential’s

transferability. We compute the surface energies for several of the low energy

reconstructions on low index germanium surfaces ((100) and (111)). The surfaces

examined are illustrated in Figures 3.6 and 6.3 (there are no structural differences

between the Si (100) 2x1 and the Ge (100) 2x1 surfaces). The energies of surfaces were

calculated from the surface area, number of atoms, bulk cohesive energy, and total

energy of the computational supercell of the reconstructed surface. Each supercell had

between 1000-2200 atoms with reconstructed top and bottom surfaces.

Ab initio simulations observe a buckling effect for the (100) 2x1 surface dimer

rows as the free energy of a (100) germanium surface is minimized [155]. The SW and

Tersoff germanium parameterizations reveal the buckling of surface dimers to be

energetically unstable. The room temperature stable surface reconstruction for the Ge

(111) surface is the c(2x8) reconstruction [156]. This configuration, in contrast to the Si

(111) 7x7 surface, does not require stacking faults or dimers, and is described by a

simple adatom model. Each c(2x8) surface unit contains 2 adatoms that stabilize the

Page 126: Bond Order Potentials for Group IV Semiconductors

126

structure by saturating the surface dangling bonds and through a charge transfer from

the adatoms to the surface [156]. The calculated surface energies relative to the

unreconstructed surfaces are compared with the ab initio/TB data in Table 6.7. Both the

SW and Tersoff potentials provide predictions for the (100) 2x1 surface that are

reasonable, however their

predictions of the (111) c(2x8)

reconstruction significantly

overestimate the surface energy.

Table 6.7 Surface Reconstruction Energies (eV/Å2)

Surface SW Tersoff DFT

(100) 2x1 -0.082 -0.072 -0.031

(111) c2x8 -0.231 -0.237 -0.012

Page 127: Bond Order Potentials for Group IV Semiconductors

127

VII. Germanium BOP Assessment

7.1 Bulk Properties

DFT calculations indicate the atomic volume increases as the structure of

germanium changes from: bcc, hcp, -Sn, fcc, sc, bc8, to dc. The BOP predicted atomic

volume trend is slightly different, increasing in the sequence: bcc, fcc, sc, -Sn, hcp, bc8,

and dc. While the atomic volume trend for the more closely packed phases differs

slightly in order, the atomic volumes for these structures are all within 7% of the DFT

predictions. These results can be found in Table 7.1.

The bond order potential proposed here correctly predicts the lowest energy

structure to be diamond cubic. The DFT calculations reveal an energy trend in

germanium crystal structures from lowest to highest energy of: dc, bc8, -Sn, sc, fcc,

bcc, and hcp. The BOP predicts cohesive energy trends from lowest to highest energy

of: dc, bc8, -Sn, sc, bcc, hcp and fcc. The BOP predictions for the cohesive energy are

listed in Table 7.2. The BOP notably captures the significant cohesive energy trend of

the experimentally observed, open structures (dc, bc8, and -Sn) with high fidelity

(within 1%). The cohesive energies of the remaining structures are within 3-10% of the

Page 128: Bond Order Potentials for Group IV Semiconductors

128

corresponding DFT values. The level of accuracy obtained for cohesive energies of the

experimentally observed phases of germanium justifies the sacrifice of accurate

predictions of the more open structure phases.

Bulk moduli data from the BOP and DFT-LDA calculations are compared in Table

7.3. The bulk moduli were calculated without internal relaxation of the selected

structures. This had a negligible impact for cubic structures, but can become more

significant for non-cubic phases [102]. The BOP predicted bulk moduli for the

equilibrium dc phase is found to be within 6%. This demonstrates that the BOP

Table 7.2 The cohesive energies, Ec (in eV), in decreasing order, of seven condensed phases of germanium as calculated by the BOP and DFT. The accurate prediction of the dc phase was a critical weight of the parameterization process so there is no variance between the DFT and BOP predictions.

Structure dc bc8 -Sn sc fcc hcp bcc

DFT -3.82 -3.69 -3.58 -3.57 -3.48 -3.48 -3.47

BOP -3.82 -3.73 -3.53 -3.45 -3.11 -3.13 -3.35

% diff 0.0 1.1 1.4 3.6 10.6 10.0 3.4

Table 7.1 The atomic volumes, V (in Å3), in decreasing order, of seven condensed phases of germanium as calculated by the BOP and DFT. The BOP predicted c/a values

for the -Sn and hcp phases are 0.46 and 1.79 respectively compared to the DFT predictions of 0.51 and 1.65 respectively. An accurate prediction of the volume of the dc phase was a large weight of the parameterization process so there is no variance between the DFT and BOP predictions.

Structure dc bc8 sc -Sn fcc bcc hcp

DFT 22.54 20.74 18.71 18.26 18.25 18.03 18.02

BOP 22.54 20.51 18.26 17.13 17.12 16.68 16.90

% diff 0.0 1.1 2.4 6.2 6.2 7.5 6.25

Page 129: Bond Order Potentials for Group IV Semiconductors

129

reasonably models the stretching and compression of the tetrahedrally coordinated dc

atomic bonds. The predicted bulk moduli for the range of structures examined is within

~10% for the sc and bc8 phases, within ~35% for the hcp phase and within ~50% for the

fcc, bcc and -Sn phases. The larger discrepancies are partially a result of the

parameterization process in which the atomic volume and cohesive energy were given a

higher weight during fitting. The discrepancies in bulk moduli observed for the higher

energy structures are similar to those encountered using other potentials [46-54]. A

visual comparison of the BOP and DFT predictions of the cohesive energy, atomic

volume (through relative volumes), and bulk moduli (through curvature) can be found in

the binding energy curves shown in Figure 7.1. The trends described in Tables 7.1-7.3

are clearly visible in Figure 7.1.

The three independent elastic constants (c11, c12, and c44) have also been

calculated for the dc structure. The predicted values are 60.81, 24.94, and 25.79 GPa for

c11, c12, and c44 respectively. These results underestimate the experimental values of

131.5, 49.5, and 68.4 GPa for c11, c12, and c44 respectively [8]. The experimental Cauchy

Table 7.3 The bulk moduli, B (in MPa), in decreasing order with dc listed first, of seven condensed phases of germanium as calculated by the BOP and DFT. A precise prediction of the dc bulk moduli was a low priority of the parameterization process; results within ~50% were considered acceptable.

Structure dc hcp -Sn sc bc8 bcc fcc

DFT 60.72 70.45 68.69 67.38 64.45 63.59 63.49

BOP 64.34 108.46 132.82 71.11 72.39 123.69 137.11

% diff 5.62 35.0 48.3 5.2 10.9 53.7 53.7

Page 130: Bond Order Potentials for Group IV Semiconductors

130

pressure (c12 – c44) for germanium is -19 GPa whereas the BOP predicts a value of -0.86

GPa. These BOP predictions are an output of the potential, i.e. the elastic constants

were not employed in the fitting process. Other interatomic potentials, such as the SW

and Tersoff shown in section 6, do not predict a negative Cauchy pressure. The

implementation of an explicit representation of the promotion energy enabled the

prediction of a negative value of the Cauchy pressure [60].

The phonon spectrum of the diamond cubic phase was also calculated. This

probes the performance of the potential at near equilibrium bond spacing conditions. A

diamond cubic crystal of 512 atoms was annealed at a simulated temperature of 300 K

and the velocities of a randomly selected 50 atom sample were tracked and used to

calculate the velocity-velocity autocorrelation function [157]. The vibrational spectrum

for the system was then calculated by taking the Fourier transform of this correlation

Page 131: Bond Order Potentials for Group IV Semiconductors

131

function. The resulting vibrational spectrum is shown in Figure 7.2. The primary

spectral peak was located at ~410 cm-1, which is reasonably close to the experimentally

reported value of 390±10 cm-1 [103].

In summary, the atomic volume and cohesive energy of germanium are well

reproduced by the BOP and compare favorably with the self consistent DFT data. The

bulk moduli predictions for the various structures examined by the BOP are comparable

in accuracy to those found using other available interatomic potentials [46-54]. The

BOP gives a negative Cauchy pressure in qualitative agreement with experiments and it

models the vibrational spectrum well.

Page 132: Bond Order Potentials for Group IV Semiconductors

132

7.2 Small Clusters

Germanium clusters can be made using laser photofragmentation methods

[158], and have been analyzed by mass spectrometry [159], ultraviolet photoelectron

spectrometry [160], and infrared spectrometry [161]. Germanium clusters have

structures that are quite different to those of the bulk material. The structure and bond

energies of small clusters are heavily reliant upon both the angular and radial

components of the interatomic potential. Examination of predicted cluster properties is

therefore a useful means of testing an interatomic potential. A conjugate gradient

method [162] was used to calculate the relaxed structure and binding energies for

numerous high symmetry configurations, and the results are listed in Table 7.4 together

with ab initio electronic structure calculations (where available) [151, 152]. We note

that the properties of germanium dimers were included in the fitting process and so

cannot be used to independently test the potential. The BOP predicted interatomic

spacing is fairly close to the expected atomic separation (2.394 Å versus ~2.3 Å

experiment). Their cohesive energy is slightly overestimated (-1.36 eV/atom versus -

1.32 eV/atom experiment) [151].

Three trimer configurations, the linear chain, the tetrahedral fragment, and the

equilateral triangle are summarized in Table 7.4. Electronic structure calculations have

Page 133: Bond Order Potentials for Group IV Semiconductors

133

Table 7.4 Germanium small cluster energetics and structure as predicted by the BOP.

Structure Bond Bond Length

(Å)

Cohesive Energy

(eV/atom)

BOP DFTa

dimer

1-2 2.394 -1.36 -1.32

trimers

1-2 2.43 -1.509 ~

1-2

1-3

2.441

4.287 -1.508 -2.24

1-2 2.568 -1.60 ~

tetramers

1-2

2-3

2.432

2.464 -1.620 ~

1-2

1-4

2.449

3.465 -2.054 ~

1-2 2.671 -1.794 ~

1-2 2.444 -1.587 ~

1-2

2-3

3-4

2.452

3.374

3.558

-2.055 -2.6

Page 134: Bond Order Potentials for Group IV Semiconductors

134

consistently found an isosceles triangle (i.e. a bent chain) to be the minimum energy

configuration with bond lengths (from the central atom) of 2.28 to 2.35 Å and a bond

angle of 81 to 86° [151, 152]. Using the BOP the linear chain and tetrahedral fragment

were observed to have nearly the same energy. The tetrahedral fragment distorted to

an angle of 122.78° from 109.7° with bond lengths of 2.441 Å from 2.446 Å with a

decrease in energy of ~0.005 eV. The equilateral triangle was found to be ~0.1 eV/atom

lower in energy than the two competing configurations, and was the minimum energy

configuration found using the BOP. The BOP was unable to minimize the structure into

the DFT predicted isosceles triangle, preferring instead the higher symmetry equilateral

triangle configuration.

Five high symmetry germanium tetramer configurations were also investigated.

They include the linear chain, the square, the tetrahedral fragment, the rhombus, and

the tetrahedron, Table 7.4. Electronic structure calculations have found the rhombus

configuration to be the lowest in energy with predicted bond lengths of 2.432 Å to

2.515 Å and 2.555 Å to 2.659 Å for the nearest neighbor (bond 1-2 in Fig. 3.4) and short

diagonal (bond 2-3 in Figure 3.4) distances respectively [151, 152]. The BOP successfully

predicts a rhombus configuration as the lowest energy configuration (although the

higher symmetry square is a close competitor). It is interesting to note that the

tetrahedral fragment is the highest in energy, illustrating the remarkable difference

between the bonding environments and preferences of small clusters and the bulk

lattice.

Page 135: Bond Order Potentials for Group IV Semiconductors

135

Overall the germanium BOP predictions compare well with ab initio calculations

performed previously by others [151, 152]. The ab initio data on germanium small

clusters available in literature was not comprehensive enough to provide a rigorous

comparison between the ab initio and BOP predictions for the range of structures

examined. However, with the exception of the germanium dimer, the cluster

predictions here are outputs of the potential, indicating that the BOP parameterization

is able to reasonably predict both bulk and cluster properties with a single parameter

set.

7.3 Melting Temperature

Prediction of a materials melting temperature (Tm) is a good test of an

interatomic potential. Near the melting temperature, the interatomic spacing is

significantly larger than the separation at room temperature to which the potential was

fitted. The melting temperature therefore provides insight about the strength of the

interatomic bond and the shape of the interatomic potential at large interatomic

separation (where the spline smoothing function may affect the interactions). Good

estimates of the melting temperature are often correlated with the accurate prediction

of surface structures, surface evaporation, and the rate of surface diffusion.

The melting temperature was estimated for the potential following an approach

of Morris et al. in which a half-liquid/half-solid supercell is allowed to achieve an

equilibrium temperature under constant pressure [120]. A large supercell (2160 atoms,

Page 136: Bond Order Potentials for Group IV Semiconductors

136

60 plans of 36 atoms each) was used, and two temperature control regions were

applied, one well above Tm and the other well below. The system was allowed to

equilibrate for 20 picoseconds at which point the supercell was a half melted and half

crystalline. The temperature control regions were then removed and the system

allowed to reach equilibrium (this was assumed to occur within 500 ps). If the resulting

atomic system retained a combination of crystalline and liquid material the uniform

temperature of the system was the melting temperature. The calculation was repeated

several times and predicted uncertainty was ±50 K.

The germanium potential predicts a melting temperature of 1115±50 K, a

temperature range below the experimentally observed Tm of 1211 K [163]. This

indicates that the BOP’s predictions of the bond strength at large separation distances is

slightly off, although not to a significant extent. The small discrepancy in melting

temperature may be related to the potential cutoff function, as believed to be the case

in other potentials [102, 164].

7.4 Point Defects

Point defects enhance diffusion rates in materials and are therefore important to

control during the synthesis of semiconductor devices [106]. For example, ion

implantation of dopants into a semiconductor substrate results in a super-saturatution

of point defects [107]. After annealing at high temperature the diffusion rate for these

defects is anomalously high; a transient effect dependent on intrinsic carrier

Page 137: Bond Order Potentials for Group IV Semiconductors

137

concentration at the annealing temperature [108]. Some progress has been made in

examining the contributions of point defects to self-diffusion through the use of ab initio

calculations, such as local density approximation [109, 110], and, to some extent,

empirical descriptions for the energy [111, 112]. However the defect migration

pathways and diffusivities are still not well established. In order to address these issues

a potential that gives a reasonable approximation for defect formation energies is

required.

The BOP predicted defect formations energies are compared to those estimated

from DFT calculations in Table 7.5. The formation energies predicted by the BOP fall

within the ranges calculated by the DFT methods (with the exception of the

tetrahedral). The lowest energy interstitial configuration has been found by DFT

calculations to be the (110)-split interstitial, followed by either the hexagonal or

tetrahedral interstitial as next

lowest in energy depending on

the exact method employed

[153, 154]. The BOP predicts the

lowest energy interstitial to be

tetrahedrally coordinated, with

the (110)-split configuration less

favorable by 0.31 eV. The

hexagonal interstitial is found to

Table 7.5 Point defect formation energy (in eV) and surface energy (in eV/ Å2) predictions by the BOP.

BOP DFT / ab initio

Defects Ef (eV)

V 1.97 1.9

IT 2.34 3.19-3.84

IX 2.65 2.29-3.55

IH 3.37 2.94-3.99

Surfaces (eV/ Å2)

(100) - 2x1 -0.051 -0.031

(111) – c2x8 -0.039 -0.012

Page 138: Bond Order Potentials for Group IV Semiconductors

138

be metastable; small distortions caused by thermal fluctuations were sufficient to

displace the interstitial atom from the hexagonal site. The DFT calculations of the

germanium vacancy have shown a distinct relaxation of the surrounding lattice atoms

inwards towards the vacancy, resulting in a 25-40% volume decrease for the vacancy

[153]. The BOP predicts the unrelaxed lattice with a vacancy to be metastable, however

with a small amount of thermal energy (1 ps anneal at 200 K) the system was able to

relax inwards. The BOP predicts the volume decrease of the vacancy to be ~50%, a

value that is comparable to the DFT calculations.

7.5 Surface Reconstructions

The local bonding environment found at the surface differs significantly from the

bonding structure of the bulk lattice phases employed in the parameterization.

Therefore the examination of surface energies is a stringent test of a potential’s

transferability. We compute the surface energies for the 2x1 surface dimer rows of the

(100) surface and the c2x8 reconstruction on the (111) surface. The surfaces examined

are illustrated in Figures 3.6 and 6.3. The energies of surfaces were calculated from the

surface area, number of atoms, bulk cohesive energy, and total energy of the

computational supercell of the reconstructed surface. Each supercell had between

1000-2200 atoms with reconstructed top and bottom surfaces.

Ab initio simulations observe a buckling effect for the (100) 2x1 surface dimer

rows as the free energy of a (100) germanium surface is minimized [155]. The Ge BOP,

Page 139: Bond Order Potentials for Group IV Semiconductors

139

like other interatomic potentials reveals the buckling of surface dimers to be

energetically unstable [73]. The room temperature stable surface reconstruction for the

Ge (111) surface is the c(2x8) reconstruction [156]. This configuration, in contrast to the

Si (111) 7x7 surface, does not require stacking faults or dimers, and is described by a

simple adatom model. Each c(2x8) surface unit contains 2 adatoms that stabilize the

structure by saturating the surface dangling bonds and through a charge transfer from

the adatoms to the surface [156]. The calculated surface energies relative to the

unreconstructed surfaces are compared with the ab initio/TB data in Table 7.5. The Ge

BOP does not explicitly account for charge transfer at a reconstructed surface and as a

result surface predictions are simplified, however the overall trends are favorable. For

increased fidelity (and increased computational cost) a charge transfer model can be

employed with the BOP. This model has been developed and can be found in reference

165.

Page 140: Bond Order Potentials for Group IV Semiconductors

140

VIII. Discussion

Like most avenues of scientific inquiry, the evolution of interatomic potentials

starts with an original basic concept, revolutionary for its time, which is developed in

incremental stages over time. The original interatomic potential employed a step

function form; if two atoms were within the cutoff distance they had a given potential

energy, if they were outside the cutoff they had no potential energy [157]. Over time,

as computational power improved, the models employed became more sophisticated.

While this dissertation makes no attempt to catalogue the full history of interatomic

potentials, it is interesting to note how far they have come. Two of the most commonly

employed interatomic potentials in use today for covalently bonded systems are the

Stillinger-Weber [46] and Tersoff [47-49] potentials. These potentials are popular

choices for silicon and germanium modeling because they are relatively simple to

implement, are accurate within certain constraints, and there exists a broad spectrum of

available literature on them [46-54]. We have examined these potential’s predictive

ability for a wide range of material properties and compared the results to experiment

and ab initio estimates. Furthermore, we have parameterized a new interatomic

potential which promises an increase in predictive validity through the use of an

Page 141: Bond Order Potentials for Group IV Semiconductors

141

analytically derived format; this potential is the Bond Order Potential, or BOP for short.

The BOP has been subjected to the same level of scrutiny as the SW and Tersoff

potentials. In this section we will give an overview of the performance of each

potential, clearly outline where the BOP predictions improve or fail to improve upon the

predictions of the SW and Tersoff, and discuss potential uses of the BOP for future

modeling.

8.1 Stillinger-Weber Overview

Our examination of a broad range of material properties reveals certain key

weaknesses in the Stillinger-Weber potential. While the potential was shown to be

proficient in the prediction of the bulk lattice properties of the diamond cubic phase,

the potential has significant difficulty in accurately modeling other lattice structures.

The inability to accurately model these non-equilibrium lattice structures is indicative of

poor potential transferability. Also troublesome for the SW potential are the energy

predictions of complicated surface reconstructions. This result leads us to believe that

the SW potential is not suitable to calculations involving free silicon or germanium

surfaces. Finally, the SW potential does not model point defects with high fidelity.

Lattice relaxation inwards at a vacancy site is observed by ab initio estimates; however

the SW potential predicts no relaxation of the lattice inwards or outwards at a vacancy

site. Additionally, the interstitial formation energies for germanium are extremely

Page 142: Bond Order Potentials for Group IV Semiconductors

142

inaccurate. These results indicate that the use of the SW potential in simulations of

defected systems will result in questionable outcomes.

The Stillinger-Weber potential does a number of things right however. It

predicts reasonable values for the elastic constants, despite not capturing a negative

Cauchy pressure, and provides a good estimate of small cluster energies. The structure

of small clusters is preferential to the tetrahedral angle, a feature not observed in ab

initio estimates. The SW potential also provides a reasonable prediction of the melting

temperature. The silicon (100) 2x1 surface reconstruction energy and structure are also

well described. Over all, the SW potential performs well in situations where the local

atom environment is tetrahedral. The SW potential does not seem to be appropriate for

use in simulations in which a significant number of non-tetrahedral atoms are present.

8.2 Tersoff Potential Overview

The Tersoff potential manages to improve upon the SW potential in a few key

areas: the energy and structure predictions of non-dc crystals of silicon and germanium,

and the defect formation energy of interstitials. Despite this, the Tersoff potential still

has a number of predictive difficulties. The first of these is the fidelity of the bulk

moduli of non-dc crystals, most notably the fcc structure. The Tersoff potential also

significantly overestimates the binding energy of the silicon trimer and predicts the

lowest energy trimer structure as the equilateral triangle. These two failures, however,

are not terribly debilitating to the potential’s usefulness. Far more troublesome for the

Page 143: Bond Order Potentials for Group IV Semiconductors

143

Tersoff potential is the prediction of a positive energy for the (111) 7x7 DAS surface

reconstruction. This would suggest that the Tersoff potential is unsuitable for

simulations that include a (111) free surface. Also problematic is the predicted melting

temperature being so high. The Tersoff potential overestimates the melting

temperature of both silicon and germanium by over one thousand Kelvin. This is a

serious issue because the accurate prediction of melting temperature indicates that the

potential reasonably models the interaction of atoms at large separation distances. It

has been suggested that temperature in simulation can be equated to real temperature

by scaling the simulated temperature to the real melting temperature outside the

simulation. In other words, if the temperature in simulation was 1500 K and the

simulated melting temperature was 2000 K while the real melting temperature 1800 K,

the real temperature you were simulating would be 1350 K. This solution is

unsatisfactory because it does not address the fact that each individual atom would

contain the equivalent energy of a particle at a much higher temperature than that

required to melt the material.

The Tersoff potential does provide an accurate prediction of the silicon and

germanium diamond cubic bulk properties. The great improvement in the predictions

of alternative bulk structures is one of the strengths of the potential. Small clusters are

neither a strength nor a weakness of the potential; however it is unlikely that the

Tersoff potential could be used to reliably predict cluster properties. Point defect

formation energies are reasonably well approximated, however it should be noted that

the Tersoff potential predicts that the lattice relaxation around a vacancy is outward,

Page 144: Bond Order Potentials for Group IV Semiconductors

144

not inward like ab initio estimates. The Tersoff potential does provide a good

description of the (100) 2x1 surface reconstruction for both silicon and germanium,

however more complicated reconstructions like the Ge (111) c2x8 are not well modeled.

Overall the Tersoff potential is adept at bulk predictions (except moduli) and the

prediction of interstitial formation energies. The failure of accurate prediction of high

temperature properties does limit its usefulness. This suggests that the Tersoff

potential is most suitable to simulations of atomistic mechanisms that occur at low

temperature within the bulk.

8.3 Improved Fidelity of the BOP Approach

The bond order potential (BOP) represents a fundamental shift in approach to

the formation of an interatomic potential. Prior potentials, such as the Stillinger-Weber

and Tersoff potentials presented here, were developed empirically. The bond order

approach differs by utilizing a functional form that has been derived from the highly

accurate tight binding methodology. This method has been discussed in detail in section

2, however it bears repeating here because of its great significance. It is well

understood by scientists and theoreticians that the predictions generated from

computer models do not necessarily reflect reality. It is of great importance to the

researching scientist who wishes to use computer modeling to be able to believe the

results he or she obtains from that computer modeling. The BOP represents an

appealing approach because it uses a format that is directly derived from the way that

Page 145: Bond Order Potentials for Group IV Semiconductors

145

atomic bonds form. Because of this, the BOP is inherently superior in terms of

predictive believability.

The BOP theory is, however, limited by the quality of the parameterization. A

parameterization of the BOP for silicon and germanium has been presented and

evaluated in the preceding sections. Table 8.1 provides a qualitative overview of each of

the three potentials’ predictions on a scale of 1 (very poor) to 5 (very good). The

method by which these numerical evaluations are assigned is discussed in appendix C.

As can be seen in this table, the BOP improves on the predictions of the SW and Tersoff

potentials in many categories. The BOP captures the bulk properties of the dc phase

very well; most notably the BOP is the only potential examined that predicts a negative

Cauchy pressure. The Ge parameterization of the BOP would have received a higher

rating for elastic constant predictions because of this but the potential is slightly soft,

i.e. the values for c11, c12, and c44 are low (although no worse than either the SW or

Tersoff). The BOP also remains highly accurate even for non-dc crystal structures, a

great improvement over the SW potential. The BOP mirrors ab initio estimations of

lattice relaxation inwards at a vacancy site, and reasonably captures the defect

formation energies of the examined interstitial configurations. The BOP accurately

predicts the melting temperature of both silicon and germanium; a considerably

improvement over the Tersoff potential. And lastly, the BOP represents a significant

step forward in predictive accuracy for complicated surface reconstructions as

evidenced by the (111) 7x7 DAS which is modeled very well by the BOP.

Page 146: Bond Order Potentials for Group IV Semiconductors

146

Table 8.1 Qualitative evaluation of the three interatomic potentials predictions for all

the properties examined in this dissertation. The properties are graded on a scale of

1-5 where 5 is very good and 1 is very poor. Appendix C details the value ranges used

to determine these qualitative assessments.

SW Tersoff BOP

Property Si Ge Si Ge Si Ge

Bulk Properties

• dc

Va 4 5 4 5 4 5

Ec 3 4 5 4 5 5

B 5 3 4 4 5 5

c11, c12, c44 3 3 3 3 4 3

phonon 4 3 3 3 4 4

• other crystals

Va 2 2 3 3 3 3

Ec 2 1 3 3 4 3

B 1 1 1 1 3 3

Small Clusters

Dimer 3 4 4 4 4 5

Trimers 3 3 1 4 3 3

Tetramers 3 3 3 3 3 4

Pentamers 3 3 3 3 4 4

Hexamers 3 3 3 3 4 4

Point Defects

Vacancy 2 2 2 2 4 5

Interstitials 3 1 4 3 4 4

Melting Temp. 4 3 1 1 5 5

Page 147: Bond Order Potentials for Group IV Semiconductors

147

Table 8.1 Continued.

SW Tersoff BOP

Property Si Ge Si Ge Si Ge

Surface Recon.

• Si surfaces

(100) 2x1 5 5 5

(113) 3x2 3 3 3

(111) 7x7 2 1 5

• Ge surfaces

(100) 2x1 3 3 4

(111) c2x8 2 2 4

These improvements do not come without a cost however. One way in which

the BOP is inferior to both the SW and Tersoff potentials is in computation time.

Simulations using the BOP consume more computational resources due to the increased

complexity of the functional format. As a result, the length of time for any simulation

using the BOP will be longer than an equivalent simulation using either the SW or

Tersoff potentials. This is not to say that the amount of time for BOP computations is

prohibitive, by no means is that the case. Molecular dynamics simulations employing

the BOP are still considerably faster than electronic structure calculations (such as DFT).

A rough estimate of the factor by which the BOP is slower than SW or Tersoff is 2-4.

While it is true that the BOP does not improve upon earlier interatomic

potentials in every way (certain features are well captured by the SW and Tersoff), the

Page 148: Bond Order Potentials for Group IV Semiconductors

148

BOP does successfully capture a much broader scope of material properties. The BOP

parameterization for silicon and germanium presented here is a much more transferable

potential than has been previously seen. As a result the applications to which it can be

put to use are many and varied. Within this dissertation is an example of one such use;

the examination if the solid phase epitaxy of silicon. This problem represents many

distinct challenges to a potential. The simulation of solid phase epitaxy involves the

modeling of complex atomic arrangements in the amorphous film and at the

amorphous/crystalline interface. The rearrangement of atoms at the a/c interface

involves the forming and breaking of atomic bonds, and so an interatomic potential

must accurately model the interaction of atoms at both long and short distances. Solid

phase epitaxy is a thermally driven process; therefore the interatomic potential must

respond and behave appropriately at a given temperature. Because the BOP is

successfully able to model these material properties it is well suited to the study of SPE.

In section 5 we presented a study of solid phase epitaxy using the BOP. We, like

others before us, observed an accelerated growth rate of the crystalline region. Where

the experimentally observed growth rate of the crystalline region can be measured in

m/s, the simulated growth rate is found to be orders of magnitude faster. Also

observed was a different activation energy for growth than that observed

experimentally; the value predicted by the BOP was closer to the activation energy of a

process related to SPE, ion beam induced epitaxial crystallization (IBIEC). Because of our

confidence in the validity of the BOP’s predictive ability, we concluded that what we

observed in simulation was a nanoscale atomistic mechanism that tells only a single

Page 149: Bond Order Potentials for Group IV Semiconductors

149

aspect of the larger story that is the SPE process. There must exist an atomistic

rearrangement mechanism that occurs on a longer timescale that limits the rate of

movement of the a/c interface. Sadly the investigation of an atomic system of such a

scale is beyond the computing resources at our disposal. This may well be resolved in

the future by continuing advances in microchip design, and the implementation of a

parallel processing version of the molecular dynamics code for the BOP.

Other avenues of investigation are open to the BOP. The investigation of rapid

crack propagation in silicon was proposed by Clint Geller at Bettis. This would require

the simulation of ~50,000 atoms, a number that is easily within reach using modern

computation techniques. Given the promising performance of the BOP in the prediction

of surface reconstructions, a wide range of surface simulations could be performed.

Also of great interest would be extending the BOP material database to include

semiconductor dopant materials such as phosphorous, boron, or arsenic. This would

require a separate parameterization for each material and the parameterization of the

binary terms. In a similar vein, it would be of great interest to parameterize the silicon-

germanium binary. The parameterization of the BOP is by no means a simple matter. It

needs to be approached in a careful manner, and often involves making choices which

material predications are more critical. In section 9 we present an examination in detail

of the parameterization process, and discuss how one would go about constructed the

Si-Ge binary potential (parallels can easily be drawn to how to develop a Si-P binary or

any other).

Page 150: Bond Order Potentials for Group IV Semiconductors

150

IX. Future Work: SiGe Alloy Parameterization

A critical feature of the developed BOP parameter sets for Si and Ge is that both

use an identical form of the bond order potential. This allows for the two material

systems to be employed together directly with the addition of a parameter set for SiGe

interactions. The development of a SiGe alloy BOP is the logical extension of the

research presented here. A BOP for SiGe alloys would allow for the simulations of many

modern complex SiGe multilayers semiconductor structures. This section will briefly

introduce the complexities involved in the parameterization of the SiGe alloy material

system.

The parameterization of a binary system presents a number of significant

challenges that are not present when fitting an elemental system. One such challenge

for the SiGe alloy system is that the Si0.5Ge0.5 alloy does not form a zinc-blende structure

like one would initially suppose (there have been reports of a stress induced ordered

Si0.5Ge0.5 lattice grown, however the ordered microstructure was not predominant in

the film and was difficult to reproduce [133]). Instead, the arrangement of silicon and

Page 151: Bond Order Potentials for Group IV Semiconductors

151

germanium atoms within the diamond cubic structure is random. Put another way,

each dc lattice site has a chance of containing either a silicon or germanium atom with a

probability directly related to their alloy concentrations. This is a direct result of the

total solid solubility of SiGe alloys. Figure 9.1 presents the phase diagram of silicon

germanium alloys. The lack of a regular crystalline structure presents a challenge for the

parameterization process.

Because fitting requires the parameters to be adjusted such that certain atomic

configurations have specific energies, one would need to provide data for specific

atomic configurations. The dearth of SiGe crystal structures forces the use of a fitting

database composed entirely of theoretical SiGe crystal configurations. Crystal

structures that likely should be employed in the fitting of the SiGe binary are: the zinc-

blende structure (ZB), the NaCl structure, and the SiGe dimer. Also of interest is the

ordered SiGe structure observed in limited deposition experiments. Energy and

Figure 9.1 Phase diagram for the silicon-germanium binary system.

Page 152: Bond Order Potentials for Group IV Semiconductors

152

structure information for these configurations should first be obtained from DFT

calculations such as those performed using VASP.

Because both the silicon and germanium BOP parameterizations employ the

same formats they can be integrated directly. As seen in Figure 9.1, silicon and

germanium exhibit total solid solubility. It is therefore reasonable to assume that the

parameters for silicon-germanium interactions can be derived from the two elemental

parameter sets. Precedence exists for this supposition; the Tersoff potential uses either

the average or geometric average of the silicon and germanium parameters for the alloy

parameters [47-49]. A similar approach may be possible with the BOP. The average of

the silicon and germanium BOP parameters should be a very reasonable first guess in

the parameterization of the silicon-germanium binary system. Appendix A discusses in

greater detail the Mathematica program employed in the fitting of the elemental

systems presented here, and would be an aid for anyone first approaching the complex

task of parameter fitting.

Page 153: Bond Order Potentials for Group IV Semiconductors

153

X. Conclusion

The modeling of large scale time dependent reassembly phenomena using highly

accurate density functional methods is computationally prohibitive. Because of this,

approximations of the covalent bonding of group IV semiconductors are necessary. To

model the complex reassembly mechanisms often seen in these materials, a molecular

dynamics approach has been employed. The predictive output of this method is highly

dependent on the quality of the interatomic potential employed. Many empirical

interatomic potentials have been proposed over the years; most notable are the

Stillinger-Weber and Tersoff potentials. The silicon and germanium parameterizations

of these potentials have been evaluated for their predictive ability for small clusters,

melting temperature, bulk properties for a wide range of crystal structures (dc, sc, fcc,

bcc, -Sn, hcp, and bc8), and the energy of low index surface reconstructions. These

potentials are shown to give reliable estimates of the bulk properties of the dc phase,

but are inadequate for the study of many other structures encountered during the

atomic assembly of both silicon and germanium.

Page 154: Bond Order Potentials for Group IV Semiconductors

154

Pettifor et al. have developed a bond order potential (BOP), derived from the

tight binding description of covalent bonding, to model atomistic systems. This

potential format is used here to propose a BOP for the group IV semiconductors silicon

and germanium. The potential addresses both the and bonding of these sp-valent

elements. A promotion energy term associated with the formation of the hybrid

orbitals is included in the formulism. The potential was parameterized using a self-

consistent database of DFT estimates and experimental data. The BOP potential’s

predictions for the cohesive energy, atomic volume, and bulk modulus of the dc, fcc,

bcc, bc8, hcp, and -Sn phases of silicon and germanium compare favorably with DFT

estimates. The BOP also gives point defect formation energies that are in good

agreement with ab initio estimates. The structure of small atomic clusters, the melting

transition temperature and the atomic structure of low index surfaces are also used to

assess the validity of the BOP potential. The functional form of the BOP’s for silicon and

germanium are the same, facilitating their eventual use for the study of the binary Si-Ge

alloy system.

In conclusion, this dissertation presents new interatomic potentials for the highly

important group IV semiconductors silicon and germanium. These potentials are

rigorously evaluated to ensure they accurately model the energy of atomic bonding in a

wide array of bonding environments. While many of the structures that were examined

are not observed experimentally, it is possible that during atomic reassembly processes

an atom may encounter a bonding environment similar to those non-experimental

phases. It is therefore a significant success to obtain interatomic potentials that are

Page 155: Bond Order Potentials for Group IV Semiconductors

155

flexible enough to simultaneously accurately model multiple phases. The BOP has been

further employed to examine the re-growth of a-Si films and identified the rate limiting

atomic scale mechanism.

Page 156: Bond Order Potentials for Group IV Semiconductors

156

Appendix A. Mathematica Fitting

In section 2.3 a brief discussion of the parameter fitting process was presented.

In this section we will examine the Mathematica codes employed in the fitting from the

perspective of fitting a SiGe binary. These Mathematica codes were designed and

written by Dewey Murdick during his graduate work on GaAs [102]. All modifications

pertaining to silicon and germanium have been made by me. The fitting code has been

split up into two primary programs: fitGSP and fitBO. These two programs call or load

additional files which are mandatory for their proper function, however only the CG-

NNL file will need to be altered for the fitting of the SiGe binary. Since modification of

the CG-NNL file is the first thing that needs to be done (it provides inputs for fitGSP and

fitBO) it will be discussed first, followed by discussions on fitGSP, fitBO, and adjusting for

promotion energy.

CG-NNL

Both the fitGSP and fitBO programs employ preconstructed crystal lattice data

(nearest neighbor lists, NNL) for each fitting structure to ease the computational cost of

the fitting algorithm. This crystal lattice data is constructed in the CG-NNL program.

Page 157: Bond Order Potentials for Group IV Semiconductors

157

This program serves two primary functions: first it generates the NNL’s that are

employed by fitGSP and fitBO, and second it outputs crystal files compatible with the

Fortran MD codes used in simulation. The uses of the second function are obvious and

will not be discussed further here. Each crystal NNL that you wish to construct has its

own subfolder in the CG-NNL program. A sample for the zinc-blende structure is shown

in Figure A.1. Similar subfolders are needed for each structure desired.

The only adjustments for the SiGe alloy system that need to be made are

changing the latparam entry and Btype information for the ZB and NaCl structures.

Additional structures can be created at this point such as the ordered SiGe alloy. The

entry for ordered SiGe can use the ZB subfolder as a template and change the atom type

positions in Btype. Structural data for the ZB and NaCl structures should be obtained

Figure A.1 Excerpt from the Mathematica program CG-NNL.nb that creates the

nearest neighbor list for the zinc blende structure of GaAs. The individual parts of

the code are labeled; additional information in the text.

cF8 / B3 (Zinc Blende); GaAs

latparam = {a -> 5.653}; Lattice Parameters

latlist = {a};

A = {{a, 0, 0}, {0, a, 0}, {0, 0, a}}; Primitive Vectors

B = {{0, 0, 0}, {1/4, 1/4, 1/4}, {0, 1/2, 1/2}, {1/4, 3/4, 3/4}, {1/2, 0, 1/2),

{3/4, 1/4, 3/4}, {1/2, 1/2, 0}, {3/4, 3/4, 3/4}}; Basis Vectors

Btype = {Ga, As, Ga, As, Ga, As, Ga, As}/. {Ga -> mat1, As -> mat2};

Nucsys = {3, 3, 3}; Crystal Size

CreateAll{“B3”, A, B, Btype, Nucsys, latparam} NNL Creation

CG-NNL.nb Mathematica Excerpt

Page 158: Bond Order Potentials for Group IV Semiconductors

158

from DFT calculations. After all necessary adjustments have been made to the crystal

structures of importance it is time to create a new input file for fitGSP and fitBO. This is

done by executing the “Write Orthogonal Form” subfolder (more specifically the

“wlist{…}”, “DeleteFile[…]”, and “Save[…]” lines).

fitGSP

As mentioned earlier in section 2.3, the BOP has been fitted using a three step

iterative process. The first fitting step is the parameterization of the GSP function. This

step obtains values for the 6 GSP parameters (r0, rc, m, n, n, nc) as well as a value for

the repulsive energy prefactor . It should be noted that the fitGSP program generalizes

the BOP format and assigns n and n the same value (this restriction can be relaxed in

the third fitting stage). The fitGSP program uses a discovery made by Albe et al. for the

Tersoff potential that the bond energy of cubic structures and dimers can be expressed

only as a function of bond length. The BOP has a similar feature if the promotion energy

term is neglected. For structures that have only the first nearest-neighbor shell within

the cutoff distance the bond orders and have a single value which remains

constant under hydrostatic strain. The fitGSP program derives parameters for the pair

terms and outputs bond order (BO) targets for the various structures considered to

fitBO. These BO targets reflect the environmental considerations of the BOP (i.e. the

angular dependencies).

Page 159: Bond Order Potentials for Group IV Semiconductors

159

A limitation on structures that can be employed in fitGSP is introduced by the

parameterization theory outlined above. Only those structures that can be fully

described by first nearest neighbors and where the angular environment of each atom

in the crystal is the same can be used. This limits the SiGe alloy structures that can be

fitted to the dimer, ZB, and NaCl structures. The theoretical ordered SiGe structure will

most likely not qualify because of the lattice anisotropy along (111) planes caused by

aligned Ge-Ge bonds. This limitation will not be detrimental in the long term.

The fitGSP program has a simple presentation. The “Setup & Functions”

subfolder should not need to be altered for fitting the SiGe binary. It is recommended

that a new “SiGe data” subfolder be created using previous examples as a template.

The data subfolder will need to adjusted significantly to reflect the properties of SiGe.

The DFT data obtained for the SiGe dimer, NaCl, and ZB structures is placed in the

“objdata” array along with weight priorities. Figure A.2 presents a sample objdata input

array for the Ge elemental system. Guidance on appropriate weight priorities is difficult

to provide. A trial and error approach to assigning weights to obtain insight into their

effects would be of benefit to a new user. Additional non 1st nearest neighbor

structures can be input as “auxiliary structures”. These structures are not fitted,

however, the fitGSP output will give estimates on the pair fitting and the corresponding

BO targets. Each auxiliary structure is placed in the “auxdata” array as a matrix

composed of structural information. An example for the Ge -Sn phase would be:

Page 160: Bond Order Potentials for Group IV Semiconductors

160

04944

122263.3"35"

028173.2"25"

0444

153314.2"15"

22

22

caA

cA

caA

The first column indicates the structure and which nearest neighbor the data refers to;

in this case “A5-1” refers to the first nearest neighbor in the -Sn structure. The second

column is the separation distance. The third column is an equation for the separation

distance as a function of the lattice constants. The fourth column is the atomic

Figure A.2 The primary data entry array found in the fitGSP.nb Mathematica fitting

program. The columns have the following definitions:

I. Crystal Structure

II. Nearest Neighbor Distance, r

III. Equation for r based on the lattice parameters

IV. Cohesive Energy per atom

V. Atomic Coordination

VI. Bulk Modulus in GPa

VII. Atomic Volume as a function of r

VIII. Fitting weight for cohesive energy

IX. Fitting weight for bulk modulus

objdata =

“dimer” 2.3 r -1.32 1 250 150

“A4” 2.4465

-3.82 4 62.17

500 200

“Ah” 2.6588 a -3.578 6 67.38 100 50

“A1” 2.9652

-3.484 12 63.49

600 100

I II III IV V VI VII VIII IX

fitGSP.nb Mathematica Excerpt

Page 161: Bond Order Potentials for Group IV Semiconductors

161

coordination of the 1st, 2nd or 3rd nearest neighbor shell. The fifth column defaults to

zero, corresponding to a feature in the code that was never finalized.

The final step before fitting the GSP parameters is choosing constraints on the

parameters. The preset variable ranges are generally sufficient for fitting, however

these can be changed if so desired (although a parameter value of zero or less would be

unphysical, this should be avoided for obvious reasons). The parameter k corresponds

to the m/n ratio (1/k = m/n) which determines the potential hardness. For group IV

semiconductors a m/n ratio target of ~2 is theoretically motivated (for example Si has a

m/n value of 1.8). To ensure a proper m/n ratio, the value of k can be constrained to 0.5

in fitGSP (this value may be adjusted later in the third stage; however it should not be

allowed to deviate significantly from the target of ~2).

Fitting the GSP parameters is an iterative process. The initial structure data for

SiGe may not provide a data trend with points that intersect a smooth GSP curve. This

makes the fitting process difficult, since all the point cannot be intersected. The initial

data must be iteratively adjusted to fit a smooth GSP curve (without significant sacrifice

of important considerations). A non-smooth GSP curve will provide unusual

abnormalities in simulation which are unintended and unrealistic, therefore it is

considered mandatory that the GSP function be smooth and continuous. The default

minimization technique uses Differential Evolution, a genetic algorithm that maintains a

population of specimens, x1… xn. At each iteration a random combination of specimens

is “mated” to form a new specimen which is compared to its progenitors. If the new

Page 162: Bond Order Potentials for Group IV Semiconductors

162

specimen represents an improvement it replaces one of the originals, if not it is

discarded. The differential evolution method is employed because it is considered quite

robust.

Once a suitably smooth GSP curve has been obtained that inTersoffects the

adjusted SiGe target values the fitGSP program will provide a set of three matrices that

are the data inputs for fitBO. These can be copy/pasted directly into the appropriate

fitBO file.

fitBO

The fitGSP program obtains parameters that determine the radially dependent

pair-wise interactions. The fitBO program determines the appropriate angular

dependencies and bond order strengths. the fitBO determines values for the following

parameters: (ds), f, k, c, p (p), p (pn), and (parameters in parenthesis are the

Mathematica equivalent). The f and k parameters are not appropriate for use with the

SiGe system; they are designed to accommodate semiconductor systems with non-half

filled valence shells (such as the III-V GaAs system). As such they are assigned values of f

= 0.5 and k = 1.0 (these values result in the valence skewing term being set to zero). For

both the Si and Ge elemental systems the angular parameter c has been set to zero; this

would be a good condition to start with for fitting SiGe, although additional angular

flexibility may be needed. If the parameter c is employed, Equation 2.23 in section 2

would become:

Page 163: Bond Order Potentials for Group IV Semiconductors

163

c

cpg

ijk

ijk

1

2cos)1(cos1

A.1

Additional parameters cprom and dprom (corresponding to the promotion energy terms

A and respectively) are listed in fitBO, however these parameters are not assigned any

significance in the Mathematica program (they are declared but never used). These two

parameters can be given any value desired, however they have been set to a default of

cprom = 0 and dprom = 1. Similarly the parameters b2n and b2o should also be

assigned values of one. These parameters were from an earlier formulation of the BOP

and are not employed in the Si and Ge potentials. After all these considerations only the

parameters , p, p, and should not be assigned any constant value.

Use the inputs from fitGSP and the preexisting samples in fitBO to construct the

input subfolder for the BO fitting. This subfolder should consist of initial variable

assumptions (detailed in previous paragraph), the three data matrices from fitGSP, and

assigned fitting weights for the energy, the BO and the pressure. The weight placed on

pressure indicates the priority on maintaining the provided structure data as a minimum

of energy. Typically the pressure weights are the highest to ensure that the correct

crystal structure is maintained. Choosing the appropriate values for the weights has

been described as an intuitive “art” rather than a science [102]. There is some truth to

this, as a significant amount of trial and error is poured into the fitting process. It is also

possible to include defect structures at this point of the fitting (i.e. a single germanium

atom in a silicon lattice or similar situations). This is done by reading in a

Page 164: Bond Order Potentials for Group IV Semiconductors

164

preconstructed data file containing the xyz lattice data for a crystal of interest. The

fitBO program can be set to calculate the energy of this system, from which you can

determine the energy of interest. There are four minimization schemes preset into

fitBO: automatic (Mathematica chooses a minimization scheme based on the nature of

the problem), Differential Evolution, Nelder Mead, and Simulated Annealing. Employing

a combination of these methods generally gives the best result.

The program fitBO outputs all the necessary parameters (including the

placeholder values for the promotion energy terms) in the “fitvalu” array. This array

lists the parameters in the following order (using the Mathematica nomenclature): b2n,

b2o, c, cprom, dprom, ds, f, k, nc, nf, ns, p, pn, r0, r1, rcut, 0, 0 and 0. This array is

the input for another Mathematica program file designed to generate a potential force

file for the Fortran MD code. Because we still need to input the promotion energy

functionality into the potential this 4th program cannot be employed. Also outputted by

the fitBO minimization subroutine is data regarding the quality of the fit. This data

includes the energy predictions for the structures provided as well as data about the BO.

A small graphic of the fitted angular dependence is also generated.

Fitting Uprom

In the third stage of the fitting process we introduce the promotion energy,

Uprom. The introduction of a non-zero promotion energy term will alter the potential

predictions in sometimes unexpected ways. It is therefore important to understand

Page 165: Bond Order Potentials for Group IV Semiconductors

165

how each parameter will influence the shape of the potential. A visual approach based

on the dimer energy curve is perhaps the simplest. Figures A.3 to A.6 compile these

effects. These images use the fitted Ge potential as their baseline. The parameters are

then adjusted by ±5% and ±10% to illustrate the nature of their influence on the shape

of the dimer energy. What follows are general comments on the nature of the changes

observed for varying each parameter individually. The effects of varying the parameters

in the many myriad combinations available are not discussed.

The bond energy coefficient terms (0, 0 and 0) behave as one would expect

when varied. Lowering 0, the repulsive energy coefficient, results in a lower energy for

Figure A.3 The effects of altering the

BOP parameters , and by small

increments. The direction of motion is

indicated by the + and – signs, where a

+ indicates the parameter was

increased and a – indicates the inverse.

+

+

+ -

-

-

Page 166: Bond Order Potentials for Group IV Semiconductors

166

all structures (i.e. stronger bonds). The reverse is also true; increasing 0 raises the

energy. A secondary effect of lowering 0 is that the separation distance, rij, at which

the potential predicts a minimum of energy also decreases. The parameters 0 and 0

influence the potential in a similar fashion, although because that are a measure of the

bond strength and not repulsion the trends are reversed. Increasing 0 and 0 results

in a lowering of the bond energy as well as shifting the energy minimum to the left

(lower NN separation). The bond energy coefficients can therefore be used to “tune”

the BOP to match the appropriate minimum energy configuration (in the case of silicon

and germanium, the dc structure). These energy terms also have a small influence on

the curvature of the potential near the minimum, although other terms will have a

greater influence (increasing 0 and decreasing 0 and 0 results in a lower curvature

and vice versa).

Figure A.4 The effects of altering the BOP promotion energy parameters A and by

small increments. The direction of motion is indicated by the + and – signs, where a

+ indicates the parameter was increased and a – indicates the inverse.

+ + - - A

Page 167: Bond Order Potentials for Group IV Semiconductors

167

The promotion energy terms A and (called cprom and dprom in Mathematica

respectively) have very similar effects on the shape of the potential to the repulsive

energy coefficient 0. This is because they are both positive terms in the potential. The

promotion energy represents the energy penalty incurred by hybridizing the atomic

orbitals; therefore as the magnitude of the promotion energy is increased the bond

energy is weakened. The promotion energy has the form

212

,11

ij

ij

prom AU

9.2

Therefore if A is very large Uprom = and if A is very small Uprom = 0. If is very large

Uprom = 0 and if is very small Uprom = 0. It is therefore important to keep a careful

balance for these parameters. For the parameterization of the SiGe binary initial values

Figure A.5 The effects of altering the BOP cutoff parameters r1 and rcut by small

increments. The direction of motion is indicated by the + and – signs, where a +

indicates the parameter was increased and a – indicates the inverse.

+ +

- - A

- -

+ +

r1 rcut

Page 168: Bond Order Potentials for Group IV Semiconductors

168

of A = 1.325 and = 8.342 would not be unreasonable (the average between the Si and

Ge elemental potential parameters).

The cutoff parameters r1 and rcut have little influence on the potential at or near

equilibrium separation distances. This is of course by design; the inclusion of a potential

cutoff is not physically motivated but required for computational purposes. By changing

these terms you can introduce non-physically motivated effects into the potential. This

is evidenced by the small separation distance around 3.4 Å in Figure A.5 for rcut in which

the dimer binding curve becomes positive. Increasing the cutoff distance also results in

Figure A.6 The effects of altering the BOP cutoff parameters r0, rc, rc and rc by

small increments. The direction of motion is indicated by the + and – signs, where a

+ indicates the parameter was increased and a – indicates the inverse.

+ +

+ +

-

-

- -

r0 rc

rc rc

Page 169: Bond Order Potentials for Group IV Semiconductors

169

the inclusion of additional atoms in the neighbor list for certain structures. This greatly

increases the computational cost of simulations for little benefit and must be avoided.

It is also important that the cutoff not be too sharp as this may result in aberrant

behavior in simulation. It is therefore important that r1 be sufficiently smaller than rcut

to provide a smooth cutoff. If r1 is too small however, structures with large NN

separation distances will be affected, most notably the fcc structure.

The parameters r0, rc, rc and rc have a profound effect on the shape of the

potential. The parameter r0 is intended to be the separation distance corresponding to

the minimum energy of the dimer; however the inclusion of the promotion energy term

prevents that. This is because the promotion energy term is not written in the form of a

GSP pair term. As a result the dimer minimum energy is located at a slightly higher

value (2.32 Å vs. 2.395 Å). Altering r0 allows one to control the position of the energy

minimum directly with little change to any other property. The parameter r0 can

therefore be used to “tune” the potential directly to the desired minimum energy

structure. The parameters rc and rc have the distinction that they are the only

parameters which were capable of altering the order of energy preference between

structures with only a 10% change in value. The dramatic changes observed in the

shape of the dimer curves in Figure A.6 alter the order of energy preference for atomic

structures. When rc is decreased, the energy minimum shifts to the right and is

lowered. As a result, the energy minimum is much closer to the fcc NN distance and the

fcc structure becomes the lowest energy structure. A small increase in rc does not

significantly shift the position of the dimer energy minimum, but does increase the

Page 170: Bond Order Potentials for Group IV Semiconductors

170

energy curve slightly at larger radius. This results in the fcc structure becoming

significantly less stable. The parameters rc and rc behave in a similar fashion. Great

care must be employed when altering these parameters.

The six GSP exponential parameters nx and ncx (where x is , or ) do not

significantly alter the shape of the potential when varied within 10% (as such they are

not shown in a separate figure). These parameters are therefore more suited for “fine

tuning” the potential than providing large blunt alterations such as can be achieved by

varying other terms such as r0. All six of these terms have a slightly larger influence on

the energy of structures with larger atomic spacing. Perhaps the most interesting

feature of these parameters is that there exists a value of r (the separation distance

between atom i and j) such that the predicted energy is invariant upon nx and ncx. This

occurs at r = r0 by design.

Using these parameter trends it is possible to incorporate a physically

meaningful promotion energy term into the pair-wise BOP obtained in the first two

fitting stages. A significant investment in time will be necessary to fit the SiGe binary

BOP. Using the suggestions laid out here should greatly assist a future researcher with

the project.

Page 171: Bond Order Potentials for Group IV Semiconductors

171

Appendix B. Density Functional Theory Calculations

A portion of the density functional theory (DFT) estimates that are presented in

this dissertation were calculated personally by the author using the Vienna Ab Initio

Simulation Package (VASP). VASP is a software package created, distributed, and

maintained by the Hafner Research Group at the University of Vienna that performs ab-

initio quantum mechanical calculations using pseudopotentials and a plane wave basis

set. The simulation approach implemented in VASP is founded on the finite

temperature local density approximation with the free energy as a variational quantity.

The VASP guide, currently available online at http://cms.mpi.univie.ac.at/vasp/, is an

excellent resource for any researcher attempting to use the software package and

served as the primary reference source for the calculations performed by the author.

The VASP guide provides significant assistance in understanding program specific

terminology, file names and control flags that control the program. References 146 to

149 provide extensive detail on the program itself and the theory behind it. This

dissertation makes use of VASP to obtain a consistent database of silicon and

germanium bulk properties for use in the fitting process. The input files employed by

Page 172: Bond Order Potentials for Group IV Semiconductors

172

VASP are INCAR, POTCAR, KPOINTS and POSCAR. These input files are described here

and the non-default values employed are shown.

B.1 INCAR

The file INCAR is the central input file of VASP. It provides the “what to do and

how to do it” for the program. There are a large number of adjustable parameters for

the researcher available within INCAR, however it is not necessary to change many of

them as the default values are typically convenient. The non-default values that were

specified in the VASP estimates performed are as follows:

ISMEAR = -5 (the recommended value for semiconductors)

ENCUT = 300 (good energy convergence was found using this value for the

cutoff energy)

B.2 POTCAR

This file contains the pseudopotential (PP) for each atomic species in the

calculation. The VASP package comes with a database of available PPs for the majority

of elements that one would consider. Variations in PPs are also available for many of

the elements, including silicon and germanium. The PP that was chosen for use here

was the projector augmented wave (PAW) potential. This was chosen because it comes

highly recommended by the developers of VASP.

Page 173: Bond Order Potentials for Group IV Semiconductors

173

B.3 KPOINTS

This input file, as its name suggests, specifies the specific k-point coordinates for

generating the k-point grid. Two methods exist for the input of the k-points, explicit and

automatic generation. The automatic k-point generation method has been selected

here because of the relatively simple nature of the calculations. In this method you are

given the option of choosing between either the Monkhorst or Gamma k mesh

generator methods. It is recommended in the VASP literature that the Monkhorst

method be chosen for cubic crystal cells and the Gamma method for hexagonal crystal

cells. A sample KPOINTS file to be used with a sc crystal would read as follows:

Automatic (first line is considered a comment line)

0 (number of k-points = 0 indicates automatic generation)

Monkhorst-Pack (indicates the use of the Monkhorst generator method)

4 4 4 (k-point mesh subdivisions)

0 0 0 (optional shift of the mesh)

B.4 POSCAR

This file contains the specific lattice information that generates the crystal.

Whereas each of the previous files were largely independent of the specific crystal being

examined, the POSCAR file must be specifically tailored to each crystal. A sample

POSCAR file to generate a fcc lattice is shown below.

fcc cubic crystal (first line is considered a comment)

3.819 (lattice constant)

Page 174: Bond Order Potentials for Group IV Semiconductors

174

1 0 0

0 1 0 (unit cell lattice vectors)

0 0 1

4 (number of atoms within the unit cell)

Direct (specifies the method atomic positions are provided)

0 0 0 (atomic positions)

0.5 0.5 0

0.5 0 0.5

0 0.5 0.5

Page 175: Bond Order Potentials for Group IV Semiconductors

175

Appendix C. Qualitative Analysis Determinants

Table C.1 This table describes the ranges used to determine the numerical value (1 to

5) assigned to each property in Table 8.1. The parameter x represents the percentage

difference between the DFT estimate and the MD prediction. The parameter

represents the numerical difference between the DFT estimate and the MD prediction.

Property 1 2 3 4 5

Va x > 10% 5% < x < 10% 2% < x < 5% 1% < x < 2% x < 1%

Ec > 0.5 0.3 < < 0.5 0.1 < < 0.3 0.02 < < 0.1 < 0.02

B x > 100% 50% < x < 100% 20% < x < 50% 6% < x < 20% x < 6%

Clusters* x > 40% 30% < x < 40% 20% < x < 30% 10% < x < 20% x < 10%

Defects** x < 30% 20% < x < 30% 10% < x < 20% 5% < x < 10% x < 5%

Tm ±400 K ±300 K ±200 K ±100 K ±50 K

Surfaces 0.1 < 0.04 < 0.02 <

* A -1 penalty was applied if the lowest energy cluster did not match the HF predicted

lowest energy structure

** A -2 penalty was applied if the lattice around the vacancy did not relax inward.

Page 176: Bond Order Potentials for Group IV Semiconductors

176

References

1. www.webelements.com

2. R.F. Pierret, Semiconductor Device Fundamentals, Addison-Wesley Publishing Company, Reading

Mass (1996).

3. J.D. Schaffer, A. Saxena, S.D. Antolovich, T.H. Sanders and S.B Warner, The Science and Design of

Engineering Materials, 2nd Ed. McGraw-Hill, Boston Mass (1999).

4. R.E. Reed-Hill and R. Abbaschian, “Physical Metallurgy Principles” 3rd Ed. PWS Publishing, Boston

Mass (1994).

5. A. Thayer, CENEAR v78 n48, 27 (2000).

6. W.B. Shockley, U.S. Patent 2502488 “Semiconductor Amplifier” Sept. 24 (1948).

7. R. Kessler, “Absent at the Creation; How One Scientist Made Off with the Biggest Invention Since

the Light Bulb”, The Washington Post Magazine, pg 16, Apr 6 (1997).

8. CRC Handbook of Chemistry and Physics 84th Ed., edited by D.R. Lide (CRC PRess, New York 2004).

9. W.H. Brattain and H.B. Briggs, Phys. Rev. 75, 1705 (1949).

10. K. Said, J. Poortmans, M. Caymax, J.F. Nijs, L. Debarge, E. Christoffel, and A. Slaoui, IEEE Trans.

Elec. Dev. 46, 2103 (1999).

11. S.A. Healy and M.A. Green, Sol. Energy Mater. Sol. Cells 28, 273 (1992).

12. T. Brunhes, P. Boucaud, S. Sauvage, F. Aniel, J.M. Lourtiozm C. Hernandez, Y. Campidelli, O.

Kermarrec, D. Bensahel, G. Faini, and I. Sagnes, Appl. Phys. Lett. 77, 1822 (2000).

13. R. Apetz, L. Vescan, A. Hartmann, C. Dieker, and H. Luth, Appl. Phys. Lett. 66, 445 (1995).

Page 177: Bond Order Potentials for Group IV Semiconductors

177

14. S.O. Kasap, Principles of Electrical Engineering Materials and Devices, Revised Ed., McGraw-Hill,

Boston (2000).

15. M. Shima, Jujitsu Sci. Tech J. 39, 78 (2003).

16. D. Paul, Physics World 14, 27 (2000).

17. G.L. Miessler and D.A. Tarr, Inorganic Chemistry 2nd ed, Prentice Hall, Upper Saddle River, NJ

(1999).

18. P.C. Zalm, G.F.A. Van De Walle, D.J. Gravesteijn, and A.A. Van Gorkum, Appl. Phys. Lett. 55, 2520

(1989).

19. B. Voigtlander and T. Weber, Phys. Rev. Lett. 77, 3861 (1996).

20. B. Voigtlander, T. Weber, P. Smilauer and D.E. Wolf, Phys. Rev. Lett. 78, 2164 (1997).

21. T. Michely, M. Hohage, S. Esch and G. Cosma, Surf. Sci. 349, L89 (1996).

22. Y.H. Xie, G.H. Gilmer, E.A. Fitzgerald and J. Michel, Columnar Structure Growth by Silicon

Molecular Beam Epitaxy, Mat. Res. Soc. Sym. Proc. 220, ed. J.C. Bean, S.S. Iyer and K.L. Wang, 41

(1991).

23. Y.W. Mo, R. Kariotis, B. Swartzentruber, M.B. Webb and M.G. Lagally, Phys. Rev. Lett. 63, 2393

(1989).

24. R. Hammers, U.K. Kohler and J.E. Demuth, Ultramicroscopy 31, 10 (1989).

25. M. Asai, H. Ueba and C. Tatsuyama, J. Appl. Phys. 58, 2577 (1985).

26. P.M.J. Maree, K. Nakagawa, F.M. Mulders and J.F. Van Der Veen, Surf. Sci. 191, 305 (1987).

27. S.S. Iyer, J.C. Tsang, M.W. Copel, P.R. Pukite and R.M. Tromp, Appl. Phys. Lett. 54, 219 (1989).

28. J.C. Bean, L.C. Feldman, A.T. Firoy, S. Nakahara and I.K. Robinson, J. Vacuum Sci. and Tech. A 2,

436 (1984).

29. R. People and J.C. Bean, Appl. Phys. Lett. 47, 322 (1985).

30. G.L. Olson and J.A. Roth, Mater. Sci. Rep. 3, 1 (1988).

31. T. Motooka, Mater. Sci. Eng. A 253, 42 (1998).

32. F. Priolo and E. Rimini, Mat. Sci. Rep. 5, 319 (1990).

Page 178: Bond Order Potentials for Group IV Semiconductors

178

33. K. Uesugi, T. Yao, T. Sato, T. Sueyoshi and M. Iwatsuki, Appl. Phys. Lett. 62, 1600 (1993).

34. N. Bernstein, M.J. Aziz, and E. Kaxiras, Phys. Rev. B 61, 6696 (2000).

35. T. Motooka, K. Nisihira, S. Munetoh, K. Moriguchi, and A. Shintani, Phys. Rev. B 61, 8537 (2000).

36. B. Weber, K. Gartner and D.M. Stock, Nuc. Ints. and Meth. in Phys. Res. B 127/128, 239 (1997).

37. B. Weber, D.M. Stock and K. Gartner, Nuc. Ints. and Meth. in Phys. Res. B 148, 375 (1999).

38. D.M. Stock, B. Weber and K. Gartner, Phys. Rev. B 61, 8150 (2000).

39. B. Weber and K. Gartner, Nuc. Ints. and Meth. in Phys. Res. B 175, 119 (2001).

40. K. Gartner and B. Weber, Nuc. Ints. and Meth. in Phys. Res. B 202, 255 (2003).

41. X.W. Zhou, R.A. Johnson and H.N.G. Wadley, Phys. Rev. B 69, 144113 (2004).

42. X.W. Zhou, H.N.G. Wadley, R.A. Johnson, D.J. Larson, N. Tabat, A. Cerezo, A.K. Petford-Long, G.D.

Smith, P.H. Clifton, R.L. Martens and T.F. Kelly, Acta Mater. 49 4005 (2001).

43. X.W. Zhou and H.N.G. Wadley, J. Appl. Phys. 87, 8487 (2000).

44. W. Zou, H.N.G. Wadley, X.W. Zhou, R.A. Johnson and D. Brownell, Phys. Rev. B 64, 174418 (2001).

45. X.W. Zhou, H.N.G. Wadley, and D.X. Wang, Comp. Mat. Sci. 39, 794 (2007).

46. F.H. Stillinger and T.A. Weber, Phys. Rev. B 31, 5262 (1985).

47. J. Tersoff, Phys. Rev. Lett. 56, 632 (1986).

48. J. Tersoff, Phys. Rev. B 37, 6991 (1988).

49. J. Tersoff, Phys. Rev. B 38, 9902 (1988).

50. X. G. Gong, Phys. Rev. B 47, 2329 (1993).

51. A. D. Mistriotis, G. E. Froudakis, P. Vendras and N. Flytzanis, Phys. Rev. B 47, 10648 (1993).

52. B.C. Bolding and H.C. Andersen, Phys. Rev. B 41, 10568 (1990).

53. D. W. Brenner, Phys. Rev. B 42, 9458 (1990).

54. K. Ding and H.C. Andersen, Phys. Rev. B 34, 6987 (1986).

55. D. A. Murdick, X. W. Zhou, and H. N. G. Wadley, Phys. Rev. B 72, 205340 (2005).

56. D. G. Pettifor and I.I. Oleinik, Phys. Rev. B 59, 8487 (1999).

57. D. G. Pettifor and I.I. Oleinik, Phys. Rev. Lett. 84, 4124 (2000).

Page 179: Bond Order Potentials for Group IV Semiconductors

179

58. D. G. Pettifor and I.I. Oleinik, Phys. Rev. B 65, 172103 (2002).

59. D. G. Pettifor, I.I. Oleinik, D. Nguyen-Manh, et al., Comp. Mat. Sci. 23, 33 (2002).

60. D.A. Murdick, X.W. Zhou, H.N.G. Wadley, D. Nguyen-Manh, R. Drautz, and D.G. Pettifor, Phys.

Rev. B 73, 045206 (2006).

61. B.J. Alder and T.E. Wainwright, J. Chem. Phys. 27, 1208 (1957).

62. B.J. Alder and T.E. Wainwright, J. Chem. Phys. 31, 459 (1959).

63. H. Balamane, T. Halicioglu, and W.A. Tiller, Phys. Rev. B 46, 2250 (1994).

64. D. Bedrov, G.D. Smith, and T.D. Sewall, J. Chem. Phys. 112, 7203 (2000).

65. S.L. Chaplot, N. Choudhury, and K.R. Rao, Amer. Mineralogist 53, 937 (1998).

66. F.F. Abraham, D. Brodbeck, R.A. Rafoy, and W.F. Rudge, Phys. Rev. Lett. 73, 272 (1994).

67. V.V. Zhakhovski, K. Nishihara, S.I. Anigimov, and N.A. Inogamon, JETP Lett. 71, 107 (2000).

68. J.W. Gibbs, Elementary Principles in Statistical Mechanics, C. Scribner’s Sons, New York (1902).

69. C. Fiolhais, F. Noqueira, and M.A.L. Marques, “A Primer in Density Functional Theory”, Springer

(2003).

70. L.C. Lew Yan Voon, Electronic and Optical Properties of Semiconductors: A Study Based on the

Empirical Tight Binding Model, Dissertation.com (1997).

71. I. Torrens, Interatomic Potentials, Academic Press (1972).

72. J.E. Lennard-Jones, Proc. of the Phys. Soc. 43, 461 (1931).

73. X.W. Zhou, private communications.

74. Y. Waseda and K. Suzuki, Z. Phys. B 20, 339 (1975).

75. J.P. Gabathuler and S. Steeb, Z. Naturforsch. 34a, 1314 (1979).

76. M. Davidovic, M. Stojic, and Dj. Jovic, J. Phys. C 16, 2053 (1983).

77. V.M. Glazov, S.N. Chizhevskaya, and N.N. Glagoleva, Liquid Semiconductors, Plenum, New York

(1969).

78. A.D. Mistriotis, N. Flytzanos, and S.C. Farantos, Phys. Rev. B 39, 1212 (1989).

Page 180: Bond Order Potentials for Group IV Semiconductors

180

79. P. Villars and L.D. Calvert, Pearson's Handbook of Crystallographic Data for Intermetallic Phases,

2nd Edition (ASM International, Materials Park, Ohio, 1991).

80. B.P. Feuston, R.K. Kalia, and P. Vashishta, Phys. Rev. B 37, 6297 (1988).

81. L.A. Bloomfield, R.R. Freeman, and W.L. Brown, Phys. Rev. Lett. 54, 2246 (1985).

82. K. Raghavachari and V. Logovinsky, Phys. Rev. Lett. 55, 2853 (1985).

83. K. Raghavachari J. Chem. Phys. 84, 5672 (1986).

84. K.E. Khor and S. Das Sarma, Phys. Rev. B 36, 7733 (1987).

85. K.E. Khor and S. Das Sarma, Phys. Rev. B 39, 1188 (1989).

86. S.C.A. Gay and G.P. Srivastava, Phys. Rev. B 60, 1488 (1999).

87. J. Cai and J. Wang, Phys. Rev. B 64, 035402 (2001).

88. Y. Yamashita, S. Machida, M. Nagao, S. Yamamoto, Y. Kakefuda, K. Mukai, and J. Yoshinobu, Jpn.

J. Appl. Phys. 41, L272 (2002).

89. J. Cho and K.S. Kim, Phys. Rev. B 69, 125312 (2004).

90. J. Ferrante, J.R. Smith, and J.H. Rose, Phys. Rev. Lett. 55, 1385 (1983).

91. J.H. Rose, J.R. Smith, and J. Ferrante, Phys. Rev. B 28, 1835 (1983).

92. A.J. Dyson and P.V. Smith, Surf. Sci. 316, 309 (1994).

93. D. Mura, L. Colombo, R. Bertoncini, and G. Mula, Phys. Rev. B 58, 10357 (1998).

94. J. Tarus and K. Nordlund, Thin Solid Films 464-465, 95 (2004).

95. T. Motooka, K. Nisihira, S. Munetoh, K. Moriguchi, and A. Shintani, Phys. Rev. B 61, 8537 (2000).

96. D.G. Pettifor, Bonding and structure of molecules and solids, Oxford science publications (Oxford

University Press, Oxford, 1995).

97. D.G. Pettifor, Springer Proc. Phys. 48, 64 (1990).

98. M. Finnis, Interatomic Forces in Condensed Matter, Oxford Series on Materials Modelling (Oxford

University Press, Oxford, 2003).

99. L. Goodwin, A.J. Skinner, and D.G. Pettifor, Europhys. Lett. 9, 701 (1989).

Page 181: Bond Order Potentials for Group IV Semiconductors

181

100. R. Drautz, D.A. Murdick, D. Nguyen-Manh, X. Zhou, H.N.G. Wadley, and D.G. Pettifor, Phys. Rev.

B 72, 144105 (2005).

101. K. Albe, K. Nordlund, J. Nord, and A. Kuronen, Phys. Rev. B 66, 035205 (2002).

102. D.A. Murdick, “Simulating the Atomic Assembly of Galium Arsenide” Dissertation (UVA Press,

Charlottesville, 2006).

103. J.C. Tsang, K. Eberl, S. Zollner, and S.S. Iyer, Appl. Phys. Lett. 61, 961 (1992).

104. K. Raghavachari and V. Logovinsky, Phys. Rev. Lett. 55, 2853 (1985).

105. K. Raghavachari J. Chem. Phys. 84, 5672 (1986).

106. L.J. Munro and D.J. Wales, Phys. Rev. B 59, 3969 (1999).

107. P.M. Fahey, P.B. Griffin, and J.D. Plummer, Rev. Mod. Phys. 61, 289 (1989).

108. A.E. Michel, W. Rausch, P.A. Ronsheim, and R.H. Kastl, Appl. Phys. Lett. 50, 416 (1987).

109. R. Car, P.J. Kelly, A. Oshiyama, and S.T. Pantelides, Phys. Rev. Lett. 52, 1814 (1984).

110. S.J. Clark and G.J. Ackland, Phys. Rev. B 56, 47 (1997).

111. H.R. Schober, Phys. Rev. B 39, 13013 (1989).

112. M. Nastar, V.V. Bulatov, and S. Yip, Phys. Rev. B 53, 13521 (1996).

113. W. Lee, S. Lee, and K.J. Chang, J. Phys. Condens. Matter 10, 995 (1998).

114. K.C. Pandey, Phys. Rev. Lett. 57, 2287 (1986).

115. Y. Bar-Yam and J.D. Joannopoulos, Phys. Rev. Lett. 52, 1129 (1984).

116. P.J. Kelly and R. Car, Phys. Rev. B 45, 6543 (1992).

117. H. Seong and L.J. Lewis, Phys. Rev. B 53, 9791 (1996).

118. S. Goedecker, T. Deutsch, and L. Billard, Phys. Rev. Lett. 88, 235501 (2002).

119. J. L. Mercer, J.S. Nelson, A.F. Wright, and E.B. Stechel, Modeling Simul. Mater. Sci. Eng. 6, 1

(1998).

120. J.R. Morris, C.Z. Wang, K.M. Ho, and C.T. Chan, Phys. Rev. B 49, 3109 (1994).

121. R.E. Schiler and H.E. Farnsworth, J. Chem. Phys. 30, 917 (1959).

Page 182: Bond Order Potentials for Group IV Semiconductors

182

122. R.M. Tromp, R.J. Hamers, and J.E. Demuth, Phys. Rev. Lett. 55, 1303 (1985); R.J. Hamers, R.M.

Tromp, and J.E. Demuth, Phys. Rev. B 34, 5343 (1986).

123. K. Takayanagi, Y. Tanishiro, M. Takahashi, and S. Takahashi, J. Vac. Sci. Technol. A 3, 1502 (1985).

124. R.A. Wolkow, Phys. Rev. Lett. 68, 2636 (1992).

125. J. Dabrowski and M. Scheffler, Appl. Surf. Sci. 56, 15 (1992).

126. Y. Fukaya and Y. Shigeta, Phys. Rev. Lett. 91, 126103 (2003).

127. A.A. Stekolnikov, J. Furthmuller, and F. Bechstedt, Phys. Rev. B 70, 045305 (2004).

128. W. Ranke, Phys. Rev. B 41, 5243 (1990).

129. J. Dabrowski, H.J. Mussig, and G. Wolff, Phys. Rev. Lett. 73, 1660 (1994).

130. J. Wang, A.P. Horsfield, D.G. Pettifor, and M.C. Payne, Phys. Rev. B 54, 13744 (1996).

131. B.A. Gillespie, X.W. Zhou, D.A. Murdick, H.N.G. Wadley, R. Drautz, and D.G. Pettifor, Phys. Rev. B

75, 155207 (2007).

132. G.X. Qian and D.J. Chadi, Phys. Rev. B 35, 1288 (1987).

133. J.C. Bean, Semiconductors and Semimetals 56, 1 (1999).

134. M.Z. Bazant, E. Kaxiras, and J.F. Justo, Phys. Rev. B 56, 8542 (1997).

135. G. Otto, G. Hobler, K. Gartner, Nuc. Ints. and Meth. Phys. Res. B 202, 114 (2003).

136. P.L. Liu, R. Yen, N. Bloembergen and R.T. Hodgson, Appl. Phys. Lett. 34, 864 (1979).

137. M.H. Brodsky, R.S. Title, K. Weiser and G.D. Pettit, Phys. Rev. B 1, 2639 (1970).

138. T.I. Kemins, M.M. Mandurah and K.C. Saraswat, J. Electrochem. Soc. 125, 927 (1978).

139. J.A. Roth and C.L. Anderson, Appl. Phys. Lett. 31, 689 (1977).

140. J.W. Mayer, L. Eriksson and J.A. Davies, Ion Implantation in Semiconductors, Academic Press,

New York (1970).

141. W.L. Brown, Mater. Res. Soc. Symp. Proc. 51, 53 (1985).

142. G.Q. Lu, E. Nygren and M.J. Aziz, J. Appl. Phys. 70, 5323 (1991).

143. X.W. Zhou, The Nitrogen Strengthening Mechanism in the Single Crystalline Fe-Ni-Cr Austenitic

Stainless Steels, Dissertation, Clemson University (1995).

Page 183: Bond Order Potentials for Group IV Semiconductors

183

144. C.R.S. da Silva and A. Fazzio, Phys. Rev. B 64, 075301 (2001).

145. K. Koharty, private communications, Oxford University (2006).

146. G. Kreese and J. Hafner, Phys. Rev. B 47, 558 (1993); 49, 14251 (1994).

147. G. Kreese and J. Furthmuller, Comp. Mater. Sci. 6, 15 (1996).

148. G. Kreese and J. Furthmuller, Phys. Rev. B 54, 11169 (1996).

149. G. Kreese and J. Hafner, J. Phys. Condens. Matter 6, 8245 (1994).

150. C.H. Bates, F. Dachille, and R. Roy, Science 147, 860 (1965).

151. W. Xu, Y. Zhao, Q. Li, Y. Xie, and H.F. Schaefer III, Mol. Phys. 20, 579 (2004).

152. G. Pacchioni and J. Koutecky, J. Chem. Phys. 84, 3301 (1986).

153. A.J.R. da Silva, A. Janotti, A. Fazzio, R.J. Baierle and R. Mota, Phys. Rev. B 62, 9903 (2000).

154. M. Dionizio Moreira, R.H. Miwa and P. Venezuela, Phys. Rev. B 70, 115215 (2004).

155. L. Spiess, A.J. Freeman and P. Soukiassian, Phys. Rev. B 50, 2249 (1994).

156. N. Takeuchi, A. Selloni and E. Tosatti, Phys. Rev. B 51, 10844 (1995).

157. L. Zhigilei, Introduction to Computational Materials Science, Lecture Notes.

158. J.R. Heath, Y. Liu, S.C. O’Brien, Q. Zhang, R.F. Curl, F.K. Tittel and R.E. Smalley, J. Chem. Phys. 83,

5520 (1985).

159. J.E. Kingcade, H.M. Nagarathna-Naik, I. Shim, and K.A. Gingerich, J. Phys. Chem. 90, 2830 (1986).

160. O. Cheshnovsky, S.H. Yang, C.L. Pettiette, M.J. Craycraft, Y. Liu, and R.E. Smalley, Chem. Phys.

Lett. 138, 119 (1987).

161. M. Scheer, R.C. Bilodeau, C.A. Brodie, and H.K. Haugen, Phys. Rev. A 58, 2844 (1998).

162. W.H. Press, Numerical Recipes in Fortran 77: The Art of Scientific Computing, 2nd ed. (Cambridge

University Press, New York, 1996).

163. G.W.C. Kaye and T.H. Laby, Tables of Physical and Chemical Constants, 15th ed. (Longman,

London, UK, 1993).

164. M. Ishimaru, K. Yoshida and T. Motooka, Phys. Rev. B 53, 7176 (1996).

165. X.W. Zhou, D.A. Murdick and H.N.G. Wadley, J. Appl. Phys. 99, 064908 (2006).