Vous êtes sur la page 1sur 122

Doctoral Thesis

Reactive Molecular Dynamic Studies of Initial Stage of Si(001) Surface Oxidation

Mauludi Ariesto Pamungkas

Nanomaterial Science and Technology UNIVERSITY OF SCIENCE AND TECHNOLOGY August 2012

Reactive Molecular Dynamic Studies of Initial Stage of Si(001) Surface Oxidation

Mauludi Ariesto Pamungkas A Dissertation Submitted in Partial Fulfillment of Requirements for the Degree of Doctor of Philosophy

July 2012 UNIVERSITY OF SCIENCE AND TECHNOLOGY Major of Nanomaterial Science and Technology Supervisor: Dr Kwang-Ryeol Lee

We hereby approve the Ph.D. thesis of Mauludi Ariesto Pamungkas


July 2012
Sign___________________ Dr. Byungki Cheong Sign___________________ Dr. Jae-Hyeok Shim Sign___________________ Prof . Ho-Seok Nam Sign___________________ Dr. Myoung-Woon Moon Sign___________________ Dr. Kwang Ryeol Lee

UNIVERSITY OF SCIENCE AND TECHNOLOGY

iii

ACKNOWLEDGEMENT
Verily! In the creation of the heavens and the earth, and in the

alternation of night and day, there are indeed signs for men of

understanding.[Quran chapter Ali Imran:190]

In the name of Allah, the Most Beneficent, the Most Merciful. All the praises and thanks to Allah, the Lord of the universe for his creating, blessings and guiding me. I express unlimited gratitude to my Father and Mother for always support and praying for me, to them I dedicate this Ph.D. thesis. Also thanks to my brothers and sister for their support and prayers. I am heartily thankful to my kind advisor, Dr.Kwang Ryeol Lee, for his guiding me patiently from the scratch in performing research and supporting over the course. In addition, I am also thankful for his concern in my personal living.

I would like to express my profound gratitude to my thesis committee, Dr.Byungki Cheong, Dr.Jae-Hyeok Shim, Prof.Ho-Seok Nam, and Dr.

iv

Myoung-Woon Moon for their detailed review, constructive criticism and excellent advice regarding my work.

My deepest gratitude to DR.Sang Phil Kim, my first teacher in molecular dynamic simulation as well as all my helpful labmates in our simulation group: DR.Min Wong Joe, Mr.Byung Hyun Kim, DR.Mina Park, Mr.Chansoo Kim, Mr.Haining Chao, as well as experimental group: Bong-Su Shin, TaeJoon Ko , and Seong Jin for their technical and scientific assistances as well as in their valuable suggestions. I would also like to thank Indonesian friend in KIST: Antun, Nova, Arens, Yudi Rahmawan, Purba Purnama, Adep Adid Dwiatmoko, Rusdi Syamsudin, Rizki, Iman Prayudi, Agus Ismail, Fathoni, Yodha, Iwa Nur Qodar, Warnadi, Muhammad Ridwan, Khairul Hudaya. To all the above individuals and to several colleagues whose names I cannot continue listing and who have assisted me on both my individual life and research works. The last but not the least, my love and gratitude to my wife and my son who have accompanied me through laughs and tears. Mauludi , Seoul , June 2012 v

Abstract
Even though silicon oxidation has been studied intensively since long time ago. Knowledge of the Si oxide layer evolution during real oxidation process is still far from well known. This thesis is aimed to provide more understanding on the initial state of oxidation of Si(001) especially what can not be obtained by previous theoretical studies. This study covers several different conditions of oxidation such as difference in temperature, shape of surface and oxidant.

Initial stage of oxidation of Si (100) surface by O2 molecules was investigated in atomic scale by molecular dynamics (MD) simulation at 300K and 1,200K without external constraint on the oxygen molecules. A reactive force field was used for the simulation to handle charge variation as well as breaking and forming of the chemical bonds associated with the oxidation reaction. Results of the present simulation are in good agreement with previous first principle calculations and experimental observations: the vi

oxygen molecules spontaneously dissociated on the Si (100) surface and reacted with Si first layer without energy barrier. The simulation also exhibited that the reacted oxygen preferentially located in the back bonds of the surface dimer. Consecutive oxidation simulation with 300 O2 molecules showed that a ballistic diffusion of oxygen atom occurs during very short time of the present oxidation simulation. The oxidation at 300K results in more stoichiometric oxide layer than that at 1,200K, because of the considerable amount of oxygen diffusions into the Si matrix even in the time scale of a few ns.

Investigation of Initial stage of oxidation of Si (100) surface by H2O molecules was also performed by molecular dynamics (MD) simulation at 300K and 1,200K without any external constraint on the water molecules. Results of the present simulation are in good agreement with the previous first principle calculations or experimental observations: most water molecules dissociate into hydroxyl oxygen (OH) and

vii

H groups and small number of water molecules adsorb in molecular form. Mechanisms of water reaction on silicon surface reported by previous
the studies, based on the density functional theory and experimental observations, were reproduced by this simulation. The present simulation further revealed that the hydrogen reaction with Si enhanced the dissociation of water. In contrast to the statistical analysis of the single water molecule reactions, oxidation reaction was enhanced at higher temperature with increased number of the supplied water molecules. The enhancement was related to water dissociation and hydroxyl decomposition at high temperature. It was suggested that the repulsion between water or hydroxyl molecules facilitated the dissociation of water molecule and hydroxyl decomposition on Si surface. Due to that mechanism, temperature dependence of early stage of wet oxidation is different from that of dry oxidation.

viii

Table of Content
Abstract ............................................................................................................vi

Table of Content ...............................................................................................ix

List of Figure ..................................................................................................xiv

List of Table ...................................................................................................xix

Introduction ...............................................................................................1

1.1

Background ...........................................................................1

1.2

Objective ...............................................................................3

1.3

Outlines of thesis ...................................................................3

Literature Review ......................................................................................5

2.1

Silicon System and Oxidation ...............................................5

2.1.1 Silicon ..............................................................................5

2.1.2 Silicon Oxide ....................................................................6

ix

2.1.3 Si(001) ..............................................................................6

2.1.4 Surface Steps and Vicinal Si(001)....................................7

2.1.5 Si/SiO2 interface ...............................................................8

2.1.6 Initial Oxidation of Si(001) ..............................................8

2.1.6.1 Initial dry oxidation of flat Si(001) ...........................8

2.1.6.2 Initial wet oxidation of flat Si(001) .........................10

2.1.6.3 Initial dry oxidation of vicinal Si(001) ....................13

2.2

Molecular dynamic ..............................................................14

2.2.1 Equation of Motion ........................................................15

2.2.2 Integrating the equations of motion................................17

2.2.3 Interatomic Potentials .....................................................18

2.2.4 Ensembles.......................................................................19

2.2.4.1 Microcanonical Ensemble .......................................19

2.2.4.2 Canonical Ensemble ................................................20 x

2.2.4.3 Isothermal and isobaric Ensemble ...........................20

2.2.4.4 Grand Canonical Ensemble .....................................21

2.2.5 Physical properties from microscopic ensemble ............21

2.2.5.1 Temperature ............................................................21

2.2.5.2 Pressure ...................................................................21

2.2.6 Structure .........................................................................22

2.2.6.1 Radial Distribution Function ...................................22

2.2.6.2 Bond Angle Distribution .........................................22

2.2.7 Simulating in Different Ensembles, Concept of

Thermostat.......................................................................23

2.2.7.1 Rescaling of Velocity .............................................24

2.2.7.2 Andersen thermostat ................................................25

2.2.7.3 Berendsen thermostat ..............................................26

2.2.8 Periodic Boundary Condition (PBC) ..............................27 xi

2.3

Potentials for Silicon and Silicon Oxide System .................27

2.3.1 Cluster Potential .............................................................28

2.3.2 Cluster functional ...........................................................28

2.3.3 Fixed and Variable charge Potential ..............................29

2.3.4 Reactive Force Field (ReaxFF) ......................................30

Computational Procedure ........................................................................33

3.1

Initial dry oxidation of flat Si(001) .....................................33

3.2

Initial wet oxidation of flat Si(001) .....................................38

3.3

Vicinal Si(001) ....................................................................41

Results and Discussion ............................................................................46

4.1

Dry Oxidation ......................................................................46

4.1.1 Single oxygen molecule reaction ...................................46

4.1.2 Statistical independent single oxygen molecule reaction

simulations .................................................................................50 xii

4.1.3 Consecutive multiple oxygen molecules reaction ..........52

4.2

Wet Oxidation .....................................................................59

4.2.1 Single water molecule reaction ......................................59

4.2.2 Statistical independent single water molecule reaction

simulations ......................................................................69

4.2.3 Consecutive multiple water molecules reaction .............71

4.3

Vicinal Surface ....................................................................85

Conclusion...............................................................................................91

References .......................................................................................................93

xiii

List of Figure
Figure 1. Radial Distribution Function of Amorphous silica. ....................34

Figure 2.

shows the simulation box for the present simulation ..................35

Figure 3.

Top view of surface with SA step ...............................................41

Figure 4.

Top view of surface with SB step ...............................................42

Figure 5.

Top view of Surface with DA step edge .....................................43

Figure 6.

Top view of DB ...........................................................................44

Figure 7.

Time evolution of single oxygen molecule reaction with different

initial oxygen molecule orientation and position on silicon surface. ..............47

Figure 8.

Time evolution of single oxygen molecule reaction when oxygen

molecule is placed on a channel of dimer rows at 300K with =45o. ...........49

Figure 9.

Top view of the oxidized surface by 300 oxygen molecules

consecutively added on the surface every 20 ps. ...........................................53

xiv

Figure 10.

Oxygen depth profile of the oxidized silicon surface after

consecutive oxidation reactions with 300 oxygen molecules. .......................55

Figure 11.

Time evolution of the charge state of silicon during

consecutive oxidation simulation at 300K .....................................................56

Figure 12.

Time evolution of the charge state of silicon during

consecutive oxidation simulation at 1,200K. ..................................................57

Figure 13.

Time evolution of single water molecule reaction when water

molecule, with two hydrogen atoms at the left side and oxygen atom at the right side of molecule at 300 K. ......................................................................60

Figure 14.

Time evolution of single water molecule reaction when water

molecule, with two hydrogen atoms on the top and oxygen atom at the bottom side of molecule at 300 K ................................................................................61

Figure 15.

Time evolution of single water molecule reaction when water

molecule, with two hydrogen atoms at the bottom and oxygen atom at the top side of molecule at 300 K ................................................................................62 xv

Figure 16.

Time evolution of single water molecule reaction when water

molecule on a silicon dimer atom with two hydrogen atoms at the left side and oxygen atom at the right side of molecule at 300 K. ................................63

Figure 17.

Time evolution of single water molecule reaction when water

molecule adsorbs molecularly at 300 K. .........................................................64

Figure 18.

Time evolution of number of adsorbed oxygen atoms in wet

oxidation and dry oxidation (inset). ................................................................72

Figure 19.

Time evolution of number of adsorbed hydrogen atoms in wet

oxidation.. ................................................................................................73

Figure 20.

Top view of the oxidized Si(001) surface by 150 water

molecule consecutively added on the surface. ................................................75

Figure 21.

Top view of the oxidized Si(001) surface by 600 water

molecule consecutively added on the surface. ................................................76

Figure 22.

Oxygen depth profile of the oxidized silicon surface after

consecutive oxidation reactions with 600 water molecules. .........................80 xvi

Figure 23.

Time evolution of (a) water molecule-water molecule reaction,

(b) water molecule-hydroxyl reaction, and (c) hydroxyl-hydroxyl reaction. ..82

Figure 24.

Time evolution of the charge state of silicon during

consecutive simulation at (a) 300 K ................................................................83

Figure 25.

Time evolution of the charge state of silicon during

consecutive simulation at 1200K. ...................................................................84

Figure 26.

Single oxygen molecule reaction on silicon surface with SA

step edge. .........................................................................................86

Figure 27.

Single oxygen molecule reaction on silicon surface with SB

step edge.. ................................................................................................87

Figure 28.

Single oxygen molecule reaction on silicon surface with DA

step edge.. ................................................................................................89

Figure 29.

Single oxygen molecule reaction on silicon surfacewith DB

step edge.. ................................................................................................90

xvii

xviii

List of Table

Table 1. Structural properties of amorphous silica calculated by molecular dynamics using ReaxFF and experimentally observed data............................37

Table 2. Properties of crystalline Si calculated by molecular statics using ReaxFF, calculated by ab initio density functional theory and experimentally observed data ...................................................................................................38

Table 3. Total uptake fraction of oxygen and the depth distribution of the oxygen obtained by statistical analysis of 1,000 single O2 reaction events. ....50

Table 4 Total uptake fraction of oxygen and the depth distribution of the oxygen obtained by statistical analysis of 1,000 single H2O molecule reaction events...............................................................................................................69

Table 5. Total uptake fraction of Hydrogen and the depth distribution of the oxygen obtained by statistical analysis of 1,000 single H2O molecule reaction events...............................................................................................................70

xix

Introduction

1.1 Background
Undoubtedly, it is the human nature that people are never satisfied with what they have. People always want to have higher electronic device performance with lower price. In order to obtain high performance, electronic devices must be able to accommodate huge number of electronic components, which means that size of electronic components have to be very small. Nowadays, we witness fantastic achievement in miniaturization of transistor as it has been predicted by Gordon More, who is one of founders of Intel Corporation, that number of transistor can be hold by an integrated circuit will be double about every two years (Moor, 1965). A traditional metaloxidesemiconductor (MOS) transistor is obtained by growing a layer of silicon dioxide (SiO2) on top of a silicon substrate by thermal oxidation process and depositing a layer of metal. As the silicon dioxide is a dielectric material, its structure is equivalent to a planar capacitor, with one of the electrodes replaced by a semiconductor. Due to the continued downscaling of the channel length, and hence reduce the gate area, the need of maintaining sufficient capacitance of the MOS gate stack was met by gradual decrease of the thickness of SiO 2 gate dielectric. Serious problems then appear when the thickness reach to very small scale in which quantum effects, such as high quantum mechanical 1

tunneling current through the gate oxide which lead to increase power consumption, become significant. Therefore, atomic knowledge of SiO2 formation or silicon oxidation process becomes vital. Even though there are intensive studies focus on silicon oxidation process, many aspects are far from well known. Initial oxidation of silicon surface attracted special interest since it is closely related to Si/SiOx interface formation which play essential role on determining properties of the gate. In addition, the formation of SiOx region for the moment unavoidable to be used even though SiO2 will be replaced by high k dielectric material(Fischer, Curioni, Billeter, & Andreoni, 2006).

Molecular dynamic simulation is a powerful theoretical method that is now routinely used to simulate the dynamics of complex physical and chemical systems. Force field or interatomic potential plays paramount important in performance of MD simulation. Reactive Force Field currently appears as influential potential widely used in various dynamical process of many elements and alloy systems due to its capability in describing charge transfer as well as forming and breaking of atomic bonding. In general, atomic simulations play role as guidance for experimental work since by atomic simulation we can work in special condition which is impractical to perform by experiment. Classical Molecular dynamic simulation also has some advantages compared to statistical ab initio method.

It can describe dynamic of the system. While its advantage compared to ab initio MD is its ability to work with huge system consist of thousands even million atoms. In case of silicon oxidation for example, by general classical molecular dynamic one can explore interaction of many oxygen molecule not only with top surface silicon but can also gain information of reactions in deep silicon layers which can not be performed by an ab initio study.

1.2 Objective
The aim of this thesis is to provide more understanding on the initial state of oxidation of Si(001). This study covers several different conditions of oxidation such as difference in temperature, shape of surface and oxidant. Particularly, the study is aimed to complement what have not been or can not be obtained by experimental works and other atomic simulation methods such as ab initio method and monte carlo method.

1.3 Outlines of thesis


This Ph.D. thesis is organized as follows: Chapter 1 contains introduction which consist of : the background, objective and outline of this thesis In Chapter 2, Basic information about silicon, silicon oxide, Si/SiOx interface, silicon oxidation process and molecular dynamic method are reviewed briefly.

Chapter 3 Describe computational method used in flat Si(001) dry oxidation simulation, flat Si(001) wet oxidation, and vicinal surface Si(001) simulation. It includes methods of force field validation. Chapter 4 discusses results of the Si(001) oxidation simulations. Chapter 5 summarizes and concludes the main finding of this work

2 Literature Review
2.1 Silicon System and Oxidation

2.1.1 Silicon
Since silicon exist plenty in nature (the second most abundant element after oxygen In Earth's crust, making up 25.7 % of the crust by mass ), remains a semiconductor at higher temperatures than the semiconductor germanium and because its native oxide (SiO2) is easily grown in a furnace and forms a good semiconductor/dielectric interface, up to now silicon and its oxide play ultimate role in various applications in microelectronics industry. Silicon has atomic radius of 1.46 A, covalent radius of 1.11 A and ionic radius of 0.4 A. It has energy gap of 1.12 eV and dielectric constant of 11.9 at 300 K and 1.2 at 0 K. Silicon has four valence electrons which form covalent bonds and form diamond crystal structure. It can be used as the intrinsic semiconductor. Substituted atoms (dopant) can dramatically change its electrical properties. In periodic table, it is classified into a group IV because of its electron valence with electronic configuration 1s2 2s2 2p6 3s2 3p2. It can be doped by

element from group III (acceptor) and V (donor). Cohesive energy of silicon is Ec=4.63 eV. While, its band gap is 1.1 eV. (Jarek Dabrowski, 2000)

2.1.2 Silicon Oxide


In nature, silicon oxide can be silicon dioxide (SiO2) which is well characterized and silicon monoxide (SiO) which is not well characterized. Gaseous SiO is produced when SiO2 is heated with silicon. SiO also can be in form of glassy brown/black amorphous solid. While SiO2, exist in amorphous as well as in many polycrystalline forms (polymorphs). The crystal structure of these polymorphs are: -quartz, -quartz, -tridymite, -cristobalite, cristobalite, keatite, coesite, and stishovite. Stishovite has local structure of sixfould distorted octahedral bonding. While, other polymorphs have fourfold tetraderal bonding for Si and two fold bridging bonding for O. SiO2 has band gap of about 9 eV and dielectric constant of 3.9. (Y. P. Li and W. Y.Ching, 1985),(Xu, 1991)

2.1.3 Si(001)
Si(001) is most used of silicon surface because it is simplest one. It is built by truncation of the bulk by a {001} crystallographic plane. The unreconstructed Si(001) has two dangling orbitals per surface atoms, each of them is filled with one electron. By reconstruction, the surface atoms form dimmers to reduce the energy of the surface. The pairs of dimers atoms bonded to each other, with each left with one dangling bond which is created

when a covalent bond is broken. The dangling bond can be passivated by hydrogen atoms. (Tao, Shanmugam, Coviello, & Kirk, 2004)

2.1.4 Surface Steps and Vicinal Si(001)


If silicon crystal is cut at an angle to one of its high symmetry planes, the resulting surface may consist of terrace of the high symmetry plane separated by a single or double atom high. Since two adjacent silicon subsurfaces have different atomic order, then there are two types of terraces and steps which are classified according to the accepted nomenclature introduced by Jim Chadi (Chadi, 1987). In case dimers on terrace are perpendicular to the step edge, the terrace is called A and it is classified as type B if dimmers on the terrace are parallel to the step edge. Step is categorized into single atom step (S) and double atom step (D). A single atom step with upper terrace type A and lower terrace type B is called step SA. On the contrary, A single atom step with upper terrace type B and lower terrace type A is called step SB. Experimental works at low temperature and ab initio studies have shown that step edge is more reactive than the terrace. (Jarek Dabrowski, 2000)

2.1.5 Si/SiO2 interface


It is well known that transition layer consist of intermediate oxidation states of Si , so called suboxides, which include Si+, Si2+, and Si3+ (Y. Takata, K. Tamasaku, T. Tokushima, D. Miwa, S. Shin, T. Ishikawa, M. Yabashi, K. Kobayashi, J. J. Kim, T. Yao, T. Yamamoto, M. Arita, H. Namatame, 2004). There are different experimental results conclusions about amount of intermediate states. Using the angle-resolved Si2p photoelectron spectra it is reported that the intermediate oxidation states are distributed over two separated layers. (Oh et al., 2001), (Hirose, Nohira, Azuma, & Hattori, 2007)

2.1.6 Initial Oxidation of Si(001)


In this part, I want to summarize fundamental aspects of initial silicon oxidation reported by many previous experimental and theoretical studies. It includes interaction of one oxygen molecule with silicon surface and features of oxidized silicon surface by oxygen molecules deposition. There are several issues appeared as a long controversy such as possible pathway of oxygen molecule adsorb on Si surface. 2.1.6.1 Initial dry oxidation of flat Si(001) A great deal of experimental investigation has been performed to resolve the oxidation state of the Si atoms in molecular level by means of various experimental techniques such as surface differential reflectance (SDR),(Yasuda et al., 2003), (Ohno, Takizawa, Koizumi, Shudo, & Tanaka,

2007a) x-ray or ultraviolet photoemission spectroscopy (XPS/UPS),(Enta et al., 2008), (Suemitsu, Enta, Miyanishi, & Miyamoto, 1999), (H. W. Yeom, Hamamatsu, Ohta, & Uhrberg, 1999), (Oh et al., 2001) scanning reflection electron microscopy (SREM),(H. Watanabe et al., 1998) in situ Auger electron spectroscopy (AES) combined with reflection high energy electron diffraction (RHEED),(Takakuwa, Ishida, & Kawawa, 2003) high resolution Rutherford backscattering spectrometry (HRBS),(Nakajima, Okazaki, & Kimura, 2001) or scanning tunneling microscope (STM).(Engel, 1993), (Chung, Yeom, Yu, & Lyo, 2006), (Itoh, Nakamura, Kurokawa, & Ichimura, 2001). The experimental observations exhibited complex behaviors of the initial oxidation phenomena, and the atomic scale computation has been expected to provide deeper insight of the reaction. Many theoretical calculations using density functional theory reported the oxygen-silicon interactions and revealed the most probable oxygen adsorption path by comparing the energy barriers of several possible paths.(Miyamoto, Yoshiyuki, 1990), (Miyamoto, Yoshiyuki, 1991), (K. Kato, Uda, & Terakura, 1998a), (K. Kato & Uda, 2000), (Fan, Zhang, Lau, & Liu, 2005), (Hemeryck, Estve, Richard, Djafari Rouhani, & Chabal, 2009), (Pasquarello, Alfredo, Mark S Hybertsen, 1998) Direct oxidation simulation using quantum molecular dynamics was also tried to understand kinetic pathway for Si oxidation.(Pasquarello, Alfredo, Mark S Hybertsen, 1998), (Bongiorno & Pasquarello, 2004), (Ciacchi & Payne, 2005a), (T Uchiyama & Tsukada,

1996), (Toshihiro Uchiyama & Tsukada, 1997). Ab initio or classical molecular dynamics (MD) simulations were employed to find the structural model of Si/SiOx interface using an artificially built interface as the starting materials or inserting O in Si-Si network.(Fischer et al., 2006), (Sarnthein, Pasquarello, & Car, 1995), (T Watanabe, 2004), (Ganster, Trglia, Lancon, & Pochet, 2010), (Bongiorno & Pasquarello, 2003), (Demkov & Sankey, 1999), (R.M. Van Ginhoven , 2007) However, most ab initio MD simulations were limited in both time scale and the size of the calculation system, which had to adopt a constrained MD condition. The classical MD simulations were also performed under the constrained condition where the oxygen atoms were impinged into Si-Si bond at the interface followed by relaxation. (T Watanabe, 2004), (Ganster et al., 2010), (Ohta, Watanabe, & Ohdomari, 2008), (Ohta, Watanabe, & Ohdomari, 2007) Although the previous theoretical works

enabled the atomistic understanding of the structures of the oxide layer itself or its interface with Si, it is still far from understanding the Si oxide layer evolution during real oxidation process. 2.1.6.2 Initial wet oxidation of flat Si(001)

In the past three decades, many research works devoted in water molecule adsorption process on silicon surface. After long controversy on the question of whether H2O adsorbs molecularly or dissociatively, it is now accepted that water molecule is mostly adsorbed on Si(001)

10

dissociatively by saturating the dangling bonds of the Si dimmer with hydroxyl (OH) and hydrogen species to form SiH and SiOH mechanism,(Niwano, 1998) even though theoretically there is a probability of water to adsorb molecularly with smaller

probability(Cho, Kim, Lee, & Kang, 2000a). Using Scanning Tunneling Microscopy, M.Chader et.al has reported that nondissociative adsorption is present at very low coverage, while dissociative adsorption is present near saturation coverage(M.Chander, Y.Z.Li , J.C.Patrin, 1993). In theoretical studies, static ab initio calculation is frequently used to study interaction of single water molecule with silicon clusters (Stefanov & Raghavachari, 1998), (Weldon, Queeney, Gurevich, Stefanov, & Chabal, 2000),(Warschkow et al., 2008). Those studies tried to provide most possible pathways for initial water-induced silicon oxidation by comparing energy of several possible pathways. Even though their finding have clarified many aspects of adsorption process of water molecules into silicon surface, but due to its limited time scale, ab initio studies still far from observing dynamics of oxygen atoms, hydrogen atom and hydroxyl groups during oxidation process. In addition, due to small scale limit, most previous ab initio calculations
11

only model single water molecule-silicon surface interaction, whereas in fact, many water molecules are deposited in wet oxidation process. Neighbor water molecules and their constituent elements and groups should make more complex interactions. For instance, Hiroyuki S Kato et.al reported hydrophobic interaction between water molecules and hydrogen groups as well as hydrophilic interaction between water molecule and hydroxyl groups (H. Kato, Kawai, Akagi, & Tsuneyuki, 2005). While many theoretical studies have been done on single water molecule interaction, no theoretical study devoted to explore oxidized silicon surface due to wet oxidation. Molecular dynamic is actually capable of handling much larger system, but conventional MD simulations can not describe chemical reaction. That is probably the reason why there is no classical MD simulation for wet oxidation of Si(001) has been reported so far. An ab initio MD(Colombi Ciacchi, Cole, Payne, & Gumbsch, 2008) has been performed to study presence of surface tensile stress which accompany thin oxide layer formation, yet it does not present atomic configuration of oxidized silicon surface structure because of small size of the system

12

2.1.6.3 Initial dry oxidation of vicinal Si(001) Although oxidation of silicon surface has been studied very intensively for a long time owing to ultimate role of silicon and its oxide in various applications in microelectronics industry, there are only limited studies focus on effects of complexities (e.g steps edges, kinks, vacancies) of the silicon surface on its oxidation process. For Si(001) surface, where different type of step edges (SA, SB, DA, and DB step edges as labeled by Chadi (Chadi, 1987) exist, there are very few studies which have shown interplay between step edge type and oxidation processes. A study, using Scanning Tunneling Microscopy and atomic force microscopy, shows enhancement of reactive sticking coefficient and oxide cluster nucleation due to high step density(Brichzin & Pelz, 1999). Later study using Scanning Tunneling Microscopy combined with ab initio calculation shows tendency of oxygen molecules to adsorb on SB edge over terraces and SA edge, even there is no oxidation observed in SA edge (Chung et al., 2006). It is believed that the preference of oxygen molecule to adsorb on SB step edge is attributed to the need of stabilization of dangling bonds and weak bonds in rebounded SB step edge (Yeo et al., 2007). Previously, an Ab Initio study of Si adatom motion has predicted that SB is a strong sink for adatoms due to strong binding while SA step is a weak sink because of weak binding and low barrier for Si adatoms to escape(Zhang, Roland,

Boguslawski, & Bernholc, 1995). 13

2.2 Molecular dynamic


Nowadays, computational simulation has been appeared as a one important method of research work. Position of computational simulation method can be considered in between experiment and analytical theory. One can regards computational simulation as theoretical experiment. On one side, it is similar to analytical theory since they are dealing with models not with the real thing''. On the other side, the procedure of verifying a model by computer simulation resembles an experiment. In both experiment and simulation, one needs to prepare sample, to give certain treatment and characterize the sample. (J.M.Haile, 1997) Since a seminal paper published in 1957 written by professor Alder (E.Wainwright, 1957)and followed by paper written by professor Aneesur Rahman (A.Rahman, 1964) who known as the father of molecular

dynamics (because his algoritm is still the basis for many codes written today), molecular dynamic simulation has been appearing as a powerful computational method to describe dynamics of a system in atomic and molecule scale due to its deterministic and simple basic theory and capability to handle large system. The main advantage of MD is its ability to produce true microscopic dynamics. In principle, with MD simulation we can extract thermodynamic information that physically possible. However, in practice it

14

will be limited by such things as relaxation time scale and the power of available computers. Implementation of MD involves the following stages: 1. Setting the parameters describing the conditions of MD simulations, such as temperature, number of atom, coordinate etc. 2. Initialization, which includes reading in the coordinates of atoms and generation of initial distribution of velocities. 3. Computation of forces 4. Integration of Newtons equation of motion 5. Repeating steps 3 and 4 until MD simulations are done for the desired timescale 6. Computation of averages

2.2.1 Equation of Motion


In principle, molecular dynamic based on Newton law related to statistical mechanic. The time evolution of a set of interacting atom is followed by integrating their equation of motion so that one can obtain complete trajectory of the atom. Started from second law of Newton Law (1) Where acceleration,

15

By integration of a with respect to time one can obtain velocity. Once more integration will result position. Velocity of atoms can be related to thermodynamic parameters through statistical mechanics. For complex motions which involve constrain, equation of motion is better written in more general form so called Lagrange equation : (

(2)

Where, L is Lagrangian. In case qi denotes a component of the Cartesian coordinates for one of the atoms, then L can be written as { }

(3)

If we substitute into equation (2), it becomes

(4)

it is nothing but Newtons second law. In case, a constrain is applied in system with general form: (5)

The resulting Lagrange equation is: (

(6)

16

Equation of motion also can be described by Hamiltonian if we change velocity in Lagrangian with generalized momentum then . Hamiltonian is defined by: Thus, the two first order equation of motion: (7) ,

(8)

+U(q)

(9)

2.2.2 Integrating the equations of motion


According to second Newtons law, if force is known, one can obtain acceleration a, and by integration of acceleration a with respect to time one can obtain velocity. Once more integration will result position. Integration process numerically can be performed by several methods. For example, by Verlet integrator:

(10)

+ .

(11)

And by leapfrog algorithm : 17

(12) ( )

(13)

with generalized equations of motion is: [ ( ) ]

(14)

(15) { [ ( ]} ) ] [ ( ) (16)

2.2.3 Interatomic Potentials


Atoms in gas, liquid, and solid interacts each other. The simplest interatomic potential model is Lennard Jones potential which describes two different type of interactions, repulsive interaction which is dominant in very short distance to avoid overlaps between two electron clouds ( in quantum terminology known as Pauli repulsion) and attractive interaction which is dominant in longer distance due to van der Waals force or dispersion force. The potential energy is described as:

[( )

( ) ]

(17)

18

The force corresponding to u(r) is: (18) Last section has explained how equation of motion derived from force. Thus, accurate form and parameters of mathematical function used to describe potential energy of the system (force field) play vital role in molecular dynamic simulation.

2.2.4 Ensembles
An ensemble is a collection of all possible system which has different microscopic states but has an identical macroscopic or thermodynamic state. In MD simulation, one can simulate several conditions according to which thermodynamic parameter is assumed as a fix parameter or variable one. 2.2.4.1 Microcanonical Ensemble Microcanonical ensemble is an ensemble for isolated system. In this ensemble, number of particle (N), volume (V), and energy are fixed, so it is also called NVE ensemble. The appropriate thermodynamic potential of this ensemble is internal energy. Thermodynamic potential is a function where all the thermodynamic properties of a system can be obtained by taking its partial derivatives with respect to its natural variables.

19

When the entropy and other natural variables (V, N) of a isolated system are fixed, the internal energy will decrease and reach a minimum state at equilibrium. 2.2.4.2 Canonical Ensemble The canonical ensemble describes the statistical distribution in a system that can exchange energy via very weak contact with a temperature bath and eventually comes to equilibrium. Thus, number of particle (N), volume (V), and temperature are reserved and this ensemble then called NVT ensemble. The appropriate potential thermodynamic is the Helmholtz free energy F.

(19)

If the temperature T and external parameters of a closed system are held constant, the Helmholtz free energy decrease and reach minimum values at equilibrium. 2.2.4.3 Isothermal and isobaric Ensemble This is a statistical ensemble where temperature T and pressure P are maintained constant. It is also called NPT ensemble. Its thermodynamic potential is Gibbs free energy.

20

2.2.4.4 Grand Canonical Ensemble In this ensemble, volume and temperature are constant (as the canonical ensemble), but is open for exchanging particles with a surrounding bath. In this case, the chemical potential of the different species has a specified average, while the instantaneous value of N the number of particles can fluctuate.

2.2.5 Physical properties from microscopic ensemble


2.2.5.1 Temperature In mechanical statistic, temperature is related to kinetic energy of atom.

(19)

Since this is only an instantaneous temperature, it will change continuously as time goes by. Accordingly, the meaningful value is temperature average: 2.2.5.2 Pressure The pressure is another thermodynamic magnitude which can be obtained by the virial theorem. The instantaneous is given by:

(20)

21

i)

(21)

Where, N, V, and T are the number of particle, volume and temperature respectively. is the force of atom i due to atom j, and is

vector between atom i and atom j. Like the case of temperature, only pressure average is meaningful.

2.2.6 Structure
Structure of system of interest is also important information one can obtain from MD simulation. Two quantities which are frequently used to describe the structure are radial distribution function and bond angle distribution function. 2.2.6.1 Radial Distribution Function Radial distribution function (RDF) or pair correlation function is a histogram of the distance between atom and its neighbors which describe how the atomic density varies as a function of the distance from one particular atom. The average coordination of atoms (the number of first neighbors) of each atom can be obtained by volume integral of the RDF from zero to the first minimum 2.2.6.2 Bond Angle Distribution Bond angle distribution is distribution of bond- angle between atoms involved in the system. It is also important because it can provide angular 22

information which can not be described by RDF. It gives complete picture of short range environment of atoms.

2.2.7 Simulating in Different Ensembles, Concept of Thermostat


In isolated system, all forces which appear in the Newton equation are related to the potential energy of the system, then the total energy of the system E = Ekin+Epot is conserved. In case the total number of atoms N and volume V are also kept constant, then The MD simulation is categorized to be performed in microcanonical ensemble (NVE). In this ensemble, temperature and pressure are variable. However, most conditions where experiments are carried out at certain temperature or certain pressure. Thus, performing a MD simulation in an ensemble other than microcanonical ensemble requires a means to keep at least one intensive quantity constant during simulation. In molecular dynamic, thermostat means a modification of the Newtonian MD scheme to generate a thermodynamic ensemble at constant temperature. There are several methods to mimic principle of thermostat in experiment. In experiment, a thermostat is a component of a control system which senses the temperature of a system so that the system's temperature is maintained near a desired setpoint.. It is done by switching heating or cooling devices on or off, or regulating the flow of a heat transfer fluid as needed, to maintain the correct temperature. 23

2.2.7.1 Rescaling of Velocity This method is the simplest way to keep temperature constant. The velocity is taken from the Maxwell Boltzman distribution.

(22)

Where, vi, is the (= x, y, z) component of the velocity of the atom i. From equation (22) that average kinetic energy of freedom is related to temperature T as:

(23)

Where, <> bracket indicates the ensemble average. Since the ensemble average correspond to the average over velocities of all atoms, then for the finite size system one may define an instantaneous temperature T(t)

(24)

Where Nf is the number of degrees of freedom. T(t) does not coincide with T used to generate velocity distribution in equation (22). It fluctuates between successive generations of random velocities. To keep the temperature T(t) =T , we can rescale velocities as

(25)

24

2.2.7.2 Andersen thermostat Equation of motion of the N particles in volume V is the Hamiltonian equation with

(26)

(27)

(28)

The coupling to a heat bath is represented by stochastic impulsive forces that act occasionally on randomly selected particles. Each atom at each integration step is subject to small probability to collide with heat bath. Between stochastic collisions (with frequency v), the system evolves according to Hamiltonian in equation (26). Each stochastic collision is an instantaneous event that affects the momentum of one particle. The value of momentum of a particle i, which is suffered from collision, is chosen from a Boltzmann distribution at temperature T. Each particle undergoes a stochastic collision with probability vt after every MD step. Successive stochastic collision is the Poisson form P(t,v)=v exp (-vt) (29)

25

2.2.7.3 Berendsen thermostat In this thermostat, system is weakly coupled with thermal bath. Such coupling can be accomplished by inserting friction terms in the equation of motion. The equation of motion for the Andersen thermostat can formally be written as: ( )

(30)

That equation of motion represents a proportional scaling of the velocities per time step in the algorithm from v to v. Where is damping constant which determine strength of coupling to heat bath, ( )

(31)

[ (

(32)

[ (

(33)

(34)

An atom at random is selected and its velocity then changed to a new value which is selected from the MaxwellBoltzman distribution

26

corresponding to the desired temperature of the simulation(Berendsen, Postma, van Gunsteren, DiNola, & Haak, 1984)

2.2.8

Periodic Boundary Condition (PBC)


Molecular Dynamics is applied to system containing thousands of

atoms. Such small system is dominated by surface effect, i.e interaction of the atoms with the container wall. While in the reality number of atoms near the surface can be neglected compare to the number of atoms contained in a macroscopic piece of matter (of the order of 1023). Therefore idea of periodic boundary condition is used where the volume V is only a small portion of bulk material. The volume V is called primary cell. It is representative of bulk material by mirror it to its neighbors over entire bulk material so that the primary cell can be periodically replicated in all direction to form a macroscopic sample. This periodicity extends to the position and momenta of the image in image cells. Due to short interaction range of potential, to apply periodic boundary condition, there is a minimum image criteria should be fulfilled. Distance between two adjacent boxes should be larger than 2 RC, where RC is a cutoff distance, to limit interaction of the same atoms in two adjacent boxes.

2.3 Potentials for Silicon and Silicon Oxide System


There were no less than 30 type of potential model (force field) proposed for silicon and its oxide. According to how pair interaction and 27

many body interaction described in force field, potentials for silicon and silicon oxide classified into two groups: cluster potentials and cluster functionals (Ercolessi, 1997). It also can be classified based on its ability to describe charge transfer. We can name fixed charge potential groups and variable charge potential groups.

2.3.1 Cluster Potential


Pioneer of this potential group is a potential developed by F.Stillinger and T.A.Weber which usually called SW potential for short (Stillinger & Weber, 1985). This potential consists of two parts, two body interaction term and three body interaction term. ( [ ( ) ) ] (35) (36) (37)

It is one of most used potential, either its original potential or extended or modified ones.

2.3.2 Cluster functional


This group of potential is pioneered by tersoff potential (Tersoff, 1988). This potential is developed based on bond order (number of chemical bonds between a pair of atoms) concept. The strength of bond is not constant

28

rather depends on the local environment. Bonding is modeled as interaction with attractive term depending on the local environment which include many body interaction. At first glance, tersoff potential looks like pair potential. ( )

(38)

Actually it is not pair potential since Bij is a function of bond order Gij. ( ) ( ) (39)

The bond ij is weakened by the presence of other ik bonds involving atom i.

2.3.3

Fixed and Variable charge Potential


Charge in first generation of potentials for silica system is kept

constant. We notice the first variable charge potential is a potential developed by Z.Jiang and R.A.Brown (Jiang & Brown, 1995). The potential is

combination and extension of Stillinger-Weber (SW) silicon 0potential and Van Beest, Kramer and Van Santen (BKS) silica potential. They include three new components: a charge transfer function, a bond softening function and ionization energy. Several other potentials then developed such as: Composite interatomic potential, extended SW (Takanobu Watanabe &

29

Ohdomari, 1999), ReaxFF (A. C. T. van Duin, Dasgupta, Lorant, & Goddard, 2001), ab initio derived augmented Tersoff (Billeter, Curioni, Fischer, & Andreoni, 2006), and charge optimized many body (COMB) (J. Yu, Sinnott, & Phillpot, 2007).

2.3.4

Reactive Force Field (ReaxFF)


This potential is designed as a bridge of empirical force field (EEF)

and quantum chemical (QC) methods. As it is widely known, QC methods work in angstrom scale, can provide electronic structure which make it applicable to all chemical systems. However they are still impractical for large system even with state of the art supercomputer because of their computational expense. On the other hand, EEF methods because of their relatively simple form can be applied to much larger system consist of millions atoms but in general can not describe reactive system. Reactive force field (ReaxFF) is built to combine capabilities of EEF and QC and minimize their limitation. Off course as a bridge its position is in between. ReaxFF can be applicable in many chemical systems with its flexible connectivity but its computational cost is more than traditional empirical force field scheme. ReaxFF is a general bond-order-dependent force field as first formulated by Tersoff. This concept was used by Brenner to construct the REBO-potential, but its transferability is limited because it is based on 30

relatively small training set and because it excludes all non-bonded interactions. In ReaxFF, bond order is calculated from interatomic distances that are updated every MD step which allows connectivity changes.

( )

( )

( )

] (40)

Where Pbo,1 and Pbo,2 are sigma bond, Pbo,3 and Pbo,4 are first pi bond, Pbo5, Pbo,6 are second pi bond respectively . All other covalent interactions such as functions describing bond angle and torsion angle, Eval and Etors are include these bond orders so that all terms dissociate properly as any bond is broken. While Bond energy is related to bond order by [ ( )

(41)

Total energy of system consists of various partial energy contributions including bonded and non-bonded interactions (van der waals, coulomb). ReaxFF describes non-bonded interactions between all atoms, irrespective of connectivity. The energy system is described as: Esystem = Ebond + Eover + Eunder + Eval + Epen + Etors + Econj + EvdWaals + ECoulomb 31 (42)

(43)

(44)

(45)

The charge on an atom depends on the molecular species. Atomic charges are adjusted with respect to connectivity and geometry. Many charge equalization methods available. ReaxFF uses Electron Equalization Method (EEM). EEM parameters were optimized against Mulliken charge distribution obtained from DFT calculations. The desired charge distribution is that which minimizes Coulomb energy from screened potential-all atom pairs calculated

(46)
[ ]

ReaxFF provides a highly transferrable method for atomistic scale simulations of complicated chemistry. It has been successfully parameterized to reproduce quantum mechanics (QM) data for wide range of system. (A. C. T. van Duin et al., 2001), (A. C. T. V. Duin et al., 2003)

32

3 Computational Procedure

3.1 Initial dry oxidation of flat Si(001)


Surface oxidation of Si needs an interatomic potential that can provide consistent description appropriate for covalent system of Si and all forms of ionic silicon oxide systems. The potential function should also correctly handle the charge distribution (or reaction) between oxygen and silicon. We used the Reactive Force Field (ReaxFF) for Si-O system, which would allow accurate simulations of the bond breaking during oxidation and the interface evolution between silica and silicon. Large-scale

Atomic/Molecular Massively Parallalized Simulator (LAMMPs) code integrated with ReaxFF was used for the present simulations.(Plimpton, 1995) The ReaxFF integrated in LAMMPS was benchmarked with amorphous silica, since the oxidation of Si surface results in an amorphous SiOx layer. Amorphous SiO2 was constructed by initially heating an -quartz up to 6,000K in NPT ensemble until the system was completely melted, and quenching the melt to room temperature within 1 ps. Figure 1 shows the partial radial distribution function (p-RDF) of the amorphous silica and Table 1 the density and atomic bond structure obtained from the amorphous SiO2. Partial RDF spectrum is almost identical to that obtained by Fogarty et al using another ReaxFF implementation (SERIALREAX).(Fogarty, Aktulga, Grama, van Duin, & Pandit, 2010) (See Fig. 7 in Ref (Fogarty et al., 2010)..) 33

As shown in Table 1, all the calculated structural characteristics of amorphous SiO2 are in good agreement with the experimental measurements.

Figure 1. Radial Distribution Function of Amorphous silica.

Slab model of single crystal Si of (100) surface was used as the substrate after relaxing at 300K for 10 ps. Figure 2 (b) shows the topview of the Si surface which exhibits the well-known surface dimers along <110> direction as marked by white lines. Reasonable agreement between calculated and observed properties of crystalline Si was also obtained as summarized in Table 2.

34

Figure 2. shows the simulation box for the present simulation

35

In order to mimic a thick substrate, Si atoms in the bottom 0.1 nm region of the simulation box was fixed. Atoms in the region of about 6.6 nm above the fixed layer were kept at a constant oxidation temperature (300K or 1200K) during simulation to provide a thermal bath in the system. Temperature of the thermal bath was rescaled at every time step (1fs) during simulations. All the atoms above the thermal bath were unconstrained with an initial temperature set to the oxidation temperature. Oxidation simulation was done by placing an O2 molecule presented by blue balls (indicated by an arrow in Fig. 2(a) and (b)) at the distance 0.12 nm above the surface, without any other external constraint. This distance was chosen from the preliminary simulations to find the largest distance for the O2 molecule to react with Si surface. Two simulation sets were carried out. In order to compare the oxidation behavior with the previous ab initio calculations or the experimental results of surface reaction, single O2 molecule reactions were investigated on randomly selected surface positions with various oxygen molecule configurations for 10 ps. 1,000 independent single molecular reactions were statistically analyzed. Secondly, we oxidized the surface by sequentially adding 300 O2 molecules on the surface to study the oxide surface layer evolution.

36

Table 1. Structural properties of amorphous silica calculated by molecular dynamics using ReaxFF and experimentally observed data ReaxFF MD Density (g/cm3) O-Si-O angle () Si-O-Si angle () Si-Si distance (nm) O-O distance (nm) Si-O distance (nm) a (Warren, 1969) b (David I.Grimley, Adrian C. Wright, Sinclair, 1990) 2.25 109.36 142 0.310 0.02 0.270 0.04 0.159 0.006 Experiment 2.20a 109.4a 144a 0.3077, 0.312 a 0.2626 b, 0.265a 0.1608 b, 0.1620a

37

Table 2. Properties of crystalline Si calculated by molecular statics using ReaxFF, calculated by ab initio density functional theory and experimentally observed data ReaxFF Silicon MD Lattice Constant (nm) Bulk Modulus (Mbar) Cohesive nergy(eV/atom) 0.532 1.12 4.56 0.545 0.98 4.84 0.543 0.99 4.63 AB Initioc Expc

c (Yin & Cohen, 1982) and references therein

3.2 Initial wet oxidation of flat Si(001)


In this report, molecular dynamic simulations were performed to simulate very initial stage of oxidation of Si(001) by water molecules. The interatomic potential used is reactive force field (ReaxFF) developed by Van Duin et.al,(A. C. T. V. Duin et al., 2003) which is integrated in Large-scale Atomic/Molecular Massively Parallalized Simulator (LAMMPS) code(Plimpton, 1995). ReaxFF was selected from many available potential for Si and SiO2 owing to its capability to describe chemical reaction (forming and breaking of bonding as well as charge transfer between constituent atoms) which is main part of oxidation process. The benchmark test of ReaxFF integrated in
38

LAMMPS has been presented in our previous report(Pamungkas, Joe, Kim, & Lee, 2011a). We performed MD simulations under constant temperature, 300 K and 1200 K, to study effect of system temperature on the oxidation process. The simulation box was built in diamond lattice structure with size of 10 a0 x10 a0 x 18 a0, where a0 is the lattice constant of crystalline silicon which was set initially to 5.43 . Some part of box (3 x a0 thick) was cut in <001> direction to create Si(001) surface and deposition area. The total number of silicon atoms was 24000. Periodic boundary condition was applied laterally. A periodic boundary condition was applied in the lateral directions. The atomic position of the bottommost layer was fixed to simulate a thick substrate. Temperature of the bottom layers (33.123 thick) was fixed at 300 K and 1200 K so that the layers acted as a thermal bath during deposition. Temperature of the thermal bath was rescaled at every time step (1fs) during simulation. All the other atoms above the thermal bath were unconstrained. The substrate was equilibrated with a thermal bath for 20 ps by the MD relaxation of the structure. Arrays of silicon dimers are developed as a result of surface reconstruction with Si-Si dimer bond length in the range of 2.28 -2.48 . That range of values encompasses diverse
39

values reported by DFT calculation,( one n XRD(B.W.Holland, C.B.Duke, 1984) and

& Doren, 1997) electron diffraction

data(Ganesh Jayaram, Xu & Marks, 1993) Procedure and route of simulations is similar to our previous dry oxidation simulations.(Pamungkas, Joe, Kim, & Lee, 2011a) Oxidation simulation was done by placing an H2O molecule at the distance 0.12 nm above the surface, without any external constraint. In order to compare the oxidation behavior with the previous ab initio calculations or the experimental results of surface reaction, single H2O molecule reactions were investigated on randomly selected surface positions with various water molecule configurations for 20 ps. 500 independent single molecular reactions were statistically analyzed. Then, simulation of wet oxidation by 600 water molecules deposition was done to silicon surface to study oxide surface layer formation.

40

3.3 Vicinal Si(001)


Here, we do not use vicinal surface with alternating smooth SA and rough SB steps as it exists in the real Si(001) surface, rather we simulate Si (001) surface with one type of step edge for each simulation box, with SA, SB, DA, and DB as depicted by figure 3,4,5 and 6 respectively so that different effects of each type of step edge can be readily observed.

Figure 3. Top view of surface with SA step

41

Figure 4. Top view of surface with SB step

42

Figure 5. Top view of Surface with DA step edge

43

Figure 6. Top view of DB

44

How SA, SB, DA, and DB defined can be understood by observing dimer raw direction as depicted by red and yellow lines at lower and upper terraces. In the present work, we investigated the oxidation by a reactive molecular dynamics simulation method using the reactive force field (ReaxFF) developed by van Duin et al which is integrated in Large-scale

Atomic/Molecular Massively Parallalized Simulator (LAMMPs) code. In order to mimic a thick substrate, Si atoms in the bottom 0.1 nm region of the simulation box was fixed. Atoms in the region of about 6.6 nm above the fixed layer were kept at a constant temperature for oxidation at 300K to provide a thermal bath in the system. Temperature of the thermal bath was rescaled at every time step (1fs) during the simulation. All the atoms above the thermal bath were unconstrained with an initial temperature set to the oxidation temperature. Oxidation simulation was started by placing an O2 molecule at the distance 0.12 nm above the surface. Complete explanations about the reason of using such values mentioned above, more detail of procedure used, and benchmark of force field we use for the simulation have been provided in our previous work.

45

4 Results and Discussion

4.1 Dry Oxidation


4.1.1 Single oxygen molecule reaction
Figure 7 shows the typical snapshots of the oxidation reaction when the oxygen molecule was placed on the dimer row of the relaxed Si (100) surface at 300K (Fig. 7(a)-(c)) and 1200K (Fig. 7(d)-(f)). Three different configurations of oxygen molecule with respect to the surface were investigated: = 90 (Fig. 7(a) and (d)), 45 (Fig. 7(b) and (e)), and 0 (Fig. 7(c) and (f)), where is the angle between the OO axis in the oxygen molecule and the surface normal. When =90 or 45, most oxygen molecules spontaneously dissociated into two oxygen atoms in at most 100 fs at both 300 and 1,200K.

46

Figure 7. Time evolution of single oxygen molecule reaction with different initial oxygen molecule orientation and position on silicon surface. For (a) 45 and 0o, respectively.

=90, (b) 45, and (c) 0o at 300K. Time evolutions at 1,200K are shown in (d), (e) and (f) for =90,

47

This simulation results support the experimental observations that the fist subsurface oxidation occurs without energy barrier(Yasuda et al., 2003),(H. Watanabe et al., 1998) . The temperature independence of the dissociation behavior will be discussed later. After dissociation, the oxygen atoms formed silanone species on the surface as indicated by arrows in Fig. 7. The silanone species as a stable intermediate phase of Si oxidation has been suggested by both infrared absorption spectroscopy and the first-principles quantum chemical calculations combined with STM study(A. Hemeryck A. Estve & Chabal, 2009)(Y. J. Chabal Krishnan Raghavachari & Garfunkel, 2002),(Hemeryck et al., 2007). The oxygen atoms then rapidly moved from the dimer row position to the channel between the dimers in the direction opposite to each other. The opposite movement would be a natural consequence of Coulombic repulsion. Both oxygen atoms were then incorporated into the back bonds of the dimer row, in agreement with the previous experimental and theoretical results(Gal-Nagy et al., 2009),(Ohno, Takizawa, Koizumi, Shudo, & Tanaka, 2007b), (H. Watanabe et al., 1998), (K. Kato & Uda, 2000), (K. Kato, Uda, & Terakura, 1998b). This reaction process occurred almost spontaneously within 0.5 ps at both 300K and 1,200K. When

=0 (Fig. 3(c) or (f)), it was frequently observed that upper oxygen atom in
the molecule leaves from the surface after the dissociation, whereas the lower oxygen atom reacts with Si atoms.

48

Figure 8. Time evolution of single oxygen molecule reaction when oxygen molecule is placed on a channel of dimer rows at 300K with

=45o.

Fig. 8 shows the oxidation process when the oxygen molecules were placed on the channel between dimmers. As expected, the reaction kinetics appeared to be dependent on the position of oxygen molecule on the Si (100) surface. Because of the longer distance from the surface silicon atoms, dissociation of the oxygen molecule took much longer than those in Fig. 7. However, the resulting atomic configuration of the oxygen incorporated Si surface is essentially the same. The spontaneous dissociation of oxygen molecule observed in the present MD simulation is consistent with recent ab initio calculations(H. Watanabe et al., 1998) (Miyamoto, Yoshiyuki, 1990)(Kato K & K, 1998)(K. Kato & Uda, 2000)(Ciacchi & Payne, 2005b) (Richard, Esteve, & Djafarirouhani, 2005). Richard et al showed that different barrierless pathways for direct dissociation of the oxygen molecule exist depending on the initial position of the molecule(Richard, Esteve, et al., 2005). They showed that low activation barrier (0.25eV) for the dissociation of oxygen molecule can be provided by the energy gain of the dimer-breaking accompanied with the chemisorption of the oxygen molecule. The ab initio 49

calculation also revealed that the dissociation of oxygen molecule on Si surface lower the energy by 4.14eV. This energy gain will generate hot atoms that can provide enough kinetics for the oxygen incorporation into the back bond of Si. The hot atom mechanism would explain the similar behavior of oxygen reaction regardless of the oxidation temperature.

4.1.2 Statistical independent single oxygen molecule reaction simulations


Table 3. Total uptake fraction of oxygen and the depth distribution of the oxygen obtained by statistical analysis of 1,000 single O2 reaction events.

Oxygen Position 300 K after 10 ps Layer 1 (0.0-0.2 nm) Layer 2 (0.2-0.4 nm) Layer 3 (0.4-0.6 nm) Layer 4 (0.6-0.8 nm) Layer 5 (0.8-1.0 nm) Total Fraction of Adsorbed Oxygen 70.76 % 20.63 % 8.23 % 0.37 % 0% 80.20 % 69.47 % 21.73 % 7.81 % 0.85% 0.13 % 76.15 % 1200 K

For statistical information of the oxygen-silicon interaction behavior, we performed statistical analysis of the reaction by simulating 1,000 independent single molecule reactions for 10 ps with randomly placing an 50

oxygen molecule on the Si (100) surface. (We observed that the system energy fluctuation due to the oxygen incorporation is settled down after 3 ps.) Table 3 shows the probability of an oxygen atom to be placed in a certain depth. In spite of the limited time scale of the MD simulation, it is evident that the oxygen could penetrate significantly at both 300 and 1200K. However, slightly higher probability of oxygen in deep layer 5 (0.8~1.0 nm in depth) was observed in 1,200K simulation. It is also obvious that the total fraction of oxygen atoms incorporated into Si substrate is dependent on the oxidation temperature. About 19.8 and 23.85 % of oxygen atoms left the Si surface at 300K and 1,200K, respectively. Figure 7 (c), (e) and (f) show the cases that one of the oxygen atoms left the Si surface after O2 dissociation. At higher temperature, higher kinetic of the oxygen molecule would increase the probability for the oxygen atom to leave without reaction. However, it must be noted that the difference in the fraction is not as significant as the difference in the oxidation temperature. This result implies that the oxidation reaction on Si (100) surface is more dependent on the initial configuration than the oxidation temperature, which is presumably due to the low energy barrier for the dissociation of the oxygen molecule (H. Watanabe et al., 1998), (Miyamoto, Yoshiyuki, 1990),(K. Kato & Uda, 2000), (Bongiorno & Pasquarello, 2004), (Ciacchi & Payne, 2005b), (Richard, Esteve, et al., 2005) . The consistency of the single O2 molecule reaction simulations with the

51

previous experimental and theoretical works supports that the reactive force field represents well the oxidation behavior of the Si surface.

4.1.3 Consecutive multiple oxygen molecules reaction

Next, we simulated the oxidation of the Si (100) surface by consecutively supplying 300 oxygen molecules on randomly chosen positions of the surface with randomly selected angle . The interval between two consecutive addition of the oxygen molecule was set to be 20 ps. Total simulation time was thus 6 ns. Before supplying each oxygen molecule, temperature of the system was rescaled to the oxidation temperature. Even if the time scale of the present simulation is not enough to consider the thermal diffusion of oxygen in the Si network, the present work revealed the initial oxidation behavior of clean Si (100) surface. Figure 9 (a) shows the atomic configuration of the oxidized surface. Oxygen atom is presented by blue ball while silicon dark brown one. It appears that the atomic structure of the surface became amorphous by the oxidation. We calculated the radial distribution function (RDF) of the oxidized surface layer of thickness 1.37 nm. Even if the RDF was of large statistical noise, it was evident that no long range order was observed. The distances of the first nearest neighbor of Si-O and O-O bonds are in good agreement with those of the amorphous SiO2 (Fig. 1). However, the first nearest neighbor distance of Si-Si bonds is closer to the 52

crystalline Si because the present simulation is in the sub oxide formation regime.

Figure 9. Top view of the oxidized surface by 300 oxygen molecules consecutively added on the surface every 20 ps. The atoms are colored by (a) the atom (oxygen in dark blue and silicon in dark brown) (b) the Mulliken charge as shown in the color index bar and (c) the coordination number.

Figure 9 (b) shows the charge distribution on the oxidized surface. Color scale in Fig. 9 (b) represents the range of the Mulliken charge values. 53

The snapshot evidently showed that the electrons transferred from silicon to oxygen during oxidation reaction. It was also noted that the charge transfer only occurred between silicon and the nearest oxygen atoms, which is in agreement with our own ab initio calculation (not shown here) that the charge transfer to oxygen occurs only from the nearest silicon atoms when an oxygen atom substitutes a silicon atom in silicon single crystal. Charge of Si increased with the number of nearest oxygen atoms, with the maximum of 1.60e with three oxygen atoms in the nearest neighbor. This value is corresponding to that of silicon dioxide(A. C. T. V. Duin et al., 2003). Colors in Fig. 9 (c) represent the coordination number of the atoms. The coordination number was estimated by the number of atoms within the radius of 0.35 nm. In the surface region, coordination number of most Si atoms is 3 or 4, whereas most oxygen atoms agglomerate around Si atoms with the coordination number of 2 (H. W. Yeom et al., 1999),(Pasquarello, Alfredo, Mark S Hybertsen, 1998), (H. W. Y. Yeom & Uhrberg, 1998) .

54

Figure 10. Oxygen depth profile of the oxidized silicon surface after consecutive oxidation reactions with 300 oxygen molecules. Each data was taken in the layer of thickness 0.3nm from the surface. Because of increased surface roughness by oxidation, data in depth 1is undependable

Initial oxidation behaviors at 300K and 1,200K were compared by analyzing the simulated oxide layer. Figure 10 shows the oxygen depth profile after the simulation. Considering the present simulation time of 6 ns, it is surprising that the oxygen diffused in the significant depth. The diffusion in this simulation would be the ballistic one as proposed by Yasuda et al (Yasuda et al., 2003). It is also obvious that the oxygen profiles are dependent on the oxidation temperature. At 300K, oxygen atoms were accumulated on the surface layer. On the other hand, oxygen atoms penetrated into deeper layer resulting in more uniform distribution of oxygen at 1,200K. This would 55

results from the strong temperature dependence of thermal diffusion of oxygen in Si network. Even if this oxygen profile is in contrast to the layerby-layer growth model(No Title, n.d.), (T. Watanabe K. Tatsumura, 2004), (Ganster et al., 2010) the present simulation result is qualitatively consistent with oxygen depth profiles obtained by high resolution XPS analysis of Si/SiOx interface(Oh et al., 2001), HRBS spectra(Nakajima et al., 2001) and classical MD simulation of Si/SiOx interface(Ohta et al., 2008).

Figure 11. Time evolution of the charge state of silicon during consecutive oxidation simulation at 300K

56

Figure 12. Time evolution of the charge state of silicon during consecutive oxidation simulation at 1,200K.

The oxygen composition profile affected the stoichiometry of the evolved oxide layer. Figure 11 and 12 show the time evolution of Si+, Si2+, Si3+, and Si4+ ion fraction in the specimen during oxidation simulation. Charge state of Si was determined by the number of nearest oxygen atoms, as suggested by the calculated core shifts in Si-SiO2 interface(Pasquarello, Hybertsen, & Car, 1996). At 300K, number of Si+ monotonically decreased as the oxidation proceeded while those of Si2+, Si3+ and Si4+ increased (See Fig. 11). The oxygen atoms accumulated on the surface layer at 300K generated a stoichiometric surface oxide layer. Actually, we could observe that Si4+ ions are accumulated in the surface layer. Recent in-situ experimental observation

57

of the oxidized Si (100) surface also reported that the 1st layer is fully oxidized at room temperature oxidation (Gal-Nagy et al., 2009). Figure 12 shows the time evolution of Si ion fraction during 1,200K oxidation simulation. It can be noted that more than 50% of the Si ions remained as Si + while the fraction of Si4+ was less than 5%. This result that the sub-oxide formation is remarkable at 1,200K would reveal that the driving force for the oxygen diffusion into Si network can be higher than the energy gain by stoichiometric oxide formation. In the present simulation, total oxidation time was limited to 6.0ns. It must be noted that the present simulation is applied only for very initial stage of the oxidation, and thermal diffusion of oxygen in long time scale is excluded. In spite of this limitation, the present simulation evidently shows the difference in the oxygen depth profile between the processes at 300 and 1,200K. In real oxidation process where thermal diffusion of oxygen occurs, the difference in the oxidation behavior at two different temperatures would be more significant than the present simulation results.

58

4.2 Wet Oxidation


4.2.1 Single water molecule reaction
In order to observe the details of the interaction between water molecule and Si surface, single water adsorption simulation was performed with various initial orientation and position of water molecule. Typical examples of 300K reaction are presented in Figures 13 to 15. The oxygen atom is presented by a dark blue ball, while two hydrogen atoms are represented by small red and orange balls, respectively. In the substrate, Si atoms composing the dimer row are denoted by green balls, and other Si atoms are presented by bright blue balls. In Fig. 13, the water molecule attached the Si surface with one hydrogen atom in the molecule. Figure 14 is the case when oxygen atom in water molecule attached the Si surface with hydrogen atoms behind. Figure 15 is the opposite case where two hydrogen atoms attached the Si surface simultaneously. Although the reaction kinetics was dependent on the position and the orientation of water molecule, most water molecule was dissociated in two steps: water dissociated into atomic hydrogen and hydroxyl (OH), followed by the dissociation of the hydroxyl into another hydrogen and oxygen. Then, the oxygen reacts with Si and placed in the back bonds of dimers. For example, Figure 13 (b) shows that the water molecule dissociated into atomic hydrogen and hydroxyl group within a few 10 fs. 59

Figure 13. Time evolution of single water molecule reaction when water molecule, with two hydrogen atoms at the left side and oxygen atom at the right side of molecule at 300 K, is placed (a) on the middle of array dimer, (b) just above one of atom dimer, and (c) channel between two dimer array.

60

Figure 14. Time evolution of single water molecule reaction when water molecule, with two hydrogen atoms on the top and oxygen atom at the bottom side of molecule at 300 K, is placed: on (a) the middle of array dimer, (b) just above one of atom dimer, (c) at a small distance from dimer atoms (d) exactly above the channel between two dimer array

61

Figure 15. Time evolution of single water molecule reaction when water molecule, with two hydrogen atoms at the bottom and oxygen atom at the top side of molecule at 300 K, is placed: on (a) the middle of array dimer, (b) just above one of atom dimer, (c) at a small distance from dimer atoms (d) exactly above the channel between two dimer array.

62

Figure 16. Time evolution of single water molecule reaction when water molecule on a silicon dimer atom with two hydrogen atoms at the left side and oxygen atom at the right side of molecule at 300 K. This snapshot is identical with Fig.1.b but it is colored by the Mulliken charge as shown in the color index bar.

63

Figure 17. Time evolution of single water molecule reaction when water molecule adsorbs molecularly at 300 K. This snapshot is identical with but it is colored by the Mulliken charge as shown in the color index bar

64

The hydroxyl group was later dissociated into atomic hydrogen and oxygen (see the snapshot at 120 fs of Fig. 13 (b)). The dissociated oxygen atom reacted with Si and positioned at the back bonds of the dimer as shown in the snapshot at 20 ps. In the present simulation, no case was observed where water molecule dissociates directly into oxygen atom and two hydrogen atoms. These results agree well with the previously suggested reaction model (Schulze, 1994), validating the interatomic potential used in the present work. Similar two step reaction was observed in many simulation cases: Fig. 13 (a) and (c), all of Fig. 14 and Fig. 15 (b) and (c). However, in some rare cases, water molecule remains without reaction in 20 ps (see Fig 15 (a) and (d)). We also observed that the dissociated oxygen or hydroxyl desorbed from the surface as shown in Fig. 14 (b) and (d). Statistical analysis of the reaction behavior during wet oxidation will be discussed later in this paper.

Figure 16 shows the details of charge variation during oxygen reaction. The atoms are colored by the Mulliken charge in the scale of the color scale bar on the right side. When the water molecule was dissociated into hydrogen and hydroxyl group (see snapshot at 30 fs of Fig 4), Si atom indicated by an arrow is positively charged by transferring the electron to the dissociated hydrogen atom (indicated by a broken line arrow). At 40fs, the hydrogen atom was neutralized by incorporating into the Si substrate, and the negatively charged oxygen in hydroxyl is in bound to the positively charged 65

Si by exchanging electrons. Figure 16 (b) shows the second step of the reaction where the hydroxyl was dissociated into hydrogen and oxygen atoms. At 120 fs, the hydroxyl started to dissociate into hydrogen and oxygen atoms. Hydrogen atom changed to neutral while the Si atom (indicated by an arrow) on the substrate is more positively charged. As the dissociation occurred, the bond length between O and Si decreased from 1.66 to 1.40 , indicating the stronger chemical reaction. After dissociation, oxygen atom was positioned at the back bond of dimer (See Fig. 13 (c) or Fig. 14 (c). Oxygen atom was not shown in the 20 ps snapshot of Fig. 16 (b).) Difference between wet and dry oxidation was evidently shown when the oxygen or water molecule was placed on the channel between two dimer rows. As shown in Fig. 13 (c), where a water molecule positioned on the channel between two dimers dissociated within few 10 fs. In the case of dry oxidation,(Pamungkas, Joe, Kim, & Lee, 2011a) much longer period is required before the molecular dissociation (see Fig. 4 in [(Pamungkas, Joe, Kim, & Lee, 2011b)]). Considering the similar bond-dissociation energy of OH-H (493.4 kJ/mol) and O=O (497 kJ/mol), this result would suggest that hydrogen is more reactive than oxygen on the Si surface. This suggestion would be also supported by the result in Fig. 14 (d) where the simulation started with the oxygen atom facing to the Si surface. It must be noted that the dissociation of water started when hydrogen atom started to react with Si

66

atoms after significant rotation of the water molecule. In this case, it took about 2 ps before the dissociation. Our simulation also revealed the possibility of hydroxyl and hydrogen species to react with Si in different dimers or even in the different dimer rows (for example, see Fig. 14 (c)). By high resolution infrared spectroscopy, it was reported that after a water molecule dissociates into hydroxyl and atomic hydrogen, the species bind to adjacent Si atoms of the same dimer, so called on-dimer (OD) configuration.(Y.J.Chabal, 1984) However, later works using STM suggested the intra dimer (ID) configuration, where hydroxyl and atomic hydrogen bind to two neighboring dimers in the same dimer row.(Hossain, Yamashita, Mukai, & Yoshinobu, 2003) More recent STM study further demonstrated the coexistence of the two configurations.(S.-yong Yu, Kim, & Koo, 2008), (S.-yong Yu, Kim, Kim, & Koo, 2011) Our simulation results would support the recent STM study.(Hossain et al., 2003), (S.-yong Yu et al., 2008), (S.-yong Yu et al., 2011) Furthermore, it would be also suggested that the hydrogen and hydroxyl can react with different dimer rows at 300K. Molecular adsorption of water was also observed in some cases. Figure 3 (d) shows the typical examples of the molecular adsorption. It is of worthy noting that hydrogen atoms are always bound to Si in the case of molecular adsorption. The strong tendency of the molecular adsorption when two hydrogen atoms of water molecule are bound to the surface Si atoms,

67

might correspond to the higher reactivity of hydrogen. Figure 5 shows the Mulliken charge of atoms during molecular adsorption. Two hydrogen atoms seem to passivate the surface dangling bonds of Si, without notable charge transfer between Si and hydrogen atoms. The present simulation would explain the previous experimental report(Ranke, 1996) that the kinetics is described as a mixture of dissociative chemisorption and non-dissociative chemisorption with the binding enthalpy of just 6 kJ/mol (0.06 eV/molecule). Recent theoretical study also showed that molecular adsorption is more favorable on hydrogenated Si surface.(Lelis-Sousa & Caldas, 2011) Therefore, the hydrogen reaction with the surface Si atoms triggers the molecular dissociation of water. Almost identical behavior of the surface reaction was observed at higher temperature, 1200 K. Water molecules dissociated into an atomic hydrogen and hydroxyl, followed by the dissociation of the hydroxyl molecule. The temperature dependent adsorption behavior was not observed in the single molecule simulations, which indicates the small energy barrier for the molecular dissociation. By the first principle calculations, Cho et al showed that the energy barrier of water dissociation on Si surface is only 0.15 eV,(Cho, Kim, Lee, & Kang, 2000b) which is even lower than that of oxygen molecule dissociation (0.25 eV).(Pamungkas, Joe, Kim, & Lee, 2011a), (Richard, Estve, & Djafari-Rouhani, 2005)They also reported that the dissociation of water molecule on Si surface lower the energy by 1.8 eV. This 68

energy gain will generate hot atoms that can provide enough kinetics for the dissociation of water molecular both at 300 and 1200K. Independent dissociation behavior from the oxidation temperature would be explained by the hot atom mechanism.

4.2.2 Statistical independent single water molecule reaction simulations

Table 4 shows probability of single oxygen atom adsorbed in silicon surface obtained from 500 times independent single water molecule adsorption simulation. Adsorbed oxygen atom almost always bond to silicon atom at first silicon layer, yet there is a small probability of oxygen atom to bond to slight deeper silicon. We also can notice that probability of oxygen adsorption at room temperature and high temperature (1200K) is very similar. This is also can be explained by very small barrier of the adsorption process (Cho, Kim, Lee, & Kang, 2000b). Table 5 demonstrates that hydrogen atoms can penetrate into very deep Si matrix both at 300 K and 1200 K. Hydrogen atom is also easy to be desorbed. Probability of hydrogen desorption at 1200 K, as expected, significantly more than that at room temperature

69

Table 4 Total uptake fraction of oxygen and the depth distribution of the
oxygen obtained by statistical analysis of 1,000 single H2O molecule reaction events Oxidation position after 20 ps Depth 1 (0.0-0.2 nm) Depth 2 (0.2-0.4 nm) Depth 3 (0.4-0.6 nm) Total fraction of adsorbed oxygen 300 K 95.60 % 3.91 % 0.49 % 81.80% 1200 K 97.79% 1.47% 0.25% 81.40%

Table 5. Total uptake fraction of Hydrogen and the depth distribution of the
oxygen obtained by statistical analysis of 1,000 single H2O molecule reaction events Hydrogen position after 20 ps Depth 1 (0.0-0.2 nm) Depth 2 (0.2-0.4 nm) Depth 3 (0.4-0.6 nm) Depth 4 (0.6-0.8 nm) Depth 5 (0.8-1.0 nm) Depth 6 (1.0-1.2 nm) Depth 7 (1.2-1.4 nm) Depth 8 (1.4-1.6 nm) Depth 9 (1.6-1.8 nm) Depth 10 (1.8-2.0 nm) Depth 11 (2.0-2.2 nm) Depth 12 (2.2-2.4 nm) Total fraction of adsorbed hydrogen 300 K 76.755 % 8.736 % 3.588 % 3.120 % 3.120 % 1.716 % 1.404 % 0.936 % 0.156 % 0.156 % 0.156 % 0.156 % 64.1 % 70 1200 K 70.849 % 10.039 % 5.598 % 4.440 % 3.088 % 2.123 % 1.158 % 0.386 % 0.772 % 0.772 % 0% 0.193 % 51.8 %

4.2.3 Consecutive multiple water molecules reaction


As a next step, we performed 600 water molecules reaction by consecutively supplying the water molecule of randomly chosen pose on randomly selected position of the Si surface. The interval between two consecutive additions of the water molecule was 20 ps. Before supplying each water molecule, temperature of the system was rescale to the oxidation temperature. Uptake behavior of oxygen and hydrogen during oxidation simulation at 300 and 1200K was presented in Fig. 18. Figure 18 shows that slightly increased amount of oxygen was adsorbed at higher temperature when more than 200 water molecules were supplied to the Si surface. This result is in contrast to that of the dry oxidation, where more oxygen atoms were adsorbed at 300K in the later stage of oxidation (see inset of Fig. 18 ). Enhanced oxidation at higher temperature with the addition of larger number of water molecule revealed the role of intermolecular interaction on the oxidation behavior, as will be discussed later. Figure 19 shows the hydrogen adsorption behavior that is strongly dependent on the temperature. Hydrogen desorption was enhanced at 1200 K due to significant thermal activation, resulting in the decreased amount of hydrogen adsorption. Energy barrier for hydrogen desorption is in the range of 2 - 3.4 eV depending on the hydrogen adsorption state.(N.Yabumoto, K.Minegishi, Y.Komne, 1990) amount of hydrogen desorption is thus expected at 300 K. 71 Negligible

Figure 18. Time evolution of number of adsorbed oxygen atoms in wet oxidation and dry oxidation (inset).

72

Figure 19. Time evolution of number of adsorbed hydrogen atoms in wet oxidation

73

Figure 20 shows the atomic configuration of the oxidized surface when 150 water molecules reacted with the Si substrate. Figure 20 (a) and (b) are respectively atomic configuration and charge distribution at 300K. Figure 20 (c) and (d) are those at 1200K. In the atomic configurations, Si, oxygen and hydrogen are respectively presented by red, blue and green balls. The charge of the atom is described by the Mulliken charge values in terms of the color scale on the right side. Figure 20 (a) and (c) show that as oxidation proceeds, surface structure became amorphous. At 300 K, many hydroxyl groups remain intact at room temperature as indicated by arrows in Fig. 20 (a). It is evident in Fig. 20 (b) that the hydrogen bound to oxygen was positively charged, as in typical hydroxyl molecule. Because of higher value of electronegativity of oxygen (3.44 in Pauling scale) than hydrogen (2.20) or Si (1.90), all oxygen atoms appear negatively charged up to 1.0e, by receiving electrons from both hydrogen and Si atoms nearby. At 1200 K, it was observed that much smaller amount of hydrogen remained on the surface.

74

Figure 20. Top view of the oxidized Si(001) surface by 150 water molecule consecutively added on the surface. The atoms are colored by (a) and (c) type of atom (oxygen in dark blue, silicon in pile green, and hydrogen in red), (b) and (d) the Mulliken charge as shown in the color index bar. Simulation is performed at (a) & (b) 300 K and (c) & (d) at 1200 K

75

Figure 21. Top view of the oxidized Si(001) surface by 600 water molecule consecutively added on the surface. The atoms are colored by (a) and (c) type of atom (oxygen in dark blue, silicon in pile green, and hydrogen in red), (b) and (d) the Mulliken charge as shown in the color index bar. Simulation is performed at (a) & (b) 300 K and (c) & (d) at 1200 K

76

The adsorbed hydrogen turned to neutral as shown in Fig. 20 (d). Furthermore, no hydroxyl molecule was found on the surface in this simulation. Figure 21 is the same figures after 600 water molecules were supplied to the Si surface. More significant amorphization occurred. Comparing with the case of 150 water molecules reaction shown in Fig. 20 (b) and (d), Si atoms were more positively charged as the transfer of electron occurs to the oxygen atoms nearby. If one compares the oxidized surfaces between 300 K and 1200 K, it would be apparent that the oxidation state of Si and the oxygen concentration on the surface was lower at 1200 K. Because the total amount of adsorbed oxygen is higher at 1200 K (Fig. 18 (a)), this result implies that significant oxygen diffusion into Si substrate occurs for short simulation time of only 12 ns. Depth profile analysis of oxygen shown in Fig. 9 revealed that the significant oxygen diffusion occurred at 1200K. Even at 300K some oxygen atoms could diffuse down to 1.0 nm. The simulated depth profile of oxygen supports the ballistic diffusion model of oxygen in the early stage of oxidation, as proposed by Yasuda et al. (Yasuda et al., 2003) In the case of hydrogen, much deeper penetration of the adsorbed hydrogen was observed, which would be a natural consequence of high mobility of hydrogen in Si lattice.

77

It is also interesting to note that most of hydroxyl molecule was decomposed on the Si surface even at 300K (see Fig. 21 (a) and (b)). Only a few hydroxyl molecules were found in the present simulation as indicated by arrows in Fig. 21. Many previous experimental works on the temperature dependence of hydroxyl decomposition on Si surface reported that the decomposition usually occurs at an elevated temperature. Schulze et al reported that hydroxyl decomposes to bridge-bonded oxygen and addition silicon hydride upon heating from 300K to 400K.(Schulze, 1994) An STM study has reported that hydroxyl is completely decomposed into O and H atoms at the temperature of higher than 563K on Si (111) surface.(R-L. Lo, IS.Hwang, 2003) It was also reported that thermal decomposition of SiOD to form Si-O-Si and Si-D mainly occurs between 500 to 750K, even if the decomposition starts at 300K.(X,-L.Zhou, C.R.Flores, and White, 1992) Other experimental works using high resolution IR spectroscopy, LEED, AES and Photoemission yield spectroscopy (PYS) reported that the elevated temperature arouses the dissociation of OH.(Y.J.Chabal, 1984), (Aissa, Zaibi, & Lacharme, 2005) However, the present simulation revealed that the decomposition of hydroxyl is also dependent on the amount of water molecules incorporated. The reaction between hydroxyl-hydroxyl or waterhydroxyl molecules may play an important role in the enhanced decomposition of hydroxyl at room temperature.

78

Previous first principle calculation predicted the repulsive interaction between water molecules on Si surface.(Cho, Kim, Lee, & Kang, 2000b) The repulsive interaction pushed each other to the Si dimers facilitating the dissociation and adsorption process. We applied the reactive MD simulation to study intermolecular interaction during oxidation. Figure 23 shows the reaction snapshots when two water molecules (Fig. 23 (a)), water and hydroxyl molecules (Fig. 23 (b)) and two hydroxyl molecules (Fig. 23 (c)) were supplied to the Si surface. We deliberately performed the simulation with the same initial orientation and position of water molecule as in Fig. 15 (d) where molecular adsorption was observed because in the single water molecule reaction. The present simulation showed that a repulsive interaction between two molecules occurred in all the combinations of Fig. 23. Higher mobility of the molecule on Si surface at higher temperature would increase the chance of repulsive reaction between molecules. This repulsion will enhance the decomposition of the molecules and thus the oxidation of Si as shown in Fig. 23. The intermolecular repulsion might explain the strong temperature dependence of the oxygen uptake in multi water molecules simulations (Fig. 18) that was in contrast to the statistical analysis of 1000 independent water molecule reaction (see Table 1). Various charge states of Si depending on the surrounding atomic configuration were evident as shown in Fig. 21 (b) or (d). The Si atoms of the oxidized substrate can be thus categorized by the Si oxidation state, such as

79

Figure 22. Oxygen depth profile of the oxidized silicon surface after consecutive oxidation reactions with 600 water molecules. Each data was taken in the layer of thickness 0.32 nm from the surface.

80

Si+, Si2+, Si3+ and Si4+. Figure 24 and 25 show the concentration variation of Si during oxidation simulation at 300 and 1200 K. The results can be compared with those of dry oxidation.(Pamungkas, Joe, Kim, & Lee, 2011a) In dry oxidation, more stoichiometric oxide rapidly formed at 300K since the oxygen atoms accumulated on the surface first layer at 300K. In contrast, oxygen atoms penetrated into deeper Si layer at 1200K which led to the larger number of Si atoms in sub-oxide state. Similar behavior was observed during wet oxidation. At 1200 K, oxygen atoms can diffuse to the deeper Si substrate (see Fig. 22) resulting in the Si atoms of sub-stoichiometric state. However, the temperature dependence was not as significant as that in dry oxidation, presumably because the number of oxygen adsorption at 300K is less than that at 1200K (Fig. 18 ). As discussed above, higher number of oxygen adsorption in wet oxidation at high temperature is related to dissociation of water molecules and decomposition of hydroxyl which are enhanced by interacting water molecules and their constituent fragments. Thus, difference of temperature dependence in wet oxidation and dry oxidation presumably stem from that mechanism

81

Figure 23. Time evolution of (a) water molecule-water molecule reaction, (b) water molecule-hydroxyl reaction, and (c) hydroxyl-hydroxyl reaction.

82

Figure 24. Time evolution of the charge state of silicon during consecutive simulation at (a) 300 K

83

Figure 25. Time evolution of the charge state of silicon during consecutive simulation at 1200K.

84

4.3 Vicinal Surface


In order to investigate oxygen molecule reaction with vicinal surface, we performed one oxygen molecule adsorption simulation on step edge of silicon surface with several different positions and different type of step edge. Figure 26, 27, 28, and 29 represent reaction of single oxygen molecule with SA, SB, DA, and DB step edge respectively. While, notation (a) represent position of oxygen molecule: above lower terrace atom, (c) represent oxygen molecule above upper terrace silicon atom, and (b) is in between (a) and (c). In general, oxygen molecule dissociate in very short time, within 60 fs. Similar to dry oxidation of silicon flat surface, oxygen atom adsorb into back bond of silicon atoms. In those aspects, different type of step edge does not lead to different adsorption process. In Figure 26.a, the two oxygen atoms, after oxygen molecule dissociation, are attracted to different direction. One atom bond to upper terrace and another atom bond to lower terrace. It is also possible that single oxygen atom or SiO leave the surface after dissociation (Figure 26.c and 29.c).

85

Figure 26. Single oxygen molecule reaction on silicon surface with SA step edge

86

Figure 27. Single oxygen molecule reaction on silicon surface with SB step edge

87

Apart from those special cases, reaction of oxygen molecule with all type of surfaces looks very similar. In contrast of this result, previous ab initio study(Yeo et al., 2007) and experimental observation using Scanning Tunneling Microscopy (STM) (Chung et al., 2006) at 80 K which proposed that , reactions with Si atoms at SB steps are dominant over other type of surface. Considering different temperature used in this present work as well as number of oxygen molecule involved in the reaction,it is then suggested for further consecutive MD simulations with several different temperatures and different step densities.

88

Figure 28. Single oxygen molecule reaction on silicon surface with DA step edge

89

Figure 29. Single oxygen molecule reaction on silicon surfacewith DB step edge

90

5 Conclusion

Reactive MD simulations of early stage of dry oxidation of Si (100) surface using the reactive force field of Si-O system showed the reaction behaviors in atomic scale which are in good agreement with the previous experimental and ab initio calculation results. The present work revealed the viability of the reactive MD technique for a massive simulation of Si oxidation in atomic scale without external constraint on the oxygen molecule. The simulation of single oxygen molecule reaction confirmed typical phenomena in silicon oxidation process such as dissociative adsorption, silanone precursor of oxidation, oxygen agglomeration and ion silicon transformation. Consecutive oxidation simulation for 6 ns exhibited the initial evolution of the surface oxide layer. At the oxidation temperature of 1,200K, significant sub-oxide layer was formed during the simulation for 6 ns, whereas the simulation at 300K showed the surface layer evolution to form more stoichiometric SiO2 layer. We have presented atomic study of the early stage of wet oxidation Si(001) using Reactive molecular dynamic simulations. It is found that dissociative and molecularly adsorptions are observed. Hydroxyl, hydrogen atom and oxygen atom of a water molecule can bond to silicon atom in different array of dimer. Temperature dependence of oxidation process observed in many water molecules adsorption simulation can be attributed to 91

the fact that many Hydroxyl groups are not decomposed at room temperature, and completely decomposed into oxygen and hydrogen atoms at high temperature as well as interaction between water molecules and their atoms and fraction constituent. Reactive MD has been implemented in observing interaction of oxygen molecule with Si(001) containing step edge. Effects of step edge type is subtle. Further simulation which involve consecutive oxygen molecule reaction in low temperature is then suggested.

92

References
A. Hemeryck A. Estve, N. R. M. D. R., & Chabal, Y. J. (2009). No Title. , Phys. Rev. B, 79, 35317. A.Rahman. (1964). Correlations in the Motion of Atoms in Liquid Argon. Phys. Rev, 136, A405. Aissa, a, Zaibi, M., & Lacharme, J. (2005). Interaction of hot Si(100)21 with water vapour. Vacuum, 79(1-2), 19-24. doi:10.1016/j.vacuum.2005.01.005 B.W.Holland, C.B.Duke, A. P. (1984). The Atomic Geometry of Si(100)-(2x1) Revisited. Surf. Sci., 140, L269-L278. Berendsen, H. J. C., Postma, J. P. M., van Gunsteren, W. F., DiNola, a., & Haak, J. R. (1984). Molecular dynamics with coupling to an external bath. The Journal of Chemical Physics, 81(8), 3684. doi:10.1063/1.448118 Billeter, S. R., Curioni, A., Fischer, D., & Andreoni, W. (2006). Ab Initio derived augmented Tersoff potential for silicon oxynitride compounds and their interface with silicon. Physical Review B, 73, 155329. doi:10.1103/PhysRevB.73.155329 Bongiorno, A., & Pasquarello, A. (2003). Atomistic structure of the Si(100) SiO[sub 2] interface: A synthesis of experimental data. Appl. Phys. Lett, 83(7), 1417. doi:10.1063/1.1604470 Bongiorno, A., & Pasquarello, A. (2004). Reaction of the Oxygen Molecule at the Si(100)-SiO2 Interface During Silicon Oxidation. Phys. Rev. Lett, 93(8), 1-4. doi:10.1103/PhysRevLett.93.086102

93

Brichzin, V., & Pelz, J. (1999). Effect of surface steps on oxide-cluster nucleation and sticking of oxygen on Si(001) surfaces. Physical Review B, 59(15), 10138-10144. doi:10.1103/PhysRevB.59.10138 Chadi, D. (1987). Stabilities of single-layer and bilayer steps on Si(001) surfaces. Physical review letters, 59(15), 1691-1694. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10035304 Cho, J.-H., Kim, K., Lee, S.-H., & Kang, M.-H. (2000a). Dissociative adsorption of water on the Si(001) surface: A first-principles study. Phys. Rev. B, 61(7), 4503. doi:10.1103/PhysRevB.61.4503 Cho, J.-H., Kim, K., Lee, S.-H., & Kang, M.-H. (2000b). Dissociative adsorption of water on the Si(001) surface: A first-principles study. Phys. Rev. B, 61(7), 4503-4506. doi:10.1103/PhysRevB.61.4503 Chung, C.-H., Yeom, H., Yu, B., & Lyo, I.-W. (2006). Oxidation of Step Edges on Si(001)-c(42). Physical Review Letters, 97(3), 2-5. doi:10.1103/PhysRevLett.97.036103 Ciacchi, L., & Payne, M. (2005a). First-Principles Molecular-Dynamics Study of Native Oxide Growth on Si(001). Physical Review Letters, 95(19), 25. doi:10.1103/PhysRevLett.95.196101 Ciacchi, L., & Payne, M. (2005b). First-Principles Molecular-Dynamics Study of Native Oxide Growth on Si(001). Physical Review Letters, 95(19), 25. doi:10.1103/PhysRevLett.95.196101 Colombi Ciacchi, L., Cole, D. J., Payne, M. C., & Gumbsch, P. (2008). Stress-Driven Oxidation Chemistry of Wet Silicon Surfaces. J. Phys. Chem. C, 112(32), 12077-12080. doi:10.1021/jp804078n David I.Grimley, Adrian C. Wright, Sinclair, R. N. (1990). N E U T R O N S C A T T E R I N G F R O M V I T R E O U S SILICA IV. Time-offlight diffraction David I. G R I M L E Y and Adrian C. W R I G H T. Journal of Non-Crystalline Solids, 119, 49-64. Demkov, A. A., & Sankey, O. F. (1999). Growth Study and Theoretical Investigation of the Ultrathin Oxide SiO 2 -Si Heterojunction. Physical Review Letters, 83, 10-13.

94

Duin, A. C. T. V., Strachan, A., Stewman, S., Zhang, Q., Xu, X., & Goddard, W. A. (2003). Reactive Force Field for Silicon and Silicon Oxide Systems. J.Phys. Chem A, 107(1), 3803. E.Wainwright, B. J. A. and T. (1957). Phase Transition for a Hard Sphere System. J. Chem. Phys, 27, 1208. Engel, T. (1993). The interaction of molecular and atomic oxygen with Si(100) and Si(111). Surface Science Reports, 18(4), 93-144. doi:10.1016/01675729(93)90016-I Enta, Y., Mun, B. S., Rossi, M., Ross, P. N., Hussain, Z., Fadley, C. S., Lee, K.-S., et al. (2008). Real-time observation of the dry oxidation of the Si(100) surface with ambient pressure x-ray photoelectron spectroscopy. Applied Physics Letters, 92(1), 012110. doi:10.1063/1.2830332 Ercolessi, F. (1997). A molecular dynamics primer. Retrieved from http://www.sissa.it/furio Fan, X., Zhang, Y., Lau, W., & Liu, Z. (2005). Adsorption of Triplet O2 on Si(100): The Crucial Step in the Initial Oxidation of a Silicon Surface. Physical Review Letters, 94(1), 1-4. doi:10.1103/PhysRevLett.94.016101 Fischer, D., Curioni, A., Billeter, S., & Andreoni, W. (2006). The structure of the SiO[sub 2]Si(100) interface from a restraint-free search using computer simulations. Appl. Phys. Lett, 88(1), 012101. doi:10.1063/1.2158520 Fogarty, J. C., Aktulga, H. M., Grama, A. Y., van Duin, A. C. T., & Pandit, S. a. (2010). A reactive molecular dynamics simulation of the silica-water interface. J. Chem. Phys, 132(17), 174704. doi:10.1063/1.3407433 Ganesh Jayaram, Xu, P., & Marks, L. D. (1993). Structure of Si(100)-(2x1) Surface Using UHV Transmission Electron Diffraction. Phys. Rev. Lett, 71(21), 3489-3493. Ganster, P., Trglia, G., Lancon, F., & Pochet, P. (2010). Molecular dynamics simulation of silicon oxidization. Thin Solid Films, 518(9), 2422-2426. Elsevier B.V. doi:10.1016/j.tsf.2009.09.144 Gal-Nagy, K., Incze, A., Onida, G., Borensztein, Y., Witkowski, N., Pluchery, O., Fuchs, F., et al. (2009). Optical spectra and microscopic 95

structure of the oxidized Si(100) surface: Combined in situ optical experiments and first principles calculations. Physical Review B, 79(4), 1-10. doi:10.1103/PhysRevB.79.045312 Hemeryck, a, Mayne, a J., Richard, N., Estve, a, Chabal, Y. J., Djafari Rouhani, M., Dujardin, G., et al. (2007). Difficulty for oxygen to incorporate into the silicon network during initial O2 oxidation of Si(100)-(2x1). The Journal of chemical physics, 126(11), 114707. doi:10.1063/1.2566299 Hemeryck, a., Estve, a., Richard, N., Djafari Rouhani, M., & Chabal, Y. (2009). Fundamental steps towards interface amorphization during silicon oxidation: Density functional theory calculations. Physical Review B, 79(3), 1-5. doi:10.1103/PhysRevB.79.035317 Hirose, K., Nohira, H., Azuma, K., & Hattori, T. (2007). Photoelectron spectroscopy studies of SiO2/Si interfaces. Prog. Surf. Sci, 82(1), 3-54. doi:10.1016/j.progsurf.2006.10.001 Hossain, M., Yamashita, Y., Mukai, K., & Yoshinobu, J. (2003). Model for C defect on Si(100): The dissociative adsorption of a single water molecule on two adjacent dimers. Phys. Rev. B, 67(15), 153307. doi:10.1103/PhysRevB.67.153307 Itoh, H., Nakamura, K., Kurokawa, A., & Ichimura, S. (2001). Initial oxidation process by ozone on Si(100) investigated by scanning tunneling microscopy. Surface Science, 482-485, 114-120. doi:10.1016/S0039-6028(00)01004-9 J.M.Haile. (1997). Molecular Dynamics Simulation: Elementry Methods. A Wiley Interscience Publication. Jarek Dabrowski, H.-J. M. (2000). Silicon Surface and Formation Interfaces. World Scientific Publishing. Jiang, Z., & Brown, R. (1995). Atomistic calculation of oxygen diffusivity in crystalline silicon. Physical review letters, 74(11), 2046-2049. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10057828 Kato, H., Kawai, M., Akagi, K., & Tsuneyuki, S. (2005). Interaction of condensed water molecules with hydroxyl and hydrogen groups on Si(001). Surf .Sci., 587(1-2), 34. doi:10.1016/j.susc.2005.04.032

96

Kato K, U. T., & K, T. (1998). No Title. Phys. Rev. Lett, 80, 2000. Kato, K., & Uda, T. (2000). Chemisorption of a single oxygen molecule on the Si(100) surface: Initial oxidation mechanisms. Physical Review B, 62(23), 15978-15988. doi:10.1103/PhysRevB.62.15978 Kato, K., Uda, T., & Terakura, K. (1998a). Backbond Oxidation of the Si(001) Surface: Narrow Channel of Barrierless Oxidation. Physical Review Letters, 80(9), 2000-2003. doi:10.1103/PhysRevLett.80.2000 Kato, K., Uda, T., & Terakura, K. (1998b). Backbond Oxidation of the Si(001) Surface: Narrow Channel of Barrierless Oxidation. Physical Review Letters, 80(9), 2000-2003. doi:10.1103/PhysRevLett.80.2000 one n , ., oren, . . (1 ). Adsorption of water on Si(100)-(21): A study with density functional theory. J. Chem. Phys, 106(6), 2426-2435. doi:10.1063/1.473346 Lelis-Sousa, R., & Caldas, M. (2011). Ab initio study of the early stages of gas-phase water oxidation of the Si(100) (21):H surface. Phys. Rev. B, 84(20), 205314. doi:10.1103/PhysRevB.84.205314 M.Chander, Y.Z.Li , J.C.Patrin, and J. H. W. (1993). Si(100)-(2 X 1) surface defects and dissociative and nondissociative adsorption of H2O studied with scanning tunneling microscopy. Phys. Rev. B, 48(4), 2493-2502. Miyamoto, Yoshiyuki, A. O. (1990). Atomic and electronic structure of oxygen on Si(100) surfaces: Metastable adsorption sites. Physical Review, 41(18), 12680. Miyamoto, Yoshiyuki, A. O. (1991). Energetic in the initial stage of oxidation of silicon. Physical Review B, 43(11), 9287-9290. Moor, G. E. (1965). Cramming more components onto integrated circuit. Electronics, 38, 8. Retrieved from http://download.intel.com/museum/Moores_Law/ArticlesPress_releases/Gordon_Moore_1965_Article.pdf N.Yabumoto, K.Minegishi, Y.Komne, and K. S. (1990). WaterAdsorbed States on Silicon and Silicon Oxide Surfaces Analysed by Using Heavy Water. Jpn. J. Appl. Phys., 29, L490-L493.

97

Nakajima, K., Okazaki, Y., & Kimura, K. (2001). Initial oxidation process on Si(001) studied by high-resolution Rutherford backscattering spectroscopy. Physical Review B, 63(11), 18-21. doi:10.1103/PhysRevB.63.113314 Niwano, M. (1998). Oxidation processes on the H2O-chemisorbed Si(100) surface studied by in-situ infrared spectroscopy. Surf. Sci., 401(3), 364. doi:10.1016/S0039-6028(98)00023-5 Oh, J., Yeom, H., Hagimoto, Y., Ono, K., Oshima, M., Hirashita, N., Nywa, M., et al. (2001). Chemical structure of the ultrathin SiO2/Si(100) interface: An angle-resolved Si 2p photoemission study. Physical Review B, 63(20), 1-6. doi:10.1103/PhysRevB.63.205310 Ohno, S., Takizawa, J., Koizumi, J., Shudo, K., & Tanaka, M. (2007a). The temperature dependence of monolayer oxidation on Si(001)-(2 1) studied with surface differential reflectance spectroscopy. Journal of Physics: Condensed Matter, 19(44), 446011. doi:10.1088/09538984/19/44/446011 Ohno, S., Takizawa, J., Koizumi, J., Shudo, K., & Tanaka, M. (2007b). The temperature dependence of monolayer oxidation on Si(001)-(2 1) studied with surface differential reflectance spectroscopy. Journal of Physics: Condensed Matter, 19(44), 446011. doi:10.1088/09538984/19/44/446011 Ohta, H., Watanabe, T., & Ohdomari, I. (2007). Strain Distribution around SiO 2 /Si Interface in Si Nanowires: A Molecular Dynamics Study. Japanese Journal of Applied Physics, 46(5B), 3277-3282. doi:10.1143/JJAP.46.3277 Ohta, H., Watanabe, T., & Ohdomari, I. (2008). Potential energy landscape of an interstitial O2 molecule in a SiO2 film near the SiO2/Si(001) interface. Phys. Rev. B, 78(15), 3-4. doi:10.1103/PhysRevB.78.155326 Pamungkas, M. A., Joe, M., Kim, B. H., & Lee, K. R. (2011a). Reactive Molecular Dynamics Simulation of Early Stage of Dry Oxidation of Si ( 001 ) Surface. J. Appl. Phys, 110(5), 053513. doi:10.1063/1.3632968 Pamungkas, M. A., Joe, M., Kim, B.-hyun, & Lee, K.-ryeol. (2011b). Reactive Molecular Dynamics Simulation of Early Stage of Dry Oxidation of Si ( 001 ) Surface. J. Appl. Phys, 110(001), 053513. Retrieved from http://jap.aip.org/resource/1/japiau/v110/i5/p053513_s1 98

Pasquarello, A., Hybertsen, M., & Car, R. (1996). Theory of Si 2p core-level shifts at the Si(001)-SiO2 interface. Physical Review B, 53(16), 1094210950. doi:10.1103/PhysRevB.53.10942 Pasquarello, Alfredo, Mark S Hybertsen, R. C. (1998). Interface structure between silicon and its oxide by first principle molecular dynamics. Nature, 396(November), 8-10. Plimpton, S. (1995). Fast Parallel Algorithms for Short Range Molecular Dynamics. J. Comput. Phys, 117(June 1994), 1. R-L. Lo, I-S.Hwang, T. T. T. (2003). Complete dissociation of water on hot silicon ( 1 1 1 ) 7 7 surface direct observation of hopping oxygen atom. Surf. Sci., 530, L302-L306. doi:10.1016/S0039-6028(03)00330-3 R.M. Van Ginhoven *, H. P. H. (2007). No Title. Nuclear Instruments and Methods in Physics Research B, 255, 183. Ranke, W. (1996). Precursor kinetics of dissociative water adsorption on the Si(001) surface. Surf. Sci., 369(1-3), 137-145. doi:10.1016/S00396028(96)00912-0 Richard, N., Esteve, a, & Djafarirouhani, M. (2005). Density functional theory investigation of molecular oxygen interacting with Si(100)-(21). Computational Materials Science, 33(1-3), 26-30. doi:10.1016/j.commatsci.2004.12.023 Richard, N., Estve, A., & Djafari-Rouhani, M. (2005). Density functional theory investigation of molecular oxygen interacting with Si(100)-(21). Comput. Mater. Sci, 33(1-3), 26-30. doi:10.1016/j.commatsci.2004.12.023 Sarnthein, ., Pasquarello, A., Review, 52(17), 690-695. Car, . (1 5). 12 6 0 1 5. Physical

Schulze, R. (1994). Room-temperature water adsorption on the Si(100) surface examined by UPS, XPS, and static SIMS. Appl. Surf. Sci, 81(4), 449-463. doi:10.1016/0169-4332(94)90050-7 Stefanov, B. B., & Raghavachari, K. (1998). Pathways for initial waterinduced oxidation of Si(100). Appl. Phys. Lett, 73(6), 824. doi:10.1063/1.122013 99

Stillinger, F., & Weber, T. (1985). Computer simulation of local order in condensed phases of silicon. Physical review. B, Condensed matter, 31(2), 5262. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9938428 Suemitsu, M., Enta, Y., Miyanishi, Y., & Miyamoto, N. (1999). Initial Oxidation of Si(100)- (21) as an Autocatalytic Reaction. Physical Review Letters, 82(11), 2334-2337. doi:10.1103/PhysRevLett.82.2334 T. Watanabe K. Tatsumura, I. O. (2004). No Title. Applied Surface Science, 237, 125. Takakuwa, Y., Ishida, F., & Kawawa, T. (2003). Phase transition from Langmuir-type adsorption to two-dimensional oxide island growth during oxidation on Si(0 0 1) surface. Applied Surface Science, 216(1-4), 133-140. doi:10.1016/S0169-4332(03)00501-4 Tao, M., Shanmugam, J., Coviello, M., & Kirk, W. P. (2004). Suppression of silicon (001) surface reactivity using a valence-mending technique. Solid State Communications, 132(2), 89-92. doi:10.1016/j.ssc.2004.07.031 Tersoff, J. (1988). New empirical approach for the structure and energy of covalent systems. Physical review. B, Condensed matter, 37(12), 69917000. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9943969 Uchiyama, T, & Tsukada, M. (1996). Atomic and electronic structures of oxygen-adsorbed Si(001) surfaces. Physical review. B, Condensed matter, 53(12), 7917-7922. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/9982245 Uchiyama, Toshihiro, & Tsukada, M. (1997). Scanning-tunneling-microscopy images of oxygen adsorption on the Si001 surface. Physical Review B, 55(15), 9356-9359. Warren, R. L. M. and B. E. (1969). The structure of vitreous silica. Journal of Applied Crystallography, 2, 164. Warschkow, O., Schofield, S. R., Marks, N. A., Radny, M. W., Smith, P. V., & Mckenzie, D. R. (2008). Water on silicon ( 001 ): C defects and initial steps of surface oxidation. Phys. Rev. B, 77(001), 201305. doi:10.1103/PhysRevB.77.201305

100

Watanabe, H., Kato, K., Uda, T., Fujita, K., Ichikawa, M., Kawamura, T., & Terakura, K. (1998). Kinetics of Initial Layer-by-Layer Oxidation of Si(001) Surfaces. Phys. Rev. Lett, 80(2), 345-348. doi:10.1103/PhysRevLett.80.345 Watanabe, T. (2004). SiO2/Si interface structure and its formation studied by large-scale molecular dynamics simulation. Applied Surface Science, 237(1-4), 125-133. doi:10.1016/j.apsusc.2004.06.044 Watanabe, Takanobu, & Ohdomari, I. (1999). Modeling of SiO2/Si( 100 ) interface structure by using extended -Stillinger-Weber potential. Thin Solid Films, 343, 370. Weldon, M. K., Queeney, K. T., Gurevich, A. B., Stefanov, B. B., & Chabal, Y. J. (2000). Si H bending modes as a probe of local chemical structure: Thermal and chemical routes to decomposition of H2O on Si ( 100 ) - ( 2 1 ) Si H bending modes as a probe of local chemical structure: Thermal and chemical routes to decomposition of H 2 O. J. Chem. Phys, 113, 2440. doi:10.1063/1.482061 X,-L.Zhou, C.R.Flores, and White, J. . (1992). Adsorption and decomposition of water on Si(100): aTPD and SSIMS study. Appl. Surf. Sci, 62, 223237. Xu, Y.-nian. (1991). Electronic and optical properties of all polymorphic forms of silicon dioxid. Phys. Rev. B, 44(20), 48-59. Y. J. Chabal Krishnan Raghavachari, X. Z., & Garfunkel, E. (2002). No Title. Phys Rev B, 66, 161315. Y. P. Li and W. Y.Ching. (1985). Band structures of all polycrystalline forms of silicon dioxide. Phys. Rev. B, 31(4), 2172-2179. Y. Takata, K. Tamasaku, T. Tokushima, D. Miwa, S. Shin, T. Ishikawa, M. Yabashi, K. Kobayashi, J. J. Kim, T. Yao, T. Yamamoto, M. Arita, H. Namatame, and M. T. (2004). A probe of intrinsic valence band electronic structure: Hard x-ray photoemission. Appl. Phys. Lett, 84, 4310. Y.J.Chabal. (1984). Hydride formation on the Si(100):H2O surface. Phys. Rev. B, 29(6), 3677-3680.

101

Yasuda, T., Kumagai, N., Nishizawa, M., Yamasaki, S., Oheda, H., & Yamabe, K. (2003). Layer-resolved kinetics of Si oxidation investigated using the reflectance difference oscillation method. Phys. Rev. B, 67(19), 195338. doi:10.1103/PhysRevB.67.195338 Yeo, J.-N., Jee, G., Yu, B., Kim, H., Chung, C.-H., Yeom, H., Lyo, I.-W., et al. (2007). Ab initio study of the oxidation on vicinal Si(001) surfaces: The step-selective oxidation. Physical Review B, 76(11), 1-8. doi:10.1103/PhysRevB.76.115317 Yeom, H. W., Hamamatsu, H., Ohta, T., & Uhrberg, R. I. G. (1999). Highresolution core-level study of initial oxygen adsorption on Si 001 : and anomalous Si 2p core-level shifts. Physical Review B, 59(16), 413416. Yeom, H. W. Y., & Uhrberg, R. (1998). High Resolution Photoemission Study of Low-temperature Oxidation on the Si ( 001 ) Surface. Japan Journal of Applied Physics, 39(7), 4460-4463. Yin, M. T., & Cohen, M. L. (1982). No Title. Phys.Rev.B, 26, 5668. Yu, J., Sinnott, S. B., & Phillpot, S. R. (2007). Charge optimized many-body potential for the Si / SiO 2 system, (December 2006), 1-13. doi:10.1103/PhysRevB.75.085311 Yu, S.-yong, Kim, H., & Koo, J.-yong. (2008). Extrinsic Nature of Point efects on the Si ( 001 ) Surface: issociated Water Molecules. Phys. Rev. Lett, 100(JANUARY), 036107. doi:10.1103/PhysRevLett.100.036107 Yu, S.-yong, Kim, Y.-sung, Kim, H., & Koo, J.-yong. (2011). Influence of Flipping Si Dimers on the Dissociation Pathways of Water Molecules on Si ( 001 ). J.Phys. Chem C, 115(001), 24800-24803. Zhang, Q., Roland, C., Boguslawski, P., & Bernholc, J. (1995). Ab initio studies of the diffusion barriers at single-height Si(100) steps. Physical review letters, 75(1), 101-104. Retrieved from http://www.ncbi.nlm.nih.gov/pubmed/10059125 van Duin, A. C. T., Dasgupta, S., Lorant, F., & Goddard, W. a. (2001). ReaxFF: A Reactive Force Field for Hydrocarbons. The Journal of Physical Chemistry A, 105(41), 9396-9409. doi:10.1021/jp004368u

102

103

Vous aimerez peut-être aussi