From 5c34f34459cff1df7493614a014d1e594f47a9b6 Mon Sep 17 00:00:00 2001 From: hackbard Date: Wed, 25 May 2011 14:45:58 +0200 Subject: [PATCH] finished inc temp sims ... --- posic/thesis/ack.tex | 2 +- posic/thesis/defects.tex | 117 ++++---- posic/thesis/md.tex | 493 +++++++++++++++---------------- posic/thesis/simulation.tex | 1 + posic/thesis/summary_outlook.tex | 2 +- posic/thesis/thesis.tex | 10 +- 6 files changed, 309 insertions(+), 316 deletions(-) diff --git a/posic/thesis/ack.tex b/posic/thesis/ack.tex index bfe39bf..479bd37 100644 --- a/posic/thesis/ack.tex +++ b/posic/thesis/ack.tex @@ -5,7 +5,7 @@ %Although Mr. Stritzker is doing experimental physics in Augsburg he gave me the opportunity to do this more or less theoretical work. %During my stays in Finland Mr. Nordlund \ldots -Thanks to \ldots +Thanks to \ldots\\[2cm] \underline{Augsburg} \begin{itemize} diff --git a/posic/thesis/defects.tex b/posic/thesis/defects.tex index f1d6d7f..a69cfbb 100644 --- a/posic/thesis/defects.tex +++ b/posic/thesis/defects.tex @@ -18,7 +18,7 @@ Respective results allow to draw conclusions concerning the SiC precipitation in For investigating the \si{} structures a Si atom is inserted or removed according to Fig. \ref{fig:basics:ins_pos} of section \ref{section:basics:defects}. The formation energies of \si{} configurations are listed in Table \ref{tab:defects:si_self} for both methods used in this work as well as results obtained by other {\em ab initio} studies \cite{al-mushadani03,leung99}. -\begin{table}[t] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c} \hline @@ -39,7 +39,7 @@ Ref. \cite{leung99} & 3.31 & 3.31 & 3.43 & - & - \\ \caption[Formation energies of Si self-interstitials in crystalline Si determined by classical potential MD and DFT calculations.]{Formation energies of Si self-interstitials in crystalline Si determined by classical potential MD and DFT calculations. The formation energies are given in eV. T denotes the tetrahedral and H the hexagonal interstitial configuration. V corresponds to the vacancy configuration. Dumbbell configurations are abbreviated by DB. Formation energies for unstable configurations are marked by an asterisk and determined by using the low kinetic energy configuration shortly before the relaxation into the more favorable configuration starts.} \label{tab:defects:si_self} \end{table} -\begin{figure}[t] +\begin{figure}[tp] \begin{center} \begin{flushleft} \begin{minipage}{5cm} @@ -107,7 +107,7 @@ The \si{} atom then begins to slowly move towards an energetically more favorabl The formation energy of \unit[3.96]{eV} for this type of interstitial is equal to the result for the hexagonal one in the original work \cite{albe_sic_pot}. Obviously the authors did not carefully check the relaxed results assuming a hexagonal configuration. In Fig. \ref{fig:defects:kin_si_hex} the relaxation process is shown on the basis of the kinetic energy plot. -\begin{figure}[t] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{e_kin_si_hex.ps} \end{center} @@ -120,7 +120,7 @@ In fact, the same type of interstitial arises using random insertions. In addition, variations exist, in which the displacement is only along two \hkl<1 0 0> axes ($E_\text{f}=3.8\,\text{eV}$) or along a single \hkl<1 0 0> axes ($E_\text{f}=3.6\,\text{eV}$) successively approximating the tetdrahedral configuration and formation energy. The existence of these local minima located near the tetrahedral configuration seems to be an artifact of the analytical potential without physical authenticity revealing fundamental problems of analytical potential models for describing defect structures. However, the energy barrier is small. -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{nhex_tet.ps} \end{center} @@ -145,8 +145,6 @@ The length of these bonds are, however, close to the cut-off range and thus are The same applies to the bonds between the interstitial and the upper two atoms in the \si{} \hkl<1 1 0> DB configuration. A more detailed description of the chemical bonding is achieved through quantum-mechanical calculations by investigating the accumulation of negative charge between the nuclei. -%\clearpage{} - \section{Carbon point defects in silicon} \subsection{Defect structures in a nutshell} @@ -158,7 +156,7 @@ Again, the displayed structures are the results obtained by the classical potent The type of reservoir of the C impurity to determine the formation energy of the defect is chosen to be SiC. This is consistent with the methods used in the articles \cite{tersoff90,dal_pino93}, which the results are compared to in the following. Hence, the chemical potential of Si and C is determined by the cohesive energy of Si and SiC as discussed in section \ref{section:basics:defects}. -\begin{table}[t] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c c} \hline @@ -178,7 +176,7 @@ Hence, the chemical potential of Si and C is determined by the cohesive energy o \caption[Formation energies of C point defects in c-Si determined by classical potential MD and DFT calculations.]{Formation energies of C point defects in c-Si determined by classical potential MD and DFT calculations. The formation energies are given in eV. T denotes the tetrahedral, H the hexagonal and BC the bond-centered interstitial configuration. S corresponds to the substitutional interstitial configuration. The dumbbell configurations are abbreviated by DB. Formation energies for unstable configurations are marked by an asterisk and are determined by using the low kinetic energy configuration shortly before the relaxation into the more favorable configuration starts.} \label{tab:defects:c_ints} \end{table} -\begin{figure}[t] +\begin{figure}[tp] \begin{center} \begin{flushleft} \begin{minipage}{4cm} @@ -297,14 +295,14 @@ The structure was initially suspected by IR local vibrational mode absorption \c Fig. \ref{fig:defects:100db_cmp} schematically shows the \ci{} \hkl<1 0 0> DB structure and Table \ref{tab:defects:100db_cmp} lists the details of the atomic displacements, distances and bond angles obtained by classical potential and quantum-mechanical calculations. For comparison, the obtained structures for both methods are visualized in Fig. \ref{fig:defects:100db_vis_cmp}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=12cm]{100-c-si-db_cmp.eps} \end{center} \caption[Sketch of the \ci{} \hkl<1 0 0> dumbbell structure.]{Sketch of the \ci{} \hkl<1 0 0> dumbbell structure. Atomic displacements, distances and bond angles are listed in Table \ref{tab:defects:100db_cmp}.} \label{fig:defects:100db_cmp} \end{figure} -\begin{table}[ht] +\begin{table}[tp] \begin{center} Displacements\\ \begin{tabular}{l c c c c c c c c c} @@ -348,7 +346,7 @@ Angles\\ \caption[Atomic displacements, distances and bond angles of the \ci{} \hkl<1 0 0> DB structure obtained by {\textsc posic} and {\textsc vasp} calculations.]{Atomic displacements, distances and bond angles of the \ci{} \hkl<1 0 0> DB structure obtained by {\textsc posic} and {\textsc vasp} calculations. The displacements and distances are given in nm and the angles are given in degrees. Displacements, distances and angles are schematically displayed in Fig. \ref{fig:defects:100db_cmp}. In addition, the equilibrium lattice constant for crystalline Si is listed.} \label{tab:defects:100db_cmp} \end{table} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \begin{minipage}{6cm} \begin{center} @@ -366,7 +364,7 @@ Angles\\ \caption{Comparison of the \ci{} \hkl<1 0 0> DB structures obtained by {\textsc posic} and {\textsc vasp} calculations.} \label{fig:defects:100db_vis_cmp} \end{figure} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[height=10cm]{c_pd_vasp/eden.eps} \includegraphics[height=12cm]{c_pd_vasp/100_2333_ksl.ps} @@ -400,7 +398,7 @@ However, strictly speaking, the Kohn-Sham levels and orbitals do not have a dire \subsection{Bond-centered interstitial configuration} \label{subsection:bc} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \begin{minipage}{8cm} \includegraphics[width=8cm]{c_pd_vasp/bc_2333.eps}\\ @@ -433,8 +431,6 @@ The blue torus, which reinforces the assumption of the $p$ orbital, illustrates In addition, the energy level diagram shows a net amount of two spin up electrons. % todo smaller images, therefore add mo image -\clearpage{} - % todo migration of \si{}! \section{Migration of the carbon interstitial} @@ -444,7 +440,7 @@ A measure for the mobility of interstitial C is the activation energy necessary The stable defect geometries have been discussed in the previous subsection. In the following, the problem of interstitial C migration in Si is considered. Since the \ci{} \hkl<1 0 0> DB is the most probable, hence, most important configuration, the migration of this defect atom from one site of the Si host lattice to a neighboring site is in the focus of investigation. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \begin{minipage}{15cm} \underline{\hkl<0 0 -1> $\rightarrow$ \hkl<0 0 1>}\\ @@ -525,7 +521,7 @@ The bond to the face-centered Si atom at the bottom of the unit cell breaks and \subsection{Migration paths obtained by first-principles calculations} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=13cm]{im_00-1_nosym_sp_fullct_thesis.ps}\\[1.5cm] \begin{picture}(0,0)(150,0) @@ -552,7 +548,7 @@ To reach the BC configuration, which is \unit[0.94]{eV} higher in energy than th This amount of energy is needed to break the bond of the C atom to the Si atom at the bottom left. In a second process \unit[0.25]{eV} of energy are needed for the system to revert into a \hkl<1 0 0> configuration. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{00-1_0-10_vasp_s.ps} %\includegraphics[width=13cm]{vasp_mig/00-1_0-10_nosym_sp_fullct.ps}\\[1.6cm] @@ -579,7 +575,7 @@ In a second process \unit[0.25]{eV} of energy are needed for the system to rever Fig. \ref{fig:defects:00-1_0-10_mig} shows the migration barrier and structures of the \ci{} \hkl<0 0 -1> to \hkl<0 -1 0> DB transition. The resulting migration barrier of approximately \unit[0.9]{eV} is very close to the experimentally obtained values of \unit[0.70]{eV} \cite{lindner06}, \unit[0.73]{eV} \cite{song90} and \unit[0.87]{eV} \cite{tipping87}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=13cm]{vasp_mig/00-1_ip0-10_nosym_sp_fullct.ps}\\[1.8cm] \begin{picture}(0,0)(140,0) @@ -612,7 +608,7 @@ Slightly increased values compared to experiment might be due to the tightend co Nevertheless, the theoretical description performed in this work is improved compared to a former study \cite{capaz94}, which underestimates the experimental value by \unit[35]{\%}. In addition, it is finally shown that the BC configuration, for which spin polarized calculations are necessary, constitutes a real local minimum instead of a saddle point configuration due to the presence of restoring forces for displacements in migration direction. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{vasp_mig/110_mig_vasp.ps} %\begin{picture}(0,0)(140,0) @@ -651,7 +647,7 @@ For this reason, the assumption that C diffusion and reorientation is achieved b %Due to applying updated constraints on all atoms the obtained migration barriers and pathes might be overestimated and misguided. %To reinforce the applicability of the employed technique the obtained activation energies and migration pathes for the \hkl<0 0 -1> to \hkl<0 -1 0> transition are compared to two further migration calculations, which do not update the constrainted direction and which only apply updated constraints on three selected atoms, that is the diffusing C atom and the Si dumbbell pair in the initial and final configuration. %Results are presented in figure \ref{fig:defects:00-1_0-10_cmp}. -%\begin{figure}[ht] +%\begin{figure}[tp] %\begin{center} %\includegraphics[width=13cm]{vasp_mig/00-1_0-10_nosym_sp_cmp.ps} %\end{center} @@ -668,7 +664,7 @@ For this reason, the assumption that C diffusion and reorientation is achieved b \subsection{Migration described by classical potential calculations} \label{subsection:defects:mig_classical} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{bc_00-1_albe_s.ps} %\includegraphics[width=13cm]{bc_00-1.ps}\\[5.6cm] @@ -730,7 +726,7 @@ If the entire transition of the \hkl<0 0 -1> into the \hkl<0 0 1> configuration Assuming equal preexponential factors for both diffusion steps, the total probability of diffusion is given by $\exp\left((2.2\,\text{eV}+0.5\,\text{eV})/k_{\text{B}}T\right)$. Thus, the activation energy should be located within the range of \unit[2.2-2.7]{eV}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=13cm]{00-1_0-10.ps}\\[2.4cm] \begin{pspicture}(0,0)(0,0) @@ -756,7 +752,7 @@ Thus, the activation energy should be located within the range of \unit[2.2-2.7] % red: ./visualize -w 640 -h 480 -d saves/c_in_si_mig_00-1_0-10_s20 -nll -0.56 -0.56 -0.8 -fur 0.3 0.2 0 -c -0.125 -1.7 0.7 -L -0.125 -0.25 -0.25 -r 0.6 -B 0.1 \label{fig:defects:cp_00-1_0-10_mig} \end{figure} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{00-1_ip0-10.ps} \end{center} @@ -776,7 +772,7 @@ As mentioned above, the starting configuration of the first migration path, i.e. Further relaxation of the BC configuration results in the \ci{} \hkl<1 1 0> configuration. Even the last two pathways show configurations almost identical to the \ci{} \hkl<1 1 0> configuration, which constitute local minima within the pathways. Thus, migration pathways involving the \ci{} \hkl<1 1 0> DB configuration as a starting or final configuration are further investigated. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{110_mig.ps} \end{center} @@ -792,7 +788,7 @@ In contrast to quantum-mechanical calculations, in which the direct transition i Thus the just proposed migration path, which involves the \hkl<1 1 0> interstitial configuration, becomes even more probable than the initially porposed path, which involves the BC configuration that is, in fact, unstable. Due to these findings, the respective path is proposed to constitute the diffusion-describing path. The evolution of structure and configurational energy is displayed again in Fig. \ref{fig:defects:involve110}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{00-1_110_0-10_mig_albe.ps} \end{center} @@ -814,15 +810,13 @@ Thus, atomic diffusion is wrongly described in the classical potential approach. The probability of already rare diffusion events is further decreased for this reason. However, agglomeration of C and diffusion of Si self-interstitials are an important part of the proposed SiC precipitation mechanism. Thus, a serious limitation that has to be taken into account for appropriately modeling the C/Si system using the otherwise quite promising EA potential is revealed. -Possible workarounds are discussed in more detail in section \ref{subsection:md:limit}. - -\clearpage{} +Possible workarounds are discussed in more detail in section \ref{section:md:limit}. \section{Combination of point defects} The study proceeds with a structural and energetic investigation of pairs of the ground-state and, thus, most probable defect configurations that are believed to be fundamental in the Si to SiC conversion. Investigations are restricted to quantum-mechanical calculations. -\begin{figure}[t] +\begin{figure}[tp] \begin{center} \subfigure[]{\label{fig:defects:combos_ci}\includegraphics[width=0.3\textwidth]{combos_ci.eps}} \hspace{0.5cm} @@ -839,7 +833,7 @@ Next to formation and binding energies, migration barriers are investigated, whi \label{subsection:defects:c-si_comb} \ci{} pairs of the \hkl<1 0 0>-type are investigated in the first part. -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c c} \hline @@ -873,7 +867,7 @@ Energetically favorable and unfavorable configurations can be explained by stres Antiparallel orientations of the second defect, i.e. \hkl[0 0 1] for positions located below the \hkl(0 0 1) plane with respect to the initial one (positions 1, 2 and 4) form the energetically most unfavorable configurations. In contrast, the parallel and particularly the twisted orientations constitute energetically favorable configurations, in which a vast reduction of strain is enabled by combination of these defects. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \subfigure[\underline{$E_{\text{b}}=-2.25\,\text{eV}$}]{\label{fig:defects:225}\includegraphics[width=0.3\textwidth]{00-1dc/2-25.eps}} \hspace{0.5cm} @@ -902,7 +896,7 @@ However, the second most favorable configuration ($E_{\text{f}}=-2.25\,\text{eV} Thus, particularly at high temepratures that cause an increase of the entropic contribution, this structure constitutes a serious opponent to the ground state. In fact, following results on migration simulations will reinforce the assumption of a low probability for C clustering by thermally activated processes. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \subfigure[\underline{$E_{\text{b}}=-2.16\,\text{eV}$}]{\label{fig:defects:216}\includegraphics[width=0.25\textwidth]{00-1dc/2-16.eps}} \hspace{0.2cm} @@ -947,7 +941,7 @@ Both configurations are unfavorable compared to far-off, isolated DBs. Nonparallel orientations, i.e. the \hkl[0 1 0], \hkl[0 -1 0] and its equivalents, result in binding energies of \unit[-0.12]{eV} and \unit[-0.27]{eV}, thus, constituting energetically favorable configurations. The reduction of strain energy is higher in the second case, where the C atom of the second DB is placed in the direction pointing away from the initial C atom. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \subfigure[\underline{$E_{\text{b}}=-1.53\,\text{eV}$}]{\label{fig:defects:153}\includegraphics[width=0.25\textwidth]{00-1dc/1-53.eps}} \hspace{0.7cm} @@ -977,7 +971,7 @@ In both configurations, the far-off atom of the second DB resides in threefold c The interaction of \ci{} \hkl<1 0 0> DBs is investigated along the \hkl[1 1 0] bond chain assuming a possible reorientation of the DB atom at each position to minimize its configurational energy. Therefore, the binding energies of the energetically most favorable configurations with the second DB located along the \hkl[1 1 0] direction and resulting C-C distances of the relaxed structures are summarized in Table~\ref{tab:defects:comb_db110}. -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c c} \hline @@ -995,7 +989,7 @@ Type & \hkl[-1 0 0] & \hkl[1 0 0] & \hkl[1 0 0] & \hkl[1 0 0] & \hkl[1 0 0] & \h \label{tab:defects:comb_db110} \end{table} % -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{db_along_110_cc_n.ps} \end{center} @@ -1018,7 +1012,7 @@ The high activation energy is attributed to the stability of such a low energy c Low barriers are only identified for transitions starting from energetically less favorable configurations, e.g. the configuration of a \hkl[-1 0 0] DB located at position 2 (\unit[-0.36]{eV}). Starting from this configuration, an activation energy of only \unit[1.2]{eV} is necessary for the transition into the ground state configuration. The corresponding migration energies and atomic configurations are displayed in Fig.~\ref{fig:036-239}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{036-239.ps} \end{center} @@ -1037,7 +1031,7 @@ In both cases the configuration yielding a binding energy of \unit[-2.25]{eV} is First of all, it constitutes the second most energetically favorable structure. Secondly, a migration path with a barrier as low as \unit[0.47]{eV} exists starting from a configuration of largely separated defects exhibiting a low binding energy (\unit[-1.88]{eV}). The migration barrier and corresponding structures are shown in Fig.~\ref{fig:188-225}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{188-225.ps} \end{center} @@ -1051,7 +1045,7 @@ As a result, C defect agglomeration indeed is expected, but only a low probabili \subsection[Combinations of the \ci{} \hkl<1 0 0> and \cs{} type]{\boldmath Combinations of the \ci{} \hkl<1 0 0> and \cs{} type} \label{subsection:defects:c-csub} -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{c c c c c c} \hline @@ -1067,7 +1061,7 @@ As a result, C defect agglomeration indeed is expected, but only a low probabili \caption{Binding energies of combinations of the \ci{} \hkl[0 0 -1] defect with a \cs{} atom located at positions 1 to 5 according to Fig.~\ref{fig:defects:combos_ci}. R corresponds to the position located at $\frac{a_{\text{Si}}}{2}\hkl[3 2 3]$ relative to the initial defect position, which is the maximum realizable distance due to periodic boundary conditions.} \label{tab:defects:c-s} \end{table} -%\begin{figure}[ht] +%\begin{figure}[tp] %\begin{center} %\begin{minipage}[t]{5cm} %a) \underline{Pos: 1, $E_{\text{b}}=0.26\text{ eV}$} @@ -1091,7 +1085,7 @@ As a result, C defect agglomeration indeed is expected, but only a low probabili %\caption{Relaxed structures of defect complexes obtained by creating a carbon substitutional at position 1 (a)), 3 (b)) and 5 (c)).} %\label{fig:defects:comb_db_04} %\end{figure} -%\begin{figure}[ht] +%\begin{figure}[tp] %\begin{center} %\begin{minipage}[t]{7cm} %a) \underline{Pos: 2, $E_{\text{b}}=-0.51\text{ eV}$} @@ -1117,7 +1111,7 @@ Fig.~\ref{fig:093-095} and \ref{fig:026-128} show structures A, B and a, b toget % A B %./visualize_contcar -w 640 -h 480 -d results/c_00-1_c3_csub_B -nll -0.20 -0.4 -0.1 -fur 0.9 0.6 0.9 -c 0.5 -1.5 0.375 -L 0.5 0 0.3 -r 0.6 -A -1 2.465 -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{093-095.ps} \end{center} @@ -1139,7 +1133,7 @@ Obviously, either the CRT algorithm fails to seize the actual saddle point struc % not satisfactory! % a b -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{026-128.ps} \end{center} @@ -1179,7 +1173,7 @@ This structure is followed by C$_{\text{s}}$ located at position 2, the lattice As mentioned earlier, these two lower Si atoms indeed experience tensile strain along the \hkl[1 1 0] bond chain, however, additional compressive strain along \hkl[0 0 1] exists. The latter is partially compensated by the C$_{\text{s}}$ atom. Yet less of compensation is realized if C$_{\text{s}}$ is located at position 4 due to a larger separation although both bottom Si atoms of the DB structure are indirectly affected, i.e. each of them is connected by another Si atom to the C atom enabling the reduction of strain along \hkl[0 0 1]. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \subfigure[\underline{$E_{\text{b}}=-0.51\,\text{eV}$}]{\label{fig:defects:051}\includegraphics[width=0.25\textwidth]{00-1dc/0-51.eps}} \hspace{0.2cm} @@ -1232,8 +1226,9 @@ For the same reasons as in the last subsection, structures other than the ground In the last section, configurations of a C$_{\text{i}}$ DB with C$_{\text{s}}$ occupying a vacant site have been investigated. Additionally, configurations might arise in IBS, in which the impinging C atom creates a vacant site near a C$_{\text{i}}$ DB, but does not occupy it. These structures are investigated in the following. + Resulting binding energies of a C$_{\text{i}}$ DB and a nearby vacancy are listed in the second row of Table~\ref{tab:defects:c-v}. -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{c c c c c c} \hline @@ -1248,7 +1243,7 @@ Resulting binding energies of a C$_{\text{i}}$ DB and a nearby vacancy are liste \caption{Binding energies of combinations of the \ci{} \hkl[0 0 -1] defect with a vacancy located at positions 1 to 5 according to Fig.~\ref{fig:defects:combos_ci}. R corresponds to the position located at $\frac{a_{\text{Si}}}{2}\hkl[3 2 3]$ relative to the initial defect position, which is the maximum realizable distance due to periodic boundary conditions.} \label{tab:defects:c-v} \end{table} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \subfigure[\underline{$E_{\text{b}}=-0.59\,\text{eV}$}]{\label{fig:defects:059}\includegraphics[width=0.25\textwidth]{00-1dc/0-59.eps}} \hspace{0.7cm} @@ -1289,14 +1284,14 @@ Strain reduced by this huge displacement is partially absorbed by tensile strain A binding energy of \unit[-0.50]{eV} is observed. The migration pathways of configuration \ref{fig:defects:314} and \ref{fig:defects:059} into the ground-state configuration, i.e. the \cs{} configuration, are shown in Fig.~\ref{fig:314-539} and \ref{fig:059-539} respectively. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{314-539.ps} \end{center} \caption{Migration barrier and structures of the transition of the initial C$_{\text{i}}$ \hkl[0 0 -1] DB and a V created at position 3 (left) into a C$_{\text{s}}$ configuration (right). An activation energy of \unit[0.1]{eV} is observed.} \label{fig:314-539} \end{figure} -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{059-539.ps} \end{center} @@ -1331,8 +1326,6 @@ Furthermore, small activation energies, even for transitions into the ground sta If the vacancy is created at position 1 the system will end up in a configuration of C$_{\text{s}}$ anyways. Based on these results, a high probability for the formation of C$_{\text{s}}$ must be concluded. -%\clearpage{} - \subsection{Combinations of \si{} and \cs} \label{subsection:si-cs} @@ -1344,7 +1337,7 @@ Thus, combinations of \cs{} and an additional \si{} are examined in the followin The ground-state of a single \si{} was found to be the \si{} \hkl<1 1 0> DB configuration. For the follwoing study the same type of self-interstitial is assumed to provide the energetically most favorable configuration in combination with \cs. -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c c} \hline @@ -1364,7 +1357,7 @@ For the follwoing study the same type of self-interstitial is assumed to provide \caption{Equivalent configurations labeled \RM{1}-\RM{10} of \hkl<1 1 0>-type Si$_{\text{i}}$ DBs created at position I and C$_{\text{s}}$ created at positions 1 to 5 according to Fig.~\ref{fig:defects:combos_si}. The respective orientation of the Si$_{\text{i}}$ DB is given in the first row.} \label{tab:defects:comb_csub_si110} \end{table} -\begin{table}[ht] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c c c c c c c c} \hline @@ -1397,7 +1390,7 @@ Thus, the compressive stress along \hkl[1 1 0] of the \si{} \hkl[1 1 0] DB is no However, even configuration \RM{1} is energetically less favorable than the \hkl<1 0 0> C$_{\text{i}}$ DB, which, thus, remains the ground state of a C atom introduced into otherwise perfect c-Si. The transition involving the latter two configurations is shown in Fig.~\ref{fig:162-097}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{162-097.ps} \end{center} @@ -1409,7 +1402,7 @@ Accordingly, the C$_{\text{i}}$ \hkl<1 0 0> DB configuration is assumed to occur However, only \unit[0.77]{eV} are needed for the reverse process, i.e. the formation of C$_{\text{s}}$ and a Si$_{\text{i}}$ DB out of the ground state. Due to the low activation energy this process must be considered to be activated without much effort either thermally or by introduced energy of the implantation process. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \includegraphics[width=0.7\textwidth]{c_sub_si110.ps} \end{center} @@ -1437,14 +1430,14 @@ Similar to what was previously mentioned, configurations of C$_{\text{s}}$ and a At higher temperatures, the contribution of entropy to structural formation increases, which might result in a spatial separation even for defects located within the capture radius. Indeed, an {em ab initio} MD run at \unit[900]{$^{\circ}$C} starting from configuration \RM{1}, which -- based on the above findings -- is assumed to recombine into the ground state configuration, results in a separation of the C$_{\text{s}}$ and Si$_{\text{i}}$ DB by more than 4 neighbor distances realized in a repeated migration mechanism of annihilating and arising Si$_{\text{i}}$ DBs. The atomic configurations for two different points in time are shown in Fig.~\ref{fig:defects:md}. -\begin{figure}[ht] +\begin{figure}[tp] \begin{center} \begin{minipage}{0.40\textwidth} -\includegraphics[width=\columnwidth]{md01.eps} +\includegraphics[width=\columnwidth]{md_vasp_01.eps} \end{minipage} \hspace{1cm} \begin{minipage}{0.40\textwidth} -\includegraphics[width=\columnwidth]{md02.eps}\\ +\includegraphics[width=\columnwidth]{md_vasp_02.eps} \end{minipage}\\ \begin{minipage}{0.40\textwidth} \begin{center} @@ -1468,7 +1461,7 @@ These results support the above assumptions of an increased entropic contributio %\section{Migration in systems of combined defects} -%\begin{figure}[ht] +%\begin{figure}[tp] %\begin{center} %\includegraphics[width=13cm]{vasp_mig/comb_mig_3-2_vac_fullct.ps}\\[2.0cm] %\begin{picture}(0,0)(170,0) @@ -1493,7 +1486,7 @@ These results support the above assumptions of an increased entropic contributio %\caption{Transition of the configuration of the C-Si dumbbell interstitial in combination with a vacancy created at position 2 into the configuration of substitutional carbon.} %\label{fig:defects:comb_mig_01} %\end{figure} -%\begin{figure}[ht] +%\begin{figure}[tp] %\begin{center} %\includegraphics[width=13cm]{vasp_mig/comb_mig_4-2_vac_fullct.ps}\\[1.0cm] %\begin{picture}(0,0)(150,0) @@ -1519,8 +1512,6 @@ These results support the above assumptions of an increased entropic contributio %\label{fig:defects:comb_mig_02} %\end{figure} -\clearpage{} - \section{Applicability: Competition of \ci{} and \cs-\si{}} \label{section:ea_app} @@ -1535,7 +1526,7 @@ Since quantum-mechanical calculations reveal the Si$_{\text{i}}$ \hkl<1 1 0> DB Empirical potentials, however, predict Si$_{\text{i}}$ T to be the energetically most favorable configuration. Thus, investigations of the relative energies of formation of defect pairs need to include combinations of C$_{\text{s}}$ with Si$_{\text{i}}$ T. Results of {\em ab initio} and classical potential calculations are summarized in Table~\ref{tab:defect_combos}. -\begin{table}[t] +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c} \hline @@ -1650,8 +1641,6 @@ In contrast, there is no obvious reason for the topotactic orientation of an agg \ifnum1=0 - - In addition, there are experimental findings, which might be exploited to reinforce the non-validity of the proposed precipitation model. High resolution TEM shows equal orientation of \hkl(h k l) planes of the c-Si host matrix and the 3C-SiC precipitate. diff --git a/posic/thesis/md.tex b/posic/thesis/md.tex index acb6d2a..0673823 100644 --- a/posic/thesis/md.tex +++ b/posic/thesis/md.tex @@ -1,63 +1,48 @@ \chapter{Silicon carbide precipitation simulations} \label{chapter:md} -The molecular dynamics (MD) technique is used to gain insight into the behavior of carbon existing in different concentrations in crystalline silicon on the microscopic level at finite temperatures. -Both, quantum-mechanical and classical potential molecular dynamics simulations are performed. -Since quantum-mechanical calculations are restricted to a few hundreds of atoms only small volumes composed of three unit cells in each direction and small carbon concentrations are simulated using the VASP code. -Thus, investigations are restricted to the diffusion process of single carbon interstitials and the agglomeration of a few dumbbell interstitials in silicon. -Using classical potentials volume sizes up to 31 unit cells in each direction and high carbon concentrations are realizable. -Simulations targeting the formation of silicon carbide precipitates are, thus, attempted in classical potential calculations only. +The molecular dynamics (MD) technique is used to gain insight into the behavior of C existing in different concentrations in c-Si on the microscopic level at finite temperatures. +The simulations are restricted to classical potential simulations utilizing the analytical EA bond order potential as described in section \ref{subsection:interact_pot}. +Parameters are chosen according to the discussion in section \ref{section:classpotmd}. -\section{Ab initio MD simulations} +At the beginning, simulations are performed, which try to mimic the conditions during IBS. +Results reveal limitations of the employed potential and MD in general. +With reference to the results of the last chapter, a workaround is discussed. +The approach is follwed and, finally, results gained by the MD simulations are interpreted drawing special attention to the established controversy concerning precipitation of SiC in Si. -No pressure control, since VASP does not support this feature in MD mode. -The time step is set to one fs. -Explain some more parameters that differ from the latter calculations ... +\section{Simulations at temperatures used in IBS} +\label{section:initial_sims} -Molecular dynamics simulations of a single, two and ten carbon atoms in $3\times 3\times 3$ unit cells of crytsalline silicon are performed. - -{\color{red}Todo: ... in progress ...} - -\section{Classical potential MD simulations} - -In contrast to the quantum-mechanical MD simulations the developed classical potential MD code is able to do constant pressure simulations using the Berendsen barostat. -The system pressure is set to zero pressure. -Due to promising advantages over the Tersoff potential the bond order potential of Erhart and Albe is used. -A time step of one fs is set. - -\subsection{Simulations at temperatures used in ion beam synthesis} -\label{subsection:initial_sims} - -In initial simulations aiming to reproduce a precipitation process simulation volumes of $31\times 31\times 31$ unit cells are utilized. +In initial simulations aiming to reproduce a precipitation process, simulation volumes of $31\times 31\times 31$ unit cells are utilized. Periodic boundary conditions in each direction are applied. -The system temperature is set to $450\, ^{\circ}\mathrm{C}$, the temperature for which epitaxial growth of 3C-SiC films is achieved by ion beam synthesis (IBS). -After equilibration of the kinetic and potential energy carbon atoms are consecutively inserted. -The number of carbon atoms $N_{\text{Carbon}}$ necessary to form a spherical precipitate with radius $r$ is given by +The system temperature is set to $450\, ^{\circ}\mathrm{C}$, the temperature for which epitaxial growth of 3C-SiC films is achieved in IBS. +After equilibration of the kinetic and potential energy, C atoms are consecutively inserted. +The number of C atoms $N_{\text{C}}$ necessary to form a spherical precipitate with radius $r$ is given by \begin{equation} - N_{\text{Carbon}}=\frac{4}{3}\pi r^3 \cdot \frac{4}{a_{\text{SiC}}^3} + N_{\text{C}}=\frac{4}{3}\pi r^3 \cdot \frac{4}{a_{\text{SiC}}^3} =\frac{16}{3} \pi \left( \frac{r}{a_{\text{SiC}}}\right)^3 \label{eq:md:spheric_prec} \end{equation} with $a_{\text{SiC}}$ being the lattice constant of 3C-SiC. -In IBS experiments the smallest precipitates observed have radii starting from 2 nm up to 4 nm. -For the initial simulations a total amount of 6000 carbon atoms corresponding to a radius of approximately 3.1 nm is chosen. -In separated simulations these 6000 carbon atoms are inserted in three regions of different volume ($V_1$, $V_2$, $V_3$) within the simulation cell. +In IBS experiments, the smallest precipitates observed have radii starting from \unit[2]{nm} up to \unit[4]{nm}. +For the initial simulations, a total amount of 6000 C atoms corresponding to a radius of approximately \unit[3.1]{nm} is chosen. +In separated simulations, the 6000 C atoms are inserted in three regions of different volume ($V_1$, $V_2$, $V_3$) within the simulation cell. For reasons of simplification these regions are rectangularly shaped. $V_1$ is chosen to be the total simulation volume. $V_2$ approximately corresponds to the volume of a minimal 3C-SiC precipitate. -$V_3$ is approximately the volume containing the necessary amount of silicon atoms to form such a precipitate, which is slightly smaller than $V_2$ due to the slightly lower silicon density of 3C-SiC compared to c-Si. -The two latter insertion volumes are considered since no diffusion of carbon atoms is expected within the simulated period of time at prevalent temperatures. -This is due to the overestimated activation energies for carbon diffusion as pointed out in section \ref{subsection:defects:mig_classical}. -For rectangularly shaped precipitates with side length $L$ the amount of carbon atoms in 3C-SiC and silicon atoms in c-Si is given by +$V_3$ is approximately the volume containing the amount of Si atoms necessary to form such a precipitate, which is slightly smaller than $V_2$ due to the slightly lower Si density of 3C-SiC compared to c-Si. +The two latter insertion volumes are considered since no diffusion of C atoms is expected within the simulated period of time at prevalent temperatures. +This is due to the overestimated activation energy for the diffusion of a \ci \hkl<1 0 0> DB, as pointed out in section \ref{subsection:defects:mig_classical}. +For rectangularly shaped precipitates with side length $L$ the amount of C atoms in 3C-SiC and Si atoms in c-Si is given by \begin{equation} - N_{\text{Carbon}}^{\text{3C-SiC}} =4 \left( \frac{L}{a_{\text{SiC}}}\right)^3 + N_{\text{C}}^{\text{3C-SiC}} =4 \left( \frac{L}{a_{\text{SiC}}}\right)^3 \text{ and} \quad - N_{\text{Silicon}}^{\text{c-Si}} =8 \left( \frac{L}{a_{\text{Si}}}\right)^3 \text{ .} + N_{\text{Si}}^{\text{c-Si}} =8 \left( \frac{L}{a_{\text{Si}}}\right)^3 \text{ .} \label{eq:md:n_prec} \end{equation} -Table \ref{table:md:ins_vols} summarizes the side length of each of the three different insertion volumes determined by equations \eqref{eq:md:n_prec} and the resulting carbon concentrations inside these volumes. -Looking at the carbon concentrations simulations can be distinguished in simulations occupying low ($V_1$) and high ($V_2$, $V_3$) concentrations of carbon. -\begin{table} +Table \ref{table:md:ins_vols} summarizes the side length of each of the three different insertion volumes determined by equations \eqref{eq:md:n_prec} and the resulting C concentrations inside these volumes. +Looking at the C concentrations, simulations can be distinguished in simulations occupying low ($V_1$) and high ($V_2$, $V_3$) concentrations of C. +\begin{table}[tp] \begin{center} \begin{tabular}{l c c c} \hline @@ -65,22 +50,23 @@ Looking at the carbon concentrations simulations can be distinguished in simulat & $V_1$ & $V_2$ & $V_3$ \\ \hline Side length [\AA] & 168.3 & 50.0 & 49.0 \\ -Carbon concentration [$\frac{1}{\text{c-Si unit cell}}$] & 0.20 & 7.68 & 8.16\\ +C concentration [$\frac{1}{\text{c-Si unit cell}}$] & 0.20 & 7.68 & 8.16\\ \hline \hline \end{tabular} \end{center} -\caption{Side lengthes of the insertion volumes $V_1$, $V_2$ and $V_3$ used for the incoorperation of 6000 carbon atoms.} +\caption{Side lengthes of the insertion volumes $V_1$, $V_2$ and $V_3$ used for the incoorperation of 6000 C atoms.} \label{table:md:ins_vols} \end{table} The insertion is realized in a way to keep the system temperature constant. -In each of 600 insertion steps 10 carbon atoms are inserted at random positions within the respective region, which involves an increase in kinetic energy. -Thus, the simulation is continued without adding more carbon atoms until the system temperature is equal to the chosen temperature again, which is realized by the thermostat decoupling excessive energy. -Every inserted carbon atom must exhibit a distance greater or equal than 1.5 \AA{} to present neighboured atoms to prevent too high forces to occur. -Once the total amount of carbon is inserted the simulation is continued for 100 ps followed by a cooling-down process until room temperature, that is $20\, ^{\circ}\mathrm{C}$ is reached. -Figure \ref{fig:md:prec_fc} displays a flow chart of the applied steps involved in the simulation sequence. -\begin{figure}[!ht] +In each of 600 insertion steps, 10 C atoms are inserted at random positions within the respective region, which involves an increase in potential energy that is partially transformed into kinetic energy. +Thus, the simulation is continued without adding more C atoms until the system temperature is equal to the chosen temperature again. +This is realized by the thermostat, which decouples excessive energy. +Every inserted C atom must exhibit a distance greater or equal to \unit[1.5]{\AA} to neighbored atoms to prevent the occurrence of too high forces. +Once the total amount of C is inserted, the simulation is continued for \unit[100]{ps} followed by a cooling-down process until room temperature, i.e. \unit[20]{$^{\circ}$C}, is reached. +Fig.~\ref{fig:md:prec_fc} displays a flow chart of the applied steps involved in the simulation sequence. +\begin{figure}[tp] \begin{center} \begin{pspicture}(0,0)(15,17) @@ -105,11 +91,11 @@ Figure \ref{fig:md:prec_fc} displays a flow chart of the applied steps involved \psframe*[linecolor=lbb](3,6.5)(11,11) \rput[lt](3.2,10.8){\color{gray}CARBON INSERTION} \rput(3,10.8){\pnode{CI}} - \rput(7,10){\rnode{9}{\psframebox{Insertion of 10 carbon aoms}}} + \rput(7,10){\rnode{9}{\psframebox{Insertion of 10 C aoms}}} \rput(7,9){\rnode{8}{\psframebox{Continue for 100 fs}}} \rput(7,8){\rnode{7}{\psframebox{$T_{\text{avg}}=T_{\text{s}} \pm1\,^{\circ}\mathrm{C}$}}} - \rput(7,7){\rnode{6}{\psframebox{$N_{\text{Carbon}}=6000$}}} + \rput(7,7){\rnode{6}{\psframebox{$N_{\text{C}}=6000$}}} \ncline[]{->}{9}{8} \ncline[]{->}{8}{7} \ncline[]{->}{7}{6} @@ -153,292 +139,308 @@ Figure \ref{fig:md:prec_fc} displays a flow chart of the applied steps involved \trput*{\scriptsize true} \end{pspicture} \end{center} -\caption[Flowchart of the simulation sequence used in molecular dnymaics simulations aiming to reproduce the precipitation process.]{Flowchart of the simulation sequence used in molecular dnymaics simulations aiming to reproduce the precipitation process. $T_{\text{s}}$ and $p_{\text{s}}$ are the preset values for the system temperature and pressure. $T_{\text{avg}}$ is the averaged actual system temperature.} +\caption[Flowchart of the simulation sequence used in MD simulations aiming to reproduce the precipitation process.]{Flowchart of the simulation sequence used in MD simulations aiming to reproduce the precipitation process. $T_{\text{s}}$ and $p_{\text{s}}$ are the preset values for the system temperature and pressure. $T_{\text{avg}}$ is the averaged actual system temperature.} \label{fig:md:prec_fc} \end{figure} -The radial distribution function $g(r)$ for C-C and Si-Si distances is shown in figure \ref{fig:md:pc_si-si_c-c}. -\begin{figure}[!ht] +The radial distribution function $g(r)$ for C-C and Si-Si distances is shown in Fig. \ref{fig:md:pc_si-si_c-c}. +\begin{figure}[tp] \begin{center} - \includegraphics[width=12cm]{sic_prec_450_si-si_c-c.ps} + \includegraphics[width=0.7\textwidth]{sic_prec_450_si-si_c-c.ps} \end{center} -\caption[Radial distribution function of the C-C and Si-Si distances for 6000 carbon atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of $450\,^{\circ}\mathrm{C}$ and cooled down to room temperature.]{Radial distribution function of the C-C and Si-Si distances for 6000 carbon atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of $450\,^{\circ}\mathrm{C}$ and cooled down to room temperature. The bright blue graph shows the Si-Si radial distribution for pure c-Si. The insets show magnified regions of the respective type of bond.} +\caption[Radial distribution function of the C-C and Si-Si distances for 6000 C atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of \unit\[450\]{$^{\circ}$C} and cooled down to room temperature.]{Radial distribution function of the C-C and Si-Si distances for 6000 C atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of \unit[450]{$^{\circ}$C} and cooled down to room temperature. The bright blue graph shows the Si-Si radial distribution for pure c-Si. The insets show magnified regions of the respective type of bond.} \label{fig:md:pc_si-si_c-c} \end{figure} -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} - \includegraphics[width=12cm]{sic_prec_450_energy.ps} + \includegraphics[width=0.7\textwidth]{sic_prec_450_energy.ps} \end{center} -\caption[Total energy per atom as a function of time for the whole simulation sequence and for all three types of insertion volumes.]{Total energy per atom as a function of time for the whole simulation sequence and for all three types of insertion volumes. Arrows mark the end of carbon insertion and the start of the cooling process respectively.} +\caption[Total energy per atom as a function of time for the whole simulation sequence and for all three types of insertion volumes.]{Total energy per atom as a function of time for the whole simulation sequence and for all three types of insertion volumes. Arrows mark the end of C insertion and the start of the cooling process respectively.} \label{fig:md:energy_450} \end{figure} -It is easily and instantly visible that there is no significant difference among the two simulations of high carbon concentration. -The first C-C peak appears at about 0.15 nm, which is compareable to the nearest neighbour distance of graphite or diamond. -The number of C-C bonds is much smaller for $V_1$ than for $V_2$ and $V_3$ since carbon atoms are spread over the total simulation volume. -These carbon atoms are assumed to form strong bonds. -This is supported by figure \ref{fig:md:energy_450} displaying the total energy of all three simulations during the whole simulation sequence. -A huge decrease of the total energy during carbon insertion is observed for the simulations with high carbon concentration in contrast to the $V_1$ simulation, which shows a slight increase. -The difference in energy $\Delta$ growing within the carbon insertion process up to a value of roughly 0.06 eV per atom persists unchanged until the end of the simulation. -The excess amount of next neighboured strongly bounded C-C bonds in the high concentration simulations make these configurations energetically more favorable compared to the low concentration configuration. -However, in the same way a lot of energy is needed to break these bonds to get out of the local energy minimum advancing towards the global minimum configuration. +It is easily and instantly visible that there is no significant difference among the two simulations of high C concentration. +Thus, in the following, the focus can indeed be directed to low ($V_1$) and high ($V_2$, $V_3$) C concentration simulations. +The first C-C peak appears at about \unit[0.15]{nm}, which is compareable to the nearest neighbor distance of graphite or diamond. +The number of C-C bonds is much smaller for $V_1$ than for $V_2$ and $V_3$ since C atoms are spread over the total simulation volume. +On average, there are only 0.2 C atoms per Si unit cell. +These C atoms are assumed to form strong bonds. +This is supported by Fig.~\ref{fig:md:energy_450} displaying the total energy of all three simulations during the whole simulation sequence. +A huge decrease of the total energy during C insertion is observed for the simulations with high C concentration in contrast to the $V_1$ simulation, which shows a slight increase. +The difference in energy $\Delta$ growing within the C insertion process up to a value of roughly \unit[0.06]{eV} per atom persists unchanged until the end of the simulation. +The vast amount of strongly bonded C-C bonds in the high concentration simulations make these configurations energetically more favorable compared to the low concentration configuration. +However, in the same way, a lot of energy is needed to break these bonds to get out of the local energy minimum advancing towards the global minimum configuration. Thus, such conformational changes are very unlikely to happen. -This is in accordance with the constant total energy observed in the continuation step of 100 ps inbetween the end of carbon insertion and the cooling process. -Obviously no energetically favorable relaxation is taking place at a system temperature of $450\,^{\circ}\mathrm{C}$. - -The C-C peak at about 0.31 nm perfectly matches the nearest neighbour distance of two carbon atoms in the 3C-SiC lattice. -As can be seen from the inset this peak is also observed for the $V_1$ simulation. -Investigating the corresponding coordinates of the atoms it turns out that concatenated and differently oriented C-Si \hkl<1 0 0> dumbbell interstitials constitute configurations yielding separations of C atoms by this distance. -In 3C-SiC the same distance is also expected for nearest neighbour silicon atoms. -The bottom of figure \ref{fig:md:pc_si-si_c-c} shows the radial distribution of Si-Si bonds together with a reference graph for pure c-Si. -Indeed non-zero $g(r)$ values around 0.31 nm are observed while the amount of Si pairs at regular c-Si distances of 0.24 nm and 0.38 nm decreases. +This is in accordance with the constant total energy observed in the continuation step of \unit[100]{ps} inbetween the end of C insertion and the cooling process. +Obviously, no energetically favorable relaxation is taking place at a system temperature of \unit[450]{$^{\circ}$C}. + +The C-C peak at about \unit[0.31]{nm} perfectly matches the nearest neighbor distance of two C atoms in the 3C-SiC lattice. +As can be seen from the inset, this peak is also observed for the $V_1$ simulation. +Investigating the corresponding coordinates of the atoms, it turns out that concatenated and differently oriented \ci{} \hkl<1 0 0> DB interstitials constitute configurations yielding separations of C atoms by this distance. +In 3C-SiC, the same distance is also expected for nearest neighbor Si atoms. +The bottom of Fig.~\ref{fig:md:pc_si-si_c-c} shows the radial distribution of Si-Si bonds together with a reference graph for pure c-Si. +Indeed, non-zero $g(r)$ values around \unit[0.31]{nm} are observed while the amount of Si pairs at regular c-Si distances of \unit[0.24]{nm} and \unit[0.38]{nm} decreases. However, no clear peak is observed but the interval of enhanced $g(r)$ values corresponds to the width of the C-C $g(r)$ peak. -In addition the abrupt increase of Si pairs at 0.29 nm can be attributed to the Si-Si cut-off radius of 0.296 nm as used in the present bond order potential. +In addition the abrupt increase of Si pairs at \unit[0.29]{nm} can be attributed to the Si-Si cut-off radius of \unit[0.296]{nm} as used in the present bond order potential. The cut-off function causes artificial forces pushing the Si atoms out of the cut-off region. -Without the abrubt increase a maximum around 0.31 nm gets even more conceivable. -For low concentrations of carbon, that is the $V_1$ simulation and early stages of the $V_2$ and $V_3$ simulations, analyses of configurations in which Si-Si distances around 0.3 nm appear and which are identifiable in regions of high disorder, which especially applies for the high concentration simulations, identify the \hkl<1 0 0> C-Si dumbbell to be responsible for stretching the Si-Si next neighbour distance. -This excellently agrees with the calculated value $r(13)$ in table \ref{tab:defects:100db_cmp} for a resulting Si-Si distance in the \hkl<1 0 0> C-Si dumbbell configuration. +Without the abrubt increase, a maximum around \unit[0.31]{nm} gets even more conceivable. +Analyses of randomly chosen configurations, in which distances around \unit[0.3]{nm} appear, identify \ci{} \hkl<1 0 0> DBs to be responsible for stretching the Si-Si next neighbour distance for low C concentrations, i.e. for the $V_1$ and early stages of $V_2$ and $V_3$ simulation runs. +This excellently agrees with the calculated value $r(13)$ in Table~\ref{tab:defects:100db_cmp} for a resulting Si-Si distance in the \ci \hkl<1 0 0> DB configuration. -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} - \includegraphics[width=12cm]{sic_prec_450_si-c.ps} + \includegraphics[width=0.7\textwidth]{sic_prec_450_si-c.ps} \end{center} -\caption{Radial distribution function of the Si-C distances for 6000 carbon atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of $450\,^{\circ}\mathrm{C}$ and cooled down to room temperature together with Si-C bonds resulting in a C-Si \hkl<1 0 0> dumbbell configuration.} +\caption{Radial distribution function of the Si-C distances for 6000 C atoms inserted into the three different volumes $V_1$, $V_2$ and $V_3$ at a temperature of \unit[450]{$^{\circ}$C} and cooled down to room temperature. Additionally the resulting Si-C distances of a \ci{} \hkl<1 0 0> DB configuration are given.} \label{fig:md:pc_si-c} \end{figure} -Figure \ref{fig:md:pc_si-c} displays the Si-C radial distribution function for all three insertion volumes together with the Si-C bonds as observed in a C-Si \hkl<1 0 0> dumbbell configuration. -The first peak observed for all insertion volumes is at approximately 0.186 nm. -This corresponds quite well to the expected next neighbour distance of 0.189 nm for Si and C atoms in 3C-SiC. -By comparing the resulting Si-C bonds of a C-Si \hkl<1 0 0> dumbbell with the C-Si distances of the low concentration simulation it is evident that the resulting structure of the $V_1$ simulation is dominated by this type of defects. -This is not surpsising, since the \hkl<1 0 0> dumbbell is found to be the ground state defect of a C interstitial in c-Si and for the low concentration simulations a carbon interstitial is expected in every fifth silicon unit cell only, thus, excluding defect superposition phenomena. -The peak distance at 0.186 nm and the bump at 0.175 nm corresponds to the distance $r(3C)$ and $r(1C)$ as listed in table \ref{tab:defects:100db_cmp} and visualized in figure \ref{fig:defects:100db_cmp}. -In addition it can be easily identified that the \hkl<1 0 0> dumbbell configuration contributes to the peaks at about 0.335 nm, 0.386 nm, 0.434 nm, 0.469 nm and 0.546 nm observed in the $V_1$ simulation. +Fig.~\ref{fig:md:pc_si-c} displays the Si-C radial distribution function for all three insertion volumes together with the Si-C bonds as observed in a \ci{} \hkl<1 0 0> DB configuration. +The first peak observed for all insertion volumes is at approximately \unit[0.186]{nm}. +This corresponds quite well to the expected next neighbor distance of \unit[0.189]{nm} for Si and C atoms in 3C-SiC. +By comparing the resulting Si-C bonds of a \ci{} \hkl<1 0 0> DB with the C-Si distances of the low concentration simulation, it is evident that the resulting structure of the $V_1$ simulation is clearly dominated by this type of defect. +This is not surpsising, since the \ci{} \hkl<1 0 0> DB is found to be the ground-state defect of a C interstitial in c-Si and, for the low concentration simulations, a C interstitial is expected in every fifth Si unit cell only, thus, excluding defect superposition phenomena. +The peak distance at \unit[0.186]{nm} and the bump at \unit[0.175]{nm} corresponds to the distance $r(3C)$ and $r(1C)$ as listed in Table~\ref{tab:defects:100db_cmp} and visualized in Fig.~\ref{fig:defects:100db_cmp}. +In addition, it can be easily identified that the \ci{} \hkl<1 0 0> DB configuration contributes to the peaks at about \unit[0.335]{nm}, \unit[0.386]{nm}, \unit[0.434]{nm}, \unit[0.469]{nm} and \unit[0.546]{nm} observed in the $V_1$ simulation. Not only the peak locations but also the peak widths and heights become comprehensible. -The distinct peak at 0.26 nm, which exactly matches the cut-off radius of the Si-C interaction, is again a potential artifact. +The distinct peak at \unit[0.26]{nm}, which exactly matches the cut-off radius of the Si-C interaction, is again a potential artifact. -For high carbon concentrations, that is the $V_2$ and $V_3$ simulation, the defect concentration is likewiese increased and a considerable amount of damage is introduced in the insertion volume. +For high C concentrations, i.e. the $V_2$ and $V_3$ simulation corresponding to a C density of about 8 atoms per c-Si unit cell, the defect concentration is likewiese increased and a considerable amount of damage is introduced in the insertion volume. The consequential superposition of these defects and the high amounts of damage generate new displacement arrangements for the C-C as well as for the Si-C pair distances, which become hard to categorize and trace and obviously lead to a broader distribution. -Short range order indeed is observed but only hardly visible is the long range order. +Short range order indeed is observed, i.e. the large amount of strong neighbored C-C bonds at \unit[0.15]{nm} as expected in graphite or diamond and Si-C bonds at \unit[0.19]{nm} as expected in SiC, but only hardly visible is the long range order. This indicates the formation of an amorphous SiC-like phase. In fact the resulting Si-C and C-C radial distribution functions compare quite well with these obtained by cascade amorphized and melt-quenched amorphous SiC using a modified Tersoff potential \cite{gao02}. -\subsection{Limitations of conventional MD and short range potentials} -\label{subsection:md:limit} +In both cases, i.e. low and high C concentrations, the formation of 3C-SiC fails to appear. +With respect to the precipitation model, the formation of C$_{\text{i}}$ \hkl<1 0 0> DBs indeed occurs for low C concentrations. +However, sufficient defect agglomeration is not observed. +For high C concentrations, a rearrangement of the amorphous SiC structure, which is not expected at prevailing temperatures, and a transition into 3C-SiC is not observed either. + +\section{Limitations of conventional MD and short range potentials} +\label{section:md:limit} -At first the formation of an amorphous SiC-like phase is unexpected since IBS experiments show crystalline 3C-SiC precipitates at prevailing temperatures. -On closer inspection, however, reasons become clear, which are discussed in the following. +Results of the last section indicate possible limitations of the MD method regarding the task addressed in this study. +Low C concentration simulations do not reproduce the agglomeration of C$_{\text{i}}$ \hkl<1 0 0> DBs. +High concentration simulations result in the formation of an amorphous SiC-like phase, which is unexpected since IBS experiments show crystalline 3C-SiC precipitates at prevailing temperatures. +Keeping in mind the results +On closer inspection, however, two reasons for describing this obstacle become evident, which are discussed in the following. The first reason is a general problem of MD simulations in conjunction with limitations in computer power, which results in a slow and restricted propagation in phase space. In molecular systems, characteristic motions take place over a wide range of time scales. -Vibrations of the covalent bond take place on the order of $10^{-14}\,\text{s}$ of which the thermodynamic and kinetic properties are well described by MD simulations. -To avoid dicretization errors the integration timestep needs to be chosen smaller than the fastest vibrational frequency in the system. +Vibrations of the covalent bond take place on the order of \unit[10$^{-14}$]{s}, of which the thermodynamic and kinetic properties are well described by MD simulations. +To avoid dicretization errors, the integration timestep needs to be chosen smaller than the fastest vibrational frequency in the system. On the other hand, infrequent processes, such as conformational changes, reorganization processes during film growth, defect diffusion and phase transitions are processes undergoing long-term evolution in the range of microseconds. -This is due to the existence of several local minima in the free energy surface separated by large energy barriers compared to the kinetic energy of the particles, that is the system temperature. +This is due to the existence of several local minima in the free energy surface separated by large energy barriers compared to the kinetic energy of the particles, i.e. the system temperature. Thus, the average time of a transition from one potential basin to another corresponds to a great deal of vibrational periods, which in turn determine the integration timestep. -Hence, time scales covering the neccessary amount of infrequent events to observe long-term evolution are not accessible by traditional MD simulations, which are limited to the order of nanoseconds. -New methods have been developed to bypass the time scale problem like hyperdnyamics (HMD) \cite{voter97,voter97_2}, parallel replica dynamics \cite{voter98}, temperature acclerated dynamics (TAD) \cite{sorensen2000} and self-guided dynamics (SGMD) \cite{wu99} retaining proper thermodynmic sampling. +Hence, time scales covering the necessary amount of infrequent events to observe long-term evolution are not accessible by traditional MD simulations, which are limited to the order of nanoseconds. +New methods have been developed to bypass the time scale problem. +The most famous appraiches are hyperdnyamics (HMD) \cite{voter97,voter97_2}, parallel replica dynamics \cite{voter98}, temperature acclerated dynamics (TAD) \cite{sorensen2000} and self-guided dynamics (SGMD) \cite{wu99}, which accelerate phase space propagation while retaining proper thermodynmic sampling. In addition to the time scale limitation, problems attributed to the short range potential exist. -The sharp cut-off funtion, which limits the interacting ions to the next neighboured atoms by gradually pushing the interaction force and energy to zero between the first and second next neighbour distance, is responsible for overestimated and unphysical high forces of next neighboured atoms \cite{tang95,mattoni2007}. +The sharp cut-off funtion, which limits the interacting ions to the next neighbored atoms by gradually pushing the interaction force and energy to zero between the first and second next neighbor distance, is responsible for overestimated and unphysical high forces of next neighbored atoms \cite{tang95,mattoni2007}. This is supported by the overestimated activation energies necessary for C diffusion as investigated in section \ref{subsection:defects:mig_classical}. -Indeed it is not only the strong C-C bond which is hard to break inhibiting carbon diffusion and further rearrengements. -This is also true for the low concentration simulations dominated by the occurrence of C-Si dumbbells spread over the whole simulation volume. -The bonds of these C-Si pairs are also affected by the cut-off artifact preventing carbon diffusion and agglomeration of the dumbbells. -This can be seen from the almost horizontal progress of the total energy graph in the continuation step, even for the low concentration simulation. -The unphysical effects inherent to this type of model potentials are solely attributed to their short range character. +Indeed, it is not only the strong C-C bond, which is hard to break, inhibiting C diffusion and further rearrengements. +This is also true for the low concentration simulations dominated by the occurrence of C-Si DBs spread over the whole simulation volume. +The bonds of these C-Si pairs are also affected by the cut-off artifact preventing C diffusion and agglomeration of the DBs. +This can be seen from the almost horizontal progress of the total energy graph in the continuation step of Fig.~\ref{fig:md:energy_450}, even for the low concentration simulation. +These unphysical effects inherent to this type of model potentials are solely attributed to their short range character. +While cohesive and formational energies are often well described, these effects increase for non-equilibrium structures and dynamics. However, since valueable insights into various physical properties can be gained using this potentials, modifications mainly affecting the cut-off were designed. One possibility is to simply skip the force contributions containing the derivatives of the cut-off function, which was successfully applied to reproduce the brittle propagation of fracture in SiC at zero temperature \cite{mattoni2007}. Another one is to use variable cut-off values scaled by the system volume, which properly describes thermomechanical properties of 3C-SiC \cite{tang95} but might be rather ineffective for the challange inherent to this study. To conclude the obstacle needed to get passed is twofold. -The sharp cut-off of the used bond order model potential introduces overestimated high forces between next neighboured atoms enhancing the problem of slow phase space propagation immanent to MD simulations. +The sharp cut-off of the employed bond order model potential introduces overestimated high forces between next neighbored atoms enhancing the problem of slow phase space propagation immanent to MD simulations, termed {\em potential enhanced slow phase space propagation} in the following. Thus, pushing the time scale to the limits of computational ressources or applying one of the above mentioned accelerated dynamics methods exclusively will not be sufficient enough. -Instead the first approach followed in this study, is the use of higher temperatures as exploited in TAD to find transition pathways of one local energy minimum to another one more quickly. +Instead, the approach followed in this study, is the use of higher temperatures as exploited in TAD to find transition pathways of one local energy minimum to another one more quickly. Since merely increasing the temperature leads to different equilibrium kinetics than valid at low temperatures, TAD introduces basin-constrained MD allowing only those transitions that should occur at the original temperature and a properly advancing system clock \cite{sorensen2000}. The TAD corrections are not applied in coming up simulations. This is justified by two reasons. -First of all a compensation of the overestimated bond strengthes due to the short range potential is expected. -Secondly there is no conflict applying higher temperatures without the TAD corrections, since crystalline 3C-SiC is also observed for higher temperatures than $450\,^{\circ}\mathrm{C}$ in IBS \cite{lindner01}. +First of all, a compensation of the overestimated bond strengths due to the short range potential is expected. +Secondly, there is no conflict applying higher temperatures without the TAD corrections, since crystalline 3C-SiC is also observed for higher temperatures than \unit[450]{$^{\circ}$C} in IBS \cite{nejim95,lindner01}. It is therefore expected that the kinetics affecting the 3C-SiC precipitation are not much different at higher temperatures aside from the fact that it is occuring much more faster. Moreover, the interest of this study is focused on structural evolution of a system far from equilibrium instead of equilibrium properties which rely upon proper phase space sampling. On the other hand, during implantation, the actual temperature inside the implantation volume is definetly higher than the experimentally determined temperature tapped from the surface of the sample. -\subsection{Increased temperature simulations} -\label{subsection:md:inct} +\section{Increased temperature simulations} +\label{section:md:inct} -Due to the limitations of short range potentials and conventional MD as discussed above elevated temperatures are used in the following. -The simulation sequence and other parameters aside system temperature remain unchanged as in section \ref{subsection:initial_sims}. -Since there is no significant difference among the $V_2$ and $V_3$ simulations only the $V_1$ and $V_2$ simulations are carried on and refered to as low carbon and high carbon concentration simulations. -Temperatures ranging from $450\,^{\circ}\mathrm{C}$ up to $2050\,^{\circ}\mathrm{C}$ are used. +Due to the limitations of short range potentials and conventional MD as discussed above, elevated temperatures are used in the following. +The simulation sequence and other parameters except for the system temperature remain unchanged as in section \ref{section:initial_sims}. +Since there is no significant difference among the $V_2$ and $V_3$ simulations only the $V_1$ and $V_2$ simulations are carried on and referred to as low C and high C concentration simulations. +Temperatures ranging from \degc{450} up to \unit[2050]{$^{\circ}$C} are used. A simple quality value $Q$ is introduced, which helps to estimate the progress of structural evolution. -In bulk 3C-SiC every C atom has four next neighboured Si atoms and every Si atom four next neighboured C atoms. -The quality could be determined by counting the amount of atoms which form bonds to four atoms of the other species. -However, the aim of the simulation on hand is to reproduce the formation of a 3C-SiC precipitate embedded in c-Si. -The amount of Si atoms and, thus, the amount of Si atoms remaining in the silicon diamond lattice is much higher than the amount of inserted C atoms. -Thus, counting the atoms, which exhibit proper coordination is limited to the C atoms. +In bulk 3C-SiC every C atom has four next neighbored Si atoms and every Si atom four next neighbored C atoms. +The quality could be determined by counting the amount of atoms, which form bonds to four atoms of the other species. +However, the aim of the simulation is to reproduce the formation of a 3C-SiC precipitate embedded in c-Si. +The amount of Si atoms and, thus, the amount of Si atoms remaining in the c-Si diamond lattice is much higher than the amount of inserted C atoms. +Thus, counting the atoms, which exhibit proper coordination, is limited to the C atoms. The quality value is defined to be \begin{equation} -Q = \frac{\text{Amount of C atoms with 4 next neighboured Si atoms}} +Q = \frac{\text{Amount of C atoms with 4 next neighbored Si atoms}} {\text{Total amount of C atoms}} \text{ .} \label{eq:md:qdef} \end{equation} By this, bulk 3C-SiC will still result in $Q=1$ and precipitates will also reach values close to one. -However, since the quality value does not account for bond lengthes, bond angles, crystallinity or the stacking sequence high values of $Q$ not necessarily correspond to structures close to 3C-SiC. +However, since the quality value does not account for bond lengthes, bond angles, crystallinity or the stacking sequence, high values of $Q$ not necessarily correspond to structures close to 3C-SiC. Structures that look promising due to high quality values need to be further investigated by other means. -\subsubsection{Low carbon concetration simulations} +\subsubsection{Low C concetration simulations} -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{tot_pc_thesis.ps}\\ -\includegraphics[width=12cm]{tot_ba.ps} +\includegraphics[width=0.7\textwidth]{tot_pc_thesis.ps}\\ +\includegraphics[width=0.7\textwidth]{tot_ba.ps} \end{center} -\caption[Si-C radial distribution and quality evolution for the low concentration simulations at different elevated temperatures.]{Si-C radial distribution and quality evolution for the low concentration simulations at different elevated temperatures. All structures are cooled down to $20\,^{\circ}\mathrm{C}$. The grey line shows resulting Si-C bonds in a configuration of substitutional C in c-Si (C$_\text{sub}$) at zero temperature. Arrows in the quality plot mark the end of carbon insertion and the start of the cooling down step. A fit function according to equation \eqref{eq:md:fit} shows the estimated evolution of quality in the absence of the cooling down sequence.} +\caption[Si-C radial distribution and evolution of quality $Q$ for the low concentration simulations at different elevated temperatures.]{Si-C radial distribution and evolution of quality $Q$ according to equation \ref{eq:md:qdef} for the low concentration simulations at different elevated temperatures. All structures are cooled down to \degc{20}. The grey line shows resulting Si-C bonds in a configuration of \cs{} in c-Si (C$_\text{sub}$) at zero temperature. Arrows in the quality plot mark the end of C insertion and the start of the cooling down step. A fit function according to equation \eqref{eq:md:fit} shows the estimated evolution of quality in the absence of the cooling down sequence.} \label{fig:md:tot_si-c_q} \end{figure} -Figure \ref{fig:md:tot_si-c_q} shows the radial distribution of Si-C bonds for different temperatures and the corresponding quality evolution as defined earlier for the low concentration simulaton, that is the $V_1$ simulation. -The first noticeable and promising change in the Si-C radial distribution is the successive decline of the artificial peak at the Si-C cut-off distance with increasing temperature up to the point of disappearance at temperatures above $1650\,^{\circ}\mathrm{C}$. -The system provides enough kinetic energy to affected atoms, which are able to escape the cut-off region. -Another important observation in structural change is exemplified in the two shaded areas. -In the grey shaded region a decrease of the peak at 0.186 nm and the bump at 0.175 nm and a concurrent increase of the peak at 0.197 nm with increasing temperature is visible. -Similarly the peaks at 0.335 nm and 0.386 nm shrink in contrast to a new peak forming at 0.372 nm as can be seen in the yellow shaded region. -Obviously the structure obtained from the $450\,^{\circ}\mathrm{C}$ simulations, which is dominated by the existence of \hkl<1 0 0> C-Si dumbbells transforms into a different structure with increasing simulation temperature. -Investigations of the atomic data reveal substitutional carbon to be responsible for the new Si-C bonds. -The peak at 0.197 nm corresponds to the distance of a substitutional carbon atom to the next neighboured silicon atoms. -The one at 0.372 nm is the distance of a substitutional carbon atom to the second next silicon neighbour along a \hkl<1 1 0> direction. -Comparing the radial distribution for the Si-C bonds at $2050\,^{\circ}\mathrm{C}$ to the resulting Si-C bonds in a configuration of a substitutional carbon atom in crystalline silicon excludes all possibility of doubt. +Fig.~\ref{fig:md:tot_si-c_q} shows the radial distribution of Si-C bonds for different temperatures and the corresponding evolution of quality $Q$ as defined above for the low concentration simulaton. +The first noticeable and promising change in the Si-C radial distribution is the successive decline of the artificial peak at the Si-C cut-off distance with increasing temperature up to the point of disappearance at temperatures above \degc{1650}. +Obviously, sufficient kinetic energy is provided to affected atoms that are enabled to escape the cut-off region. +Additionally, a more important structural change is observed, which is illustrated in the two shaded areas in Fig.~\ref{fig:md:tot_si-c_q}. +% +In the grey shaded region a decrease of the peak at \unit[0.186]{nm} and of the bump at \distn{0.175} accompanied by an increase of the peak at \distn{0.197} with increasing temperature is visible. +Similarly, the peaks at \distn{0.335} and \distn{0.386} shrink in contrast to a new peak forming at \distn{0.372} as can be seen in the yellow shaded region. +Obviously, the structure obtained from the \degc{450} simulations, which is dominated by the existence of \ci{} \hkl<1 0 0> DBs, transforms into a different structure with increasing simulation temperature. +Investigations of the atomic data reveal \cs{} to be responsible for the new Si-C bonds. +The peak at \distn{0.197} corresponds to the distance of a \cs{} atom and its next neighbored Si atoms. +The one at \distn{0.372} constitutes the distance of a \cs{} atom to the second next Si neighbor along a \hkl<1 1 0> direction. +Comparing the radial distribution for the Si-C bonds at \degc{2050} to the resulting Si-C bonds in a configuration of a \cs{} atom in c-Si excludes all possibility of doubt. The resulting bonds perfectly match and, thus, explain the peaks observed for the increased temperature simulations. -To conclude, by increasing the simulation temperature, the \hkl<1 0 0> C-Si dumbbell characterized structure transforms into a structure dominated by substitutional C. +To conclude, by increasing the simulation temperature, the structure characterized by the \ci{} \hkl<1 0 0> DB structure transforms into a structure dominated by \cs{}. -This is also reflected in the quality values obtained for different temperatures. -While simulations at $450\,^{\circ}\mathrm{C}$ exhibit 10 \% of fourfold coordinated carbon simulations at $2050\,^{\circ}\mathrm{C}$ exceed the 80 \% range. -Since substitutional carbon has four next neighboured silicon atoms and is the preferential type of defect in elevated temperature simulations the increase of the quality values become evident. +This is likewise reflected in the quality values obtained for different temperatures. +While simulations at \degc{450} exhibit \perc{10} of fourfold coordinated C, simulations at \degc{2050} exceed the \perc{80} range. +Since \cs{} has four nearest neighbored Si atoms and is the preferential type of defect in elevated temperature simulations, the increase of the quality values become evident. The quality values at a fixed temperature increase with simulation time. -After the end of the insertion sequence marked by the first arrow the quality is increasing and a saturation behaviour, yet before the cooling process starts, can be expected. -The evolution of the quality value of the simulation at $2050\,^{\circ}\mathrm{C}$ inside the range in which the simulation is continued at constant temperature for 100 fs is well approximated by the simple fit function +After the end of the insertion sequence marked by the first arrow, the quality is increasing and a saturation behaviour, yet before the cooling process starts, can be expected. +The evolution of the quality value of the simulation at \degc{2050} inside the range, in which the simulation is continued at constant temperature for \unit[100]{fs}, is well approximated by the simple fit function \begin{equation} f(t)=a-\frac{b}{t} \text{ ,} \label{eq:md:fit} \end{equation} -which results in a saturation value of 93 \%. -Obviously the decrease in temperature accelerates the saturation and inhibits further formation of substitutional carbon. +which results in a saturation value of \perc{93}. +Obviously, the decrease in temperature accelerates the saturation and inhibits further formation of \cs{}. \label{subsubsection:md:ep} -Conclusions drawn from investigations of the quality evolution correlate well with the findings of the radial distribution results. +Conclusions drawn from investigations of the quality evolution correlate well with findings of the radial distribution results. -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{tot_pc2_thesis.ps}\\ -\includegraphics[width=12cm]{tot_pc3_thesis.ps} +\includegraphics[width=0.7\textwidth]{tot_pc2_thesis.ps}\\ +\includegraphics[width=0.7\textwidth]{tot_pc3_thesis.ps} \end{center} -\caption[C-C and Si-Si radial distribution for the low concentration simulations at different elevated temperatures.]{C-C and Si-Si radial distribution for the low concentration simulations at different elevated temperatures. All structures are cooled down to $20\,^{\circ}\mathrm{C}$. Arrows with dashed lines mark C-C distances of \hkl<1 0 0> dumbbell combinations and those with solid lines mark C-C distances of combinations of substitutional C. The dashed line corresponds to the distance of a substitutional C with a next neighboured \hkl<1 0 0> dumbbell.} +\caption[C-C and Si-Si radial distribution for the low concentration simulations at different elevated temperatures.]{C-C and Si-Si radial distribution for the low concentration simulations at different elevated temperatures. All structures are cooled down to $20\,^{\circ}\mathrm{C}$. Arrows with dashed lines mark C-C distances of \hkl<1 0 0> DB combinations and those with solid lines mark C-C distances of combinations of substitutional C. The dashed line corresponds to the distance of a substitutional C with a next neighbored \hkl<1 0 0> DB.} \label{fig:md:tot_c-c_si-si} \end{figure} -The formation of substitutional carbon also affects the Si-Si radial distribution displayed in the lower part of figure \ref{fig:md:tot_c-c_si-si}. -Investigating the atomic strcuture indeed shows that the peak arising at 0.325 nm with increasing temperature is due to two Si atoms directly bound to a C substitutional. -It corresponds to the distance of second next neighboured Si atoms along a \hkl<1 1 0>-equivalent direction with substitutional C inbetween. -Since the expected distance of these Si pairs in 3C-SiC is 0.308 nm the existing SiC structures embedded in the c-Si host are stretched. +The formation of \cs{} also affects the Si-Si radial distribution displayed in the lower part of Fig.~\ref{fig:md:tot_c-c_si-si}. +Investigating the atomic strcuture indeed shows that the peak arising at \distn{0.325} with increasing temperature is due to two Si atoms that form direct bonds to the \cs{} atom. +The peak corresponds to the distance of next neighbored Si atoms along the \hkl<1 1 0> bond chain with C$_{\text{s}}$ in between. +Since the expected distance of these Si pairs in 3C-SiC is \distn{0.308}, the existing SiC structures embedded in the c-Si host are stretched. -In the upper part of figure \ref{fig:md:tot_c-c_si-si} the C-C radial distribution is shown. +In the upper part of Fig.~\ref{fig:md:tot_c-c_si-si} the C-C radial distribution is shown. The total amount of C-C bonds with a distance inside the displayed separation range does not change significantly. -Thus, even for elevated temperatures agglomeration of C atoms neccessary to form a SiC precipitate does not take place within the simulated time scale. -However, with increasing temperature a decrease of the amount of next neighboured C pairs can be observed. -This is a promising result gained by the high temperature simulations since the breaking of these diomand and graphite like bonds is mandatory for the formation of 3C-SiC. -A slight shift towards higher distances can be observed for the maximum above 0.3 nm. -Arrows with dashed lines mark C-C distances resulting from \hkl<1 0 0> dumbbell combinations while the arrows with solid lines mark distances arising from combinations of substitutional C. -The continuous dashed line corresponds to the distance of a substitutional C with a next neighboured \hkl<1 0 0> dumbbell. -By comparison with the radial distribution it becomes evident that the shift accompanies the advancing transformation of \hkl<1 0 0> dumbbells into substitutional C. -Next to combinations of two substitutional C atoms and two \hkl<1 0 0> dumbbells respectively also combinations of \hkl<1 0 0> dumbbells with a substitutional C atom arise. -In addition, structures form that result in distances residing inbetween the ones obtained from combinations of mixed defect types and the ones obtained by substitutional C configurations, as can be seen by quite high g(r) values to the right of the continuous dashed line and to the left of the first arrow with a solid line. -For the most part these structures can be identified as configurations of one substitutional C atom with either another C atom that practically occupies a Si lattice site but with a Si interstitial residing in the very next surrounding or a C atom that nearly occupies a Si lattice site forming a defect other than the \hkl<1 0 0>-type with the Si atom. -Again, this is a quite promising result, since the C atoms are taking the appropriate coordination as expected in 3C-SiC. -However, this is contrary to the initial precipitation model proposed in section \ref{section:assumed_prec}, which assumes that the transformation into 3C-SiC takes place in a very last step once enough C-Si dumbbells agglomerated. +Thus, even for elevated temperatures, agglomeration of C atoms neccessary to form a SiC precipitate does not take place within the simulated time scale. +However, with increasing temperature, a decrease of the amount of next neighbored C pairs can be observed. +This is a promising result gained by the high-temperature simulations since the breaking of these diomand and graphite like bonds is mandatory for the formation of 3C-SiC. +Obviously, acceleration of the dynamics occurred by supplying additional kinetic energy. +A slight shift towards higher distances can be observed for the maximum located shortly above \distn{0.3}. +Arrows with dashed lines mark C-C distances resulting from \ci{} \hkl<1 0 0> DB combinations while arrows with solid lines mark distances arising from combinations of \cs. +The continuous dashed line corresponds to the distance of \cs{} and a next neighbored \ci{} \hkl<1 0 0> DB. +% +Obviously, the shift of the peak is caused by the advancing transformation of the C$_{\text{i}}$ DB into the C$_{\text{s}}$ defect. +Next to combinations of two \cs{} atoms or \ci{} \hkl<1 0 0> DBs, combinations of \ci{} \hkl<1 0 0> DBs with a \cs{} atom arise. +In addition, structures form that result in distances residing inbetween the ones obtained from combinations of mixed defect types and the ones obtained by \cs{} configurations, as can be seen by quite high $g(r)$ values in between the continuous dashed line and the first arrow with a solid line. +For the most part, these structures can be identified as configurations of \cs{} with either another C atom that basically occupies a Si lattice site but is displaced by a \si{} atom residing in the very next surrounding or a C atom that nearly occupies a Si lattice site forming a defect other than the \hkl<1 0 0>-type with the Si atom. +Again, this is a quite promising result since the C atoms are taking the appropriate coordination as expected in 3C-SiC. +%However, this is contrary to the initial precipitation model proposed in section \ref{section:assumed_prec}, which assumes that the transformation into 3C-SiC takes place in a very last step once enough C-Si DBs agglomerated. To summarize, results of low concentration simulations show a phase transition in conjunction with an increase in temperature. -The C-Si \hkl<1 0 0> dumbbell dominated struture turns into a structure characterized by the occurence of an increasing amount of substitutional C with respect to temperature. -Although diamond and graphite like bonds are reduced no agglomeration of C is observed within the simulated time resulting in the formation of isolated structures of stretched SiC, which are adjusted to the c-Si host with respect to the lattice constant and alignement. -It would be conceivable that by agglomeration of further substitutional C atoms the interfacial energy could be overcome and a transition into an incoherent SiC precipitate could occur. +The \ci{} \hkl<1 0 0> DB dominated struture turns into a structure characterized by the occurence of an increasing amount of \cs{} with respect to temperature. +Clearly, the high-temperature results indicate the precipitation mechanism involving an increased participation of \cs. +Although diamond and graphite like bonds are reduced, no agglomeration of C is observed within the simulated time. +Isolated structures of stretched SiC, which are adjusted to the c-Si host with respect to the lattice constant and alignement, are formed. +It would be conceivable that by agglomeration of further \cs{} atoms the interfacial energy could be overcome and a transition from a coherent and stretched SiC structure into an incoherent and partially strain-compensated SiC precipitate could occur. -{\color{red}Todo: Results reinforce the assumption of an alternative precipitation model as already pointed out in the defects chapter.} +\subsubsection{High C concetration simulations} -\subsubsection{High carbon concetration simulations} - -\begin{figure}[!ht] +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{12_pc_thesis.ps}\\ -\includegraphics[width=12cm]{12_pc_c_thesis.ps} +\includegraphics[width=0.7\textwidth]{12_pc_thesis.ps}\\ +\includegraphics[width=0.7\textwidth]{12_pc_c_thesis.ps} \end{center} -\caption[Si-C and C-C radial distribution for the high concentration simulations at different elevated temperatures.]{Si-C (top) and C-C (bottom) radial distribution for the high concentration simulations at different elevated temperatures. All structures are cooled down to $20\,^{\circ}\mathrm{C}$.} +\caption[Si-C and C-C radial distribution for the high concentration simulations at different elevated temperatures.]{Si-C (top) and C-C (bottom) radial distribution for the high concentration simulations at different elevated temperatures. All structures are cooled down to \degc{20}.} \label{fig:md:12_pc} \end{figure} -Figure \ref{fig:md:12_pc} displays the radial distribution for Si-C and C-C pairs obtained from high C concentration simulations at different elevated temperatures. +Fig.~\ref{fig:md:12_pc} displays the radial distribution for Si-C and C-C pairs obtained from high C concentration simulations at different elevated temperatures. Again, in both cases, the cut-off artifact decreases with increasing temperature. Peaks that already exist for the low temperature simulations get slightly more distinct for elevated temperatures. -This is also true for peaks located past distances of next neighbours indicating an increase in the long range order. -However this change is rather small and no significant structural change is observeable. -Due to the continuity of high amounts of damage atomic configurations remain hard to identify even for the highest temperature. -Other than in the low concentration simulation analyzed defect structures are no longer necessarily aligned to the primarily existing but succesively disappearing c-Si host matrix inhibiting or at least hampering their identification and classification. -As for low temperatures order in the short range exists decreasing with increasing distance. -The increase of the amount of Si-C pairs at 0.186 nm could be positively interpreted since this type of bond also exists in 3C-SiC. -On the other hand the amount of next neighboured C atoms with a distance of approximately 0.15 nm, which is the distance of C in graphite or diamond, is likewise increased. -Thus, higher temperatures seem to additionally enhance a conflictive process, that is the formation of C agglomerates, instead of the desired process of 3C-SiC formation. -This is supported by the C-C peak at 0.252 nm, which corresponds to the second next neighbour distance in the diamond structure of elemental C. -Investigating the atomic data indeed reveals two C atoms which are bound to and interconnected by a third C atom to be responsible for this distance. -The C-C peak at about 0.31 nm, wich is slightly shifted to higher distances (0.317 nm) with increasing temperature still corresponds quite well to the next neighbour distance of C in 3C-SiC as well as a-SiC and indeed results from C-Si-C bonds. -The Si-C peak at 0.282 nm, which is pronounced with increasing temperature is constructed out of a Si atom and a C atom, which are both bound to another central C atom. -This is similar for the Si-C peak at approximately 0.35 nm. +This is also true for peaks located past distances of next neighbors indicating an increase in the long range order. +However, this change is rather small and no significant structural change is observeable. +Due to the continuity of high amounts of damage, atomic configurations remain hard to identify even for the highest temperature. +Other than in the low concentration simulation, analyzed defect structures are no longer necessarily aligned to the primarily existing but succesively disappearing c-Si host matrix inhibiting or at least hampering their identification and classification. +As for low temperatures, order in the short range exists decreasing with increasing separation. +The increase of the amount of Si-C pairs at \distn{0.186} could be positively interpreted since this type of bond also exists in 3C-SiC. +On the other hand, the amount of next neighbored C atoms with a distance of approximately \distn{0.15}, which is the distance of C in graphite or diamond, is likewise increased. +Thus, higher temperatures seem to additionally enhance a conflictive process, i.e. the formation of C agglomerates, obviously inconsistent with the desired process of 3C-SiC formation. +This is supported by the C-C peak at \distn{0.252}, which corresponds to the second next neighbor distance in the diamond structure of elemental C. +Investigating the atomic data indeed reveals two C atoms, which are bound to and interconnected by a third C atom, to be responsible for this distance. +The C-C peak at about \distn{0.31}, wich is slightly shifted to higher distances (\distn{0.317}) with increasing temperature still corresponds quite well to the next neighbor distance of C in 3C-SiC as well as a-SiC and indeed results from C-Si-C bonds. +The Si-C peak at \distn{0.282}, which is pronounced with increasing temperature, is constructed out of a Si atom and a C atom, which are both bound to another central C atom. +This is similar for the Si-C peak at approximately \distn{0.35}. In this case, the Si and the C atom are bound to a central Si atom. -To summarize, the amorphous phase remains though sharper peaks in the radial distributions at distances expected for a-SiC are observed indicating a slight acceleration of the dynamics due to elevated temperatures. +To summarize, the amorphous phase remains. +Though, sharper peaks in the radial distributions at distances expected for a-SiC are observed indicating a slight acceleration of the dynamics due to elevated temperatures. \subsubsection{Conclusions concerning the usage of increased temperatures} -Regarding the outcome of both, high and low concentration simulations at increased temperatures, encouraging conclusions can be drawn. -With the disappearance of the peaks at the respective cut-off radii one limitation of the short range potential seems to be accomplished. -In addition, sharper peaks in the radial distributions lead to the assumption of expeditious structural formation. -The increase in temperature leads to the occupation of new defect states, which is particularly evident but not limited to the low carbon concentration simulations. +Regarding the outcome of both, high and low C concentration simulations at increased temperatures, encouraging conclusions can be drawn. +With the disappearance of the peaks at the respective cut-off radii, one limitation of the short range potential seems to be accomplished. +In addition, sharper peaks in the radial distribution functions lead to the assumption of expeditious structural formation. +The increase in temperature leads to the occupation of new defect states, which is particularly evident but not limited to the low C concentration simulations. + +The question remains, whether these states are only occupied due to the additional supply of kinetic energy and, thus, have to be considered unnatural for temperatures applied in IBS or whether the increase in temperature indeed enables infrequent transitions to occur faster, thus, leading to the intended acceleration of the dynamics and weakening of the unphysical quirks inherent to the potential. +As already pointed out in section~\ref{section:defects:noneq_process_01} and section~\ref{section:defects:noneq_process_02}, IBS is a non-equilibrium process, which might result in the formation of the thermodynamically less stable \cs{} and \si{} configuration. +Indeed, 3C-SiC is metastable and if not fabricated by IBS requires strong deviation from equilibrium and low temperatures to stabilize the 3C polytype. +In IBS, highly energetic C atoms are able to generate vacant sites, which in turn can be occupied by highly mobile \ci{} atoms. +This is in fact found to be favorable in the absence of the \si{}, which turned out to have a low interaction capture radius with the \cs{} atom and very likely prevents the recombination into a thermodynamically stable \ci{} DB for appropriate separations of the defect pair. +Results gained in this chapter show preferential occupation of Si lattice sites by \cs{} enabled by increased temperatures supporting the assumptions drawn from the defect studies of the last chapter. -{\color{blue} -The question remains whether these states are only occupied due to the additional supply of kinetic energy and, thus, have to be considered unnatural for temperatures applied in IBS or whether the increase in temperature indeed enables infrequent transitions to occur faster, thus, leading to the intended acceleration of the dynamics and weakening of the unphysical quirks inherent to the potential. -As already pointed out in section~\ref{section:defects:noneq_process_01} on page~\pageref{section:defects:noneq_process_01} and section~\ref{section:defects:noneq_process_02} on page~\pageref{section:defects:noneq_process_02} IBS is a nonequilibrium process, which might result in the formation of the thermodynamically less stable substitutional carbon and Si self-interstitital configuration. -Indeed 3C-SiC is metastable and if not fabricated by IBS requires strong deviation from equilibrium and/or low temperatures to stabilize the 3C polytype \cite{}. -In IBS highly energetic C atoms are able to generate vacant sites, which in turn can be occupied by highly mobile C atoms. -This is found to be favorable in the absence of the Si self-interstitial, which turned out to have a low interaction capture radius with a substitutional C atom very likely preventing the recombination into thermodynamically stable C-Si dumbbell interstitials for appropriate separations of the defect pair. -Results gained in this chapter show preferential occupation of Si lattice sites by substitutional C enabled by increased temperatures supporting the assumptions drawn from the defect studies of the last chapter. +Thus, it is concluded that increased temperatures is not exclusively usefull to accelerate the dynamics approximatively describing the structural evolution. +Moreover it can be considered a necessary condition to deviate the system out of equilibrium enabling the formation of 3C-SiC, which is obviously realized by a successive agglomeration of \cs{}. -Thus, employing increased temperatures is not exclusively usefull to accelerate the dynamics approximatively describing the structural evolution. -Moreover it can be considered a necessary condition to deviate the system out of equilibrium enabling the formation of 3C-SiC obviously realized by a successive agglomeration of substitutional C. -} +\ifnum1=0 -\subsection{Valuation of a practicable temperature limit} -\label{subsection:md:tval} +\section{Valuation of a practicable temperature limit} +\label{section:md:tval} The assumed applicability of increased temperature simulations as discussed above and the remaining absence of either agglomeration of substitutional C in low concentration simulations or amorphous to crystalline transition in high concentration simulations suggests to further increase the system temperature. -So far, the highest temperature applied corresponds to 95 \% of the absolute silicon melting temperature, which is 2450 K and specific to the Erhart/Albe potential. +So far, the highest temperature applied corresponds to 95 \% of the absolute Si melting temperature, which is 2450 K and specific to the Erhart/Albe potential. However, melting is not predicted to occur instantly after exceeding the melting point due to additionally required transition enthalpy and hysteresis behaviour. -To check for the possibly highest temperature at which a transition fails to appear plain silicon is heated up using a heating rate of $1\,^{\circ}\mathrm{C}/\text{ps}$. -Figure \ref{fig:md:fe_and_t} shows the free energy and temperature evolution in the region around the transition temperature. -Indeed a transition and the accompanying critical behaviour of the free energy is first observed at approximately 3125 K, which corresponds to 128 \% of the silicon melting temperature. -The difference in free energy is 0.58 eV per atom corresponding to $55.7 \text{ kJ/mole}$, which compares quite well to the silicon enthalpy of melting of $50.2 \text{ kJ/mole}$. +To check for the possibly highest temperature at which a transition fails to appear plain Si is heated up using a heating rate of $1\,^{\circ}\mathrm{C}/\text{ps}$. +Fig.~\ref{fig:md:fe_and_t} shows the free energy and temperature evolution in the region around the transition temperature. +Indeed a transition and the accompanying critical behaviour of the free energy is first observed at approximately 3125 K, which corresponds to 128 \% of the Si melting temperature. +The difference in free energy is 0.58 eV per atom corresponding to $55.7 \text{ kJ/mole}$, which compares quite well to the Si enthalpy of melting of $50.2 \text{ kJ/mole}$. The late transition probably occurs due to the high heating rate and, thus, a large hysteresis behaviour extending the temperature of transition. To avoid melting transitions in further simulations system temperatures well below the transition point are considered safe. -According to this study temperatures of 100 \% and 120 \% of the silicon melting point could be used. +According to this study temperatures of 100 \% and 120 \% of the Si melting point could be used. However, defects, which are introduced due to the insertion of C atoms are known to lower the transition point. Indeed simulations show melting transitions already at the melting point whenever C is inserted. -Thus, the system temperature of 95 \% of the silicon melting point is considered the maximum limit. -\begin{figure}[!t] +Thus, the system temperature of 95 \% of the Si melting point is considered the maximum limit. +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{fe_and_t.ps} +\includegraphics[width=0.7\textwidth]{fe_and_t.ps} \end{center} -\caption{Free energy and temperature evolution of plain silicon at temperatures in the region around the melting transition.} +\caption{Free energy and temperature evolution of plain Si at temperatures in the region around the melting transition.} \label{fig:md:fe_and_t} \end{figure} -\subsection{Long time scale simulations at maximum temperature} +\section{Long time scale simulations at maximum temperature} -As discussed in section~\ref{subsection:md:limit} and~\ref{subsection:md:inct} a further increase of the system temperature might help to overcome limitations of the short range potential and accelerate the dynamics involved in structural evolution. +As discussed in section~\ref{section:md:limit} and~\ref{section:md:inct} a further increase of the system temperature might help to overcome limitations of the short range potential and accelerate the dynamics involved in structural evolution. Furthermore these results indicate that increased temperatures are necessary to drive the system out of equilibrium enabling conditions needed for the formation of a metastable cubic polytype of SiC. -A maximum temperature to avoid melting is determined in section \ref{subsection:md:tval} to be 120 \% of the Si melting point but due to defects lowering the transition point a maximum temperature of 95 \% of the Si melting temperature is considered usefull. +A maximum temperature to avoid melting is determined in section \ref{section:md:tval} to be 120 \% of the Si melting point but due to defects lowering the transition point a maximum temperature of 95 \% of the Si melting temperature is considered usefull. This value is almost equal to the temperature of $2050\,^{\circ}\mathrm{C}$ already used in former simulations. Since the maximum temperature is reached the approach is reduced to the application of longer time scales. This is considered usefull since the estimated evolution of quality in the absence of the cooling down sequence in figure~\ref{fig:md:tot_si-c_q} predicts an increase in quality and, thus, structural evolution is liekyl to occur if the simulation is proceeded at maximum temperature. @@ -449,55 +451,52 @@ To speed up the simulation the initial simulation volume is reduced to 21 Si uni The 100 ps sequence after C insertion intended for structural evolution is exchanged by a 10 ns sequence, which is hoped to result in the occurence of infrequent processes and a subsequent phase transition. The return to lower temperatures is considered seperately. -\begin{figure}[!t] +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{c_in_si_95_v1_si-c.ps}\\ -\includegraphics[width=12cm]{c_in_si_95_v1_c-c.ps} +\includegraphics[width=0.7\textwidth]{c_in_si_95_v1_si-c.ps}\\ +\includegraphics[width=0.7\textwidth]{c_in_si_95_v1_c-c.ps} \end{center} \caption{Si-C (top) and C-C (bottom) radial distribution for low concentration simulations at 95 \% of the potential's Si melting point at different points in time of the simulation.} \label{fig:md:95_long_time_v1} \end{figure} -\begin{figure}[!t] +\begin{figure}[tp] \begin{center} -\includegraphics[width=12cm]{c_in_si_95_v2.ps} +\includegraphics[width=0.7\textwidth]{c_in_si_95_v2.ps} \end{center} \caption{Si-C and C-C radial distribution for high concentration simulations at 95 \% of the potential's Si melting point at different points in time of the simulation.} \label{fig:md:95_long_time_v2} \end{figure} -Figure \ref{fig:md:95_long_time_v1} shows the evolution in time of the radial distribution for Si-C and C-C pairs for a low C concentration simulation. +Fig.~\ref{fig:md:95_long_time_v1} shows the evolution in time of the radial distribution for Si-C and C-C pairs for a low C concentration simulation. Differences are observed for both types of atom pairs indeed indicating proceeding structural changes even well beyond 100 ps of simulation time. Peaks attributed to the existence of substitutional C increase and become more distinct. This finding complies with the predicted increase of quality evolution as explained earlier. -More and more C forms tetrahedral bonds to four Si neighbours occupying vacant Si sites. +More and more C forms tetrahedral bonds to four Si neighbors occupying vacant Si sites. However, no increase of the amount of total C-C pairs within the observed region can be identified. -Carbon, whether substitutional or as a dumbbell does not agglomerate within the simulated period of time visible by the unchanging area beneath the graphs. +C, whether substitutional or as a DB does not agglomerate within the simulated period of time visible by the unchanging area beneath the graphs. -Figure \ref{fig:md:95_long_time_v2} shows the evolution in time of the radial distribution for Si-C and C-C pairs for a high C concentration simulation. +Fig.~\ref{fig:md:95_long_time_v2} shows the evolution in time of the radial distribution for Si-C and C-C pairs for a high C concentration simulation. There are only small changes identifiable. A slight increase of the Si-C peak at approximately 0.36 nm attributed to the distance of substitutional C and the next but one Si atom along \hkl<1 1 0> is observed. In the same time the C-C peak at approximately 0.32 nm corresponding to the distance of two C atoms interconnected by a Si atom along \hkl<1 1 0> slightly decreases. Obviously the system preferes a slight increase of isolated substitutional C at the expense of incoherent C-Si-C precipitate configurations, which at a first glance actually appear as promising configurations in the precipitation event. -On second thoughts however, this process of splitting a C atom out of this structure is considered necessary in order to allow for the rearrangement of C atoms on substitutional lattice sites on the one hand and for C diffusion otherwise, which is needed to end up in a structure, in which one of the two fcc sublattices is composed out of carbon only. +On second thoughts however, this process of splitting a C atom out of this structure is considered necessary in order to allow for the rearrangement of C atoms on substitutional lattice sites on the one hand and for C diffusion otherwise, which is needed to end up in a structure, in which one of the two fcc sublattices is composed out of C only. For both, high and low concentration simulations the radial distribution converges as can be seen by the nearly identical graphs of the two most advanced configurations. Changes exist ... bridge to results after cooling down to 20 degree C. {\color{red}Todo: Cooling down to $20\,^{\circ}\mathrm{C}$ by $1\,^{\circ}\mathrm{C/s}$ in progress.} -{\color{red}Todo: Remember NVE simulations (prevent melting).} - -\subsection{Further accelerated dynamics approaches} +% todo evtl in ausblick -Since longer time scales are not sufficient \ldots +%\subsection{Further accelerated dynamics approaches} +%{\color{red}Todo: self-guided MD?} +%{\color{red}Todo: ART MD?\\ +%How about forcing a migration of a $V_2$ configuration to a constructed prec configuration, determine the saddle point configuration and continue the simulation from this configuration? +%} -{\color{red}Todo: self-guided MD?} - -{\color{red}Todo: other approaches?} - -{\color{red}Todo: ART MD?\\ -How about forcing a migration of a $V_2$ configuration to a constructed prec configuration, determine the saddle point configuration and continue the simulation from this configuration? -} +\fi \section{Conclusions concerning the SiC conversion mechanism} + diff --git a/posic/thesis/simulation.tex b/posic/thesis/simulation.tex index 04b4f18..0ebb1f8 100644 --- a/posic/thesis/simulation.tex +++ b/posic/thesis/simulation.tex @@ -169,6 +169,7 @@ Clearly, a competent parameter set is found, which is capabale of describing the % ref for experimental values! \section{Classical potential MD} +\label{section:classpotmd} The classical potential MD method is much less computationally costly compared to the highly accurate quantum-mechanical method. Thus, the method is capable of performing structural optimizations on large systems and MD calulations may be used to model a system over long time scales. diff --git a/posic/thesis/summary_outlook.tex b/posic/thesis/summary_outlook.tex index 66facbe..7bdc537 100644 --- a/posic/thesis/summary_outlook.tex +++ b/posic/thesis/summary_outlook.tex @@ -1,3 +1,3 @@ -\chapter{Summary and Outlook} +\chapter{Summary and conclusions} \label{chapter:summary} diff --git a/posic/thesis/thesis.tex b/posic/thesis/thesis.tex index 0b32cf1..599f944 100644 --- a/posic/thesis/thesis.tex +++ b/posic/thesis/thesis.tex @@ -37,8 +37,12 @@ % shortcuts \newcommand{\si}{Si$_{\text{i}}${}} -\newcommand{\ci}{Ci$_{\text{i}}${}} -\newcommand{\cs}{Ci$_{\text{s}}${}} +\newcommand{\ci}{C$_{\text{i}}${}} +\newcommand{\cs}{C$_{\text{s}}${}} +\newcommand{\degc}[1]{\unit[#1]{$^{\circ}$C}{}} +\newcommand{\distn}[1]{\unit[#1]{nm}{}} +\newcommand{\dista}[1]{\unit[#1]{\AA}{}} +\newcommand{\perc}[1]{\unit[#1]{\%}{}} % (re)new commands \newcommand{\printimg}[5]{% @@ -114,7 +118,7 @@ \include{d_tersoff} \include{vasp_patch} \include{code} -\include{publications} +%\include{publications} \backmatter{} \include{literature} -- 2.20.1