From: hackbard Date: Wed, 15 Jun 2011 13:35:11 +0000 (+0200) Subject: started to merge ... X-Git-Url: https://hackdaworld.org/gitweb/?p=lectures%2Flatex.git;a=commitdiff_plain;h=5c4559cc0c6a9122d8e354adb8153a0b21d6fa24 started to merge ... --- diff --git a/posic/publications/sic_prec_merge.tex b/posic/publications/sic_prec_merge.tex new file mode 100644 index 0000000..e069fb9 --- /dev/null +++ b/posic/publications/sic_prec_merge.tex @@ -0,0 +1,557 @@ +\documentclass[prb,twocolumn,superscriptaddress,a4paper,showkeys,showpacs]{revtex4} +\usepackage{graphicx} +\usepackage{subfigure} +\usepackage{dcolumn} +\usepackage{booktabs} +\usepackage{units} +\usepackage{amsmath} +\usepackage{amsfonts} +\usepackage{amssymb} + +% additional stuff +\usepackage{miller} + +\begin{document} + +\title{Combined ab initio and classical potential simulation study on the silicon carbide precipitation in silicon} +\author{F. Zirkelbach} +\author{B. Stritzker} +\affiliation{Experimentalphysik IV, Universit\"at Augsburg, 86135 Augsburg, Germany} +\author{K. Nordlund} +\affiliation{Department of Physics, University of Helsinki, 00014 Helsinki, Finland} +\author{J. K. N. Lindner} +\author{W. G. Schmidt} +\author{E. Rauls} +\affiliation{Department Physik, Universit\"at Paderborn, 33095 Paderborn, Germany} + +\begin{abstract} +Atomistic simulations on the silicon carbide precipitation in bulk silicon employing both, classical potential and first-principles methods are presented. +These aime to clarify a controversy concerning the precipitation mechanism as revealed from literature. +% +For the quantum-mechanical treatment, basic processes assumed in the precipitation process are calculated in feasible systems of small size. +The migration mechanism of a carbon \hkl<1 0 0> interstitial and silicon \hkl<1 1 0> self-interstitial in otherwise defect-free silicon using density functional theory calculations are investigated. +The influence of a nearby vacancy, another carbon interstitial and a substitutional defect as well as a silicon self-interstitial has been investigated systematically. +Interactions of various combinations of defects have been characterized including a couple of selected migration pathways within these configurations. +Almost all of the investigated pairs of defects tend to agglomerate allowing for a reduction in strain. +The formation of structures involving strong carbon-carbon bonds turns out to be very unlikely. +In contrast, substitutional carbon occurs in all probability. +A long range capture radius has been observed for pairs of interstitial carbon as well as interstitial carbon and vacancies. +A rather small capture radius is predicted for substitutional carbon and silicon self-interstitials. +We derive conclusions on the precipitation mechanism of silicon carbide in bulk silicon and discuss conformability to experimental findings. +% +Furthermore, results of the accurate first-principles calculations on defects and carbon diffusion in silicon are compared to results of classical potential simulations revealing significant limitations of the latter method. +An approach to work around this problem is proposed. +Finally, results of the classical potential molecular dynamics simulations of large systems are discussed, which allow to draw further conclusions on the precipitation mechanism of silicon carbide in silicon. +\end{abstract} + +\keywords{point defects, migration, interstitials, first-principles calculations, classical potentials, molecular dynamics, silicon carbide, ion implantation} +\pacs{61.72.J-,61.72.-y,66.30.Lw,66.30.-h,31.15.A-,31.15.xv,34.20.Cf,61.72.uf} +\maketitle + +% -------------------------------------------------------------------------------- +\section{Introduction} + +The wide band gap semiconductor silicon carbide (SiC) is well known for its outstanding physical and chemical properties. +The high breakdown field, saturated electron drift velocity and thermal conductivity in conjunction with the unique thermal and mechanical stability as well as radiation hardness makes SiC a suitable material for high-temperature, high-frequency and high-power devices operational in harsh and radiation-hard environments\cite{edgar92,morkoc94,wesch96,capano97,park98}. +Different modifications of SiC exist, which solely differ in the one-dimensional stacking sequence of identical, close-packed SiC bilayers\cite{fischer90}. +Different polytypes exhibit different properties, in which the cubic phase of SiC (3C-SiC) shows increased values for the thermal conductivity and breakdown field compared to other polytypes\cite{wesch96}, which is, thus, most effective for high-performance electronic devices. +Much progress has been made in 3C-SiC thin film growth by chemical vapor deposition (CVD) and molecular beam epitaxy (MBE) on hexagonal SiC\cite{powell90,fissel95,fissel95_apl} and Si\cite{nishino83,nishino87,kitabatake93,fissel95_apl} substrates. +However, the frequent occurrence of defects such as twins, dislocations and double position boundaries remains a challenging problem. +Apart from these methods, high-dose carbon implantation into crystalline silicon (c-Si) with subsequent or in situ annealing was found to result in SiC microcrystallites in Si\cite{borders71}. +Utilized and enhanced, ion beam synthesis (IBS) has become a promising method to form thin SiC layers of high quality and exclusively of the 3C polytype embedded in and epitactically aligned to the Si host featuring a sharp interface\cite{lindner99,lindner01,lindner02}. + +However, the process of the formation of SiC precipitates in Si during C implantation is not yet fully understood. +\begin{figure} +\begin{center} +\subfigure[]{\label{fig:hrem:c-si}\includegraphics[width=0.48\columnwidth]{tem_c-si-db.eps}} +\subfigure[]{\label{fig:hrem:sic}\includegraphics[width=0.48\columnwidth]{tem_3c-sic.eps}} +\end{center} +\caption{High resolution transmission electron microscopy (HREM) micrographs\cite{lindner99_2} of agglomerates of C-Si dimers showing dark contrasts and otherwise undisturbed Si lattice fringes (a) and equally sized Moir\'e patterns indicating 3C-SiC precipitates (b).} +\label{fig:hrem} +\end{figure} +High resolution transmission electron microscopy (HREM) studies\cite{werner96,werner97,eichhorn99,lindner99_2,koegler03} suggest the formation of C-Si dimers (dumbbells) on regular Si lattice sites, which agglomerate into large clusters indicated by dark contrasts and otherwise undisturbed Si lattice fringes in HREM, as can be seen in Fig.~\ref{fig:hrem:c-si}. +A topotactic transformation into a 3C-SiC precipitate occurs once a critical radius of 2 nm to 4 nm is reached, which is manifested by the disappearance of the dark contrasts in favor of Moir\'e patterns (Fig.~\ref{fig:hrem:sic}) due to the lattice mismatch of \unit[20]{\%} of the 3C-SiC precipitate and c-Si. +The insignificantly lower Si density of SiC ($\approx \unit[4]{\%}$) compared to c-Si results in the emission of only a few excess Si atoms. +In contrast, investigations of strained Si$_{1-y}$C$_y$/Si heterostructures formed by IBS and solid-phase epitaxial regorowth\cite{strane94} as well as MBE\cite{guedj98}, which incidentally involve the formation of SiC nanocrystallites, suggest an initial coherent precipitation by agglomeration of substitutional instead of interstitial C. +Coherency is lost once the increasing strain energy of the stretched SiC structure surpasses the interfacial energy of the incoherent 3C-SiC precipitate and the Si substrate. +These two different mechanisms of precipitation might be attributed to the respective method of fabrication. +While in CVD and MBE surface effects need to be taken into account, SiC formation during IBS takes place in the bulk of the Si crystal. +However, in another IBS study Nejim et~al.\cite{nejim95} propose a topotactic transformation that is likewise based on the formation of substitutional C. +The formation of substitutional C, however, is accompanied by Si self-interstitial atoms that previously occupied the lattice sites and a concurrent reduction of volume due to the lower lattice constant of SiC compared to Si. +Both processes are believed to compensate one another. +% +Solving this controversy and understanding the effective underlying processes will enable significant technological progress in 3C-SiC thin film formation driving the superior polytype for potential applications in high-performance electronic device production. +It will likewise offer perspectives for processes that rely upon prevention of precipitation events, e.g. the fabrication of strained pseudomorphic Si$_{1-y}$C$_y$ heterostructures\cite{strane96,laveant2002}. + +Atomistic simulations offer a powerful tool to study materials on a microscopic level providing detailed insight not accessible by experiment. +% +A lot of theoretical work has been done on intrinsic point defects in Si\cite{bar-yam84,bar-yam84_2,car84,batra87,bloechl93,tang97,leung99,colombo02,goedecker02,al-mushadani03,hobler05,sahli05,posselt08,ma10}, threshold displacement energies in Si\cite{mazzarolo01,holmstroem08} important in ion implantation, C defects and defect reactions in Si\cite{tersoff90,dal_pino93,capaz94,burnard93,leary97,capaz98,zhu98,mattoni2002,park02,jones04}, the SiC/Si interface\cite{chirita97,kitabatake93,cicero02,pizzagalli03} and defects in SiC\cite{bockstedte03,rauls03a,gao04,posselt06,gao07}. +However, none of the mentioned studies consistently investigates entirely the relevant defect structures and reactions concentrated on the specific problem of 3C-SiC formation in C implanted Si. +% but mattoni2002 actually did a lot. maybe this should be mentioned!^M +In fact, in a combined analytical potential molecular dynamics and ab initio study\cite{mattoni2002} the interaction of substitutional C with Si self-interstitials and C interstitials is evaluated. +However, investigations are, first of all, restricted to interaction chains along the \hkl[1 1 0] and \hkl[-1 1 0] direction, secondly lacking combinations of C interstitials and, finally, not considering migration barriers providing further information on the probability of defect agglomeration. + +In particular, molecular dynamics (MD) constitutes a suitable technique to investigate their dynamical and structural properties. +Modelling the processes mentioned above requires the simulation of a large number of atoms ($\approx 10^5-10^6$), which inevitably dictates the atomic interaction to be described by computationally efficient classical potentials. +These are, however, less accurate compared to quantum-mechanical methods and their applicability for the description of the physical problem has to be verified beforehand. +The most common empirical potentials for covalent systems are the Stillinger-Weber\cite{stillinger85}, Brenner\cite{brenner90}, Tersoff\cite{tersoff_si3} and environment-dependent interatomic potential\cite{bazant96,bazant97,justo98}. +These potentials are assumed to be reliable for large-scale simulations\cite{balamane92,huang95,godet03} on specific problems under investigation providing insight into phenomena that are otherwise not accessible by experimental or first-principles methods. +Until recently\cite{lucas10}, a parametrization to describe the C-Si multicomponent system within the mentioned interaction models did only exist for the Tersoff\cite{tersoff_m} and related potentials, e.g. the one by Gao and Weber\cite{gao02} as well as the one by Erhart and Albe\cite{albe_sic_pot}. +All these potentials are short range potentials employing a cut-off function, which drops the atomic interaction to zero in between the first and second nearest neighbor distance. +It was shown that the Tersoff potential properly describes binding energies of combinations of C defects in Si\cite{mattoni2002}. +However, investigations of brittleness in covalent materials\cite{mattoni2007} identified the short range character of these potentials to be responsible for overestimated forces necessary to snap the bond of two neighbored atoms. +In a previous study\cite{zirkelbach10}, we determined the influence on the migration barrier for C diffusion in Si. +Using the Erhart/Albe (EA) potential\cite{albe_sic_pot}, an overestimated barrier height compared to ab initio calculations and experiment is obtained. +A proper description of C diffusion, however, is crucial for the problem under study. + +In this work, a combined ab initio and empirical potential simulation study on the initially mentioned SiC precipitation mechanism has been performed. +% +By first-principles atomistic simulations this work aims to shed light on basic processes involved in the precipitation mechanism of SiC in Si. +During implantation defects such as vacancies (V), substitutional C (C$_{\text{s}}$), interstitial C (C$_{\text{i}}$) and Si self-interstitials (Si$_{\text{i}}$) are created, which play a decisive role in the precipitation process. +A systematic investigation of density functional theory (DFT) calculations of the structure, energetics and mobility of carbon defects in silicon as well as the influence of other point defects in the surrounding is presented. +% +Furthermore, highly accurate quantum-mechanical results have been used to identify shortcomings of the classical potentials, which are then taken into account in these type of simulations. + +% -------------------------------------------------------------------------------- +\section{Methodology} +% ----- DFT ------ +The first-principles DFT calculations have been performed with the plane-wave based Vienna ab initio Simulation package (VASP)\cite{kresse96}. +The Kohn-Sham equations were solved using the generalized-gradient exchange-correlation functional approximation proposed by Perdew and Wang\cite{perdew86,perdew92}. +The electron-ion interaction is described by norm-conserving ultra-soft pseudopotentials\cite{hamann79} as implemented in VASP\cite{vanderbilt90}. +Throughout this work, an energy cut-off of \unit[300]{eV} was used to expand the wave functions into the plane-wave basis. +%Sampling of the Brillouin zone was restricted to the $\Gamma$-point. +To reduce the computational effort sampling of the Brillouin zone was restricted to the $\Gamma$-point, which has been shown to yield reliable results\cite{dal_pino93}. +The defect structures and the migration paths were modelled in cubic supercells with a side length of \unit[1.6]{nm} containing $216$ Si atoms. +Formation energies and structures are reasonably converged with respect to the system size. +The ions and cell shape were allowed to change in order to realize a constant pressure simulation. +The observed changes in volume were less than \unit[0.2]{\%} of the volume indicating a rather low dependence of the results on the ensemble choice. +Ionic relaxation was realized by the conjugate gradient algorithm. +Spin polarization has been fully accounted for. + +% ------ Albe potential --------- +For the classical potential calculations, defect structures were modeled in a supercell of nine Si lattice constants in each direction consisting of 5832 Si atoms. +Reproducing the SiC precipitation was attempted by the successive insertion of 6000 C atoms (the number necessary to form a 3C-SiC precipitate with a radius of $\approx 3.1$ nm) into the Si host, which has a size of 31 Si unit cells in each direction consisting of 238328 Si atoms. +At constant temperature 10 atoms were inserted at a time. +Three different regions within the total simulation volume were considered for a statistically distributed insertion of the C atoms: $V_1$ corresponding to the total simulation volume, $V_2$ corresponding to the size of the precipitate and $V_3$, which holds the necessary amount of Si atoms of the precipitate. +After C insertion the simulation has been continued for \unit[100]{ps} and is cooled down to \unit[20]{$^{\circ}$C} afterwards. +A Tersoff-like bond order potential by Erhart and Albe (EA)\cite{albe_sic_pot} has been utilized, which accounts for nearest neighbor interactions realized by a cut-off function dropping the interaction to zero in between the first and second nearest neighbor distance. +The potential was used as is, i.e. without any repulsive potential extension at short interatomic distances. +Constant pressure simulations are realized by the Berendsen barostat\cite{berendsen84} using a time constant of \unit[100]{fs} and a bulk modulus of \unit[100]{GPa} for Si. +The temperature was kept constant by the Berendsen thermostat\cite{berendsen84} with a time constant of \unit[100]{fs}. +Integration of the equations of motion was realized by the velocity Verlet algorithm\cite{verlet67} and a fixed time step of \unit[1]{fs}. +For structural relaxation of defect structures the same algorithm was used with the temperature set to 0 K. + +The formation energy $E-N_{\text{Si}}\mu_{\text{Si}}-N_{\text{C}}\mu_{\text{C}}$ of a defect configuration is defined by choosing SiC as a particle reservoir for the C impurity, i.e. the chemical potentials are determined by the cohesive energies of a perfect Si and SiC supercell after ionic relaxation. +% +This corresponds to the definition utilized in another study on C defects in Si\cite{dal_pino93} that we compare our results to. +% +Migration and recombination pathways have been investigated utilizing the constraint conjugate gradient relaxation technique\cite{kaukonen98}. +% +While not guaranteed to find the true minimum energy path, the method turns out to identify reasonable pathways for the investigated structures. +% +Time constants of \unit[1]{fs}, which corresponds to direct velocity scaling, and \unit[100]{fs}, which results in weaker coupling to the heat bath allowing the diffusing atoms to take different pathways, were used for the Berendsen thermostat for structural relaxation within the migration calculations utilizing classical potentials. +% +The binding energy of a defect pair is given by the difference of the formation energy of the complex and the sum of the two separated defect configurations. +Accordingly, energetically favorable configurations result in binding energies below zero while unfavorable configurations show positive values for the binding energy. +The interaction strength, i.e. the absolute value of the binding energy, approaches zero for increasingly non-interacting isolated defects. + +\section{Results} + +\subsection{Carbon and silicon defect configurations} + +\begin{figure} +\begin{minipage}[t]{0.32\columnwidth} +\underline{Si$_{\text{i}}$ \hkl<1 1 0> DB}\\ +\includegraphics[width=\columnwidth]{si110.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{Si$_{\text{i}}$ hexagonal}\\ +\includegraphics[width=\columnwidth]{sihex.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{Si$_{\text{i}}$ tetrahedral}\\ +\includegraphics[width=\columnwidth]{sitet.eps} +\end{minipage}\\ +\begin{minipage}[t]{0.32\columnwidth} +\underline{Si$_{\text{i}}$ \hkl<1 0 0> DB}\\ +\includegraphics[width=\columnwidth]{si100.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{Vacancy}\\ +\includegraphics[width=\columnwidth]{sivac.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{C$_{\text{s}}$}\\ +\includegraphics[width=\columnwidth]{csub.eps} +\end{minipage}\\ +\begin{minipage}[t]{0.32\columnwidth} +\underline{C$_{\text{i}}$ \hkl<1 0 0> DB}\\ +\includegraphics[width=\columnwidth]{c100.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{C$_{\text{i}}$ \hkl<1 1 0> DB}\\ +\includegraphics[width=\columnwidth]{c110.eps} +\end{minipage} +\begin{minipage}[t]{0.32\columnwidth} +\underline{C$_{\text{i}}$ bond-centered}\\ +\includegraphics[width=\columnwidth]{cbc.eps} +\end{minipage} +\caption{Configurations of Si and C point defects in Si. Si and C atoms are illustrated by yellow and gray spheres respectively. Bonds are drawn whenever considered appropriate to ease identifying defect structures for the reader. Dumbbell configurations are abbreviated by DB.} +\label{fig:sep_def} +\end{figure} +Table~\ref{table:sep_eof} summarizes the formation energies of relevant defect structures for the EA and DFT calculations. +The respective structures are shown in Fig.~\ref{fig:sep_def}. +% +\begin{table*} +\begin{ruledtabular} +\begin{tabular}{l c c c c c c c c c} + & Si$_{\text{i}}$ \hkl<1 1 0> DB & Si$_{\text{i}}$ H & Si$_{\text{i}}$ T & Si$_{\text{i}}$ \hkl<1 0 0> DB & V & C$_{\text{s}}$ & C$_{\text{i}}$ \hkl<1 0 0> DB & C$_{\text{i}}$ \hkl<1 1 0> DB & C$_{\text{i}}$ BC \\ +\hline + \multicolumn{10}{l}{Present study} \\ + VASP & 3.39 & 3.42 & 3.77 & 4.41 & 3.63 & 1.95 & 3.72 & 4.16 & 4.66 \\ + Erhart/Albe & 4.39 & 4.48$^*$ & 3.40 & 5.42 & 3.13 & 0.75 & 3.88 & 5.18 & 5.59$^*$ \\ + \multicolumn{10}{l}{Other ab initio studies} \\ + Ref.\cite{al-mushadani03} & 3.40 & 3.45 & - & - & 3.53 & - & - & - & - \\ + Ref.\cite{leung99} & 3.31 & 3.31 & 3.43 & - & - & - & - & - & - \\ + Ref.\cite{dal_pino93,capaz94} & - & - & - & - & - & 1.89\cite{dal_pino93} & x & - & x+2.1\cite{capaz94} +\end{tabular} +\end{ruledtabular} +\caption{Formation energies of C and Si point defects in c-Si determined by classical potential and ab initio methods. The formation energies are given in electron volts. T denotes the tetrahedral and BC the bond-centered configuration. Subscript i and s indicates the interstitial and substitutional configuration. Dumbbell configurations are abbreviated by DB. Formation energies for unstable configurations obtained by classical potential MD 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{table:sep_eof} +\end{table*} + +Although discrepancies exist, calssical potential and first-principles methods depict the correct order of the formation energies with regard to C defects in Si. +Substitutional C (C$_{\text{s}}$) constitutes the energetically most favorable defect configuration. +Since the C atom occupies an already vacant Si lattice site, C$_{\text{s}}$ is not an interstitial defect. +The quantum-mechanical result agrees well with the result of another ab initio study\cite{dal_pino93}. +Clearly, the empirical potential underestimates the C$_{\text{s}}$ formation energy. +The C interstitial defect with the lowest energy of formation has been found to be the C-Si \hkl<1 0 0> interstitial dumbbell (C$_{\text{i}}$ \hkl<1 0 0> DB), which, thus, constitutes the ground state of an additional C impurity in otherwise perfect c-Si. +This finding is in agreement with several theoretical\cite{burnard93,leary97,dal_pino93,capaz94} and experimental\cite{watkins76,song90} investigations. +However, to our best knowledge, no energy of formation based on first-principles calculations has yet been explicitly stated in literature for the ground-state configuration. +% +Astonishingly EA and DFT predict almost equal formation energies. +There are, however, geometric differences with regard to the DB position within the tetrahedron spanned by the four neighbored Si atoms, as already reported in a previous study\cite{zirkelbach10}. +Since the energetic description is considered more important than the structural description, minor discrepancies of the latter are assumed non-problematic. +The second most favorable configuration is the C$_{\text{i}}$ \hkl<1 1 0> DB followed by the C$_{\text{i}}$ bond-centered (BC) configuration. +For both configurations EA overestimates the energy of formation by approximately \unit[1]{eV} compared to DFT. +Thus, nearly the same difference in energy has been observed for these configurations in both methods. +However, we have found the BC configuration to constitute a saddle point within the EA description relaxing into the \hkl<1 1 0> configuration. +Due to the high formation energy of the BC defect resulting in a low probability of occurrence of this defect, the wrong description is not posing a serious limitation of the EA potential. +A more detailed discussion of C defects in Si modeled by EA and DFT including further defect configurations can be found in our recently published article\cite{zirkelbach10}. + +Regarding intrinsic defects in Si, classical potential and {\em ab initio} methods predict energies of formation that are within the same order of magnitude. +% +However discrepancies exist. +Quantum-mechanical results reveal the Si$_{\text{i}}$ \hkl<1 1 0> DB to compose the energetically most favorable configuration closely followed by the hexagonal and tetrahedral configuration, which is the consensus view for Si$_{\textrm{i}}$ and compares well to results from literature\cite{leung99,al-mushadani03}. +The EA potential does not reproduce the correct ground state. +Instead the tetrahedral defect configuration is favored. +This limitation is assumed to arise due to the cut-off. +In the tetrahedral configuration the second neighbors are only slightly more distant than the first neighbors, which creates the particular problem. +Indeed, an increase of the cut-off results in increased values of the formation energies\cite{albe_sic_pot}, which is most significant for the tetrahedral configuration. +The same issue has already been discussed by Tersoff\cite{tersoff90} with regard to the description of the tetrahedral C defect using his potential. +% h unstable +% todo - really do this?!?! +The hexagonal configuration is not stable within the classical potential calculations opposed to results of the authors of the potential\cite{albe_sic_pot}. +In the first two pico seconds, while kinetic energy is decoupled from the system, the \si{} seems to condense at the hexagonal site. +The formation energy of \unit[4.48]{eV} is determined by this low kinetic energy configuration shortly before the relaxation process starts. +The \si{} atom then begins to slowly move towards an energetically more favorable position very close to the tetrahedral one but slightly displaced along the three coordinate axes. +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. +As has been shown, variations of this defect 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\cite{zirkelbach09}. +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 among the artificial and tetrahedral configuration is found to be smaller than \unit[0.2]{eV}. +Hence, these artifacts have a negligible influence in finite temperature simulations. +% nevertheless ... +While not completely rendering impossible further, more challenging empirical potential studies on large systems, these artifacts have to be taken into account in the following investigations of defect combinations. + +% todo - where to put this! +% spin polarization and C mobility +Instead of giving an explicit value of the energy of formation, Capaz et al.\cite{capaz94}, investigating migration pathways of the C$_{\text{i}}$ \hkl<1 0 0> DB, find this defect to be \unit[2.1]{eV} lower in energy than the BC configuration. +The BC configuration is claimed to constitute the saddle point within the C$_{\text{i}}$ \hkl[0 0 -1] DB migration path residing in the \hkl(1 1 0) plane and, thus, interpreted as the barrier of migration for the respective path. +However, the present study indicates a local minimum state for the BC defect if spin polarized calculations are performed resulting in a net magnetization of two electrons localized in a torus around the C atom. +Another DFT calculation without fully accounting for the electron spin results in the smearing of a single electron over two non-degenerate Kohn-Sham states and an increase of the total energy by \unit[0.3]{eV} for the BC configuration. +Regardless of the rather small correction of \unit[0.3]{eV} due to the spin, the difference we found is much smaller (\unit[0.94]{eV}), which would nicely compare to experimentally observed migration barriers of \unit[0.70-0.87]{eV}\cite{lindner06,tipping87,song90}. +However, since the BC configuration constitutes a real local minimum another barrier exists, which is about \unit[1.2]{eV} in height. +% +Indeed, Capaz~et~al. propose another path and find it to be the lowest in energy\cite{capaz94}, in which a C$_{\text{i}}$ \hkl[0 0 -1] DB migrates to a C$_{\text{i}}$ \hkl[0 -1 0] DB located at the neighbored Si lattice site in \hkl[1 1 -1] direction. +Calculations in this work reinforce this path by an additional improvement of the quantitative conformance of the barrier height (\unit[0.90]{eV}) to experimental values. +A more detailed description can be found in a previous study\cite{zirkelbach10}. +Next to the C$_{\text{i}}$ BC configuration the vacancy and Si$_{\text{i}}$ \hkl<1 0 0> DB have to be treated by taking into account the spin of the electrons. +For the vacancy the net spin up electron density is localized in caps at the four surrounding Si atoms directed towards the vacant site. +In the Si$_{\text{i}}$ \hkl<1 0 0> DB configuration the net spin up density is localized in two caps at each of the two DB atoms perpendicularly aligned to the bonds to the other two Si atoms respectively. +No other configuration, within the ones that are mentioned, is affected. + +% todo- where to put mobility +Concerning the mobility of the ground state Si$_{\text{i}}$, an activation energy of \unit[0.67]{eV} for the transition of the Si$_{\text{i}}$ \hkl[0 1 -1] to \hkl[1 1 0] DB located at the neighbored Si lattice site in \hkl[1 1 -1] direction is obtained by first-principles calculations. +Further quantum-mechanical investigations revealed a barrier of \unit[0.94]{eV} for the Si$_{\text{i}}$ \hkl[1 1 0] DB to Si$_{\text{i}}$ H, \unit[0.53]{eV} for the Si$_{\text{i}}$ \hkl[1 1 0] DB to Si$_{\text{i}}$ T and \unit[0.35]{eV} for the Si$_{\text{i}}$ H to Si$_{\text{i}}$ T transition. +These are of the same order of magnitude than values derived from other ab initio studies\cite{bloechl93,sahli05}. + + +\subsection{Formation energies of C$_{\text{i}}$ and C$_{\text{s}}$-Si$_{\text{i}}$} + +As has been shown in a previous study\cite{zirkelbach11a}, the energetically most favorable configuration of C$_{\text{s}}$ and Si$_{\text{i}}$ is obtained for C$_{\text{s}}$ located at the neighbored lattice site along the \hkl<1 1 0> bond chain of a Si$_{\text{i}}$ \hkl<1 1 0> DB. +However, the energy of formation is slightly higher than that of the C$_{\text{i}}$ \hkl<1 0 0> DB, which constitutes the ground state for a C impurity introduced into otherwise perfect c-Si. + +For a possible clarification of the controversial views on the participation of C$_{\text{s}}$ in the precipitation mechanism by classical potential simulations, test calculations need to ensure the proper description of the relative formation energies of combined structures of C$_{\text{s}}$ and Si$_{\text{i}}$ compared to C$_{\text{i}}$. +This is particularly important since the energy of formation of C$_{\text{s}}$ is drastically underestimated by the EA potential. +A possible occurrence of C$_{\text{s}}$ could then be attributed to a lower energy of formation of the C$_{\text{s}}$-Si$_{\text{i}}$ combination due to the low formation energy of C$_{\text{s}}$, which is obviously wrong. + +Since quantum-mechanical calculations reveal the Si$_{\text{i}}$ \hkl<1 1 0> DB as the ground state configuration of Si$_{\text{i}}$ in Si it is assumed to provide the energetically most favorable configuration in combination with C$_{\text{s}}$. +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 VASP and EA calculations are summarized in Table~\ref{tab:defect_combos}. +\begin{table} +\begin{ruledtabular} +\begin{tabular}{l c c c} + & C$_{\text{i}}$ \hkl<1 0 0> & C$_{\text{s}}$ \& Si$_{\text{i}}$ \hkl<1 1 0> & C$_{\text{s}}$ \& Si$_{\text{i}}$ T\\ +\hline + VASP & 3.72 & 4.37 & 4.17$^{\text{a}}$/4.99$^{\text{b}}$/4.96$^{\text{c}}$ \\ + Erhart/Albe & 3.88 & 4.93 & 5.25$^{\text{a}}$/5.08$^{\text{b}}$/4.43$^{\text{c}}$ +\end{tabular} +\end{ruledtabular} +\caption{Formation energies of defect configurations of a single C impurity in otherwise perfect c-Si determined by classical potential and ab initio methods. The formation energies are given in electron volts. T denotes the tetrahedral and the subscripts i and s indicate the interstitial and substitutional configuration. Superscripts a, b and c denote configurations of C$_{\text{s}}$ located at the first, second and third nearest neighbored lattice site with respect to the Si$_{\text{i}}$ atom.} +\label{tab:defect_combos} +\end{table} +Obviously the EA potential properly describes the relative energies of formation. +Combined structures of C$_{\text{s}}$ and Si$_{\text{i}}$ T are energetically less favorable than the ground state C$_{\text{i}}$ \hkl<1 0 0> DB configuration. +With increasing separation distance the energies of formation decrease. +However, even for non-interacting defects, the energy of formation, which is then given by the sum of the formation energies of the separated defects (\unit[4.15]{eV}) is still higher than that of the C$_{\text{i}}$ \hkl<1 0 0> DB. +Unexpectedly, the structure of a Si$_{\text{i}}$ \hkl<1 1 0> DB and a neighbored C$_{\text{s}}$, which is the most favored configuration of a C$_{\text{s}}$ and Si$_{\text{i}}$ DB according to quantum-mechanical calculations\cite{zirkelbach11a}, likewise constitutes an energetically favorable configuration within the EA description, which is even preferred over the two least separated configurations of C$_{\text{s}}$ and Si$_{\text{i}}$ T. +This is attributed to an effective reduction in strain enabled by the respective combination. +Quantum-mechanical results reveal a more favorable energy of fomation for the C$_{\text{s}}$ and Si$_{\text{i}}$ T (a) configuration. +However, this configuration is unstable involving a structural transition into the C$_{\text{i}}$ \hkl<1 1 0> interstitial, thus, not maintaining the tetrahedral Si nor the substitutional C defect. + +Thus, the underestimated energy of formation of C$_{\text{s}}$ within the EA calculation does not pose a serious limitation in the present context. +Since C is introduced into a perfect Si crystal and the number of particles is conserved in simulation, the creation of C$_{\text{s}}$ is accompanied by the creation of Si$_{\text{i}}$, which is energetically less favorable than the ground state, i.e. the C$_{\text{i}}$ \hkl<1 0 0> DB configuration, for both, the EA and ab initio treatment. +In either case, no configuration more favorable than the C$_{\text{i}}$ \hkl<1 0 0> DB has been found. +Thus, a proper description with respect to the relative energies of formation is assumed for the EA potential. + +\subsection{Carbon mobility} +\label{subsection:cmob} + +To accurately model the SiC precipitation, which involves the agglomeration of C, a proper description of the migration process of the C impurity is required. +As shown in a previous study\cite{zirkelbach10}, quantum-mechanical results properly describe the C$_{\text{i}}$ \hkl<1 0 0> DB diffusion resulting in a migration barrier height of \unit[0.90]{eV}, excellently matching experimental values of \unit[0.70-0.87]{eV}\cite{lindner06,tipping87,song90} and, for this reason, reinforcing the respective migration path as already proposed by Capaz et~al.\cite{capaz94}. +During transition a C$_{\text{i}}$ \hkl[0 0 -1] DB migrates towards a C$_{\text{i}}$ \hkl[0 -1 0] DB located at the neighbored lattice site in \hkl[1 1 -1] direction. +However, it turned out that the description fails if the EA potential is used, which overestimates the migration barrier (\unit[2.2]{eV}) by a factor of 2.4. +In addition a different diffusion path is found to exhibit the lowest migration barrier. +A C$_{\text{i}}$ \hkl[0 0 -1] DB turns into the \hkl[0 0 1] configuration at the neighbored lattice site. +The transition involves the C$_{\text{i}}$ BC configuration, which, however, was found to be unstable relaxing into the C$_{\text{i}}$ \hkl<1 1 0> DB configuration. +If the migration is considered to occur within a single step, the kinetic energy of \unit[2.2]{eV} is sufficient to turn the \hkl<1 0 0> DB into the BC and back into a \hkl<1 0 0> DB configuration. +If, on the other hand, a two step process is assumed, the BC configuration will most probably relax into the C$_{\text{i}}$ \hkl<1 1 0> DB configuration resulting in different relative energies of the intermediate state and the saddle point. +For the latter case a migration path, which involves a C$_{\text{i}}$ \hkl<1 1 0> DB configuration, is proposed and displayed in Fig.~\ref{fig:mig}. +\begin{figure} +\begin{center} +\includegraphics[width=\columnwidth]{110mig.ps} +\end{center} +\caption{Migration barrier and structures of the \hkl[0 0 -1] DB (left) to the \hkl[0 -1 0] DB (right) transition involving the \hkl[1 1 0] DB (center) configuration. Migration simulations were performed utilizing time constants of \unit[1]{fs} (solid line) and \unit[100]{fs} (dashed line) for the Berendsen thermostat.} +\label{fig:mig} +\end{figure} +Approximately \unit[2.24]{eV} are needed to turn the C$_{\text{i}}$ \hkl[0 0 -1] DB into the C$_{\text{i}}$ \hkl[1 1 0] DB located at the neighbored lattice site in \hkl[1 1 -1] direction. +Another barrier of \unit[0.90]{eV} exists for the rotation into the C$_{\text{i}}$ \hkl[0 -1 0] DB configuration for the path obtained with a time constant of \unit[100]{fs} for the Berendsen thermostat. +Roughly the same amount would be necessary to excite the C$_{\text{i}}$ \hkl[1 1 0] DB to the BC configuration (\unit[0.40]{eV}) and a successive migration into the \hkl[0 0 1] DB configuration (\unit[0.50]{eV}) as displayed in our previous study\cite{zirkelbach10}. +The former diffusion process, however, would more nicely agree with the ab initio path, since the migration is accompanied by a rotation of the DB orientation. +By considering a two step process and assuming equal preexponential factors for both diffusion steps, the probability of the total diffusion event is given by $\exp(\frac{\unit[2.24]{eV}+\unit[0.90]{eV}}{k_{\text{B}}T})$, which corresponds to a single diffusion barrier that is 3.5 times higher than the barrier obtained by ab initio calculations. + +Accordingly, the effective barrier of migration of C$_{\text{i}}$ is overestimated by a factor of 2.4 to 3.5 compared to the highly accurate quantum-mechanical methods. +This constitutes a serious limitation that has to be taken into account for modeling the C-Si system using the otherwise quite promising EA potential. + +\subsection{Molecular dynamics simulations} + +Fig.~\ref{fig:450} shows the radial distribution functions of simulations, in which C was inserted at \unit[450]{$^{\circ}$C}, an operative and efficient temperature in IBS\cite{lindner99}, for all three insertion volumes. +\begin{figure} +\begin{center} +\subfigure[]{\label{fig:450:a} +\includegraphics[width=\columnwidth]{sic_prec_450_si-si_c-c.ps} +} +\subfigure[]{\label{fig:450:b} +\includegraphics[width=\columnwidth]{sic_prec_450_si-c.ps} +} +\end{center} +\caption{Radial distribution function for C-C and Si-Si (Fig.~\ref{fig:450:a}) as well as Si-C (Fig.~\ref{fig:450:b}) pairs for C inserted at \unit[450]{$^{\circ}$C}. In the latter case the resulting C-Si distances for a C$_{\text{i}}$ \hkl<1 0 0> DB are given additionally and the Si-C cut-off distance is marked by an arrow. Insets in Fig.~\ref{fig:450:a} show magnified regions of the respective distribution functions.} +\label{fig:450} +\end{figure} +There is no significant difference between C insertion into $V_2$ and $V_3$. +Thus, in the following, the focus is on low ($V_1$) and high ($V_2$, $V_3$) C concentration simulations only. + +In the low C concentration simulation the number of C-C bonds is small, as can be seen in the upper part of Fig.~\ref{fig:450:a}. +On average, there are only 0.2 C atoms per Si unit cell. +By comparing the Si-C peaks of the low concentration simulation with the resulting Si-C distances of a C$_{\text{i}}$ \hkl<1 0 0> DB in Fig.~\ref{fig:450:b} it becomes evident that the structure is clearly dominated by this kind of defect. +One exceptional peak at \unit[0.26]{nm} (marked with an arrow in Fig.~\ref{fig:450:b}) exists, which is due to the Si-C cut-off, at which the interaction is pushed to zero. +Investigating the C-C peak at \unit[0.31]{nm}, which is also available for low C concentrations as can be seen in the upper inset of Fig.~\ref{fig:450:a}, reveals a structure of two concatenated, differently oriented C$_{\text{i}}$ \hkl<1 0 0> DBs to be responsible for this distance. +Additionally, in the inset of the bottom part of Fig.\ref{fig:450:a} the Si-Si radial distribution shows non-zero values at distances around \unit[0.3]{nm}, which, again, is due to the DB structure stretching two neighbored Si atoms. +This is accompanied by a reduction of the number of bonds at regular Si distances of c-Si. +A more detailed description of the resulting C-Si distances in the C$_{\text{i}}$ \hkl<1 0 0> DB configuration and the influence of the defect on the structure is available in a previous study\cite{zirkelbach09}. + +For high C concentrations, the defect concentration is likewise increased and a considerable amount of damage is introduced in the insertion volume. +A subsequent superposition of defects generates new displacement arrangements for the C-C as well as Si-C pair distances, which become hard to categorize and trace and obviously lead to a broader distribution. +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 hardly visible is the long range order. +This indicates the formation of an amorphous SiC-like phase. +In fact, 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}. + +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. +On closer inspection two reasons for describing this obstacle become evident. + +First of all, there is the time scale problem inherent to MD in general. +To minimize the integration error the discretized time step must be chosen smaller than the reciprocal of the fastest vibrational mode resulting in a time step of \unit[1]{fs} for the investigated materials system. +Limitations in computer power result in a slow propagation in phase space. +Several local minima exist, which are separated by large energy barriers. +Due to the low probability of escaping such a local minimum, a single transition event corresponds to a multiple of vibrational periods. +Long-term evolution, such as a phase transformation and defect diffusion, in turn, are made up of a multiple of these infrequent transition events. +Thus, time scales to observe long-term evolution are not accessible by traditional MD. +New accelerated methods have been developed to bypass the time scale problem retaining proper thermodynamic sampling\cite{voter97,voter97_2,voter98,sorensen2000,wu99}. + +However, the applied potential comes up with an additional limitation, as previously mentioned in the introduction. +%The cut-off function of the short range potential limits the interaction to nearest neighbors, which results in overestimated and unphysical high forces between neighbored atoms. +The cut-off function of the short range potential limits the interaction to nearest neighbors. +Since the total binding energy is, thus, accommodated within this short distance, which according to the universal energy relation would usually correspond to a much larger distance, unphysical high forces between two neighbored atoms arise. +While cohesive and formational energies are often well described, these effects increase for non-equilibrium structures and dynamics. +This behavior, as observed and discussed for the Tersoff potential\cite{tang95,mattoni2007}, is supported by the overestimated activation energies necessary for C diffusion as investigated in section \ref{subsection:cmob}. +Indeed, it is not only the strong, hard to break C-C bond inhibiting C diffusion and further rearrangements in the case of the high C concentration simulations. +This is also true for the low concentration simulations dominated by the occurrence of C$_{\text{i}}$ \hkl<1 0 0> DBs spread over the whole simulation volume, which are unable to agglomerate due to the high migration barrier. + +\subsection{Increased temperature simulations} + +Due to the problem of slow phase space propagation, which is enhanced by the employed potential, pushing the time scale to the limits of computational resources or applying one of the above mentioned accelerated dynamics methods exclusively might not be sufficient. +Instead, higher temperatures are utilized to compensate overestimated diffusion barriers. +These are overestimated by a factor of 2.4 to 3.5. +Scaling the absolute temperatures accordingly results in maximum temperatures of \unit[1460-2260]{$^{\circ}$C}. +Since melting already occurs shortly below the melting point of the potential (2450 K)\cite{albe_sic_pot} due to the presence of defects, a maximum temperature of \unit[2050]{$^{\circ}$C} is used. + +Fig.~\ref{fig:tot} shows the resulting radial distribution functions for various temperatures. +\begin{figure} +\begin{center} +\subfigure[]{\label{fig:tot:si-c} +\includegraphics[width=\columnwidth]{tot_pc_thesis.ps} +} +\subfigure[]{\label{fig:tot:si-si} +\includegraphics[width=\columnwidth]{tot_pc3_thesis.ps} +} +\subfigure[]{\label{fig:tot:c-c} +\includegraphics[width=\columnwidth]{tot_pc2_thesis.ps} +} +\end{center} +\caption{Radial distribution function for Si-C (Fig.~\ref{fig:tot:si-c}), Si-Si (Fig.~\ref{fig:tot:si-si}) and C-C (Fig.~\ref{fig:tot:c-c}) pairs for the C insertion into $V_1$ at elevated temperatures. For the Si-C distribution resulting Si-C distances of a C$_{\text{s}}$ configuration are plotted. In the C-C distribution dashed arrows mark C-C distances occurring from C$_{\text{i}}$ \hkl<1 0 0> DB combinations, solid arrows mark C-C distances of pure C$_{\text{s}}$ combinations and the dashed line marks C-C distances of a C$_{\text{i}}$ and C$_{\text{s}}$ combination.} +\label{fig:tot} +\end{figure} +In Fig.~\ref{fig:tot:si-c}, the first noticeable and promising change observed for the Si-C bonds is the successive decline of the artificial peak at the cut-off distance with increasing temperature. +Obviously, sufficient kinetic energy is provided to affected atoms that are enabled to escape the cut-off region. +Additionally, a more important structural change was observed, which is illustrated in the two shaded areas in Fig.~\ref{fig:tot:si-c}. +Obviously, the structure obtained at \unit[450]{$^{\circ}$C}, which was found to be dominated by C$_{\text{i}}$, transforms into a C$_{\text{s}}$ dominated structure with increasing temperature. +Comparing the radial distribution at \unit[2050]{$^{\circ}$C} to the resulting bonds of C$_{\text{s}}$ in c-Si excludes all possibility of doubt. + +The phase transformation is accompanied by an arising Si-Si peak at \unit[0.325]{nm} in Fig.~\ref{fig:tot:si-si}, which 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 \unit[0.308]{nm} the existing SiC structures embedded in the c-Si host are stretched. + +According to the C-C radial distribution displayed in Fig.~\ref{fig:tot:c-c}, agglomeration of C fails to appear even for elevated temperatures, as can be seen on the total amount of C pairs within the investigated separation range, which does not change significantly. +However, a small decrease in the amount of neighbored C pairs can be observed with increasing temperature. +This high temperature behavior is promising since breaking of these diamond- 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 \unit[0.3]{nm}. +Arrows with dashed lines mark C-C distances resulting from C$_{\text{i}}$ \hkl<1 0 0> DB combinations while arrows with solid lines mark distances arising from combinations of C$_{\text{s}}$. +The continuous dashed line corresponds to the distance of C$_{\text{s}}$ and a neighbored C$_{\text{i}}$ 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. +Quite high g(r) values are obtained for distances 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 C$_{\text{s}}$ with either another C atom that basically occupies a Si lattice site but is displaced by 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. + +Fig.~\ref{fig:v2} displays the radial distribution for high C concentrations. +\begin{figure} +\begin{center} +\subfigure[]{\label{fig:v2:si-c} +\includegraphics[width=\columnwidth]{12_pc_thesis.ps} +} +\subfigure[]{\label{fig:v2:c-c} +\includegraphics[width=\columnwidth]{12_pc_c_thesis.ps} +} +\end{center} +\caption{Radial distribution function for Si-C (Fig.~\ref{fig:v2:si-c}) and C-C (Fig.~\ref{fig:v2:c-c}) pairs for the C insertion into $V_2$ at elevated temperatures. Arrows mark the respective cut-off distances.} +\label{fig:v2} +\end{figure} +\begin{figure} +\begin{center} +\includegraphics[width=\columnwidth]{2050.eps} +\end{center} +\caption{Cross section along the \hkl(1 -1 0) plane of the atomic structure of the high concentration simulation for a C insertion temperature of \unit[2050]{$^{\circ}$C}.} +\label{fig:v2as} +\end{figure} +A cross-section along the \hkl(1 -1 0) plane of the atomic structure for a C insertion temperature of \unit[2050]{$^{\circ}$C} is shown in Fig.~\ref{fig:v2as}. +The amorphous SiC-like phase remains. +No significant change in structure is observed. +However, the decrease of the cut-off artifact and slightly sharper peaks observed with increasing temperature, in turn, indicate a slight acceleration of the dynamics realized by the supply of kinetic energy. +However, it is not sufficient to enable the amorphous to crystalline transition. +In contrast, even though bonds of neighbored C atoms could be partially dissolved in the system exhibiting low C concentrations, the amount of neighbored C pairs even increased in the latter case. +Moreover, the C-C peak at \unit[0.252]{nm} in Fig.~\ref{fig:v2:c-c}, which gets slightly more distinct, equals the second nearest neighbor distance in diamond and indeed is made up by a structure of two C atoms interconnected by a third C atom. +Obviously, processes that appear to be non-conducive are likewise accelerated in a system, in which high amounts of C are incorporated within a short period of time, which is accompanied by a concurrent introduction of accumulating, for the reason of time non-degradable damage. +% non-degradable, non-regenerative, non-recoverable +Thus, for these systems even larger time scales, which are not accessible within traditional MD, must be assumed for an amorphous to crystalline transition or structural evolution in general. +% maybe put description of bonds in here ... +Nevertheless, some results likewise indicate the acceleration of other processes that, again, involve C$_{\text{s}}$. +The increasingly pronounced Si-C peak at \unit[0.35]{nm} in Fig.~\ref{fig:v2:si-c} corresponds to the distance of a C and a Si atom interconnected by another Si atom. +Additionally, the C-C peak at \unit[0.31]{nm} in Fig.~\ref{fig:v2:c-c} corresponds to the distance of two C atoms bound to a central Si atom. +For both structures the C atom appears to reside on a substitutional rather than an interstitial lattice site. +However, huge amounts of damage hamper identification. +The alignment of the investigated structures to the c-Si host is lost in many cases, which suggests the necessity of much more time for structural evolution to maintain the topotactic orientation of the precipitate. + +\section{Discussion and Summary} + +Investigations are targeted at the initially stated controversy of SiC precipitation, i.e. whether precipitation occurs abruptly after enough C$_{\text{i}}$ agglomerated or after a successive agglomeration of C$_{\text{s}}$ on usual Si lattice sites (and Si$_{\text{i}}$) followed by a contraction into incoherent SiC. +Results of a previous ab initio study on defects and defect combinations in C implanted Si\cite{zirkelbach11a} suggest C$_{\text{s}}$ to play a decisive role in the precipitation of SiC in Si. +To support previous assumptions MD simulations, which are capable of modeling the necessary amount of atoms, i.e. the precipitate and the surrounding c-Si structure, have been employed in the current study. + +In a previous comparative study\cite{zirkelbach10} we have shown that the utilized empirical potential fails to describe some selected processes. +Thus, limitations of the employed potential have been further investigated and taken into account in the present study. +We focussed on two major shortcomings: the overestimated activation energy and the improper description of intrinsic and C point defects in Si. +Overestimated forces between nearest neighbor atoms that are expected for short range potentials\cite{mattoni2007} have been confirmed to influence the C$_{\text{i}}$ diffusion. +The migration barrier was estimated to be larger by a factor of 2.4 to 3.5 compared to highly accurate quantum-mechanical calculations\cite{zirkelbach10}. +Concerning point defects, the drastically underestimated formation energy of C$_{\text{s}}$ and deficiency in the description of the Si$_{\text{i}}$ ground state necessitated further investigations on structures that are considered important for the problem under study. +It turned out that the EA potential still favors a C$_{\text{i}}$ \hkl<1 0 0> DB over a C$_{\text{s}}$-Si$_{\text{i}}$ configuration, which, thus, does not constitute any limitation for the simulations aiming to resolve the present controversy of the proposed SiC precipitation models. + +MD simulations at temperatures used in IBS resulted in structures that were dominated by the C$_{\text{i}}$ \hkl<1 0 0> DB and its combinations if C is inserted into the total volume. +Incorporation into volumes $V_2$ and $V_3$ led to an amorphous SiC-like structure within the respective volume. +To compensate overestimated diffusion barriers, we performed simulations at accordingly increased temperatures. +No significant change was observed for high C concentrations. +The amorphous phase is maintained. +Due to the incorporation of a huge amount of C into a small volume within a short period of time, damage is produced, which obviously decelerates structural evolution. +For the low C concentrations, time scales are still too low to observe C agglomeration sufficient for SiC precipitation, which is attributed to the slow phase space propagation inherent to MD in general. +However, we observed a phase transition of the C$_{\text{i}}$-dominated into a clearly C$_{\text{s}}$-dominated structure. +The amount of substitutionally occupied C atoms increases with increasing temperature. +Entropic contributions are assumed to be responsible for these structures at elevated temperatures that deviate from the ground state at 0 K. +Indeed, in a previous ab initio MD simulation\cite{zirkelbach11a} performed at \unit[900]{$^{\circ}$C} we observed the departing of a Si$_{\text{i}}$ \hkl<1 1 0> DB located next to a C$_{\text{s}}$ atom instead of a recombination into the ground state configuration, i.e. a C$_{\text{i}}$ \hkl<1 0 0> DB. + +% postannealing less efficient than hot implantation +Experimental studies revealed increased implantation temperatures to be more efficient than postannealing methods for the formation of topotactically aligned precipitates\cite{eichhorn02}. +In particular, restructuring of strong C-C bonds is affected\cite{deguchi92}, which preferentially arise if additional kinetic energy provided by an increase of the implantation temperature is missing to accelerate or even enable atomic rearrangements. +We assume this to be related to the problem of slow structural evolution encountered in the high C concentration simulations due to the insertion of high amounts of C into a small volume within a short period of time resulting in essentially no time for the system to rearrange. +% rt implantation + annealing +Implantations of an understoichiometric dose at room temperature followed by thermal annealing results in small spherical sized C$_{\text{i}}$ agglomerates at temperatures below \unit[700]{$^{\circ}$C} and SiC precipitates of the same size at temperatures above \unit[700]{$^{\circ}$C}\cite{werner96}. +Since, however, the implantation temperature is considered more efficient than the postannealing temperature, SiC precipitates are expected -- and indeed are observed for as-implanted samples\cite{lindner99,lindner01} -- in implantations performed at \unit[450]{$^{\circ}$C}. +Implanted C is therefore expected to occupy substitutionally usual Si lattice sites right from the start. + +Thus, we propose an increased participation of C$_{\text{s}}$ already in the initial stages of the implantation process at temperatures above \unit[450]{$^{\circ}$C}, the temperature most applicable for the formation of SiC layers of high crystalline quality and topotactical alignment\cite{lindner99}. +Thermally activated, C$_{\text{i}}$ is enabled to turn into C$_{\text{s}}$ accompanied by Si$_{\text{i}}$. +The associated emission of Si$_{\text{i}}$ is needed for several reasons. +For the agglomeration and rearrangement of C, Si$_{\text{i}}$ is needed to turn C$_{\text{s}}$ into highly mobile C$_{\text{i}}$ again. +Since the conversion of a coherent SiC structure, i.e. C$_{\text{s}}$ occupying the Si lattice sites of one of the two fcc lattices that build up the c-Si diamond lattice, into incoherent SiC is accompanied by a reduction in volume, large amounts of strain are assumed to reside in the coherent as well as at the surface of the incoherent structure. +Si$_{\text{i}}$ serves either as a supply of Si atoms needed in the surrounding of the contracted precipitates or as an interstitial defect minimizing the emerging strain energy of a coherent precipitate. +The latter has been directly identified in the present simulation study, i.e. structures of two C$_{\text{s}}$ atoms and Si$_{\text{i}}$ located in the vicinity. + +It is, thus, concluded that precipitation occurs by successive agglomeration of C$_{\text{s}}$ as already proposed by Nejim et~al.\cite{nejim95}. +This agrees well with a previous ab initio study on defects in C implanted Si\cite{zirkelbach11a}, which showed C$_{\text{s}}$ to occur in all probability. +However, agglomeration and rearrangement is enabled by mobile C$_{\text{i}}$, which has to be present at the same time and is formed by recombination of C$_{\text{s}}$ and Si$_{\text{i}}$. +In contrast to assumptions of an abrupt precipitation of an agglomerate of C$_{\text{i}}$\cite{werner96,werner97,eichhorn99,lindner99_2,koegler03}, however, structural evolution is believed to occur by a successive occupation of usual Si lattice sites with substitutional C. +This mechanism satisfies the experimentally observed alignment of the \hkl(h k l) planes of the precipitate and the substrate, whereas there is no obvious reason for the topotactic orientation of an agglomerate consisting exclusively of C-Si dimers, which would necessarily involve a much more profound change in structure for the transition into SiC. + +% ---------------------------------------------------- +\section*{Acknowledgment} +We gratefully acknowledge financial support by the Bayerische Forschungsstiftung (DPA-61/05) and the Deutsche Forschungsgemeinschaft (DFG SCHM 1361/11). +%Meta Schnell is greatly acknowledged for a critical revision of the present manuscript. + +% --------------------------------- references ------------------- + +\bibliography{../../bibdb/bibdb}{} +\bibliographystyle{h-physrev3} + +\end{document}