1: % as.tex
2: %\documentclass {revtex4}
3: \documentclass[prb,twocolumn,showpacs,showkeys,groupedaddress,amsmath,amssymb]{revtex4}% Physical Review B
4: %
5: %\usepackage{Bm}
6: \usepackage {amssymb}
7: \usepackage {amsmath}
8: %\usepackage[dvips]{epsfig}
9: \usepackage{graphics,amssymb}
10: \usepackage{natbib}
11: \usepackage{revsymb}
12:
13: % Some other (several out of many) possibilities
14: %\documentclass[preprint,aps]{revtex4}
15: %\documentclass[preprint,aps,draft]{revtex4}
16:
17: \usepackage{graphicx}% Include figure files
18: \usepackage{dcolumn}% Align table columns on decimal point
19: \usepackage{bm}% bold math
20: %
21: \newcommand{\DS}{\displaystyle}
22:
23: \begin{document}
24:
25:
26: \title{Mechanisms of arsenic clustering in silicon\\}
27: % Force line breaks with \\
28:
29:
30: \author{F. F. Komarov}
31: \affiliation{Department of Physical Electronics, Belarusian State
32: University, 1 Kurchatov
33: Str., Minsk 220064, Belarus \\}%
34: %e-mail: KomarovF@bsu.by
35:
36: \author{O. I. Velichko} %\email[]{oleg_velichko@lycos.com}
37: \affiliation{Department of Physics, Belarusian State University on
38: Informatics and Radioelectronics, 6 P. Brovka Str., Minsk 220013,
39: Belarus
40: \\}
41:
42:
43: \author{V. A. Dobrushkin} \email[]{dobrush@dam.brown.edu}
44: \affiliation{Brown University, Division of Applied Mathematics,
45: Providence, Rhode Island 02912, USA
46: \\}
47:
48: \author{A. M. Mironov}
49: \affiliation{Institute of Applied Physics Problems, 7 Kurchatov
50: Str., Minsk 220064, Belarus
51: \\}
52: % e-mail: MironovA@bsu.by
53:
54: \begin{abstract}
55:
56: A model of arsenic clustering in silicon is proposed and analyzed.
57: The main feature of the proposed model is the assumption that
58: negatively charged arsenic complexes play a dominant role in the
59: clustering process. To confirm this assumption, electron density
60: and concentration of impurity atoms incorporated into the clusters
61: are calculated as functions of the total arsenic concentration at
62: a temperature of 1050\,$^{\circ}$C. A number of the negatively
63: charged clusters incorporating a point defect and one or more
64: arsenic atoms $(\text{DAs}_{1})^{-}$, $(\text{DAs}_{1})^{2-}$,
65: $(\text{DAs}_{2})^{-}$, $(\text{DAs}_{2})^{2-}$,
66: $(\text{DAs}_{3})^{-}$, $(\text{DAs}_{3})^{2-}$,
67: $(\text{DAs}_{4})^{-}$, and $(\text{DAs}_{4})^{2-}$ are
68: investigated. It is shown that for the doubly negatively charged
69: clusters or for clusters incorporating more than one arsenic atom the
70: electron density reaches a maximum value and then monotonically
71: and slowly decreases as total arsenic
72: concentration increases. In the case of cluster $(\text{DAs}_{2})^{2-}$, the
73: calculated electron density agrees well with the experimental
74: data. Agreement with the experiment confirms the conclusion that
75: two arsenic atoms participate in the cluster formation. Among all
76: present models, the proposed model of clustering by formation of
77: $(\text{DAs}_{2})^{2-}$ gives the best fit to the experimental
78: data and can be used in simulation of high concentration arsenic
79: diffusion.
80: \end{abstract}
81:
82: \pacs{61.72.Ji, 61.72.Tt, 61.72.Ss, 85.40.Ry}% PACS
83: %Doping and impurity implantation in germanium and silicon % Classification Scheme.
84:
85: \keywords{clusters; annealing; doping effects; arsenic; silicon}
86:
87: %Use showkeys class option if keyword
88: %display desired
89:
90: \maketitle
91:
92: \section{Introduction}
93:
94: Using low energy high dose arsenic ion implantation, one can
95: produce the active regions of modern integrated microcircuits
96: characterized by very shallow junctions and high dopant
97: concentrations. Thermal annealing is applied after implantation
98: for the arsenic activation and damage reduction. During the
99: initial stage of annealing, the arsenic atoms become electrically
100: active, occupying the substitutional positions. Then a fraction of
101: impurity atoms incorporates into the clusters, thereby decreasing
102: the layer conductivity.
103: \cite{Tsukamoto_80,Tsai_80,Fahey_89,Nobili_94,
104: Solmi_98,Krishnamoorthy_98,Nobili_99,Uematsu_00,
105: Nobili_01,Solmi_01,Solmi_03} Due to clustering, the concentration
106: of charge carriers is less than the total impurity concentration
107: because the clustered dopant atoms do not serve as a source of
108: free charge carriers. We can observe the clustering phenomenon in
109: many cases of the formation of highly doped semiconductor
110: structures; for example, in the region with high impurity
111: concentration during thermal diffusion of arsenic
112: atoms.\cite{Fair_73,Fair_Weber_73,Murota_79} Deactivation of the
113: electrically active dopant atoms also indicates that clustering
114: can occur during thermal treatment of the supersaturated arsenic
115: layers created by ion implantation with subsequent thermal or
116: laser annealing.
117: \cite{Nobili_94,Nobili_99,Parisini_90,Luning_92,Rousseau_94,
118: Rousseau_98,Solmi_02} Moreover, a reverse annealing or transient
119: increase of the carrier density can be observed if a doped
120: specimen, previously annealed at temperature $T_{1}$, is further
121: annealed at higher temperature $T_{2}$. \cite{Solmi_00} It is
122: supposed that transient dissolution of a part of the clusters
123: occurs during such a reverse annealing. \cite{Solmi_00}
124:
125:
126: At very high dopant concentrations exceeding the solid solubility
127: of As in Si, a significant fraction of arsenic atoms can form
128: precipitates.
129: \cite{Fahey_89,Nobili_94,Solmi_98,Krishnamoorthy_98,Uematsu_00,Solmi_01,
130: Parisini_90, Nobili_83,La Via_91} The typical distributions of the
131: total dopant concentration and electron density after thermal
132: annealing of silicon substrates heavily implanted by As were
133: presented by Nobili et al. \cite{Nobili_94} (Fig. 7) and by Solmi
134: \cite{Solmi_01} (Fig. 1c). As can be seen from Fig. 1c,
135: \cite{Solmi_01} the carrier concentration reaches its saturation
136: value $n_{e}=3.57\times10^{8}$ $\mu $m$^{{\rm - }{\rm 3}}$,
137: whereas the total arsenic concentration increases to the
138: solubility limit $C_{sol}=3.27\times10^{9}$ $\mu $m$^{{\rm - }{\rm
139: 3}}$ \cite{Nobili_01} and continues to increase in the region
140: adjoint to the surface. It is supposed that the cluster formation
141: occurs at high dopant concentrations,
142: \cite{Tsukamoto_80,Tsai_80,Fahey_89,Nobili_99,Nobili_01,Solmi_01,Fair_73,
143: Parisini_90,Luning_92} more precisely, at the dopant
144: concentrations approximately ranging from $n_{e}$ to $10\,n_{e}$.
145: If dopant concentration exceeds the solubility limit $C_{sol}$,
146: precipitation occurs. \cite{Nobili_94,Solmi_01} Arsenic clustering
147: and precipitation have attracted evergrowing attention of the
148: researchers, as they are greatly important for the silicon
149: integrated circuits technology. Both clusters and precipitates do
150: not serve as a source of charge carriers and the total arsenic
151: concentration may be higher than the electron density by about two
152: orders of magnitude. The clusters and precipitates are metastable
153: and can dissociate under subsequent thermal treatments. This means
154: that stringent requirements should be placed upon the accuracy of
155: the clustering models used to calculate the carrier concentrations
156: in heavily doped silicon layers.
157:
158: \section{Analysis of clustering phenomena}
159:
160: The development of a model for As clustering necessitates a
161: consistent analysis of the experimental data and theoretical
162: substantiation of the processes proceeding in heavily doped
163: silicon layers.
164:
165: \subsection{Clustering during thermal diffusion
166: and annealing of ion-implanted layers}
167:
168:
169: A great number of clustering models with different types of
170: clusters have been developed and proposed since 1973.
171: \cite{Tsukamoto_80,Tsai_80,Fahey_89,Fair_73,Parisini_90,Luning_92,
172: Guerrero_82,Berding_98,Mueller_03} The variety of the proposed
173: models reflects discrepancies in the experimental results
174: \cite{Nobili_83} and a great uncertainty in the nature of the
175: clusters discussed. For example, inactive cluster $\text{VAs}_{2}$
176: was considered in \cite{Fair_73} to explain the difference between
177: the total and electrically active dopant concentrations for
178: arsenic thermally diffused from a constant source. For the best
179: fit to the diffusion profiles of ion-implanted arsenic, it was
180: suggested that the electrically inactive arsenic atoms were
181: incorporated into the clusters with four arsenic atoms per
182: cluster. \cite{Tsukamoto_80} According to Tsai et al.
183: \cite{Tsai_80} who have also investigated diffusion in the
184: ion-implanted layers, clustering of arsenic may be expressed as
185:
186: \begin{equation} \label{Cluster_Reaction_Tsai}
187: 3\text{As}^{+}+e^{-}{\stackrel{\mbox{Annealing}}{\qquad
188: \longleftrightarrow \qquad }}
189: \text{As}_{3}^{2+}{\stackrel{\mbox{25\,$^{\circ}$C}}{\quad
190: \longrightarrow \quad }} \text{As}_{3},
191: \end{equation}
192:
193:
194: \noindent where $\text{As}^{ +} $ is the substitutionally
195: dissolved arsenic atom participating in the cluster formation;
196: $e^{ - }$ is the electron.
197:
198: As follows from reaction (\ref{Cluster_Reaction_Tsai}), the
199: clustered As atoms are electrically active at annealing
200: temperatures, but are neutral at room temperature.
201:
202: It was found in \cite{Uematsu_00} that the formation of clusters
203: $(\text{As}_{2}\text{V})^{\times}$ occurs due to the following
204: reaction, which
205:
206: \begin{equation} \label{Cluster_Reaction_Uematsu}
207: 2\text{As}^{+}+\text{V}^{2-}\longleftrightarrow
208: (\text{As}_{2}\text{V})^{\times}
209: \end{equation}
210:
211: \noindent is most liable to cause deactivation because the
212: concentration of inactive As species increases to the fourth power
213: of the active As concentration. \cite{Tsukamoto_80,Murota_79} Here
214: $\text{V}^{2-}$ is the doubly negatively charged vacancy
215: participating in the clustering.
216:
217: Indeed, if a local equilibrium is assumed, it follows from the mass action
218: law for reaction (\ref{Cluster_Reaction_Uematsu}) that
219:
220: \begin{equation} \label{MAL_Uematsu}
221: C^{A}=A\tilde{C}^{V\times}\chi^{2}C^{2},
222: \end{equation}
223:
224: \begin{equation} \label{Constant_Uematsu}
225: A=H^{A} C_{i}^{V\times} n_{i}^{2},
226: \end{equation}
227:
228: \begin{equation} \label{Rel_vacancy_concentration}
229: \tilde{C}^{V\times}=\frac{C^{V\times}}{C_{i}^{V\times}},
230: \end{equation}
231:
232: \noindent where $C^{A}$ and $C$ are the concentrations of the
233: impurity atoms incorporated into clusters and substitutionally
234: dissolved arsenic atoms, respectively; $H^{A}$ is the constant of
235: the local equilibrium for reaction
236: (\ref{Cluster_Reaction_Uematsu}); $C^{V\times}$ and
237: $C_{i}^{V\times}$ are the actual and the equilibrium
238: concentrations of neutral vacancies, respectively;
239: $\chi=\displaystyle \frac{n}{n_{i}}$ is the electron density
240: normalized to the concentration of intrinsic charge carriers in a
241: semiconductor during thermal treatment $n_{i}$.
242:
243: One would expect from (\ref{MAL_Uematsu}) that reaction
244: (\ref{Cluster_Reaction_Uematsu}) yields the fourth power
245: dependence $C^{A}\propto C^{4}$, since for high concentrations
246: $C>>n_{i}$ it might be reasonable to assume $\displaystyle
247: \chi=\frac{n}{n_{i}}\thickapprox \frac{C}{n_{i}}$. However, as
248: seen from Fig. 7 (Ref. \cite{Nobili_94}) and Fig. 1c (Ref.
249: \cite{Solmi_01}), the electron density $n\thickapprox n_{e}=const$
250: in the region of concentrations associated with active clustering.
251: Consequently, reaction (\ref{Cluster_Reaction_Uematsu}) yields the
252: second power dependence $C^{A}\backsim C^{2}$, which contradicts
253: the experimental data. \cite{Tsukamoto_80,Murota_79}
254:
255:
256: Because of the vacancy participating in cluster formation, silicon
257: self-interstitials were in turn ejected or left behind during
258: clustering. \cite{Krishnamoorthy_98} This assumption had been made
259: much earlier in the study \cite{Parisini_90,Rousseau_94} of
260: impurity deactivation within the laser-annealed layers. However,
261: the results obtained in \cite{Krishnamoorthy_98} show that the
262: point defects induced during clustering and/or precipitation make
263: no contribution to the enhanced transient diffusion of arsenic
264: implanted at low energy, while these defects did coalesce to form
265: extended defects at the projected range. Analysis of the defect
266: microstructure has revealed that the transition between arsenic
267: clustering and SiAs precipitation is not abrupt, pointing to
268: possible coexistence of arsenic clusters and SiAs precipitates
269: (although precipitates are not directly observed).
270: \cite{Krishnamoorthy_98}
271:
272:
273: The model \cite{Tsai_80} was generalized by Solmi and Nobili
274: \cite{Solmi_98}
275:
276: \begin{equation} \label{Cluster_Reaction_Solmi}
277: m\text{As}^{+}+e^{-}{\stackrel{\mbox{Annealing}}{\qquad
278: \longleftrightarrow \qquad }}
279: \text{As}_{m}^{(m-1)+}{\stackrel{\mbox{25\,$^{\circ}$C}}{\quad
280: \longrightarrow \quad }} \text{As}_{m}
281: \end{equation}
282: \noindent to take into account a saturation behavior of the
283: carrier density with increasing dopant concentration. For
284: diffusion at a temperature of 1050\,$^{\circ}$C, the best agreement
285: between the modeled and experimental curves describing diffusivity
286: against total arsenic concentration was achieved for $m=4$. And at
287: 900\,$^{\circ}$C, the neutral cluster model \cite{Fair_73} provided
288: a better agreement with the experimental data. The reaction
289: (\ref{Cluster_Reaction_Solmi}) was also used in \cite{Nobili_01}
290: to describe deactivation in the arsenic-doped layers, which were
291: formed by ion implantation and high temperature annealing of
292: silicon-on-insulator specimens. However, Solmi et al.
293: \cite{Solmi_03} have recently used the following reaction
294:
295: \begin{equation} \label{Cluster_Reaction_Solmi_03}
296: m\text{As}^{+}+\text{V}\longleftrightarrow \text{As}_{m}+\text{I}
297: \end{equation}
298:
299: \noindent to model the transient arsenic diffusion. Here
300: $m$ is assumed to have the values between 2 and 4 to take into
301: account the fact that the As cluster was formed around a vacancy with
302: the subsequent injection of the self-interstitial.
303:
304:
305: It is important to note that reaction
306: (\ref{Cluster_Reaction_Solmi}) was first theoretically studied by
307: Guerrero et al. in paper, \cite{Guerrero_82} which will be
308: considered below.
309:
310: \subsection{Clustering during deactivation of supersaturated
311: arsenic layers formed by laser annealing}
312:
313: Let us consider the clustering models used for the explanation of
314: deactivation in the supersaturated arsenic layers formed by ion
315: implantation and subsequent laser annealing. In \cite{Luning_92}
316: it was found that a cluster consisting of four positively charged
317: arsenic atoms and two additional negative charges gave the best
318: least-squares fit with respect to the experimental data obtained
319: in the case of deactivation within the supersaturated arsenic
320: layers created by ion implantation and subsequent laser annealing.
321: If two negative charges are associated with a doubly charged
322: vacancy, this model gains support from the theoretical
323: calculations of Pandey et al. \cite{Pandey_88} Such a cluster has
324: an intuitive geometrical and structural appeal --- four arsenic
325: atoms tetrahedrally arranged around a vacancy may relax inward,
326: thus relieving lattice strains due to the As size effect.
327: \cite{Luning_92} Using this cluster species, a single-step
328: clustering reaction was postulated in \cite{Luning_92}
329:
330: \begin{equation} \label{Cluster_Reaction_Luning}
331: 4\text{As}^{+}+\text{V}^{2-}\longleftrightarrow
332: (\text{As}_{4}\text{V})^{2+}.
333: \end{equation}
334:
335:
336: The assumption of self-interstitial ejection during clustering was
337: supported by the experimental data obtained by Parisini et al.
338: \cite{Parisini_90} who have observed that extended defects at the
339: projected range were extrinsic or interstitial in nature. The reaction
340:
341: \begin{equation} \label{Cluster_Reaction_Parisini}
342: (\text{As}_{2})^{2+}+2e^{-}\longrightarrow
343: (\text{As}_{2}\text{V})^{+}+\text{I}^{-}
344: \end{equation}
345:
346: \noindent was proposed in \cite{Parisini_90} only for the initial
347: step of the deactivation process in arsenic implanted silicon
348: specimens, first laser-annealed and then thermally annealed. Here
349: $\text{I}^{-}$ is the negatively charged self-interstitial. The
350: results obtained by Rousseau et al. \cite{Rousseau_94} confirm the
351: conclusion that arsenic is indeed deactivated by vacancies, with a
352: concurrent injection of self-interstitials. The latter follows
353: from the observed significant enhancement of the diffusion in the
354: buried boron layer underneath the As structure subjected to
355: deactivation. Investigations of the enhanced diffusion in buried
356: boron layers were continued in, \cite{Rousseau_98} where it was
357: proposed that small arsenic clusters of various sizes were formed
358: around a vacancy during deactivation with injection of the
359: associated interstitial into the bulk.
360:
361: Considering that the majority of the models described in this
362: section are based essentially on the experimental investigations
363: of the defect subsystem, we first concern ourselves with these
364: experimental results.
365:
366:
367: \subsection{Experimental investigations of defect subsystem}
368:
369: The methods of transmission electron spectroscopy (TEM),
370: \cite{Nobili_94,Krishnamoorthy_98,Parisini_90,La
371: Via_91,Dokumachi_95} extended x-ray-absorption fine-structure
372: (EXAFS), \cite{Parisini_90,Erbil_86,Allain_92,d'Acapito_04}
373: Rutherford backscattering (RBS), \cite{La Via_91,Brizard_93}
374: positron lifetime measurements
375: \cite{Makinen_89,Lawther_95,Myler_96,Szpala_96} combined with the
376: measurements of electron momentum distributions,
377: \cite{Polity_98,Saarinen_99,Ranki_03,Ranki_Saarinen_03,Ranki_04}
378: and other methods \cite{Nobili_83,Herrera-Gomez_96} are commonly
379: used to investigate silicon specimens doped by As. For example,
380: TEM observations show that As-lased samples are completely free of
381: any visible defects. \cite{Dokumachi_95} On the other hand, TEM
382: analysis reveals the presence of extended defects after subsequent
383: thermal treatment. \cite{Parisini_90,Dokumachi_95} For example,
384: small \{113\} interstitial loops begin to appear beyond the
385: plateau region of the As profile after low-temperature treatment.
386: \cite{Parisini_90} After high-temperature thermal treatment
387: ($>$850\,$^{\circ}$C), a low density of very small As-related
388: precipitates (about 2 nm in diameter) is observed. The
389: precipitated As fraction cannot be responsible for the total
390: amount of inactive As. \cite{Parisini_90}
391:
392: It was established that EXAFS measurements provided particular
393: details about the local atomic structure around the dopant and
394: confirmed the fact that As atoms are substitutional after laser
395: annealing. \cite{Erbil_86,Allain_92} Note that, according to the
396: x-ray standing-wave spectroscopy measurements, the As atoms
397: remained at the substitutional positions even after 85\% of the
398: electrical activity has been lost due to thermal annealing of the
399: laser melted layers. \cite{Herrera-Gomez_96} On the other hand, a
400: large amount of nonsubstitutional arsenic was detected by
401: Rutherford back-scattering after diffusion from $\text{TiSi}_{2}$
402: source. \cite{La Via_91} These defects are interpreted as
403: precipitates, probably formed due to the stress induced by
404: $\text{TiSi}_{2}$ layer. Indeed, when thermal annealing is carried
405: out after the complete removal of titanium silicide (i.e., without
406: stress), one can observe the activation of all the arsenic atoms
407: diffused in the silicon substrate. \cite{La Via_91} Thus, such
408: full activation indicates the significant influence of stress on
409: the formation of precipitates.
410: %
411: % very important for future analysis
412: %
413:
414: According to EXAFS measurements, \cite{Allain_92} subsequent
415: thermal annealing (350\,\,--\,750\,$^\circ $C) of the As-lased
416: layers leads to the formation of $VAs_{m}$ clusters including up
417: to $7\pm4$ As atoms around the vacancy. However, at high
418: temperatures it was observed that the number of the first
419: neighbors of Si has tended back to 4 atoms.
420: %
421: % very important for this analysis
422: %
423: This is probably due to the precipitates observed by TEM over
424: this temperature range. Quantitative analysis of EXAFS data for
425: the ultra-low energy implanted layers after rapid or spike
426: annealing \cite{d'Acapito_04} has revealed a site for As that was
427: different from the pure substitutional one, suggesting the presence
428: of clusters of As atoms coupled to vacancies. All the observed
429: phenomena, namely, a low value of the total coordination number in
430: the annealed samples, small $\text{As--Si}$ bond length, and the
431: presence of $\text{As--As}$ coordinations, may be explained by
432: the assumption that clustering of As ions with vacancies takes
433: place to form $\text{VAs}_{m}$ structures.
434:
435: Ion channeling and RBS were used in \cite{Brizard_93} to
436: complement the EXAFS and electron microscopy results. In has been
437: found that the deactivated As atom is displaced with respect to
438: the lattice sites and, moreover, the average displacement is
439: constant over the temperature range 450\,\,--\,900\,$^{\circ}$C,
440: being equal to $0.23 \pm 0.06$ \AA. Also, in \cite{Brizard_93} it
441: was proposed that the formation of larger clusters
442: $\text{V}_{n}\text{As}_{m}$ ($n\thickapprox 5$ and $m\thickapprox
443: 10$) occurred starting from $\text{VAs}_{2}$ or $\text{VAs}_{4}$.
444: The formation of these larger clusters is in agreement with a
445: decrease in the first nearest neighbors of As. At higher
446: temperatures, an increase in the number of the first nearest
447: neighbors of As to 4 was observed by EXAFS measurements.
448: \cite{Allain_92} This means that the cluster species may differ at
449: intermediate (350\,\,--\,750\,$^{\circ}$C) and higher
450: temperatures.
451:
452: Very interesting information about clustering was obtained during
453: the experimental investigations of the doped layers by positron
454: annihilation spectroscopy
455: \cite{Lawther_95,Myler_96,Saarinen_99,Ranki_03,Ranki_Saarinen_03,Ranki_04}
456: and by positron annihilation spectroscopy combined with the
457: Hall-effect/resistivity measurements. \cite{Ranki_03} The defect
458: subsystem was generally investigated before and after thermal
459: treatment. To support these measurements, the positron lifetimes
460: and core electron momentum distributions were calculated for
461: different vacancy-donor complexes. \cite{Ranki_04} Lawther et al.
462: \cite{Lawther_95} have investigated the doped layers melted by an
463: excimer laser to obtain the profile with constant As concentration
464: and a sharp fall-off at a depth of 0.2 $\mu$m. Arsenic
465: deactivation was initiated by annealing at 750\,$^{\circ}$C for 15
466: s or by conventional thermal treatment for 2 h. It was found that
467: $\text{VAs}_{m}$ complexes with the average values of $m$ greater
468: than 2 caused arsenic deactivation in heavily doped Si. Myler et
469: al. \cite{Myler_96} have studied the silicon layers (with arsenic
470: concentration 4$\times 10^{8}$ $\mu \text{m}^{3}$) fully activated
471: by laser melting and subsequently annealed for 15 s at 500 and
472: 750\,$^{\circ}$C. The changes in the positron annihilation spectra
473: after the thermal treatment at 500 or 750\,$^{\circ}$C were also
474: attributed to the formation of
475: $\text{As}_{m}\text{Si}_{4-m}\text{V}_{ac}$ complexes. In this
476: experiment, the impossibility of determining the number of
477: impurity atoms incorporated in the impurity-vacancy complex was
478: established. Ranki at al. \cite{Ranki_03} have studied the samples
479: implanted by As during MBE growth at 450\,$^{\circ}$C. The samples
480: were annealed in $\text{N}_{2}$, either by RTA at 900\,$^{\circ}$C
481: for 10\,\,--\,170 s or furnace annealing at
482: 800\,\,--\,900\,$^{\circ}$C for two minutes. Based on the positron
483: annihilation and Hall-effect/resistivity experiments, Ranki et al.
484: \cite{Ranki_03} have concluded that the dominant defect in
485: As-grown samples was $\text{VAs}_{3}$. The measurements
486: demonstrated high concentrations (above $10^{5}$ $\mu
487: \text{m}^{-3}$) of $\text{VAs}_{3}$ complexes in heavily-doped
488: silicon. Moreover, larger V-As complexes, probably
489: $\text{V}_{2}\text{As}_{5}$, may occur together with
490: $\text{VAs}_{3}$ at high As concentrations. A relative amount of
491: $\text{V}_{2}\text{As}_{5}$ increases with annealing. This cluster
492: is even dominant after annealing at 800\,$^{\circ}$C. The
493: $\text{VAs}_{3}$ and $\text{V}_{2}\text{As}_{5}$ complexes become
494: unstable at 800\,$^{\circ}$C and 900\,$^{\circ}$C, respectively,
495: then their concentrations decrease. Cluster reconstruction may
496: occur during cooling , starting from $\text{VAs}_{1}$ to
497: $\text{VAs}_{2}$ and then to $\text{VAs}_{3}$. The clusters
498: $\text{VAs}_{2}$ and $\text{VAs}_{3}$ may be in turn transformed
499: to the $\text{V}_{2}\text{As}_{5}$ complexes. Moreover, some
500: $\text{VAs}_{4}$ and other vacancy complexes may be also formed,
501: but with much lower concentrations.
502:
503:
504: High energy electron irradiation is commonly used in positron
505: experiments for the generation of nonequilibrium vacancies, mobile
506: even at room temperature and liable to interact with the dopant
507: atoms. \cite{Saarinen_99,Ranki_02,Ranki_Saarinen_03,Ranki_04} For
508: example, positron experiments \cite{Saarinen_99} were performed
509: both with the As-grown samples doped by arsenic in concentrations
510: of $10^{7}$ and $10^{8}$ $\mu \text{m}^{-3}$ and samples subjected
511: to 2 MeV electron irradiation at 300 K. It was established that
512: heavily As-doped silicon contained $\text{VAs}_{3}$ complex as a
513: native defect. Before irradiation, the concentration of
514: $\text{VAs}_{3}$ was equal to $\sim 10^{7}$ $\mu \text{m}^{-3}$ at
515: a doping level of $10^{8}$ $\mu \text{m}^{-3}$. After electron
516: irradiation, one can observe the pairs ($\text{VAs}_{1}$) formed
517: by a vacancy and a single impurity atom. It was demonstrated in
518: \cite{Ranki_02} that the migration of these $\text{VAs}_{1}$
519: defects started at around 450 K, leading to formation of
520: $\text{VAs}_{2}$ defects. And, in turn, these defects were
521: transformed to $\text{VAs}_{3}$ defects at 700 K. The
522: $\text{VAs}_{3}$ defects were stable at 700 K, representing the
523: dominant vacancy-impurity cluster in heavily doped n-type Si at
524: this temperature. The formation of larger $\text{VAs}_{2}$ and
525: $\text{VAs}_{3}$ complexes was significantly dependent on the As
526: concentration. This interpretation of the experimental data was
527: confirmed by the results obtained in. \cite{Ranki_Saarinen_03} In
528: these experiments, the electron irradiated samples were annealed
529: isochronally (30 min) at 300\,\,--\,1220 K. Irradiation of the
530: heavily doped Si samples has produced mainly the vacancy-donor
531: pairs ($\text{VAs}_{1}$) with a small concentration of
532: divacancies. Considering that, after irradiation, $\text{VAs}_{1}$
533: concentration was much higher than the initial concentration of
534: $\text{VAs}_{3}$, no signals from $\text{VAs}_{3}$ were observed
535: for the As-irradiated As-doped sample. From the core-region
536: electron momentum distribution measurements, it was found
537: \cite{Ranki_Saarinen_03} that defects in As-irradiated samples may
538: be identified as $\text{VAs}_{1}$, in samples annealed at 600 K
539: --- as $\text{VAs}_{2}$, and in those annealed at 775 K --- as
540: $\text{VAs}_{3}$. It is well known that a drastic drop in the
541: conductivity occurs when heavily doped Si is annealed at
542: temperatures between 400\,$^{\circ}$C and 500\,$^{\circ}$C. This
543: deactivation of the dopants is partially reversible by annealing
544: at 800\,\,--\,1000\,$^{\circ}$C. Taking into account these
545: experimental data, Ranki et al. \cite{Ranki_Saarinen_03} have
546: suggested that the formation and annealing of $\text{VAs}_{3}$ at
547: 700 and 1100 K, respectively, are responsible for the observed
548: behavior of the conductivity. The experiments conducted in
549: \cite{Ranki_02} were completed by studies in \cite{Ranki_04}
550: associated with thermal treatment up to 1220 K. It has been found
551: that dissociation of $\text{VAs}_{3}$ began at 1100 K, and
552: at 1220 K these defects were annealed away.
553:
554:
555: Thus, the analyzed data show that complexing of As atoms with
556: vacancies occurs in the layers heavily doped by arsenic. Based on
557: the experimental data, various $\text{V}_{m}\text{As}_{n}$
558: clusters are possible but $\text{VAs}_{2}$ and $\text{VAs}_{3}$
559: are the most likely. Taking clustering into account, one can
560: explain the phenomenon of compensation at high doping levels.
561: Nevertheless, some aspects are not clearly understood. For
562: example, $\text{VAs}_{3}$ starts to dissociate at 1100 K. This
563: annealing temperature agrees well with the observation that the
564: number of Si first neighbors tends back to 4 atoms
565: \cite{Allain_92} at temperatures higher than 750\,$^{\circ}$C. At
566: the same time, clustering occurs at temperatures higher than
567: 750\,$^{\circ}$C as well. \cite{Solmi_01} This means that such a
568: high temperature clustering is hardly explained by
569: $\text{VAs}_{3}$ complexes.
570:
571: % grown by MBE and doped layers irradiated by electrons
572:
573: \subsection{First principles studies}
574:
575: Along with various experimental investigations, a number of the
576: theoretical {\it ab initio} calculations have been performed to
577: explain deactivation of arsenic atoms in silicon. From the
578: calculations of Pandey et al., \cite{Pandey_88} it follows that
579: $\text{VAs}_{4}$ complex including a vacancy surrounded by four
580: arsenic atoms is energetically favored over both substitutional,
581: isolated As in Si and substitutional $\text{Si\,-As}_{4}$
582: configurations. This cluster is electrically inactive, being
583: responsible for arsenic deactivation and structural changes in
584: heavily doped silicon. Larger defect clusters (e.g.,
585: $\text{V}_{2}\text{As}_{4}$) should also form during annealing,
586: whereas $\text{VAs}_{4}$ is only the first step in the clustering
587: process. These investigations were continued in,
588: \cite{Ramamoorthy_96} where general $\text{V}_{n}\text{As}_{m}$
589: complexes have been considered. The formation energies of the
590: vacancy-donor complexes $\text{VAs}_{1}$, $\text{VAs}_{2}$,
591: $\text{VAs}_{3}$, and $\text{VAs}_{4}$ were found to be 2.47,
592: 0.82, -0.53, and -2.39 eV, respectively. Moreover, it was proposed
593: that the complex $\text{VAs}_{2}$ was mobile, as were the
594: $\text{VAs}_{1}$ pairs. As these pairs moved, they reacted with
595: other defects to form larger, immobile complexes. When considering
596: the mobile $\text{VAs}_{2}$ complexes, Ramamoorthy and Pantelides
597: \cite{Ramamoorthy_96} have tried to explain the coupled diffusion
598: phenomenona and clustering. It was supposed that $\text{VAs}_{2}$
599: and $\text{VAs}_{3}$ were the dominant complexes in the
600: deactivated specimens near the enhanced-diffusion threshold.
601: \cite{Larsen_86,Larsen_93} A high rate of As diffusion observed
602: during rapid thermal annealing \cite{Larsen_86} was due to
603: $\text{VAs}_{2}$ migration over a very short time period
604: (6\,\,-\,60 s), when no extensive clustering could occur. During
605: their motion, $\text{VAs}_{2}$ complexes reacted with other
606: defects and formed larger, immobile complexes, which decreased the
607: As diffusivity.
608:
609: Nevertheless, further investigation is necessary because some
610: theoretical results disagree with the experimental data. For
611: example, in \cite{Chadi_97,Citrin_03} it was stated that the
612: formation of $\text{VAs}_{3}$ or $\text{VAs}_{4}$ defects was
613: exothermic, but, according to the first-principles calculations,
614: \cite{Ramamoorthy_96} buildup of $\text{VAs}_{1}$ or
615: $\text{VAs}_{2}$ structures was endothermic. This means that even
616: at low arsenic concentrations all impurity atoms must form
617: $\text{VAs}_{3}$ or $\text{VAs}_{4}$ clusters after long-time
618: treatment. In point of fact, at low dopant concentrations, one can
619: observe intensive diffusion by means of $\text{VAs}_{1}$ pairs
620: rather than clustering.
621:
622: Berding et al., \cite{Berding_98} and Berding and Sher
623: \cite{Berding_Sher_98} have used the electronic quasichemical
624: formalism to calculate a free energy of various clusters,
625: including $\text{As\,Si}_{4}$, $\text{VAs}_{4}$,
626: $\text{VAs}_{3}\text{Si}_{1}$, $\text{VAs}_{2}\text{Si}_{2}$. Here
627: $\text{As\,Si}_{4}$ is the arsenic atom in the substitutional
628: position. The neutral cluster composed of threefold-coordinated
629: second-neighbor arsenic atoms DP(2) was proposed by Chadi et al.
630: \cite{Chadi_97} as an alternative to exothermic $\text{VAs}_{4}$.
631: The clusters DP(2), $\text{V}_{2}\text{As}_{6}$, and
632: $\text{Si}_{8}$ were also included in the calculations performed
633: in. \cite{Berding_Sher_98} In contrast to, \cite{Ramamoorthy_96}
634: Berding et al., \cite{Berding_98} Berding and Sher
635: \cite{Berding_Sher_98} take into account the ionized states and
636: entropy of the cluster formation. The entropy is unfavourable for
637: the formation of a large defect complex such as $\text{VAs}_{4}$.
638: Therefore, the complete free-energy calculation is needed to
639: determine the role of $\text{VAs}_{4}$ in deactivation. Based on
640: the complete free-energy calculation, it was found that cluster
641: $\text{VAs}_{4}$ was neutral, in agreement with the previous
642: findings. \cite{Pandey_88} Also, the energy of this complex (-1.62
643: eV) was in a rough agreement with the value given in.
644: \cite{Pandey_88} The $\text{VAs}_{3}$ and $\text{VAs}_{2}$
645: clusters were found to have one and two acceptor levels within the
646: energy gap. Consequently, their formation energy may be
647: effectively decreased when the Fermi energy is near the
648: conduction-band edge. The equilibrium cluster concentrations
649: depending on the temperature and concentration of dopant atoms are
650: obtained using the minimized free energy of the system. For all
651: the dopant concentrations and temperatures considered
652: (400\,\,--\,1000\,$^{\circ}$C), three classes of clusters are
653: dominant under equilibrium conditions: $\text{Si}_{8}$,
654: $\text{AsSi}_{4}$, and $\text{VAs}_{4}$. In all cases, the Fermi
655: energy was near the conduction-band edge. Predominantly,
656: $\text{VAs}_{3}$ and $\text{VAs}_{2}$ were singly and doubly
657: ionized, respectively. At low arsenic concentrations (up to $\sim
658: 5\times10^{6}$ $\mu$\text{m}$^{-3}$), noticeable deactivation was
659: absent for temperatures higher than 500\,$^{\circ}$C. With the
660: arsenic concentration raised to $5\times10^{7}$ $\mu
661: $\text{m}$^{-3}$ and higher, significant deactivation was
662: predicted, mainly due to the formation of $\text{VAs}_{4}$
663: clusters. Both concentration and temperature influenced the
664: contributions of different defects. For example, the concentration
665: of $\text{VAs}_{4}$ clusters has reached the concentration of
666: isolated arsenic atoms in the lattice $\text{As\,Si}_{4}$ for
667: $\thicksim 5\times10^{8}$ $\mu $\text{m}$^{-3}$ at a temperature
668: of 700\,$^{\circ}$C and for $\thicksim 2\times10^{9}$ $\mu
669: $\text{m}$^{-3}$ at a temperature of 1000\,$^{\circ}$C.
670:
671: In \cite{Berding_98,Berding_Sher_98} the authors have attempted to
672: explain the effect of the electron concentration saturation at
673: high doping levels. However, the calculations performed in these
674: papers show that saturation is not reachable even at very high
675: arsenic concentrations. For example, saturation was not reached
676: with the arsenic concentration of $2\times10^{9}$ $\mu $m$^{-3}$
677: at a temperature of 700\,$^{\circ}$C, which is in disagreement
678: with the experimental data (see Fig. 2 from \cite{Berding_98}). It
679: is important that, according to, \cite{Berding_Sher_98} DP(2)
680: clusters proposed by Chadi et al. in \cite{Chadi_97} may be
681: present as a deactivating species, insignificantly contributing to
682: the deactivation under conditions of full equilibration.
683:
684: These new defects, first mentioned in \cite{Chadi_97} and called
685: DP, represent a pair of two three-fold coordinated donor atoms.
686: The lowest energy DP structures are DP(2) and DP(4), where the
687: donor atoms occupy either second- or fourth-neighbor Si sites
688: along the $<110>$ direction. It was shown that DP(2) exist in the
689: stable electrically active 2+ charge state that is donating
690: electrons to the conduction band or in the metastable neutral
691: charge state capturing two electrons from the Fermi sea. A very
692: important conclusion made in \cite{Chadi_97} is a threefold
693: coordination of each dopant atom in DP(2) in the neutral trap
694: state. Thus, relaxation of the neighboring Si atoms creates an
695: increased volume, which is consistent with the experimental
696: observations. Therefore, such an increased volume around the donor
697: atoms is possible without vacancies. The main ideas of
698: \cite{Chadi_97} were further developed in. \cite{Citrin_03} Citrin
699: et al. \cite{Citrin_03} have proposed a new class of defects,
700: called the donor-pair-vacancy-interstitials and composed of two
701: dopant donor atoms near the displaced Si atom that is forming a
702: vacancy-interstitial pair. This defect, which is like a hybrid of
703: the donor-pair and Frenkel-pair defects, is denoted DP(i)V-I. For
704: the case of Sb, the data of annular dark-field (``Z-contrast'')
705: scanning transmission electron microscopy (ADF-STEM) clearly
706: demonstrate that, in heavily Sb-doped silicon grown at low
707: temperatures, the primary deactivating defect contains only two Sb
708: atoms. \cite{Voyles_02} Thus, energetically favorable
709: $\text{VSb}_{3}$ and $\text{VSb}_{4}$ may be ruled out from the
710: deactivation process. Only $\text{VSb}_{2}$, DP(2), and DP(4) may
711: be considered as liable candidates. Using additional STEM
712: measurements, x-ray absorption data, and first-principles
713: calculations, Voyles et al. \cite{Voyles_02} have shown that
714: neither $\text{VSb}_{2}$ nor DP(2) and DP(4) defects are important
715: in heavily Sb-doped Si. At that time, DP(2)V-I and DP(4)V-I were
716: in line with the available experimental results, including
717: positron annihilation spectroscopy data. \cite{Citrin_03} Both
718: DP(2)V-I and DP(4)V-I are independent out of the pre-existing
719: vacancy population, which is in conformity with the observations
720: of. \cite{Solmi_02}
721:
722: Nevertheless, as follows from the experimental data,
723: \cite{Solmi_01} a deactivation mechanism can be different in the
724: As- and Sb-doped silicon. The mechanism of arsenic clustering
725: calls for further investigation. In, \cite{Mueller_03} Mueller et
726: al. have investigated the electronic structure and charge states
727: of various vacancy-impurity clusters using the first-principles
728: density-functional theory (DFT). It was found that the
729: $\text{VAs}_{1}$ complex can trap up to two conduction electrons
730: at a high n-doping level. Both $\text{VAs}_{2}$ and
731: $\text{VAs}_{3}$ act as a single-electron trap center. In
732: agreement with,
733: \cite{Pandey_88,Ramamoorthy_96,Berding_98,Berding_Sher_98} Mueller
734: et al. \cite{Mueller_03} have found that in $\text{VAs}_{4}$ the
735: fifth valence electrons of all four As atoms are strongly bound
736: and the complex is expected to remain neutral for any position of
737: the Fermi level. Note that according to the calculations of,
738: \cite{Berding_98} the cluster $\text{VAs}_{2}$ has two acceptor
739: levels, whereas Solmi et al. \cite{Solmi_00} have found from the
740: carrier mobility measurements that the complex to be electrically
741: neutral at room temperature. In \cite{Xie_99} the first-principles
742: calculations were performed to explain the coupled diffusion
743: phenomenona and clustering in heavily arsenic-doped silicon. As
744: contrast to, \cite{Ramamoorthy_96} it was found that the
745: $\text{VAs}_{2}$ cluster is less mobile due to its high migration
746: barrier. However, these clusters can contribute to the diffusion
747: of As at elevated temperatures.
748:
749: Proceeding from the presented analysis, our understanding of the
750: clustering mechanisms is neither complete nor firmly established.
751: This concerns the number of arsenic atoms in the cluster as well as
752: its structure. However, a study based on the first-principles
753: shows that clusters of a certain kind have an acceptor level and can
754: be negatively charged. Let us consider the models for the clusters
755: with different charge states.
756:
757: \subsection{Models based on the mass action law}
758:
759: The models based on thermodynamic formalism are usually used for
760: simulation of high concentration dopant diffusion taking place
761: during semiconductor processing. Assuming a local thermodynamic
762: equilibrium among substitutionally dissolved arsenic, the dopant
763: atoms incorporated into clusters, and electrons, one can use the
764: mass action law to calculate the concentration of clustered dopant
765: atoms. \cite{Tsai_80,Fair_73,Fair_Weber_73,Guerrero_82} The charge
766: conservation law for the reaction of cluster formation under the
767: assumption of local charge neutrality is also useful for
768: describing the clustering phenomenon. A widespread model of Tsai
769: et al. \cite{Tsai_80} was the first to account for the saturation
770: of the charge carriers at increasing doping levels.
771: \cite{Guerrero_82} It is interesting to compare this model with
772: the latest experimental data. The mass action law for the reaction
773: (\ref{Cluster_Reaction_Tsai}) at the diffusion temperature yields
774: \cite{Tsai_80}
775:
776: \begin{equation} \label{MAL_Tsai}
777: C^{Cl}=A^{Cl}nC^{3},
778: \end{equation}
779:
780: \noindent where $C^{Cl}$ and $n$ are the concentrations of
781: clusters and electrons at the annealing temperature, respectively;
782: $A^{Cl}$ is the constant of local thermodynamic equilibrium. After
783: cooling, all clusters become neutral and $C\approx n_{R}$ if $C\gg
784: n_{i}$. Here $n_{R}$ is the concentration of electrons at room
785: temperature. It follows from the assumption of local charge
786: neutrality that the electron concentration at the annealing
787: temperature for $C\gg n_{i}$ is
788:
789: \begin{equation} \label{Electron_equation_Tsai}
790: n\approx C+2C^{Cl}=C+2A^{Cl}nC^{3}.
791: \end{equation}
792:
793: Solving this equation, Tsai et al. \cite{Tsai_80} obtained
794:
795: \begin{equation} \label{Electron_Tsai}
796: n=\frac{C}{1-2A^{Cl}C^{3}}
797: \end{equation}
798:
799: \noindent and
800:
801: \begin{equation} \label{Total_Tsai}
802: C^{T}=C+\frac{3A^{Cl}C^{4}}{1-2A^{Cl}C^{3}}=\frac{C+A^{Cl}C^{4}}{1-2A^{Cl}C^{3}}.
803: \end{equation}
804:
805: From Eq.(\ref{Total_Tsai}) it is inferred that under condition
806:
807: \begin{equation} \label{Csat_Tsai}
808: C=C_{sat}=\frac{1}{\sqrt[3]{2A^{Cl}}}
809: \end{equation}
810:
811: \noindent both the concentration of arsenic atoms incorporated
812: into clusters and the total concentration of arsenic atoms
813: approach infinity. The value $n_{Re}=C_{sat}$ is interpreted in
814: \cite{Tsai_80} as a maximum level of electron concentration. From
815: the experimental data it is seen that a maximum equilibrium
816: carrier concentration at the annealing temperature in the arsenic
817: implanted layers may be described by the following relation
818: \cite{Solmi_01}
819:
820: \begin{equation} \label{Electron_Solmi}
821: n_{e}=1.3\times10^{11}\exp\left(-\frac{0.42 eV}{k_{B}T}\right).
822: \end{equation}
823:
824: Using Eqs.\;(\ref{Electron_Solmi}) and (\ref{Csat_Tsai}), the
825: value of $A^{Cl}$ may be derived to be
826:
827: \begin{equation} \label{Constant_Tsai}
828: A^{Cl}=\frac{1}{2n_{e}^{3}}.
829: \end{equation}
830:
831: To illustrate, $n_{e}=3.57\times10^{8}$ $\mu $m$^{-3}$ at a
832: temperature of 1050\,$^{\circ}$C and $A^{Cl}=1.10\times10^{-26}$
833: $\mu $m$^{9}$.
834:
835: The values of $n_{R}$ as a function of the total arsenic
836: concentration calculated by the model of Tsai et al.
837: \cite{Tsai_80} are shown in Fig. 1. The experimental values of
838: $n_{R}$ based on the data from \cite{Solmi_98} is also given for
839: comparison. As seen from Fig. 1, saturation of the electron
840: concentration for the model \cite{Tsai_80} is observed at the
841: extremely high total As concentrations, whereas the experimental
842: data show that the carrier concentration reaches its maximum at
843: $C^{T}\approx 8\times10^{8}$ $\mu $m$^{-3}$. Thus, the model
844: \cite{Tsai_80} is inconsistent with the experimental data.
845: Moreover, as follows from the experimental data of Solmi and
846: Nobili, \cite{Solmi_98} the value of $n_{e}$ obtained at room
847: temperature corresponds to the peculiarity of the curve for
848: diffusivity as a function of the total arsenic concentration.
849: Because this feature characterizes impurity diffusion at the
850: annealing temperature, it was concluded in \cite{Solmi_98} that
851: $n_{e}$ had a physical meaning. Thus, the model \cite{Tsai_80}
852: that includes cluster neutralization during cooling is also
853: inconsistent with the experimental data for arsenic diffusivity.
854: Indeed, the electron concentration at the annealing temperature
855: calculated by this model and associated with the peculiarity of
856: the diffusivity is three times greater than $n_{e}$.
857:
858:
859: \begin{figure}[ht]
860: \centering {
861: \begin{minipage}[ht]{8.6 cm}
862: {\includegraphics[ scale=0.74]{Fig1.eps}}
863:
864: \end{minipage}
865: }\caption{Calculated electron concentration against the total
866: arsenic concentration for different models of clustering: solid
867: line --- model of Tsai et al. (Ref. \cite{Tsai_80}), dash-dotted
868: line --- neutral clusters $\text{VAs}_{2}$, and dashed line ---
869: neutral clusters $\text{VAs}_{4}$. The experimental data (circles)
870: are taken from Solmi and Nobili \cite{Solmi_98} for the diffusion
871: at a temperature of 1050\,$^{\circ}$C. \label{As_Cluster1}}
872: \end{figure}
873:
874:
875:
876: Fig. 1 presents the functions $n=n(C^{T})$ calculated for the
877: cases when neutral clusters $VAs_{2}$ and $VAs_{4}$ are formed. In
878: these cases, saturation of the electron density is not observed and
879: the calculations do not agree with the experimental data. The
880: absence of saturation for the electron concentration corresponds
881: to the first-principles calculations in \cite{Berding_98} (Fig. 2)
882: and \cite{Berding_Sher_98} (Fig. 3). As seen from these figures,
883: no saturation is observed at arsenic concentrations when $VAs_{4}$
884: clustering is prevailing.
885:
886: The reactions
887:
888: \begin{equation} \label{Reaction_Guerrero}
889: m\text{As}^{+}+k e^{-}\rightleftarrows \text{As}_{m}^{r}
890: \end{equation}
891:
892: \noindent were used in the basic model for the clustering of As in
893: silicon proposed by Guerrero et al. \cite{Guerrero_82} Here $m$
894: and $k$ are the numbers of arsenic atoms and electrons
895: participating in the clustering process; $r$ is the electric
896: charge of a cluster. The function $C=C(C^{T})$ was investigated by
897: means of the mass action law combined with the charge conservation
898: law for the cluster formation reaction on the assumption of local
899: charge neutrality
900:
901: \begin{equation} \label{LCN_Guerrero}
902: n=C+rC^{Cl}=C+rA^{Cl}C^{m}n^{k}.
903: \end{equation}
904:
905: It was supposed in \cite{Guerrero_82} that the charge of clusters
906: becomes zero when the sample is cooled down to room
907: temperature. Therefore, $C$ is interpreted as the concentration of
908: electrically active arsenic at room temperature, which equals
909: the electron density $n$ in the sample at room temperature. Four
910: main conclusions have been drawn from the basic model:
911:
912: 1. When the electrons do not participate in the reaction
913: (\ref{Reaction_Guerrero}), i.e., $k=0$, or the clusters are
914: electrically neutral at high temperature, i.e., $r=0$, $C$ is a
915: monotonically increasing function of $C^{T}$ and also there is no
916: finite limit to $C$ and electron density $n$ at room temperature.
917:
918: 2. If the clusters are positively charged at a high temperature
919: ($r\geq 1$) and exactly one electron takes part in the reaction
920: (\ref{LCN_Guerrero}), i.e., $k=1$, $C$ is a monotonically
921: increasing function of $C^{T}$, which approaches a finite
922: ``saturation'' value $C_{max}$ with increasing $C^{T}$.
923:
924: As can be see from the reaction (\ref{Cluster_Reaction_Tsai}), the
925: model of Tsai et al. \cite{Tsai_80} satisfies these conditions. On
926: the other hand, from our calculations for this model (see Fig. 1),
927: it follows that saturation of the electron concentration is
928: observed at extremely high total As concentrations. It means that
929: the saturation value obtained in such a manner is inconsistent
930: with the experimental data.
931:
932:
933: 3. Provided that the clusters are positively charged at high
934: temperatures ($r\geq 1$) and there are two or more electrons
935: taking part in the reaction (\ref{Reaction_Guerrero}), i.e.,
936: $k\geq 2$, $C$ increases with increasing $C^{T}$ only up to a
937: finite value $C^{T}_{max}$. A maximum concentration value
938: $C=C_{max}$ is observed for $C^{T}=C^{T}_{max}$. With further
939: increase in the total arsenic concentration $C^{T}>C^{T}_{max}$
940: the concentration of the substitutionally dissolved arsenic atoms
941: $C$ and hence electron density $n$ at room temperature, are
942: decreased.
943:
944: This behavior of $n$ is at variance with the experimental data,
945: because experimentally measured $n$ is approximately constant
946: with increasing $C^{T}$ (see Fig. 1).
947:
948: 4. If the clusters are negatively charged at high temperatures
949: ($r\leq -1$), then $C=C(C^{T})$ behaves as in case 1, i.e.,
950: $C$, and hence electron density $n$ at room temperature,
951: monotonically increase without limit for increasing $C^{T}$. This
952: behavior of $n$ also conflicts with the experimental data.
953:
954: In what follows, the basic model is extended to take into account
955: the different charge states of vacancies participating in the
956: cluster formation. \cite{Guerrero_82} It is shown that the main
957: conclusions from the basic model are valid for the case when
958: vacancy is involved in the formation of clusters. Thus, the
959: generalized model of arsenic clustering proposed in
960: \cite{Guerrero_82} is not in agreement with the experimental data
961: for all possible cases of the cluster formation considered in this
962: model. In our opinion, this contradiction is due to the assumption
963: that charged clusters become neutral during cooling. As follows
964: from, \cite{Solmi_98} the hypothesis that deactivation is taking
965: place during cooling of the substrates contradicts the
966: experimental data.
967:
968: Evidently, the models of arsenic clustering should give an
969: adequate explanation not only for saturation of the electron
970: density, but also for the plateau formed on the carrier
971: concentration profile after annealing of the ion implanted layers.
972: \cite{Solmi_98,Solmi_01} As follows from the above analysis, there
973: is no model satisfying these requirements.
974:
975: The performed analysis enables us to formulate the purpose of this
976: study: the development of a more adequate model %for arsenic clustering
977: to explain the plateau phenomenon of the charge carrier profile at
978: high concentration arsenic diffusion.
979:
980: %For example, at higher temperatures concentration of isolated
981: %arsenic atom in the lattice $AsSi_{4}$ prevails on the
982: %concentration of $VAs_{4}$.
983:
984: \section{\textbf{Model}}
985:
986: To develop a model for arsenic clustering that can describe the
987: saturation phenomenon of the electron density, we assume that not
988: only neutral, but also negatively charged, arsenic clusters are
989: formed. The possibility for the formation of the negatively
990: charged arsenic clusters is confirmed by the calculations in.
991: \cite{Berding_98,Mueller_03} It has been shown by Berding et al.
992: \cite{Berding_98} that clusters $VAs_{3}$ and $VAs_{2}$ have one
993: and two acceptor levels, respectively. The calculations carried
994: out by Mueller et al. \cite{Mueller_03} also show that for the
995: electron densities $n>10^{4}$ $\mu \text{m}^{-3}$ the majority of
996: complexes $\text{VAs}_{2}$ will be negatively charged. According
997: to, \cite{Mueller_03} the complex $\text{VAs}_{3}$ exhibits an
998: acceptor level. This cluster is singly ionized at elevated Fermi
999: levels, resulting in a loss of four mobile carriers for the
1000: formation of a $\text{VAs}_{3}$ aggregate.
1001:
1002: It is interesting that saturation of the electron density is
1003: observed in the layers of GaAs heavily doped by silicon. To
1004: explain this phenomenon, it was supposed (see, for example
1005: \cite{Greiner_84,Velichko_93}) that silicon atoms in heavily doped
1006: GaAs are dissolved substitutionally not only in the Ga sublattice
1007: (donor $\text{Si}_{\text{Ga}}^{+}$), but also in the As sublattice
1008: (acceptor $\text{Si}_{\text{As}}^{-}$). It was supposed in
1009: \cite{Adler_95} that a maximum electrical conductivity in Sb-doped
1010: silicon is limited by the defect formation at high dopant
1011: concentrations. The Sb-vacancy pairs were considered as defect
1012: dominating at high doping levels. Also, it was supposed that this
1013: defect acts as an electron acceptor and is responsible for
1014: saturation of charge carriers at high doping levels. Although a
1015: mechanism of such a saturation of the electron density in silicon
1016: heavily doped by Sb is different from that in As-doped silicon,
1017: \cite{Solmi_01} the idea of the limited electron density due to
1018: compensation by the acceptors is a very fruitful one. For example,
1019: it has been supposed in \cite{Velichko_Dobrushkin_Pakula_05} that
1020: phosphorus clusters are negatively charged. Due to this
1021: assumption, saturation of the electron density during high
1022: concentration phosphorus diffusion was explained.
1023:
1024:
1025: As distinct from the models of arsenic clustering considered in,
1026: \cite{Tsai_80,Guerrero_82} we assume that arsenic clusters have
1027: the same negative charge both at high diffusion temperatures and
1028: at room temperature. As follows from the analysis of the
1029: experimental data and theoretical investigations, the structure of
1030: arsenic clusters under diffusion temperature is still not
1031: conclusively established. Therefore, it is assumed that a point
1032: defect $\text{D}_{1}$ can be involved in the cluster formation and
1033: another defect $\text{D}_{2}$ can be generated during clustering.
1034: In this case, a reaction for the formation and dissolution of
1035: clusters can be written as
1036:
1037: \begin{equation}\label{Reaction_Basic}
1038: m\text{As}^{+} + m_{1}\text{D}^{r_{1}}+ ke^{ -}
1039: {\longleftrightarrow} \Phi^{r_{Cl} } + m_{2}\text{D}^{r_{2}} ,
1040: \end{equation}
1041:
1042: \noindent where A is the substitutionally dissolved arsenic atom
1043: $\text{As}^{+}$; $e^{-}$ is the electron; $\Phi$ is the cluster
1044: formed; $m_{1}$ and $m_{2}$ are respectively the numbers of
1045: defects $\text{D}_{1}$ and $\text{D}_{2}$ participating in the
1046: cluster formation; $r_{1}$, $r_{2}$, and $r_{Cl}$ are the charge
1047: states of defect $\text{D}_{1}$, defect $\text{D}_{2}$, and
1048: cluster $\Phi^{r_{Cl} }$, respectively.
1049:
1050: The charge conservation law for the chemical reaction
1051: (\ref{Reaction_Basic}) is of the form
1052:
1053: \begin{equation} \label{CCL_Basic}
1054: m + m_{1} z_{1} - k = z_{Cl}+ m_{2} z_{2} ,
1055: \end{equation}
1056:
1057: \noindent where $z_{1}$ and $z_{2}$ are the charges of defects
1058: $\text{D}_{1}$ and $\text{D}_{2}$, respectively; $z_{Cl}$ is the
1059: charge of cluster $\Phi^{r_{Cl} }$ in terms of the elementary
1060: charge.
1061:
1062: The mass action law for reaction (\ref{Reaction_Basic}) yields
1063:
1064: \begin{equation} \label{MAL_Basic}
1065: K_{L} \;C^{\DS{m}} \;(C_{D1}^{r_{1}} )^{\DS{m_{1}}}
1066: \;\chi^{\DS{k}}n_{i}^{\DS{k}} = K_{R} \;C_{\Phi}
1067: \;(C_{D2}^{r_{2}} )^{\DS{m_{2}}} ,
1068: \end{equation}
1069:
1070: \noindent where $K_{L}$ is the rate of chemical reaction
1071: (\ref{Reaction_Basic}) in the forward direction; $K_{R}$ is the
1072: rate of this reaction in the opposite direction; $C$ is the
1073: concentration of impurity atoms participating in reaction
1074: (\ref{Reaction_Basic}) ; $C_{D1}^{r_{1}}$ is the concentration of
1075: defects in the charge state $r_{1}$ participating in the cluster
1076: formation; $C_{D2}^{r_{2}}$ is the concentration of defects in
1077: the charge state $r_{2}$ generated during clustering; $C_{\Phi}$
1078: is the concentration of the clusters in the charge state $r_{Cl}$;
1079: $\chi$ is the electron density normalized to the concentration of
1080: the intrinsic charge carriers in a semiconductor during diffusion
1081: $n_{i}$.
1082:
1083: Due to a high dopant concentration, the approximation of local
1084: charge neutrality can be used for the evaluation of $\chi$
1085:
1086:
1087: \begin{eqnarray} \label{chi}
1088: \chi = \DS{\frac{n}{n_{i}}} &=& \frac{1}{2n_{i}} \biggl[ C+ z_{Cl}
1089: \; C_{\Phi} - N_{B} \nonumber \\
1090: &+& \left. \sqrt {\left( C +z_{Cl} \; C_{\Phi} - N_{B}
1091: \right)^{2} + 4n_{i}^{2}}\, \right]\, ,
1092: \end{eqnarray}
1093:
1094:
1095: \noindent where $N_{B}$ is the summarized concentration of
1096: acceptors; $n_{ie}$ is the effective concentration of the
1097: intrinsic charge carriers at a diffusion temperature calculated
1098: for a high doping level.
1099:
1100: It follows from equation (\ref{Reaction_Basic}) that the
1101: concentration of impurity atoms incorporated into clusters $C^{A}
1102: = mC_{\Phi}$. Moreover, considering highly mobile electrons, the
1103: mass action law for the reaction of defect conversion between the
1104: charge states is valid
1105:
1106: \begin{equation} \label{MAL_Electron_1}
1107: C_{D1}^{r_{1}} = h_{D1}^{r_{1}} \;n^{\DS{-z_{1}}} \,
1108: C_{D1}^{\times} = h_{D1}^{r_{1}} \, n_{i}^{\DS{ - z_{1}}} \;\chi
1109: ^{\DS{ - z_{1}}} \, C_{D1}^{\times} \quad { ,}
1110: \end{equation}
1111:
1112: \begin{equation} \label{MAL_Electron_2}
1113: C_{D1}^{r_{2}} = h_{D2}^{r_{2}} \;n^{\DS{-z_{2}}} \,
1114: C_{D2}^{\times} = h_{D2}^{r_{2}} \, n_{i}^{\DS{ - z_{2}}} \;\chi
1115: ^{\DS{ - z_{2}}} \, C_{D2}^{\times} \quad { ,}
1116: \end{equation}
1117:
1118: \noindent where $h_{D1}^{r_{1}}$ and $h_{D2}^{r_{2}}$ are the
1119: constants for the local thermodynamic equilibrium in the reactions
1120: when neutral defects $\text{D}_{1}^{\times}$ and
1121: $\text{D}_{2}^{\times}$ are converted into charge states $r_{1}$
1122: and $r_{2}$, respectively.
1123:
1124:
1125: Substituting (\ref{MAL_Electron_1}) and (\ref{MAL_Electron_2})
1126: into (\ref{MAL_Basic}) and taking into consideration that $C_{AC}
1127: = mC_{Cl}$, we obtain
1128:
1129: \begin{equation} \label{CA}
1130: C_{AC} = K\; \tilde {C}_{D} \;\chi ^{\displaystyle{(
1131: \;k\;+r_{2}m_{2}-r_{1}m_{1})}} \;C^{\displaystyle{m}} ,
1132: \end{equation}
1133:
1134: \noindent where
1135:
1136: \begin{equation} \label{Constant}
1137: K = \frac{m\;K_{L} \,(h_{D1 \DS{r_{1}}})^{\DS{m_{1}}} \,
1138: n_{i}^{\DS{(k+r_{2}m_{2}-r_{1}m_{1})}} \,(C_{D1
1139: \DS{\times}}^{eq})^{\DS{m_{1}}}} {K_{R} \,(h_{D2
1140: \DS{r_{2}}})^{\DS{m_{2}}} \,(C_{D2 \DS{\times}}^{eq})^{\DS{m_{2}}}
1141: \, },
1142: \end{equation}
1143:
1144:
1145: \begin{equation} \label{Rel_Defect}
1146: \tilde {C}_{D}=\frac{(\tilde {C}_{D1})^{\DS{m_{1}}}}{(\tilde
1147: {C}_{D2})^{\DS{m_{2}}}} \, , \quad \tilde {C}_{D1}= \frac{{C}_{D1
1148: \DS{\times}}}{{C}_{D1 \DS{\times}}^{eq}} \, ,\quad \tilde
1149: {C}_{D2}= \frac{{C}_{D2 \DS{\times}}}{{C}_{D2 \DS{\times}}^{eq}}
1150: \, .
1151: \end{equation}
1152:
1153: \noindent Here $C_{AC}$ is the concentration of clustered arsenic
1154: atoms and $C_{T} = C + C_{AC}$ is the total arsenic concentration.
1155: The parameter $K$ has a constant value depending on the
1156: temperature of diffusion. This value can be extracted from the
1157: best fit to the experimental data. The quantities ${C}_{D1
1158: \DS{\times}}^{eq}$ and ${C}_{D2 \DS{\times}}^{eq}$ are the
1159: equilibrium concentrations of corresponding defects in the neutral
1160: charge state.
1161:
1162:
1163:
1164: \section{\textbf{Results of calculations}}
1165:
1166: Analysis of Eqs. (\ref{chi}) and (\ref{CA}) shows that saturation
1167: of the electron density will be observed for the formation of
1168: singly negatively charged clusters incorporating one arsenic atom
1169: and a lattice defect. The calculations of the functions $C =
1170: C(C^{T})$, $C^{A} = C^{A}(C^{T})$, and $n = n(C^{T} )$ for this
1171: case are presented in Fig. 2. For comparison with the experimental
1172: data, \cite{Solmi_98} the processing temperature was chosen to be
1173: 1050\,$^{\circ}$C ($n_{i} $ = 1.1665$\times10^{7}$ $\mu$m$^{-3}$).
1174: To calculate the functions $C = C\left( C^{T} \right)$,
1175: $C^{A}(C^{T})$, and $n(C^{T})$, a system of nonlinear equations
1176: (\ref{chi}), (\ref{CA}) was solved numerically by Newton's method.
1177: In the case of singly negatively charged clusters
1178: $(\text{DAs})^{-}$ incorporating a lattice defect and one arsenic
1179: atom we adopt $r_{Cl}$ = -1 and $m_{1} $ = 1. Eq.
1180: (\ref{CCL_Basic}) for $m$ = 1 yields that $- z_{1}m_{1} -
1181: z_{2}m_{2} + k$ = 2, and the system (\ref{CA}), (\ref{chi}) is of
1182: the form:
1183:
1184: \begin{equation} \label{CA1}
1185: C^{A} = K\;\tilde {C}_{D} \;\chi ^{2}\;C \; ,
1186: \end{equation}
1187:
1188: \begin{eqnarray} \label{chi1}
1189: \chi &=& \frac{1}{2n_{i}} \biggl[ C - K \tilde{C}_{D} \chi^{2}C -
1190: N_{B}+ \nonumber \\
1191: &+& \left. \sqrt {\left( C - K \tilde {C}_{D} \chi ^{2}C - N_{B}
1192: \right)^{2} + 4n_{ie}^{2}}\quad\right] { .}
1193: \end{eqnarray}
1194:
1195: As can be seen from Fig. 1, the concentration of charge carriers
1196: in the saturation region is approximately equal to the value
1197: $n_{e}=$ 3.5665$\times $10$^{8}$ $\mu$m$^{-3}$. In Fig. 2 for
1198: $C^{T}\rightarrow+\infty$, the electron concentration reaches
1199: saturation at the value $n \sim $ 3.765$\times $10$^{8}$
1200: $\mu$m$^{-3}$ associated with the plateau of the charge carrier
1201: profile given in Fig. 1. The value of parameter $K\tilde{C_{D}}$
1202: for this case was chosen as $ 1.0698\times
1203: 10^{{ -4}}$ a. u. to provide a fit to the experimental value of
1204: $n_{e}$. As seen from Fig. 2, the function $n=n(C^{T})$ is similar
1205: to that predicted by the model of Tsai et al. \cite{Tsai_80}
1206: Saturation of the electron density occurs at extremely high values
1207: of the total impurity concentration $C^{T}$, which is in conflict
1208: with the experimental data.
1209:
1210: %\begin{figure}
1211: %\includegraphics{}
1212: %\caption{\label{}}
1213: %\end{figure}
1214:
1215: \begin{figure}[ht]
1216: \begin{minipage}[ht]{8.6 cm}
1217: {\includegraphics[ scale=0.74]{Fig2.eps}}
1218: \end{minipage}
1219:
1220: \caption{Calculated concentrations of substitutionally dissolved
1221: arsenic atoms (solid line), clustered arsenic atoms (dashed line),
1222: and electron density (dotted line) against the total dopant
1223: concentration for the formation of singly negatively charged
1224: clusters incorporating one arsenic atom. The measured electron
1225: density (circles) is taken from Solmi and Nobili \cite{Solmi_98}
1226: for the diffusion at a temperature of 1050\,$^{\circ}$C.
1227: \label{As_Cluster2}}
1228: \end{figure}
1229:
1230:
1231: In Fig. 3a the system of nonlinear equations (\ref{chi}),
1232: (\ref{CA}) is solved numerically for the formation of singly
1233: negatively charged clusters $(DAs_{2})^{-}$ incorporating a
1234: lattice defect and two arsenic atoms ($r_{Cl} $ = -1, $m_{1} $ =
1235: 1, and $m$ = 2). Then, from (\ref{CCL_Basic}), (\ref{chi}), and
1236: (\ref{CA}) it follows that:
1237:
1238: \begin{equation} \label{CA2}
1239: C^{A} = K\;\tilde {C}_{D} \;\chi ^{3}\;C^{2} \, ,
1240: \end{equation}
1241:
1242: \begin{eqnarray} \label{chi2}
1243: \chi &=& \frac{1}{2n_{i}} \Biggl[ C - \DS{\frac{1}{2}}K
1244: \tilde{C}_{D} \chi^{3}C^{2} - N_{B}+ \nonumber \\
1245: &+& \left. \sqrt {\left( C - \DS{\frac{1}{2}}K \tilde {C}_{D} \chi ^{3}C^{2} - N_{B}
1246: \right)^{2} + 4n_{ie}^{2}}\, \right] .
1247: \end{eqnarray}
1248:
1249:
1250: The calculations for $n=n(C^{T})$ show that, with the increasing
1251: total concentration of impurity atoms $C^{T}$, the concentration of
1252: charge carriers $n$ reaches its maximum value $n_{max}$
1253: =3.750$\times 10^{8}$ $\mu$m$^{-3}$ at $C^{T}_{\max} \sim
1254: 1.634\times$10$^{9}$ $\mu$m$^{-3}$ and then decreases
1255: monotonically. At maximum the concentration of the
1256: substitutionally dissolved arsenic atoms $C_{\max}$ is
1257: $7.943\times$10$^{8}$ $\mu$m$^{-3}$. The parameter
1258: $K\tilde{C_{D}}$ used in this calculation is equal to
1259: 4.0$\times$10$^{-14}$ $\mu$m$^{3}$. The calculated curve agrees
1260: with the experimental function $n=n(C^{T})$ much better than those
1261: obtained by the model of Tsai et al. \cite{Tsai_80} and the models
1262: of neutral $(\text{DAs}_{2})$ and neutral $(\text{DAs}_{4})$
1263: clusters.
1264:
1265: \begin{figure}[ht]
1266: \begin{minipage}[ht]{8.6cm}
1267: {\includegraphics[ scale=0.74]{Fig3a.eps}} \centerline{(a)}
1268: \end{minipage}
1269:
1270: \begin{minipage}[t]{8.6cm}
1271: {\includegraphics[ scale=0.74]{Fig3b.eps}}
1272:
1273: \centerline{(b)}
1274: \end{minipage}
1275:
1276: \caption{Calculated concentrations of substitutionally dissolved
1277: arsenic atoms (solid line), clustered arsenic atoms (dashed line),
1278: and electron density (dotted line) against the total dopant
1279: concentration for the formation of singly (a) and doubly (b)
1280: negatively charged clusters incorporating two arsenic atoms. The
1281: measured electron density (circles) is taken from Solmi and Nobili
1282: \cite{Solmi_98} for the diffusion at a temperature of
1283: 1050\,$^{\circ}$C. \label{As_Cluster3a-b}}
1284: \end{figure}
1285:
1286: The calculations of $n=n(C^{T})$ and $C^{A}=C^{A}(C^{T})$ were
1287: also performed for the clusters $(\text{DAs}_{2})^{2-}$,
1288: $(\text{DAs}_{3})^{-}$, $(\text{DAs}_{3})^{2-}$,
1289: $(\text{DAs}_{4})^{-}$, and $(\text{DAs}_{4})^{2-}$. And similar
1290: behavior of $n=n(C^{T})$ was observed for all these cases. Also,
1291: the concentration of charge carriers increases with increasing
1292: $C^{T}$, reaches a maximum, and then decreases monotonically. This
1293: decrease is more pronounced with an increased number of arsenic atoms
1294: in the cluster. Besides, the position of maximum carrier
1295: concentration is shifted to smaller dopant concentrations with
1296: increases in the negative charge of the cluster.
1297:
1298: It follows from comparison of the calculated distributions
1299: $n(C^{T})$ with the experimental data that the best fit is
1300: observed for the doubly negatively charged cluster
1301: $(\text{DAs}_{2})^{2-}$ (see Fig. 3b). To obtain the curves shown
1302: in Fig. 4, we solve the system (\ref{chi}), (\ref{CA}) having the
1303: following form for the clusters $(\text{DAs}_{2})^{2-}$ \, :
1304:
1305: \begin{equation} \label{CA22}
1306: C^{A} = K\;\tilde {C}_{D} \;\chi ^{4}\;C^{2} \; ,
1307: \end{equation}
1308:
1309: \begin{eqnarray} \label{chi22}
1310: \chi &=& \frac{1}{2n_{i}} \Biggl[ C - K \tilde{C}_{D}
1311: \chi^{4}C^{2} - N_{B}+ \nonumber \\
1312: &+& \left. \sqrt {\left( C - K \tilde {C}_{D}
1313: \chi ^{4}C^{2} - N_{B}
1314: \right)^{2} + 4n_{ie}^{2}} \, \right] \, .
1315: \end{eqnarray}
1316:
1317: At maximum value of $n_{max}=3.694\times$10$^{8}$ $\mu$m$^{-3}$,
1318: the concentration of the substitutionally dissolved arsenic atoms
1319: $C_{\max}$ is equal to $7.943\times$10$^{8}$ $\mu$m$^{-3}$. The
1320: calculated curve $n=n(C^{T})$ agrees well with the experimental
1321: data, although only one fitting parameter $K\tilde{C_{D}}=$
1322: 0.67$\times$10$^{-15}$ $\mu$m$^{3}$ has been used. Full agreement
1323: is reached in the region of the transition to saturation of the
1324: electron density $n$. This means that two arsenic atoms are
1325: incorporated in the cluster, and this arsenic cluster is doubly
1326: negatively charged, at least at the total dopant concentrations
1327: $C^{T}\leq$ 1.0$\times$10$^{9}$ $\mu$m$^{-3}$. Minor differences
1328: in values of the theoretical curve and experimental data for
1329: $C^{T}>$ 2.3$\times$10$^{9}$ $\mu$m$^{-3}$ can arise due to heavy
1330: doping effects (change of the silicon zone structure, changes in
1331: the constants of forward $K_{L}$ and backward $K_{R}$ reactions,
1332: etc). It is clear from expressions (\ref{chi}), (\ref{CA}) that
1333: variation of the defect concentration $\tilde {C}_{D}$ in the
1334: region of high impurity concentrations also influences the
1335: concentration of impurity atoms incorporated in the clusters and,
1336: hence, the concentration of charge carriers. A change of $\tilde
1337: {C}_{D}$ due to generation (absorption) of the defects in the
1338: region with high arsenic concentration is quite possible. For
1339: example, such a generation can occur due to the lattice expansion.
1340: Besides, it is possible that a high doping level leads to error in
1341: the determination of the experimental values for electron density,
1342: and the data for $C^{T}>$ 2.3$\times$10$^{9}$ $\mu$m$^{-3}$ may be
1343: incorrect.
1344:
1345: Figs. 4 and b show the functions $C = C(C^{T})$, $C^{A} =
1346: C^{A}(C^{T})$, and $n = n(C^{T})$ calculated for clusters
1347: $(\text{DAs}_{3})^{-}$ and $(\text{DAs}_{4})^{-}$, respectively.
1348: As can be seen from these figures, fitting to the experimental
1349: data is worse in comparison with Fig. 3b, especially for the total
1350: concentrations greater than 2.0$\times$10$^{9}$ $\mu$m$^{-3}$.
1351:
1352:
1353: \begin{figure}[ht]
1354: \begin{minipage}[ht]{8.6cm}
1355: {\includegraphics[ scale=0.74]{Fig4a.eps}} \centerline{(a)}
1356: \end{minipage}
1357:
1358: \begin{minipage}[ht]{8.6 cm}
1359: {\includegraphics[ scale=0.74]{Fig4b.eps}} \centerline{(b)}
1360: \end{minipage}
1361:
1362:
1363: \caption{Calculated concentrations of substitutionally dissolved
1364: arsenic atoms (solid line), clustered arsenic atoms (dashed line),
1365: and electron density (dotted line) against the total dopant
1366: concentration for the formation of singly negatively charged
1367: clusters incorporating three (a) and four (b) arsenic atoms. The
1368: measured electron density (circles) are taken from Solmi and
1369: Nobili \cite{Solmi_98} for the diffusion at a temperature of
1370: 1050\,$^{\circ}$C. \label{As_Cluster4a-b}}
1371: \end{figure}
1372:
1373:
1374: Note that multiple clustering can occur, too, and different clusters
1375: can coexist in the equilibrium state. For example, the clusters
1376: incorporating different numbers of arsenic atoms
1377: $(\text{DAs}_{1})^{-}$ and $(\text{DAs}_{2})^{-}$ or clusters in
1378: different charge states $(\text{DAs}_{2})^{\times}$,
1379: $(\text{DAs}_{2})^{-}$, and $(\text{DAs}_{2})^{2-}$ can coexist.
1380: As follows from the calculated curves for the electron density,
1381: when clusters $(\text{VAs}_{1})^{-}$ (Fig. 2) and
1382: $(\text{VAs}_{2})^{\times}$ (Fig. 1) are formed, it may be
1383: expected that the fall of the electron density in the plateau region
1384: will be less pronounced in the case of multiple clustering.
1385: However, this problem requires further investigation.
1386:
1387: \section{\textbf{Conclusions}}
1388:
1389: Based on analysis of the experimental results and theoretical
1390: calculations for clustering of arsenic atoms in silicon, it was
1391: shown that the developed clustering models fail to explain the
1392: available experimental data, especially the effect of electron
1393: density saturation at high arsenic concentrations. Therefore, a
1394: new model of arsenic clustering was proposed and analyzed. The
1395: main feature of the proposed model is the assumption that
1396: negatively charged arsenic complexes play a dominant role in the
1397: clustering process. To confirm this assumption, the concentration
1398: of the impurity atoms incorporated in clusters and electron
1399: density were calculated as a function of the total arsenic
1400: concentration at a temperature of 1050\,$^{\circ}$C. Different
1401: cases of the formation of negatively charged clusters
1402: incorporating a point defect and one arsenic atom
1403: $(\text{DAs}_{1})^{-}$, $(\text{DAs}_{1})^{2-}$ or more arsenic
1404: atoms $(\text{DAs}_{2})^{-}$, $(\text{DAs}_{2})^{2-}$,
1405: $(\text{DAs}_{3})^{-}$, $(\text{DAs}_{3})^{2-}$,
1406: $(\text{DAs}_{4})^{-}$, and $(\text{DAs}_{4})^{2-}$ were
1407: investigated. It was shown that in the case of
1408: $(\text{DAs}_{1})^{-}$ the concentration of charge carriers
1409: reaches saturation with increase in the total arsenic
1410: concentration and concentration of the substitutionally dissolved
1411: impurity atoms. However, saturation was observed for very high
1412: values of the total arsenic concentration, conflicting with the
1413: experimental data. In the cases of doubly negatively charged
1414: cluster $(\text{DAs}_{1})^{2-}$ and clusters incorporating more
1415: than one arsenic atom, the electron density reached its maximum
1416: value, then decreased monotonically and slowly with an increase
1417: in the total arsenic concentration. The electron density
1418: calculated for the formation of $(\text{DAs}_{2})^{2-}$ clusters
1419: agrees well with the experimental data and confirms the conclusion
1420: that two arsenic atoms participate in the cluster formation. Minor
1421: difference between the theoretical curve and experimental values
1422: for $C^{T}>$ 2.3$\times$10$^{9}$ $\mu$m$^{-3}$ may be explained by
1423: various heavy doping effects or uncertainty due to the
1424: experimental errors.
1425:
1426: Thus, the plateau on the experimental distribution of the charge
1427: carriers, characterized by a practically constant value of the
1428: electron density in the region of high doping level, may be
1429: attributed to the negatively charged clusters
1430: $(\text{DAs}_{2})^{2-}$. The parameters describing arsenic
1431: clustering at a temperature of 1050\,$^{\circ}$C were determined
1432: by fitting the calculated values of the electron density to the
1433: experimental data. The proposed model of clustering by the
1434: formation of the negatively charged $(\text{DAs}_{2})^{2-}$ system
1435: gives the best fit to the experimental data among all the
1436: present-day models and may be used in simulation of high
1437: concentration arsenic diffusion.
1438:
1439: \newpage
1440:
1441: \begin{thebibliography}{99}
1442:
1443: %Furnace annealing
1444:
1445: %Ion-implanted layers
1446: \bibitem{Tsukamoto_80} K. Tsukamoto, Y. Akasaka, and K. Kuima,
1447: Jpn. J. Appl. Phys. \textbf{19}, 87 (1980).
1448:
1449: \bibitem{Tsai_80} M. Y. Tsai, F. F. Morehead, J. E. E. Baglin,
1450: and A. E. Michel, J. Appl. Phys. \textbf{51}, 3230 (1980).
1451:
1452: \bibitem{Fahey_89} P. M. Fahey, P. B. Griffin, and
1453: J. D. Plummer, Rev. Mod. Phys. \textbf{61}, 289 (1989).
1454:
1455: \bibitem{Nobili_94} D. Nobili, S. Solmi, A. Parisini, M. Derdour,
1456: A. Armigliato, and L. Moro, Phys. Rew. B
1457: \textbf{49}, 2477 (1994).
1458:
1459: \bibitem{Solmi_98} S. Solmi and D. Nobili, J. Appl. Phys.
1460: \textbf{83}, 2484 (1998).
1461:
1462: \bibitem{Krishnamoorthy_98} V. Krishnamoorthy, K. Moller, K. S. Jones,
1463: D. Venables, J. Jackson, and L. Rubin, J. Appl. Phys. \textbf{84},
1464: 5997 (1998).
1465:
1466: \bibitem{Nobili_99} D. Nobili, S. Solmi, M. Merli, and J. Shao,
1467: J. Electrochem. Soc., \textbf{146}, 4246 (1999).
1468:
1469: \bibitem{Uematsu_00} M. Uematsu, Jpn. J. Appl. Phys., Part. 1
1470: \textbf{39}, 1006 (2000).
1471:
1472: \bibitem{Nobili_01} D. Nobili, S. Solmi, and J. Shao, J.
1473: Appl. Phys. \textbf{90}, 101 (2001).
1474:
1475: \bibitem{Solmi_01} S. Solmi, in {\itshape Encyclopedia of Materials:
1476: Science and Technology}, edited by K. H. J. Buschow, R. W. Cahn, M. C.
1477: Flemings, B. Ilschner, E. J. Kramer, S. Mahajan, and P.
1478: Veyssi\`{e}re (Elsevier Science Ltd., 2001) p. 2331.
1479:
1480: \bibitem{Solmi_03} S. Solmi, M. Ferri, M. Bersani, D. Giubertoni, and
1481: V. Soncini, J. Appl. Phys. \textbf{94}, 4950 (2003).
1482:
1483: %Thermal diffusion
1484:
1485: \bibitem{Fair_73} R. B Fair and G. R. Weber, J. Appl. Phys.
1486: \textbf{44}, 273 (1973).
1487:
1488: \bibitem{Fair_Weber_73} R. B Fair and G. R. Weber, J. Appl. Phys.
1489: \textbf{44}, 280 (1973).
1490:
1491: \bibitem{Murota_79} J. Murota, E. Arai, K. Kobayashi, and
1492: K. Kubo. J. Appl. Phys. \textbf{50}, 804 (1979).
1493:
1494: %Deactivation after laser anneling
1495:
1496: \bibitem{Parisini_90} A. Parisini, A. Bourret, A. Armigliato,
1497: M. Servidori, S. Solmi, R. Fabbri, J. R. Regnard, and J. L. Allain,
1498: J. Appl. Phys. \textbf{67}, 2320 (1990).
1499:
1500: \bibitem{Luning_92} S. Luning, P. M. Rousseau, P. B. Griffin, P. G. Carey,
1501: and J. D. Plummer, Tech. Dig. Int. Electron. Devices Meet. 457 (1992).
1502:
1503: \bibitem{Rousseau_94} P. M. Rousseau, P. B. Griffin, and
1504: J. D. Plummer, Appl. Phys. Lett. \textbf{65}, 578 (1994).
1505:
1506: \bibitem{Rousseau_98} P. M. Rousseau, P. B. Griffin, W. T. Fang,
1507: and J. D. Plummer, J. Appl. Phys. \textbf{84}, 3593 (1998).
1508:
1509: \bibitem{Solmi_02} S. Solmi, M. Attari, and D. Nobili, Appl.
1510: Phys. Lett. \textbf{80}, 4774 (2002).
1511:
1512: %Reverse annealing
1513:
1514: \bibitem{Solmi_00} S. Solmi, D. Nobili, and J. Shao, J. Appl.
1515: Phys. \textbf{87}, 658 (2000).
1516:
1517: %Precipitation
1518:
1519: \bibitem{Nobili_83} D. Nobili, A. Carabelas, G. Celotti, and
1520: S. Solmi, J. Electrochem. Soc. \textbf{130}, 922 (1983).
1521:
1522: \bibitem{La Via_91}F. La Via, V. Privitera, S. Lombardo, C. Spinella,
1523: V. Raineri, E. Rimini, P. Baeri, and G. Ferla, J. Appl. Phys.
1524: \textbf{69}, 726 (1991).
1525:
1526: \bibitem{Guerrero_82}
1527: E. Guerrero, H. P\"{o}tzl, R. Tielert, M. Grasserbauer, and G.
1528: Stingeder, J. Electrochem. Soc. \textbf{129},1826 (1982).
1529:
1530: \bibitem{Pandey_88} K. C. Pandey, A. Erbil, G. S. Cargill, III,
1531: R. F. Boehme, and David Vanderbilt, Phys. Rev. Lett. \textbf{61},
1532: 1282 (1988).
1533:
1534: \bibitem{Berding_98}
1535: M.A. Berding, A. Sher, and M. van Schilfgaarde, Appl. Phys. Let.
1536: \textbf{72}, 1492 (1998).
1537:
1538: \bibitem{Mueller_03} D. C. Mueller, E. Alonso, and W. Fichtner,
1539: Phys. Rev. B \textbf{68}, 045208 (2003).
1540:
1541: % Low-temperature treatments
1542:
1543: % Experimental investigations of defect subsystem
1544:
1545: % TEM
1546:
1547: \bibitem{Dokumachi_95} 0. Dokumaci, P. Rousseau, S. Luning,
1548: V. Krishnamoorthy, K. S. Jones, and M. E. Law,
1549: J. Appl. Phys. \textbf{78}, 828 (1995).
1550:
1551: % EXAFS
1552:
1553: \bibitem{Erbil_86} A. Erbil, W. Weber, G. S. Cargill III, and R. F.
1554: Boehme, Phys. Rev. B \textbf{34}, 1392 (1986).
1555:
1556: \bibitem{Allain_92} J. L. Allain, J. R. Regnard, A. Bourret,
1557: A. Parisini, A. Armigliato, G. Tourillon, and S. Pizzini, Phys. Rev. B
1558: \textbf{46}, 9434 (1992).
1559:
1560: \bibitem{d'Acapito_04} F. d'Acapito, C. Maurizio, and M. Malvestuto,
1561: Mater. Sci. Eng., B \textbf{114-115}, 386 (2004).
1562:
1563: % RBS
1564:
1565: \bibitem{Brizard_93} C. Brizard, J. R. Regnard, J. L. Allain,
1566: A. Bourret, M. Dubus, A. Armigliato, and A. Parisini,
1567: J. Appl. Phys. \textbf{75}, 126 (1994).
1568:
1569: % positron
1570:
1571: \bibitem{Makinen_89} J. M\"{a}kinen, C. Corbel, P. Hautoj\"{a}rvi,
1572: P. Moser, and F. Pierre, Phys. Rev. B \textbf{39}, 10162 (1992).
1573:
1574: \bibitem{Lawther_95} D.W. Lawther, U. Myler, P. J. Simpson,
1575: P. M. Rousseau, P. B. Griffin, and J. D. Plummer,
1576: Appl. Phys. Lett. \textbf{67}, 3575 (1995).
1577:
1578: \bibitem{Myler_96} U. Myler, R. D. Goldberg, A. P. Knights,
1579: D. W. Lawther, and P. J. Simpson, Appl. Phys. Lett. \textbf{69},
1580: 3333 (1996).
1581:
1582: \bibitem{Szpala_96} S. Szpala, P. Asoka-Kumar, B. Nielsen, J. P. Peng,
1583: S. Hayakawa, and K. G. Lynn, Phys. Rev. B \textbf{54}, 4722
1584: (1996).
1585:
1586: % positron + electron moments
1587:
1588: \bibitem{Polity_98} A. Polity, F. B\H{o}rner, S. Huth, S. Eichler, and
1589: R. Krause-Rehberg, Phys. Rev. B \textbf{58}, 10363 (1998).
1590:
1591: \bibitem{Saarinen_99} K. Saarinen, J. Nissil\H{a}, H. Kauppinen,
1592: M. Hakala, M. J. Puska, P. Hautoj\"{a}rvi, and C. Corbel, Phys.
1593: Rev. Lett. \textbf{82}, 1883 (1999).
1594:
1595: \bibitem{Ranki_02} V. Ranki, J. Nissil\"{a}, and K. Saarinen,
1596: Phys. Rev. Lett. \textbf{88}, 105506 (2002).
1597:
1598: \bibitem{Ranki_03} V. Ranki, K. Saarinen, J. Fage-Pedersen,
1599: J. Lundsgaard Hansen, and A. Nylandsted Larsen, Phys. Rev. B
1600: \textbf{67}, 041201(R) (2003).
1601:
1602: \bibitem{Ranki_Saarinen_03} V. Ranki and K. Saarinen, Physica B
1603: \textbf{340-342}, 765 (2003).
1604:
1605: \bibitem{Ranki_04} V. Ranki, A. Pelli, and K. Saarinen,
1606: Phys. Rev. B \textbf{69}, 115205 (2004).
1607:
1608: % x-ray standing-wave spectroscopy
1609:
1610: \bibitem{Herrera-Gomez_96} A. Herrera-G\"{o}mez, P. M. Rousseau,
1611: G. Materlik, T. Kendelewicz, J. C. Woicik, P. B. Griffin,
1612: J. Plummer, and W. E. Spicer, Appl. Phys. Lett. \textbf{68}, 3090 (1996).
1613:
1614: % First principles
1615:
1616: \bibitem{Ramamoorthy_96} M. Ramamoorthy and S. T. Pantelides,
1617: Phys. Rev. Lett. \textbf{76}, 4753 (1996).
1618:
1619: \bibitem{Chadi_97} D. J. Chadi, P. H. Citrin, C. H. Park, D. L. Adler,
1620: M. A. Marcus, and H.-J. Gossmann, Phys. Rev. Lett. \textbf{79}, 4834 (1997).
1621:
1622: \bibitem{Citrin_03} P.H. Citrin, P.M. Voyles, D.J. Chadi, and D.A. Muller,
1623: Physica B. \textbf{340-342}, 784 (2003).
1624:
1625: \bibitem{Larsen_86} A. Nylandsted Larsen, S. Yu. Shiryaev, E. S. S\o
1626: rensen, and P. Tidemand-Peterson, Appl. Phys. Lett. \textbf{48}, 1805 (1986).
1627:
1628: \bibitem{Larsen_93} A. Nylandsted Larsen, K. Kyllesbech Larsen,
1629: P. E. Andersen, and B. G. Svensson, J. Appl. Phys. \textbf{73},
1630: 691 (1993).
1631:
1632: \bibitem{Berding_Sher_98} M. A. Berding and A. Sher, Phys. Rev.
1633: B. \textbf{58}, 3853 (1998).
1634:
1635: % ADF-STEM experiments
1636: \bibitem{Voyles_02} P. M. Voyles, D. A. Muller, J. L. Grazul, P. H. Citrin,
1637: and H.-J. L. Gossmann, Nature, \textbf{416}, 826 (2002).
1638:
1639: % First principles
1640:
1641: \bibitem{Xie_99} J. Xie and S. P. Chen, Phys. Rev. Lett.
1642: \textbf{83}, 1795 (1999).
1643:
1644: \bibitem{Greiner_84} M. E. Greiner and J. F. Gibbons. Appl.
1645: Phys. Lett. \textbf{44}, 750 (1984).
1646:
1647: \bibitem{Velichko_93} O. I. Velichko, A. A. Egorov, and S. K. Fedoruk,
1648: Inzh.-Fiz. Zhurnal {\bf65}, 567 (1993) [J. Eng. Phys. Thermophys.
1649: {\bf65}, 1091 (1993)]. [Russian journal reference with English
1650: translation]
1651:
1652: \bibitem{Adler_95}D. L. Adler, J. D. Chadi, M. A. Marcus, H.-J. Gossmann,
1653: and P. H. Citrin, Bull. Amer. Phys. Soc. {\bf40}, 396 (1995).
1654:
1655: \bibitem{Velichko_Dobrushkin_Pakula_05}
1656: O. I. Velichko, V. A. Dobrushkin, L. Pakula, Mater. Sci. Eng. B
1657: {\bf123}, 176 (2005).
1658:
1659: \end{thebibliography}
1660:
1661: \end{document}
1662: