1: \documentclass[prb,a4paper,preprint,footinbib,showpacs]{revtex4}
2: \usepackage[latin1]{inputenc}
3: \usepackage{amssymb}
4: \usepackage{amsmath}
5: \usepackage{amsbsy}
6: \usepackage{bm}
7: \usepackage{graphicx}
8: \begin{document}
9:
10: \title{Electromigration-Induced Step Meandering on Vicinal Surfaces:
11: Nonlinear Evolution Equation}
12: \author{ Matthieu Dufay}
13: \affiliation {Laboratoire Mat\'eriaux et Micro\'electronique de Provence,
14: Aix-Marseille Universit\'e and CNRS,
15: Facult\'e des Sciences et Techniques de Saint-J\'er\^ome,
16: Case 151,
17: 13397 Marseille Cedex 20, FRANCE}
18:
19:
20: \author{ Jean-Marc Debierre}
21: \affiliation {Laboratoire Mat\'eriaux et Micro\'electronique de Provence,
22: Aix-Marseille Universit\'e and CNRS,
23: Facult\'e des Sciences et Techniques de Saint-J\'er\^ome,
24: Case 151,
25: 13397 Marseille Cedex 20, FRANCE}
26:
27:
28: \author{ Thomas Frisch}
29: \affiliation {Institut de Recherche sur les Ph\'enom\`nes Hors
30: Equilibre,
31: Aix-Marseille Universit\'e, Ecole Centrale Marseille, and CNRS,\\
32: 49, rue Joliot Curie,
33: BP 146,
34: 13384 Marseille Cedex 13, FRANCE}
35:
36: \begin{abstract}
37: We study the effect of a constant electrical field applied on
38: vicinal surfaces such as the Si$(111)$ surface. An electrical
39: field parallel to the steps induces a meandering instability with a nonzero phase shift.
40: Using the Burton-Cabrera-Frank model, we extend
41: the linear stability analysis performed by Liu, Weeks and Kandel
42: (Phys. Rev. Lett. {\bf 81}, p.2743, 1998) to the nonlinear regime
43: for which the meandering amplitude is large. We derive an
44: amplitude equation for the step dynamics using a highly nonlinear
45: expansion method. We investigate numerically two
46: limiting regimes (small and large attachment lengths)
47: which both reveal long-time coarsening dynamics.
48: \end{abstract}
49:
50:
51: \pacs{ 66.30.Qa, 47.20.Hw}
52:
53:
54:
55: \maketitle
56:
57: \section{Introduction}
58:
59:
60: Stepped crystal surfaces exhibit a number of different
61: morphological instabilities likely to play an important role during
62: crystal
63: growth\cite{pimpinelli98,saito98,jeong99,politi00,yagi01,pierre-louis03}.
64: Furthermore, the ability to control the growth of faceted stepped
65: crystal surfaces may be of considerable importance when
66: manufacturing electronic and optoelectronic devices \cite{stangl04}.
67: These morphological instabilities occur not only during growth and
68: evaporation but also under the influence of an electrical field,
69: as on the well-studied Si$(111)$ surface of a semiconductor. Surface
70: electromigration instabilities may also arise in metals, where they
71: are an important source of failure in microelectronic devices at
72: metal-metal interfaces and also an interesting tool for pattern
73: formation \cite{schimschak97, kuhn05}. One of the most studied
74: instability, known as step bunching, arises on the Si$(111)$
75: surfaces from the biased diffusion (drift) of adatoms under the
76: influence of an external driving force such as an electrical
77: constant field
78: \cite{latyshev89,stoyanov91,liu98bis,stoyanov00,metois01}. Step
79: bunching is a one-dimensional instability which has been explained
80: within the framework of the Burton-Cabrera Frank equations
81: \cite{burton51} in terms of displacements of steps and terraces.
82: Recent experimental and theoretical studies of step bunching
83: revealed several difficulties, like the complex role of step
84: transparency, the Ehrlich-Schwoebel barriers, the effect of
85: substrate temperature, and the variations of the adatom mobility
86: with the distance to the steps \cite{saul02,krug05,chang06,
87: pierre-louis06}.
88:
89: In the present study, a constant electrical field is applied
90: along the mean step direction of a train of synchronized steps (all identical up to a
91: constant phase-shift). An experimental study of a
92: comparable system was recently reported \cite{degawa01} and it was shown in
93: this work that a two-dimensional step meandering instability takes place.
94: The linear analysis of this problem was previously performed by Liu and co-authors
95: who predicted the occurrence of synchronized
96: meandering \cite{liu98}. We perform here a nonlinear
97: analysis of this instability in order to describe the long time
98: behavior of the in-phase meandering mode. In particular we
99: show the appearance of the coarsening regime in which steps undulations
100: increase. This paper is organized as follows. In the next section, we present
101: a model based on the Burton-Cabrera-Frank
102: equations \cite{burton51}. In the third section, we
103: perform the linear analysis, which serves as a basis for
104: the nonlinear analysis. A general nonlinear
105: evolution equation including the effects of the repulsive step-step interactions
106: is derived in section IV. The results of numerical simulations of this nonlinear equation
107: are presented and discussed in section V, while conclusions and perspectives
108: are postponed to section VI.
109:
110: Before presenting our model, we shortly review previous work
111: concerning nonlinear equations for the time evolution of
112: synchronized steps. The step meandering instability was originally
113: predicted theoretically by Bales and Zangwill \cite{bales90} for a
114: vicinal surface under growth. Its origin is the asymmetry between
115: the descending and ascending currents of adatoms. As shown by Bales
116: and Zangwill, a straight train of step may become morphologically
117: unstable during MBE growth if the kinetic attachment at the steps is
118: asymmetrical: this is the Ehrlich-Schwoebel effect. It was shown
119: that the most dangerous mode corresponds to a zero phase-shift
120: \cite{pimpinelli94}. Nonlinear extensions of this work have shown
121: that the meander evolution can be described by amplitude equations
122: displaying diverse behaviors. Close to the instability threshold,
123: starting from the Burton-Cabrera-Frank (BCF) model, it was proved
124: \cite{bena93} that the step position in the presence of desorption
125: (evaporation) obeys the Kuramoto-Sivashinsky equation. The ultimate
126: stage of this dynamics is thus spatio-temporal chaos. In the case
127: of negligible desorption with strong or moderate Ehrlich-Schwoebel
128: effect, it was found that the step amplitude obeys a highly nonlinear
129: equation. \cite{pierre-louis98,kallunki00,gillet00} This equation
130: cannot be derived from a weakly nonlinear analysis but is based on
131: the assumption that the slope of the steps is of order unity.
132: Instead of spatio-temporal chaos, a regular pattern is revealed: the
133: lateral modulation wavelength is fixed while the transverse
134: amplitude of the step deformation (meandering amplitude) increases.
135: Elasticity or diffusion anisotropy can also influence the meander
136: dynamics. \cite{paulin01, danker04} It was recently shown in
137: context of the step meandering instabilities during growth on
138: Si$(001)$ \cite{frisch06,frisch06a} that the nonlinear dynamics is
139: driven by a conserved Kuramoto-Sivashinsky equation. This
140: equation was already mentioned in Ref. \cite{gillet00} on the basis
141: of symmetry arguments but was not derived there, because of a
142: different scaling of the Ehrlich-Schwoebel effect. Step meandering
143: was also studied in the context of electromigration
144: \cite{sato00,sato03,sato05} using analytic linear analysis and
145: kinetic Monte Carlo simulations. It is therefore of importance to
146: extend the work of Liu {\it et al\/} \cite{liu98} and to develop
147: analytical tools to describe the nonlinear regime of the meandering
148: instability.
149:
150:
151:
152:
153: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
154: \section{Model}
155: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
156:
157: \subsection{Validity range and notations}
158: %%%%%%%%%%%%%%%%%%%%%
159:
160: The geometry of the problem is sketched in Fig. (\ref{Fig:Geom}).
161: Initially, all the step edges are directed along the $X$ axis, and equidistant
162: from a distance $L_0$.
163: A constant electrical field $E$ is applied in the positive $X$ direction.
164: To investigate the resulting meandering instability, we use a two-dimensional version
165: of the BCF model. The terraces are numbered sequentially in the step-down direction.
166: In our notations, the $n$-th terrace is bordered by
167: the two steps numbered $n$ and $n+1$. To distinguish between quantities defined anywhere
168: on the terrace and quantities defined at steps only, we will use an upper index $n$ in the first case
169: and a lower index $n$ in the second. For instance, the adatom concentration on
170: the $n$-th terrace is denoted $C^n$, while the equilibrium concentration at step $n$
171: reads $C^{eq}_n$.
172: \begin{figure}
173: \includegraphics[width=0.7\textwidth]{Fig1}
174: \caption{ \label{Fig:Geom} Schematic representation of a small
175: portion of a vicinal surface showing three steps. On the top view
176: (above), the electrical field $\bf E$ is represented. On the side
177: view (below), two attachment mechanisms are illustrated. }
178: \end{figure}
179:
180: In practice, it is usually assumed that the concentrations are not
181: explicitly time-dependent, so that $C^n=C^n(X,Y)$. This quasi-static
182: approximation is valid provided that the diffusion coefficient $D_s$
183: of the adatoms on the terraces is sufficiently large that diffusion
184: takes place on time scales shorter than those for step motion. The
185: diffusion bias introduced by the electrical field can be quantified
186: by the ratio of the thermal energy $k_BT$ to the electrical energy
187: $\vert Z^*e \vert E\ell_E$. Balancing the two terms defines the electrical length
188: as
189: \begin{equation}
190: \ell_E=\frac{k_B T}{\vert Z^*e\vert E}\ ,
191: \label{Eq:Xi}
192: \end{equation}
193: where $k_B$ is the Boltzmann constant, $T$ the absolute temperature, $Z^*$ the effective
194: atomic charge number, and $e$ the electron electrical charge.
195:
196: For the sake of simplicity, we set both deposition and evaporation
197: of adatoms to zero here, whereas experiments are usually performed with
198: a small but nonzero net flux.
199: Introducing both effects in our model is straightforward and would not affect
200: qualitatively the results obtained within the zero flux assumption.
201:
202: On the $n$-th terrace, the quasi-static biased diffusion equation
203: reduces to
204: \begin{equation}
205: D_s(\partial_{XX} +\partial_{YY}) C^n-(D_s/\ell_E) \partial_{X} C^n=0,
206: \label{Eq:DiffDim}
207: \end{equation}
208: and the adatom flux is
209: \begin{equation}
210: \mbox{\bf J}^n =D_s(1/\ell_E-\partial_X, -\partial_Y)C^n.
211: \label{Eq:Flux}
212: \end{equation}
213:
214: The boundary conditions for Eq. (\ref{Eq:DiffDim}) are obtained by
215: writing mass conservation at all the points $\mbox{\bf R}_n=(X,
216: Y_n)$ and $\mbox{\bf R}_{n+1}=(X, Y_{n+1})$ located on both edges of
217: the terrace. In the present model, we assume that the adatom
218: attachment/detachment kinetic coefficients are the same on the upper
219: and lower side of a given step, $\nu^+=\nu^-=\nu$. We further
220: restrict ourselves to temperature ranges where direct mass exchange
221: between adjacent terraces (transparency) can be neglected. At
222: $\mbox{\bf R}_n$, the boundary condition thus reads
223: \begin{equation}
224: \mbox{\bf J}^n(\mbox{\bf R}_n) \cdot \mbox{\bf u}_n=-\nu[C^n(\mbox{\bf R}_n)-C^{eq}_n],
225: \label{Eq:CLleft}
226: \end{equation}
227: where $\mbox{\bf u}_n$ represents the normal unit vector pointing
228: in the step-down direction.
229: Alternatively, at $\mbox{\bf R}_{n+1}$, the second boundary condition is
230: \begin{equation}
231: \mbox{\bf J}^n (\mbox{\bf R}_{n+1})\cdot \mbox{\bf u}_{n+1}=\nu[C^n(\mbox{\bf R}_{n+1})-C^{eq}_{n+1}].
232: \label{Eq:CLright}
233: \end{equation}
234: Writing mass conservation at any point $\mbox{\bf R}_n$ along step $n$,
235: we obtain the normal (component along $\mbox{\bf u}_n$) step velocity,
236: \begin{equation}
237: V_n=\Omega_s\nu \big[C^n(\mbox{\bf R}_n)+C^{n-1}(\mbox{\bf R}_n)-2C^{eq}_n\big].
238: \label{Eq:Vnorm}
239: \end{equation}
240: In this equation, $\Omega_s$ is the adatom area and we neglect adatom
241: diffusion along the step.
242:
243:
244:
245: \subsection{Nondimensional version of the governing equations}
246: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
247:
248: The terrace width $L_0$ provides a natural length scale for the problem.
249: Another possibity is the electrical length $\ell_E$ defined in Eq. (\ref{Eq:Xi}).
250: However, since $\ell_E$ is likely to diverge as the electrical field goes to zero, $L_0$
251: is prefered. Setting
252: \begin{equation}
253: x=\frac{X}{L_0}, \quad y=\frac{Y}{L_0}, \quad\frac{1}{\eta}=\frac{\ell_E}{L_0}, \quad c^n=\frac{C^n}{C_0},
254: \label{Eq:Adim1}
255: \end{equation}
256: where $C_0\Omega_s$ is the fraction of adsorption sites occupied by adatoms,
257: we get the nondimensional form of Eqs. (\ref{Eq:DiffDim}-\ref{Eq:Vnorm}).
258: The quasistatic diffusion equation reads
259: \begin{equation}
260: (\partial_{xx} +\partial_{yy} -\eta \partial_{x} )c^n=0,
261: \label{Eq:ADiff}\end{equation}
262: and the adatom flux is
263: \begin{equation}
264: \mbox{\bf j}^n =(\eta-\partial_x, -\partial_y)c^n.
265: \label{Eq:AFlux}
266: \end{equation}
267: In this equation, the nondimensional flux vector is defined as
268: \begin{equation}
269: \mbox{\bf j}^n = \frac{L_0}{D_s C_0}{\mbox{\bf J}^n},
270: \label{Eq:AFV}
271: \end{equation}
272: so that the physical time is rescaled by the characteristic time
273: \begin{equation}
274: t_0=\frac{1}{D_s C_0}.
275: \label{Eq:AT}
276: \end{equation}
277: Using both the time and space scale factors, we obtain the first,
278: \begin{equation}
279: \mbox{\bf j}^n(\mbox{\bf r}_n) \cdot \mbox{\bf u}_n=-\rho[c^n(\mbox{\bf r}_n)-c^{eq}_n],
280: \label{Eq:ACL1}
281: \end{equation}
282: and the second boundary condition,
283: \begin{equation}
284: \mbox{\bf j}^n (\mbox{\bf r}_{n+1}) \cdot \mbox{\bf u}_{n+1}=\rho[c^n(\mbox{\bf r}_{n+1})-c^{eq}_{n+1}].
285: \label{Eq:ACL2}
286: \end{equation}
287: The nondimensional number,
288: \begin{equation}
289: \rho=\frac{\nu L_0}{D_s},
290: \label{Eq:ARho}
291: \end{equation}
292: inversely proportional to the characteristic length $d=D_s/\nu$,
293: indicates which mechanism between diffusion and attachment governs
294: the time evolution of the steps. The normal velocity of step $n$ now
295: takes the form,
296: \begin{equation}
297: v_n=\sigma\rho\big[c^n(\mbox{\bf r}_n)+c^{n-1}(\mbox{\bf r}_n)-2c^{eq}_n\big],
298: \label{Eq:AVnorm}
299: \end{equation}
300: where
301: \begin{equation}
302: \sigma=\frac{\Omega_s}{L_0^2}
303: \label{Eq:ASigma}
304: \end{equation}
305: is the ratio of the two basic areas in the problem.
306:
307:
308:
309: \subsection{Equilibrium concentration}
310: %%%%%%%%%%%%%%%%%%%%%%%%%%%
311:
312: We now derive a detailed expression of the equilibrium concentration
313: $c^{eq}_n$ at step $n$. A quite general form is
314: \begin{equation}
315: c^{eq}_n=C^{eq}_n/C_0=\exp\Big(\frac{M}{k_BT}\Big)=1+\frac{M}{k_BT}+\dots
316: \label{Eq:ECeq}
317: \end{equation}
318: We will use the thermal energy $k_BT$ as the energy scale, and define
319: the nondimensional chemical potential as
320: \begin{equation}
321: \mu=\frac{M}{k_B T}.
322: \label{Eq:EAmu}
323: \end{equation}
324: Within the nondimensional description presented in previous section,
325: the position of step $n$ is represented by a function $y_n(x)$, and we define
326: the relative position as
327: \begin{equation}
328: \zeta_n(x)=y_n(x)-n.
329: \label{Eq:En}
330: \end{equation}
331: Following Paulin and coworkers, \cite{paulin01} we introduce a
332: nondimensional free energy functional for step $n$,
333: \begin{equation}
334: f_n=f_n^R+f_n^I.
335: \label{Eq:Ef}
336: \end{equation}
337: The first term is due to the step stiffness,
338: \begin{equation}
339: f_n^R=\beta \int_n ds.
340: \label{Eq:EfR}
341: \end{equation}
342: Here $\int_n ds$ is the integral of the curvilinear abscissa along the whole step $n$
343: (total step length) and $\beta=B (L_0/k_BT)$, where $B$ is the step-stiffness of
344: the material.
345: The second term sums up the step-step repulsive energies assumed
346: to vary as the inverse square distance,
347: \begin{equation}
348: f_n^I = \frac{\alpha}{2} \int_n \Big[\Big(\frac{1}{l_n^+}\Big)^2+\Big(\frac{1}{l_n^-}\Big)^2\Big]\ ds,
349: \label{Eq:EfI}
350: \end{equation}
351: where $\alpha=A/(k_BT L_0)$, and $A$ is the step interaction
352: coefficient. As shown in Fig. (\ref{Fig:lnpm}), the lengths
353: $l_n^+$ and $l_n^-$ are the shortest distances between steps
354: $(n,n+1)$ and steps $(n,n-1)$.
355: Thus the previous relation gives only an approximate value of the
356: total repulsion energy. Finally, the chemical potential is
357: obtained by a functional derivation of the free energy,
358: \begin{equation}
359: \mu=\sigma \Big(\frac{\delta f_n}{\delta \zeta_n}\Big).
360: \label{Eq:Emu}
361: \end{equation}
362: \begin{figure}
363: \includegraphics[width=0.5\textwidth]{Fig2} \caption{
364: \label{Fig:lnpm} Shortest distances between a given step and its
365: two closest neighbors, in the general case. The tangents to steps
366: $n-1$ and $n+1$ are drawn at two points having the same abscissa
367: $x$. }
368: \end{figure}
369:
370: %%%%%%%%%%%%%%%%
371: \section{Linear stability analysis}
372: %%%%%%%%%%%%%%%%%
373:
374: Repulsions between steps prevent them
375: from intersecting one-another and edge stiffness limits their curvature.
376: However, the possibility that the shapes of two consecutive steps are
377: weakly correlated in phase or amplitude remains open.
378: Since the general problem is quite difficult to solve in practice, we
379: limit the present study to the simple case of a synchronized
380: train of steps.
381:
382:
383: Starting with straight steps, separated by a unit distance, we introduce
384: an harmonic perturbation of amplitude $\epsilon\ll 1$, wave number $q$,
385: and phase-shift $\phi$,
386: \begin{equation}
387: \zeta_n(x)=\epsilon \exp(iqx+\omega t+in\phi).
388: \label{Eq:Lzeta}
389: \end{equation}
390: Looking for solutions of the nondimensional diffusion equation under the form
391: \begin{equation}
392: c^n(x,y)=1+c_1^n(y)\zeta_n(x),
393: \label{Eq:Lcn}
394: \end{equation}
395: we obtain
396: \begin{equation}
397: c_1^n(y)=A_1^n e^{ry}+B_1^n e^{-ry},
398: \label{Eq:LC1}
399: \end{equation}
400: with
401: \begin{equation}
402: r=\sqrt{q^2+i\eta q}.
403: \label{Eq:LR}
404: \end{equation}
405: To derive the dispersion relation, we first express
406: the chemical potential using Eqs. (\ref{Eq:Ef}-\ref{Eq:Emu}).
407: In practice, the step curvature is small, and
408: an accurate approximation of the step-step distances is given by
409: \begin{equation}
410: l_n^+ = \frac{1+\zeta_{n+1}-\zeta_n}{\sqrt{1+(\partial_x
411: \zeta_{n+1})^2}} \quad \mbox{and} \quad
412: l_n^-=\frac{1+\zeta_{n}-\zeta_{n-1}} {\sqrt{1+(\partial_x
413: \zeta_{n-1})^2}}
414: \label{Eq:elln}
415: \end{equation}
416: To the leading order in the perturbation amplitude $\epsilon$, we find
417: the chemical potential from Eqs. (\ref{Eq:Ef}-\ref{Eq:Lzeta},\ref{Eq:elln}),
418: \begin{equation}
419: \mu= \sigma \zeta_n g(q,\phi),
420: \end{equation}
421: where
422: \begin{equation}
423: g(q,\phi)= (\alpha + \beta) \ q^2 + 6 \alpha(1-\cos{\phi}).
424: \end{equation}
425: Introducing this result in Eqs. (\ref{Eq:ACL1},\ref{Eq:ACL2},\ref{Eq:AVnorm}),
426: the following dispersion relation is finally obtained,
427: \begin{equation}
428: \frac{\omega(q,\phi)}{2 \sigma r}= \frac{q \frac{\eta}{\rho}
429: \sin{\phi} + \sigma (\cos{\phi} - \cosh{r} - \frac{r}{\rho}
430: \sinh{r}) g(q,\phi) }{(1+\frac{r^2}{\rho^2}) \sinh{r} + 2
431: \frac{r}{\rho} \cosh r}. \label{Eq:Reldisp}
432: \end{equation}
433: This equation is similar to the dispersion relation derived in Ref. \cite{liu98}.
434: In the remaining of this section, we will neglect the step-step
435: interactions, $\alpha=0$, to avoid unnecessarily complicated
436: relations. Keeping the actual value of $\alpha$ would only
437: introduce small quantitative changes. The growth rate is defined as
438: $\Gamma(q,\phi)=\mbox{\rm Re}(\omega)$. A meandering instability is
439: thus expected for positive values of $\Gamma$. Experimentally, the
440: electrical field is a weak perturbation, so that, according to
441: Eq. (\ref{Eq:Adim1}), the parameter $\eta$ takes small values
442: ($\simeq10^{-8}-10^{-4}$). We thus expect the wave numbers of the
443: corresponding instabilities to verify $q\ll 1$. As we do not know
444: {\em a priori} the relative magnitude of $\eta$ and $q$, we first
445: use a general expansion in which both quantities are small and
446: considered equivalently. This leads to
447: \begin{eqnarray}
448: \Gamma &=& \frac{2}{2+\rho} \ \sigma \eta \ \sin{\phi} \ q
449: -\frac{2\rho}{2+\rho} \ \sigma^2 \beta \big(1-\cos{\phi} \big) \
450: q^2
451: \nonumber \\
452: &-&\frac{1}{3}\frac{\rho^2+6\rho+6}{\rho \big(2+\rho\big)^2}
453: \ \sigma \eta \ \sin{\phi} \ q^3 \nonumber \\
454: &-& \frac{1}{3} \frac{\rho^2+\big(\rho^2+6\rho+6
455: \big)\big(1+\cos{\phi} \big)}{\big(2+\rho \big)^2} \ \sigma^2
456: \beta \ q^4\nonumber \\
457: &+&\dots
458: \end{eqnarray}
459: Although the two last terms are of higher order, we nevertheless
460: keep them since the lowest order terms vanish when $\phi=0$. A first
461: necessary condition for a maximally unstable mode, $(\partial
462: \Gamma/\partial \phi)_q =0$, gives the following phase-shift:
463: \begin{equation}
464: \phi^*(q)=\tan^{-1}{\Big(\frac{\eta}{\sigma \rho \beta q}\Big)}.
465: \label{Eq:Phimax}
466: \end{equation}
467: \begin{figure}
468: \includegraphics[width=0.6\textwidth]{Fig3}
469: \caption{ \label{Fig:linstab} Plot of the growth rate
470: $\Gamma(q,\phi^*)=Re(\omega)$ obtained from the linear stability
471: analysis Eq. (\ref{Eq:Reldisp}) as a function of the perturbation
472: wave number $q$.}
473: \end{figure}
474: The corresponding growth rate, $\Gamma(q,\phi^*)$, is plotted as a
475: function of the wave number $q$ in Fig. (\ref{Fig:linstab}):
476: a meandering instability
477: arises as soon as the electrical field is nonzero. \cite{liu98}
478: For $q\gg \eta$, it is possible to find relations between $q$ and
479: $\eta$ in different ranges of wave numbers. The most unstable mode
480: $q=q_{max}$ is obtained by introducing the second condition
481: $(\partial \Gamma/ \partial q)_\phi =0$. Together with Eq.
482: (\ref{Eq:Phimax}), we get
483: \begin{equation}
484: q_{max}=\bigg(\frac{1}{2 \rho} \bigg)^{\frac{1}{2}}
485: \bigg(\frac{1}{2 + \rho} \bigg)^{\frac{1}{6}}
486: \bigg(\frac{\eta}{\sigma \beta} \bigg)^{\frac{2}{3}},
487: \label{Eq:Qmax}
488: \end{equation}
489: and the absolute maximum of the amplification factor is
490: \begin{equation}
491: \Gamma_{max}=\Gamma(q_{max},\phi^*)=\frac{\eta^2}{\beta \rho (2+\rho)}.
492: \end{equation}
493: %Thus, we will use the scaling relations
494: %\begin{equation}
495: %q\sim \eta^{\frac{1}{2}}, \quad\mbox{\rm for}\quad q\sim q_0,
496: %\label{Eq:Eta1}
497: %\end{equation}
498: %and
499: %\begin{equation}
500: %q\sim \eta^{\frac{2}{3}}, \quad\mbox{\rm for}\quad q\sim q_{max},
501: %\label{Eq:Eta2}
502: %\end{equation}
503: %which suggest that the range of unstable modes is quite large.
504: Finally, the marginal mode $q=q_0$ is deduced from
505: the condition $\Gamma(q,\phi^*)=0$:
506: \begin{equation}
507: q_0=\bigg(\frac{1}{\rho} \bigg)^{\frac{1}{4}} \bigg(\frac{1}{2 +
508: \rho} \bigg)^{\frac{1}{4}} \bigg(\frac{\eta}{\sigma \beta}
509: \bigg)^{\frac{1}{2}}.
510: \label{Eq:Qmarginal}
511: \end{equation}
512: The scalings of $q_{max}$ and $q_0$ with $\eta$
513: thus suggest that the range of unstable modes is quite large.
514: The limit of weak electrical fields ($\eta\ll 1$) is relevant
515: for the experimental work reported in Ref. \cite{degawa01}, in
516: which step meandering is observed.
517: In addition, the maximum growth rate being small, the
518: situation is favorable for a nonlinear analysis of the
519: meandering instability which is presented in the next section.
520:
521:
522: %%%%%%%%%%%%%%%%%%%
523: \section{Nonlinear analysis}
524: %%%%%%%%%%%%%%%%%%
525:
526: \subsection{Local coordinates}
527: %%%%%%%%%%%%%%%%%%
528: As illustrated in Fig. (\ref{Fig:local}), we consider the case of
529: steps which are all identical up to a translation in an oblique
530: direction $\tilde z$ rotated by an angle $\theta$ with respect to
531: axis $y$. The amplitudes of two successive steps have thus the
532: following property:
533: \begin{equation}
534: \zeta_{n+1}(x-\tan \theta)=\zeta_n(x).
535: \label{Eq:Zeta}
536: \end{equation}
537: In the linear regime defined by Eq. (\ref{Eq:Lzeta}), we have
538: \begin{equation}
539: \theta\simeq \tan^{-1}(\phi/q).
540: \label{Eq:Theta}
541: \end{equation}
542: Since we want to explore nonlinear dynamics, the amplitude of the meanders
543: may reach values of order unity for which $\zeta_n(x)$ is no more
544: a single-valued function in the original frame of reference $(x, y)$.
545: For this reason, our nonlinear model makes use of a non-orthogonal frame of reference
546: $(\tilde x, \tilde z)$, defined as
547: \begin{eqnarray}
548: \tilde{x}&=&x+y \tan{\theta}, \nonumber \\
549: \tilde{z}&=&\frac{y}{\cos \theta},
550: \end{eqnarray}
551: as shown in Fig. (\ref{Fig:local}).
552: With this change of coordinates, the step shape function becomes
553: \begin{equation}
554: \xi (\tilde x)=\frac{\zeta_n(x-n\tan \theta)}{\cos \theta},
555: \end{equation}
556: where the $n$ index can be omitted because all the steps are
557: identical in the new frame. We further define a local frame of
558: reference $(\chi, \psi)$ by moving the $\tilde x$ and $\tilde z$
559: axes along the step,
560: \begin{eqnarray}
561: \chi &=&\tilde{x}, \nonumber \\
562: \psi &=& \tilde{z} - \xi (\tilde x)= \tilde z - \xi (\chi).
563: \end{eqnarray}
564: In the local frame, the partial derivatives transform as
565: \begin{eqnarray}
566: \partial_x&=&\partial_\chi-(\partial_\chi \xi) \partial_\psi, \nonumber \\
567: \partial_y&=&\tan \theta \ \partial_\chi +\Big(\frac{1}{\cos \theta}
568: -\tan \theta \ \partial_\chi \xi \Big)\ \partial_\psi,
569: \end{eqnarray}
570: where $\partial_\chi \xi =\partial \xi/ \partial \chi$.
571: The second derivatives are derived from these expressions.
572: \begin{figure}
573: \includegraphics[width=0.5\textwidth]{Fig4}
574: \caption{ A set of steps identical up to a translation along the
575: $\tilde z$ axis. The local variables used in the nonlinear
576: analysis are $\tilde z=y/(\cos \theta)$ and $\tilde x =x+y \tan
577: \theta$. } \label{Fig:local}
578: \end{figure}
579:
580: \subsection{BCF equations for the local coordinates}
581: %%%%%%%%%%%%%%%%%%%%%%%%%%
582: Introducing the relations for the partial derivatives into Eq. (\ref{Eq:ADiff}), one gets
583: the quasi-static diffusion equation for the local coordinates $(\chi,\psi)$,
584: \begin{eqnarray}
585: 0&=&[\partial_{\chi\chi} + p^2 \partial_{\psi\psi} +
586: (\eta \cos^2 \theta\ \partial_\chi \xi - \partial_{\chi\chi} \xi) \partial_\psi \nonumber \\
587: &+& 2 (\sin \theta- \partial_\chi \xi) \partial_{\chi\psi} -
588: \eta \cos^2 \theta \ \partial_\chi] c(\chi,\psi),
589: \label{Eq:BCF0}
590: \end{eqnarray}
591: where
592: \begin{equation}
593: p(\chi) = \sqrt{(1- \sin \theta\ \partial_\chi \xi)^2 +(\cos \theta\ \partial_\chi \xi)^2}.
594: \label{Eq:BCFP}
595: \end{equation}
596: Note that the step index $n$ is purposely omitted because of
597: translational invariance.
598: It is easier to express the vectorial quantities in the base of the two
599: unit vectors of the initial frame $(x,y)$. The adatom flux reads
600: \begin{eqnarray}
601: \mathbf{j} &=&(j_x,j_y)=c(\chi,\psi) \Big( -\partial_\chi + (\partial_\chi \xi) \partial_\psi + \eta \ , \nonumber \\
602: &-&\tan \theta \ \partial_\chi - \frac{1}{\cos \theta}\partial_\psi + (\partial_\chi \xi) \tan \theta\
603: \partial_\psi \Big),
604: \label{Eq:BCFJ}
605: \end{eqnarray}
606: and the unit normal vector to the step,
607: \begin{equation}
608: \mathbf{u}=\frac{1}{p} \Big(- \cos \theta \ \partial_\chi \xi,
609: 1- \sin \theta\ \partial_\chi \xi \Big).
610: \label{Eq:unit}
611: \end{equation}
612: The two boundary conditions take on very simple forms,
613: \begin{eqnarray}
614: \mathbf{j} \cdot \mathbf{u} &=& - \rho(c - c^{eq})
615: \quad \mbox{at} \quad \psi=0, \nonumber \\
616: \mathbf{j} \cdot \mathbf{u} &=&+ \rho (c - c^{eq}) \quad
617: \mbox{at} \quad \psi=\frac{1}{\cos \theta}.
618: \label{Eq:BCFL}
619: \end{eqnarray}
620: The expression of the local curvature is needed to complete
621: these boundary conditions. We obtain
622: \begin{equation}
623: \kappa(\chi) = - \frac{\cos \theta}{p^3} \partial_{\chi\chi} \xi.
624: \end{equation}
625:
626: The normal velocity is deduced from Eq. (\ref{Eq:AVnorm}),
627: \begin{equation}
628: v(\chi)=\frac{\partial_t \xi}{p}=\sigma\rho\bigg[c(\chi,0)+c\Big(\chi,\frac{1}{\cos \theta}\Big)-2c^{eq}(\chi)\bigg],
629: \label{Eq:Vchi}
630: \end{equation}
631: where $c^{eq}(\chi)\simeq 1+\mu(\chi)$. An expression for the chemical potential is
632: obtained through the functional derivative of the free energy functional, Eqs. (\ref{Eq:Ef}-\ref{Eq:Emu}).
633: In the oblique frame of reference, we have now,
634: \begin{equation}
635: \mu=\frac{\sigma}{\cos \theta} \bigg(\frac{\delta f}{\delta \xi}\bigg),
636: \end{equation}
637: and the curvilinear length element is $ds=p \ d\chi$.
638: As illustrated in Fig. (\ref{Fig:lpm}) the shortest step-step distances are
639: defined in a slightly different way in this frame: the tangents to the
640: two adjacent steps are drawn at a given value of $\tilde x$. Adapting
641: Eq. (\ref{Eq:elln}) to this new definition, we obtain
642: \begin{equation}
643: l^+ = \frac{\frac{1}{\cos \theta}+\xi_+ -\xi}{p_+} \quad \mbox{and} \quad
644: l^- = \frac{\frac{1}{\cos \theta}+\xi -\xi_-}{p_-},
645: \label{Eq:ellpm}
646: \end{equation}
647: where
648: \begin{equation}
649: p_\pm = \sqrt{(1- \sin \theta\ \partial_{\chi} \xi_\pm)^2 + (\cos \theta\ \partial_{\chi} \xi_\pm)^2}.
650: \label{Eq:ppm}
651: \end{equation}
652: Note that it is necessary to keep the amplitudes $\xi_-$, $\xi$, and $\xi_+$
653: of three successive steps to perform the functional derivation. After derivation,
654: we set $\xi=\xi_\pm$, so that $l_+=l_-=l$, and,
655: \begin{equation}
656: \mu(\chi)=\sigma \kappa \bigg\{\beta + \frac{\alpha}{l^2 \cos^2 \theta}
657: \Big[ p^2+(\partial_\chi \xi-\sin \theta)^2\Big] \bigg\}.
658: \label{Eq:muchi}
659: \end{equation}
660: The chemical potential sums up the contributions of the step
661: stiffness and of the step-step interactions.
662:
663:
664: \begin{figure}
665: \includegraphics[width=0.5\textwidth]{Fig5}
666: \caption{ \label{Fig:lpm} Shortest distances between a given step
667: and its two closest neighbors, in the local frame. The case of
668: translational invariance along $\tilde z$ is represented here. The
669: tangents to the adjacent steps are drawn at two points having the
670: same abscissa $\tilde x$. }
671: \end{figure}
672:
673:
674: \subsection{Small parameter expansion}
675: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
676: The aim of this paragraph is to establish a nonlinear equation for
677: the time evolution of a step.
678:
679: \subsubsection{\label{Sec:C1} Scaled variables}
680: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
681: In the linear analysis presented above, we have shown that the wave
682: number $q_{0}\sim \eta^{1/2}$ is
683: small as compared to unity in the limit of a weak electrical field.
684: We thus introduce a small parameter $\epsilon\ll 1$, such as
685: \begin{equation}
686: \eta =\epsilon^2 \eta_2. \label{Eq:Eta2}
687: \end{equation}
688: and $\eta_2$ is of order unity. As a consequence, we define the
689: slow space variable
690: \begin{equation}
691: x=\epsilon \chi.
692: \end{equation}
693: Note that the slow $x$ variable used hereafter differs from the fast
694: $x$ variable discussed in section II. This should not introduce any
695: confusion, since only the new $x$ appears in the following. At the
696: marginal wave number $q=q_0$, the linear analysis results of Eqs. (\ref{Eq:Phimax},
697: \ref{Eq:Qmarginal}, \ref{Eq:Theta}) give the
698: following relation between the inclination angle $\theta$ and the
699: non-dimensional number $\rho$,
700: \begin{equation}
701: \theta\simeq\tan^{-1}{\bigg(\Big(\frac{2+\rho}{\rho}\Big)^{\frac{1}{2}}\bigg)}.
702: \end{equation}
703: Since $\rho$ can be
704: large or small depending on the parameters, $\theta$ can
705: take arbitrary values. The boundary conditions
706: given in Eq. (\ref{Eq:BCFL}) are applied at $\psi=0$ and
707: $\psi=1/\cos{\theta}$, so that the space variable $\psi$ is simply
708: equal to $z$. Accordingly, one defines the meander amplitude as $h(x) =
709: \xi(\chi)$ and the normal velocity as $\tilde{v}(x)=v(\chi)$.
710:
711: \subsubsection{Order by order expansion}
712: %%%%%%%%%%%%%%%%%%%%%%
713: The unknown variables, concentration and shape function are
714: expressed as power expansions of the scaling parameter $\epsilon$,
715: \begin{eqnarray}
716: c(x,z)&=&c_0(x,z) + \epsilon c_1(x,z) + \epsilon^2 c_2(x,z) +\dots \nonumber \\
717: h(x) &=&\epsilon^{-1}h_{-1}(x)+\epsilon^0 h_0(x)+\epsilon
718: h_1(x)+\dots \nonumber \\
719: \tilde{v}(x)&=&\epsilon^3 \tilde{v}_3(x)
720: \end{eqnarray}
721: Introducing this expression of $h(x)$ in Eq. (\ref{Eq:BCFP}), the
722: following development is found for $p$,
723: \begin{equation}
724: p(x)=p_0(x)+\epsilon p_1(x) +\epsilon^2 p_2(x)+ \dots,
725: \end{equation}
726: with
727: \begin{equation}
728: p_0(x)=\sqrt{(1- \sin \theta \ \partial_x h_{-1})^2+( \cos \theta
729: \ \partial_x h_{-1})^2 }.
730: \end{equation}
731: We obtain in a similar way the equilibrium concentration
732: \begin{equation}
733: c^{eq}(x)=1+\epsilon c_1^{eq}(x) + \epsilon^2 c_2^{eq}(x) + \dots,
734: \end{equation}
735: where
736: \begin{equation}
737: c_1^{eq}(x)=- \sigma \cos{\theta} \frac{\partial_{xx}
738: h_{-1}}{p_0^3} \Big[\beta+\alpha p_0^2 \Big(2
739: p_0^2-\cos^2{\theta}\Big)\Big].
740: \end{equation}
741: We now solve order by
742: order the nondimensional equations obtained by introducing the scaled
743: variables defined above into
744: Eqs.(\ref{Eq:BCF0}, \ref{Eq:BCFJ}, \ref{Eq:BCFL}, \ref{Eq:Vchi}). The results
745: obtained at order $i$ are used to derive the equations at order
746: $i+1$.
747: \vskip10pt \centerline{\bf order 0}
748: \noindent The diffusion equation reduces to
749: \begin{equation}
750: p_0^2 \ \partial_{zz} c_0 = 0.
751: \end{equation}
752: We look for solutions of the form
753: \begin{equation}
754: c_0(x,z)=a_0(x) z + b_0(x),
755: \end{equation}
756: which imposes
757: \begin{eqnarray}
758: p_0^2 \ a_0 &-& \rho p_0 \cos{\theta} \big[b_0-1\big] = 0,
759: \nonumber \\
760: p_0^2 \ a_0 &+& \rho p_0 \cos{\theta} \big[b_0-1\big] +\rho p_0 \
761: a_0 = 0,
762: \end{eqnarray}
763: for the boundary conditions, so that,
764: \begin{eqnarray}
765: a_0(x)&=&0, \nonumber \\
766: b_0(x)&=&1, \nonumber \\
767: c_0(x,z)&=&1.
768: \end{eqnarray}
769: At this order, the velocity is found to be zero.
770:
771: \vskip10pt \centerline{\bf order 1} \noindent Diffusion equation:
772: \begin{equation}
773: p_0^2 \ \partial_{zz} c_1 = 0.
774: \end{equation}
775: Solution:
776: \begin{equation}
777: c_1 (x,z)=a_1 (x) z + b_1 (x).
778: \end{equation}
779: Boundary conditions:
780: \begin{eqnarray}
781: p_0^2 \ a_1 &-& \rho p_0 \cos{\theta} \big[b_1-c_1^{eq}\big] = 0,
782: \nonumber \\
783: p_0^2 \ a_1 &+& \rho p_0 \cos{\theta} \big[b_1-c_1^{eq}\big] +
784: \rho p_0 \ a_1 = 0.
785: \end{eqnarray}
786: Solution:
787: \begin{eqnarray}
788: a_1(x)&=&0, \nonumber \\
789: b_1(x)&=&c_1^{eq}(x), \nonumber \\
790: c_1(x,z)&=&c_1^{eq}(x).
791: \end{eqnarray}
792: At this order, the velocity is found to be zero.
793:
794:
795: \vskip10pt \centerline{\bf order 2} \noindent Diffusion equation:
796: \begin{equation}
797: p_0^2 \ \partial_{zz} c_2=0.
798: \end{equation}
799: Solution:
800: \begin{equation}
801: c_2(x,z)=a_2(x)z+b_2(x).
802: \end{equation}
803: Boundary conditions:
804: \begin{eqnarray}
805: p_0^2 \ a_2 &-& \rho p_0 \cos{\theta} \big[b_2-c_2^{eq}\big] +
806: f_2=0,
807: \nonumber \\
808: p_0^2 \ a_2 &+& \rho p_0 \cos{\theta} \big[b_2-c_2^{eq}\big] + f_2
809: +\rho p_0 \ a_2=0,
810: \end{eqnarray}
811: with
812: \begin{equation}
813: f_2(x) = \eta_2 \cos^2{\theta} \ \partial_x h_{-1} +
814: (\sin{\theta}-\partial_x h_{-1}) \partial_x c_1^{eq}.
815: \end{equation}
816: Concentration:
817: \begin{eqnarray}
818: a_2(x)&=&-\frac{2}{p_0(x)} \ \frac{f_2(x)}
819: {2 p_0(x) + \rho}, \nonumber \\
820: b_2(x)&=&c_2^{eq}(x) - \frac{a_2 (x)}{2 \cos{\theta}}, \nonumber \\
821: c_2(x,z)&=&c_2^{eq}(x) + \Big(z-\frac{1}{2 \cos{\theta}}\Big) \
822: a_2 (x).
823: \end{eqnarray}
824: Zero normal velocity.
825:
826: \vskip10pt \centerline{\bf order 3} \noindent Diffusion equation:
827: \begin{equation}
828: p_0^2 \ \partial_{zz} c_3 - 2 p_0^2 \ d_3=0.
829: \end{equation}
830: with
831: \begin{equation}
832: d_3(x)=\frac{2 \big(\partial_x h_{-1} - \sin{\theta} \big)
833: \partial_x a_2 + a_2 \ \partial_{xx} h_{-1} - \partial_{xx} c_1^{eq}}{2 \ p_0^2}.
834: \end{equation}
835: Solution:
836: \begin{equation}
837: c_3(x,z)=d_3(x) z^2 + a_3(x)z +b_3(x).
838: \end{equation}
839: Boundary conditions:
840: \begin{eqnarray}
841: p_0^2 \ a_3 &-& \rho p_0 \cos{\theta} \big[b_3-c_3^{eq}\big] +
842: f_3=0,\\
843: p_0^2 \ a_3 &+& \rho p_0 \cos{\theta} \big[b_3-c_3^{eq}\big]
844: + f_3 + g_3 + \rho p_0 \ a_3=0,\nonumber
845: \end{eqnarray}
846: with
847: \begin{eqnarray}
848: f_3(x) &=& \frac{\partial_x h_{-1} - \sin{\theta}}{2 \cos{\theta}}
849: \ \partial_x a_2 - \partial_x h_{0} \ \partial_x c_1^{eq}\nonumber \\
850: &+& \Big[ 2 \Big( \partial_x h_{-1}-\sin{\theta} \Big) \
851: \partial_x h_0 - p_0 p_1 \Big] \ a_2 \nonumber \\
852: &+& \Big[ \Big( c_1^{eq}-\frac{p_1}{p_0} \Big)
853: \partial_x h_{-1} + \partial_x h_0 \Big] \ \eta_2 \cos^2{\theta} \nonumber \\
854: &+& \Big( \partial_x h_{-1} - \sin{\theta}\Big)
855: \Big(\frac{p_1}{p_0} \ \partial_x c_1^{eq} -
856: \partial_x c_2^{eq} \Big),
857: \end{eqnarray}
858: and
859: \begin{equation}
860: g_3(x)=\frac{p_0(2 p_0 + \rho)}{\cos{\theta}} \ d_3-
861: \frac{\partial_xh_{-1} - \sin{\theta}}{\cos{\theta}} \ \partial_x
862: a_2.
863: \end{equation}
864: Concentration:
865: \begin{eqnarray}
866: a_3(x)&=&- \frac{1}
867: {p_0} \ \frac{g_3+2 f_3}{2 p_0 + \rho} \nonumber \\
868: b_3(x)&=&c_3^{eq}+ \frac{\rho f_3 - p_0 g_3}
869: {\rho p_0 \cos{\theta} \ \big(2 p_0 + \rho\big)}\nonumber \\
870: c_3(x,z)&=&d_3(x) z^2 + a_3(x)z +b_3(x)
871: \end{eqnarray}
872: The normal velocity is nonzero for the first time at this order.
873: Its expression is derived by using Eq. (\ref{Eq:Vchi}) together
874: with the scaling relations of section \ref{Sec:C1},
875: \begin{eqnarray}
876: \tilde{v}_3(x)&=&\frac{\sigma}{p_0 \cos^2{\theta}} \partial_x
877: \bigg[\frac{2 \cos^2{\theta}+ \rho p_0}{p_0(\rho+2p_0)} \partial_x
878: c_1^{eq}\nonumber
879: \\ &-&2\eta_2 \cos^2{\theta}\
880: \frac{\sin{\theta}-\partial_x h_{-1}}{p_0(\rho+2p_0)} \ \partial_x
881: h_{-1} \bigg]. \label{Eq:V3}
882: \end{eqnarray}
883: %%%%%%%%%%%%%%%%%%%
884: \subsection{Amplitude equation}
885: We finally obtain the following amplitude equation using Eqs.
886: (\ref{Eq:Vchi},\ref{Eq:V3}):
887: \begin{eqnarray}
888: \partial_t H
889: &=& \frac{\sigma}{\cos^2{\theta}}
890: \partial_x \bigg[\frac{2 \cos^2{\theta}+ \rho p_0}{p_0(\rho+2p_0)} \partial_x c_1^{eq}
891: \nonumber \\
892: &-&2\eta_2 \cos^2{\theta}\ \frac{\sin{\theta}-\partial_x
893: H}{p_0(\rho+2p_0)} \partial_x H \bigg], \label{Eq:Amplitude}
894: \end{eqnarray}
895: where
896: \begin{equation}
897: p_0(x)=\sqrt{(1-\sin \theta \ \partial_x H)^2+( \cos \theta \
898: \partial_x H)^2 },
899: \end{equation}
900: and
901: \begin{equation}
902: c_1^{eq}(x)=- \sigma \cos{\theta} \ \frac{\partial_{xx} H}{p_0^3}
903: \Big[\beta+\alpha p_0^2 \Big(2 p_0^2-\cos^2{\theta}\Big)\Big].
904: \end{equation}
905: Here $H(x)=h_{-1}(x)$ and the time is rescaled such as $\epsilon^4
906: t\rightarrow t$.
907: This amplitude equation is the central result of our study.
908: As expected, this equation ensures mass
909: conservation since its right hand side is a derivative of a mass
910: current.
911:
912:
913:
914:
915:
916: %%%%%%%%%%%%%%%%%%%%%%%%
917: \section{Numerical simulations and discussion}
918: %%%%%%%%%%%%%%%%%%%%%%%%
919: The time evolution of vicinal surfaces is obtained by integrating
920: numerically Eq. (\ref{Eq:Amplitude}). While the simulations are performed
921: in the oblique frame $(x,z)$, the system is represented in the
922: laboratory orthogonal frame $(x,y)$. Solving this stiff partial
923: differential equation necessitates the use of an adaptive time step.
924: A single step with periodic boundary conditions is simulated in
925: practice. The whole vicinal surface is obtained by reproducing this
926: step periodically along the $\tilde z$ direction. The elastic
927: interactions included in our model are not only justified from a
928: purely physical point of view but are also a necessary ingredient in
929: realistic numerical simulations. Indeed, test simulations performed
930: without elastic interactions systematically resulted in step
931: crossings at late times.
932:
933: \begin{figure}
934: \includegraphics[width=0.6\textwidth]{Fig6}
935: \caption{ \label{Fig:onestep} Numerical simulation of Eq.
936: (\ref{Eq:Amplitude}). Time evolution of a single step for
937: $\rho=0.001$ and $\theta=0.8$. The step is systematically shifted
938: in time (given by the lower axis). The electrical field is applied
939: in the positive $x$ direction. }
940: \end{figure}
941: \begin{figure}
942: \includegraphics[width=0.6\textwidth]{Fig7}
943: \caption{ \label{Fig:Onestep} Same as in Fig. (\ref{Fig:onestep})
944: for $\rho=20$ and a larger system width. }
945: \end{figure}
946: We first compare the dynamics of one step in two physical regimes defined
947: by the values of the nondimensional number $\rho=\frac{\nu L_0}{D_s}$.
948: For $\rho>1$, the system dynamics is diffusion-limited, while
949: it is attachment-limited for $\rho<1$. All the parameters ($\alpha$, $\beta$, $\eta_2$,
950: $\sigma$) entering Eq. (\ref{Eq:Amplitude}) are set to unity here, and
951: Figs. (\ref{Fig:onestep}) and (\ref{Fig:Onestep}) show the time
952: evolution of a single step for $\rho=0.001$ and $\rho=20$,
953: respectively. At short times, the steps are rather similar in shape
954: for both values of $\rho$. Calculating the wave length emerging at
955: short times, we find that it increases with $\rho$ as predicted by
956: the linear stability analysis. Alternatively, the growth rate
957: $\Gamma$ is found to decrease with $\rho$. At late times, after
958: coarsening has set in, the step shapes differ strongly: a
959: single-valued function is found in the laboratory frame for
960: $\rho=0.001$, while long overhangs are visible for $\rho=20$. In
961: both cases, the electrical field triggers local facetting of the
962: steps which look like asymmetrical saw-teeth. Ultimately,
963: the meander amplitude saturates to a finite value in a finite size system.
964: \begin{figure}
965: \centerline{\includegraphics[width=0.48\textwidth]{Fig8a}
966: \includegraphics[width=0.48\textwidth]{Fig8b}}
967: \centerline{\includegraphics[width=0.48\textwidth]{Fig8c}
968: \includegraphics[width=0.48\textwidth]{Fig8d}}
969: \caption{ \label{Fig:vicinale} Top view of a vicinal surface
970: computed at different times for the same parameters as in Fig.
971: (\ref{Fig:onestep}): a) $t=230$, b) $t=1200$, c) $t=8300$, d)
972: $t=1.75\times 10^5$. The step down direction is rigthwards while
973: the electrical field direction is downwards.}
974: \end{figure}
975:
976: \begin{figure}
977: \centerline{\includegraphics[width=0.47\textwidth]{Fig9a}
978: \includegraphics[width=0.48\textwidth]{Fig9b}}
979: \centerline{\includegraphics[width=0.48\textwidth]{Fig9c}
980: \includegraphics[width=0.48\textwidth]{Fig9d}}
981: \caption{ \label{Fig:Vicinale} Top view of a vicinal surface
982: computed at different times for the same parameters as in Fig.
983: (\ref{Fig:Onestep}): a) $t=6.4\times 10^3$, b) $t=1.6\times 10^5$,
984: c) $t=8\times 10^5$, d) $t=3\times 10^6$.
985: The step down direction is rigthwards while
986: the electrical field direction is downwards.}
987: \end{figure}
988:
989: The time evolution of two vicinal surfaces is displayed in
990: Figs.(\ref{Fig:vicinale}) and (\ref{Fig:Vicinale}). Dark regions
991: correspond to a high step density in which the electrical field is
992: essentially oriented in the step-down direction, while it is mainly
993: oriented in the step-up direction in the low step density regions.
994: This result is consistent with the well-known step bunching observed
995: for Si (111) when the heating current is applied perpendicular to
996: the steps, in the step-down direction \cite{latyshev89}.
997:
998:
999: According to the nature of the material, the surface orientation and
1000: the temperature range, physical parameters such as the diffusion
1001: coefficient may vary a lot. In addition, they are not always known
1002: with a great accuracy. For example, for a Si$(111)$ surface, four
1003: acceptable sets of physical parameters are given in table I of ref.
1004: \cite{liu98bis}, of which set B seems particularly consistent with
1005: the experimental observations. For this particular set of physical
1006: parameters, Eq. (\ref{Eq:ARho}) gives
1007: $d=L_0/\rho=5\times10^{-7}\mbox{\rm m}$. A miscut angle of one
1008: degree, then results in $\rho\simeq 0.03$, thus
1009: attachment/detachment-limited dynamics. Note that with the parameter
1010: sets A, C, or D, and/or a different miscut angle, $\rho$ may vary in
1011: wide range, both below and above one. Our model is valid in both
1012: cases and it predicts rather different step shapes at long times, as
1013: just discussed. Experimental observation of vicinal
1014: surfaces under an electrical field parallel to the initial steps
1015: could possibly give an indication on the magnitude of the
1016: nondimensional number $\rho$ which governs the system dynamics.
1017:
1018: %%%%%%%%%%%%%%%%%%%%%%%%
1019: \section{Conclusion and perspectives}
1020: %%%%%%%%%%%%%%%%%%%%%%%%
1021:
1022: In summary, we have studied the meandering instability
1023: induced by a constant electrical field
1024: initially parallel to a train of straight steps.
1025: The time evolution of the meanders is described
1026: by a nonlinear amplitude equation which we
1027: have derived through an asymptotic expansion.
1028: Numerical simulations have been performed
1029: both in the attachment/detachment-limited ($\rho\ll 1$)
1030: and the diffusion-limited ($\rho\gg 1$) regimes. At large times, overhangs
1031: are observed in the latter case only.
1032:
1033:
1034: It is very instructive to compare our results with an experimental
1035: study of step meandering on Si (111) vicinal surfaces, in which
1036: the orientation of the electrical field $E$ is taken different
1037: from the step-down direction \cite{degawa01}. When $E$ is set
1038: parallel to the steps, as in the present study, a similar step
1039: meandering effect is observed but the steps bend in the opposite
1040: direction as compared to our model. This apparent contradiction is
1041: in fact not unexpected because the experiments are performed at
1042: $T=1100^\circ \mbox{\rm C}$. Indeed, in this intermediate range
1043: of temperature $(1000^\circ \mbox{\rm C}-1180^\circ \mbox{\rm
1044: C})$, the steps have been argued to become transparent to the
1045: diffusing adatoms \cite{degawa01}. The underlying physics is thus
1046: expected to differ from the one introduced in our model
1047: (impermeable steps) and an opposite direction of bending is not
1048: contradictory. In the light of this discussion, new experiments
1049: performed at temperatures slightly higher than $T=1180^\circ
1050: \mbox{\rm C}$ or slightly lower than $T=1000^\circ \mbox{\rm C}$
1051: would be desirable to test our model.
1052:
1053: In the present model, consecutive steps are
1054: assumed identical up to a given phase-shift. Removing
1055: this phase constraint would allow a realistic description of
1056: experiments on a large scale. However, this can
1057: hardly be envisaged on the basis of the present method and a quite
1058: different point of view should be considered, such as a continuous
1059: limit approach. In addition, it would
1060: be helpful to include the step transparency in order to
1061: compare the resulting model to the experiments in the
1062: intermediate range of temperatures.
1063:
1064:
1065: %%%%%%%%%%%%
1066: \acknowledgments
1067: %%%%%%%%%%%%%
1068: It is a pleasure to acknowledge F. Leroy, J. J. M\'etois, C. Misbah, P. M\H uller, and A. Verga for fruitful discussions.
1069:
1070: %%%%%%%%%%%%%%
1071: \bibliography{DDF}
1072:
1073: %%%%%%%%%%%%%%
1074:
1075: %\cleardoublepage
1076:
1077:
1078: \end{document}
1079: