1: \documentstyle[aps,prl,epsfig,amsmath,multicol]{revtex}
2:
3: \topmargin=-0.4in
4:
5: \newcommand{\be}{\begin{eqnarray}}
6: \newcommand{\ee}{\end{eqnarray}}
7:
8: \begin{document}
9:
10: \title{Defect Structures in the Growth Kinetics of the
11: Swift-Hohenberg Model}
12: \author{Hai Qian and Gene F. Mazenko}
13: \address{James Franck Institute and Department of
14: Physics, University of Chicago, Chicago, Illinois 60637}
15: \date{Oct. 18, 2002}
16: \maketitle
17: \begin{abstract}
18: The growth of striped order resulting from a quench of
19: the two-dimensional Swift-Hohenberg model is studied in the
20: regime of a small control parameter and quenches to zero
21: temperature. We introduce an algorithm for finding and identifying the disordering
22: defects (dislocations, disclinations and grain boundaries) at a given
23: time. We can track
24: their trajectories separately.
25: We find that the coarsening of the defects and lowering of the effective
26: free energy in the system are governed by a growth law $L(t)\approx t^{x}$
27: with an exponent $x$ near $1/3$. We obtain scaling for the correlations
28: of the nematic order parameter with the same growth law. The scaling for the
29: order parameter structure factor is governed, as found by others, by
30: a growth law with an exponent smaller than $x$ and near to $1/4$. By comparing two systems with
31: different sizes, we clarify the finite size effect. We find that the
32: system has a very low density of disclinations compared to that for
33: dislocations and fraction of points in grain boundaries. We also measure
34: the speed distributions of the defects at different times and find that
35: they all have power-law tails and the average speed decreases as a power law.
36: \end{abstract}
37: \draft
38: \pacs{PACS numbers: 05.70.Ln, 64.60.Cn, 64.75.+g, 98.80.Cq}
39:
40: \begin{multicols}{2}
41:
42: \section{Introduction}
43:
44: What are the defects which control the long-time
45: ordering of systems growing a striped pattern?
46: This question arises in a variety of physical
47: contexts\cite{BN}. Here we are motivated by
48: the recent experiments\cite{Harrison,H2002}
49: investigating the ordering of a two dimensional
50: diblock copolymer system.
51: The system studied offers a physical realization of
52: the ordering in an isotropic two-dimensional smectic material.
53: In these experiments they
54: found that the late-time ordering
55: satisfies scaling with a growth law $L\approx t^{x}$ with $x=1/4$
56: and the final stages of ordering are governed by the
57: annihilation of sets of disclination quadrapoles.
58: In this paper we address the question:
59: Is the ordering in this physical system described by the
60: Swift-Hohenberg (SH) model\cite{SH}, the simplest model
61: one can construct to govern the ordering in
62: stripe forming systems?
63:
64: We investigate the growth kinetics of the Swift-Hohenberg
65: model for a small control parameter ($\epsilon=0.1$)
66: in two dimensions and quenches to zero temperature.
67: It is this regime which appears most likely to correspond
68: to the experimental situation. In large $\epsilon$ regime the system
69: evolves to a glassy state.
70: We focus primarily on the
71: defect structures generated in the ordering of the
72: system. In the most naive picture
73: of this ordering process one can think in terms of an
74: initial local layering,
75: as in a smectic, in some direction. This ordering can be
76: disrupted by point defects: dislocations and disclinations.
77: This suggests a coarsening picture with annihilating point defects
78: similar to the case of the XY model\cite{BRAY} and a
79: growth law with exponent $x=1/2$. This simple picture is not seen
80: in simulations.
81: We find, in agreement with the numerical results
82: of Hou \textit{et al.} \cite{HSG}
83: and Boyer and Vi\~nals \cite{BVdc},
84: that the
85: defect structures for the SH model are dominated by grain boundaries which
86: persist for long times. Unlike the case of an XY model, the ordering is not
87: dominated by annihilation of isolated point defects. These are
88: observed but are not the dominant structures.
89:
90: We find numerically,
91: at late times after finite size effects enter, that the system becomes
92: anisotropic, and the grain boundaries shrink. In this case
93: one sees a cross-over
94: to an effective growth exponent $x=1/2$.
95:
96: We give below a detailed numerical study of the
97: statistical properties of the defects disrupting
98: striped pattern formation
99: in the SH model.
100: In order to carefully discuss the defects we need a reliable
101: filter for finding them. We present an algorithm which effectively
102: locates defects and grain boundaries for any control parameter $\epsilon$.
103: We can distinguish between
104: grain boundaries and other defects, and track their trajectories separately.
105: We compare this method to the other approaches used in earlier
106: work in appendix A.
107:
108: There are a number of ways of characterizing the degree of ordering in
109: these systems: (i) Counting the number and size of defects and their
110: evolution with time. (ii) Monitoring the lowering of the average
111: effective driving free energy as a function of time. (iii) Evaluation
112: of the nematic order parameter correlation function and its associated
113: scaling behavior. (iv) Evaluation
114: of the order parameter structure factor and its associated
115: scaling behavior. We find that (i), (ii), and (iii) can all
116: be characterized by a single growth law with the exponent near $1/3$, while
117: the order parameter scaling, as found by others, is characterized
118: by a growth law with the exponent near $1/5$.
119:
120: \section{Swift-Hohenberg Model}
121:
122: The Swift-Hohenberg model for a scalar order parameter, $\psi$,
123: is specified by the equation of motion
124: \be
125: \frac{\partial \psi}{\partial t}
126: =\epsilon \psi -\psi^{3}-\left(q_{0}^{2}+\nabla^{2}\right)^{2}\psi+\zeta
127: \ ,
128: \ee
129: where $\epsilon$ is a positive control parameter,
130: $q_{0}$ is the magnitude of an ordering wavenumber and $\zeta$ is the
131: Gaussian noise satisfying $\langle\zeta({\bf r},t)\zeta({\bf
132: r}',t')\rangle=2\Gamma\delta({\bf r}-{\bf r}')\delta(t-t')$, where the
133: noise strength $\Gamma$ is proportional to the final temperature
134: governing the system after a quench. We
135: will focus here on quenches to zero temperature where we can set
136: $\Gamma$ and the noise $\zeta$ to zero.
137: We are interested in the growth kinetics problem where we prepare
138: this system initially in a completely disordered state. We then allow
139: the system to evolve forward in time to form a striped pattern. For
140: example one
141: could choose
142: \be
143: \langle \psi ({\bf x},t_{0})\psi ({\bf y},t_{0})\rangle
144: =\Psi_{0}^{2}\,\delta\left({\bf x}- {\bf y}\right) \ ,
145: \ee
146: where $\Psi_0^2$ is a constant. However the precise form of the initial
147: conditions is not important \cite{BRAY}.
148:
149: This model can be formulated as a Langevin equation driven
150: by an effective Hamiltonian:
151: \be
152: {\cal H}_{E}=\int d^{2}x\left\{- \frac{\epsilon}{2}\psi^{2}
153: +\frac{1}{4}\psi^{4}
154: +\frac{1}{2}\left[\left(q_{0}^{2}+\nabla^{2}\right)\psi\right]^{2}\right\}
155: \label{eq:3} \ .
156: \ee
157: If we introduce
158: \be
159: E(t)=\langle {\cal H}_{E}\rangle_{t}\ ,
160: \ee
161: where the average is over an ensemble of initial conditions,
162: then $E(t)$ is lowered as
163: the system orders in a striped pattern
164: with wavenumber $q_{0}$.
165:
166: Eventually
167: the system approaches an ordered
168: state described approximately by the single-mode approximation \cite{PM79}
169: where, assuming layering along the z-direction,
170: \be
171: \psi_0 = A_0\, \cos q_{0}z \ .
172: \ee
173: If we put this ansatz into Eq.(\ref{eq:3}), assume that the system is an
174: integral
175: number of wavelengths in the z-direction, and minimize with respect to the
176: amplitude $A_0$, we obtain the results,
177: \begin{eqnarray}\displaystyle
178: A_{0}^{2}&=&\frac{4\epsilon}{3} \ ,\\
179: \langle \,\psi_0^2 \,\rangle&=& \frac{q_0}{2\pi}\int_0^{2\pi/q_0}(A_0\, \cos
180: q_0z)^2\,dz = \frac{2 \epsilon}{3}\ ,
181: \end{eqnarray}
182: and
183: \be
184: E_{eq}=-\frac{\epsilon^{2}}{6}S\ ,
185: \ee
186: where $S$ is the area of the system. Pomeau and Manneville\cite{PM79}
187: have shown
188: that this is a very good approximation for the ``ground'' state even for
189: moderately large values of $\epsilon$. In the growth kinetics context
190: the approach to
191: equilibrium is monitored by
192: \be
193: \Delta E(t)\equiv E(t)-E_{eq}\propto L_{E}^{-1}(t) \ ,
194: \ee
195: and
196: \be
197: \Delta \psi^{2}(t)\equiv
198: \langle\,\psi_0^2\,\rangle-\langle \psi^{2}\rangle_{t}\propto
199: L_{\psi}^{-1}(t) \ ,
200: \ee
201: where $L_E(t) \propto L_{\psi}(t)$ \cite{NT}.
202:
203: Another measure of the ordering in the system is given by
204: considering the director field
205: \be
206: \hat{n}({\bf x})=\frac{\nabla \psi ({\bf x})}{|\nabla \psi ({\bf x})|}\ ,
207: \ee
208: and the associated nematic order parameter
209: \be
210: Q_{\alpha\beta}=Q_0\,\left[\hat{n}_{\alpha}\hat{n}_{\beta}-\frac{1}{2}\,\delta_{\alpha\beta}\right]
211: \ .
212: \label{eq:12}
213: \ee
214: In two dimensions, however, all of the information in this
215: order parameter is contained in the quantity $\cos 2\theta$
216: where $\hat{n}=(\cos \theta , \sin \theta)$. It is
217: easy to show, for example, that
218: \begin{eqnarray}
219: &&\,C_{nn}({\bf x},{\bf y},t)
220: \equiv 2\langle \,\mathrm{Tr}\, Q({\bf x},t)Q({\bf y},t)\,\rangle_{t}
221: \nonumber \\
222: &&=
223: \left\langle\, \cos\left[ (\varphi ({\bf x},t)-\varphi ({\bf
224: y},t)\right]\,\right\rangle_{t} \ .
225: \end{eqnarray}
226: where
227: \be
228: \varphi ({\bf x},t)=2\theta ({\bf x},t)
229: ~~~.
230: \ee
231: If we define
232: \be
233: \hat{B}_{x}=\hat{n}_{x}^{2}-\hat{n}_{y}^{2}
234: \label{eq:15}
235: \ee
236: \be
237: \hat{B}_{y}=2\hat{n}_{x}\hat{n}_{y}
238: \label{eq:16}
239: \ee
240: then
241: \be
242: C_{nn}({\bf x},{\bf y},t)=\langle \hat{B}({\bf x},t)
243: \cdot \hat{B}({\bf y},t)\rangle_{t}
244: ~~~.
245: \ee
246:
247: The nematic order parameter correlation function, $C_{nn}$,
248: was shown by Christensen and Bray \cite{CB} to obey scaling
249: in the conventional form
250: \be
251: C_{nn}({\bf r},t)=F\left(r/L_{n}(t)\right) \ ,
252: \ee
253: where ${\bf r}={\bf x}-{\bf y}$.
254: Elder, Vi\~nals and Grant \cite{EVG} showed that
255: the scaling of the order parameter structure factor
256: \be
257: S(k,t)=\langle \,\left|\psi_{{\bf k}}(t)\right|^{2}\rangle
258: =L_{s}(t)F_1((k-q_{0})L_{s}(t)) \ ,
259: \label{eq:19}
260: \ee
261: differs from that observed
262: in ordering system without stripes: $S(k,t)=L^{2}(t)F_2(kL(t))$.
263:
264: \section{Review of Previous Work}
265:
266: The early work on this problem focused on
267: establishing the final equilibrium state reached after a
268: quench.
269: This is a
270: two dimensional system and by forming stripes one has a broken
271: continuous symmetry. The behavior of the system at non-zero
272: temperatures, as for the two dimensional X-Y model, requires, as pointed
273: out by Toner and Nelson \cite{TN81} a treatment of both long wavelength
274: fluctuations in the layers and free defects. Above a Kosterlitz-Thouless
275: type transition one has an isotropic phase while below this transition
276: one has a phase
277: with persistent orientational order.
278:
279: In an early paper,
280: Elder, Vi\~nals, and Grant\cite{EVG}
281: carried out a numerical analysis leading to the scaling
282: solution given by Eq.(19).
283: Working with fixed $\epsilon=0.25$ they looked at the system's
284: ordering as a function
285: of noise strength $\Gamma$. They found a qualitative
286: difference between low noise and high noise. For the large noise case
287: they found a rapid
288: (exponential) relaxation to the asymptotic stationary state and a
289: power-law approach for the lower noise case.
290: Their results are in agreement with the picture due to
291: Toner and Nelson that one has a
292: transition to an isotropic state for large enough noise. There is no real
293: ordering in the isotropic state
294: and this is why there is exponential decay to the equilibrium state.
295: In the ordered state one has scaling and a power-law growth law which,
296: for small noise, they found to have an exponent $x_s=1/4$. They found
297: a smaller exponent $x_s=1/5$
298: at low temperatures, but they had less statistics and there
299: appeared to be "difficulty removing defects". They argued for
300: a late time cross over to the {\it expected} $x=1/2$ but they did not
301: see this.
302:
303: Cross and Meiron \cite{CM} also studied the SH model
304: numerically in the absence of noise.
305: They found a $x_s=1/4$ for $\epsilon =0.25$. The dynamics appear to
306: freeze for higher $\epsilon$.
307: They looked at the defect structure but in a qualitative way noting
308: the existence of domain walls rather than a set of isolated point
309: defects.
310: The theoretical discussion in their paper is based on the phase-field
311: approximation
312: \be
313: \frac{\partial\phi}{\partial t}=\left(D_{\parallel}\nabla^{2}_{\parallel}
314: +D_{\perp}\nabla^{2}_{\perp}\right)\phi \ ,
315: \ee
316: which from the most naive point of view suggests a growth law with exponent
317: $x=1/2$. They discuss some selection mechanisms which could lead
318: $D_\parallel$ and $D_\perp$ to adjust themselves to zero and reduce $x$
319: to $1/3$ or $1/4$. They concluded that they did ``not have a good theoretical
320: understanding of these results''
321: and suggested that the defects in the problems should be
322: treated explicitly.
323:
324: Hou, Sasa, and Goldenfeld (HSG) \cite{HSG} confirmed previous
325: numerical results which showed
326: for $\epsilon =0.25$, $x_s=1/5$ with zero noise and $x_s=1/4$ with nonzero
327: noise as obtained from the structure factor scaling. They went further
328: and used a simple method to identify domain walls and measure their lengths
329: (more about this below). They measured excess energy, $\Delta E(t)$,
330: and the domain wall length and found that they show the same scaling
331: exponents $1/4$ at zero noise and $0.3$ at non-zero noise. The energy
332: does go to the lowest order in $\epsilon$ value of $-\epsilon^2/6$ in
333: the noiseless limit. They find
334: ``defects are indeed the driving force behind the coarsening process due
335: to its dominant contribution to the excess energy.'' They suggest that
336: the phase field approach gives the wrong exponent because it does not
337: include the effects of defects.
338: For larger $\epsilon$ (=0.75) they found much slower logarithmic
339: growth. The system seems to become glassy.
340:
341: Christensen and Bray \cite{CB} also carried out numerical work
342: on the SH model for $\epsilon =0.25$
343: and found $x_s=1/5$ for zero noise and $x_s=1/4$ for nonzero noise.
344: From scaling of the {\it director} correlation function they find
345: exponents are $0.25$ and $0.30$ for zero and nonzero noise.
346: They suggest that there is a cross over to $x=1/2$ at very long times.
347: The theory they developed does not include defects.
348:
349: Boyer and Vi\~nals \cite{BVdc} point out
350: "Near the bifurcation threshold, the evolution of disordered configurations is
351: dominated by grain boundary motion through a background of largely immobile
352: curved
353: stripes". They find for small $\epsilon$ an exponent $x=1/3$ which they
354: interpret as arising from a
355: law of grain boundary motion \cite{BVgm}.
356: Elsewhere \cite{BVgbp} they also point out for larger
357: values of $\epsilon$ the dynamics cross over to a frozen state
358: with quenches to zero temperature. This glassy behavior is associated
359: with grain boundary pinning.
360:
361: \section{Numerical Results for SH Model}
362:
363: We present here our numerical results for the SH equation. We follow
364: the numerical prescriptions of Bray and Christensen \cite{CB}.
365: We use the finite difference
366: scheme on two dimensional lattice of sizes $256\times 256$ and
367: $512\times 512$ with periodic boundary conditions. We set
368: $\epsilon=0.1$, $\Delta r = \pi/4$ and $\Delta t = 0.03$. We replace
369: $\partial_t \psi({\bf r},t)$ by
370: $\left(\psi_{ij}^{n+1}-\psi_{ij}^{n}\right)/\Delta t$, and $\nabla^2
371: \psi({\bf r},t)$ by
372: \begin{equation}
373: \nabla^2\psi_{ij}=\frac{1}{(\Delta
374: x)^2}\left[\frac{2}{3}\sum_{NN}+\frac{1}{6}\sum_{NNN}-\frac{10}{3}\right]\psi_{
375: ij}
376: \end{equation}
377: where $NN$ and $NNN$ mean the nearest neighbors and next-nearest
378: neighbors respectively.
379: By choosing the proper scale of time, space and the order parameter, we
380: can set $q_0=1$. The systems have eight grid points per wavelength. We used uniformly distributed random initial conditions.
381:
382: For the smaller $256\times 256$ systems we were able to follow the ordering
383: process to very late stages. Some of the independent trials proceed
384: to a final state where we have a set of well aligned layers.
385:
386: In Fig. 1 we plot $\Delta E(t)$ and $\Delta \psi^{2}(t)$ for an
387: ensemble of runs on a
388: $256\times 256$ lattice. We note that there are two regimes
389: where $L_{E}$ defined by Eq.(9) is described by different
390: exponents. For $t_{s}< t <t_{c}$ ($t_s \approx 300$ and $t_c \approx 9000$)
391: we find $x_{E} \approx 0.3$, while for
392: $t > t_{c}$ we find $x_{E}\approx 0.5$. The cross over at
393: $t \approx t_{c}$ appears to be due to the finite size effects, as we
394: discuss below.
395: For $t > t_{c}$ the system is effectively anisotropic and we find an
396: effective exponent
397: $x_E$ near to $1/2$.
398:
399: \begin{figure} % Fig. 1
400: \begin{center}
401: \includegraphics[scale=.43]{fig/fig_1.eps}
402: \end{center}
403: \caption{\footnotesize $\Delta E(t)$ and $\Delta \psi^2(t)$ for a
404: $256\times 256$
405: system. Straight lines are used to fit different parts. Averaged over $40$ trials.}
406: \end{figure}
407:
408: In Fig. 2 we plot $\Delta E(t)$ and $\Delta \psi^{2}(t)$
409: for a $512\times 512$ system. In this case we see that
410: $t_{c}$ has been extended to much larger values and we have
411: not been able to follow the ordering process to completion.
412: Our fits to $\Delta E(t)$ and $\Delta \psi^{2}(t)$ in the
413: regime $t_{s}< t <t_{c}$ again gives, to higher accuracy,
414: $x_{E}=x_{\psi}=1/3$.
415:
416: \begin{figure} % Fig. 2
417: \begin{center}
418: \includegraphics[scale=.43]{fig/fig_2.eps}
419: \end{center}
420: \caption{\footnotesize $\Delta E(t)$ and $\Delta \psi^2(t)$ for a
421: $512\times 512$
422: system. The data for $t<t_s$ are not shown. The straight lines are used
423: to guide eyes. Averaged over $57$
424: different trials.}
425: \end{figure}
426:
427: To probe directly the stripes' increasingly orientational order,
428: we measure the {\it nematic}
429: order parameter correlation function $C_{nn}(\mathbf{r},t)$
430: in the $512\times 512$ system. The results,
431: averaged over $57$ runs, are shown in Fig. 3. We obtain
432: scaling with a correlation
433: length obeying the growth law $L_{n} \propto
434: t^{0.36}$. We can estimate the time $t_c$ when the cross-over
435: begins in this larger
436: system as follows. The system becomes anisotropic and one
437: expects cross-over when the correlation length $L_n$
438: grows to be some substantial fraction of a lateral dimension
439: of the system. In terms of ratios we can write
440: \begin{equation}
441: \frac{L_n(t_{c}(512))}{L_n(t_c(256))}\approx \frac{512}{256}=\left[\frac{t_c(51
442: 2)}{t_c(256)}\right]^{1/3}
443: ~~~.
444: \end{equation}
445: In the $256\times 256$ system $t_c (256)\sim 9000$, so we
446: obtain $t_c(512)\sim 60000$.
447: Notice that in Fig. 2 the effective exponent $x_E$ begins to increase at
448: the time $50000\sim 70000$, which is consistent with our estimate.
449:
450: \begin{figure} % Fig. 3
451: \begin{center}\includegraphics[scale=.31]{fig/fig_3.eps}\end{center}
452: \caption{\footnotesize Time evolution of the correlation function
453: $\displaystyle
454: C_{nn}(r,t)$ in
455: $512\times 512$ SH system
456: illustrated with times $6\times 10^3,1.2\times
457: 10^4,1.8\times 10^4$ increasing from left to right. We extract the time
458: evolution of the correlation length $L(t)$ by monitoring the $r_\alpha
459: (t)$ for which $C_{nn}(r_\alpha(t))=\alpha$, where we choose
460: $\alpha = \{0.3,0.4,0.5,0.6,0.7\}$. The scaling exponent $x_n$ is
461: extracted from the log-log plot insert of $r_\alpha(t)$ v.s. t by fitting it
462: with a straight line. Averaged over $57$ trials.}
463: \end{figure}
464:
465: \begin{figure} % Fig. 4
466: \begin{center}\includegraphics[scale=.31]{fig/fig_4.eps}\end{center}
467: \caption{\footnotesize The structure factor $S(k,t)=\langle
468: |\psi(\mathbf{k},t)|^2 \rangle$ in the $512\times 512$
469: system. The log-log plot of $S(q_0,t)$ v.s. $t$ can be fit to $t^x$ with
470: $x=0.24$. The scaling collapse of the structure factor was obtained with
471: $x=0.24$ as the scaling exponent. Averaged over $57$ independent trials.}
472: \end{figure}
473:
474: In Fig. 4 we plot the structure factor $S({\bf k},t)$ and show
475: that scaling holds in the form given by Eq.(\ref{eq:19}) with a growth law
476: characterized by an exponent $x_{s}=0.24$ as shown in the insert.
477: Our results here agree with those found previously that the
478: exponent governing the growth law for the structure factor is
479: significantly smaller than that governing the nematic order
480: parameter.
481:
482: \section{Defect Structures and dynamics}
483:
484: In Fig. 5 we show a typical configuration for the Swift-Hohenberg
485: model for a quench to zero
486: temperature after a time $12000$ for a $512\times 512$ system. Notice the
487: rather complicated structure which includes dislocations,
488: disclinations and grain (domain) boundaries. Our main focus in this
489: paper is to study the statistics of these defects.
490: In appendix A we discuss an algorithm for picking out the defects and
491: tracking their motions.
492:
493: \begin{figure} % Fig. 5
494: \begin{center}\includegraphics[scale=.53]{fig/fig_5.eps}\end{center}
495: \caption{\footnotesize A typical configuration for the SH model for a
496: quench to zero temperature after $t=12000$ in a $512\times 512$ system. The
497: black points correspond to $\psi({\bf x})>0$, and the
498: white points to $\psi({\bf x})<0$. }
499: \end{figure}
500:
501: If we look at Fig. 5 we see that it shows a complicated situation
502: with a variety of different defect structures which one can
503: identify by eye at the length scale of several layer spacings.
504: At a more fundamental level we need a way of identifying which
505: points in space, at the level of each site on the numerical
506: grid, are part of a defect.
507: At the shortest length
508: scale in the problem
509: the order parameter is $Q_{\alpha\beta}$
510: defined by Eq.(\ref{eq:12}).
511: For this two-dimensional system this can be replaced
512: by the vector order parameter ${\bf\hat{B}}$ defined by
513: Eqs.(\ref{eq:15}) and (\ref{eq:16}). The assumption is that all of
514: the defects in the system
515: can be built up from the $\pm\frac{1}{2}$ disclinations in the
516: director field $\hat{n}$ which translate into vortices with charge
517: $\pm 1$ for the field ${\bf\hat{B}}$.
518: We identify these defects by looking for the cores of the vortices.
519: We can find the cores of the defects by looking from those sites
520: where ${\bf\hat{B}}$ is changing rapidly.
521: We can define
522: \be
523: A=\sum_{\alpha ,\beta}\left(\nabla_{\alpha}B_{\beta}\right)^{2}
524: \ee
525: and
526: identify defect points as those sites where $A$ is larger than some
527: value. Notice that $A$ can also be written in the form
528: \be
529: A=4\sum_{\alpha ,\beta}\left(\nabla_{\alpha}n_{\beta}\right)^{2}
530: =\left(\nabla_{\alpha}\varphi\right)^{2}
531: ~~~.
532: \ee
533: The precise numerical determination of $A$ is discussed in
534: appendix A. Notice that $A$ is proportional to the
535: gradient energy for an isotropic nematic.
536:
537: In analyzing their
538: experimental data Harrison, et al. \cite{Harrison,H2002} found a set
539: of {\it fundamental} disclinations and from these built up
540: dislocations as bound disclinations with opposite
541: charge. They used this procedure to identify a large
542: dislocation density. Most of the fundamental disclinations
543: went into forming these dislocations since in the end the
544: ratio of dislocations to the remaining disclinations was
545: about ten to one. In our case
546: the situation is complicated by the grain boundaries. We first
547: separate the defects into compact point defects and larger
548: grain boundaries. For the point defects we determine the
549: topological charge by taking the usual phase-angle path
550: integral around the center of mass of the defect. Those
551: defects with plus or minus unit charge are identified as
552: disclinations, while those with zero charge are dislocations. Then we
553: can track the motion of each single defect.
554:
555: \begin{figure} % Fig. 6
556: \begin{center}\includegraphics[scale=0.31]{fig/fig_6.eps}\end{center}
557: \caption{\footnotesize A typical example of the track of a dislocation's
558: ``mass center'', see appendix A. The dislocation moves along the arrow. It starts at
559: $t=1590$ and disappears at $t=29220$. Notice the small arcs along the
560: curve, the diameters of the arcs are about $2\pi$, which is equal to two layer spacing.}
561: \end{figure}
562:
563: As an example of the method we show in Fig. 6
564: the path of a dislocation. We see that some dislocations travel
565: over long distances during
566: very long times. It seems that the dislocations are more stable
567: when compared to
568: grain boundaries and disclinations. Our simulations on $256\times 256$ system show
569: that after the annihilation of point defects
570: and grain boundaries, some dislocations still exist in the system.
571: Our simulations show that there are
572: also dislocations which are pinned and move little.
573:
574: The number of disclinatins is quite small. And we notice that they are
575: rather immobile, which is consistent with Boyer and Vi\~nals' discussion \cite{BVdc}.
576:
577: The most important motion of grain boundaries is that they can move over
578: long distances and combine with other grain boundaries.
579: As shown in Fig. 7, two grain boundaries can combine to
580: form a larger grain boundary. Thus the number of grain boundaries
581: decreases while their average
582: size increases. This process happens on a time scale of the
583: order $1000$ dimensionless time units.
584:
585: \begin{figure} % Fig. 7
586: \begin{center}
587: \includegraphics[scale=.28]{fig/fig_7.eps}
588: \end{center}
589: \caption{\footnotesize The combination of two grain boundaries in a
590: $512\times 512$ system. The portion shown is $200\times
591: 200$. From left to right and top to bottom, the
592: times are $t = 2880$, $2955$, $3030$, $3105$, $3180$, $4710$. Not all the
593: points in the grain boundaries are shown.}
594: \end{figure}
595:
596: \begin{figure} % Fig. 8
597: \begin{center}
598: \includegraphics[scale=.33]{fig/fig_8.eps}
599: \end{center}
600: \caption{\footnotesize The motion of a grain boundary in a $512\times
601: 512$ system. The portion shown is $200\times
602: 200$. From left to right and top to bottom, the times
603: are $t=11415$, $13665$, $15915$, $18165$, $20415$, $22665$, $24915$, $27165$,
604: $29415$. Again not all the points in the grain boundary are shown.}
605: \end{figure}
606:
607: As shown in Fig. 8, one grain boundary can sweep across a quite large
608: area. At the same time its size decreases. This process occurs on a
609: time scale of the order $10000$. According to our
610: observations, the grain boundaries' motions also relieve the stripe
611: curvatures through disclination annihilations. After one grain boundary
612: passes through a disclination, the disclination disappears. This is
613: consistent with Boyer and Vi\~nals' prediction \cite{BVdc}.
614:
615: \begin{figure} % Fig. 9
616: \begin{center}\includegraphics[scale=.31]{fig/fig_9.eps}\end{center}
617: \caption{\footnotesize The number of points in point defects and grain
618: boundaries in the $256\times 256$ system. The data for all defects and
619: domain walls are averaged over $40$ trials. The others are averaged
620: over 38 trials.}
621: \end{figure}
622:
623: Next we focus on the statistics of the
624: defects generated by the model.
625: In Fig. 9 we plot the total number of points in grain
626: boundaries, dislocations and disclinations separately for the $256\times 256$
627: system.
628: We see that the grain boundaries dominate. In the
629: scaling regime ($t_s<t<t_c$), we see that the number of points
630: corresponding to grain boundaries and all defect points, the curves $a$
631: and $b$ can be fit to $\sim
632: t^{-1/3}$. At late stages the disclinations
633: disappear, while the dislocations and grain boundaries persist. The number of
634: disclinations decreases much faster than the other defects.
635:
636: \begin{figure} % Fig. 10
637: \begin{center}\includegraphics[scale=.31]{fig/fig_10.eps}\end{center}
638: \caption{\footnotesize The average number of grain boundaries and the average
639: number of points in a grain boundary. $256\times 256$ system
640: averaged over $40$ trials.}
641: \end{figure}
642:
643: In Fig. 10 we plot the average number of grain boundaries $\bar{n}$ and the
644: average size of a grain boundary $\bar{l}$ for the $256\times 256$ system. We
645: use the number of points in one grain boundary
646: as a measure of its size. For $t_s<t<t_c$, $\bar n$ decreases but $\bar l$
647: increases due to the combining of grain boundaries. The
648: shrinkage of their sizes is not as important as the combinations. However for
649: $t_c \sim 9000$ the correlation length $L_n$ is the same order as the
650: system's size, and the large grain boundaries stop growing. After that
651: the shrinkage is important \cite{SHRINK}. In the scaling regime ($t_s<t<t_c$),
652: $\bar n
653: \sim t^{-0.45}$, and $\bar{l} \sim t^{0.13}$. So $\bar n \bar{l} \sim
654: t^{-1/3}$, which is consistent with Fig. 9.
655:
656: \begin{figure} % Fig. 11
657: \begin{center}\includegraphics[scale=.31]{fig/fig_11.eps}\end{center}
658: \caption{\footnotesize The average number of points in defects and grain
659: boundaries in the $512\times 512$ system. Most of the points are in grain
660: boundaries. All
661: the curves can be fit to $t^{-y}$. The data for defects and domain
662: walls are averaged over $57$ trials, and the other data are averaged
663: over 20 trials.}
664: \end{figure}
665:
666: \begin{figure} % Fig. 12
667: \begin{center}\includegraphics[scale=.31]{fig/fig_12.eps}\end{center}
668: \caption{\footnotesize $\bar n$ and $\bar{l}$ v.s. time in the $512\times 512$
669: system. The data are averaged over $57$ independent trials.}
670: \end{figure}
671:
672: In Fig. 11 we plot the total number of points in grain
673: boundaries, dislocations and disclinations separately for the $512\times
674: 512$ system. The number of points in disclinations is much smaller than
675: that of dislocations and at very late stages it decreases to the order
676: of 1, which in fact indicates the disappearance of disclinations. Now the
677: scaling regime extends to much longer times. The domain walls and
678: dislocations' scaling exponents are both $1/3$, which is same to the
679: scaling of the energy. However, the scaling
680: exponent of disclinations, given by 0.57 is much larger. So we conclude that
681: disclinations are not the dominant structures in SH system.
682:
683: In Fig. 12 we plot the number of grain boundaries $\bar{n}$ and the
684: average size of a grain boundary $\bar{l}$ for $512\times 512$
685: system. All of our data falls in the scaling regime..
686: The plot of the average
687: number $\bar n$ of grain boundaries versus time $t$, can be fit to $\bar n \sim
688: t^{-0.49}$ and the average size $\bar{l}$ v.s. time $t$, can be fit by
689: $\bar{l} \sim t^{0.17}$.
690: So we have $\bar n \bar{l} \sim L^{-1} \sim t^{-1/3}$, which is
691: consistent with the results shown in Fig. 11.
692:
693: Although we did not count
694: the number of disclinations directly, it is proportional to the number
695: of lattice points in disclinations,
696: i.e. $t^{-0.57}$. This is because the average number of lattice points in one point
697: disclination, which is about $10\sim20$, is quite stable during the
698: simulation. By the same reasoning, we find that the number of dislocations
699: is proportional to $t^{-1/3}$. It is interesting to note that the number of grain
700: boundaries scales as $t^{-0.49}$. This exponent is near to that for
701: disclinations.
702:
703: The number of grain
704: boundaries is about 5 at $t\sim 70000$, the number of disclinations is
705: about $0$ or $1$ at $t\sim 50000$,
706: and the number of dislocations is on the order of 10 at $t\sim 50000$, as can be seen in
707: Figs. 11 and 12.
708:
709: \begin{figure} % Fig. 13
710: \begin{center}\includegraphics[scale=0.85]{fig/fig_13.eps}\end{center}
711: \caption{\footnotesize A typical example of probability density $P$
712: v.s. speed $v$ for the speed distribution of
713: dislocations at time $1350$ for the $512\times 512$ system. The distributions at other
714: times have approximately the same shape. Averaged over 56 trials.}
715: \end{figure}
716:
717: Since we can track the motion of each defect, we can
718: measure their speeds. We define the speed of
719: each as the speed of its mass center (see appendix A for
720: the definition of ``mass center''). If in a time $\Delta \tau$, the mass center
721: travels over a distance $\Delta d$, then the speed is $v=\Delta
722: d/\Delta \tau$. If
723: $\Delta \tau$ is small enough, we found that $v=0$ has the biggest probability.. If $\Delta\tau$ is large enough, all the details are
724: coarse-grained and we observe a continuous distribution of the speed and
725: the largest probability appears at a non-zero speed, as is shown in
726: Fig. 13 where $\Delta \tau =60$. We measured the speed distributions of
727: domain walls and point defects separately. As we have already seen, for
728: point defects the number of disclinations is much smaller than that of
729: dislocations, so what we measure in the latter case is in fact the speed distribution of dislocations.
730:
731: The speed distribution has
732: a long tail which decreases as a power law. The numerical fits at
733: different times give us different exponents. However the tail exponents
734: at different times do distribute in a narrow region, as is shown in
735: Fig. 14. The exponents of grain boundaries are quite different from those
736: of point defects. If we ignore the exponents at very early times
737: when grain
738: boundaries just begin to form and at the very late times when the grain
739: boundaries have already disappeared, the mean value of the grain
740: boundaries' exponents is $-1.50$, and that of
741: point defects' exponents is $-2.10$.
742:
743: \begin{figure} % Fig. 14
744: \begin{center}\includegraphics[scale=0.31]{fig/fig_14.eps}\end{center}
745: \caption{\footnotesize The power law exponents of the speed distribution's
746: tail at different times. Ignoring the data points at very early times and
747: very late times, the mean value of the exponents is $-1.5$ for grain
748: boundaries, $-2.1$ for point defects and $-1.7$ for all the defects.
749: Averaged over $61$
750: independent trials.}
751: \end{figure}
752:
753: \begin{figure} % Fig. 15
754: \begin{center}\includegraphics[scale=0.31]{fig/fig_15.eps}\end{center}
755: \caption{\footnotesize The average speed for defects and grain
756: boundaries. From top to bottom, the curves have the form $v_0\times
757: t^{-x}$ with $v_0$ being a constant. From top to bottom, $x=0.35\pm 0.01$, $x=0.41\pm 0.01$ and
758: $x=0.48\pm0.01$. Averaged over $56$
759: independent trials.}
760: \end{figure}
761:
762: We
763: also measured the average speed of the point defects and grain
764: boundaries as a function of time after the quench,
765: as is shown in Fig. 15. The average speed of point defects decreases as $\sim
766: t^{-0.48}$;
767: the average speed of grain boundaries goes as $\sim t^{-0.35}$. The
768: scattering of the points at late stages is due to the small data sample
769: at those times..
770:
771: %\newpage
772:
773: \section{Summary}
774:
775: We have studied the dynamics of the defect structures in the SH model
776: after quenches to zero temperature with a control parameter of
777: $\epsilon=0.1$. We find in agreement with earlier workers that the
778: kinetics in the ordering regime, before finite-size effects enter,
779: are dominated by the existence of moving and coalescing grain
780: boundaries. In this regime the average size of these grain boundaries is
781: growing and they are relatively mobile. Under the influence of finite
782: size effects these grain boundaries shrink, the system becomes
783: anisotropic and the ordering process speeds up.
784:
785: We also measured the speed distribution of all structures that appear in
786: the system. The average speed is decreasing as a power law and the
787: distributions show a power-law behaviors at large speeds.
788: However we can only get a rough estimate due to the poor statistics.
789:
790: Let us return to the question of whether the SH model gives a
791: good description of the physical system studied by Harrison,
792: et al.\cite{Harrison,H2002}. The SH model,
793: for small
794: control parameter $\epsilon$, does give coarsening with an
795: exponent in roughly the same range as in the experiment
796: ($1/4\sim 1/3$). The ordering is constrained to be slower
797: then the picture where one has
798: a simple point defect pair annihilation process. However,
799: the defect structures in the SH model and experiment
800: appear quite different. The disclination quadrapole annihilations
801: seen in experiments are not observed in the late stages of the
802: evolution of the
803: SH model. In the SH model grain boundaries dominate the
804: evolution in the scaling regime, but these structures appear
805: to play a limited role in the experiments.
806: We must make clear that our numerical results are for systems
807: with many fewer roll periods compared to the experimental systems
808: ($10^{2}$ compared to $10^{5}$), so it is possible that things
809: change as we increase the size of the ordering system. However, our
810: study of 256 and 512 systems shows that they differ only in
811: the time when the finite-size effect enters. This indicates that
812: a even larger system will display the same behavior except that the
813: finite-size effect enters at a even later time.
814: So we conclude that the SH model does not
815: give a physically faithful description of the ordering in the
816: experimental system.
817: This raises the provocative question:
818: Are there many different types of scenarios for ordering striped
819: systems? We will address this question by looking at other
820: competing models for striped formation elsewhere.
821:
822: \vspace{5mm}
823:
824: Acknowledgments: We thank Dr. C. Harrison for providing
825: us with a copy of his thesis. This work was supported by the National
826: Science Foundation under Contract NO. DMR-0099324.
827:
828: %\newpage
829:
830: \appendix
831:
832: \section{The algorithm for picking out defects and finding
833: domain walls}
834:
835: Hou, \textit{et al.} \cite{HSG} proposed a method, the HSG method, to measure
836: the length of grain boundaries.
837: They computed the quantity $A^2 \equiv \psi^2+(\nabla \psi)^2/q_0^2$, and if
838: the calculated $A^2$ is bigger than an upper filter $0.7\times
839: Max(A^2)+0.3\times
840: AVG(A^2)$ or smaller than a lower filter $0.7 \times Min(A^2)+0.3\times
841: AVG(A^2)$,
842: that point is counted as belonging to a domain wall. When $\epsilon$ is
843: small, this method gives quite good results.
844:
845: However, when $\epsilon$ increases, the original filters are no longer
846: applicable. They fail to pick out most of the points and the filters
847: must be re-chosen.
848: For example, at $\epsilon=0.6$, the filters $0.5\times Max(A^2)+0.5\times
849: AVG(A^2)$ and $0.5 \times Min(A^2)+0.5\times AVG(A^2)$ can give a
850: satisfying result; while for $\epsilon = 0.75$, $0.4\times Max(A^2)+0.6\times
851: AVG(A^2)$ and $0.4 \times Min(A^2)+0.6\times AVG(A^2)$ are the better
852: choices. Sometimes this method is unable to pick out all the defects for
853: any choice of filter.
854:
855: We introduce here a
856: method which works for all $\epsilon$ and picks
857: out all of the defects and nothing more.
858:
859: First let us define some useful quantities. Suppose the system is
860: discrete on the $x\-y$ plane, with ${\bf x}=(i,j)$ (square lattice). At a
861: fixed time,
862: starting from the order
863: parameter field $\psi ({\bf x})=\psi_{i,j}$, we can define a director
864: field $\hat{n}({\bf x})$
865: as given by Eq.(11)
866: where $\nabla \psi({\bf x})$ is defined by the usual finite difference
867: scheme, i.e.
868: \begin{eqnarray}
869: \nabla \psi({\bf x})= \left(\frac{\psi_{i+1,j}-\psi_{i-1,j}}{2\,\Delta
870: r},\frac{\psi_{i,j+1}-\psi_{i,j-1}}{2\,\Delta r}\right) \ ,
871: \end{eqnarray}
872: where $\Delta r$ is the lattice space of the system.
873: In two-dimensional cases the nematic order parameter,
874: $Q_{\alpha \beta}$ is completely specified by the angle
875: \begin{equation}
876: \varphi ({\bf x}) = 2\, \theta ({\bf x})\ ,
877: \end{equation}
878: where
879: \begin{equation}
880: \theta ({\bf x}) = \arctan
881: \left( \frac{\hat{n}_y({\bf x})}{\hat{n}_x({\bf x})}\right)\ .
882: \end{equation}
883:
884: Rather than using $\varphi ({\bf x})$ given by the two equations
885: above we introduce some local smoothing.
886: First we compute
887: \begin{eqnarray}
888: &&\hat{B}_{y}= \sin \varphi({\bf x})=2 \hat{n}_x({\bf x}) \hat{n}_y({\bf x})\ ,
889: \nonumber \\
890: && \hat{B}_{x}=\cos \varphi({\bf x})=2\hat{n}_x({\bf x})^2-1\ .
891: \end{eqnarray}
892: Then we smooth these two fields using the iterative process:
893: \begin{equation}
894: f_{(n+1)}(i,j)=\frac{1}{2}f_{(n)}(i,j)
895: +\frac{1}{8}\sum_{(i',j')\in NN}f_{(n)}(i',j')\ ,
896: \end{equation}
897: where $f_{(n)}$ is $\sin\varphi$ or $\cos \varphi$ after $n$ iterations,
898: and $NN$ means the $4$ nearest neighbors of $(i,j)$ on
899: the square lattice. This process will suppress the small fluctuations
900: of $\varphi({\bf x})$ away from the defects, while the variation of
901: of $\varphi({\bf x})$ near a defect core remains large. Our calculations
902: show that $5$ iterations provides a sufficiently smooth set of
903: fields for our purposes.
904: In the next step, we calculate
905: $\varphi({\bf x})$ from $\sin \varphi({\bf x})$ and $\cos
906: \varphi({\bf x})$ using
907: \begin{equation}
908: \varphi({\bf x})=\arctan \left[\frac{\sin
909: \varphi({\bf x})}{\cos\varphi({\bf x})}\right]\ ,
910: \end{equation}
911: where we adopt the convention that $-\pi < \varphi({\bf x}) <\pi$.
912:
913: In picking a filter we want to look at the spatial variation
914: of the $\varphi({\bf x})$ field.
915: $\nabla \varphi({\bf x})$ can be evaluated as for $\nabla
916: \psi({\bf x})$. However, there is a subtlety here. For example, if
917: $\varphi_{i+1,j}=\pi-\delta \phi_1$ and $\varphi_{i-1,j}=-\pi+\delta
918: \phi_2$, where $\delta \phi_1$ and $\delta \phi_2$ are small angles,
919: then the difference between the nematic tensor
920: $Q_{\alpha\beta}(i+1,j)$ and $Q_{\alpha\beta}(i-1,j)$ should be a small
921: quantity. But
922: $\displaystyle(\varphi_{i+1,j}-\varphi_{i-1,j})/2
923: =\pi-(\delta\phi_1+\delta\phi_2)/2 \sim \pi$, which
924: means that if we calculate the change rate of $\varphi({\bf x})$ in
925: exactly the same way as Eq.(A1), we will get a wrong answer in this
926: context. To avoid such a
927: problem, we define the difference between $\varphi({\bf x})$ and
928: $\varphi({\bf x'})$ as the
929: quantity with the smallest absolute value among the choices
930: $\varphi({\bf x}')-\varphi({\bf x})$ and
931: $\varphi({\bf x}')-\varphi({\bf x})\pm 2\pi$. And we use this quantity
932: in determining
933: \be
934: A({\bf x})=|\nabla
935: \varphi({\bf x}) |^2
936: \ee
937: which is the key quantity in our analysis.
938:
939: \begin{figure} % Fig. 16
940: \begin{center}\includegraphics[scale=.4]{fig/fig_16.eps}\end{center}
941: \caption{\footnotesize In the lower graph, the vector field $(\cos
942: \varphi(\vec{x}),\sin \varphi(\vec{x}))$ for the order parameter field
943: shown above. The
944: components of the vector field have been smoothed over $5$ iterations. The
945: lattice spacing is $\pi/4$, which means there are $8$ points in one
946: period of the layers. Not all the vectors on the lattice are shown.}
947: \end{figure}
948:
949: Our method
950: is based on the observation that at the core region of a defect
951: (dislocation, disclination or part of a domain wall) the angle
952: field $\varphi({\bf x})$ changes
953: rapidly, while in the region away from the defect's core the
954: $\varphi({\bf x})$ field is
955: rather smooth, as can be seen in Fig. 16. Thus we can conclude with
956: confidence that those points with larger change rates of
957: $\varphi({\bf x})$ must belong to some defect's core region or a part
958: of a grain boundary.
959:
960: \begin{figure} % Fig. 17
961: \begin{center}\includegraphics[scale=0.45]{fig/fig_17.eps}\end{center}
962: \caption{\footnotesize In the lower graph, the scalar field
963: $A(\vec{x})$ is plotted. This corresponds to
964: the order parameter field shown above. $A(\vec{x})$ is sharply peaked at the core regions of the defects.}
965: \end{figure}
966:
967: We find that $A({\bf x})\approx 0$ away from defects,
968: but increases very rapidly in the vicinity of any defect. An example
969: is given in Fig. 17. Therefore as long as $A({\bf x})$ is large
970: enough, we
971: can identify the point ${\bf x}=(i,j)$ as part of the core region of a
972: defect. Naturally we set up a threshold $A_0$, and any point with
973: $4(\Delta r)^2\cdot A({\bf x})>A_0$ is counted as belonging to some defect's core. Because the
974: value of $A({\bf x})$ is much larger in the defects' cores than at any other
975: places, a range of values of $A_0$ can be used to find the positions of the defects core
976: regions. With a smaller threshold the program will pick out more points
977: in the core regions,
978: and with a larger one it will pick out fewer points in the core regions. Our
979: experience shows that if $A_0$ takes the value of $2\sim 10$, the
980: program picks out the same defects cores and grain boundaries.
981: As is shown in Fig. 18 and Fig. 19, it
982: picks out all the defects without irrelevant points.
983:
984: \begin{figure} % Fig. 18
985: \begin{center}\includegraphics[scale=0.63]{fig/fig_18.eps}\end{center}
986: \caption{\footnotesize Identification of all the defects in a $512\times
987: 512$
988: system ($\epsilon = 0.1$) with a threshold $A_0=3.5$. At each
989: defect core region, the $A$ field for many
990: points exceeds the threshold. The red points belong to domain walls,
991: the green ones belong to dislocations and the blue ones belong to disclinations.}
992: \end{figure}
993:
994: \begin{figure} % Fig. 19
995: \begin{center}\includegraphics[scale=0.63]{fig/fig_19.eps}\end{center}
996: \caption{\footnotesize Identification of all the defects in a $512\times
997: 512$
998: system ($\epsilon = 0.5$) again with a threshold $A_0=3.5$. Apparently the defect's density in this system is greater than
999: the density in Fig. 18. The domain walls are much smaller for the system
1000: with larger $\epsilon$. There
1001: are no domain walls for $\epsilon > 0.6$.}
1002: \end{figure}
1003:
1004: After we have used the above algorithm to pick out the points in the core
1005: regions
1006: of the point defects and grain boundaries, we can distinguish between these
1007: two structures. The difference between them is obvious. The point defects are
1008: compact in space
1009: while the grain boundaries are ramified.
1010:
1011: First, we must group the points we have identified according to whether
1012: or not they are in the same structure. The points in one point defect core
1013: or grain
1014: boundary are picked out because the director field changes drastically
1015: on those sites. They are very near to each other. However, they may not be
1016: neighbors. So we define a filter $a_0$, and
1017: when any two points' distance is less than $a_0$, they are supposed to
1018: be in the same defect or grain boundary's core region. We use the
1019: cluster multiple labeling method of Hoshen and Kopelman
1020: \cite{HK} to pick out
1021: such point clusters. Thus given the
1022: system's status at any time, we can find those sets of points
1023: corresponding to each individual defect or grain boundary.
1024:
1025: Now we measure the approximate size of these structures and then
1026: distinguish between point defects and grain boundaries. We use the number of
1027: the points in the set as the size of the corresponding
1028: structure. This approximation
1029: reflects the actual size of the defect or grain
1030: boundary quite well. Then we define a filter $l_0$, and when the structure's
1031: size is larger than $l_0$, we regard the corresponding structure as a
1032: grain boundary, otherwise it's taken to be a point defect (dislocation or
1033: disclination).
1034:
1035: We employed $a_0=5\,\Delta r$ where $\Delta r$ is the
1036: lattice spacing and $l_0=18$. The results are quite satisfying.
1037:
1038: After we have picked out the point defects, we can devide them into
1039: disclinations and dislocations. We
1040: follow Harrison's method \cite{Harrison,H2002}. Given the angle field
1041: $\varphi({\bf x})$ computed in Eq.(A6) after the smoothing process, we do an integral of the
1042: variation of $\varphi({\bf x})$ over
1043: a counterclockwise close path around the ``mass-center'' of a point
1044: defect, which is defined below. The condition for a defect to be a
1045: disclination is
1046: \begin{equation}
1047: \oint \frac{\partial \varphi}{\partial s}\,ds = \pm 2\pi
1048: \end{equation}
1049: The integral is zero if the defect is a dislocation. To make the
1050: computation easier, we choose a $16\times 16$ square with the
1051: mass-center at its center as the integration route.
1052:
1053: To record the motion of one single defect or grain boundary, we track
1054: the motion of the corresponding point set's ``mass center'', which is
1055: defined as follows. Suppose the point set has $n$ points with
1056: coordinates ${\bf r}_i,\,i=1,2,...,n$. Then the ``mass center'' of the
1057: point set is defined as $\displaystyle {\bf r}=\sum_i {\bf r}_i/n$, just
1058: like the usual mass center definition in classical mechanics but with
1059: all masses equal to one.
1060:
1061: In the evolution of the SH model, we sample the
1062: system every $500$ time steps (in our case this is equal to 15
1063: dimensionless time units), which is a quite short-time period in the
1064: simulation. We then identify all the dislocations, disclinations and grain boundaries,
1065: distinguish among them, and compute their {\it centers~of~mass}.
1066: Suppose at time $t_1$, we have the set of {\it mass~ centers}
1067: $P=\{{\bf p}_i,\, i=1,2,...,n_1\}$, and at time $t_2$, the ``mass center''
1068: set is $Q=\{{\bf q}_j,\, j=1,2,...,n_2\}$; usually $n_1 \ne n_2$. Define
1069: $d_{PQ}(i,j)=|\,{\bf p}_i-{\bf q}_j|$. We assume that the
1070: defects and grain boundaries do not move much in such a short time
1071: period. So if there exist two integers $k \in [1,n_1]$ and $l \in
1072: [1,n_2]$, such that
1073: \begin{equation}
1074: d_{PQ}(k,l)=\min_{j \in [1,n_2]}
1075: d_{PQ}(k,j)=\min_{i \in [1,n_1]} d_{PQ}(i,l) \ ,
1076: \end{equation}
1077: it is quite
1078: reasonable to believe that ${\bf p}_k$ and ${\bf q}_l$ are just the same
1079: defect's or grain boundary's ``mass center'' at two successive times. Using
1080: this method, we are able to find out the trajectories of the ``mass
1081: centers'' as time goes on. Not all points in $P$ and $Q$ can be grouped
1082: into such pairs. On the one hand, this is because $n_1 \ne n_2$; on the
1083: other hand, this is also due to the criterion (A9) applied onto
1084: ${\bf p}_k$ and ${\bf q}_l$. Physically, this is consistent with the
1085: phenomena of the defect annihilation and the combination, split and shrinkage
1086: of
1087: grain boundaries.
1088:
1089: \section{Measurement of the Nematic Correlation Function}
1090:
1091: To probe the stripes' increasingly orientational order, we define the
1092: correlation function which is similar to the one employed by
1093: Christensen and Bray\cite{CB}.
1094: \begin{eqnarray}
1095: \lefteqn{C_{nn}({\bf r},t)=\frac{1}{N^2} \sum_{{\bf x}} \langle \,
1096: \cos\,[\varphi({\bf x}+{\bf r},t)-\varphi({\bf x},t)] \,\rangle {}} \nonumber
1097: \\
1098: & & {}= \frac{1}{N^2} \sum_{{\bf x}} \langle \, \cos
1099: \varphi({\bf x}+{\bf r},t)\cdot \cos \varphi({\bf x},t)\, \rangle + {}
1100: \nonumber \\
1101: & & {}+ \frac{1}{N^2}
1102: \sum_{{\bf x}} \langle \, \sin
1103: \varphi({\bf x}+{\bf r},t) \cdot \sin({\bf x},t) \,\rangle \ ,
1104: \label{eq:26}
1105: \end{eqnarray}
1106: where $N^2$ is the area of the system and the angular brackets denote
1107: the statistical average over different initial conditions. The
1108: definition of the angle $\varphi({\bf x})$ is given in appendix A.
1109:
1110: Now in Eq.(\ref{eq:26}) the function has been split into two parts
1111: which have the
1112: same form
1113: \begin{equation}
1114: G({\bf r},t)=\frac{1}{N^2} \sum_{{\bf x}} \langle \,
1115: f({\bf x}+{\bf r},t)\cdot f({\bf x},t) \,\rangle \ ,
1116: \label{eq:27}
1117: \end{equation}
1118: with $f({\bf x},t)=\cos \varphi({\bf x},t)$ and $f({\bf x},t)= \sin
1119: \varphi({\bf x},t)$ separately. Eq.(\ref{eq:27}) can be easily calculated by
1120: fast
1121: Fourier transformation (FFT). First FFT $f({\bf r},t)$\,
1122: to obtain its Fourier components
1123: $\tilde{f}({\bf k},t)$. Then $\tilde{G}({\bf k},t)= \langle \,
1124: |\tilde{f}({\bf k},t)|^2 \, \rangle $, and inverse Fourier transformation
1125: gives $G({\bf r},t)$. We
1126: compute the two parts in Eq.(\ref{eq:27}) separately, and then
1127: add to obtain the
1128: correlation function $C_{nn}({\bf r},t)$.
1129:
1130: \begin{thebibliography}{99}
1131:
1132: %1
1133: \bibitem{BN} C. Bowman and A. C. Newell,
1134: Rev. Mod. Phys. {\bf 70}, 289 (1998).
1135:
1136: %2
1137: \bibitem{Harrison} C. Harrison, D. H. Adamson, Z. Cheng,
1138: J. M. Sebastian, S. Sethuraman, D. A. Huse, R. A. Register and P. M. Chaikin,
1139: Science {\bf 290}, 1558 (2000).
1140:
1141: %3
1142: \bibitem{H2002} C. Harrison, Z. Cheng, S. Sethuraman,
1143: D. A. Huse, P. M. Chaikin,, D. A. Vega, J. M. Sebastian,,
1144: R. A. Register, and D. H. Adamson,
1145: Phys. Rev. E{\bf 66}, 011706 (2002).
1146:
1147: %4
1148: \bibitem{SH} J. Swift and P. C. Hohenberg, Phys. Rev. A \textbf{15},
1149: 319 (1977).
1150:
1151: %5
1152: \bibitem{BRAY} A. J. Bray, Adv. Phys. \textbf{43}, 357 (1994).
1153:
1154: %6
1155: \bibitem{HSG} Q. Hou, S. Sasa, and N. Goldenfeld, Physica A {\bf 239},
1156: 219 (1997).
1157:
1158: %7
1159: \bibitem{BVdc} D. Boyer and J. Vi\~nals, Phys. Rev. E \textbf{64}, 050101
1160: (2001).
1161:
1162: %8
1163: \bibitem{PM79} Y. Pomeau and P. Manneville, J. de Phsique -- Lett.,
1164: \textbf{40} (23), L-609 (1979).
1165:
1166: %9
1167: \bibitem{NT} Assume that the total area of defects and grain
1168: boundaries is $a(t)$, and $\langle\psi^2\rangle$ in those regions is a constant
1169: $\chi^2$, which is smaller than $\langle \psi^2_0\rangle$. Then approximately
1170: $\displaystyle \langle\psi^2\rangle_t
1171: =\frac{1}{S}\,\left[\langle\psi_0^2\rangle\cdot\left(S-a(t)\right)+\chi^2\cdot
1172: a(t)\right]=\langle\psi_0^2\rangle -
1173: a\cdot\frac{\langle\psi_0^2\rangle-\chi^2}{S}$, i.e. $\Delta \psi^2(t)=
1174: \langle\psi_0^2\rangle-\langle \psi^{2}\rangle_{t} \propto a(t)$.
1175: Since the energy should
1176: concentrate on the defects and grain boundaries, it is reasonable to
1177: expect that $a(t) \propto \Delta
1178: E(t) \propto L_E^{-1}(t)$, which is supported by the simulations.
1179: So $\Delta \psi^2(t)\propto
1180: L_{E}^{-1}(t)$. $L_E(t)$ and $L_{\psi}(t)$ are expected to have the same
1181: growth exponent. This is also supported by our simulations.
1182:
1183: %10
1184: \bibitem{CB} J. J. Christensen and A. J. Bray, Phys. Rev. E \textbf{58},
1185: 5364 (1998).
1186:
1187: %11
1188: \bibitem{EVG} K. R. Elder, J. Vi\~nals, and M. Grant, Phys. Rev. Lett. {\bf
1189: 68},
1190: 3024 (1992).
1191:
1192: %12
1193: \bibitem{TN81} J. Toner and D. R. Nelson, Phys. Rev. B {\bf 23}, 316 (1981).
1194:
1195: %13
1196: \bibitem{CM} M. C. Cross and D. I. Meiron, Phys. Rev. Lett. {\bf 75},
1197: 2152 (1995).
1198:
1199: %14
1200: \bibitem{BVgm} D. Boyer and J. Vi\~nals, Phys. Rev. E \textbf{63}, 061704
1201: (2001).
1202:
1203: %15
1204: \bibitem{BVgbp} D. Boyer and J. Vi\~nals, Phys. Rev. E \textbf{65},
1205: 046119 (2002).
1206:
1207: %16
1208: \bibitem{SHRINK} One may doubt that the domain walls break up at this
1209: time so that their sizes decrease. However if this was true, the number
1210: of domain walls would increase, which is obviously wrong, as is shown in
1211: Fig. 10. So the main effect that decreases the size of the domain walls
1212: is the shrinkage.
1213:
1214: %17
1215: \bibitem{HK} H. Gould and J. Tobochnik, \textit{An Introduction
1216: to Computer Simulation Methods, Part
1217: 2}, Addison-Wesley Publishing Company (1988).
1218:
1219: \end{thebibliography}
1220:
1221: \end{multicols}
1222: \end{document}
1223:
1224:
1225:
1226:
1227:
1228:
1229:
1230: